You are on page 1of 367

Mathematics for Industry 16

OsamiMatsushita
MasatoTanaka
HiroshiKanki
MasaoKobayashi
PatrickKeogh

Vibrations
of Rotating
Machinery
Volume 1. Basic Rotordynamics:
Introduction to Practical Vibration
Analysis
Mathematics for Industry

Volume 16
Editor-in-Chief
Masato Wakayama (Kyushu University, Japan)

Scientic Board Members

Robert S. Anderssen (Commonwealth Scientic and Industrial Research Organisation, Australia)


Heinz H. Bauschke (The University of British Columbia, Canada)
Philip Broadbridge (La Trobe University, Australia)
Jin Cheng (Fudan University, China)
Monique Chyba (University of Hawaii at Mnoa, USA)
Georges-Henri Cottet (Joseph Fourier University, France)
Jos Alberto Cuminato (University of So Paulo, Brazil)
Shin-ichiro Ei (Hokkaido University, Japan)
Yasuhide Fukumoto (Kyushu University, Japan)
Jonathan R.M. Hosking (IBM T.J. Watson Research Center, USA)
Alejandro Jofr (University of Chile, Chile)
Kerry Landman (The University of Melbourne, Australia)
Robert McKibbin (Massey University, New Zealand)
Andrea Parmeggiani (University of Montpellier 2, France)
Jill Pipher (Brown University, USA)
Konrad Polthier (Free University of Berlin, Germany)
Osamu Saeki (Kyushu University, Japan)
Wil Schilders (Eindhoven University of Technology, The Netherlands)
Zuowei Shen (National University of Singapore, Singapore)
Kim-Chuan Toh (National University of Singapore, Singapore)
Evgeny Verbitskiy (Leiden University, The Netherlands)
Nakahiro Yoshida (The University of Tokyo, Japan)

Aims & Scope

The meaning of Mathematics for Industry (sometimes abbreviated as MI or MfI) is different


from that of Mathematics in Industry (or of Industrial Mathematics). The latter is restrictive: it
tends to be identied with the actual mathematics that specically arises in the daily management
and operation of manufacturing. The former, however, denotes a new research eld in mathematics
that may serve as a foundation for creating future technologies. This concept was born from the
integration and reorganization of pure and applied mathematics in the present day into a fluid and
versatile form capable of stimulating awareness of the importance of mathematics in industry, as
well as responding to the needs of industrial technologies. The history of this integration and
reorganization indicates that this basic idea will someday nd increasing utility. Mathematics can
be a key technology in modern society.
The series aims to promote this trend by (1) providing comprehensive content on applications of
mathematics, especially to industry technologies via various types of scientic research,
(2) introducing basic, useful, necessary and crucial knowledge for several applications through
concrete subjects, and (3) introducing new research results and developments for applications of
mathematics in the real world. These points may provide the basis for opening a new mathematics
oriented technological world and even new research elds of mathematics.

More information about this series at http://www.springer.com/series/13254


Osami Matsushita Masato Tanaka

Hiroshi Kanki Masao Kobayashi


Patrick Keogh

Vibrations of Rotating
Machinery
Volume 1. Basic Rotordynamics:
Introduction to Practical Vibration Analysis

123
Osami Matsushita Hiroshi Kanki
The National Defense Academy Kobe University
Yokosuka Kobe
Japan Japan
and and
Hitachi, Ltd. (retired) Mitsubishi Heavy Industries, Ltd. (retired)
Tokyo Tokyo
Japan Japan

Masato Tanaka Masao Kobayashi


The University of Tokyo IHI Corporation
Tokyo Yokohama
Japan Japan

Patrick Keogh
The University of Bath
Bath
UK

ISSN 2198-350X ISSN 2198-3518 (electronic)


Mathematics for Industry
ISBN 978-4-431-55455-4 ISBN 978-4-431-55456-1 (eBook)
DOI 10.1007/978-4-431-55456-1

Library of Congress Control Number: 2016947398

Springer Japan 2017


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microlms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specic statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer Japan KK
Preface

The development of new rotating machines often encounters unexpected vibration


problems. In many reported cases these vibration problems have had to be solved
before the hydraulic performance of rotating machinery such as turbines, pumps,
and compressors could be assessed.
The v_BASE Databook compiled by the Japan Society of Mechanical Engineers
(JSME) contains a collection of vibration problems actually experienced in
industry. Its rst edition, published in 1994, includes 300 case studies. Almost
two-thirds of the data case studies included in the rst edition are related to rotating
machinery, about half of which involve resonance. The authors learned the fol-
lowing lessons associated with the case studies:
1. Serious issues caused by large of rotational energy and centrifugal force:
A seemingly slight problem in rotor dynamics may produce very dangerous
vibration in a body rotating at high speed, the same as the dynamics of a car
running at high speed on a highway. Furthermore, apparently vibration is
reproduced unless any repair is undertaken. This vibration problem, occurring
once, can be eliminated only by making specic improvements at the designing
and manufacturing stages, which may entail signicant effort and cost for the
nal correct solution. Our rotor dynamics require us to be more practical in
solving the issue to avoid encountering such a dangerous situation.
2. Knowledge versus practical experience:
Rotating machinery engineers must be accomplished with skills to achieve good
machine operation, typically through balancing. Many engineers possess good
theoretical knowledge of unbalance resonance phenomena through their basic
engineering education. However, they may have apparent difculty in applying
that knowledge on sitefor example, practical eld balancing by scoping two
waveform signals together with rotational pulses and rotor vibration, and
reading the amplitude and phase difference without data analyzers. Such dif-
culty is most easily overcome by appropriate learning and experience to apply
the knowledge to eld practice without hesitation. Field experience in

v
vi Preface

eliminating vibration for troubleshooting based on variety of measurements and


analyses offer the best educational opportunity. Experience is good teacher and
the eld is a good class.
3. Inertial (stationary) or rotating coordinate systems?
Rotor vibration is characterized by the gyroscopic effect. The term gyroscopic
effect is related to the inertial coordinate system. On the other hand, the same
phenomenon observed in the rotating coordinate system is called the Coriolis
effect as per the theory of blade vibration. This example shows the importance
of a unied and seamless understanding of both dynamics described in the
inertial and rotating frames of reference. The bridge to connect the knowledge
gap between rotor dynamics and rotating structure dynamics is provided in this
book. The introduction of the complex displacement for the analysis of rotor
whirling motion may also be an effective way to facilitate such understanding.
An effective approach to analyzing various phenomena thus benets the
understandable explanation for the correction of rotor vibration problems. The
purpose of the present book is to describe the general mechanisms of resonance and
self-excited vibration by using models that are as simple as possible, thereby
forming a common basis for addressing various vibration problems. The study
of these models will also be useful for enhancing intuitive ability. The authors have
placed special emphasis on the conciseness of the mathematical formulations in the
process of solving vibration problems. The entire content of the book is within the
realm of linear vibration theory, but the authors believe, based on their experiences
on-site and in consulting, that the book provides the sufcient body of knowledge
needed for practical engagement. The book will also prove useful as
test-preparation material for the ISO Machinery Condition Analyst (Vibration), i.e.,
an examination organized by the JSME in Japan.
Although the authors have endeavored to eliminate errors and provide appro-
priate emphasis, critical comments from readers are welcome. We are grateful to the
authors of the literature cited in this book. Special thanks are also due to Prof.
Hiroyuki Fujiwara (National Defense Academy, Japan) and Dr. Naohiko Takahashi
(Hitachi, Ltd.) for reviewing the manuscript and providing valuable comments.
The authors deeply thank Mr. Fumito Shinkawa, President of Shinkawa Electric
Co., Ltd. for his support for our efforts.
We also wish to thank Springer for its support in publishing this work.

Yokosuka, Japan On behalf of the authors


May 2016 Osami Matsushita
Contents

1 Introduction of Rotordynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Vibration Problems in Rotating Machinery . . . . . . . . . . . . . . . . . 1
1.1.1 Varieties of Rotating Machinery . . . . . . . . . . . . . . . . . . 1
1.1.2 Bearings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.3 Defects in Various Elements and Induced
Vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.4 Rotordynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Types of Vibration in Rotating Machinery . . . . . . . . . . . . . . . . . 8
1.3 Classication of Vibration by Mechanism of Occurrence . . . . . . 9
1.4 Simplifying Complicated Phenomena . . . . . . . . . . . . . . . . . . . . . 11
2 Basics for a Single-Degree-of-Freedom Rotor . . . . . . . . . . . . . . . . . . 13
2.1 Free Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.1 Natural Frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Calculation of Spring Constant . . . . . . . . . . . . . . . . . . . 14
2.1.3 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.4 Mass Effects of Spring on Natural Frequency . . . . . . . . 15
2.2 Damped Free Vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2.1 Mass-Spring-Viscous Damped System . . . . . . . . . . . . . 18
2.2.2 Measurement of Damping Ratio . . . . . . . . . . . . . . . . . . 20
2.2.3 Phase Lead/Lag Corresponding to Damping Ratio . . . . 24
2.3 Unbalance Vibration of a Rotating Shaft . . . . . . . . . . . . . . . . . . 25
2.3.1 Complex Displacement and Equation of Motion . . . . . . 25
2.3.2 Complex Amplitude of Unbalance Vibration . . . . . . . . . 26
2.3.3 Resonance Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3.4 Nyquist Plot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.5 Bearing Reaction Force at Resonance . . . . . . . . . . . . . . 30
2.3.6 Transmissibility of Unbalance Vibration to
Foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 32

vii
viii Contents

2.4 Evaluation of Q-Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33


2.4.1 Q-Value Criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.4.2 Measurement of Q-Value by the Half Power Point
Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4.3 Measurement of Q-Value Using a Nyquist Plot . . . . . . . 34
2.4.4 Re-evaluation of Q-Value for Rapid Acceleration . . . . . 36
2.4.5 Vibration in Passing Through a Critical Speed . . . . . . . 39
3 Modal Analysis of Multi-Degree-of-Freedom Systems . . . . . . . . . . . 41
3.1 Equation of Motion for a Multi-dof System . . . . . . . . . . . . . . . . 41
3.1.1 Multiple Mass Systems . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.2 Equation of Motion for a Two-dof System . . . . . . . . . . 42
3.1.3 Equation of Motion for a Multi-dof System . . . . . . . . . 43
3.2 Modal Analysis (Normal Mode Method) . . . . . . . . . . . . . . . . . . 46
3.2.1 Eigenvalue Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2.2 Orthogonality. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2.3 Reduced Order Modal Model . . . . . . . . . . . . . . . . . . . . 47
3.2.4 Vibration Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.3 Modal Analysis of Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.3.1 Natural Frequencies and Eigenmodes . . . . . . . . . . . . . . 53
3.3.2 Correspondence of the Modal Analyses
for Multi-dof Systems and Continua . . . . . . . . . . . . . . . 53
3.3.3 Reduced Modal Models . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.3.4 Modal Eccentricity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.4 Physical Models from Reduced Modal Models. . . . . . . . . . . . . . 60
3.4.1 Modal Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.4.2 Equivalent Mass Method . . . . . . . . . . . . . . . . . . . . . . . . 62
3.5 Approximation of Natural Frequencies . . . . . . . . . . . . . . . . . . . . 63
3.5.1 Rayleighs Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.5.2 Method Using Inuence Coefcients . . . . . . . . . . . . . . . 65
3.5.3 Dunkerleys Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.5.4 Iterative Method (Power Method) [B4] . . . . . . . . . . . . . 68
3.5.5 Stiffness Matrix Method . . . . . . . . . . . . . . . . . . . . . . . . 70
3.5.6 Transfer Matrix Method . . . . . . . . . . . . . . . . . . . . . . . . 73
4 Mode Synthesis and Quasi-modal Method. . . . . . . . . . . . . . . . . . . . . 79
4.1 Mode Synthesis Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1.1 Why Mode Synthesis? . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1.2 Guyan Reduction Method . . . . . . . . . . . . . . . . . . . . . . . 80
4.1.3 Mode Synthesis Models . . . . . . . . . . . . . . . . . . . . . . . . 84
4.2 Quasi-modal Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.2.1 Principle of the Quasi-modal Model . . . . . . . . . . . . . . . 90
4.2.2 Examples of Quasi-modal Models . . . . . . . . . . . . . . . . . 97
4.3 Plant Transfer Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Contents ix

5 Unbalance and Balancing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105


5.1 Unbalance in a Rigid Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.1.1 Static Unbalance and Dynamic Unbalance . . . . . . . . . . 105
5.1.2 Static Unbalance and Couple Unbalance . . . . . . . . . . . . 107
5.1.3 Adverse Effects of Unbalance Vibration . . . . . . . . . . . . 107
5.1.4 Residual Permissible Unbalance in a Rigid
Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 108
5.2 Field Single-Plane Balancing (Modal Balancing) . . . . . . . . .... 111
5.2.1 Relationships among Rotational Pulse,
Unbalance and Vibration Vector . . . . . . . . . . . . . . . . . . 111
5.2.2 Linear Relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.2.3 Identifying the Inuence Coefcient . . . . . . . . . . . . . . . 114
5.2.4 Correction Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.3 Balancing by the Inuence Coefcient Method . . . . . . . . . . . . . 117
5.4 Modal Balancing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5.5 n-Plane Balancing or (n + 2)-Plane Balancing? . . . . . . . . . . . . . 126
5.5.1 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
5.5.2 Number of Correction Planes Needed
for Universal Balancing . . . . . . . . . . . . . . . . . . . . . . . . . 131
5.5.3 What Is the 2 in the (n + 2)-Plane Method? . . . . . . . 132
5.6 Balancing of a Rotor Supported by Magnetic Bearings . . . . . . . 136
5.6.1 Balancing by Feed-Forward (FF) Excitation . . . . . . . . . 136
5.6.2 Case Study: Centrifugal Compressor Supported
by AMBs [VB245] . . . . . . . . . . . . . . . . . . . . . . . . .... 141
5.7 Balancing without Rotational Pulses . . . . . . . . . . . . . . . . . . .... 143
5.7.1 Four Run Method . . . . . . . . . . . . . . . . . . . . . . . . . .... 143
5.7.2 Balancing by Placing a Trial Mass
at a Regular Phase Pitch . . . . . . . . . . . . . . . . . . . . . . . . 146
5.8 Solution of Two-Plane Balancing . . . . . . . . . . . . . . . . . . . . . . . . 147
5.8.1 Principle of Calculation . . . . . . . . . . . . . . . . . . . . . . . . . 147
5.8.2 In-Phase and Out-of-Phase Balancing . . . . . . . . . . . . . . 149
6 Gyroscopic Effect on Rotor Vibrations . . . . . . . . . . . . . . . . . . . . . . . 153
6.1 Rotordynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
6.2 Gyroscopic Moment and the Motion of a Top . . . . . . . . . . . . . . 155
6.2.1 Gyroscopic Moment . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
6.2.2 Equation of Motion of a Top and Whirling
Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
6.3 Natural Vibration of a Rotor System . . . . . . . . . . . . . . . . . . . . . 158
6.3.1 Natural Frequency of Whirling . . . . . . . . . . . . . . . . . . . 158
6.3.2 Inuence of the Gyroscopic Factor . . . . . . . . . . . . . . . . 160
6.3.3 Calculation of the Natural Frequency
of Whirling in Multi-dof Rotor System . . . . ......... 162
x Contents

6.4 Unbalance Vibration and Resonance . . . . . . . . . . . . . . . . . . . . . . 163


6.4.1 Condition for Unbalance Resonance
and Critical Speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
6.4.2 Resonance Curves for Unbalance Vibration . . . . . . . . . 165
6.4.3 Calculation of Critical Speed of a Multi-dof
Rotor System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
6.5 Vibration and Resonance with Base Excitation . . . . . . . . . . . . . . 168
6.5.1 Resonance Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 168
6.5.2 Forced Vibrational Solution for Base Excitation . . . . . . 170
6.5.3 Resonance Curves and Whirling Trajectories . . . . . . . . 172
6.5.4 Case Study: Aseismic Evaluation of a High-Speed
Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
6.6 Ball Passing Vibration and Resonance Due to Ball
Bearing Defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
6.6.1 Ball Bearing Specications . . . . . . . . . . . . . . . . . . . . . . 176
6.6.2 Excitation by a Recess on Outer Race . . . . . . . . . . . . . . 176
6.6.3 Excitation by a Recess on Inner Race . . . . . . . . . . . . . . 177
6.6.4 Resonance Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 178
6.6.5 Case Study: Hard Disk Drive (HDD) [VB218] . . . . . . . 178
7 Approximate Evaluation for Eigenvalues of Rotor-Bearing
Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 181
7.1 Equation of Motion for a Single-Degree-of-Freedom
Rotor System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 181
7.2 Vibration Characteristics of a Symmetrically Supported
Rotor System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
7.2.1 Natural Frequencies of a Conservative System . . . . . . . 184
7.2.2 Effects of Non-conservative System Parameters . . . . . . 185
7.2.3 Parameter Survey. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
7.3 Natural Frequencies of a Rotor Supported by Anisotropic
Bearings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
7.3.1 Natural Frequency of a Conservative System . . . . . . . . 189
7.3.2 Elliptical Whirling of a Conservative System . . . . . . . . 190
7.3.3 Inuence of Gyroscopic Effect . . . . . . . . . . . . . . . . . . . 192
7.3.4 Shape of Elliptical Whirling Orbit . . . . . . . . . . . . . . . . . 193
7.3.5 Effects of Non-conservative Parameters . . . . . . . . . . . . . 195
7.3.6 Parameter Survey. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
7.4 Vibration Characteristics of a Jeffcott Rotor . . . . . . . . . . . . . . . . 199
7.4.1 Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
7.4.2 Vibration Characteristics . . . . . . . . . . . . . . . . . . . . . . . . 200
7.4.3 Real Mode Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
7.4.4 Complex Mode Analysis . . . . . . . . . . . . . . . . . . . . . . . . 204
7.5 Analysis of Characteristics of Unbalance Vibration . . . . . . . . . . 205
7.5.1 Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
Contents xi

7.5.2 Unbalance Vibration of an Isotropically Supported


Rotor System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
7.5.3 Unbalance Vibration of a Rotor Supported
by Anisotropic Bearings . . . . . . . . . . . . . . . . . . . . . . . . 206
7.6 Case Study: Vibrations of a Flexible Rotor
with Cylindrical Bearings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.6.1 Critical Speed Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.6.2 Calculation of Complex Eigenvalues
and Q-Values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
7.6.3 Root Loci . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
7.6.4 Resonance Curves for Unbalance Vibration . . . . . . . . . 211
8 Rotor System Evaluation Using Open-Loop Characteristics . . .... 213
8.1 Open-Loop Analysis of a Single-dof System . . . . . . . . . . . .... 213
8.1.1 Open-Loop Frequency Response
of a Single-dof System . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.1.2 Measurement of Open-Loop Frequency Response . . . . . 221
8.2 Modal Open-Loop Frequency Response . . . . . . . . . . . . . . . . . . . 222
8.2.1 Modal Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
8.2.2 Modal Open-Loop Frequency Response . . . . . . . . . . . . 224
8.3 Open-Loop Frequency Response of a Jeffcott Rotor . . . . . . . . . . 228
8.3.1 Series Coupling and Phase Lead Function . . . . . . . . . . . 228
8.3.2 Open-Loop Frequency Response . . . . . . . . . . . . . . . . . . 229
8.3.3 Gain Cross-Over Frequency and Phase Margin . . . . . . . 230
8.3.4 Precision of Approximate Solutions . . . . . . . . . . . . . . . 232
8.3.5 Optimal Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
8.3.6 Frequency Response . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
9 Bridge Between Inertial and Rotational Coordinate Systems . . .... 241
9.1 Vibration Waveforms (Displacement and Stress Caused
by Strain) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
9.2 Natural Frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
9.3 Resonance Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
9.4 Representation of Equation of Motion . . . . . . . . . . . . . . . . . . . . 246
9.4.1 Gyroscopic Moment and Coriolis Force . . . . . . . . . . . . 246
9.4.2 Case Study: Multi-blade Fan
(Sirocco Fan) [VB55] . . . . . . . . . . . . . . . . . . . . . . .... 248
10 Vibration Analysis of Blade and Impeller Systems . . . . . . . . . . . . . . 253
10.1 Natural Frequencies of Rotating Structure Systems. . . . . . . . . . . 253
10.1.1 Natural Frequencies of a Thin Disk . . . . . . . . . . . . . . . . 253
10.1.2 Natural Frequencies of Blades . . . . . . . . . . . . . . . . . . . . 257
10.1.3 Vibration Analysis of Cyclic Symmetry Structural
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
10.1.4 General Vibration Analysis of Blades and Impellers
in a Rotational Coordinate System . . . . . . . . . . . . . . . . 266
xii Contents

10.2 Vibration and Resonance of Blades and Impellers . . . . . . . .... 268


10.2.1 Conditions for Blade-Shaft Coupled Vibration . . . .... 268
10.2.2 Natural Vibration Modes of Blades
and Blade Wheels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
10.2.3 External Forces Acting on Blades and Impellers . . . . . . 269
10.2.4 Resonance Conditions of Blades . . . . . . . . . . . . . . . . . . 270
10.2.5 Criterion of Blade Resonance: Campbell Diagram . . . . 273
10.2.6 Case Study: Resonance in Impeller Blades of
Centrifugal Compressor [VB958] . . . . . . . . . . . . . .... 278
10.3 Blade/Impeller Vibrations Excited at Stationary Side . . . . . .... 281
10.3.1 Difference in Excitation Methods
and Resonance Conditions . . . . . . . . . . . . . . . . . . . .... 281
10.3.2 Representation of Vibration of Blades
and Impellers in an Inertial Coordinate System . . . .... 281
10.3.3 Resonance Condition 1 . . . . . . . . . . . . . . . . . . . . . .... 283
10.3.4 Resonance Condition 2 . . . . . . . . . . . . . . . . . . . . . .... 285
11 Stability Problems in Rotor Systems . . . . . . . . . . . . . . . . . . . . . .... 287
11.1 Unstable Vibration Due to Internal Damping
of a Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
11.1.1 Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
11.1.2 Stability Condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
11.1.3 Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
11.2 Unstable Vibration of an Asymmetric Rotor System . . . . . . . . . 293
11.2.1 Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
11.2.2 Overview of Vibration in an Asymmetric
Rotating Shaft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
11.2.3 Simulation of Vibration of Asymmetric Rotor . . . . . . . . 303
11.3 Vibration Due to Thermal-Bow by Contact Friction . . . . . . . . . . 306
11.3.1 Thermal-Bow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
11.3.2 Thermal-Bow Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
11.3.3 Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
11.3.4 Physical Interpretation of Stability . . . . . . . . . . . . . . . . . 310
11.3.5 Simulation of Thermal-Bow Induced Vibration . . . . . . . 312
11.4 Thermal-Bow Induced Vibration of an Active Magnetic
Bearing Equipped Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
11.4.1 Thermal-Bow Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
11.4.2 Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
11.4.3 Physical Interpretation of Stability . . . . . . . . . . . . . . . . . 317
11.4.4 Simulation of Thermal Bow Induced Vibrations . . . . . . 318
12 Rotor Vibration Analysis Program: MyROT . . . . . . . . . . . . . . . . . . 321
12.1 Data on Rotor Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
12.1.1 Rotor Drawing and Discretization . . . . . . . . . . . . . . . . . 321
12.1.2 Data Organization of a Rotor System . . . . . . . . . . . . . . 323
Contents xiii

12.2 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326


12.2.1 Matrices in the Original System . . . . . . . . . . . . . . . . . . 326
12.2.2 Reduced Matrices in the Guyan Method . . . . . . . . . . . . 326
12.2.3 Matrices in the Mode Synthesis Models . . . . . . . . . . . . 329
12.2.4 Discretization of Beam Elements . . . . . . . . . . . . . . . . . . 331
12.3 Analyses Corresponding to Job Commands . . . . . . . . . . . . . . . . 332
12.3.1 Analysis Menu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
12.3.2 Analysis Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
12.3.3 EDIT Screen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
Appendices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
Chapter 1
Introduction of Rotordynamics

Abstract This book explains various phenomena and mechanisms that induce
vibrations in rotating machinery, based on theory and eld experiences (Chaps.
112 in volume 1). It also provides guidance in undertaking diagnosis and
implementing effective countermeasures against various vibration problems in the
eld (in volume 2). Consequently, volume 1 is intended mainly for beginners and
students, while volume 2 is mainly for design engineers and practitioners. This
chapter of volume 1 emphasizes the subtlety of vibration problems in rotating
machinery and the importance of reliable technologies that help to stabilize and
reduce vibrations. It also outlines a wide variety of rotating machinery, vibration
problems found in the eld, and mathematical approaches to analyze vibration
problems. In high-speed rotating machinery, the steady rotating state corresponds to
a stationary equilibrium condition with a high rotational energy. Vibration brings
the machine into a dynamic state. If the rotating system becomes unstable in the
dynamic state, resulting in self-excited vibration, the machine enters a very dan-
gerous operational condition. Since the energy source of self-excited vibration in a
rotating system is provided by the spin of the rotor, the only way to avoid this
dangerous situation is to stop the energy source, for example, by shutting down the
power source in the case of a motor driven system. Vibrations caused by an external
force, unless kept small enough, may also lead to a serious problems through
contact between the rotor and the stationary part (stator).

 
Keywords Rotor Bearing Mechanism of vibration  Rotor modeling  Rotor

vibration simulation m BASE

1.1 Vibration Problems in Rotating Machinery

1.1.1 Varieties of Rotating Machinery

A rotating machine consists of a rotor (comprising a shaft and disks), and bearings
with associated casings/housings that support the rotor. Fluid performance based
machines used in the energy and process industries, conventionally called

Springer Japan 2017 1


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_1
2 1 Introduction of Rotordynamics

turbomachines, are well-known rotating machines. Examples of turbomachines


include: steam turbine-generator sets (Fig. 1.1), gas turbines (Fig. 1.2), jet engines,
waterwheels, compressors, fans, pumps, centrifuges, electric motors, and step-up/
reduction gear units.

(a) LP turbine rotor (Ref. Toshiba Review)

ALP BLP
HP IP
GEN

(b) Rotor line

Fig. 1.1 Steam turbine-generator set

Fig. 1.2 Gas turbine


1.1 Vibration Problems in Rotating Machinery 3

While these are large-sized high-power machines that rotate with high energy
levels, there are also small-sized low-power machines for industrial or home
applications, such as electric motors, gyroscopes, agitators, grinding machines,
spinning machines, marine and automotive power trains, automotive internal
combustion engines, reciprocating machines, lathes, vacuum pumps, fans, washing
machines (Fig. 1.3), compressors for air conditioners (Fig. 1.4) and refrigerators.

Fig. 1.3 Washing machine

Fig. 1.4 Air conditioner compressor


4 1 Introduction of Rotordynamics

Fig. 1.5 Current type of hard


disk (3.5 in.)

The increase of the fluid performance of turbomachines requiring drastically


high pressure and/or extremely large capacity is commonly set as a goal for
engineering projects, which challenges the development of new turborotors. The
turbomachines characterizing these high specications are planed from the view-
points of fluid performance and fluid dynamics with less consideration of
mechanical vibrations. We note that the smooth rotation test with low vibration
level is absolutely necessary before the fluid performance test at the rated speed
operation is executed, otherwise, large vibration at rated operation might negate the
fluid performance test. For example, of the 12 case studies concerning new chal-
lenging developments of rotors such as compressors, pumps, etc., introduced in the
October 2009 issue of the Japanese journal Turbomachinery, more than half were
focused on how to solve vibration problems in rotating machinery, which required
much effort to attenuate vibrations before the fluid performance test.
In addition to conventional machines, recent developments in mechatronics have
promoted new rotational machinery, particularly for computer-related devices and
devices in the eld of information technology, such as drives for hard disks
(Fig. 1.5), optical disks and compact disks, video tape recorders, polygonal mirrors
and hybrid engines. They are progressing rapidly toward a smaller size, lower
thicknesses, higher speeds or higher density information. Most of them are char-
acterized by thin rotating disks where the gyroscopic effect is often important.
Though these innovative products are often specied to have a short life cycle,
dynamical design for vibration reduction will invariably be required from engineers.

1.1.2 Bearings

A bearing is located on the stationary part to support the rotor and to ensure its
smooth rotation. The rotating shaft in the bearing portion is called the journal (for
radial bearings) or disk (for thrust bearings). Bearings may be classied in a number
of varieties, including rolling element bearings (ball bearings, roller bearings),
sliding bearings (cylindrical bearings, multi-lobe bearings, tilting pad bearings,
etc.), and magnetic bearings (with actively controlled electromagnet or
passive/non-controlled permanent magnet), according to applications. Their char-
acteristics [1] are summarized in Table 1.1.
1.1 Vibration Problems in Rotating Machinery 5

Table 1.1 Comparison of various types of bearings

Rolling element bearing Sliding bearing Active magnetic bearing

Inferior for impact load. Suitable for impact load and Most suitable for light load
A deep groove / angular heavy load. with high speed.
Load contact and cone roller Acceptable specific loads are Acceptable load pressures are
bearing, etc., can support the approximately: approximately:
load both in radial and thrust Radial dir. : less than 5 MPa Radial dir. : 0.3 to 0.5 MPa
directions. Thrust dir. : less than 7 MPa Thrust dir. : less than 0.8 MPa

Static friction coefficient is Static friction coefficient is Small


Friction as small as 10 -3 ~10 -2 as large as 10 -2 ~10 -1
Dynamic friction coefficient is almost the same as 10 -3

Depending on centrifugal Depending on turbulent flow Depending on strength of


Speed limit force and lubrication, etc. transition and overheating AMB rotor material. Generally,
of oil film. Generally,
5
DN < 210 mm rpm V < 120 m/s V < 200 m/s

Stiffness and Large stiffness and no Large stiffness and high Stiffness and damping are low,
damping damping. damping. but widely controllable.

Noise Comparatively large Comparatively small Small


Lubricant Grease in general. Oil in general. Not required.
Life can be estimated using Infinite life in hydrodynamic Nearly permanent.
Life and fatigue strength of material. operation. Seizure and wear
breakage Seizure breakage may occur are main causes of breakage.
at high-speed rotation. Flaking may occur due to high
load.
Installation Comparatively sensitive. Comparatively insensitive. Insensitive because bearing
error clearance is large.

Influence Influential on life, wear, Comparatively less influential. Less influential.


contamination especially noise.

By using grease / oil Leakage from lubricant oil Maintenance free in general,
Maintenance lubrication, maintenance is circulating system need to be except some electronics parts.
easy. checked and stopped.

Mass-produced standard Generally in-house production. Expensive because


Cost bearing are inexpensive and Comparatively inexpensive. custom-made manufacture
interchangeable. Arbitrary bearing dimensions. is still the mainstream.

The reaction force on a bearing is often represented by spring (stiffness) and


damper elements in a mechanical model using linear vibration characteristics. In
these terms, the reaction force is generally characterized as follows for each type of
radial bearing:
(1) Rolling bearings: The reaction force is isotropic (radially symmetric). Stiffness
is relatively high, but damping is negligible or nil.
(2) Sliding bearings: The reaction force is anisotropic (radially asymmetric)
because the rotor weight is supported by an oil lm in the bearing gap at a
slightly eccentric position with respect to the center axis of the bearing. Since
the oil lm also dissipates vibrational energy, resulting in vibration reduction,
6 1 Introduction of Rotordynamics

a well-damped rotor system can be completed if the design is effective. Since


stiffness and damping are very high in this type of bearing, they are thus
required to be predicted in high-precision calculations.
(3) Magnetic bearings: The reaction force is isotropic if the journal is tuned to
coincide with the bearing having an radially symmetric electromagnetic force.
Both stiffness and damping are controllable over usable ranges. Errors in
control may be compensated by adjusting control parameters in situ.
While the reaction force is nonlinear in general, a linear approximation (i.e.,
modeled by spring and damper elements) around the static equilibrium point may
be used for design analysis. A small nonlinearity may, however, be evident in a
normal vibration waveform, which can grow into a large amplitude waveform
featuring nonlinear sources.

1.1.3 Defects in Various Elements and Induced Vibration

A rotating machine comprises a variety of elements such as shafts, disks, blades,


nozzles, drums, joints, bearings, seals, coils and gears, each of which undergoes
fluid forces from the process working gas/liquid or lubrication, electromagnetic
forces and so on. Therefore, a rotor is practically never free from the possibility of
generating vibration during its operation.
Rotor vibration causes various other problems in rotating machinery, which may
make the situation worse. Increased cyclic stress and fatigue failure, collision of the
rotor with stationary parts, seized bearings, vibrating force transmission to sta-
tionary parts, and induced vibration of peripheral units, are examples of defects
caused by vibration of the rotor.
To achieve vibration reduction of a rotor, the design of rotating machinery
should take analysis of such factors into account to avoid vibration under normal
operating conditions. When vibration occurs in a machine, it is absolutely necessary
to nd its cause and take appropriate countermeasures. It is a prerequisite for
engineers, therefore, to understand dynamic characteristics of machine elements,
particularly with regard to possible causes for induced vibration of the rotor.
Figure 1.6 shows an example of elemental technical issues in the development of
a large-scale power generator. Many factors, including large sliding bearings, seals
and slot structures to hold the coil, are among important issues for avoiding rotor
vibration.

1.1.4 Rotordynamics

The purpose of rotordynamics is to establish methodologies of vibration


reduction/suppression. This enables design, appropriate operation and maintenance
procedures of rotating machinery by elucidating causes of vibration from the
1.1 Vibration Problems in Rotating Machinery 7

High strength High strength Water cooled Slidable coil


end coil material stator coil end support

Optimal L/D for


rotor vibration

Optimal cross-section
for coil capacity
23 inch large
size bearing

Half-split type Iron loss reduction Compact High strength


seal ring by flux shunt frame terminal box

Fig. 1.6 Technical issues for 1000 MW generator development (Ref. Toshiba review)

viewpoints of excitation force and mechanisms of excitation, natural frequencies of


modes excited by external influences on the rotor.
During 19501970, many Japanese manufacturers of rotating machinery learned
know-how for solving individual vibration problems from their technological
partners in Europe and the U.S. Increases in size and speed of rotating machinery
since the 1960s led to development and accumulation of original technology during
19701980.
Since then Japanese industry has relied largely on its own technology, and now it
is ranked among the highest in the eld of rotordynamics.
Many cases of vibration in rotating machinery were thought to be simple
troubles and remained obscure, even within departments of the same company.
However, in April 1972 it was a surprise when the Journal of the JSME (Japan
Society of Mechanical Engineers) published papers from Dr. Kazuhiro Shiraki [2]
and his group, which elucidated 132 case histories on vibration problems with root
cause analyses, countermeasures and their effects. Subsequently, successor engi-
neers realized the need for openness as a means to overcome vibration problems
experienced in industry. They established a Vibration Engineering Database
Committee [3] (m BASE) in the JSME in 1991, which has been organizing yearly
meetings for collecting and discussing vibration case studies, and compiling them
into database book/CD (Fig. 1.7). It is highly desirable for engineers to share
solutions of vibration problems and lessons learned from them. The idea for
troubleshooting came from the fundamental understanding of eld data that enables
them to rectify vibration phenomena, rather than from fragmentary or supercial
knowledge learned in the classroom.
Some of the m BASE data have been presented at conferences [46]. Recently,
many m BASE data related to rotordynamics are planned to be part of a Website of
the JSME.
8 1 Introduction of Rotordynamics

Fig. 1.7 m_BASE databook


[3]

1st edition
1994
2nd edition
CD-ROM
2002
3rd edition
CD-ROM
2011

1.2 Types of Vibration in Rotating Machinery

From a practical point of view, frequently encountered rotor vibration modes can be
categorized as follows:
(1) Bending vibration of the shaft
The shaft whirls in its rotation plane (the plane is perpendicular to the rotational
axis) while maintaining its deflection mode that varies along the axis. The whirling
shaft mode is measured as being rigid mode and/or flexible in nature. Most
vibration problems of the shaft can be described by these bending modes.
(2) Torsional vibration of the shaft
Torsional deformation arises when the vibrations of each part of the shaft are
twisted along the rotation axis. The occurrence of the torsional vibration problem is
relatively small compared with bending vibration in rotating machinery, excluding
reciprocating and/or geared machines. Vibrational changes in electromagnetic
forces in motors or in the torque of the load may also induce torsional vibrations.
(3) Longitudinal vibration of the shaft
This mode of vibration usually does not appear due to the high longitudinal
stiffness of the shaft, but there have been cases due to a collision with a static part
due to local resonance, and an excessive dynamic load in a thrust bearing.
(4) Vibration of rotating structure
Bending vibration of a rotating structure, such as turbine blades or pump and
compressor impellors, is a very important issue. Three-dimensional nite element
analysis is commonly used to assess vibrations in this category using detailed cal-
culation of natural frequencies and prediction of resonance amplitudes based on the
actual geometry of specic blades and impellers. The stiffness increase under cen-
trifugal force is taken into account. The rotating structure is usually analyzed by
assuming that the center of the structure is xed, being free from the shaft movement.
The coupling effect of shaft-blade vibration must be considered in some cases.
1.3 Classication of Vibration by Mechanism of Occurrence 9

1.3 Classication of Vibration by Mechanism


of Occurrence

Figure 1.8 illustrates the mechanisms of vibration occurrence from the viewpoint of
the equation of motion and the associated parameters. The terms in this gure are
explained below:
(1) Equation of motion The equation of motion describes a rotor system, where
M (mass), C (damping) and K (stiffness) are linear matrices and ef is a small
nonlinear term. These terms constitute a free vibration system. For forced
vibration, there is a term for unbalance force U, which is generated at a
frequency synchronous to the rotational speed X. External harmonic excitation
F(mt) with an exciting frequency m may also be included.
(2) Forced vibration system The spinning rotor under an unbalance force U or a
periodic external force F vibrates steadily at a frequency synchronous with the
external force. When the frequency of the external force coincides with the
natural frequency of the rotor, resonance occurs and the vibration amplitude
becomes maximal.
The lower half of the gure shows details of the forced vibration. The resonance of
the unbalance vibration synchronous with the rotational speed occurs at a rotational
speed X  xn, which is called the critical speed of the rotor because it represents a
potentially dangerous operating point. Resonance under harmonic excitation with an
external frequency m occurs at m = xn. This type of resonance should be prevented by
avoiding the resonance condition or reducing the sensitivity to resonance.
Resonance due to the existence of the nonlinear term at Nm = xn (where N is a
multiple or fraction) is dependent on the right hand side term ef.

1
Stable Q-factor =
2
Free vibration Unstable Self-excited vibration

n = K
= eigen frequency Limit Adding damping
M
cycle is effective
= C
= damping ratio
2 MK

System description for vibration analysis


.. . .
MX + CX + KX = U 2 cos ( t ) + F (vt ) + f ( X, X, vt )
Resonance Nonlinear resonance N v = n
= n N = 2,3 Ultra-harmonic
Q=
1 v = n Forced
2 N = 1/2,1/3 Sub-harmonic Resonance detuning
Forced vibration self-excited is effective
vibration
Structure resonance : External force frequency v = n
Unbalance resonance : Rotational speed = n

Fig. 1.8 Vibration mechanisms and parameters


10 1 Introduction of Rotordynamics

(3) Free vibration system The upper half of Fig. 1.8 deals with free vibrations.
Vibration is attenuated in a stable system, while self-excited vibration appears
in an unstable system. Damping is positive in the former, and negative in the
latter. The frequency of the resulting vibration is at a natural frequency of the
rotor system in both cases.
An example of self-excited vibration is the gradual growth of amplitude even if
no external force is applied. It is well known that a window shade/blind may vibrate
frequently under certain conditions of wind velocity. Once the self-excited vibration
happens, it is difcult to prevent it. Note that the only essential countermeasure is to
provide the system with a positive damping factor. This solution is generally dif-
ferent from that to reduce resonance in forced vibrations.
The amplitude of the self-excited vibration that grows in an unstable linear
system is limited by the effect of the nonlinear term, resulting in a limit cycle as
noted in the gure. In terms of dynamics, this state of negative damping corre-
sponds to an undesirable cyclic state where the restoring force (output) acts with a
certain delay against the rotor motion (input). This delay between the input and
output in the mechanical dynamics is sometimes qualied as being in a negative
spiral.
(4) Parametrically excited oscillation This is a type of self-excited vibration.
The vibration amplitude may grow under certain conditions with triggered by
slight changes of the parameters in time. The equation of motion includes time
dependent mass, spring and damping parameters in a rotating system.
(5) Nonlinear vibration Vibrations caused by mechanisms involving nonlinear
effects/elements (shown on the upper right and lower right of the gure) is also
important. Typical examples include rubbing vibration between a rotor and a
stator, and the limit cycle of oil whip appearing after the onset of self-excited
vibration.
Figure 1.9 shows vibration phenomena and their causes classied according to
the categories described previously. The variety of problems indicates the reality
that the design optimization of actual machines is often difcult.
Wording related to mechanical vibration is summarized in ISO 2041 [7], which
is denitely a prerequisite for mechanical engineering experts.

( Phenomena ) ( Causes )
Forced vib. Unbalance vib. Residual unbalance, Thermal unbalance

Resonance Critical speed, Blade resonance, Foundation resonance, Torsional resonance

Flow induced vib. Rotating stall, Blade passing freq.

Others Motor electromagnetic force, Gear mesh excitation

Free vib. Oil-whip Sliding bearing, Oil-film characteristics


( unstable ) Flow induced instability Labyrinth seal, Oil-film bearing, Impeller seal

Internal friction whirl Shrink-fit

Fig. 1.9 Typical subjects to develop new compressors (note vib. = vibration, freq. = frequency)
1.4 Simplifying Complicated Phenomena 11

1.4 Simplifying Complicated Phenomena

Since a rotating machine in a steady state has a high rotational energy, even small
and localized problems tend to be amplied and lead to signicant vibration
problems for the entire system. It has been reported that a cupful of condensed
water caused vibration of a whole turbine-generator set, and an elastic rubber
coupling introduced for ease of centering rotating shafts may generate difcult
vibrations in coupled machines.
However, it is true that most of these complicated vibration problems can be
described by simple mechanical models, which permit identication of the causes.
Appropriate countermeasures can actually eliminate vibration effectively. Engineers
dealing with rotating machinery should have the ability to construct models that
explain the mechanisms of how the resonance occurs or the self-excited vibration is
generated, along with knowledge, experience and intuition.
Figure 1.10 illustrates an engineer contemplating various factors, such as
unbalance and fluid force, causing vibration in a machine with use of a
single-degree-of-freedom system. He/she tries to describe a story for solving the

Vibration mode Causes Themal


Rubbing
deformation

Unbalance Fluid
m* force
Internal Bearing
Calculation k* damping
Natural frequency damping
amplitude

waveforms 0 1 2 speed
1X
frequency

FFT spectrum
speed
Campbell diagram

Fig. 1.10 Cause identication using simple model with measurement data
12 1 Introduction of Rotordynamics

4th

[10 m p-p] Frequency [Hz]


3rd

2nd
1st

Amplitude
Rotating machine
Eigen mode
( simulation ) Speed [rpm]
Campbell diagram

rotor
Mathematical
MyROT solution
Brg Brg FE mode
( Bearing ) model synthesis
model
Beam element model

Fig. 1.11 Full model simulation using 1D-FE Method

problem by using a simple vibrating system in combination with knowledge,


experience, other advice, and, of course, referring to our v_BASE database bank.
Simulation is another important technique in designing a rotor. Large-sized
rotating machinery requires particularly detailed design calculations. Figure 1.11
shows an example of the modeling of a rotating shaft system with multiple disks by
beam elements. This is a nite element method (1D-FEM) together with a vibration
analysis program taking the dynamic stiffness of the bearings into account. The
authors have developed MyROT (See Chap. 12), a vibration analysis program for
rotating shaft systems based on 1D-FEM. Today, 1D-Finite element analysis is a
standard tool in rotor design for calculations rather than 3D-FEM. 1D is enough as
we do not need to employ a steam-hammer to crack a nut?
The 1D-FEM simulation is also a useful means to quantify the recommended
countermeasures against vibration once the reason has been identied qualitatively
by simple modeling. Results of simulations and countermeasures should be accu-
mulated in a database as intellectual property for the future reference. Nevertheless
it is said that vibration problems will not disappear forever, since no engineer can
predetermine every cause of vibration when during the process of challenging
machine design. We hope that the next generation never makes the same mistakes,
at least in the eld of mechanical vibration, especially rotordynamics.
Paper [8] is recommended for readers to review the state-of-art vibration tech-
nology in Japan.
Chapter 2
Basics for a Single-Degree-of-Freedom
Rotor

Abstract This chapter species the denitions, calculation and measurement of


basic vibration properties: natural frequency, modal damping, resonance and
Q-value (Q-factor).
Basic properties featuring in a vibrating system, which are obtained from the free
vibration waveform, are:
Natural frequency fn [Hz], or natural angular (or, circular) frequency xn =2p
fn[rad/s]
Damping ratio f [], or logarithmic decrement d = 2p f []
Using these parameters, the resonance caused by forced excitation can be pre-
dicted with
Resonance frequency (critical speed in unbalanced vibration) = natural fre-
quency fn [Hz]
Resonance sensitivity Q = 1/(2f ) []
Since separation of resonance or reduction of the Q-value are fundamental
requirements in the vibration design of rotating machinery, the placements of a
natural frequency and the damping ratio are very important design indices.


Keywords Single-dof Natural frequency  Damping ratio  Equivalent mass 
 
Bode plot Nyquest plot Q-value

2.1 Free Vibrations

2.1.1 Natural Frequency

Undamped free vibration in a single-dof (degree-of-freedom) system consisting of a


mass m [kg] and a spring constant k [N/m] is featured by the natural angular
(circular) frequency xn , given by
p
xn k=m rad=s 2:1

Springer Japan 2017 13


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_2
14 2 Basics for a Single-Degree-of-Freedom Rotor

Fig. 2.1 (Undamped) Free v0


x
vibration wave form
period
T [s]
k x x0

t
m

mg ( = 0 ) fn =
1
[Hz]
T
..
mx + kx = 0
..
x + n2 x = 0 n = 2 fn [rad/s]

which can be converted to the natural frequency by

fn xn =2p Hz 2:2

The term natural frequency will be also used hereafter for the accurate term
natural circular frequency according to the convention in industry (Fig. 2.1).

2.1.2 Calculation of Spring Constant

The spring constant k [N/m] is the reciprocal of deflection per unit load, and is
determined from a static deflection calculation involving the strength of the material.
Several formulae [9] to obtain spring constants are summarized in Table 2.1.

Table 2.1 Example of spring constant [9]

shaft system shaft system

(1) cantilever P (3) thrust of bar


EI EA
3EI ka = P
k= 3 l
l l
(a) circular cross-section
(a) (b)
(4) simple supported beam at both sides
I = d 4 64 P
h
d

b 3 EIl
(b) rectangular cross-section k= EI
l12 l 22
I = bh 3 12
l1 = l 2 = l 2 cantilever l1 l2
(2) torsion of bar 48 EI l
GJ T
GJ k=
kt = l3
l
d4 (5) overhang P
(a) J =
d 32
3EI
k=
b 1 b b4 a2 (l + a ) l a
J = 0.2 ) ab 3
a 1 12 a 4
(b) (
a 3
2.1 Free Vibrations 15

x (t)
k
a /2 3 / 2
x (t) = a cos n t
m = nt
0 2
v (t) = a n sin n t
1 1
T = m v = m ( a n sin n t ) 2
2
Vmax Tmax E 2 2
1 1
V = kx2= k ( a cos n t ) 2
1 2 T T 2 2
ka
2 V V V 1 1
T + V = k a 2 = m a 2 n2
2 2

Fig. 2.2 The conservation of energy

2.1.3 Conservation of Energy

A system of moving objects is called conservative if the sum E of the kinetic energy
T and potential energy (strain energy) V of the objects remains constant, i.e.,
T + V = constant. Figure 2.2 shows that the kinetic energy and potential energy in
a conservative system is complementary and their sum is thus constant. The
maximum kinetic energy Tmax is equal to the maximum potential energy Vmax:

Tmax Vmax 2:3

This relationship determines the natural circular frequency of the system:


p
2Tmax maxn 2 ; 2Vmax ka2 ! xn k=m 2:4

Since free vibration is a movement with respect to a static equilibrium point and
may be termed as dynamic behavior, the energy changes are represented as the
deviations from the energy at that point. The equilibrium point in Fig. 2.2 is the
point where gravity balances the reaction of the spring via the expansion of the
spring. Therefore, the waveform of the free vibration under the gravity would
remain the same even under zero gravity.

2.1.4 Mass Effects of Spring on Natural Frequency

Calculation of natural frequency is straightforward under an ideal condition of a


massless spring. In actual cases, however, the natural frequency is lower than the
ideal value due to the added mass of the spring. Ignoring this effect in design is
dangerous because it often results in optimistic solutions having higher natural
frequency than in reality.
16 2 Basics for a Single-Degree-of-Freedom Rotor

Fig. 2.3 Example of spring k


mass effect m
1

x
= y y
l
y

In the mass m and spring k system shown in Fig. 2.3, let ms be the mass of the
coil spring and y the displacement of the tip of the spring. The displacement d of
any part of the spring can be represented by the linear equation

d x=ly  ny 2:5

The kinetic energy of the system is the sum of that for the mass at the tip and the
distributed mass of the spring mass with a line density (mass per unit length) ql :
0 1
Z1 Z1
m 1 m ms 1 ms  2
T y_ 2 q1 d_ 2 dx y_ 2 @ n2 dnA m y_ 2:6
2 2 2 2 2 3
0 0

Therefore, the formula for natural angular frequency xn becomes


s
k
xn 2:7
m ms =3

Thus, one third of the spring mass is added to reduce the natural frequency.
Example 2.1 Figure 2.4 shows several examples of added mass effect, when
considering the spring stiffness of a uniform bar. Conrm the factor as the added
mass of spring for each case.
  
Note: (a) dn n2 3  n2y, (b) dn ny, (c) dn n 3  4n2
h i
1  2n3 Un  12y, (d) dn 4 3  4nn2  1  2n3 Un  12 y,
where the step function is U(t) = 0 for t < 0 and U(t) = 1 for t  0.
Example 2.2 Figure 2.5 shows a cantilever, l in length and ms in mass. Find the
equivalent mass meq of the point located at a distance al from the supporting point
for a = {1, 0.9, 0.8, 0.7}.
2.1 Free Vibrations 17

(a) l
(b) l

ms 0.24 Ips 1 / 3
m Ip

3EI y GJ
bending k= torsional kt =
l3 l

(c) ms 0.49 (d) ms 0.37

m y m y
48EI 192 EI
k = k =
bending l3 bending l3

Fig. 2.4 Added mass of spring (equivalent mass)

Fig. 2.5 Deflection curve y


l
al

ms = Al

Answer
If the displacement is unity at the point al, the deflection curve d is dened:

n2 3a  n n  a3  x
d Un  a n 2:8
2a3 2a3 l

The deflection curve is third order up to the point al and linear beyond it. The
equivalent mass meq is given by

Z1
meq ms d2 dn 2:9
0

Thus meq = {0.25, 0.33, 0.47, 0.70}ms for the conditions given.
18 2 Basics for a Single-Degree-of-Freedom Rotor

2.2 Damped Free Vibration

2.2.1 Mass-Spring-Viscous Damped System

If a viscous damping coefcient c [Nm/s] is added to a mass-spring system as


shown in Fig. 2.6, the system becomes non-conservative and its vibration is
observed as a damped vibration waveform. For example, the impulse response is
given by

xt a efxn t sin qt 2:10

The damping is evaluated by the damping ratio f, dened as the viscous


damping coefcient c [Ns/m] divided by the critical viscous damping coefcient
p
cc 2 mk [Ns/m]:
c c c
f p  2:11
cc 2 mk 2mxn

Figure 2.7 compares damped free vibration waveforms for different damping
ratios.

Fig. 2.6 Damped vibration a0 Example of = 0.05


system a1
a2
x a3 a
4 a
5 a6 a7 a
8
m 1 2 3 4 5 [s]

k c 8T

T = 2 / q 2 / n
.. . .. .
mx + cx + kx = 0 x + 2 n x + n2 x = 0

Fig. 2.7 Damped vibration = 0.05


waveform in impulse test 2
0.1
0.3
1.5 0.5
mq x ( t )

1
1.0
0.5 0.7

0 1 2 3 4 5
2 n [s]
2.2 Damped Free Vibration 19

The frequency of the damped free vibration is the damped natural angular fre-
quency q given by
q
q xn 1  f2 2:12

which is a value close to the undamped natural angular frequency xn when the
damping ratio is not too large.
If Eq. (2.10) is rewritten using the characteristic root (complex eigenvalue) k as

xt a Imekt  2:13

the characteristic root is given by solving the characteristic equation


k2 2fxn k x2n 0:
q
k  a  jq fxn  jxn 1  f2 2:14

In design practice, the characteristic root k a jq is rst obtained by complex


eigenvalue analysis of a multi-degree-of-freedom system representing the rotor
system. The vibration characteristics (natural circular frequency xn and damping
ratio f) are then calculated from the characteristic root:
p
xn a2 q2 jkj
2:15
f a=jkj

The system is stable when the real part of the complex eigenvalue a fxn is
negative, and unstable if it is positive. A waveform for each case is shown in
Fig. 2.8.

(a) Stable (b) Unstable


x v0 x

e n t
x0 e n t x (t )

0 t 0 t

e n t
= jq = jq e n t
( < 0) ( > 0)

Example of = 0.1 Example of = 0.1

Fig. 2.8 Stable/unstable system


20 2 Basics for a Single-Degree-of-Freedom Rotor

2.2.2 Measurement of Damping Ratio

The damping ratio f is an important parameter (dimensionless) in a vibrating system


as well as the natural circular frequency xn . It can be estimated by observing the
damped free vibration waveform. The amplitude envelope ~ A of damped vibration
(2.10) is represented by

~
At efxn t 2:16

Since the frequency of the vibration is q and the period 2p=q, as indicated in
Fig. 2.6,

~
A0 1; ~
A2p=q efxn 2p=q 2:17

The ratio of the two terms is

2pf
p
e 1f
2
~
A0=~
A2p=q efxn 2p=q 2:18

This means that the ratio of amplitudes in consecutive cycles is a function of


damping ratio f only. The latter can therefore be identied by calculating the
former.
The series of peak heights a1, a2, a3, , measured for consecutive cycles as
shown in Fig. 2.6, is geometric:

2pf
p
a0 a1 a2 an 1  f2
1\     e 2:19
a1 a2 a3 an 1

The damping ratio is obtained by taking the natural logarithms of each term,
considering that is generally small enough to be approximated as follows::

an 2pf
Logarithmic decrement : d ln p  2pf 2:20
an 1 1  f2

d 1 an
Damping ratio : f ln 2:21
2p 2p an 1

In practice, a semi-logarithmic plot is constructed for the series of peak heights ai


measured for consecutive cycles as shown in Fig. 2.9a, and the damping ratio f is
calculated from the gradient ma of the straight line tted to the plot as follows:
2.2 Damped Free Vibration 21

(a) 5
100 2
1.36
50 bn 3 mb = = 0.272 = 0.003 3
5
9

1.36
1
peak amplitude ai

10 an

log ai
1.17
5
1.25
yn
1 0
4 1 m a = 1.17 = 0.129 = 0.047 2
0.5 9
1.25
2 my = = 0.313 = 0.23
4
0.1 -1
0 1 2 3 4 5 6 7 8 9
measured number
Log Graph
(b)
40
bi 1
3 hb = = 1.82 = ln hb = 0.0032
b i+1 30 2

30
yi bn
ai , y i , bi

2 hy =
y i+1 = 2.054
2
20 = ln h y = 0.23
2 an
yn
a
10 1 h a = a i = 1.345
i+1
1
= ln h a = 0.045
2
0
0 5 10 15 20
a i+1 , y i+1 , b i+1

Ratio Graph

Fig. 2.9 Data plot

logan =an 1 ma
f 2:22
2p log e 2:73

where ma is the gradient for a cycle. Note that the vertical axis log ai is indicated on
the right ordinate of Fig. 2.9a, hence

log a0  log an
ma log an  log an 1
n

In Fig. 2.9b, another description of damped waveforms is drawn, where the


ordinate axis indicates ai and the abscissa axis ai+1. Since the slope is h  an =an 1 ,
the damping ratio f is then given by
22 2 Basics for a Single-Degree-of-Freedom Rotor

d 1 an 1
f ln ln h 2:23
2p 2p an 1 2p

To improve the accuracy, the average value of the slope is recommended, which is
known by the straight-line tting. Note that it is not necessary for this straight line to
pass through the origin, because the offset from the origin depends on the friction effect.
Example 2.3 Find the damping ratio by measuring the peak amplitudes in Fig. 2.6
with a ruler.
Answer
The peak amplitude readings for each cycle are: 21.5, 16, 11.5, 8.5, , which are
plotted as in Fig. 2.9. In Fig. 2.9a, the gradient ma of the line tted is
1.17/9 = 0.129, leading to an identied value f 0:129=2:73 0:047.
In the case of Fig. 2.9b, the slope is approximated as ha = 1.345, thus giving
d lnha 0:296 ! f d=2p 0:0472.
Example 2.4 Find the damping ratio for a strongly damped vibration waveform
shown in Fig. 2.10 by reading peak heights y0, y1, y2, for consecutive half
cycles.
Answer
The peak amplitude readings for each half cycle are: 14.5, 8, 4, 2.5, , which are
plotted as in Fig. 2.9. In Fig. 2.9a, the gradient ma of the line tted is 1.25/
(4/2) = 0.63, leading to an identied value f 0:63=2:73 0:23.
In the case of Fig. 2.9b, the slope is approximated as hy = 2.054, thus giving
d 2 lnha 1:44 ! f d=2p 0:23.
Example 2.5 Find the damping ratio for a poorly damped vibration shown in
Fig. 2.11 by reading the amplitude envelope height b1, b2, b3, at a constant time
interval Dt [s] in (a) of the gure and determine the natural frequency fn using the
magnied time domain waveform in (b).
Answer
The magnied graph (b) gives fn = 30 Hz. The readings of amplitude envelope
height at an interval of t = 1 s in graph (a) are: 46, 26, 14, 8, , which are plotted
as in Fig. 2.9.

Fig. 2.10 Large damping example of = 0.2


y0
vibration waveform 3 y n : measuring at each half cycle
2
y2
1 y4
0
1 y5 2 [s]
1 y3
2 y1
3
2.2 Damped Free Vibration 23

(a) (b)
b0 = 46
2
b1 = 26 f n [Hz]
1.5 b2 = 14
b3 = 8 b5 = 2 1.5
b4 = 4
1 1 [s]
1.0 1.1 1.2
0.5
0.5 t [s]
Zoomed plot area

1 2 3 4 5 [s]
Example of = 0.003 Zoomed plot

Fig. 2.11 Small damping vibration waveform

Fig. 2.12 Estimation of


damping ratio B A
160 [Hz]

18

5.5
amplitude

[s]

0 0.1 0.2 0.3 0.4 0.5

0 2 4 6 8 10
2
frequency [ 10 Hz]

In Fig. 2.9a, the gradient ma 1:36=5 Dt fn 0:009 gives an identied


value f 0:009=2:73 0:00033.
In the case of Fig. 2.9b, the slope is approximated as hb = 1.82, thus giving
d lnha =30 0:02 ! f d=2p 0:0032:
Example 2.6 Find the damping ratio of the system using its impulse response
waveform A and FFT analysis result B in Fig. 2.12.
Answer
The amplitude (p-p) 2a0 = 18 at 0 s and 2a1 = 5.5 at Dt 0:5 s, and the natural
frequency fn = 160 Hz are read. The damping ratio is identied as follows:

  
1 a0 1 a0 1 18 d
d ln ln ln 0:015 ! f 0:0024
n a1 Dtfn a1 0:5 160 5:5 2p
24 2 Basics for a Single-Degree-of-Freedom Rotor

2.2.3 Phase Lead/Lag Corresponding to Damping Ratio

Control force (e.g. electromagnetic force) via a control unit in response to a dis-
placement of a mass is generated, for example, by the feedback system of mechatronic
equipment. In the block diagram of Fig. 2.13a, the mass having the plant transfer
function Gp is the controlled object, and the controller transfer function, Gr, represents
the characteristics from the displacement input x through to the control force u.
The displacement x(t) as input and control force u(t), i.e., reaction force, as
output are expressed as follows in terms of the natural frequency xn:

xt a cos xn t
2:24
ut f0 cosxn t /

As shown in Fig. 2.13b, the phase difference / between the input and output
signals is positive in phase-lead control and negative in phase-lag control.
Considering that

ut f0 cosxn t / f0 cos xn t cos /  f0 sin xn t sin / 2:25

if u(t) is equivalent to the force exerted by the spring k and viscous damping c

ut  kxt c_xt ka cos xn t  caxn sin xn t 2:26

comparison of Eqs. (2.25) and (2.26) gives

ka f0 cos / ; caxn f0 sin / 2:27

x
(a) u x (b) displacement
1 1 a
Gp=
ms 2
frequency n t

electromagnet controller u x
electromagnet controller

controller G r u
f0
Block diagram
t
+
(c) 1 lead

f0 u
ca n

ka t

2 lag

x(t) and u(t)

Fig. 2.13 Phase lead/lag


2.2 Damped Free Vibration 25

It is clear from Fig. 2.13c or the second equation of Eq. (2.27) that phase
lead/lag means positive/negative damping, and / > 0 and / < 0 relate to stable
damping vibration and unstable self-excited vibration, respectively. The stability of
the system is thus dependent on the phase lead/lag of the reaction force u(t) with
respect to the natural frequency vibration x(t).
Using Eq. (2.27), the damping ratio of a stable system may be directly estimated
from the phase lead / > 0:

c cxn f0 sin / 1
f p tan / 2:28
2 mk 2k 2f0 cos / 2

Note that phase lead/lag means positive/negative damping and that the damping
ratio may be directly estimated from the phase lead / > 0.
Example 2.7 If the displacement signal x(t) at the natural frequency and the cor-
responding control force signal are measured as shown in Fig. 2.13 , what
damping ratio f is expected?
Answer
The gure gives / 40
, which leads to f 1=2 tan 40
0:4.

2.3 Unbalance Vibration of a Rotating Shaft

2.3.1 Complex Displacement and Equation of Motion

Consider a rotating shaft system, vertically supported by bearings at both ends,


having a rotating disk at the midpoint as shown in Fig. 2.14a. Since the rotor spins
in a horizontal plane with the rotating speed X, gyroscopic moment are negligible
and the effect of gravity are considered to be small. Assume that the shaft has
stiffness, but no mass.

x G = x + cos t
(a) (b) Y (c)
y G = y + sin t y
2
m
k G
yG G
S G t
S S
y
m x
x, y | SG | =
c
t
O x xG X

Rotor system Coordinate system Dynamic model

Fig. 2.14 Unbalanced rotor


26 2 Basics for a Single-Degree-of-Freedom Rotor

y y
G ( j ) = A = ae j
e jt 2 z(t) = Ae j t x
G(s) =
x s 2 + 2 n s + n2
a

rotational vector orbit

Fig. 2.15 Block diagram of unbalance vibration

Consider an inertial coordinate system O-XY xed in the space with the origin
coinciding with the shaft center at rest as shown in Fig. 2.14b. The shaft is xed at
the disk centroid S. Let the vibration displacement of the shaft, as measured by
displacement sensors, be (x, y). The gravity center G of the disk is at a distance e
from the centroid S. For a rotational angular velocity X, the angle formed by the
S-G axis, xed to the shaft, and the O-X axis is the angle of rotation Xt.
Taking viscous damping c into account, a single-dof model consisting of mass,
spring and viscous damping as shown in Fig. 2.14c, can be constructed for the rotor
equation of motion derived from Newtons second law. For the center of gravity of
the rotor fxG x e cos Xt; yG y e sin Xtg , the reaction forces of spring and
damping are proportional to the displacement of the centroid S (x, y). Therefore,

mxG kx  c_x


2:29
myG ky  c_y

which can be rewritten as

mxG c_x kx meX2 cos Xt


2:30
myG c_y ky meX2 sin Xt

Using the complex displacement z = x + jy for simplicity, the equation of


motion is written as

mz c_z kz meX2 ejXt ! z 2fxn z_ x2n z eX2 ejXt 2:31

where x2n k=m ; 2fxn c=m.


A block diagram (Fig. 2.15) can be constructed with the rotation vector ejXt as
the input and the unbalance vibration as the output. The transfer function G(s) is
dened from the above equation as shown in the gure.

2.3.2 Complex Amplitude of Unbalance Vibration

An unbalance vibration is observed as a forward whirling motion synchronized with


the rotation angular velocity X including an orbit radius a and a phase lag angle u
2.3 Unbalance Vibration of a Rotating Shaft 27

output
input output
rotational signal 2 unbalance vibration a
G(s) = 2
cos t s + 2 n s + n2 xp = a cos( t )
1
= Re[ Ae j t ]
j input
G ( j ) = A = ae

Fig. 2.16 Unbalance vibration waveform

behind the rotating unbalance direction (see Fig. 2.15 for the denition of a and u).
This motion is described by

zt  AejXt  aeju ejXt aejXtu 2:32

where A  aeju  a\  u is the complex amplitude, which is equal to the value


of the transfer function G(s) with s jX:

eX2 ep2
A GjX 2:33
X2 x2n 2jfxn X 1  p2 2jfp

where p X=xn is the dimensionless rotating speed.


When observing the orbit from one stationary direction, the whirling motion
appears as a repetition of approaching and leaving motion, i.e., vibration. For
example, the vibrations in the X and Y directions are thus represented by

xt a cosXt  u ReAejXt 
2:34
yt a sinXt  u ImAejXt 

The block diagram of Fig. 2.16 corresponds to the motion in the X direction.
The observed phase lag angle u represents the phase difference between the cosine
input waveform and the output waveform x(t).

2.3.3 Resonance Curves

Graphs of the vibration amplitude a and phase difference u of unbalance vibration


versus rotational speed are called resonance curves or Bode plots. Figures 2.17a, b
show examples with various damping ratios f:
(1) Amplitude
As seen in Fig. 2.17a, the amplitude approaches zero at low rotational speeds
p X=xn 1; while it approaches the mass eccentricity e (which is relatively
small) at high rotational speeds p X=xn 1: The amplitude increases rapidly as
X approaches xn . The peak amplitude is innite if f 0, for an undamped system,
28 2 Basics for a Single-Degree-of-Freedom Rotor

(a) = 0.05 (c)


5 0.1

amplitude a [ ]
Im
4
0.2
0.3 6 4 2 1.0 0 2 4 6
3 0.4
0.71 Re
2 1.0 2 =
1 0.2 0.71
0 4 0.4
0 0.5 1 1.5 2 2.5 3 0.3
0.1
p [ / n ] 6

5
0.0
Amplitude curve
8
(b)

=
= 0.05 10
phase difference [ ]

0
0.1
0.2 0.3
90 0.4 0.71 Nyquist plot ( polar plot )
1.0

180
0 0.5 1 1.5 2 2.5 3
p [ / n ]
Phase curve

Fig. 2.17 Unbalance vibration response

but decreases as the damping ratio f increases. The peak amplitude is called the
resonance amplitude ap, and the corresponding rotating frequency is called the
critical speed. They are approximated by:

ap 1 1
Q-value or Q-factor) : Q p  2:35
e 2f 1  f 2 2f

X 1
Critical speed: p p  1 2:36
xn 1  2f2

(2) Phase difference


The curves in Fig. 2.17b show that the phase difference u proceeds from 0
to
180
as the rotational speed p X=xn increases. The phase lag u at the critical
speed p 1X xn is always 90
. The phase lag changes slowly in strongly
damped systems around the resonance point, but it changes rapidly in poorly
damped systems.

2.3.4 Nyquist Plot

A Nyquist plot is a graphical representation of the complex amplitude of Eq. (2.33)


in the complex plane. It is also called a vector locus or polar plot, because it
displays the complex number as a vibration vector (amplitude a \ phase u) in
polar coordinates.
2.3 Unbalance Vibration of a Rotating Shaft 29

(a) (b) (c)


Y Y 90 Y
180
0
t
G t
y
G S
t y
y
S S
G
O x O x x
X X O X

p << 1 low speed p 1 critical speed p >> 1 high speed

Fig. 2.18 Unbalance whirl motion

The plot in Fig. 2.17c shows that the amplitude gradually increases while
drawing a clockwise circle tangent to the real axis after starting at the origin. This
starting vector direction is set to the real axis (0
), because the phase difference of
the vibration response with respect to the unbalance force is zero, i.e., they both
have the same phase.
At low rotational speeds (p X=xn 1, a slight lag of the vibration response
after the unbalance force appears in the vibration vector pointing to the 0 degree
direction, i.e., the real axis. At the critical speed (p X=xn  1, the vibration
vector lags by 90
and passes the zenith of the circular trajectory. At high rotational
speeds (p X=xn 1, the amplitude decreases again, lagging by 180
, and
approaches the point e; 0 from the negative direction of the real axis.
Example 2.8 Whirling motion
Figure 2.18 shows three instantaneous states of a rotor whirling by unbalance
vibration. Find the complex amplitudes in (a) to (c) in the gure.
Answer
Since the complex amplitude = OS/OGe\  u, (a) 1:0e\  18
, (b) 3:2e\  90

and (c) 2e\  169


.
Example 2.9 Waveform of unbalance vibration
Figure 2.19 shows comparisons of the input cosine waves and the corresponding
unbalance vibration waveforms measured by an oscilloscope and their magnied
views for measuring phase differences. Find the complex amplitudes in (a)(c) in
the gure.
Answer
Since the complex amplitude = a=ee\  u, (a) 0:3e\  15
, (b) 5:9e\  90
and
(c) 1:5e\  175

30 2 Basics for a Single-Degree-of-Freedom Rotor

(a) (b) vibration eccentricity (c)


vibration eccentricity vibration eccentricity

eccentricity eccentricity eccentricity

vibration vibration vibration



( almost in-phase ) ( lag 90 )
( almost out-of-phase )
p << 1 low speed rotation p 1 critical speed p >> 1 high speed rotation

Fig. 2.19 Amplitude and phase related to unbalance eccentricity

2.3.5 Bearing Reaction Force at Resonance

The spring-viscous damping model of Fig. 2.14c can also be applied to a bearing,
and permits one to describe how the unbalance force is balanced with bearing force
at resonance. At the critical speed where the phase difference u  90
, the rotor
vibration is represented by

zt  Ap ejXt  ap ej90 t ejXt jap ejXt 2:37

Substituting this into Eq. (2.31),

k  mX2 Ap jXcAp meX2 2:38

Considering Ap jap , the equation for static balance is obtained by setting the
unbalance force to the imaginary axis as shown in Fig. 2.20b:

jmeX2  jXcap k mX2 ap 0 2:39

Figure 2.20a shows the instantaneous rotor position farthest from the origin
through the X axis in a whirling orbit at the critical speed. The phase of the
vibration vector Ap = OS is lagging by 90
behind the unbalance direction SG.
2.3 Unbalance Vibration of a Rotating Shaft 31

(a) Y (b) m 2 (c)


G Q times Fb
m 2
t X ka p G
m 2a p
O A S S ma p 2

Fb = ap k2 + ( c ) 2 c ap

Whirling Force balance


displacement vector

Fig. 2.20 Force balance at resonance point

Figure 2.20b illustrates the balance of forces at this point in time.


p
Horizontal directionreal axis : k mX2 0 ! Xc k=m xn
mexn e
Vertical directionimaginary axis : meX2  Xcap 0 ! ap
c 2f


q
Bearing reaction forcejFb j
k jXcAp
ap k2 cX2
q
map xc 1 4f2  map X2c
2

2:40

The balance in the horizontal direction gives the critical speed and that in the
vertical direction the resonance amplitude. Figure 2.20 indicates that the input
unbalance force is balanced by the damping force at resonance. Increasing the
viscous damping coefcient reduces accordingly the resonance amplitude. It can be
stated that
The resonance amplitude ap is approximately Q times as great as the mass
eccentricity e, and
The bearing reaction force at resonance Fb is approximately Q times as great as
the unbalance force meX2 .

Example 2.10 The force balance is represented by a closed arrangement of vectors


as shown in Fig. 2.20c. Conrm that the three states of Fig. 2.18 can similarly be
represented by the closed form as in Fig. 2.21.
Answer
The angle formed by the amplitude vector a (reset on real axis) and the eccentricity
vector e is the phase lag u. Since the force balance is mX2 a meX2 eju Fb 0;
the bearing reaction force vector Fb is added to the sum of the two vectors points as
being back to the origin, as shown in Fig. 2.21.
32 2 Basics for a Single-Degree-of-Freedom Rotor

(a) (b) (c)

ka
a

m 2
Fb m 2
G Fb c a G Fb
G
S m 2 S m 2 a S
m 2 a m 2 a

p << 1 p1 p >> 1

Fig. 2.21 Force balance

2.3.6 Transmissibility of Unbalance Vibration


to Foundation

The transmissibility T that is the ratio of unbalance vibration to the foundation of


the machine, normalized with the unbalance force at the critical speed mex2n , is
written, considering Eq. (2.33), as
q
jFb j jkx c_xj p j1 2jfpj 2p 2
1 2fp2
T q 2:41
mex2n mex2n j1  p2 2jfpj
1  p2 2 2fp2

This ratio T is shown graphically in Fig. 2.22. When passing through the critical
speed, both of the amplitude and transmission ratio are large and the peak value is
p
T  Q. T always has the value 2 at p X=xn 2 . It should be noted that at
high rotational speeds T increases with increasing speed even if the amplitude
decreases. Reduction of this transmissibility to the foundation is only possible by
decrease of e, i.e., by having a rotor that is well-balanced.

Fig. 2.22 Transmissibility of 5


unbalance force to base = 0.05 0
4 0.1 1.
transmissibility

1
0.2 0.7
3 0.3
0.4
2

1
2
0
0 0.5 1 1.5 2 2.5 3
p
2.4 Evaluation of Q-Value 33

20 0.025

Mm = 10
Modal sensi- Expected running
17.5 0.028 conditions
tivity range
15 0.033 Very low Very smooth resonant
E A
sensitivity speed; difficult to detect

damping ratio
12.5 0.04
Q-value

Low Smooth, low and


B

5
10 0.05 sensitivity stable vibration

m 7.
D M = Moderate Acceptable, moderate and
7.5 0.067
=5 C
C sensitivity slightly unsteady vibration
5
m

0.1
M

B Sensitive to unbalance;
High Field balancing may be
2.5 2.5 0.2 D
A Mm = sensitivity required
0 0.0
0.8 0.9 1.0 1.1 1.2 1.3 1.4 Very high Too sensitive to unbalance;
E
sensitivity to be avoided
R = service speed / critical speed
1/ R
1.3 1.2 1.1 1.0 0.9 0.8 0.7

Fig. 2.23 Modal sensitivity (Q-value) criteria

2.4 Evaluation of Q-Value

2.4.1 Q-Value Criterion

The Q-value is a universal and measurable index for evaluation of the vibration
response characteristics of a rotating machine. It is very important to ascertain how
low the Q-value can be maintained. The ISO 10814 [10] standard provides
guidelines for Q-value design for various rotors. An example is shown in Fig. 2.23.
The abscissa is rated rotational speed, non-dimensionalized with the critical speed,
and the ordinate is the tolerated Q-value. The chart is usually used to evaluate the
Q-value of the resonance mode near the rated speed.
As well as ISOs Q-value evaluation, the API standard [11] is strong in the oil
and gas eld of turbo-machinery industry.
Example 2.11 Conrm that a machine with a rated rotational speed of 11,000 rpm
and a critical speed of 10,000 rpm requires approximately Q 9 for a B-zone
specication. Note that (rated rotational speed)/(critical speed) is 1.1.
Example 2.12 If the rotational speed of the machine increases slowly, so that the
abscissa in Fig. 2.23 is traced very slowly from left to right through the critical
speed, the machine may be considered to always be in the steady state, and the
Q-value at (rotational speed)/(critical speed) = 1 should be evaluated in the whole
range. Conrm that the machine requires Q 5 for a B-zone specication.
34 2 Basics for a Single-Degree-of-Freedom Rotor

Fig. 2.24 Measuring n


1
Q-value (ISO 10814)
0.8
0.707 n
70% Q= = n

amplitude
0.6
2 1

0.4

0.2

0
1 2
speed

2.4.2 Measurement of Q-Value by the Half Power Point


Method

The ISO standard species procedures to measure the Q-value from a resonance
curve for forced vibration by the half power point method. Consider a measured
resonance curve shown in Fig. 2.24, where the maximum amplitude is amax at the
resonance point xn . The difference DX of the frequencies X1 and X2 of the
half-power points, i.e., the intersect to give the half power amplitudes
p
a70 amax = 2  0:7amax

which permits determination of the Q-value by

1 xn
Q 2:42
2f DX

Example 2.13 Measure the Q-value for the resonance curve shown in Fig. 2.24.
Answer
Draw a horizontal line through the 70 % amplitude. Use a length scale to obtain
xn = 20 mm and DX 4 mm: Q = 20/4 = 5 and f 0:1 are thus obtained.

2.4.3 Measurement of Q-Value Using a Nyquist Plot

This is also a method specied in the ISO standard. Suppose a vector trajectory that
describes a counterclockwise circle (Fig. 2.25). Connect the starting point of the
circle, i.e., the origin, and the point xn corresponding to the peak amplitude. Draw
two lines from the origin on both sides of this line at angles of 45
each, and let the
intersections with the vector trajectory be X45 and X135 . The Q-value is given by
2.4 Evaluation of Q-Value 35

Fig. 2.25 Measuring 0.85


Q-value (ISO 10814) 0.90

45 = 0.95
45
45

1.00

1.20
135 = 1.15
p = / n = 1.05

1.10 n 45
Q= = 4.99
n2 245

Fig. 2.26 Measuring 3 450 3 300


Q-value (ISO 10814) 100
3 600 3 210
3 150
80
60
3 060
40
45

20
y [ m ]

3 000
0
40 80 120 160 200 x [ m ]
20

45

40
2 940
60
80
2790 2 850
100
2 550 2 700 2 730
n 45
[r/min] Qn =
n2 245

N1 N 45 Q1

3 000 r/min 2 710 r/min 4.91

1 xn xn X45
Q 2 2:43
2f X135  X45 xn  X245

Note: The phase angle is theoretically concerned with phase lead. The trajectory
of the Nyquist plot proceeds clockwise due to the phase lag of the response.
However, measuring instrument (called vector monitor) usually operates on the
basis of a phase lag, and gives a counterclockwise trajectory as shown in Fig. 2.26.
Example 2.14 Measure the Q-value for the Nyquist plot shown in Fig. 2.25.
Answer
Interpolation based on p X=xn in the gure gives X45 0:95 and X135 1:15.
Therefore Q = 1/(1.15 - 0.95) = 5.
36 2 Basics for a Single-Degree-of-Freedom Rotor

Example 2.15 Measure the Q-value for the Nyquist plot shown in Fig. 2.26.
Answer
With the critical speed xn 3; 000 rpm and X45 2; 710 rpm, Eq. (2.43) gives
Q = 4.91.

2.4.4 Re-evaluation of Q-Value for Rapid Acceleration

The Q-value of Fig. 2.23 evaluated from the resonance curve may be used for cases
in which the critical speed is passed so slowly such that the angular acceleration is
negligible and that the system is almost in steady state. In contrast, if the critical
speed is passed rapidly, there is not enough time for the vibration to attain a steady
state. This means that the peak amplitude is smaller in such a case. In other words,
the Q-value apparently decreases by rapid acceleration of rotation. ISO provides a
chart (Fig. 2.27) to nd equivalent damping ratios by re-evaluating the apparent
decrease in Q-value.
Example 2.16 Determination of zone according to the Q-value criterion
Fill in the blanks [ ? ] below.
(1) If a rotating machine, with a rated rotational speed of 3,000 rpm, 1st critical
speed of 2,730 rpm and damping ratio f 0:04, is slowly accelerated (and
start/stop is infrequent), zone [ ? ] is assigned to it according to Fig. 2.23,
because Q = 12.5 and gR 3; 000=2; 730 1:1.

25
acceleration
deceleration
a=0
20
4 10 3 a = A / n 2 = constant
3 A = acceleration (1/ s 2 )
8 10
equivalent Q-value

15
16 10 3
32 10 3
10
64 10 3
6.3
5

0 0.05 0.1 0.15


= 0.04 Measured damping ratio

10 5 3.3
measured Q-value

Fig. 2.27 Equivalent Q-value in acceleration and deceleration (ISO 10814)


2.4 Evaluation of Q-Value 37

(2) If the same machine is slowly accelerated up to 3,000 rpm, gR should be


regarded as 1 because operation at the critical speed cannot be excluded.
According to Fig. 2.23, zone [ ? ] should be assigned in this case, which notes
that the operation at the critical speed is very dangerous.
(3) If the same machine is accelerated rapidly from 1,000 to 30,000 rpm over the
critical speed 2,730 rpm, where f 0:04 and Q = 12.5,
Acceleration: A 2p30000  1000=60=1:161 2615 s2 , and

2615
Constant : a 32 103 :
2730 2p=602

An equivalent Q-value of 6.3 is re-evaluated according to Fig. 2.27, as indicated


by arrows. Since gR 1 because the machine passes the critical speed during the
acceleration, zone [?] is assigned to the machine for Q = 6.3 according to Fig. 2.23.
Answer
(1) C, (2) E, (3) C
Example 2.17 In Fig. 2.28a, consider a rotor of a total length l = 750 mm, con-
sisting of a disk of mass m = 40 kg xed to the right end, and uniform shaft of mass

j t
(b) m 2 e

ms x
m
3

(a)
= 10 m
l = 750
kb h2 cb h 2
m 2 e
j t
l b = 500
D

1
L Equivalent model (x displacement)

m x
m s = 30 kg xb (c) m 2/ h e
j t
40 kg


ms
kb cb m + xb
3
h2
( h = lb / l ) x vibration

Rotor system kb cb

( x b = hx )

Equivalent model (xb displacement)

Fig. 2.28 Rotor and equivalent model system


38 2 Basics for a Single-Degree-of-Freedom Rotor

300 3 103

250 amplitude of disk 1

amplitude of disk [ m ]

bearing load [ N ]
200 2
Q = 33

150
bearing load Fb 2
100 1

50

0 0
0 10 20 30 40 50 60
speed [rps]

Fig. 2.29 Unbalance vibration

ms = 30 kg. The shaft has a pinned support at the left end, and a bearing support at
a distance of lb = 500 mm from the left end (spring constant kb = 10 MN/m and
viscous damping coefcient cb = 10.3102 Ns/m). The resonance curve and
bearing reaction force for the unbalance vibration of the rotor are shown in
Fig. 2.29. Answer the following questions.
(1) The single-dof model equivalent to the system shown in Fig. 2.28a is shown
in Figs. 2.28b, c. The former model with respect to the displacement x of the
shaft tip includes 1/3 times of the shaft mass. The latter is another model with
respect to the displacement xb of the bearing portion (xb = hx). Obtain the
equation of motion for both models.
(2) Measure the critical speed Xc , the peak amplitude apeak of the disk, and the
Q-value of the system in the resonance curve .
(3) Find the equivalent mass meq using the displacement xb of the bearing portion.
(4) Find the peak displacement ap of vibration at the bearing portion.
(5) Find the peak load Fb on the bearing.
(6) Assuming that the journal diameter D = 100 mm and length L = 50 mm, nd
the journal surface pressure P at resonance.
(7) Find the acceleration A necessary to halve the surface pressure by rapid
acceleration.

Answer
(1) The equation of motion of Fig. 2.28a: I h cb l2b h_ kb l2b h meX2 lejXt ,

Zl 
qAl 2  ms  2
where I ml 2
qAz2 dz m l m l
3 3
0
2.4 Evaluation of Q-Value 39

The equation of motion of Fig. 2.28b: Introduce h x=l into the above
equation and divide it by l;

m ms =3x cb x_ kb xl2b =l2 meX2 ejXt

The equation of motion of Fig. 2.28c: Introduce h xb =lb into the above
equation and divide it by lb;

m ms =3l2b =l2xb cb x_ b kb xb l=lb meX2 ejXt

(2) Xc 47 Hz, apeak 267 mm, Q 33, f 0:015.


(3) meq the modal mass of the linear mode; when setting the bearing journal
displacement xb 1
750=5002 m ms =3 m ms =h2 112:5 kg:
(4) The peak displacement ap of vibration at the bearing portion
ap = 267 500/750 = 178 m.
(5) The peak load Fb on the bearing; jFb j meq ap X2 = 112.5 178 106
(2 p 47)2 = 1744 N, which corresponds to the peak of the bearing reaction
force curve .
(6) The surface pressure P jFb =DLj = 1744/(0.1 0.05) = 3.4 105
Pa = 0.34 MPa.
(7) Aiming at a point (Q = 15 on y-axis, f = 0.015 on x-axis) on Fig. 2.27,
a = 6 103 may be chosen; this leads to an acceleration rate
A ax2n 5221=s2  83 Hz=s, meaning that the critical speed must be
passed through in approximately 50/83  0.5 s.

2.4.5 Vibration in Passing Through a Critical Speed

Unlike in the steady state, reduced resonance peak amplitude and transient
amplitude changes are observed when passing the critical speed rapidly. For
example, calculation predicts the vibration shown in Fig. 2.30 for a system with
vibration and amplitude

steady state amplitude


10
Q = 8.1 |z ( t) |

p = / n
0

x (t)
10
0 1 2

Fig. 2.30 Unbalance vibration wave form in acceleration f 0:04; a 16 103 ; e 1


40 2 Basics for a Single-Degree-of-Freedom Rotor

14 Im
a = 10 -3
-3
12 2 10 steady state
steady state amplitude -3 10
4 10
10 8 10 -3 in acceleration
amplitude | z ( t ) |

-3
16 10
8
Re
6 -10 10
Im
4 steady state
10
2
in acceleration
0
0 0 .5 1 1.5 2
p= /n Re
-10 10

Fig. 2.31 Constant a and unbalance vibration envelope and polar plot

f 0:04 and acceleration a 16 103 [1/s2]. The x-directional vibration


waveform x(t) and the radius of whirling |z(t)| shown in the gure along with the
amplitude in steady state. The peak at the critical speed indicates that Q = 8.1,
which is in accordance with the ISO guideline shown in Fig. 2.27.
Figure 2.31 compares envelopes and polar plots for different values of
constant a. The amplitude approaches the steady state amplitude in slow acceler-
ation, but the polar plot remains considerably different from that for the steady state:
it resembles a circle deformed downwards, and corresponds to smaller amplitude.
Acceleration for balancing purposes should be as slow as possible because an
accurate polar plot is needed.
Chapter 3
Modal Analysis
of Multi-Degree-of-Freedom Systems

Abstract The preceding chapter dealt with the basics of rotor vibrations con-
cerning a single-degree-of-freedom (single-dof, 1-dof) system. An actual machine
should, however, be analyzed as a multi-degree-of-freedom (multi-dof) system
where multiple masses are arranged according to the shape of the rotor shafting.
The equation of motion for such a system is represented using matrices. Eigenvalue
analysis of the multi-dof system gives the natural frequencies and eigenmodes.
These are important factors in rotor design because they represent the resonance
frequencies and the vibration mode shapes at critical speeds. This chapter also
discusses modal analysis, in which a multi-dof system is reduced to an assemblage
of single-dof systems utilizing the orthogonal condition of the eigenmodes. In other
words, a complicated actual system is simplied to a set of simple 1-dof systems
corresponding to each mode. In addition, a simple estimation method of the natural
frequency and the damping ratio is presented based on the orthogonality condition.

 
Keywords Multi-dof Eigenvalue Eigenmode Modal analysis   Equivalent
 
mass Stiffness matrix method Transfer matrix method

3.1 Equation of Motion for a Multi-dof System

3.1.1 Multiple Mass Systems

The actual rotor of Fig. 3.1b may be represented as a single-dof system as shown in
Fig. 3.1a, with m equal to about 1/2 of the entire mass distributed over the shaft and
disks and a spring constant k corresponding to the flexural rigidity or stiffness of
shaft. This gives an intuitive understanding of the lowest mode vibration, but it is of
too low an order to achieve appropriate accuracy.
A model for higher precision may involve a discrete multi-dof system, i.e.,
a multiple-mass system, as shown in Fig. 3.1c, in which each disk stage is regarded
as a mass point. It also permits analysis of more than one eigenmode. The method of
modal analysis of the equation of motion for a multi-dof system is now described.

Springer Japan 2017 41


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_3
42 3 Modal Analysis of Multi-Degree-of-Freedom Systems

(a) m
k k

(b)

(c) m1 m2 m3 m4 m5 m6 m7 m8 m9 m10 m11

Fig. 3.1 Modeling of an actual rotor

3.1.2 Equation of Motion for a Two-dof System

Firstly, consider a two-dof system shown in Fig. 3.2. Let the vibration displacements
of the masses m1 and m2 be x1 and x2, respectively. The indices are set by node
numbers. The spring reaction force (internal force) is illustrated in Fig. 3.2b, c.
When the mass m1 is moved by x1 to the right, the spring k1 pulls the mass to the left
with a reaction force k1x1, and the spring k12 pushes the mass to the left with a
reaction force k12 x1  x2 . This gives the equation of motion for m1 shown in the
gure according to Newtons second law. Similar analysis applies to the mass m2.
The equation of the motion for the entire system is thus written in the matrix form:

KX 0;
MX 3:1

where
    

m 0 k k12 k12 x1
M 1 ; K 1 ; X :
0 m2 k12 k12 k2 x2

k1 k 12 k2
m1 m2

x1 x2
(a)
x1 x2

k1 x1 m1 k 12 ( x 1 x 2 ) k 12 ( x 2 x 1 ) m2 k2 x2
.. ..
m1 x 1 = k 1 x 1 k 12 ( x 1 x 2 ) m2 x 2 = k 12 ( x 2 x 1 ) k 2 x 2
(b) (c)

Fig. 3.2 2-dof system


3.1 Equation of Motion for a Multi-dof System 43

The mass matrix, M, contains the masses placed at diagonal positions. The
stiffness matrix, K, is symmetric and positive denite, consisting of spring constants
arranged by the following rules:
The spring constants connecting the mass to ground, denoted as k1 and k2, are
added only to the diagonal elements that correspond with the node numbers
(1 and 2),
The spring constant connecting two masses, denoted as k12, is added to the
diagonal elements that correspond with the node numbers (1 and 2), and sub-
tracted from the non-diagonal elements of (1,2) and (2,1).
The operation called superposition allows systematic construction of the
stiffness matrix. The matrix K in Eq. (3.1) can be constructed by superposing each
of the springs, k1, k12 and k2, in sequence according to:
     
k1 0 k12 k12 0 0
K : 3:2
0 0 k12 k12 0 k2

3.1.3 Equation of Motion for a Multi-dof System

The equation of motion for a discrete multi-dof system, for example, as appropriate
from Finite Element analysis, generally has a matrix form. The equation for the
discrete system shown in Fig. 3.3, containing an external force F and a displace-
ment sensor signal y, is

DX_ KX BFt
MX
3:3
y CX

where M is the mass matrix (n  n), D the damping matrix (n  n), K the stiffness
matrix (n  n),
X the displacement vector (n  1), F the external force component (1  1),
B the input matrix (n  1) = [0, 0, 1, 0, ]t
(a column matrix with 1 only for the node excited and 0 at all other elements),
and C the output matrix (1  n) = [0,..0, 1, 0, ]
(a row matrix with 1 only for the sensor node and 0 at all other elements).
The procedure of deriving the equation of motion in Finite Element Analysis is
followed as below for the multi-dof system shown in Fig. 3.3.
(1) Number of dimensions n: Numbers are rst assigned to the mass of each node.
Hence each node has a single-dof. The total dimension is n-dof.
44 3 Modal Analysis of Multi-Degree-of-Freedom Systems

f
c1 c 12

m1 m2 m3 m4 mn
k1 k 12
y
x1 x2 x3 x4 xn

Fig. 3.3 Multi-degree of freedom mass-spring-damper system

Fig. 3.4 Additional mass m

mi mj

xi xj

Note: In the case of the rotating shaft shown in Fig. 3.1, where each node has
two motions (corresponding to the bending deflection and tilting angle of each
shaft portion), the total dimension is thus 2n-dof.
(2) Displacement vector X = [ x1, x2, , xn]t: A row vector consisting of the
displacement of each mass in the node number order.
(3) Mass matrix M = Diagonal [m1, m2, , mn]: The masses of each node are
arranged diagonally. If a mass Dm is added to the ith node (see Fig. 3.4), we
should add Dm to the (i, i) element of the matrix.
(4) Stiffness matrix K: Constructed by superimposing all spring elements
step-by-step in the stiffness matrix as principally stated in Eq. (3.2). Initially,
the matrix is a zero matrix.
(5) Damping matrix D: Constructed in the same way as the stiffness matrix, K,
using individual viscous damping elements.
(6) Input matrix B: A column matrix where the elements are 1 for the nodes under
external force and 0 otherwise.
(7) Output matrix C: A row matrix where the elements are 1 for the sensor nodes
and 0 otherwise.

Example 3.1 Construct the stiffness matrix by superposition for the system of
Fig. 3.5.

Fig. 3.5 Superposition k4

k1 k2 k3

1 2 3
3.1 Equation of Motion for a Multi-dof System 45

Answer

2 3 2 3 2 3 2 3
k1 0 0 k2 k2 0 0 0 0 k4 0 k4
6 7 6 7 6 7 6 7
k 4 0 0 0 5 4 k2 k2 0540 k3 k3 5 4 0 0 0 5
0 0 0 0 0 0 0 k3 k3 k4 0 k4
2 3
k1 k2 k4 k2 k4
6 7
4 k2 k2 k3 k3 5
k4 k3 k3 k4

Example 3.2 Find the stiffness matrix for the system of Fig. 3.6, which is identical
with that of Fig. 3.5 except for the different node numbers.
Answer

2 3
k2 k3 k3 k2
K 4 k3 k3 k4 k4 5
k2 k4 k1 k2 k4

Example 3.3 Draw the diagram for the multiple mass-spring system corresponding
to the equation of motion:
2 32 3 2 32 3 2 3
24 0 0 x1 6 2 1 x1 0
4 0 8 0 54 x2 5 4 2 2 0 54 x2 5 4 0 5
0 0 1 x3 1 0 1 x3 0

Answer
Fig. 3.7.

Fig. 3.6 3-dof system k4

k1 k2 k3

3 1 2

Fig. 3.7 3-dof system x2


2
8
3
24
1
1
x1
x3
46 3 Modal Analysis of Multi-Degree-of-Freedom Systems

3.2 Modal Analysis (Normal Mode Method)

3.2.1 Eigenvalue Analysis

For a solution of the equation of motion in undamped free vibration (D = F = 0),


suppose a natural frequency is xn and set

X /ejxn t 3:4

where / is the eigenmode (termed eigenvector in linear algebra).


Substituting Eq. (3.4) into Eq. (3.3), the corresponding eigenvalue problem of
this M-K system is:
x2n M/ K/ 3:5

which gives pairs of eigen solutions, i.e., eigenvalues x2n and eigenvectors /:

x2n ; / ) x21 ; /1 ; x22 ; /2 ; . . .; x2i ; /i . . . 3:6

The eigenvectors of an undamped system are called normal modes.

3.2.2 Orthogonality

A modal matrix U is formed by locating the column array of eigenvectors /:

U /1 /2    /n  3:7

Since the mass matrix M and stiffness matrix K are real symmetric and positive
denite, the eigenvalues are positive x2n [ 0. The eigenvectors are real and
mutually orthogonal with respect to the mass matrix M and the stiffness matrix K:
 
mi i j ki i j
/ti M/j ; /ti K/j 3:8
0i 6 j 0i 6 j

Therefore, a congruent transformation of the mass and stiffness matrices by


using the modal matrix leads to the off-diagonal terms being zero, resulting in
diagonal matrices:

Ut MU diagonalm1 m2 . . .mn   M  : modal mass matrix


3:9
Ut KU diagonalk1 k2 . . .kn   K  : modal stiffness matrix

A modal mass mi* and modal stiffness ki* are obtained for each x2i ; /i . Since
the lumped mass matrix M is diagonal, the modal mass mi /ti M/i is equal to the
3.2 Modal Analysis (Normal Mode Method) 47

P
sum of (mass)  (mode displacement)2 for each node mi nj1 mj /2ith;j .
Similarly, the modal stiffness ki /ti K/i is the sum of the strain energy stored at
each spring and total strain energy indicates that ki mi x2i .

3.2.3 Reduced Order Modal Model

Transformation from the physical coordinates, X, to modal coordinates (often ter-


med normal coordinates), g, which reflect the weighting contribution of each
eigenmode to be included in actual displacement X, is dened by
2 3 2 3
x1 g1
6 . 7 6 7
X  4 .. 5 /1 g1 /2 g2    /l gl /1    /l 4 ... 5  Ug 3:10
xn gl

The number l is called the mode truncation. Substituting Eq. (3.10) into
Eq. (3.3) and performing the congruent transformation yields the equation of
motion in the modal coordinates. If the damping matrix can be transformed
similarly,

M  gt D gt
_ K  gt B Ft
3:11
yt C gt

where
modal input coefcient B Ut B    bi /ti B   
= modal displacement of the node at which the excitation acts,
modal output coefcient C  CU    ci C/i    and
= modal displacement at the sensor location
modal damping

D Ut DU
! diagonal   di /ti D/i     diagonal   2fi xi mi   

It is practical to assume that D* can be diagonal for weakly damped vibration


systems, although this is not generally and mathematically true, except for Rayleigh
damping where damping is proportional to mass and stiffness D aM bK.
Figure 3.8 shows a mechanical model of the system including the modal
parameters obtained above (modal mass mi*, modal damping ratio fi and natural
frequency xi ). The multi-dof system shown in Fig. 3.3 is thus transformed to a
48 3 Modal Analysis of Multi-Degree-of-Freedom Systems

mode truncation
l -th
F (t)
modal
input
constant
1t B 2t B lt B nt B b i = i B
t

modal mass
m 1 m 2 m l m n m i = i t M i

1 2 l n
1 2 l n k i = m i i 2
1 2 l n c i = 2 i i m i

modal
output
C1 C2 Cl Cn constant
+ + + c i = C i
+ + + y (t)

Fig. 3.8 Modal model for multi-degree of freedom system (normal coordinates)

parallel sum of single-dof systems. This is called a reduced modal model because
the employed modes are truncated to a certain number, l, which is smaller than the
original matrix dimension n.

3.2.4 Vibration Response

In the transfer function description, the modal response of each mode is one of the
parallel relations from F(t) to y(t) as indicated in Fig. 3.8:

gi s bi
 2 i 1l 3:12
Fs mi s 2fi xi s x2i

The response of the sensor y, which is the sum of these modal responses, is therefore

ys X l
ci bi

3:13
Fs i1
mi s 2fi xi s x2i
2

No matter how numerous the mode number, l, may be, the response calculation
is simple because all are the single-dof systems. The block diagram for the s domain
between the input F and output y has a parallel structure as shown in Fig. 3.9.
3.2 Modal Analysis (Normal Mode Method) 49

1
b 1* c 1*
m 1* ( s 2 + 2 1 1 s + 12 )

F (s) 1 y (s)
b 2* c 2* +
m 2* ( s 2 + 2 2 2 s + 22 )


1
b l* c l*
m l* ( s 2 + 2 l l s + l 2 )

Fig. 3.9 Modal transfer function

In the mechanical model (Fig. 3.8) and block diagram (Fig. 3.9).
(1) If any modal input coefcient bi* is zero, the mode is not excited by external
activity and is therefore uncontrollable. The system is thus controllable only if
all the modal inputs coefcients bi* are non-zero.
(2) If any modal output coefcient ci* is zero, the mode vibration is not detected
by the sensor even if the machine resonates. The system is thus observable
only if all the modal output coefcients ci* are non-zero.
Note: Instead of using the modal response Eq. (3.13), the transfer function from
the excitation F to the sensor y is obtained directly by solving the equation of
motion:

Sensor response ys CMs2 Ds K1 BFs 3:14

Fig. 3.10 3-dof system f (t)


1 0.2 2

6 4 4
12 12 12 4

x1 x2 x3 y(t)

3 3
2

1
0
1
3
2

5
50 3 Modal Analysis of Multi-Degree-of-Freedom Systems

f(t) = cos t modal


input
impulse
2 1 2 constant
b i*
modal
mass 96 10 160
m i*
96 1 40 2 1 440 3
modal
output
1 = 1 2 = 2 3 = 3 constant
3 1 3
1 = 0.12 2 = 0.08 3 = 0.033 c i*
++ + +
displacement
sensor
y = x3

Fig. 3.11 Modal model (Impulse)

Example 3.4 Answer the questions below on the three-mass system shown in
Fig. 3.10.
(1) Find the mass matrix M, stiffness matrix K, input matrix B, and output
matrix C.
(2) Solve eigenvalue problem and conrm the corresponding eigenvectors
shown in Fig. 3.10.
(3) Derive the modal model shown in Fig. 3.11, assuming that the damping
matrix is diagonal.
(4) Find the transfer function G(s) between the input f and output y using the
modal model.
(5) Conrm Fig. 3.12 as the frequency response amplitude using the transfer
function.
(6) Estimate the peak amplitude of the frequency response using the Q-value.
(7) Conrm Fig. 3.13a as the acceleration waveform of the impulse response by
modal analysis.
(8) Conrm Fig. 3.13b as the natural frequency components by FFT analysis of
the impulse response waveform.
(9) Indicate the uncontrollable node in the system.
(10) Indicate the unobservable node in the system.
(11) Find the complex eigenvalues k of the system, calculate the exact damping
ratio by f Rek=Absk, and compare with the damping ratios assumed
in (3) above.
3.2 Modal Analysis (Normal Mode Method) 51

300

250 peak
amplitude [ mm ] 200
peak

150
peak
100 3rd only

50
1st only
2nd only
0
0 1 2 3 4
frequency [rad/s]

Fig. 3.12 Resonance curves and peak values

(a) (b)
300
acceleration [mm/s2]

200 1.4 2nd 3rd


1.2
100
acc. [mm/s2]

1.0 1st
T = 502.4 [s]
0 0.8
10 20 30 502.4 0.6 N = 1 024
100 [s] 0.4
0.2
200 0 1 2 3 4 5
freq. [rad/s]
300
Acceleration waveform FFT

Fig. 3.13 Impulse test

Answer

2 3
24 12 0
6 7
M diagonal 6 4 4  K 4 12 24 12 5
0 12 16
1 2 3
1:2 0:2 0
6 7
B 1 0 0 t ; C 0 0 1 ; D 4 0:2 0:2 05
0 0 2
52 3 Modal Analysis of Multi-Degree-of-Freedom Systems
2 3
2 1 2
2 x2n 1 4 9 ; xn 1 2 3 ; U 43 0 5 5
3 1 3

M  Ut MU diagonal 96 10 160 

K U MU diagonal 96 40 1440 
t

B 2 1 2 t ; C  3 1 3 
2 3 2 3
3 1 0:2 0:2 0 22:2 4:2 20:6
 t6 7 6 7
D U 4 0:2 0:2 0 5U 4 4:2 3:2 2:6 5
0 0 2 20:6 2:6 31:8
f f 0:12 0:08 0:033 g
 diagonal 22:2 3:2 31:8 
ys 2=96 1=10 2=160
4 Gs  3 2 1 2 3 2
Fs s 0:23s 1 s 0:32s 4 s 0:199s 9

(5) Using G(s) = y(s)/F(s) with Eq. (3.14), the transformation Gjx gives
Fig. 3.12.
(6) Peak values C B =2f=M  x2n {270, 156, 63} [mm] agree with dots  in
Fig. 3.12.
(7) Acceleration response a(t) = L1[s2G(s)] gives Fig. 3.13a.
(8) FFT using 1024 points sampled for 502 [s] gives Fig. 3.13b.
(9) Shift the external force f to the node x2 in Fig. 3.10. The excitation induces
no response of the second system, because x2 = 0 in the second mode. The
node x2 is therefore uncontrollable.
(10) For the same reason, a sensor y set at node x2 does not detect the resonance of
the second mode. The node x2 is therefore unobservable.
 
(11) The characteristic equation k2 M kD K  0 yields.

k f 0:12 j1 0:16 j2 0:1 j3 g and


.
f  Rek=Absk f 0:116 0:08 0:033 g
These exact values for are in good agreement with the assumed values in (3).

Note: Accuracy of modal response peak Fig. 3.12 shows three modal reso-
nance curves of Eq. (3.13), denoted by the ne lines, and the exact resonance curve
calculated by Eq. (3.14), denoted by the bolder line. It is clear that the two cal-
culation methods give almost the same peak values concerning the rst, second and
third modes, thus indicating the validity of the approximate diagonalization of the
damping matrix.
3.3 Modal Analysis of Beams 53

3.3 Modal Analysis of Beams

3.3.1 Natural Frequencies and Eigenmodes

Ignoring the complexity of the conguration of shaft shapes, an actual rotating shaft
may also be modeled simply as an equivalent uniform shaft (i.e., a continuum).
Vibration formulae of various continua are described in a number of textbooks.
The natural frequencies and eigenmodes of a beam are shown in Table 3.1 for
different boundary conditions, including graphical representations of the eigen-
modes, normalized by the maximum amplitude = 1, and corresponding modal
masses m*.
Let us comment on how to calculate the modal mass. Through the mode normal
function formulae, /n; k, given in Table 3.1 is the simplest form, their maximum
amplitudes are not always unity. If not unity, divide the given formula by the
maximal amplitude, which is searched by inserting possible values n into the given
formula. For example, in the case of free-free boundary condition on the top left
in Table 3.1(1), we must set the amplitude at both ends to unity. The normalized
mode function is then given by /n; k=/0; k and, therefore, the modal masses
are calculated as follows:

Z1

m qAl f/n; k=/0; kg2 dn ! m f 0:25 0:25 0:25 gqAl
0

3.3.2 Correspondence of the Modal Analyses for Multi-dof


Systems and Continua

The number of degrees of freedom of a continuum is innite; it therefore has an


innite number of natural frequencies and eigenmodes. The natural frequencies are
listed in the increasing order. The modes are mutually orthogonal. For example, the
rst and second eigenmodes are designated as k1 and k2 , and therefore the normal
functions are denoted by /1 n; k1 and /2 n; k2 , where n x=l is dimensionless
position along the axis. The orthogonality conditions are written as

Z1
/21 n; k1 dn 6 0
0
3:15
Z1
/1 n; k1 /2 n; k2 dn 0
0
54 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Table 3.1 Free vibration characteristics of beam


2 EI modal mass m* = m0 Al EI : flexural rigidity
frequency f = [Hz]
2l 2 A : mass density [kg/m3] l : length [m] A : cross-section area [m2]
boundary free-free Y (0) = Y (0) = Y (l ) = Y (l ) = 0 fixed-fixed Y ( 0 ) = Y ( 0 ) = Y (l ) = Y (l ) = 0
4.730 7.853 10.996 4.730 7.853 10.996
1.0 3rd 1.0 1st
2nd
0.5 1st 0.5 2nd
mode 0 0
0.5 1.0 3rd 0.5 1.0
0.5 0.5
1.0 1.0
normal cosh + cos sinh + sin cosh cos sinh sin
= =
function cosh cos sinh sin (1) cosh cos sinh sin (2)
m0 0.25 0.25 0.25 0.396 0.439 0.437
boundary pinned-free Y (0) = Y (0) = Y (l ) = Y (l ) = 0 fixed-pinned Y (0) = Y (0) = Y (l ) = Y (l ) = 0
3.927 7.069 10.210 3.927 7.069 10.210
1.0 1.0 1st
1st
0.5 0.5 2nd
2nd 3rd
mode 0 0
0.5 1.0 0.5 1.0
0.5 3rd 0.5
1.0 1.0
normal sinh sin cosh cos sinh sin
= + =
function sinh sin (3) cosh sinh (4)
m0 0.25 0.25 0.25 0.439 0.437 0.438
boundary pinned-pinned Y (0) = Y (0) = Y (l ) = Y (l ) = 0 fixed-free Y (0) = Y (0) = Y (l ) = Y (l ) = 0
2 3 1.875 4.694 7.855
1.0 1st 1.0 3rd 2nd 1st
0.5 2nd 0.5
mode 0 0
0.5 1.0 0.5 1.0
0.5 0.5
3rd
1.0 1.0
normal cosh cos sinh sin
= sin =
function (5) cosh + cos sinh + sin (6)
m0 0.5 0.5 0.5 0.25 0.25 0.25
boundary roller-roller Y (0) = Y (0) = Y (l ) = Y (l ) = 0 roller-fixed Y (0) = Y (0) = Y (l ) = Y (l ) = 0
2 3 2.365 5.498 8.639
1.0 1.0 1st
0.5 1st
0.5
3rd
mode 0 0
0.5 1.0 0.5 1.0
0.5 3rd 0.5
1.0 2nd 1.0 2nd
normal cosh cos
= cos =
function (7) cosh cos (8)
m0 0.5 0.5 0.5 0.396 0.437 0.438
boundary pinned-roller Y (0) = Y (0) = Y (l ) = Y (l ) = 0 free-roller Y (0) = Y (0) = Y (l ) = Y (l ) = 0
/2 3 / 2 5 / 2 2.365 5.498 8.639
1.0 2nd 1.0
1st 3rd
0.5 0.5 1st
mode 0 0
0.5 1.0 0.5 1.0
0.5 3rd 0.5
2nd
1.0 1.0
normal cosh + cos sinh + sin
= sin =
function (9) cosh + cos sinh sin (10)
m0 0.801 0.502 0.513 0.25 0.25 0.25
3.3 Modal Analysis of Beams 55

Considering these orthogonality conditions, the modal analysis of continuous


beams is similar to that of multi-dof systems as shown in Table 3.2. The modal
mass of a uniform beam of density q, cross-sectional area A, and length l, is

Z1

M qAl Ut Udn diagonal   ; mi ;   : 3:16
0

where the modal matrix : U  /1 n; k1 /2 n; k2    , and modal mass:


R1
mi  qAl 0 /2i n; ki dn
The modal analysis of multi-dof systems described in the preceding section is
thus valid also for continuous beams, as shown in Table 3.2. Inner products in the
former are replaced by integrals in the latter.

3.3.3 Reduced Modal Models

Consider any interval [l1, l2] in a continuous beam under a distributed excitation
force F(x, t) = B(x)w(t) and a displacement sensor y to measure the vibration at a
point x = l3 in the beam as shown in Fig. 3.14.
The response in the modal coordinate can be calculated for each gi , as shown in
Table 3.2, and is given by adding an appropriate assumption of the modal damping
ratio fi ;
 
mi gi t 2fi xi g_ i t x2i gi t bi wt i 1l 3:17
R l2
where the modal input coefcient bi  l1 Bx/i xdx.

Table 3.2 Modal analysis for multi-mass system and continuous system

Multi-degree of system : matrix Continuous system : differential equation


equation .. 2y 4y
MX+ KX= BF( t ) A 2
+ EI = F ( x, t )
of motion t x4
eigen pair n , n : normal mode n, n( x) : normal function
l

orthgonality
congruence it M j = ij
integral
0 Ai (x) j (x)dx = ij
transform. l
it K j = ij 0 EI''i (x) ''j (x)dx = ij
1

modal y ( x) = 1( x)1(t) + 2( x)2(t) + = X = [1 , 2 , ] 2
coordinate

modal l
parameters mi*= it M i , k i*= mi* i2 mi* = Ai2(x)dx, k i*= mi* i2
0

modal analysis mi* (i + 2 i ii + i2i ) = it BF (t ) mi* (i + 2 i ii + i2i ) = i ( x ) F ( x, t )dx


.. . .. . l
0
56 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Fig. 3.14 Distributed F (x,t) = B (x) w (t)


excitation of a beam
y (x,t)

l1
y
l2
l3
l

The response, y, of the displacement sensor is the sum of modal vibration


responses, gi :

X
n
yt /i l3 gi t 3:18
i1

where /i (i = l3) at x = l3 is the modal output coefcient to the sensor.

This leads to the modal model of the continuous beam, as shown in Fig. 3.15. The
number of modes must, evidently, be truncated at an appropriate modal number n.
This model for the continuum is essentially identical to Fig. 3.8 for the multi-dof

mode
truncation
w (t)

l l l
l12 B ( x ) 1( x ) dx l12 B ( x ) 2( x ) dx l12 B ( x ) n ( x ) dx agitation
factor

m 1 m 2 mn

1 2
1 1 2 2 n n n

modal
1( l 3 ) 2( l 3 ) n( l 3 )
output
+ +
+ + y (t)

Fig. 3.15 Modal model


3.3 Modal Analysis of Beams 57

system. Instead of the time domain, the modal analysis gives the response of the
displacement sensor in the s-domain of Laplace transformation:
R l2
X
n / l
i 3 l1 Bx/i xdx
ys ws 3:19
i1
mi s2 21i xi s x2i

3.3.4 Modal Eccentricity

When a uniform rotating shaft has distributed mass eccentricity ex shown in


Fig. 3.16, the distributed excitation force per unit length F(x, t) can be written as

Fx; t q1 xexX2 ejXt 3:20

where q1 x is the shaft mass per unit length.


The modal unbalance U* for the mode / is

Z1 Z1

U q1 xex/xdx l q1 nen/ndn n x=l 3:21
0 0

Dividing this by the modal mass m* gives the modal eccentricity e :

Z1
 1
e  q1 nen/ndn 3:22
m
0

The modal unbalance vibration, , corresponding to a mode, / {natural frequency


xn , modal damping ratio f}, can be written using the complex amplitude, Ag , as

e X 2 
g Ag ejXt ! Ag  3:23
s 2fxn s xn sjX
2 2

(a) (b) (c) (d)


(x) 0 0 1

90

180

Fig. 3.16 Unbalance distribution


58 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Example 3.5 Consider the unbalance distributions shown in Fig. 3.16 acting upon
the shaft eigenmode under the pinned-pinned boundary (Table 3.1(5)). Figure 3.16
includes four cases: (a) uniform distribution en 1, (b) uniformly twisted dis-
tribution en 1ejpn=2 between 0 and 90, (c) uniformly twisted distribution
en 1ejpn between 0 and 180, and (d) triangular distribution en n. Find the
modal eccentricity e for the rst three eigenmodes for each case.
Answer
Considering that the eigenmodes, U  /1 /2 /3  sin pn sin 2pn
R1
sin 3pn and modal masses, mi qAl 0 /2i ndn f 0:5 0:5 0:5 gqAl,
R1
(a) e 1=mi 0 qAlen/i ndn f 4=p 0 4=3p g f 1:27 0 0:42 g
(b) e f 1:2\45
0:48\45
0:31\45
g
(c) e f 1\90
0:85 0 g
(d) e f 0:64 0:32 0:21 g

Example 3.6 As shown in Fig. 3.17, consider a uniform steel beam, 10 mm in


diameter and 500 mm in length, with simple supports at both ends,
impulse-excitation at l1 =l 1=4 and shaft vibration measurement at l3 =l 5=6.
(1) Find the rst three natural frequencies and recheck the corresponding
parameters of Fig. 3.18.
(2) Find the sensor response, y, in terms of the s-domain of the Laplace
transformation.

(a) impulse (b)


1.0 1
f (x) 0.5
0
0.5 1.0
l sensor 0.5
l1= 4 3 2
5l
l3= 6
y 1.0

(c) (d)
20 dB
1 T = 1 [s]
30 N = 2 048
1 2
40 3

10 20 30 40 50 50
[ms]
1 60

0 100 200 300 400 500 600 700 800


[Hz]
waveform FFT

Fig. 3.17 Impulse test


3.3 Modal Analysis of Beams 59

impulse
(t)

0.71 1 0.71 B

1/2 1/2 1/2 m i Al


1 = 2 = 3 =
80 Hz 1 320 Hz 2 720 Hz 3

1 = 0.08 2 = 0.05 3 = 0.015

C 0.5 0.87 1

++ ++ y (t)

Fig. 3.18 Modal model (Impulse)

(3) Assuming the modal damping ratios f = {0.08 0.05 0.015} for the rst three
modes, derive the impulse response waveform of Fig. 3.17c.
(4) Perform the frequency analysis to obtain Fig. 3.17d.

Answer
(1) The natural frequencies of a beam with simple supports at both ends:
xn = {80 320 720} Hz follow by reference to Table 3.1(1).
Eigenmode functions: U sin pn=2 sin pn sin 3pn=2  (see Fig. 3.17b).
R1
Modal masses: mi qAl 0 sin ipn2 dx qAl=2 ! qAlf 1=2 1=2 1=2 g
Modal input coefcient = mode displacement at the excited point:

B fsin ip=4g f 0:71 1 0:71 g

Sensor output coefcient = mode displacement at the sensor point:

C  fsin 5ip=6g f 0:5 0:87 1 g

These parameters are seen in Fig. 3.18.


(2) Substituting the impulse excitation w(s) = L[dt] = 1, obtained by the
Laplace transform of Dirac delta function dt, into Eq. (3.19) yields

5pi ip
X
3 sin sin
ys 6 4 3:24
qAl 2
i1 s 2fi xi s x2i
2
60 3 Modal Analysis of Multi-Degree-of-Freedom Systems

(3) The inverse Laplace transform of Eq. (3.24) gives the waveform of Fig. 3.17c
as the impulse response.
(4) The FFT with N = 2048 samples of the waveform for a time window T = 1 s
and displaying 800 lines gives Fig. 3.17d. The plot shows spectral peaks at
{80, 320, 720} Hz corresponding to the natural frequencies. The FFT of
impulse response waveforms is often used in the practical design process
because it permits quick identication of the layout of all natural frequencies.

3.4 Physical Models from Reduced Modal Models

3.4.1 Modal Mass

The modal mass mi* corresponding to the ith eigenvector /i is dened as follows:

mi /ti M/i m1 /2i1 m2 /2i2 m3 /2i3    3:25

where, for a 3-dof system, the mass matrix and eigenvector have the forms
M diagonal m1 m2 m3 ; /i ui1 ui2 ui3 t .
What is important about the eigenvector is not its element values, but the mutual
ratio between element values. In fact, their values are commonly determined from
two types of normalization:
(I) the maximum value in the eigenvectors /i is unity, and
(II) the modal mass /ti M/i 1, that is normalized by the mass matrix.
In the cantilever shown in Fig. 3.19, column is the rst mode /1 normalized
to the maximum magnitude = 1 at the top of the building, providing the modal
mass qAl=4. Dividing /1 (column ) by the square root of the modal mass qAl=4
p
(or scaling by 2= qAl) yields the values shown in column of /1 as normalized
by the mass matrix.
Example 3.7 Concerning the second mode /2 of the cantilever shown in Fig. 3.19,
column is normalized by the maximum values = 1. Conrm that column is
normalized by the mass matrix.
Answer
The modal mass is also qAl=4 for the second mode /2 by referring to xed-free of
p
Table 3.1. The values in column are obtained by scaling column by 2= qAl.
As shown in Fig. 3.8, the modal model is represented by a set of single-dof
systems corresponding to each mode. Assuming that the ith mode has the dis-
placement = 1 at node j, the physical coordinate xj(t) of the node displacement is
3.4 Physical Models from Reduced Modal Models 61

1 2 1 2
a 1 1.000 1.000 2.000 2.000 a
2
top of 2 0.862 0.524 1.725 1.047 top of

beam 3 0.725 0.070 1.451 0.140 beam

4 0.591 0.317 1.182 0.634
1
b 5 0.461 0.590 0.922 1.179 b

mid- 6 0.340 0.714 0.679 1.427 mid-
point
7 0.230 0.684 0.460 1.367 point

8 0.136 0.526 0.273 1.052
9 0.064 0.301 0.128 0.602
1
10
0.017 0.093


0.034 0.185 Al
1 2 3 4
max amp. = 1 modal mass = 1

Fig. 3.19 Normalization of eigenmodes

approximately equal to the ith modal coordinate gi t as far as frequencies near the
natural frequency xi are concerned:

xj t  u1j g1 t u2j g2 t    uij gi t     uij gi t gi t 3:26

where uij 1. This applies to each mode (i = 1,, l).


The ith modal mass of the corresponding single-dof system can thus be regarded
as a mass physically equivalent to the ith mode seen from the node j. Thus the
correspondence between a modal mass and a physically equivalent mass can be
used so long as the ith mode vibration at the natural frequency xi is predominant.
For example, the equivalent masses for the cantilever shown in Fig. 3.19 are as
follows:
Equivalent masses seen from the tip node 1 are:
rst mode m1eq qAl=4 0:25qAl and
second mode m2eq qAl=4 0:25qAl
Equivalent masses seen from the midpoint node 5 are:
rst mode m1eq qAl=41=0:4612 1:18qAl and
second mode m2eq qAl=41=0:5902 0:72qAl
This equivalent masses can be measured by a technique, termed the equivalent
mass method, as described in the following section.
62 3 Modal Analysis of Multi-Degree-of-Freedom Systems

3.4.2 Equivalent Mass Method

The equivalent mass was introduced by Seto [12] who used the reduced order
physical model in an experimental manner to identify the equivalent mass corre-
sponding to an eigenmode.
First, the natural frequency of the ith mode xi is measured; then a weight Dm is
added to a node station number j at which this mode presents a large amplitude. The
lowered natural frequency xi is then measured. The two values are related by the
approximate equation
s r

 keq meq Dm
xi xi  xi 1  3:27
meq Dm meq Dm 2meq

The equivalent mass is therefore identied by

xi Dm
meq 3:28
xi  xi 2

In practice, data pairs {xi ; Dm} are collected and plotted as in Fig. 3.20. The
unit of frequency xi may be either [rad/s] or [Hz].
For higher precision, the gradient a of the tted line is used for identication:
xi xi
a ! meq kg for jth node station 3:29
2meq 2a

This equivalent mass meq obtained for the jth node station can be transformed to
another node station k, by measuring the amplitudes simultaneously at the nodes
j and k (aj and ak, respectively) in a resonance mode and converting as

Equivalent mass meq aj =ak 2 kg for kth node station 3:30

4.592
4.58
m
0.01
4.56 m
f [Hz]

4.54 0.03 k
n

1/ f n
4.52 =3

4.5
0 0.01 0.02 0.03
m [kg]

Fig. 3.20 Evaluation of an equivalent mass


3.4 Physical Models from Reduced Modal Models 63

Example 3.8 The data shown in Fig. 3.20 represent the decrease of 1st mode
natural frequency by adding mass at the tip node no = 1 of the cantilever shown in
Fig. 3.19.
(1) Find the modal equivalent mass.
(2) Convert the result to the equivalent mass for the midpoint node no = 5.

Answer
(1) a = 3 [Hz/kg] and x1 = 4.592 [Hz] yields meq = x1 =2a = 0.765 [kg]
(2) meq = 0.765(1/0.461)2 = 3.6 [kg]

3.5 Approximation of Natural Frequencies

3.5.1 Rayleighs Method

Suppose that, the eigenvalue problem for a multi-dof M-K system is

x2 M/ K/ 3:31

where M = Diagonal[m1 m2 m3 ] and K is a positive denite symmetric matrix,


with
eigenvalues x2 : x21 ; x22 ; x23 ;    and
eigenvectors / : /1 ; /2 ; /3 ;   
as solutions. Since the eigenvectors are mutually orthogonal with respect to the
mass matrix M and stiffness matrix K, the eigenvalue for the ith mode is given by
the Rayleigh quotient

/ti K/i
x2i 3:32
/ti M/i

If approximate eigenvectors are known, the eigenvalues may be obtained from


Eq. (3.32) to a precision dependent on the precision of the eigenvectors.
It is common practice to estimate the lowest-order natural frequency by
assuming the deflection due to gravity as the rst eigenmode. Since the deflection
due to gravity y y1 ; y2 ; y3 ;    may be calculated by

Ky m1 m2 m3    t g
64 3 Modal Analysis of Multi-Degree-of-Freedom Systems

where g is the acceleration of gravity, Eq. (3.32) can be rewritten as


P
yt Ky yt m1 m2 m3    t g g m i yi
x21 t P 3:33
y My yt My m i yi 2

Since any vector /a is a linear combination of eigenvectors normalized by the


mass matrix (/t M/ 1; /t K/ x2 ) :

/a c1 /1 c2 /2 c3 /3    3:34

the corresponding quadratic forms, considering the orthogonality condition of


eigenvectors, are
/ta K/a c21 x21 c22 x22 c23 x23   
3:35
/ta M/a c21 c22 c23   

If this vector /a is close to the kth eigenvector /k , i.e., /a  /k , the Rayleigh


quotient is

/ta K/a c21 x21 c22 x22 c23 x23    c2k x2k
x2a
/ta M/a c21 c22 c23    c2k

In other words,

c21 x21 c22 x22 c23 x23


2 2  1
c2 x2 c2k x2k ck x k
x2a x2k k 2 k 2 x2k for k max
c1 =ck c2 =ck c23 =c2k    1
2 2
3:36
c22 x22 c23 x23 c2 x 2
1 2 2
2 2    k2 k2
c1 x1 c1 x1 c1 x 1
x2a x21 x21 for k1
1 c22 =c21 c23 =c21    c2k =c21

Equation (3.36) is suitable for estimation of the kth maximum eigenvalue


because it converges rapidly when xi xk and ci ck  1. It is clear that the
estimated value may be somewhat lower than the true value. On the other hand,
the rst (k = 1) eigenvalue may be somewhat higher than the true value.
Example 3.9 For a system shown in Fig. 3.21,
(1) Find the exact eigenvalues and eigenvectors.
(2) Assuming that the rst eigenmode is the deflection due to gravity, estimate the
eigenvalue.
(3) Assuming the third eigenmode /a 1 1 1 t , estimate the corresponding
eigenvalue.
3.5 Approximation of Natural Frequencies 65

Fig. 3.21 3-dof system 20 60 18 18


20 6 9

1.5 1.5 1 12 = 1
1 22 = 4
0
0.3
1.5 1 32 = 16

Answer

2 3 2 3
20 0 0 80 60 0
6 7 6 7
M 4 0 6 0 5; K 4 60 78 18 5 ! x2 f 1 4 16 g;
0 0 9 0 18 36
(1) 2 3
1:5 0:3 1:5
6 7
U 4 1:5 0 6 5
1 1 1

(2) Using Eq. (3.33),


2 3 2 3 2 3
1 1 67
g
Ky M 4 1 5g ! y K 1 M 4 1 5g 4 68 5
64
1 1 50

20 67 6 68 9 50
) x2a 64 1:0046 x21 1 true value
20 672 6 682 9 502

(3) Using Eq. (3.36),

20 12 60 22 18 22 18 12


) x2a 10\x23 16 true value
20 6 9

3.5.2 Method Using Influence Coefcients

The inverse matrix of the stiffness matrix K is called the influence coefcient
matrix: a K 1 . A general eigenvalue problem can be so reconstructed that the
matrix of eigenvalue problem is symmetric by use of the following transformation:
66 3 Modal Analysis of Multi-Degree-of-Freedom Systems

1 1
x2 M/ K/ ! / K 1 M/ ! 2 M/ MK 1 M/ 3:37
x2 x

Solving this yields

eigenvalues1=x2 : 1=x21 1=x22 1=x23 . . .; and

eigenvectors/: /1 /2 /3 . . .

The eigenvectors are the reciprocal of the Rayleigh quotients:

/ti M/i /ti M/i


x2i 3:38
/i MK 1 M/i
t t
/i MaM/i

If a vector /a is an approximation of the lowest order eigenvector as shown in


Eq. (3.34), the Rayleigh quotients yield

/ta M/a c21 c22 c23   


x2a
/a MK 1 M/a c21 =x21 c22 =x22 c23 =x23   
t
3:39
c21 c22 c23   
x21 2 x21
c1 c22 x21 =x22 c33 x21 =x23   

where /a  /1 .
It should be noted that Eq. (3.39) is suitable for estimation of the minimum
eigenvalue x1 , and that the estimation is somewhat greater than the true value, as
stated in Example 3.9(2).
Example 3.10 Assuming that the rst eigenmode of the system shown in Fig. 3.21
is the deflection due to gravity, evaluate the eigenvalue by Eq. (3.38).
Answer

2 3 2 3 2 3
23=640 1=32 1=64 67 1:34
6 7 g 6 7 6 7
K 1 4 1=32 1=24 1=24 5 ! y 4 68 5 ! y 4 1:36 5
64
1=64 1=24 11=288 50 1
t
y My
x2a t 1:0012
y MK 1 My
3.5 Approximation of Natural Frequencies 67

3.5.3 Dunkerleys Formula

An eigenvalue problem involving an influence coefcient matrix [a] can be written


as
2 32 3
a11 a12 a13  m1 0 0 
6 7 6 0 7
1 6 a21 a22 a23 76 m2 0 7
/ aM/ 6 a31 a32 a33 7 6 7 / 3:40
x 2 4 54 0 0 m3 5
.. .. .. .. .. .. .. ..
. . . . . . . .

The corresponding characteristic equation is


 
 a11 m1  1=x2 a12 m2 a13 m3    

 a22 m2  1=x2    
 a21 m1 a23 m3
 
 a31 m1 a32 m2 a33 m3  1=x2    

 .. .. .. . .  3:41
 . . . .

n
n1
1 1
a11 m1 a22 m2 a33 m3       0
x2 x2

This is an nth order polynomial equation with respect to the root 1=x2 . Since
the coefcient of the (n 1)th order term multiplied by 1 is the sum of the
characteristic roots if the coefcient in the nth order term is 1, the following
equation, termed the root formula, holds:

1 1 1
   a11 m1 a22 m2 a33 m3    3:42
x21 x22 x23

If x1 x2 \x3   , the left side of Eq. (3.42) can be approximated to the rst
term. The term aii mi in the right side is the reciprocal of the eigenvalue x2ii of a
component system where mi is the only mass in the system. The lowest-order
natural frequency x1 can thus be estimated by

1 1 1 1
 3:43
x21 x211 x222 x233

It is clear that this estimation is somewhat smaller than the true value.
Example 3.11 Estimate the natural frequency of a system shown in Fig. 3.22 using
Dunkerleys formula.
Answer
The eigenvalues for each component system in Fig. 3.22 are listed below, where //
represents the series spring:
68 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Fig. 3.22 Each independent 20 60 18 18


system 20
112 = 32/23

20 60 18 18
6
222 = 4

20 60 18 18
9
332 = 32/11

x211 20 60==18==18=20 640=23=20 32=23


x222 20==60 18==18=6 24=6 4
x233 20==60==18 18=9 288=11=9 32=11

The estimate is thus: x21 1=23=32 1=4 11=32 0:762.


(Note that this estimation is considerably smaller than the true value = 1 of
Fig. 3.21.)

3.5.4 Iterative Method (Power Method) [B4]

In a eigenvalue problem concerning the coefcient matrix is A,

k/ A/ 3:44

Any vector, /a , may be written as a linear sum of eigenvectors, as indicated in


Eq. (3.34). Performing the multiplication of the coefcient matrix A to the right side
of the above equation gives

A/a c1 k1 /1 c2 k2 /2 c3 k3 /3   
A2 /a AA/a c1 k21 /1 c2 k22 /2 c3 k23 /3   
..
.
A /a c1 kn1 /1 c2 kn2 /2 c3 kn3 /3   
n

If k1 is the maximum eigenvalue,

An /a kn1 c1 /1 c2 k2 =k1 n /2 c3 k3 =k1 n /3    kn1 c1 /1 3:45


3.5 Approximation of Natural Frequencies 69

Fig. 3.23 Power method


in = in A in = 2
out
1 1

in = out

the coefcients of ki =k1 n converge to zero as n ! ; the right side of Eq. (3.45)
therefore converges to the eigenvector, /1 . Thus, assuming any eigenvector ini-
tially, and repeating the multiplication of the coefcient matrix, A, as above, gives
convergence to the eigenvector corresponding to the maximum eigenvalue.
In practice, as shown in Fig. 3.23, a normalized input eigenvector /in is given.
The normalization is made as a certain element is unity, e.g., the last element = 1 in
this gure. The initial /in is multiplied by the coefcient matrix. The vector obtained
is normalized each time to be checked its convergence. The convergence resulting
from this repetition automatically gives the eigenvector and eigenvalue x2 .
Note that an algorithm with A M 1 K gives the convergence to the maximum
eigenvalue, otherwise with A K 1 M convergence is to the minimum eigenvalue.
Example 3.12 Find the minimum natural frequency of the system shown in
Fig. 3.21 by the iteration method.
Answer

2 3
46 12 9
1 6 7
A K 1 M 4 40 16 12 5 ! k/ A/ k 1=x2
64
20 8 22
2 3 2 3 2 3 2 3 2 3 3:46
1 1:34 1:34 1:4 1:5
6 7 6 7 6 7 6 7 6 7
A4 1 5 0:784 1:36 5 ! A4 1:36 5 0:9334 1:46 5    ! 1:04 1:5 5
1 1 1 1 1

If an eigenvalue, k1 , and eigenvector, /1 , are obtained in an eigenvalue problem


of nth order, an (n 1)th order eigenvalue problem which gives another eigenvalue
can be constructed as follows. For a space / x1 ; x2 ; x3 ;   t spanned by
n eigenvectors, if one of eigenvectors /1 u11 ; u12 ; u13 ;   t is known, the
dimension is reduced by one by imposing the condition of orthogonality of these
two vectors concerning the mass matrix.

/t1 M/ u11 u12 u13    M x1 x2 x3    t 0 3:47

The specic procedure is seen in the following example.


70 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Example 3.13 Starting from the minimum eigenvalue and eigenvector for Example
3.12 (three-dimensional eigenvalue problem), derive the two-dimensional eigen-
value problem to obtain the next smallest eigenvalue.
Answer
Equation (3.47) becomes in this case

/t1 M/ 1:5 1:5 1 M x1 x2 x3 t 30x1 9x2 9x3 0 3:48

Thus the coordinate transformation, T32, for reduction from the


three-dimensional space {x1, x2, x3 } to the two-dimensional space {x1, x2} is
dened as
2 3 2 3
x1 9=30 9=30    
4 x2 5 4 1 5 x2 x
0  T32 2 3:49
x3 x3
x3 0 1

The congruent transformation of Eq. (3.44) using this transformation matrix T32
yields
  
x2 x2
t
kT32 T32 T32
t
AT32 !
x3 x3
       3:50
x2 x2 1=16 0 x2
k T32
t
T32 1 T32
t
AT32
x3 x3 1=32 1=4 x3

Solving this eigenvalue problem gives the eigenvalue k = {1/4, 1/16}, and
therefore x2 1=k f4; 16g.

3.5.5 Stiffness Matrix Method

An elastic shaft element (Fig. 3.24) with having density q = 0, length L and flexural
rigidity EI has four nodal displacement variables: bending deflections d1 ; d2 and
tilting angles h1 ; h2 . The deflection mode shape between the nodes is therefore
described by a three-dimensional polynomial equation, i.e., a cubic curve.
Assuming that the node masses are m1 and m2 and the transverse moments of inertia
Id1 and Id2, and dening the shearing forces Q and moments N at the ends, with
corresponding indices, as shown in the gure, the equation of motion for a beam
element is written as follows:
3.5 Approximation of Natural Frequencies 71

Fig. 3.24 Beam FE model 1 2


EI
N1 m1 = 0 m2 N2
Q1 I d1 I d2 Q2

L
1 i 2
1 i 2

2 32 3 2 32 3
m1 0 0 0 d1 12 6L 12 6L d1
6 0 Id1 0 0 7 6 h1 7 EI 6 6L 4L2 6L 2L2 7 6 h1 7
6 76 7 6 76 7
4 0 0 m2 0 54 d2 5 L3 4 12 6L 12 6L 54 d2 5
0 2 0 30 Id2 h2 6L 2L2 6L 4L2 h2
Q1
6 N1 7
6 7
4 Q2 5 3:51
N2

The rst diagonal matrix term of Eq. (3.51) is called the mass matrix and the
second the stiffness matrix. It is basically the same concept for the discretization by
the 3-D FEM (nite element method).
Consider now a system containing two disks (m1 = m, m2 = 2 m, Id1 = Id2 = 0)
as shown in Fig. 3.25. The equation of motion for the entire system, with the
displacement of the ith node ( di and hi i = 1 to 4) as the state variables, is obtained
by superposing the mass and stiffness matrices of three elements:

EI
+ 3
0 l 12 6 12 6 0 0 0 0 1
0 6 4 6 2 0 0 0 0 l1

m1 2 12 6 24 0 12 6 0 0 2
( I / l 2 )l 6 2 0 8 6 2 0 0 l 2
d1 2
3:52
m2 3 0 0 12 6 24 0 12 6 3
( I / l 2 )l 0 0 6 2 0 8 6 2 l3
d2 3

0 0 0 0 0 12 6 12 6 4

0


0 0 0 0 6 2 6 4 l 4
= [ Q1 N4 / l]
t
+ N1 / l 0 0 0 0 Q4

Applying the boundary conditions d1 d4 N1 N4 0 and Idi = 0, the


equations of the rst and seventh rows of Eq. 3.52 are eliminated. Change of the
order of the state variables leads to
72 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Fig. 3.25 An inboard rotor 1 2 3 4

EI
m1 m2
m 2m
L
l l l
1 2 3 4
1 2 3 4

m1 2 24 12 6 0 6 0 2 0

m2 3 12 24 0 6 0 6 3 0
0 EI 6 0 4 2 0 0 l1 0
+ 3 = 3:53
0 l 0 6 2 8 2 0 l 2 0
0 0 0 0 2 8 2 l3 0

0 0 0 0 0 2 4 l 4 0

d2 and d3 are called master coordinates; h1 to h4 slave coordinates. The slave


coordinates are interpolated from variations of the master coordinates. Dividing the
stiffness matrix, K66, in Eq. (3.53) into four sub-matrices:

24 12 6 0 6 0
12 24 0 6 0 6

EI 6 0 4 2 0 0 EI K 22 K 24
K 66 3 K 66 3 t 3:54
l 0 6 2 8 2 0 l K 24 K 44
6 0 0 2 8 2

0 6 0 0 2 4
the equation becomes
1 t
lh1 lh2 lh3 lh4 t K44 K24 d2 d3  t : 3:55

Substituting this into the equation of motion in the master coordinates yields
 
1 t 6 8 7
K22  K24 K44 K24 3:56
5 7 8

Considering that l = L/3, the two-dimensional equation of motion is written in


the form:
     
m1 0 d2 162EI 8 7 d2
d3 0 3:57
0 m2 5L3 7 8 d3
3.5 Approximation of Natural Frequencies 73

Solving the corresponding eigenvalue problem yields the eigenvalue x2 = El/


(mL3) [21.4, 367.4].

3.5.6 Transfer Matrix Method

Consider a basic transfer element consisting of an elastic shaft, L in length and EI in


flexural rigidity, with a disk m in mass and Id in transverse moment of inertia at the
right end (Fig. 3.26). Taking Eq. (3.51) into account, the equation of motion for
this system in s domain is written as
0 2 3 2 312 3 2 3
0 12 6L 12 6L d1 Q1
B 6 0 7 EI 6 6L 4L2 6L 2L2 7 C6 7 6 7
Bdiagonal6 2 7 6 7C6 h1 7 6 N1 7
@ 4 ms 5 L3 4 12 6L 12 6L 5A 4 d2 5 4 Q2 5
Id s2 6L 2L2 6L 4L 2
h2 N2
3:58

Letting s jxn to obtain the natural frequency, the equation can be rewritten to
represent the relationship between the state variables of the left end
f d1 h1 N1 Q1 g and those of the right end f d2 h2 N2 Q2 g:
2 3 2 3 32
d2 1 L L2 =2EI L3 =6EId1
6h 7 6 L=EI L2 =2EI 76 h 7
6 27 6 0 1 76 1 7
6 76 76 7
4 N2 5 4 0 Id x2n 1  LId xn =EI L  L Id xn =2EI 54 N1 5
2 2 2

Q2 mx2n Lmx2n L2 mx2n =2EI 1 L3 mx2n =6EI Q1

 Tm EI; L; m; Id d1 h1 N1 Q1 t
3:59

For the system of Fig. 3.25, the transfer functions are thus
Tm12 = Tm(EI, L/3, m, 0) for the section ,
Tm23 = Tm(EI, L/3, 2m, 0) for the section , and
Tm34 = Tm(EI, L/3, 0, 0) for the section .

Fig. 3.26 TM model of a 1 2


beam element
N1 m N2
EI
Q1 Q2
Id
L
1 i 2
1 i 2
74 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Using the transfer functions, the entire system is represented by


2 3 2 3 2 32 3
d4 d1 t11 t12 t13 t14 d1
6 h4 7 6 h1 7 6 t21 t22 t23 t24 76 h1 7
6 7 Tm34  Tm23  Tm12 6 7  6 76 7 3:60
4 N4 5 4 N1 5 4 t31 t32 t33 t34 54 N1 5
Q4 Q1 t41 t42 t43 t44 Q1

Applying the boundary conditions d1 d4 N1 N4 0 to Eq. (3.60), it is


found that
    
0 t t14 h1
12 3:61
0 t32 t34 Q1

should be satised. Therefore, setting X L3 mx2 =EI, the characteristic equation


is


 2
L2 1 X2 
   1 2X X 
2X

 t12 t14   81 39366
EI 6 2187 2125746 

 t32 t34   X2  0:
2 2 X X 
 mLx 3 729 1
27 39366 
3:62

Solving this equation yields the eigenvalues x2 EIml3 f21:4; 367:4g.


Example 3.14 Assuming that the disk in Fig. 3.27 has a mass m and a transverse
moment of inertia Id mr 2 , where r = L/15, conrm that the natural frequency xa
with the rotational inertia Id taken into account is lower than the hypothetical natural
frequency x0 where only the mass is considered, by
(1) the stiffness matrix method,
(2) the transfer matrix method,
(3) the influence coefcient method (Eq. 3.38, assuming the deflection due to
gravity), and
(4) Dunkerleys formula.

Answer
(1) Applying the boundary conditions for the xed left end and free right end, i.e.,
d1 h1 N2 Q2 0, to Eq. (3.51),

Fig. 3.27 Cantilever


EI m
Id
L
1 2
1 2
3.5 Approximation of Natural Frequencies 75

     
m 0 d2 EI 12 6L d2
0 Id h2 L3 6L 4L2 h2
0

Therefore

2:97EI 3EI
x2a \x20 
mL3 mL3

(2) Applying the boundary conditions to the left and right transfer functions
obtained from Eq. (3.59),
2 3 2 32 3
d2 1 L L2 =2EI L3 =6EI 0
6 h2 7 6 0 1 L=EI L2 =2EI 76 0 7
6 76 76 7
405 4 0 Id x2n 1  LId x2n =EI L  L2 Id x2n =2EI 54 N1 5
0 mx2n Lmx2n L2 mx2n =2EI 1 L3 mx2n =6EI Q1

Solving the characteristic equation


 
 1  LId x2 =EI L  L2 Id x2 =2EI 
 n n 0
 L2 mx2 =2EI 1 L3 mx2n =6EI 
n

yields the same answer as in (1) above.


(3) The influence coefcient matrix is
 1  
L3 12 6L L3 2 3=L
a
EI 6L 4L2 6EI 3=L 6=L2

Using the deflection due to gravity,


   
m mgL3 2
y a g
0 6EI 3=L

Therefore

2:97EI 3EI
x2a \x20 
mL3 mL3

(4) When the mass only is considered,

1 3EI
x20
a1; 1 mL3

When the rotational inertia only is considered,


76 3 Modal Analysis of Multi-Degree-of-Freedom Systems

Fig. 3.28 An outboard rotor 1 2 3 4


EI
m1 m2

L
l l l
1 2 3 4
1 2 3 4

1 EI 152 EI 1 2:96EI
x2d ! x2a
a2; 2Id Id L mL3 1 1 mL3
2
2
x0 xd

Example 3.15 Find the natural frequency of the two-mass system shown in
Fig. 3.28 (m1 = m2 = m) by
(1) the stiffness matrix method,
(2) the transfer matrix method,
(3) the influence coefcient method, assuming the deflection due to gravity, and
(4) Dunkerleys formula.

Answer
(1) The equation of motion concerning eight state variables includes the same
stiffness matrix of Eq. (3.52) and the mass matrix = Diagonal{0, 0, m1, 0, 0, 0,
m2, 0}. Eliminating two rows corresponding to d1 = d3 = 0, the resulting
equation of motion, reduced to six state variables is
2 3 2 32 3 2 3
0 4 6 2 0 0 0 lh1 0
6 m1 d2 7 6 6 24 0 6 0 0 7 6 d2 7 6 0 7
6 7 6 76 7 6 7
6 0 7 EI 6 2 0 8 2 0 0 7 6 7 6 7
6 7 6 76 lh2 7 6 0 7
6 0 7 l3 6 0 6 2 8 6 2 7 6 lh3 7 6 0 7
6 7 6 76 7 6 7
4 m2 d4 5 4 0 0 0 6 12 6 54 d4 5 4 0 5
0 0 0 0 2 6 4 lh4 0

or, rearranged,
2 3 2 32 3 2 3
m1 d2 24 0 6 0 6 0 d2 0
6 m2 d4 7 6 0 12 0 0 6 6 76 d4 7 6 0 7
6 7 6 76 7 6 7
6 0 7 EI 6 6 0 4 2 0 0 7 6 7 6 7
6 7 6 76 lh1 7 6 0 7
6 0 7 l3 6 0 0 2 8 2 7
0 76 lh2 7
6 6 7
6 7 6 7 607
4 0 5 4 0 6 0 2 8 2 54 lh3 5 405
0 0 6 0 0 2 4 lh4 0
3.5 Approximation of Natural Frequencies 77

Partitioning the stiffness matrix into master coordinates of d2 ; d4 and slave


coordinates of h1 ; h2 ; h3 ; h4 ) gives
   
24 0 6 0 6 0
K22 ; K24 ;
0 12 0 0 6 6
2 3
4 2 0 0
62 2 07
6 8 7
K42 K24
t
; K44 6 7
40 2 8 25
0 0 2 4

The equivalent shaft stiffness is


 
1 4 12 3
K22  K24 K44 K42 :
5 3 2

For the two-mass system the equation of motion is:


     
m1 0 d2 32EI 12 3 d2
0
0 m2 d4 5L3 3 2 d4

Therefore, eigenvalues are

EI
x2 f 7:48 82:1 g
mL3

(2) Unsolvable because the boundary condition exits at the midpoint of node 3
and interferes with the transfer matrix operation.
(3) The influence coefcient matrix is
 1  
5L3 12 3 L3 1=48 1=32
a
32EI 3 2 EI 1=32 1=8

Using the deflection due to gravity


   
m mgL3 1
y a g ;
m 96EI 9

7:66EI
the rst natural frequency x21 is obtained from use of Eq. (3.38).
mL3
1 3EI
(4) For a system with m1 only, x211 and
a1; 1m mL3
1 8EI
for a system with m2 only, x222 ,
a2; 2m mL3
78 3 Modal Analysis of Multi-Degree-of-Freedom Systems

An approximate eigenvalue is given by Eq. (3.43):

1 48EI 6:86EI
x2a
1 1 7mL3 mL3
2
2
x11 x22

Note: Influence coefcient matrix An influence coefcient relates the beam


deflection due to a unit load. Therefore it can be calculated directly from a
deflection curve. Figure 3.29 shows an example of a shaft deflection function
yn; a; P, i.e., the deflection y measured at the position n under the weight P acting
on the position a. Using these deflection curves, we can obtain the influence
coefcient matrix.
In the case of Example 3.15, the influence coefcient matrix is obtained as
2

3
1 1 1 3
y
6 2 2; ; 1 y ; ;1 7  
6 2 2 7 L
3 1=48 1=32
a  6

7
4 3 1 3 3 5 EI 1=32 1=8
y ; ;1 y ; ;1
2 2 2 2

yielding the same answer as for (3).


In the case of Fig. 3.25, the influence coefcient matrix is obtained as shown
below, and the inverse matrix of which is identical to the stiffness matrix in
Eq. (3.57):
2

3
1 1 1 2
y
6 3 3; ; 1 y ; ;1 7  
6 3 3 7 L3 8 7
a  6

7
4 2 1 2 2 5 486EI 7 8
y ; ;1 y ; ;1
3 3 3 3
 
162EI 8 7
and the inverse matrix is K a1 .
5L3 7 8

Fig. 3.29 Deflection curves Ut 0t \0; Ut 1t 0


Chapter 4
Mode Synthesis and Quasi-modal Method

Abstract Modal analysis consists of superposing eigenmodes /i, weighted


according to the modal coordinates i, through all modes. The modal coordinates no
longer have the representation of the original physical coordinates. This chapter
discusses methods to reduce the order of the system while preserving the physical
coordinates as far as possible. One such method is based on Guyan reduction, in
which only the relatively important nodes are chosen out of numerous nodes in a
nely meshed model. The static deflection modes are developed to reduce the
system matrices. The reduced system consists of the physical coordinates of the
chosen nodes. Mode synthesis is another method. Here the most important nodes
are treated with the Guyan reduction method, while the other nodes are considered
as internal to the system and undergo modal analysis. Mode synthesis gives a model
containing both physical and modal coordinates. Since, however, the mass matrix
of this model is not diagonal, no equivalent model of the multiple mass system can
be derived. The quasi-modal method is a solution that gives a physical model
equivalent to the reduced model obtained by mode synthesis. A convenient model
providing an appropriate physical meaning is obtained. In addition, a procedure will
be presented in which the response of a bearing journal to a force acting on the rotor
is created by the mode synthesis model as a plant transfer function.

 
Keywords 1D-FEM Eigenvalue solution Master and slave designation Guyan 
  
reduction Mode synthesis model Deflection mode Quasi-modal model

4.1 Mode Synthesis Models

4.1.1 Why Mode Synthesis?

The steps to derive some reduced order models [13, 14] by using the numerically
exact nite element model are shown in Fig. 4.1. The modal model described in
Chap. 3 is obtained as a set of single-dof systems by conversion to the modal
coordinates corresponding to the normal eigenmodes of the M-K system. While the

Springer Japan 2017 79


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_4
80 4 Mode Synthesis and Quasi-Modal Method

Equivalent n-dof
1-dof system
system
Finite Element (FE) Model ( Equivalent mass )
(Seto Model)

M-K system Modal model Set of reduced


Normal modes ( Modal coordinates ) 1-dof systems
Forced deflection modes Guyan reduction (Boundary=Master)
at master coordinates

Forced deflection modes Quasi-modal model


Mode synthesis model
Eigen modes of (Boundary+Absolute
(Boundary+Modal coordinates)
internal system coordinates)

Fig. 4.1 System reduction technique

obtained reduced modal model is simple and practical, it has the following
difculties:
(1) It presupposes that there is no change in eigenmode by the input, e.g. in the
case of external excitation or feed-forward control. Then, to be precise, the
model is not applicable to problems in which the eigenmode of the system
may be influenced by a change of state variables or state feedback control.
(2) It contains modal coordinates only. The absence of physical coordinates
makes connection and coupling with adjacent vibration systems or control
systems inconvenient.
The Guyan reduction model and the constrained mode synthesis [14] dis-
cussed in this chapter are methods to overcome such difculties. The former is a
generalization of the approach of Eq. (3.54), i.e. division into master and slave
coordinates, and reduction to a system corresponding only to the master coordi-
nates. The latter gives a mixed system of important master nodes represented by
physical coordinates and other slave nodes (the internal system) represented by
modal coordinates.

4.1.2 Guyan Reduction Method

In actual vibration analysis, a modern automatic meshing capability provides a very


large number of elements to describe the precise system conguration. In fact, it is
most likely to be too large for vibration analysis. For example, a beam element of
FEM model (Fig. 4.2a, b of a rotor) is divided by many meshing nodes as repre-
sented by the white circles. This division is too ne to study a few modes of real
importance only. The Guyan reduction method involves choosing a few important
nodes, called master nodes, as indicated by the red lled circles in Fig. 4.2b.
All nodes on which reaction forces (bearing reaction force, feedback control force,
etc.) act should be assigned as master nodes. Nodes on which external forces (e.g.
unbalance) may act, but not necessarily so, could be master nodes.
4.1 Mode Synthesis Models 81

(a) sensor brg. (bearing) thrust brg. sensor


disk

Side 1 Side 2
(b)

1 1
(c)

1 1 2

Fig. 4.2 Selection of master nodes and forced deflection modes

The coordinates of the master nodes, X1 x11 x12    t ; are called boundary
coordinates, and those of the other nodes, X2 x21 x22    t , are internal
coordinates. The equation of motion of the M-K system under a boundary reaction
force vector, Q, and external force, F, can thus be written, by separating boundary
coordinates, X1, and internal coordinates, X2, as
         
M1 0 1
X K11 K12 X1 QX1 ; X2 F1 t
4:1
0 M2 X2 K t
12 K22 X2 0 F2 t

The damping term is ignored here for simplicity.


Here the static deflection mode, di, is introduced, as shown in Fig. 4.2c. The
static deflection mode, where one of the boundary coordinates is displaced by a unit
length (=1) while the others are constrained (=0), is determined rst, and then the
same procedure is applied to other boundary coordinates consecutively to obtain d1,
d2, . These are deflection modes due to imposing forced displacements at each
boundary coordinate, according to
    
K11 K12 E Keq
t 4:2
K12 K22 d 0

where E is a unit matrix and the deflection matrix, d d1 d2    ; is composed


of the columns of the static deflection modes.
The right hand side of Eq. (4.2) represents the force vectors needed to move a
boundary coordinates by a unit length, therefore, it contains spring constants,
specied here by equivalent shaft stiffness Keq. Solving this, the deflection matrix,
d, and equivalent shaft stiffness, Keq, are obtained:
82 4 Mode Synthesis and Quasi-Modal Method

1
d  K22 t
K12 E 4:3
1
Keq K11  K12 K22 t
K12 E 4:4

The following coordinate transformation is dened using these static deflection


modes:
2 3
1 0  2 3
  x11  
X1 60 1    76 7 E
6
4 ... .. 7 x
. . 54 12 5 X  Tg X1 4:5
X2 . . .. d 1
.
d1 d2 

Applying this to the equation of motion Eq. (4.1) and performing a congruent
transformation, the equation of motion containing boundary coordinates only is
obtained:

1 Keq X1 Q1 X1 ; X_ 1 F1 t dt F2 t
Md X 4:6

where, Md M1 dt M2 d is called the effective mass.


Since the boundary force Q1 is represented by physical coordinates in this
equation, the addition of nonlinear reaction forces or feedback forces acting to the
boundary via Q1 is straightforward.
This technique is frequently used in practical vibration analysis for the benet of
the signicant reduction of the number of nodes down to the master node without
losing the high precision. Note that the eigenvalues obtained from the Guyan
reduced model may be somewhat higher than those from the original model.
Example 4.1 Consider a spring kb acting on a cantilever as shown in Fig. 4.3a.
(1) Construct a single-dof Guyan model, shown in Fig. 4.3b, in which y1 of the
cantilever end is chosen as the only displacement.
(2) Find an approximate value of the natural frequency at the free end (kb = 0).

Answer
R1
(1) Since a static deflection of the beam d1 3  nn2 =2; Md qAl 0 d21 dn
0:24qAl; Keq 3EI=l3 , and superimposing the spring constant kb yields

Mdy1 Keq kb y1 0 4:7

Md is the effective mass sensed by the end of the cantilevered beam and it is
the same to the effective mass of the cantilever shown in Fig. 2.4.
(2) Approximate solution k2a 3=0:241=2 3:575  exact value k2e f3:516g;
referring to Table 3.1(6).
4.1 Mode Synthesis Models 83

Fig. 4.3 Cantilever and (a)


1-dof system 1
1 kb

(b) y1

0.24
3 EI l 3

Example 4.2 Assume that n x=l for the cantilever shown in Fig. 4.4.
(1) Construct a 2-dof model by Guyan reduction, involving displacements y1 and
y2 at the positions n = 1 and n = 0.5, respectively.
p
(2) Find an approximate value of the natural frequency, xa k2a =l2 EI=qA , at
the free end boundary.

Answer
(1) For Fig. 4.4a, where the cantilever is subject to load P at the free end and
supported simply at the midpoint, the deflection curve Y1 is
7 Pl3
Y1 n d1 n; and the deflection mode
96" EI
     #
8 1 1 3 1
d1 n 3n n 
2
5 n U n :
7 2 2 2
For Fig. 4.4b, where the cantilever is subject to load P at the midpoint and
supported simply at the end, the deflection curve Y2 is
7 Pl3
Y2 n d2 n; and the deflection mode
" EI
768
   #
8 2 1 3 1
d2 n n 9  11n 16 n  U n :
7 2 2

(a) (b)
l l
l/2 l/2
3P P 3 Pl P 7 Pl 3
2 16 768 EI
y2 =0 y1 =0
x x x

3 Pl y2
5P 1 y1
11P 5P
16 7 Pl 3 2
2 96 EI 16 16
Load at right, pinned at center Load at center, pinned at right

Fig. 4.4 Deflection curve caused by load P


84 4 Mode Synthesis and Quasi-Modal Method

Therefore, the equivalent mass matrix for the two-mass system is

Z1    
d21 d1 d2 0:137 0:088
M22 qAl dn qAl ;
d1 d2 d22 0:088 0:445
0

and the equivalent stiffness matrix


   
EI 96=7 5=16  768=7 96=7 240=7
K22 4:8
l3 5=2  96=7 768=7 240=7 768=7

(2) Calculation of the eigenvalues for this two-mass system yields the approxi-
mate solutions k2a f 3:52 22:3 g  exact values k2e f 3:516 22 g;
referring to Table 3.1(6).

Remarks This example shows that the precision of the Guyan reduction is
improved by increasing the number of master nodes.

4.1.3 Mode Synthesis Models

Let us designate the master nodes in the Guyan reduction method to the most
important nodes, such as points of connection with adjacent structural systems or
points of feedback (e.g. bearing reaction forces). The remaining nodes are con-
sidered as included in the internal system.
This method is hereafter explained using a model shown in Fig. 4.5 without
losing sight of general features. The shaft has a simple support at the left end, and a
spring element with a variable boundary value kb at the right end. Specically, the
coupling with a control system assumes a state feedback control u acting at the right
end. The right end is then chosen as the boundary coordinate, i.e. the most
important (master) node. All other coordinates form the internal system. In addition,
an external force distribution F(x, t) is assumed to act on the shaft.
Two kinds of mode shown in Fig. 4.6 are considered with the notation of n = x/l.

Fig. 4.5 Uniform shaft l


kb
y (x, t)
u
x
... ...
F (x, t) y1
4.1 Mode Synthesis Models 85

Fig. 4.6 Modes for mode 1


1
synthesis
x
0 =
0.5 1 l
y1
3 2
1

(1) the static shaft deflection mode d caused by a unit forced displacement of the
boundary coordinates, that is the mode shape of a straight line with the pinned
displacement (=0) at the left end and unity (y1 1 at the right end:

dn 4:9

We can lift up the right end of the shaft to the unit displacement without any
reaction force. Then, the equivalent spring constant:

Keq 0 4:10

(Note: The boundary spring kb at the right end is ignored at rst, and super-
imposed later directly on the completed reduced order model), and
(2) the eigenmodes / of the internal system under the boundary condition,
assuming y1 0 at the right end, are
Eigenmodes for simple support at both ends:
/i sinipn i 1; 2; 3 4:11

(Note: Three eigenmodes are employed for modal analysis of the internal
system).
Using these two types of mode, the coordinate transformation for mode synthesis
is dened:
X
n
yx; t dxy1 t gi t/i x
i1

dx /1 x /2 x /3 x  y1 g1 g2 g3 t 4:12
   
y1 y1
 d U W
g g

where y(x, t) is the vibration displacement of the position x of the beam, and y1(t) is
the vibration displacement of the right end.
The congruent transformation of Eq. (4.12) gives the equation of motion in the
mode synthesis coordinate system {y1, }:
         
Md Mc y1 Keq 0 y1 1 Fd
ut 4:13
Mct Mg g 0 Kg g 0 F/
86 4 Mode Synthesis and Quasi-Modal Method

where
  Z1 Z1  
Md Mc d2 dU
 qAl W Wdn qAl
t
dn 4:14
Mct Mg Ut d Ut U
0 0

The effective mass sensed at the boundary coordinate is


Z1
Md  qA d2 xdx qAl=3
0

The equivalent shaft stiffness is Keq = 0 (because the deflection mode is a rigid
mode, a straight line and no strain energy is stored).
The modal mass for the internal system (with simple supports at both ends) is

Z1
 
Mg  qAl Ut Udn diagonal    ; mi ;    qAl  diagonal1=2; 1=2; 1=2
0

Z  
1  
 1 1 1
Coupled mass: Mc  qAl dnUndn    ; mci ;    qAl ; ;
0 p 2p 3p
Natural circular frequency of the internal system is
s
k2 EI
xz 2 k p; 2p; 3p; referring to Table 3.1(5)
l qA
 
Modal stiffness of the internal system: Kg diagonal    ; ki mi x2zi ;   
  Z1  
Fd dnF n; t
Modal external force: l dn
F/ Ut nF n; t
0
Therefore the mass matrix Mw and stiffness matrix Kw of the mode synthesis
model are
2 3
  1=3 1=p 1=2p 1=3p
Md Mc 6 1=p 1=2 0 0 7
Mw  qAl6
4 1=2p
7 4:15
Mct Mg 0 1=2 0 5
1=3p 0 0 1=2

and
  " #
Keq 0 EI p4 2p4 3p4
Kw  3 diagonal 0; ; ; 4:16
0 Kg l 2 2 2
4.1 Mode Synthesis Models 87

Since the boundary spring reaction force at the right end is

u kb x 4:17

superimposing the boundary spring constant kb on the stiffness matrix yields the
stiffness matrix Kw for the mode synthesis model for the entire system:
" #  
EI p4 2p4 3p4 EI
Kw 3 diagonal Kb ; ; ; kb Kb 3 4:18
l 2 2 2 l

Features of this model can be summarized as follows:


(1) The matrices Md and Keq represent the effective mass sensed at the right end of
the system and the equivalent shaft stiffness, respectively, as with the Guyan
method. The effective mass is 1/3 times the total mass in this example.
(2) The mass matrix M and stiffness matrix K for the internal system are
diagonal (as with modal analysis) because the corresponding internal eigen-
modes are orthogonal with respect to the mass and stiffness matrices.
(3) Since the deflection mode d and internal eigenmodes / are not orthogonal, the
congruent matrix Mw is obtained as their inner products, concerning the mass
matrix, which implies non-zero edge elements Mc.
(4) This edge diagonal mass matrix is the principal difference of the mode syn-
thesis from the modal analysis where all congruent matrices are diagonal.
(5) Connection with a boundary element is performed by superposition of the
stiffness matrices as in Eq. (4.18). For example, if the control reaction force
us Gr sy1 s of the transfer function Gr(s) in response to a boundary
displacement y1 is the action, kb is simply replaced by Gr(s).
For further information, readers are recommended as a minimum to review
original works [1820] and refer to some applications [2125].

Example 4.3 Solve the eigenvalue problem x2n MW W KW W for the system shown
in Fig. 4.5. Conrm Fig. 4.7 by indicating the dependence of the natural frequency

Fig. 4.7 Natural frequency 100


map 3rd
50
1 EI

l 2 A

20
10 2nd
n =


2

1st
2

1
1 10 100 1000 10000
EI
stiffness k b = Kb 3
l
88 4 Mode Synthesis and Quasi-Modal Method

xn on the boundary spring constant kb, and check the values for the left and right
ends.
Answer
It is called the natural frequency map because it shows the relationship between xn
and kb..
Left end (the boundary condition is simple supportfree) gives, referring to
Table 3.1(3):

k f 3:927 7:069 10:21 g ! k2 f 15:4 50 104 g

Right end (both the boundary conditions are simple supports) gives, referring to
Table 3.1(5):

k f1 2 3 g ! k2 f 1 4 9 gp2

Example 4.4 Construct the mode synthesis model for the cantilever system shown
in Fig. 4.8a. Conrm that the mode expansion gives the modes in Fig. 4.8b, c.
Conrm Fig. 4.9 by indicating the dependence of the natural frequency xn on the
boundary spring constant kb, and check the values for the left and right ends.
Answer
The conversion modes for the mode synthesis, with the designated boundary
conditions for the right end, are the following two types:

Fig. 4.8 Modes for mode


synthesis
(a) kb
Al u

3 EI
m = Al /4 , k eq =
l3

(b)

fixed-free mode deflection mode

1
(c)

3
2

fixed-pinned modes
4.1 Mode Synthesis Models 89

Fig. 4.9 Natural frequency 100


map
50
3rd

1 EI

l 2 A
20
2nd
10

n =


2
5
1st
2

1
1 10 100 1000 10000
EI
stiffness k b = Kb 3
l

(1) Deflection mode shown in Fig. 4.8b:


Deflection curve with the right end lifted up by unit length = 1:
d 3  nn2 =2
Equivalent shaft stiffness: Keq 3EI=l3
(2) Eigenmodes for the internal system shown in Fig. 4.8c as shown in Table 3.1
(4):
The rst three eigenmodes /n for the simple support at the right end are used:
/n n /0 n=/0max

cosh kn  cos kn sin hkn  sin kn


/0 n  ;
cosh k sin hk
where k f 3:927 7:069 10:21 g;
/0max f 0:06 0:00261 0:000112 g and n x=l

The equation of motion for the mode synthesis model of Eq. (4.13) includes:
Mass matrix:
2 3
  33=140 0:23 0:131 0:091
Md Mc 6 0:23 0:431 0 0 7
Mw  qAl6
4 0:131
7 4:19a
Mct Mg 0 0:425 0 5
0:091 0 0 0:432

Stiffness matrix:
 
EI EI
Kw 3 diagonal3 Kb ; 1020:5; 1062; 4693 kb K b 3 4:19b
l l
90 4 Mode Synthesis and Quasi-Modal Method


Solving the eigenvalue problem x2n MW W KW W corresponding to Eqs. (4.19a
and 4.19b) yields the natural frequency xn for a variable boundary spring constant
kb shown in Fig. 4.9. The values for the left and right ends in the map are as
follows:
Left end (the boundary condition is xedfree) gives, from Table 3.1(6),

k f 1:875 4:694 7:855 g ! k2 f 3:515 22:03 61:7 g

Right end (boundary condition: xedsimple support) gives, from Table 3.1(4),

k f 3:927 7:069 10:21 g ! k2 f 15:4 50 104 g

4.2 Quasi-modal Models

No physical model equivalent to the mode synthesis model can be constructed


because the mass matrix Mw is not diagonal due to the existence of the mass
coupling (non-zero elements) along its edge. This section presents a method to
construct a mechanical model equivalent to the mode synthesis model. The
obtained model is called a quasi-modal model [17], though the name is not yet
accepted by standards, because it is similar to the modal model as stated in Fig. 3.8.

4.2.1 Principle of the Quasi-modal Model

The existence of the non-zero elements located at the edges of the mass matrix Mw
in the mode synthesis model is due to the inertial force between the boundary
coordinates and the modal coordinates in the internal mass system. These non-zero
coupling terms come from the fact that the modal coordinates are relative coor-
dinates, as seen in Eq. (4.12). So virtual absolute coordinates, n, corresponding to
the relative modal coordinates, , are introduced here. The relative coordinates are
assumed to be proportional to the difference between these virtual absolute coor-
dinates and the boundary absolute coordinates y1:

gi ai ni  y1 i 1; 2; 3 4:20

The coefcients, ai, are yet to be determined. The corresponding coordinate


transformation matrix T is dened as
   
y1 y
T 1 4:21
g n
4.2 Quasi-Modal Models 91

2 3
1 0 0 0 2 3 2 3
6 a1 g1 n1
a1 0 07
where T6
4 a2
7; g  4 g2 5; n  4 n2 5
0 a2 05
g3 n3
a3 0 0 a3

Substituting into Eq. (4.13), it follows from the congruent transformation, with
ai = (1st row, i-th column element of Mc)/(i-th row, i-th column element of M),
with

ai f 2=p 1=p 2=3p g; 4:22

that the following equation is obtained with a diagonal mass matrix:


     
y1 y1 u
Mn n Kn n 0 4:23

With the equivalent mass of the internal system and coupled spring constant, we
obtain the congruent mass matrices:
 
P3
Mn  T t Mw T diagonal Md  mni mn1 mn2 mn3
i1
 
1 49 2 1 2 4:24
Mn  T Mw T diagonal 
t
3 18p2 p2 2p2 9p2
qAl
0:173 0:608 0:152 0:067 
3

and stiffness matrix:


2 3
P3
6 k b K eq k ni k n1 k n2 k n3 7
6 i1 7
6 kn1 0 7
Kn  T Kw T 6
t kn1 0 7
6 7
4 kn2 0 kn2 0 5
kn3 0 0 kn3 4:25
2 3
Kb =p2 28 2 8 18
 
EI 2 66 2 2 0 0 7
7 EI
3p 6 7 kb Kb 3
l 4 8 0 8 0 5 l
18 0 0 18

Although the stiffness matrix Kn is not diagonal, the mass matrix Mn is so, thus
permitting construction of a multiple mass model, called a quasi-modal model, as
shown in Fig. 4.10.
The obtained quasi-modal model shows a parallel sum of single-dof systems
(similar to the modal model Fig. 3.11), which represent the eigenmode components
92 4 Mode Synthesis and Quasi-Modal Method

Fig. 4.10 Quasi-modal 2 3


model
1

m1 m2 m3
m 61% 15% 7%

1 2 3
K k 1 2 k 2 8 k 3 18

m
equivalent mass 17% ys
M = Al / 3
m = ratio to M kb
EI
K = ratio to 3 2
l

of the internal system, connected with the boundary mass. The diagonal elements of
the mass matrix are the ratios with which the effective mass M = qAl/3 sensed at the
right end is divided to each mass point ni. Review the division
f mn1 60:8 % mn2 15:2 % mn3 6:7 % g distributed at each mass point and
the remaining mass Dm = 17.3 %, which is placed at the boundary point. Because
the sum is 100 %, the effective mass, Md, is thus preserved.
Features of the quasi-modal model are summarized as follows:
(1) The (1,1) element of the mass matrix, Mn, of Eq. (4.24) represents the mass
corresponding to the forced deflection mode at the boundary physical point y1.
The effective mass, Md, obtained by the mode synthesis is partially divided by
the equivalent masses, mni , of the single-dof systems corresponding to the
chosen eigenmodes of the internal system. When considering the boundary
point y1 as mother and the internal modal coordinates ni as children, the
the total mass of the mother and children remains at the effective mass Md.
(2) The child single-dof systems are supported by springs kn, and connected to
the mother coordinates. But child systems are orthogonal to one another
and have no mutual connection.
(3) The 4-mass system of the quasi-modal model can be reduced to a 3-mass
system by bringing the third child back to the mother, as shown in
Fig. 4.11a, leaving the two eigenmodes of the internal system. Similarly 2- or
1-mass systems can be constructed as shown in Fig. 4.11b, c by reverting
child masses to mother mass.

Example 4.5 Conrm that the quasi-modal model Fig. 4.12 corresponds to the
mode synthesis model given in Example 4.4.
4.2 Quasi-Modal Models 93

Fig. 4.11 Mother-children (a) (b) (c)


modeling
61% 15% 61%

39% 100%
m 24%

kb kb kb

3 dof 2 dof 1 dof

Fig. 4.12 Quasi-modal 1


model of cantilever 2 3

m1 m2 m3
m 52% 17% 8%

1 2 3
K k 1 29 k 2 100 k 3 208

equivalent mass
23%
M = 33 Al
y1
140
EI
m = ratio to M k eq k b = Kb
l3
EI
K = ratio to 3
l

Answer
The coordinate transformation matrix to the quasi-modal model is:

2 3
1 0 0 0
6 0:533 0:533 0 0 7
T6
4 0:307
7
0 0:307 0 5
0:211 0 0 0:211
94 4 Mode Synthesis and Quasi-Modal Method

Mass matrix:
2 3
0:0538 0 0 0
6 7
6 0 0:123 0 0 7
Mn  T t Mw T qAl6 7
4 0 0 0:0401 0 5
0 0 0 0:0192

33=140 q Al diagonal 0:23 0:52 0:17 0:08 

Stiffness matrix:
2 3
Kb 3 337 29 100 208
EI 6 29 29 0 0 7
Kn  T Kw T 3 6
t 7
l 4 100 0 100 0 5
208 0 0 208

Example 4.6 Consider the system shown in Fig. 4.13 supported at both ends by
bearing springs (kb each). Choosing the bearings as the boundary coordinates
X1 = {x11, x12} and assigning the internal coordinates X2 to the rest, and assuming
that the bearing springs are to be superimposed after the rotor model is completed,
construct the mode synthesis model and quasi-modal model of the system by the
following procedures:
(1) Dene a mode synthesis transformation matrix U including modes d1 and d2 in
Fig. 4.13a as the deflection modes by forced displacement of the boundary
node and three eigenmodes /1, /2 and /3 of the internal system. Derive the
equation of motion for the mode synthesis model.
(2) Replace the deflection modes by the translation dp and tilting dt modes
(Fig. 4.13b), and redene the mode synthesis transformation matrix U
(Fig. 4.13c). Derive the corresponding equation of motion for the mode
synthesis model.
(3) Construct the quasi-modal model for (2) above and draw the mechanical
system.
(4) Construct a critical speed map showing the relationship between the natural
frequency xn and the spring constant kb.

Answer
(1) With n x=l; W 1  n n sin pn sin 2pn sin 3pn  and Keq = 0
Natural frequency of the internal system, i.e. anti-resonant frequency as seen
in Eq. (4.27):
s
k2 EI
xz 2 ki p; 2p; 3p
l qA
4.2 Quasi-Modal Models 95

Fig. 4.13 Rotor supported (a)


by 2 bearings and modes for x11 X2 x12
mode synthesis
,,A l

m = Al
kb 1 3 kb
Id = Al
12
EI
1 k b = Kb 3
l 2

1 1

x11 x12

(b) p

1 t

(c) 1
2

(see Table 3.1(5))


2 3 2 3
2 1 6=p 3=p 2=p x11
6 7 6 7
Z 1 6 1 2 6=p 3=p 2=p 7 6 x12 7
qAl 6
6 6=p 6=p
7 6 7
MW qAl Wt Wdn 3 0 0 7 6
7; Xg 6 g1 7
7
6 66 7 6 7
0 4 3=p 3=p 0 3 0 5 4 g2 5
2=p 2=p 0 0 3 g3

 
KW diagonal kb kb p4 =2 8p4 81p4 =2 EI=l3

Equation of motion for the mode synthesis model: MW X g KW Xg 0


(2) Using the translational displacement xp x11 x12 =2 and tilting displace-
ment xt x11 x12 =2
96 4 Mode Synthesis and Quasi-Modal Method

2 3 2 3
6 0 12=p 0 4=p xp
6 7 6 7
Z 1 6 0 2 0 6=p 0 7 6 xt 7
qAl 6
6 12=p
7 6 7
MW qAl Wt Wdn 0 3 0 0 7 ; X  6 g1 7
6 6 7 6 7
g
6 7 6 7
0 4 0 6=p 0 3 0 5 4 g2 5
4=p 0 0 0 3 g3
  
KW diagonal 0 0 p4 =2 8p4 81p4 =2 EI=l 3 KW
diagonal 2kb 2kb p =2 8p 81p =2 EI=l3
4 4 4

Equation of motion: MW X g KW Xg 0
(3) Coordinate transformation to the quasi-modal model:
2 3
1 0 0 0 0
6 7
6 0 1 0 0 0 7
6 7
6
T 6 4=p 0 4=p 0 0 7
7
6 7
4 0 2=p 0 2=p 0 5
4=3p 0 0 0 4=3p

 
Mn T t MW T qAl 0:1 1=3 2=p2 0:81 2=p2 0:09

2 3
2Kb 80p2 0 8p 0 72p2
6 0 2Kb 32p2 0 32p2 0 7
6 7
Kn T t MW

T EI 6 8p2 0 7
l 6
3 0 8p2 0 7;
4 0 32p2 0 32p2 0 5
72p2 0 0 0 72p2
t
Xn  xp xt n1 n2 n3 

The equation of motion for the quasi-modal model, Mn X n K  Xn 0; is


n
separated into the translational and tilting systems. The mechanical model for
the translational system fxp ; n1 ; n3 g is shown in Fig. 4.14a, and that for the
tilting system fxt ; n2 g in Fig. 4.14b.
The quasi-modal model thus converts a part of the effective mass, Md, into the
equivalent mass of the eigenmode of the internal system, and the mass of the
entire system is constant. This is a physically convenient feature of the model.

(4) Solving the eigenvalue problem x2n MW / KW kb / with the bearing spring
kb as the parameter yields Fig. 4.15.
As for the ends of the map, we consider that the left end is in a freefree
boundary condition. Table 3.1(1) gives k = {4.73, 7.853, 10.996}; therefore
k2 = {22.4, 61.4, 121}.
4.2 Quasi-Modal Models 97

Fig. 4.14 Quasi-modal (a) (b)


model 1 3 2
0.81 0.09 0.6

8 2 72 2 32 2

xp xt mass
EI
stiffness 0.1 0.4
m
l3 3
mass m

kb kb kb kb

translation tilting motion

Fig. 4.15 Critical speed map 100

50
1 EI

l 2 A

20

10
n =


2

1
1 10 100 1000 10000
EI
stiffness k b = Kb 3
l

The right end is in a simply supportedsimply supported boundary condition.


Table 3.1(5) gives
k = {p, 2 p, 3 p}; therefore k2 = {p2, 4 p2, 9 p2}.

4.2.2 Examples of Quasi-modal Models

Examples of quasi-modal models for various systems are shown in Table 4.1.
98 4 Mode Synthesis and Quasi-Modal Method

Table 4.1 Examples of quasi-modal modeling [9]

(a) tilting (b) cantilever (c) 2 bearings (d) torsion

1 1 d2
m = Id = A l
3 Ip = A l
3
m m = 0.24 m m = A l 12 8

f f f1 f2 f

defection modes caused by forced displacement input


1

k eq = 0 k eq = 3 2 k eq = 0
k eq = { 0 , 0 }

eigenmodes of inner system only


c1 c1 c1 c1

c3 c3 c3 c3
c2 c2 c2 c2

, 3 , 5
i = { , 2 , 3 } i = {3.93, 7.07 , 10.2 } i = { , 2 , 3 } i = { }
2 2 2
2i EI 2i EI 2i EI i G
i = i = i = i =
l2 A l2 A l2 A l

m = 1 3 m m = 33 140 m = 0.24 m

m1 m = 0.61 m1 m = 0.52 m1 m = 081 m1 I p = 0.81


m2 m = 0.15 m2 m = 0.17 m2 I d = 0.61 m2 I p = 0.09
m3 m = 0.07 m3 m = 0.08 m3 m = 0.09 m3 I p = 0.03

m m = 0.17 m m = 0.23 m m = 0.1 Ip I p = 0.07


Id Id = 0.39

m m m
m m 1
m m Id m
0.61 0.15 0.52 0.17 0.81 0.81
0.07 0.08 0.61 0.09
Ip
2 f
1 2 3 1 2 3 1 2 3
0.09
Ip
3
k eq = 3 0.07 I p
k eq = 0 0.17m 0.23m 0.03
f f m = 0.1 m Ip
f1 Id = 0.39 Id f2

Ref. to Fig. 410 Ref. to ex.45 Ref. to ex.46


4.3 Plant Transfer Function 99

4.3 Plant Transfer Function

This section discusses the plant transfer function y1/u for a single boundary coor-
dinate y1 in the system shown in Fig. 4.5. The equation of motion of Eq. (4.13),
corresponding to mode synthesis model, is rewritten using values of md and keq,
instead of the matrix representations, Md and Keq, respectively, to emphasize the
one-dimensional matrices.
If the eigenmodes of the internal system are normalized by mass,

Z1
Mg qAl Ut Udn 1 unit matrix
0
Kg diagonal[ x2z1    x2zi      x2z 
diagonal matrix including xz natural frequencies of the internal system

Therefore, the equation of motion for the mode synthesis model is


       
md Mc y1 k 0 y1 1
:: eq ut 4:26
Mct 1 g 0 x2z  g 0

where
2 3 2 3
md mc1 mc2 mc3 keq 0 0 0
  6m 7  keq  6 7
md Mc 6 c1 1 0 0 7 0 6 0 x2z1 0 0 7
6 7; 6 7
Mct 1 4 mc2 0 1 0 5 0 x2z  4 0 0 x2z2 0 5
mc3 0 0 1 0 0 0 x2z3
md qAl=3

keq md x2d (setting non-zero here for the sake of generality, though keq = 0 in
Fig. 4.5)
The plant transfer function is therefore

y1 s s2 x2z1 s2 x2z2 s2 x2z3


Gp s  4:27
us Ds

where the denominator:



md s2 x2 mc1 s2 mc2 s2 mc3 s2
d
mc1 s 2
s x2z1
2
0 0
Ds 
4:28
mc2 s 2
0 s2 x2z2 0
mc3 s2 0 0 s2 x2z3
100 4 Mode Synthesis and Quasi-Modal Method

The numerator of the plant transfer function contains the natural frequencies xzi
(i = 13). They are the natural frequencies of the system whose boundary coor-
dinate is constrained, corresponding to the modes in which the boundary is a node.
In other words, the plant does not respond to excitation of these frequencies xzi .
These anti-resonant (or zero) frequencies correspond to the frequencies at the right
end of the natural frequency map (Fig. 4.7).
On the other hand, the denominator contains the resonance frequencies for a free
boundary (without bearing, u = 0), or pole frequencies. They correspond to the
frequencies at the left end of the map (Fig. 4.7). Let xpi (i = 03) be the pole
frequencies. Taking into account that the square of the coupled mass is the child
mass in the quasi-modal model:

m2c i mn i i 1; 2; 3; 4:29

the denominator of the plant transfer function is written as


!
X
3




Ds md  mn i s2 x2p0 s2 x2p1 s2 x2p2 s2 x2p3 4:30
i1

Comparison of the constant terms in Eqs. (4.28) and (4.30) yields


!
X
3
md  mni x2p0 x2p1 x2p2 x2p3 md x2d x2z1 x2z2 x2z3 4:31
i1

The lowest-order natural frequency xp0 at the left end of the natural frequency
map is approximately equal to the lowest-order natural frequency xd in the Guyan
reduction model. The plant transfer function is therefore presented in following
equation [17]




s=xz1 2 1 s=xz2 2 1 s=xz3 2 1
Gp s
2
2
2
md s2 x2d s=xp1 1 s=xp2 1 s=xp3 1




s=xz1 2 1 s=xz2 2 1 s=xz3 2 1


2
2
2
keq s=xd 2 1 s=xp1 1 s=xp2 1 s=xp3 1




s=xz1 2 1 s=xz2 2 1 s=xz3 2 1

2
2
2
md s2 keq s=xp1 1 s=xp2 1 s=xp3 1
4:32

Specically in Fig. 4.5, where keq = 0, i.e., xd = 0,


4.3 Plant Transfer Function 101





s=xz1 2 1 s=xz2 2 1 s=xz3 2 1
Gp s
2
2
2 4:33
md s2 s=xp1 1 s=xp2 1 s=xp3 1

In the series of the frequencies involved in the ascending order, the pole fre-
quencies xp and anti-resonant frequencies xz occur alternately:

xd xp0 \xz1 \xp1 \xz2 \xp2 \xz3 \xp3    4:34

Example 4.7 Construct the Bode plot of the plant transfer function for the system
shown in Fig. 4.5 and estimate the gain at x = 1.
Answer
s
1 EI qAl
With x0 2 ; md and keq = 0
l qA 3
 
Anti-resonant frequencies: xz p2 4p2 9p2 x0 f 10 40 90 gx0 ;
referring to Table 3.1(5)  
Pole frequencies: xp 3:9272 7:0692 10:212 x0 f 150:4 50
104gx0 ; referring to Table 3.1(3)
Substituting these into Eq. (4.33) gives the Bode plot shown in Fig. 4.16 .
Since Gp x 1=md x2 at low frequencies, Gp = 3 = 9.5 dB at x = 1.
Example 4.8 Construct the Bode plot of the plant transfer function for the can-
tilever system shown in Fig. 4.8 and estimate the gain at x = 1.
Answer
s
1 EI 33 EI
With x0 2 ; md qAl and keq 3 3
l qA 140 l

40
1 pinned
20 at left y1
1

1 Gr
EI
| Gp | dB 3
l

0 u
20 2 fixed
at left
40
y1
60 2 Gr
u
80
1 2 5 10 20 50 100 200
1 EI
frequency = 2 0 2
l A

Fig. 4.16 Plant transfer function of beam y1/u


102 4 Mode Synthesis and Quasi-Modal Method
 
Anti-resonant frequencies: xz 3:9272 7:0692 10:212 x0
f 150:4 50 104 gx0 ; referring

to Table 3.1(4) 
Pole frequencies: xp 1:8752 4:6942 7:8552 x0 f 3:52 22 61:7 gx0 ;
referring to Table 3.1(6)
Substituting these into Eq. (4.32) gives the Bode plot shown in Fig. 4.16 .
Since Gp 1=kd at low frequencies, the transfer function shows flat character-
istics Gp = 1/3 = 9.5 dB at x = 1.
Example 4.9 The response in the s domain of the system shown in Fig. 4.13 to a
translation input up and a tilting input ut are, using the mass matrix Mw and stiffness
matrix Kw of the mode synthesis model constructed in Example 4.6(2),
3 2 2 3 2 3
xp s up s 1 0
6 xt s 7 6 ut s 7 6 0 17    
6 7 6 7 6 7 up s up s
MW s2 KW 6 7 6 7 6
6 g1 s 7 6 0 7 6 0 077 u s  B 4:35
4 g2 s 5 4 0 5 4 0 ut s
05 t
g3 s 0 0 0

(1) Construct the Bode plots of the plant transfer functions for the translational
system Gpp xp =up and the tilting system Gtt xt =ut .
(2) How are the zero and pole frequencies of the plant transfer function related to
the natural frequency map (Fig. 4.15)?
Answer
(1) With C = Bt, the plant transfer function is derived from CMW s2 KW 1 B
Diagonal Gpp Gtt ; which yields the Bode plot shown in Fig. 4.17.
(2) The anti-resonant frequencies are related to the natural frequencies of the right
end (simply supported - simply supported) in Fig. 4.15.

40

20 1 1 1
Tilt G tt
2 1 1
1
EI
|Gp | dB 3
l

0 xp xt
x1 x2
20 Grp Grt
Translation up ut
40 G pp u1 u2
1 1 1
60
2 1 1
80
1 2 5 10 20 50 100 200
1 EI
frequency = 0 2
2

l A

Fig. 4.17 Plant transfer function


4.3 Plant Transfer Function 103
 
Translational system: xz p2 9p2 x0 f 10 90 gx0 , referring to
Table 3.1(5)
Tilting system: xz 4p2 x0 40x0 , referring to Table 3.1(5).
The pole frequencies are related to the natural frequencies of the left end
(free-free boundary) in Fig. 4.15:
Translational system: xp f4:732 112 gx0 f22:4 121gx0 , referring to
Table 3.1(1).
Tilting system: xp f7:852 gx0 f61:4gx0 , referring to Table 3.1(1).
Chapter 5
Unbalance and Balancing

Abstract Most of the cases of rotor vibration problems come from excessive
vibration or resonance due to unbalance. A quick remedy to compensate for it is
balancing. Balancing is based on the assumptions of a linear relationship between
input (unbalance) and output (vibration), namely,
The amplitude of unbalance vibration is proportional to the level of unbalance,
and
A shift in the angular position of unbalance on a rotor results in a corresponding
phase shift of the vibration waveform.
While this linearity is the sole theoretical concept to explain how balancing works,
practical methods of balancing include various alternatives based on operators
experiences. Readers are expected to explore the numerical examples prepared so as
to experience a wide variety of techniques.

  
Keywords Unbalance Imbalance Balancing Influence coefcient method 
 
Modal balance (n + 2) plane balance Universal balance

5.1 Unbalance in a Rigid Rotor

5.1.1 Static Unbalance and Dynamic Unbalance

When a rotor has a centroid S at a distance of e from the center of gravity G, the
centrifugal force or unbalance force is

F meX2 N] 5:1

which acts on the shaft. The quantity U = me is called the unbalance and e is the
eccentricity. In principle, the direction of eccentricity can be known by placing the
rotor on smoothly movable (near frictionless) bearings, because the rotor becomes
stationary with the center of gravity G beneath the centroid S as shown in Fig. 5.1a.

Springer Japan 2017 105


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_5
106 5 Unbalance and Balancing

Fig. 5.1 Static unbalance


(a)
S

G

0 (S G ) , = 0

(b) b a
Fl = F Fr = F
a+b a+b

G

a b

2
F = m

In other words, it is not necessary to rotate the rotor to identify the unbalance. This
type of unbalance is called static unbalance.
Figure 5.1b illustrates the translational reaction force at the bearing caused by the
unbalance. This is an alternating force that changes the direction as the rotor rotates.
When a rotating disk has a centroid S that coincides with the center of gravity
G and the principal axis of inertia, passing through the center of gravity, is tilted at
an angle s with respect to the rotational axis as shown in Fig. 5.2a, the moment

Fig. 5.2 Dynamic unbalance


(a)

G=S

= 0 (S = G ),
0

(b) M
Fl =
l
l

G
M
M Fr =
l
2
M = ( Ip Id )
5.1 Unbalance in a Rigid Rotor 107

(a) m (b) m

r r

r
mr = m a
m

Fig. 5.3 Static unbalance and couple unbalance

M Ip  Id sX2 N m] 5:2

acts on the shaft. This moment can be detected not by simply placing the rotor on
bearings, but by rotating it. This type of unbalance is hence called dynamic
unbalance. Figure 5.2b illustrates the alternating reaction force on the bearings
generated by rotation of the rotor.

5.1.2 Static Unbalance and Couple Unbalance

The static unbalance of a disk with a mass m and an eccentricity e can be repre-
sented by a small mass Dm at a radius r (Fig. 5.3a):

Static unbalance: U me Dmr kg m] 5:3

The moment which causes the dynamic unbalance of a disk tilted by s is


equivalent to two out-of-phase static unbalances on either side of the disk being
separated by distance a (Fig. 5.3b):

Ip  Id s Dmra kg m2  5:4

This pair of out-of-phase static unbalances is called couple unbalance.


While an actual rotor has both static and dynamic unbalances, they are thus nally
represented by static unbalance at two locations (planes). The instrument to detect
equivalent static unbalance at those two planes is the rigid balancer (described later).
Balancing is an operation to eliminate unbalance thus detected, in practice by
attaching appropriate masses to, or trimming excess material from the rotor.

5.1.3 Adverse Effects of Unbalance Vibration

Static unbalance and moment are often treated as scalars indicating the magnitudes
of which are shown above. However, they are actually vectors because they
108 5 Unbalance and Balancing

unbalance shaft vibration bearing vibration casing vibration floor vibration

flexible shaft deflection bearing load instrument damage

(rubbing) shaft stress oil film rupture


disturbance to precision machinery

reduction in life breakage

Fig. 5.4 Harmful influences caused by unbalance

correspond to specic phases of the rotor cross-section, and the direction of which
alternates in synchronization with rotation of the rotor. The alternating bearing
reaction force generated by unbalance is thus transferred to the bearing mounts and
floor, and may act as a source of excitation.
Figure 5.4 shows various adverse effects of unbalance vibration. It is clear that
unbalance vibration is the root of all inconveniences, or, in other words, balance is
of prime importance in rotating machinery for the alleviation of problems.

5.1.4 Residual Permissible Unbalance in a Rigid Rotor

In the manufacturing process of rotors, balancing is made for each part as a rigid
body. At the nal stage after assembling rotating parts, the nal balancing is made
for this assembled rotor as a rigid body before being placed inside the casing.
Two types of dynamic balancers for rigid rotors (Fig. 5.5) are commercially
available:
(1) Hard-bearing balancing machine: The rigid shaft is supported by rigid char-
acteristics of bearings, where bearing reaction force is measured and converted
to residual unbalance.

(a) (x) :unknown eccentricity (b)


(x ) S1 (x ) S2

FB1 G FB2 G

Hard type Soft type

Fig. 5.5 Dynamic balancer


5.1 Unbalance in a Rigid Rotor 109

Table 5.1 Balance quality grades (ISO 1940-1:2003)

G 250 Crankshaft drives, inherently unbalanced, rigidly mounted


G 100 Complete reciprocating engines for cars, trucks and locomotives
G 40 Cars: wheels, wheel rims, wheel sets, drive shafts
Crankshaft drives, inherently balanced, elastically mounted
Agricultural machinery, Crankshaft drives, inherently balanced, rigidly mounted
G 16
Crushing machines, Drive shafts (cardan shafts, propeller shafts)
G 6.3 Aircraft gas turbines, Centrifuges (separators, decanters)
Electric motors and generators (of at least 80 mm shaft height), of maximum rated
speeds up to 950 r/min, Electric motors of shaft heights smaller than 80 mm
Fans, Gears, Machinery general, Machine-tools, Paper machines, Process plant machines
Pumps, Turbo-chargers, Water turbines
Compressors, Computer drives
G 2.5 Electric motors and generators (of at least 80 mm shaft height), of maximum rated
speeds above 950 r/min
Gas turbines and steam turbines, Machine-tool drives, Textile machines
G 1.0 Audio and video drives, Grinding machine drives
G 0.4 Gyroscopes, Spindles and drives of high-precision systems

(2) Soft-bearing balancing machine: The rigid shaft is supported by soft charac-
teristics of bearings, where bearing vibration is measured and converted to
residual unbalance.
The ISO 1940-1 standard denes balance quality grade G as eccentricity
(mm)  angular rotational speed (rad/s), as shown in Table 5.1:

eper lm] N rpm] eN


Balance quality grade: G 2p mm/s] 5:5
1000 60 9550

where N = operational speed [rpm], and eper = residual permissible unbalance


(eccentricity) [lm].
The permissible unbalance Uper is therefore

Uper eper m lm kg g mm 5:6

Uper is the permitted value in single-plane balancing; it is separated into two


planes for evaluation in two-plane balancing. A recent trend is to take the bearing
planes for convenience in the control of vibration transmission and bearing loads;
here Uper is split into UperA and UperB according to the equation for static unbalance
(Fig. 5.6). The two values should satisfy the inequalities regarding UperA; B =Uper
shown in the gure; otherwise the value should be replaced by the corresponding
boundary value.
Although the bearing plane is preferable for the plane to be evaluated, both sides
of the rotor body are still often chosen for the practical correction plane. In this case
110 5 Unbalance and Balancing

(a) UperA, B (b) UperA, B


0.3 < < 0.7 0.3 < < 1.3
Uper Uper

Uper
UperB

UperA
Uper

UperB
UperA

CM CM

A B A B

LA LB L LB
L LA

Inboard rotor Outboard rotor

Fig. 5.6 UperA Uper LB =L; UperB Uper LA =L at bearing planes


UperA

(a) (b)
Uper1

Uper1,2 =UperA,B L

UperA
Uper1
1/1 Uper1,2 = U
b perA,B
UperB

Uper2
Uper2

L/b

UperB
1/1 L/b

CM
A B

A B
LA LB L
L b
Correction planes inside of bearings Correction planes outside of bearings

Fig. 5.7 Allocation of Uper to two correction planes

Uper is split into the permitted values Uper1 and Uper2 for the correction planes as
shown in Fig. 5.7.
Example 5.1 Find the balance quality grade G for e 10 lm and N = 6,000 rpm.
Answer

G 10  103  2p6000=60 6:28  6:3

Example 5.2 Find the permissible unbalance for the steam turbine rotor shown in
Fig. 5.7a with a total mass = 3,600 kg, N = 4,950 rpm, LA = 8 m, LB = 10 m and
L = 18 m.
Answer
Unbalance quality grade G = 2.5 is recommended.
Permissible residual unbalance: eper 2:5  1000=2p4950=60 4:8 lm.
Permissible residual unbalance: Uper 4:8  3600 17:3  103 g mm.
Permissible residual unbalance at the left bearing plane: UperA 17:3  103 
10=18 9:6  103 g mm.
5.1 Unbalance in a Rigid Rotor 111

Permissible residual unbalance at the right bearing plane: UperB 17:3  103 
8=18 7:7  103 g mm.
Permissible residual unbalance converted for the left correction plane:
Uper1 UperA 9:6  103 g mm.
Permissible residual unbalance converted for the right correction plane:
Uper2 UperB 7:7  103 g mm.

5.2 Field Single-Plane Balancing (Modal Balancing)

5.2.1 Relationships among Rotational Pulse, Unbalance


and Vibration Vector

As shown in Fig. 2.17, the unbalance vibration response of Eq. (2.33) can be rep-
resented by Bode or Nyquist (polar) plots. The phase difference in the plot is with
respect to the direction of the unbalance eccentricity e, which is, however, unknown.
To determine the phase, a pulse sensor and a vibration sensor are utilized as
shown in Fig. 5.8. A once-per-turn mark (notch or keyway) is provided on the rotor
to generate rotational pulse signal when it rotates, and is used as the reference of the
rotor phase. Comparison of the rotational pulse and rotor vibration waveforms
permits detection of the phase difference / of the peaks as seen from the pulse, as
well as the vibration amplitude a. In practice, a vector monitor (Fig. 5.8) is used to
construct Bode or Nyquist plots.
Note: Here the polarity of the sensor signal is assumed to be positive when the
rotor approaches the sensor and negative when it moves away. The phase / is
positive when advancing and negative when lagging (the Nyquist plot proceeds
clockwise). In practice, however, these polarities may be reversed. Any actual
system needs polarity check before starting balancing.

positive negative
ac = a cos
vibration vector
as = a sin
A = ae j Nyquist plot
as
Y r ac
to
o ni as A
a
rm
cto
ve ac
G

Bode plot
t
S a
pulse amplitude a

X phase
O difference
vibration
speed

Fig. 5.8 Unbalance vibration and vector monitor


112 5 Unbalance and Balancing

(a)
1 2 3
pulse
45
- 45
x

Phase relationship between pulse and unbalance

(b) 5
1 = 2 = 3
(c)

phase difference [ ]
4
0
amplitude

1
3
2 3
90
2
1
0 180
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
p = / n p = / n
Amplitude curve Phase curve

(d) Im
lead
6 4 2 0 2 4 6
Re

2 lag

3 4 2

1 6

Nyquist plot ( polar plot )

Fig. 5.9 Bode plot and Nyquist plot of unbalanced vibration (at f = 0.1)

Figure 5.9 shows examples of response plots obtained in vibration analysis


using the pulse as the reference.
If the unbalance and the rotating pulse are in the same direction as shown in
this example, the Bode and Nyquist plots are similar to those of Fig. 2.17.
If the unbalance phase is advancing with respect to the mark (by 45 in the
rotation direction in this example), the phase curve shifts upward by the
unbalance phase difference, and the Nyquist plot rotates around the origin in
the direction of phase lead.
5.2 Field Single-Plane Balancing (Modal Balancing) 113

(a) Y Y 90 Y
U

0
U

18
0 t
G
t
G S S
A
S t
G
A A
O X O X O X
U
1 2 3
(b)
(c) Im
positive negative negative
1
Re
90
a
a
3
2
1 3 2

Fig. 5.10 Whirling motion of rotor and vibration waveform

If the unbalance phase is lagging with respect to the mark (by 45 in the
direction opposite to rotation in this example), the phase curve shifts down-
ward by the unbalance phase difference, and the Nyquist plot rotates around
the origin in the direction of phase lag.

The initial direction of the Nyquist plot thus permits prediction of the phase of
the unbalance. In an example shown in Fig. 5.10, which presents stroboscopic
views, synchronized with the rotation, of a rotor section before, during, and
after resonance, each showing the directions of the unbalance U, rotating mark,
and vibration vector A.
At a low rotational speeds (before resonance), the unbalance U and vibration
vector A are in approximately the same direction.
At the critical speed (during resonance), the vibration vector A lags by 90
behind the unbalance U. The vibration peak appears here to be 90 behind that
before resonance.
At a high rotational speeds (after resonance), the direction of the vibration
vector A is approximately opposite to the unbalance U.
The corresponding vibration in the Nyquist plot starts, as shown in Fig. 5.10c, in the
45 direction to upper right. The unbalance phase can thus be known, in principle,
from the initial direction of the corresponding Nyquist plot, or from the phase
difference between the rotating pulse and the peaks of vibration at low rotational
speed, but this is not really a practical approach. In fact, the influence coefcient
method described in the following section is more effective to identify the unbalance.
114 5 Unbalance and Balancing

5.2.2 Linear Relationship

When a rotor has an eccentricity e at a phase h with respect to the key mark, the
unbalance force is meejh X2 ejXt . Since it is in a linear relationship with the unbalance
vibration vector A, from Eq. (2.33)

eejh X2
A  Ga jXeejh 5:7
X x2n 2fxn X
2

X2
Ga s  ! Ga jX
s2 x2n 2fxn s
5:8
X2 A
 : influence coefficient
X x2n 2jfXxn eejh
2

The balancing procedure requires identication of the unknown magnitude e and


phase h of the eccentricity when the vibration vector A is measurable and the
influence coefcient Ga jX from the unbalance force to the measured vibration
vector is unknown.

5.2.3 Identifying the Influence Coefcient

Assume that a vibration vector A of an initial vibration, as shown in Fig. 5.11,


has been measured at a rotational speed X near the resonance frequency
(p X=xn 0:95 in this case).
In order to measure the influence coefcient, the rotor is rotated with a known
trial mass Dm attached at a radius r and phase ht , as shown in Fig. 5.12a, and the
vibration vector B is measured at the same rotational speed X as above . With the
total mass m,
 
Dmt jht
eejh
re X2  
m Dmt jht
B Ga jX ee jh
re 5:9
X2 x2n 2j1xn X m

Fig. 5.11 Modal balance 1 30

2 1 1 2 3 4
C
1 1
2 20 B A
3 p = 0.95
2
4
AC = 3
5 AB
5.2 Field Single-Plane Balancing (Modal Balancing) 115

Y Y
(a) (b)


30 30

Pulse X Pulse X

c
t 20
130 130

3m t
m t
Trial mass Correction mass

Fig. 5.12 Balance masses

Now the vibrations before and after attaching the trial mass are compared. The
difference between Eqs. (5.7) and (5.9) indicates the influence of the trial masses:
   
Dmt jht Dmt jht
B  A Ga jX re ! Ga jX B  A= re 5:10
m m

The difference B A is called the effective vector. The influence coefcient


Ga jX is the effective vector per unit eccentricity.

5.2.4 Correction Mass

The correction mass Dmc to be eventually attached at the radius r and a phase hc
should produce vibration A to cancel the original vibration A:

Dmc jhc Dmc jhc A


A Ga jX re ! re 5:11
m m Ga jX

The solution of Eq. (5.11) is too perfunctory. Instead, we discuss here how the
trial masses nally should be modied as the correction mass. Dividing Eq. (5.11)
by Eq. (5.10),
 
A
Dmc rejhc Dmt rejht Dmt rejht aejha 5:12
BA

where A=B  A aejha .


116 5 Unbalance and Balancing

Therefore,

Dmc aDmt ; hc ht ha 5:13

Equation (5.13) indicates that multiplying the trial mass by a and shifting the
phase by ha will yield the effective vector of A, being sought. This complex
calculation is needed only at a single rotational speed near the critical speed, and
can readily be performed with a scale and a protractor. The procedure is shown in
the following example.
Example 5.3 Modal balance Consider a rotor with an unbalance eccentricity
e 1 mm and a phase h 30 . The Nyquist plot for the unbalance vibration of
this rotor is shown in Fig. 5.11, with the point A for the rotational speed
p X=xn 0:95. If the unbalance is unknown, how can it be identied?
Answer
The initial direction of the Nyquist plot indicates the unbalance phase. Since the
gure suggests h 30 , it is plausible to attach a mass in the opposite direction,
i.e. around 150 . Here we begin with a trial mass Dmt at a radius r (equivalent to
an eccentricity of Dmt r=m 0:3 mm) and an angle (Fig. 5.12a). This yields the
Nyquist plot Fig. 5.11, with the point B for rotational speed p = 0.95. The
reduced circle suggests that the phase is fairly good, but the mass is not sufcient.
Since the attachment of the trial mass moves the initial vibration vector from
point A to point B, the effective vector is AB [corresponding to B A in
Eq. (5.10)]. This can be made the ideal vector AC directed to the origin [corre-
sponding to A in Eq. (5.11)] by multiplying the magnitude by |AC|/|AB| and
shifting the phase by \BAC . Since the segment AC is in the clockwise direction
with respect to the segment AB, the angle is negative. In this case, therefore,

a jACj/jABj 3; hc \ABC 20 :

By replacing the trial mass by a correction mass Dmc 3Dmt at the phase
130  20 150 ; as shown in Fig. 5.12b the vibration is minimized, i.e. bal-
ancing is completed as shown in Fig. 5.13. The gure shows the resonance

Fig. 5.13 Balancing process


5
1 initial
4
2 trial
amplitude

3
3 correction
2

0
0 0.5 1 1.5 2 2.5 3
p= n
5.2 Field Single-Plane Balancing (Modal Balancing) 117

curves corresponding to these steps. The initial unbalance was thus identied as
0.9 mm eccentricity and 150 180 30 phase.
Note that, as shown in this example, we need three rotation tests for vibration
measurement, i.e. the initial run, trial run and correction run. Some people call them
three balancing runs.

5.3 Balancing by the Influence Coefcient Method

The influence coefcient method described previously for single-plane balancing


can now be extended to multi plane balancing. Consider a system shown in
Fig. 5.14 in which vibration is measured at n sensor points for balancing m cor-
rection planes. The initial vibration A0 = amplitude \ phase is measured at a
rotational speed close to the critical speed:

A0  A10 A20    An0 t 5:14

Then the vibration A1 when a trial mass Wt1 = mass \ phase is attached to the
correction plane 1 is measured at the same rotational speed:

A1  A11 A21    An1 t

The change in vibration by attaching the trial mass Wt1 to the correction plane 1,
i.e., the effective vector DA1 is

DA1  A1  A0 A11  A10 A21  A20  An1  An0 t

Dividing this by the trial mass gives the influence coefcients a1:

a1  a11 a21  am1 t DA1 =Wt1 A1  A0 =Wt1

where, in aij , i is the sensor number and j is the correction plane number.
Repeating this procedure for each correction plane gives all the influence
coefcients

U1 U4 .....
correction plane U2 U3 Um
m

S3
.....
S1 S2
vibration sensor
n
Sn

Fig. 5.14 Number of correction plane m and vibration sensor n


118 5 Unbalance and Balancing

a1 a2    am :

Since balancing consists in attaching a correction mass Wci to each correction


plane to cancel the initial vibration,

Ai0 ai1 Wc1 ai2 Wc2    aim Wcm i 1 n 5:15

or, in a matrix representation,

A a Wc 5:16

where a  a1 a2    am   aij  (influence coefcient matrix) and

Wc  Wc1 Wc2    Wcm t correction masses:

The solution of Eq. (5.16) is straightforward if n = m:

Wc a1 A0 5:17

If n > m, the least squares method can be used:

Wc at a1 at A0 5:18

If n < m, sensors should be added or correction planes decreased so that n


m.
The correction mass distribution can be represented by a ratio Hc, which describes
how many times and how many degrees the trial mass distribution attached to each
correction plane has to be multiplied and advanced. Hc is obtained by replacing

a ! K DA1 DA2    DAm ; Wc ! H c 5:19

and using Eq. (5.16) through Eq. (5.18). The correction masses are determined as
2 3
Wt1 0 0 0
6 0 Wt2 0 0 7
6 7
Wc 6 .. .. .. .. 7Hc 5:20
4 . . . . 5
0 0 0 Wtm

The following should be noted when applying this balancing method:


(1) The method is not suitable for quick use on site because calculation involves
the inverse manipulation of complex matrices. However, a simplied bal-
ancing system is commercially available which requires a pocket calculator
only.
(2) Most of commercial systems based on this method are called computer-aided
automatic balancing systems.
5.3 Balancing by the Inuence Coefcient Method 119

1 050
150 600 150 #3
sensor sensor

50
S1 #2* S2

200

300
kb cb #1 kb cb 60
#2

k b = 10 7 N/m , cb = 10 3 N.s/m

Fig. 5.15 Rotor to be balanced

(3) We need m + 1 balancing runs of rotation tests for vibration measurement, i.e.
the initial run and m times of trial run. Vibration measurement in each bal-
ancing run should be made at exactly the same rotational speed.
(4) Vibration should preferably be read using the waveform passed a rotational
synchronizing lter rather than the raw (non-ltered) waveform.
(5) Instead of the measurement of the rotational shaft vibration, balancing is also
possible, in principle, by measuring pairs of pulse signals and acceleration
vibration waveforms transferred to the stationary housing of the machine.

Example 5.4 Consider the balancing of a rotor shown in Fig. 5.15 with three
correction planes #1, #2 and #3 by measuring vibration with two sensors S1 and S2
at the left and right bearings No. 1 and No. 2. Before balancing, the rotor vibration
was represented by a resonance curve indicated as Before shown in Fig. 5.16.
Complex amplitudes (amplitude [lm] \ phase  ) were measured at rotational
frequencies of 46 and 63 Hz. Table 5.2 summarizes the vibration records in 0
(initial run) to 3 (three trial runs). Conrm that the influence coefcient method
gives the correction masses for balancing Run 4 in the table.
Answer
The correction mass distribution is calculated from ratios of the trial mass
distributions:

Fig. 5.16 Comparison of 500


vibration before and after
correction 400
amplitude [m ]

sensor S1
300

200 Before
sensor S2
100 After

0
0 20 40 60 80 100
speed [Hz]
120 5 Unbalance and Balancing

Table 5.2 Vibration log l m\

Run No speed No1. bearing No2. bearing speed No1. bearing No2. bearing

Run 0 rotation test before balancing


unknown unbalance { #1, #2*, #3 } = 10g { 0, 90, 180} 100mm

A0 46Hz 34 73 73 110 63Hz 104 7 43 165

Run 1 measure the vibration with trial mass 5g 90 on correction plane #1

A1 46Hz 19 70 59 122 63Hz 105 18 41 +173


Run 2 measure the vibration with trial mass 5g 90 on correction plane #2

A2 46Hz 19 74 51 134 63Hz 104 5 42 175


Run 3 measure the vibration with trial mass 0.5g 180 on correction plane #3

A3 46Hz 34 76 76 113 63Hz 107 7 43 166

Run 4 rotation test for confirmation of balance Wc = { 10g 166 , 8g 90, 10g } 0.95 *

) Correction mass is 95% of calculated mass, because of being adopted to make clear peak amplitudes.

K  K1 K2 K3 
2 3 2 3
19\70  34\73 19\74  34\73
6 59\122  73\110 7 6 51\134  73\110 7
6 7 6 7
K1  6 7 K2  6 7
4 105\18  104\7 5
  4 104\5  104\7 5
41\173  43\165 42\175  43\165
2   3
34\76  34\73
6 76\113  73\110 7
6 7
K3  6 7
4 107\7  104\7 5
43\166  43\165
2 3
34\73 2 3
6 73\110 7 2\76
6 7 6 7
Eq: 5:16 ! 6 7 KHc ! Hc 4 1:5 5
4 104\7 5
20
43\165
2 3 2 3
5\90 0 0 10g\166
6 7 6 7
! Wc 4 0 5\90 0 5Hc 4 7:5g\90 5
0 0 0:5\180 10g

The curve representing the vibration after balancing is given in Fig. 5.16.
Example 5.5 Consider balancing of the same rotor under the same conditions, but
considering the in-phase and out-of-phase modes at each critical speed, with pairs
of trial masses (Fig. 5.17) using the three correction planes #1, #2 and #3. Table 5.3
summarizes the vibration records in each balancing run. The resonance curve before
balancing is the curve indicated as Before shown in Fig. 5.18. Complex
5.3 Balancing by the Inuence Coefcient Method 121

sensor S1 sensor S 2

#1 #2 #3
1 1 1
j 90
Run 1 5g 100 mm e

2 j 180
Run 2 1g 100 mm e
1

Fig. 5.17 Set of correction masses

Table 5.3 Vibration log lm\

Run No speed No1. bearing No2. bearing speed No1. bearing No2. bearing

Run 0 rotation test before balancing ( unknown unbalance is same as Table 52 )

A0 46Hz 34 73 73 110 63Hz 104 7 43 165


Run 1 measure vibration with trial masses = { 1, 1, 1 } 5g 90 on each correction plane

A1 46Hz 9 110 58 157 63Hz 104 3 44 +170

Run 2 measure vibration with trial masses = { 2, 0, 1 } 1g 180 on each correction plane

A2 46Hz 33 78 73 110 63Hz 90 7 36 160


rotation test for confirmation of balance
Run 3
Wc = { 12.4g 156, 4g 55, 9.2g 14 } 0.95

amplitudes (amplitude [lm] \ phase  ) were measured at rotational frequencies of


46 and 63 Hz. Dene the correction mass distribution.
Answer
The correction mass distribution is calculated as ratios to the trial mass distribution
again. From Eq. (5.16),
2 3 2 3
34\73 9\  110  34\73 33\78  34\73
6 73\110 7 6 58\  157  73\110 73\110  73\110 7
6 7 6 7
6 76 7H c
4 104\  7 5 4 104\3  104\  7 90\  7  104\  7 5
43\  165 44\170  43\  165 36\  160  43\  165
2 3 2 3
  1 2   12:3g\  15
0:77\35 6 7 5\  90 
0 6 7
! Hc ! Wc 4 1 0 5 Hc 4 4g\  55 5
7\8 0 1 
1 1 9:2g\  14

This answer is slightly different from the answer obtained in Example 5.4.
The curve representing the vibration after balancing is given in Fig. 5.18.
122 5 Unbalance and Balancing

Fig. 5.18 Comparison of 500


vibration before and after
correction 400

amplitude [m ]
sensor S1
300

200 Before
sensor S2
100 After

0
0 20 40 60 80 100
speed [Hz]

5.4 Modal Balancing

A single eigenmode is predominant in each critical speed. This method eliminates


only the n-th modal unbalance that induces such a critical mode [27].
For a beam shown in Fig. 5.19a, for example, balancing against vibration near a
critical speed corresponding to the rst bending mode can be performed without
affecting the second mode by choosing the midpoint of the shaft as the correction
plane. Similarly, for the second mode, correction at two planes in out-of-phase
(couple balancing masses) is recommended. Alternatively, balancing in the in-phase
plane for the rst mode and in out-of-phase plane for the second mode is also
possible (Fig. 5.19b). The pair ratio of the correction masses should be determined
beforehand taking the mode shapes to be balanced into account.
In practice, balancing work is performed for each consecutive critical speed
mode, e.g., 1st, 2nd, , within the rated rotational speed range according to a
predetermined pair ratio. The influence coefcients are identied using polar plots,
etc., to calculate the correction masses.

(a) (b)

1st mode 1st mode

2nd mode 2nd mode

3 planes required 2 planes required

Fig. 5.19 Typical modal correction masses


5.4 Modal Balancing 123

1
1st 1.0

2nd
shape of mode 0.5

0
200 400 600 800 1000 [mm]

3rd
1 #1 #2 #3
sensor
sensor S1 S2

kb 5cb kb cb
k b = 10 7 N/m , cb = 10 3 N.s/m

Fig. 5.20 Predicted mode shapes

Example 5.6 In order to reduce the resonance sensitivity of the second mode as
compared with Figs. 5.15 and 5.16, the rotor system is modied by increasing cb by
5 times at the left bearing as shown in Fig. 5.20, which includes the expected
eigenmode of each order. Apply the modal balance to this rotor as the amplitude
should not exceed 150 lm.
Answer
Let us start with the balance mass ratio of {1, 1, 1} at {#1, #2, #3} (i.e. in-phase type)
for the rst mode. The ratio to the masses for the second mode {h1, 0, h3} at { #1, #2,
#3}, which should not affect the rst mode, is determined as out-of-phase as follows:
the inner product of the rst mode and {h1, 0, h3} = 0
0:5h1 1:0h3 0 ! h1 2h3 ! the ratio to the mass distribution for the
second mode = {2, 0, 1}.
After starting the rotation under this condition, rst, the balancing Runs 0 and 1
shown in Table 5.4 are executed. The resonance curves of Run 0 are shown in

Table 5.4 Vibration log lm\

Run No description ( modal balance for first critical speed )

rotation test before balancing ( unknown unbalance is same as Table 52 )

measure resonance curve, Fig.521(1) Run 0 (Before),


Run 0
run up to 48Hz which approximately equals to 1st critical speed

measure polar plot of sensor S 2 , Fig.522 Run 0, choosing as larger vibration

attach trial masses = { 1, 1, 1 } 5g 90 on each correction plane


Run1
measure polar plot of sensor S 2 , Fig.522 Run 1, up to 48Hz
124 5 Unbalance and Balancing

200
S1
160
amplitude [m ]

S2
120 (After)
S1 S2
S2 Run 4
80 (Before)
S1 Run 2
Run 0 S2
40 S1

Run 2
0
0 20 40 60 80 100 0 20 40 60 80 100
speed [Hz] speed [Hz]
(1) 1st mode balance (2) 2nd mode balance

Fig. 5.21 Comparison of before and after balancing process

Run 0 S2
[ m ]
48 Hz
49 100 A
+35

46
48 Hz
B 50

Run 1 44 AC
Ratio = 0.72+35
AB
S2 55 Hz
49
C
100 50 55 Hz 50 100 [ m ]
44
46

Fig. 5.22 Vibration from sensor S2 (around 1st critical speed)

Fig. 5.21(1). Comparison of the polar plots for balancing Runs 0 and 1, as shown in
! !
Fig. 5.22 giving AC = AB  0:7\ 35 , indicates that the in-phase trial masses
should be multiplied by 0.7 and the phase should be shifted by 35 in the rotational
direction.
So the correction mass distribution of 5 g  0:7 3:5 g is attached at
90 35 55 , that is f 1 1 1 g3:5 g\55 at { #1, #2, #3} (in-phase).
After starting the rotation under this condition, the balancing Runs 2 and 3 in
Table 5.5 are executed. The resonance curves of Run 2 are shown in Fig. 5.21(2).
Comparison of the polar plots for the two balancing Runs 2 and 3, as shown in
! !
Fig. 5.23 giving AC = AB  2\  50 , indicates that the out-of-phase trial masses
5.4 Modal Balancing 125

Table 5.5 Vibration log lm\

Run No description ( modal balance for 2nd critical speed )

rotate again with the correction masses for 1st mode = { 1, 1, 1 } 3.5g 55

passing the 1st critical speed as shown in Fig.521


Run 2
run up to 68Hz which approximately equals to 2nd critical speed
measure polar plot of sensor S 1 , Fig.523 Run 2, choosing as larger vibration

attach trial masses = { 2, 0, -1 } 3g 130 on each correction plane


Run 3
measure polar plot of sensor S 1 , Fig.523 Run 3, up to 68Hz

Fig. 5.23 Vibration from [ m ]


sensor S1 (around 2nd critical
speed) 50
AC
Ratio = 2.250
AB
48
50 C 50 100 150 [ m ]
48

63
50 63
Run 3 Run 2
S1 S1
66
100 66
0

68 Hz
5

B
150 68 Hz
A

should be multiplied by 2 and the phase shifted by 50 in the anti-rotational


direction.
Finally, the correction mass of 3 g  2 6 g is attached at 130  50 180 ,
that is f 2 0 1 g6 g\180 at { #1, #2, #3} (out-of-phase), as indicated in
Table 5.6. They are attached and a balancing Run 4 is executed. The curve of the
vibration amplitude after balancing demonstrates that the system is now in a good
state of balance, as shown in Fig. 5.21(2) Run 4.
Table 5.6 Vibration log lm\

Run No description ( rotating test to confirm final balance )

attach correction masses for 2nd mode = { 2, 0, -1 } 6g 180


Run 4 measure resonance curve, Fig.521(2) Run 4 (After)
confirm well balanced state with small amplitude
over the complete speed range
126 5 Unbalance and Balancing

The total of the correction mass distribution attached is the sum of those in the
Runs 2 and 4:

Wc f 1 1 1 g3:5 g\55 f 2 0 1 g6 g\180 ;


10:4 g\164 3:5 g\55 8:5 g\20 

which is approximately the same solution as in Example 5.5.

5.5 n-Plane Balancing or (n + 2)-Plane Balancing?

5.5.1 Comparison

Two methods have been proposed to determine the number of correction planes
needed for balancing an elastic rotor to be rotated up to the n-th critical speed:
(1) n-plane method: Since shaft vibration can be minimized in each mode at least
by single-plane balancing, we need n-planes for the rotation up to the n-th
critical speed, as treated in examples above.
(2) (n + 2)-plane method: In addition to shaft vibration, bearing reaction force
may be also minimized in each critical speed. This needs two more correction
planes for balancing of the rigid rotor [28, 29].
Only one action using a single correction plane sufces for reducing unbalance
vibration predominant near the critical speed corresponding to an eigenmode. If
n critical speeds have to be passed, or operation near to them is required, n cor-
rection planes are needed, as shown in Fig. 5.24.
Consider a rotor with simple supports at both ends and uniformly distributed
eccentricity e shown in Fig. 5.25a. Balancing for the half-sine mode shape at the
rst critical speed can be made, as the case of n = 1, by attaching a correction mass
Uc on the correction plane at the midpoint of the shaft:

3 plane method or 3+2 =5 plane method


amplitude

speed rated speed

Fig. 5.24 n-plane method and (n + 2)-plane method balancing


5.5 n-Plane Balancing or (n + 2)-Plane Balancing? 127

Fig. 5.25 Example of R1 = 0.18 m 2 R2 = 0.18 m 2


balance of uniform shaft (a)

U c = 0.64 m

n plane method
R1 = 0 R2 = 0
(b)


U 1 = 0.18 m Uc = 0.64 m U 2 = 0.18 m

(n + 2) plane method

ZL
x 2qALe
Uc qAesinp dx 0:64me m qAL 5:21
L p
0

The correction mass Uc is certainly effective in reducing the vibration at the rst
critical speed.
As for the loads R1 and R2 indicating the bearing reaction forces, they are
obtained from the equation of static equilibrium, because the unbalance vibration is
a problem of static mechanics seen in a rotating coordinate system:
2 L 3
Z  
X2 4 L5 1 1
R1 R2 xqAedx  Uc  meX2 0:18meX2 5:22
L 2 2 p
0

The reaction forces persist though smaller than 0:5meX2 before balancing. They
can be canceled out by addition of two more correction masses, U1 and U2, for
compensating the residual R1 and R2 of the reaction forces, attached nearby the two
bearings. As shown in Fig. 5.25b, balancing with n + 2 = 3 planes:

fU1 ; Uc ; U2 g f0:18; 0:64; 0:18gme

gives an ideal balance where both vibration and reaction force approach zero.
The (n + 2)-plane balancing thus eliminates bearing reaction forces and there-
fore prevents the transfer of rotor vibration to the foundation. It is a method suitable
for equipment sensitive to floor vibration, such as home electric appliances and
high-precision instruments. It also maintains good balance of the rotor even when
128 5 Unbalance and Balancing

U = rotary piston motor

upside

#1 bearing #2 bearing S1

U = 0.8 g m
5 6
20 30 30 20 100

1 2 4

60
16

3
U1
kb cb kb cb U2 U3

k b = 10 7 N/m c b = 10 3 N s/m

Fig. 5.26 Calculated model of compressor rotor [VB117]

the spring or damping characteristics of the bearings change due to temperature


fluctuation or other factors. In other words, the (n + 2)-plane balancing is not
affected by boundary conditions and can therefore be regarded as a universal bal-
ancing method.
Example 5.7 [VB117] Figure 5.26 shows a compressor for an air conditioner,
which is a simplied model for calculation. It is assumed that the only unbalance
U is the eccentric mass of the rotary piston. Three correction planes are chosen: the
lower end of this vertical rotor U1, and the two ends of the motor U2 and U3.
Vibration is measured with a sensor S1 at the top of the motor for calculation of
influence coefcients.
The curves in Fig. 5.27 shows examples of the amplitude of journal vibration
at the bearing #1 and that of unbalance vibration measured with S1.
(1) Rigid body balancing with U2 and U3 sufces even for the rotor unbalance
U in conventional low-speed operation around 50 Hz. Determine U2 and U3.
(2) Demonstrate that this method does not prevent increase in vibration at a high
rotational speed around 90 Hz.
(3) Achieve the universal balance with three-plane balancing in the case of n = 1.
(Note n + 2 = 1 + 2 = 3 planes required.)

Answer
 
U U2 U3 0 U2 1:5U
(1) therefore (see Fig. 5.28)
50U2 150U3 0 U3 0:5U
(2) The curves and in Fig. 5.27 show the shaft vibration in the case of
before and after the rigid body balancing. Their comparison indicates that the
journal vibration #1 is slightly reduced by the balancing at lower speeds, but
5.5 n-Plane Balancing or (n + 2)-Plane Balancing? 129

50
#1
high-speed rotation
40
amplitude [m] low-speed rotation
30
0 before balancing
#1
20
1 after rigid 2 plane
balancing S1
10 S 1

3 after 3 plane
#1
balancing
0
0 20 40 60 80 100 S1 120
speed [Hz]

Fig. 5.27 Unbalanced shaft vibration (sensor S1 and bearing #1)

U 0.5U

bearing bearing 1.5U

1 rigid 2 plane balancing

U U

bearing bearing 2U

2 3 plane trial masses ratio

U 0.18U

0.32U bearing bearing 0.86U

3 3 plane balancing

Fig. 5.28 2-plane balancing and n + 2 = 3-plane balancing

rather increased at higher speeds. The same fact is seen in the bearing reaction
forces of Fig. 5.29 and .
The vibration vector at 90 Hz as measured by the sensor S1, before and after
rigid body balancing, must be recorded here for calculation of the influence
coefcient.
S1 vibration before the balancing: A0 29 lm\13 at 90 Hz
S1 vibration after the balancing: A1 118 lm\7 at 90 Hz.
130 5 Unbalance and Balancing

200
high-speed rotation
180
low-speed rotation
reaction force [NdB] 160 g
ncin #1
#2
r e bala #2
140
befo
0 #1
cing
120
e balan
ne
ter 3 plan
pla 3 af
100 2
id #2
rig ng 2
er
80 aft lanci
1 ba #1
60
0 20 40 60 80 100 120
speed [Hz]

Fig. 5.29 Bearing reaction force

(3) The three-plane balancing is performed by superimposing a correction mass


U1 for the third plane on the existing two-plane correction masses (U2 and U3)
for rigid body balancing obtained above. The ratio of these three masses to be
attached without affecting the rigid balance is

U1 U2 U3 0
50U1 50U2 150U3 0

Then, fU1 ; U2 ; U3 g f1; 2; 1g (see Fig. 5.28).


The influence coefcient should be determined for the case of this ratio of
correction masses. When trial masses fU1 ; U2 ; U3 g f1; 2; 1gU are
then added, S1 vibration of A2 487 lm\6 at 90 Hz is measured.
Then, the correction ratio: Hc A1 =A2  A1 0:32\3  0:32:
The total correction masses are thus, as shown in Fig. 5.28:

fU1 ; U2 ; U3 g f0; 1:5; 0:5gU  0:32f1; 2; 1gU


f0:32; 0:86; 0:18gU 5:23

As seen in Figs. 5.27 5.29, the three-plane balancing reduces both


vibration and bearing reaction force over the entire speed range.
(4) (different solution) The modes for mode-synthesis transformation (Fig. 5.30)
can be used for this purpose:

d1 1:33 1 0:5 0 0:33 2 t


Rigid body mode: 5:24
d2 0:33 0 0:5 1 1:33 3 t

Simple support mode for bearings: /1 0:16 0 0:19 0 0:49 3:7 t


5:25
5.5 n-Plane Balancing or (n + 2)-Plane Balancing? 131

3.7
3 3
2
2
mode shapes 1.33
1.33 1 1
1
1 0.5
0.16 0.49
0
0.33 0.19
0.33
1
#1 #2
1
bearing bearing 2
0 50 100 150 200 mm

Fig. 5.30 Modes for mode-synthesis transformation

Assuming correction mass distribution and unbalance in vector Wc,

Wc U 1 0 U 0 U2 U 3 t

therefore, the following inner products should vanish, if each modal excitation is
eliminated:

dt1 Wc dt2 Wc /t1 Wc 0 !


5:26
fU1 ; U2 ; U3 g f0:32; 0:86; 0:18gU

5.5.2 Number of Correction Planes Needed for Universal


Balancing

An interpretation of (n + 2)-planes is presented in terms of a critical speed map. In


an example of the map shown in Fig. 5.31, the abscissa is the normalized bearing
stiffness and the ordinate the critical speed normalized by the rst natural frequency
of free-free bending mode. The curves represent the rst, second, third, order
critical speeds, respectively. The left soft support system corresponds to a magnetic
bearing, the middle a sliding bearing, and the right-rigid support system of a ball
bearing, for example.
The rst and second order critical speed curves have a straight line at the left,
corresponding to the rigid body modes, and a saturated part at the right, corre-
sponding to the bending modes with simple-simple support boundaries. Regarding
the rst and second curves, the beginning of curvature in the middle means that the
eigenmode transits from the rigid modes to bending modes. The third and higher
order curves show a transition of the bending mode from free-free support
boundaries at the left to simple-simple support boundaries at the right in the gure.
The dotted curves in the gure indicate the number of correction planes needed
from the viewpoint of (n + 2)-plane balancing, which are drawn by considering
132 5 Unbalance and Balancing

n plane method ( n +2) plane method

5
5 planes
4th 4 4
4 4 planes
3rd 3
(2+2)
1

0.5 3 3 3 planes
critical speed

(1+2)

2 planes 2
2 planes
0.1
2 planes (0+2)
0.05
ed
d
2n

pe
ls
ica
rit
tc
1s

0.01
Case No. 1 2 3 4 5 6 7 8 9 10
bearing rigidity (logarithmic axis)

Fig. 5.31 Correction planes needed for universal balancing

how many bending mode components are included to pass each critical speed. The
height of the vertical bars corresponds to the possible range of the operational
speed. The low-speed operation in Example 5.7(1) is Case 7 in the gure, for which
two-plane balancing is enough, while the high-speed operation in Example 5.7(3) is
Case 8, for which three-plane balancing is recommended.
The left part of the map applies to rotor systems with soft supports. Here the
number of correction planes equals to that of critical speeds and, in this aspect, may
be said to be the same as the n-plane method [26, 27], stated in Sect. 5.5.1.

5.5.3 What Is the 2 in the (n + 2)-Plane Method?

The 2 in the (n + 2)-plane method is generally considered to be the number of


correction planes needed for balancing a rigid rotor supported by two bearings.
Then a question arises: do we need to add only 2 planes even to a rotor supported
by three bearings? The answer is no, because the 2 should be understood as the
number of bearings involved.
The (n + 2)-plane method should rather be read as the (n + number of bearings)
[23] plane method. So it becomes (n + 3)-plane method for a rotor supported by
three bearings.
t
According to the mode synthesis theory, the modal external force Fd F/  in
Eq. (4.13) with an unknown unbalance U(x) is expressed as
5.5 n-Plane Balancing or (n + 2)-Plane Balancing? 133

dt1 Ux; dt2 Ux. . .dtm Ux ! 0


5:27
for the deflection modes m number of bearings

/t1 Ux; /t2 Ux    /tn Ux ! 0


5:28
for the bending mode with simple support bearings

The universal balance, being independent of the dynamic characteristics of


bearings, is attained when these values are all zero. For this purpose of satisfying
Eqs. (5.27) and (5.28), the number of correction planes is clearly n + m, i.e.,
n + number of bearings, as stated above.
It is also known that, for m = 2 or more, if Eq. (5.27) is satised, rigid body
balance is automatically attained.
Note: The different solution to Example 5.7 determines the correction mass
distribution using the transformation mode for mode synthesis, which should be
unique for given bearing positions, i.e., they are dependent solely on the rotor
geometry and not on the bearing constants. In contrast, the correction mass dis-
tribution obtained using influence coefcients is dependent on the bearing dynamic
coefcients as far as it is derived from vibration measurement.
(1) However, the (n + 2)-plane balancing using the influence coefcients should
eventually yield the same result as the universal balance (vibration ! 0 and
bearing force ! 0) regardless of the bearing characteristics.
(2) Balancing based on eliminating the right hand side of the mode synthesis
model of Eq. (4.13) is intended for compatible reduction of journal vibration
and shaft bending vibration. It makes the shaft vibration and bearing reaction
force approach zero, simultaneously.

Example 5.8 Consider a rotor with two bearings with eccentricity e uniformly
distributed between 0 and 90 (Fig. 5.32a). The shaft vibration is to be measured
with a sensor S1 at the midpoint of the shaft to determine the influence coefcient.
The left end Wc1, the midpoint Wc2, and the right end Wc3 are chosen as the
correction planes.
The curves for the S1 resonance of unbalance vibration and bearing reaction
force under these conditions are shown as curve (: before balancing) in Fig. 5.33.
(1) Single-plane balancing using the influence coefcient based on the measure-
ment of vibration at the midpoint with S1 at X 7 leads to the correction mass:
Single-plane fWc1 ; Wc2 ; Wc3 g f0; 0:72\45 ; 0gme (see Fig. 5.32c).
Reconrm the calculations result of the curve in Fig. 5.33.
(2) Three-plane balancing using the influence coefcient based on the measure-
ment of vibration at the left bearing, midpoint (S1) and right bearing at X 7
leads to the correction masses.
Three-planes fWc1 ; Wc2 ; Wc3 g f 0:2\6 ; 0:6\45 ; 0:2\84 gme
(see Fig. 5.32d).
Reconrm calculations result of the curve in Fig. 5.33.
134 5 Unbalance and Balancing

(a) 1 1 2
u( ) (b)

S1
l
(c)
k c k c
-0.72 45
Wc 2
-0.2 84 ( m )
W c3
(d)
48 EI
k= 3
c = k /100
l
U( ) = Al Exp ( j 90 ) Wc 1 Wc2
-0.2 6 -0.6 45 ( m )

Fig. 5.32 n = 1-plane balancing and n + 2 = 3-plane balancing

(a) (b)
) [ NdB ]

dB
35 60
1 before balancing
amplitude [ m ]

30 1 before balancing 40
25 20
3
EI

20 2 1 plane balancing 2 1 plane balancing


l

0
15
reaction force (

10 3 3 plane balancing 20
5 40
3 3 plane balancing
0 60
0 2 4 6 8 10 0 2 4 6 8 10

1 EI 1 EI
speed speed
l3 A l3 A
shaft vibration bearing reaction force

Fig. 5.33 Shaft vibration and bearing reaction force

(3) Using the mode for mode synthesis model shown in Fig. 5.32b, perform
three-plane balancing and conrm that the result is the same as in (2) above.

Answer
(1) and (2) omitted.

(3) Using Un qAleej90 n; d1 n; d2 1  n, /1 sinpn; m qAl and
n x=l
2 3 2 32 3 2 3 2 3
Z1 Und1 1 0:5 0 Wc1 Wc1 0:19\6
4 Und2 5dn 4 0 0:5 1 54 Wc2 5 0 ! 4 Wc2 5 4 0:6\45 5me;
0 Un/1 0 1 0 Wc3 Wc3 0:19\84

which is the same as (2) above.


5.5 n-Plane Balancing or (n + 2)-Plane Balancing? 135

Fig. 5.34 n + 3 = 5-plane


balancing

l l

kb cb kb cb kb cb

Wc 1 Wc 2 Wc 3 Wc 4 Wc 5
0.188 0.6 0.425 0.6 0.188 Al

Note: See Fig. 5.33. Single-plane balancing successfully reduces the shaft
vibration to zero, but reduction of the bearing reaction force by around 20 dB may
not be sufcient. Three-plane balancing reduces this by 60 dB or more, thus
achieving an ideal balance, where both vibration and bearing reaction force
approach zero.
Example 5.9 A rotor supported by three bearings with eccentricity e uniformly
distributed in the shaft (Fig. 5.34) is given. Taking into account the bending mode
up to the second order (n = 2) for simple support at the bearing portions, perform
(n + 3 = 5)-plane balancing. Use the correction planes Wc1, Wc2, , Wc5 shown in
the gure.
Answer
Deflection modes d1 ; d2 ; d3 and bending modes /1 ; /2 for simple support at the
bearings are determined (Fig. 5.35). The correction mass distribution is calculated
as follows:
2 3 2 32 3
d1 1 0:41 0 0:09 0 Wc1
6 7 6 76 7
Z2l 6 d2 7 6 0 0:69 0 0:69 0 76 Wc2 7
6 7 6 76 7
qAe6 d
6 37
7dx 6 0 0:09
6 0 0:41 17 6 7
76 Wc3 7
6 7 6 76 7
0 4 /1 5 4 0 0:96 0 0:96 0 54 Wc4 5
/2 1 0 0 1 0 Wc5
2 3 23
Wc1 0:188
6 7 6 7
6 c2 7 6 0:6 7
W
6 7 6 7
0!6 7 6 7
6 Wc3 7 6 0:425 7qAle
6 7 6 7
4 Wc4 5 4 0:6 5
Wc5 0:188

The sum of the correction masses is equal to 2qAle, which is also the sum of
unbalance distributed in the rotor. The moments are also balanced so that rigid body
balance is thus achieved automatically.
136 5 Unbalance and Balancing

1 1

amplitude
modes of 2
deformation 0.5
3 bearings 3
0
0.5 1 1.5 2l

1
1
0.5
amplitude
bending modes
simply supported 0
at bearings 0.5 1 1.5 2 l
0.5
2
1

Fig. 5.35 Modes for mode-synthesis transformation

5.6 Balancing of a Rotor Supported by Magnetic Bearings

5.6.1 Balancing by Feed-Forward (FF) Excitation

The active magnetic bearing (AMB) is represented by the symbol shown in


Fig. 5.36. Electromagnetic force Fb, being proportional to the command signal
which the controller (CNTR), generated in association with the displacements
measured at S1 and S2, acts on the bearing journal. This feedback force maintains
the journal at the center of the bearing, thus keeping the rotor in stable levitation.
The characteristics of the controller are represented by a transfer function, of which
the specic form in this example is given in the gure (CNTR).

1 10 g 0
1 mode 0.62
0.63 AMB2
10 g 90

AMB1
0.59
0.88 10 g 180
2 j ( t + 1) 2 j ( t + 2 )
U1 e U2 e
+ +
U1 U2 U3
U4 F b1 U5 F b2

S1 S2

2 s +1
CNTR CNTR = 3 10 5 + CNTR
s s + 1
( )
at = 1/ 2 90 = 0.05

Fig. 5.36 FF excitation of magnetic bearing to compensate unbalance


5.6 Balancing of a Rotor Supported by Magnetic Bearings 137

When superimposing a whirling signal UejX h , synchronized with rotation on


the command signal, a whirling force acts on the bearing journal. When the
whirling signal multiplied by the square of the rotational speed is set for
UX2 ejX h as the centrifugal force signal, an exciting force acts on the journal as
an equivalent unbalance. This is a feed-forward (FF) force independent of the
displacement measured. Superimposition of FF force is a convenient method,
because it is relatively easy to perform and does not affect the stability of the
system.
This FF excitation, which is called open-loop balancing [36], can be utilized to
nd the influence coefcients for determining the correction masses for systems
supported by AMBs. This method is uniquely applied to AMB equipped rotors, as
explained below. If we install AMBs additionally in rotors supported by sliding
bearings, this FF balancing technique is also applicable for obtaining similar
benets.
A rotor supported by AMBs generally shows two critical speeds corresponding
to translation and tilt in the rigid body mode, because the support stiffness of AMB
is lower than, for example, that of sliding bearings. The resonance curve before
balancing is illustrated in Fig. 5.37. For balancing, FF excitation is executed at a
constant rotational speed lower than the critical speed and the influence coefcient
is determined.
In this example, the left and right bearings were excited at 8 and 13.5 Hz, just
below the rst and second critical speeds as indicated by the curve written by Run 0
(before balancing). This yielded the data shown in Table 5.7. After the vibration
vector A0 was measured before excitation, rotation was continued with FF excita-
tion at AMB1 by a signal equivalent to 5 g\45 , which was expected to reduce
vibration, and the vibration vector A1 was measured. This was followed by FF
excitation at AMB2 by a signal equivalent to 5 g\0 and the vibration vector A2 was
measured. All of these three measurements were done only at the before balanc-
ing rotation written as Run 0 in Fig. 5.37.

Fig. 5.37 Rigid modal 250


balance rigid mode
bearing amplitude [ m ]

200
13.5 Hz
150 before balancing,
8 Hz no excitation
after
100 Run 0 rigid balancing
S1
S2 Run 1
50

0
0 5 10 15 20 25 30
speed [Hz]
138 5 Unbalance and Balancing

Table 5.7 Vibration log lm\

Run 0 speed AMB 1 AMB 2 speed AMB 1 AMB 2

initial vibration ( around critical speed of rigid mode ) (unknown unbalance is same as Tab.52)
A0 8Hz 22 40 38 90 13.5Hz 101 56 29 118
FF excitation @ AMB 1 corresponding to 5g 45

A1 8Hz 21 18 31 92 13.5Hz 125 68 43 107

FF excitation @ AMB 2 corresponding to 5g 0


A2 8Hz 22 23 28 60 13.5Hz 86 58 26 132

corresponding to corresponding to
10.2 g -163 12 g -32
10.3 g -162 3.86 g -88 10.3 g -12

1 rigid modal U 4 U1 U2 U5 U3
balance

150 300 300 150 150

3.57 g 88

2 bending modal
balance

1.19 g 92 corresponding to 2.38 g 92


corresponding to 2.8 g 74
6.9 g 88
10 g 1
7.5 g -88

3 = 1 + 2
compositioned
correction mass
10 g -168

Fig. 5.38 Modication of balance masses

The correction masses for balancing the rigid body mode vibrations can now be
calculated. For this purpose, we rstly consider virtual correction masses {U4, U5}
which are obtained from the result during FF excitation at AMB potions, as shown
in Fig. 5.38. Then, {U4, U5} are converted to the correction masses {U1, U2, U3}
for normal 3-plane balancing.
5.6 Balancing of a Rotor Supported by Magnetic Bearings 139

The virtual correction masses {U4, U5} by AMBs are rstly calculated as
follows:
2 3
22\40
6 38\90 7
6 7
Eq: 5:16 ! 6 7 KHc
4 101\56 5
29\118
 
2:04\118
! Hc
2:40\32
     
U4 5\45 0 10:2g\163
! H c ! Fig: 5:38
1
U5 0 5\0 12g\32

where
2 3
21\18  22\40 22\23  22\40
6 31\92  38\90 28\60  38\90 7
K6
4 125\68  101\56
7
86\58  101\56 5
43\107  29\118 26\132  29\118

Next, the masses for three-plane balancing {U1, U2, U3} are calculated by the
equivalent conversion from {U4, U5} under the following three conditions.
To complete the rigid body balancing.
The total sum of the translation forces 0 ! U1 U2 U3 U4 U5 , and
The total sum of the moments 0 ! 300U1  300U2  600U3
450U4  450U5 .
To not affect the rst bending mode, whose shape is shown in Fig. 5.36:
the mode excitation force 0 ! 0:63U1  0:88U2 0:62U3 0:
The solution gives the three-plane balancing for the critical speed in the rigid
body mode:

f U1 U2 U3 g f 10:3g\162 3:86g\88 10:3g\12 g


! Fig: 5:38
1

These correction masses give the resonance curve shown in Fig. 5.37 as Run 1.
It is clear that a perfect rigid body balance has been attained.
Example 5.10 The system above exhibited the resonance curve shown in Fig. 5.39
as Run 1 at a rotational speed close to the bending mode. FF excitation at 152 Hz in
Run 1 was performed, which yielded the data shown in Table 5.8. Perform 3-plane
balancing for the bending mode using these data.
140 5 Unbalance and Balancing

Fig. 5.39 Bending modal 500


balance 152 Hz

bearing amplitude [ m]
400
rigid balance,
no excitaion bearing mode
300
Run 1
200
Run 2
S1 3-plane balancing
100
S2

0
100 120 140 160 180 200
speed [Hz]

Table 5.8 Vibration log l m\

Run 1 speed AMB 1 AMB 2

initial vibration ( around the critical speed of bending mode)


A3 152Hz 243 121 113 60

out-of-phase FF excitation { +1 , 1 } corresponding to 2g 90


A4 152Hz 119 120 57 62

in-phase FF excitation { +1 , +1 } corresponding to 4g 90


A5 152Hz 151 120 69 60

Answer

   
243\121 1:08\7
Eq: 5:16 !  KH c ! H c
113\60 1:21\6
          
U4 1 1 2\90 0 6:9g\88
! Hc ! Fig: 5:38
2
U5 1 1 0 4\90 2:8g\74

where
 
119\  120  243\  121 151\  120  243\  121
K
57\62  113\60 69\60  113\60

In the same way, these virtual correction masses {U4, U5} are equivalently
converted to 3-plane correction masses {U1, U2, U3} by the following three
conditions.
5.6 Balancing of a Rotor Supported by Magnetic Bearings 141

This bending mode balancing does not excite the translation and tilt rigid body
modes:
U1 U2 U3 0
300U1  300U2  600U3 0

Equalizing {U4, U5} obtained by FF and the 3-plane masses for the rst bending
mode:

0:63U1  0:88U2 0:62U3 U4  0:59U5

The solution gives the three-plane balancing for the third critical speed:

f U1 U2 U3 g f 1:19g\92 3:57g\88 2:38g\92 g ! Fig: 5:38


2

These balancing masses give the resonance curve shown in Fig. 5.39 as Run 2,
which shows a reduced vibration. Finally, the correction masses in Runs 1 and 2 may
be added for simplifying the correction mass distribution shown in Fig. 5.38.

5.6.2 Case Study: Centrifugal Compressor Supported


by AMBs [30, VB245]

A rotor system for a centrifugal compressor supported by AMBs, its peripheral


equipment and the control system are shown in Fig. 5.40a, and the details of the
control system in (b). The AMB control system is usually constructed as a closed
loop feedback system (FB/CNTR) with measured displacement {x, y} as the input
and the bearing force {Fx, Fy} as the output. The unbalance vibration of the system is
represented by the resonance curve in Fig. 5.41 and the polar plot in Fig. 5.42.
The system went through the two critical speeds for the rigid body (translation
and tilt) modes around 5,000 rpm with low vibration because the balancing was

(a) (b)
AMB AMB

x + AMB
FB/CNTR - X
+ Fx
gap double phase
FB levitation and function generator
FF excitation FF excitation
pulse sin/cos Fy
gap AMB control system +
FB/CNTR - Y
double phase monitor y +
pulse function generator
sin/cos
AMB system Control system ( levitation by FB and excitation by FF )

Fig. 5.40 AMB equipped compressor and bending modal balance


142 5 Unbalance and Balancing

100
1 Run 0 before balancing 1
2 Run 0 + FF excitation
amplitude [ m] 3 Run 1 after 3-plane balancing 2

50 ON

OFF
3

0
0 5 10 14
speed [ 1000 rpm]

Fig. 5.41 Resonance curves

Fig. 5.42 Polar plots 1 Run 0 before balancing


[m] 2 Run 0 + FF excitation
60 3 Run 1 after 3-plane balancing
trial excitation 1
Trial
40

1
trial excitation 2 ON excitation after optimal tuning
20

2
3

0 20 O 40 60 [m]
FF

complete, but not the third critical speed at about 11,500 rpm. In this situation, sine
and cosine signals, being synchronous with the rotational pulse, were generated by
a two-phase oscillator with the variable phase shifter. In addition to the centering
position control of FB/CNTR by AMBs, this FF signal was superimposed on the
controller output before power ampliers. The gain and phase differences of the
electromagnetic force synchronous with the rotation acting on the journal were
adjusted to reduce vibration. This FF excitation is illustrated by a trial (effective)
vector starting from the double circle ( ON) in Fig. 5.42, which was adjusted as
to forward the effective vector to the origin.
Figures 5.41 and 5.42 are the amplitudes of vibration under FF excitation
after the adjustment, which show the effect of such adjustment to reduce vibration.
The reduction obtained from the excitation data of this example indicated that the
correction mass for the bending mode balancing is equal to fU4 ; U5 g
f230; 230g g mm at the AMB portion as shown in Fig. 5.43a. The equivalent
conversion of Example 5.10 above was applied to the three correction planes {U1,
U2, U3} in the middle of the shaft, and a correction mass ratio of U1 :U2 :U3
0:55:1:0:  0:45 was chosen not to affect the rigid body mode balance. The
conversion calculation gave the correction masses: fU1 ; U2 ; U3 g f276;
512; 235g g mm. In practice, the balance is attained by grinding appropriate parts
of the impellers.
5.6 Balancing of a Rotor Supported by Magnetic Bearings 143

(a) correction planes


(b)
230 gmm 230 gmm
276 235 gmm
U1 U2 U3

AMB AMB
AMB AMB
U4 U5
3rd bending mode 512

Quantity of FF excitation balance Quantity of balancing correction

Fig. 5.43 Balancing masses

The curves for Run 1 in Figs. 5.41 and 5.42 are the result of the three-plane
balancing conrmed by rotating the shaft. Though a small vibration amplitude
remained, sufcient balance was generally achieved. FF excitation thus greatly
reduces the duration of the balancing process when using magnetic bearings.

5.7 Balancing without Rotational Pulses

5.7.1 Four Run Method [B22]

This section describes a balancing method using three trial runs without rotational
pulses, i.e., using the vibration amplitudes measured from the vibration waveform.
The balancing masses are determined by the amplitudes in totally four balancing
Runs, 0 to 3, called four-run method, where a trial mass is attached to the three
points on periphery of the rotor (Fig. 5.44) one by one.
Four amplitudes are measured:
Initial vibration amplitude |A0| before balancing
Vibration amplitude |A1| with a trial mass at a phase .
Vibration amplitude |A2| with the same trial mass at a phase leading by
120 .
Vibration amplitude |A3| with the same trial mass at a phase lagging by
120 .
The absolute values noted above are used to indicate that the phase of the
complex amplitude is unknown.

Fig. 5.44 Rotor section A 0 amplitude of initial vibration

2 U
A2
120

1 U

A1
120

A3
3 U
144 5 Unbalance and Balancing

The amplitudes are mutually related as

A0 aU A1

A0 aUej120 A2 5:29
j120
A0 aUe A3

where a is the influence coefcient. Rewriting these for the effective vector aU:

aU A0 A1
 
aU A0 ej120 A2 ej120 5:30
j120 j120
aU A0 e A3 e

Since the phase is actually unknown (replaced by the question mark?),


Equation (5.30) should be rewritten as

aU A0 jA1 jej?



aU A0 ej120 jA2 jej? 5:31
j120
aU A0 e jA3 je j?

Equation (5.31) can be solved graphically. It is hereafter assumed that |


A0| = 30 lm, |A1| = 49.0 lm, |A2 | = 18.1 lm, and |A3| = 34.3 lm in the case of
Fig. 5.45.
As shown in Fig. 5.46, A0 is the left-pointing vector and an |A0| radius circle are
drawn rst. A circle #1 centered on point A with radius |A1| is then constructed. The
right side of the rst equation of Eq. (5.31) should be on this circle. Similarly, the
right side of the second equation should be on a circle #2 centered on point B that is
120 apart from A, clockwise with radius |A2|, and the right side of the third equation
on a circle #3 centered on point C that is 120 apart from A, counterclockwise with
radius |A3|. Since each of the right sides equals to the effective vector aU, the three
circles should intersect each other at one as indicated by the point P, which means the
effective vector OP by attaching a trial mass U at the phase .
The correction mass can now be so chosen that the vector generates the initial
vibration A0, i.e., it points to A.

Fig. 5.45 Example of OA 2


calculation P 18.1
=
156 |A 0 |
= Wc = U
OP
49.0 0
12
1
12
0
A0 = 30 m 34.3
3
5.7 Balancing without Rotational Pulses 145

Im
18.1 m
#1 #2
B
49 A2
.0
A m direction
1
of rotation
= +156 P
A 0 Re
A O phase lag

30 m
A 0 = 30 m
A 1 = 49.0 34
C .3
A 2 = 18.1 m
A 3 = 34.3

A3
#3

Fig. 5.46 4-run method

The correction mass Wc should have a magnitude of U  jA0 j=jOPj and a phase
\POA apart from in the rotation direction. In this example, Wc 1:5U\156
apart from in the rotation direction.
Example 5.11 Figure 5.47a shows the same balancing operations as above
(Fig. 5.45) in a different run sequence. Determine the correction mass and conrm
that the result is the same as above.

Im

(a) (b) #2
m
3
A 0 = 30 m 2 34.3 .
34
Wc B
18.1
36 120 18.1 m
1 P
#1 30 m
120
Re
A A0 O
3 49.0 36 #3
m
Rotor section 49.0

Fig. 5.47 4-run method (in case of changing test run sequence)
146 5 Unbalance and Balancing

#1 Im

13.4 m

49.0
(a) (b)
#2

m
A 0 = 30 m 2 13.4 B

135 Wc P
49.0
1
21 +156
Re
90
A A0 O
Lag
30 m
3 42.3
42.3 m
Rotor section #3
C

Fig. 5.48 4-run method (in case of uneven angles)

Answer
The construction shown in Fig. 5.47b leads to Wc 1:5U\36 apart from in the
rotation direction.
Example 5.12 Consider a rotor section shown in Fig. 5.48a, to which a trial mass
is attached at three points on the periphery, sequentially. Assuming that |
A0| = 30 lm, |A1| = 49 lm, |A2| = 13.4 lm, and |A3| = 42.3 lm, determine the
correction mass Wc. Note that the angular locations of the trial mass are not
equidistant.
Answer
The construction shown in Fig. 5.48b leads to Wc 1:5U\156 .

5.7.2 Balancing by Placing a Trial Mass at a Regular Phase


Pitch

Suppose that the measured amplitude of rotor vibration is related to the angular
position, changed by a 15-degree increment, of a trial mass U as shown in
Fig. 5.49. The system is balanced at 150 where the amplitude is at a minimum.
The trial mass proves to be insufcient considering that the amplitude curve is in the
form jA0  initial amplitudej b. The balancing mass Wc here is 1:5U\150 .
5.7 Balancing without Rotational Pulses 147

Fig. 5.49 Rotor vibration is 100


related to the angular position. A0
80 balance mass = U

amplitude [ m ]
(Regular phase pitch) in the b
case of relatively smaller trial
60
mass phase
40 b

20
A0
0
0 120 240 360
rotor angle [ ]

Fig. 5.50 Rotor vibration is 100


related to the angular position. A0
80 balance mass = U
amplitude [ m ]
(Regular phase pitch) in the b
case of relatively larger trial
60 phase
mass
b
40

20
A0
0
0 120 240 360
rotor angle [ ]

Similarly, for an amplitude curve where the trial mass U is moved with a regular
phase pitch as shown in Fig. 5.50, the trial mass is excessive since the curve
indicates that |A0| < b in this case. The balancing mass Wc is here 0:83U\94 .

5.8 Solution of Two-Plane Balancing

Solution for balancing using two or more correction planes is difcult because it
requires calculation of the inverse of complex matrices. It is somewhat simplied,
however, by using scientic/electronic calculators available for matrix operations in
this section.

5.8.1 Principle of Calculation [B22]

Suppose balancing of a rotor shown in Fig. 5.51 was performed at the near (N) and
far (F) ends, which yielded the data shown in Table 5.9.
Run 0: initial vibration N and F measured
Run 1: vibration N1 and F1 measured with a trial mass Wtn attached to the N end
Run 2: vibration N2 and F2 measured with a trial mass Wtf attached to the F end
148 5 Unbalance and Balancing

W tf
side F
W tn
side N (Far)
(Near)

phase
phase

A = N1 N

N1 N = 4 20 O F2
N 0
0 F1
O B = F2 F
F
B = N2 N A = F1 F
N2
F = 6 300

Fig. 5.51 Calculation of 2-plane balancing and vibration vector (scale lm)

Table 5.9 Measured vibration data of 2-plane balancing

Run side N m side F m trial weight


No. amplitude phase symbol amplitude phase symbol g

0 4 20 N 6 300 F
1 6 110 N1 4 210 F1 Wtn = 5 g 0
2 8 290 N2 2 3 F2 Wtf = 5 g 0

Complex coefcients a and b are introduced as follows:

F1  F aN1  N  aA
5:32
N2  N bF2  F  bB

Representing the correction masses Wcn and Wcf to be attached to the two planes,
respectively, in terms of the ratios h and / to the trial masses:

Wcn h Wtn
5:33
Wcf / Wtf
5.8 Solution of Two-Plane Balancing 149

The ratios are dened by the condition so that the correction masses should
generate opposite values of the initial vibration amplitude:
         
N N1  N N2  N h N A bB h
 ! 5:34
F F1  F F2  F / F aA B /
      
h 1 B bB N 1 bF  N=A
) 5:35
/ 1  abAB aA A F 1  ab aN  F=B

Example 5.13 Determine the correction masses using the data on vibration mea-
sured at the two ends in Table 5.9.
Answer
Referring to vector manipulations in Fig. 5.51:

N 4\20 ; F 6\300
N1  N  A 7\144 ; F1  F  aA 7:5\155 ; a aA=A 1:07\11
F2  F  B 5:5\102 ; N2  N  bB 9\265 ; b bB=B 1:64\163
aN 4:28\31 ; bF 9:84\103 ; ab 1:75\174
bF  N 10\125
h 0:52\345
1  abA 2:75\356 7\144
aN  F 7:25\86
/ 0:48\348
1  abB 2:75\356 5:5\102
5:36

f Wcn Wcf g f hWtn /Wtf g f 2:6 g\345 2:4 g\348 g 5:37

5.8.2 In-Phase and Out-of-Phase Balancing

The vibration vectors of a rigid rotor were represented in the preceding section by
the amplitudes N and F actually measured at the right and left ends, but they can
also be represented by the in-phase (parallel mode) P component and out-of-phase
(tilting/conical mode) T component. The component P is the average of N and F,
and the component T indicates their difference, which is processed as shown in
Eq. (5.38):

P N F=2 N PT
, 5:38
T N  F=2 F PT
150 5 Unbalance and Balancing

Fig. 5.52 Representation of phase


vector by right/left and
in-phase/out-of-phase (scale F T N = P + T = 4 20
lm) F = P T = 6 300
N

0
O
P N+F
P = = 3.9 330
2
NF
T = = 3.3 83
F 2

Figure 5.52 shows the corresponding graphical solution, where addition or


subtraction of vectors is performed graphically.
Trial masses are also considered in pairs. The ratio of the right and left in-phase
trial mass is Wtp f 1 1 gwtp , and that for out-of-phase trial mass is
Wtt f 1 1 gwtt . The trial mass ratio is not uniquely dened; any ratio will do as
far as the two are independent. It is, however, usually convenient to use the in-phase
or out-of-phase proportion of related mode shapes.
Balancing using the in-phase/out-of-phase pair of masses is generally very useful
because either in-phase or out-of- phase unbalance are predominant in the majority
of cases. For example, Eq. (5.37) of the correction masses determined indepen-
dently for the F and N ends is equivalent to in-phase masses of about 2:5 g\346 .
Rigid body balancers provide an option of the N/F and P/T representations.
Example 5.14 Conrm the data in Table 5.10 for P/T balancing that are converted
from the data in Table 5.9 for N/F balancing:
Run 0: In-phase and reverse phase vibration before balancing
Run 1: P and T vibration with in-phase trial mass distribution
Wtp f1; 1g  5g\0
Run 2: P and T vibration with reverse phase trial mass distribution
Wtt f1; 1g  5g\0

Table 5.10 Measured vibration data of 2-plane (in-phase/out-of-phase) balancing

Run in-phase m out-of-phase m trial mass


No. amplitude phase symbol amplitude phase symbol g

0 3.9 330 P 3.3 83 T


1 3.8 180 P1 3.3 282 T1 Wtp = { 1 , 1} 5 g 0
2 4.0 116 P2 11 86 T2 Wtt = { 1 , 1} 5 g 0
5.8 Solution of Two-Plane Balancing 151

Answer 1

P N F=2; T N  F=2
Np N N1  N N2  N N1 N2  N; Fp F1 F2  F;
Nt N N1  N  N2  N N1  N2 N; Ft F1  F2 F;
P1 Np Fp =2; T1 Np  Fp =2;
P2 Nt Ft =2; T2 Nt  Ft =2

Answer 2
Using matrix forms
      
P 1 1 1 N N1  N N2  N
; K influence coefficient matrix
T 2 1 1 F F1  F F2  F
               
P1 1 1 1 N 1 P2 1 1 1 N 1
K ; K
T1 2 1 1 F 1 T2 2 1 1 F 1

Example 5.15 After measuring the P/T balancing data of Table 5.10, nd the
correction masses to be attached at the N and F planes and conrm that they are
equal to the answer of Example 5.13.
Answer
Letting the correction mass ratios h and / for in-phase trial mass Wtp, and that for
out-of-phase trial mass Wtt, respectively, the answer is given in a similar manner
through Eqs. (5.34) and (5.35):
         
P P1  P P2  P h P A bB h
 !
T T1  T T 2  T / T aA B /
      
h 1 B bB P 1 bT  P=A
)
/ 1  abAB aA A T 1  ab aP  T=B

Instead of matrix equation, we may nd the answer in a graphical manner.


Referring to Fig. 5.53, specic values are:

P 3:9\330 ; T 3:3\83
P1  P  A 7:4\165 ; T1  T  aA 6:6\273 ; a aA=A 0:89\108
T2  T  B 7:6\87 ; P2  P  bB 7:5\133 ; b bB=B 0:99\45
bT  P 7\141
h 0:52\348
1  abA 1:82\348 7:4\165
aP  T 0:32\16
/ 0:02\301
1  abB 1:82\348 7:6\87
5:39
152 5 Unbalance and Balancing

W tp W tt T2
1 1 1

5g 5g B
1
phase
P2
T = 3.3 83

B
P1
0 0
O O A
A P = 3.9 330 phase
T1

Fig. 5.53 Calculation of vibration vector (scale lm)

f Wcn Wcf g h Wtp / Wtt f 2:7g\346 2:5g\350 g 5:40

Note that the answer agrees well with Eq. (5.37).

Note: In-phase unbalance is predominant in this system as indicated by the large


h and the small /. As shown in this typical example, the P and T representation
generally provides easier physical understanding for balancing.
Note: Since rotor balancing is one of very traditional subjects in the eld of
turbo-machinery, many technical papers and experience reports are to be found.
The balancing theory is simple, but a wide variety of procedures exist in practice. In
this eld, Refs. [3151] provide further information and details on specic prob-
lems. The reader is referred to these to experience the range of problems that relate
to the techniques discussed in this chapter.
Chapter 6
Gyroscopic Effect on Rotor Vibrations

Abstract This chapter discusses the gyroscopic effect characterizing rotordynam-


ics as being different from the structural dynamics, associated with the non-rotating
parts of the rotor system, such as the casing and foundation. A top spinning at a
high speed whirls slowly in a tilted position. Similarly, a rotor of a rotating machine
whirls while rotating around the driven shaft axis. The spinning top does not fall
due to a moment, generated by the gyroscopic effect, which is proportional to the
rotational speed. This gyroscopic effect of a rotor system appears as the
self-centering tendency during rotation, which may be considered as an increase in
the centering stiffness. It is absolutely essential to understand the influence of the
gyroscopic effect on the natural frequency and the resonances in the frequency
response in rotating machinery vibrations.

Keywords Gyroscopic effect  Whirl motion  Forward and backward reso-



nances Unbalance excitation  Base excitation  Ball passing frequency

6.1 Rotordynamics

Vibration of a system may be assessed by structural dynamics as shown in Fig. 6.1.


The general form of the equation of motion of the vibration system consisting of
mass, spring (stiffness) and damping elements is

Mx Cx_ Kx Ft 6:1

where M is the mass matrix, K the stiffness matrix and C the damping matrix of the
structural system, and F(t) the external force acting on the system.
A unidirectional free vibration of the undamped M-K system is represented by

xt /a cos xt 6:2

where / is an eigenvector, called a normal mode, a is the amplitude and x is the


natural frequency.

Springer Japan 2017 153


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_6
154 6 Gyroscopic Effect on Rotor Vibrations

vibration in
one direction Y
straight line

Fig. 6.1 Featured vibration of structural system

Characteristics of free vibration of the entire M-K-C structure system may be


determined by complex eigenvalue analysis, which gives the natural frequency and
damping ratio for each eigenmode. For forced vibration, modal analysis allows
denition of single-dof responses for each mode having its own modal external
force and modal damping ratio. The sum of all single-dof responses represents the
vibration response of the entire system. Vibration occurs in the same direction as
the excitation.
Vibration of a rotating body is governed by the rotordynamics, which are
characterized by the gyroscopic effect accompanying the rotation (spin). This effect
features in the rotor vibration by a whirl motion in the XY plane, as seen in Fig. 6.2,
instead of a unidirectional motion. Regarding the XY plane as a complex plane, the
general form of the equation of motion is represented using the complex dis-
placement z = x + jy, comprising the displacement x in the X direction and y in the
Y direction. The equation of motion is as follows:

Mx XG_y Kx Kb x Cb x_ Fx t
! Mz  jXG_z Kz Kb z Cb z_ Fz t
My  XG_x Ky Kb y Cb y_ Fy t
6:3

whirling vibration

forward backward

= speed of rotor
z = x+jy =complex amplitude

Fig. 6.2 Featured vibration of rotor system


6.1 Rotordynamics 155

where X is the angular velocity of the rotor, M the mass matrix of the rotor system,
K the stiffness matrix of the rotating shaft, G the gyroscopic matrix of the system
(symmetric matrix). Kb and Cb are the stiffness and damping matrices of the
bearings supporting the rotor, and Fz t Fx t jFy t is the external force
associated with whirling motion.
The vibrational solution of Eq. (6.3) can be written as

zt  xt jyt /aejxt 6:4

Equation (6.4) describes the whirling motion of the shaft center as seen in a
stationary reference frame. The motion is said to be forward whirl if the whirling
and spinning directions are the same, and backward whirl if the whirling direction is
opposite to the spinning direction. The whirling direction corresponds to the sign of
the whirling frequency, as seen in Fig. 6.2. The forward and backward vibration
responses at resonance should therefore be different.

6.2 Gyroscopic Moment and the Motion of a Top

6.2.1 Gyroscopic Moment

Consider a top consisting of a disk with a shaft through its center. When it whirls,
while spinning around the shaft, the gyroscopic moment acting on the top prevents
it from falling (Fig. 6.3). If the friction at the support of the top and the air resis-
tance are both negligible, the top will process indenitely. The gyroscopic moment
is a conservative, as is the inertial force or the spring force.

Fig. 6.3 Gyroscopic moment


motion of spinning top

slope angle
(tilting mode)
156 6 Gyroscopic Effect on Rotor Vibrations

Consider a top spinning at an angular velocity X and whirling around the H axis of
an inertial frame XYH, xed in the space, with an angular velocity of precession, x,
at an inclination/tilting angle, h. Assume that the disk has a polar moment of inertia Ip
around the H0 axis of the coordinate system xed to the rotating body and a transverse
moment of inertia Id around the diametral axis through the center of gravity, and that
the distance between the center of gravity and the support of the shaft is h.
The gyroscopic moment,

Mg Ip Xhx 6:5

acts on the X 0 axis to bring the spin axis H0 closer to the whirl axis H, by decreasing
the tilting angle h.
Consider now the projection of the motion of the top onto the XY planes. The
displacement d and tilting angle h of the disk are dened in a similar manner to the
deformation of a beam. Indices x and y are given to the projection onto the X-H and
Y-H planes, respectively (Fig. 6.4), i.e. dx and hx : the displacement and tilting
angle of the disk projected onto the X-H plane; dy and hy : the displacement and
tilting angle of the disk projected onto the Y-H plane.
When the axis of the top is in the Y-H plane with u 90 , X 0 and X axes
coincide. Considering that
hhx xdy d_ x hh_ x gives hx h_ x , the gyroscopic moment is Mgx
Ip Xh_ x in Fig. 6.4b.
Since Mgx My ,

My Ip Xh_ x 6:6

Similarly, when the axis of the top is in the X-H plane with u 0 (the Y 0 and X
axes coincide), hhx xdx d_ y hh_ y gives hx h_ y which yields Mgy
Ip Xh_ y in Fig. 6.4a,
Since Mgy Mx ,

Mx Ip Xh_ y : 6:7

(a) (b) Mgx


Y X
x = h x y = h y
Mgy

x y
Mx My
X Y
Fx Fy
X- plane Y- plane

Fig. 6.4 Coordinates for rotordynamic system


6.2 Gyroscopic Moment and the Motion of a Top 157

6.2.2 Equation of Motion of a Top and Whirling Solution

Considering the transverse moment of inertia I1 Id mh2 seen from the support
point as well as the gyroscopic moment obtained above and the moment of gravity,
the equations of motion of a top are written as

I1 hx Mx mghhx I1 hx Ip Xh_ y mghhx


! 6:8
I1 hy My mghhy I1 hy  Ip Xh_ x mghhy

By adding the second equation of Eq. (6.8) with j (imaginary unit) to the rst
equation, and introducing the complex displacement:

h hx jhy 6:9

the equation of motion using the complex displacement becomes:

I1 h  jXIp h_ mghh 6:10

Writing the solution of this equation as

h hx jhy aejxt 6:11

the characteristic equation can be derived to determine the natural frequency x:

I1 x2  XIp x mgh 0 6:12

Although this equation has no real roots at low speeds, it has two real roots for
higher speeds, X (Fig. 6.5):

Fig. 6.5 Characteristic root


of spinning top

*
See references [B18] for the equation of motion of a top derived by Lagranges equation.
158 6 Gyroscopic Effect on Rotor Vibrations

q
XIp  XIp 2  4I1 mgh
x  fx1 ,x2 g [ 0 6:13
2I1

which exist for X2  4I1 mgh=Ip2 ; 0\x1 \x2 .


The two positive roots fx1 ; x2 g of the characteristic equation thus indicate the
angular velocity of possible whirling motions. They are seen as vibration in the X or
Y directions. Equation (6.13) corresponds to the empirical knowledge that a top
falls without whirling when spinning at angular velocities below a lower threshold
limit. We see the top whirling as the combination of the precession movement of
low frequency x2 and the nutation movement of high frequency x1 at the high
spinning frequency X over than a certain limitation.

6.3 Natural Vibration of a Rotor System

6.3.1 Natural Frequency of Whirling

Consider a single-disk rotor system akin to a top laid out along the horizontal axis
as shown in Fig. 6.6. The rotating shaft is assumed to be rigid and massless. The
left end, corresponding to the support of the top, is supported by a bearing as simple
support (pinned), while a bearing with a spring constant k is provided to the right
side. The disk is located outside the right-side bearing, hence the system is referred
to as an overhung rotor.
Instead of the gravity moment in Eq. (6.10), substituting the moment M around
the left end produced by the reaction force of the added spring:

M kl2 h 6:14

Fig. 6.6 1-disk rotor system

disk ( m , I d , I p , eccentricity G )
I 1 = I d + m lo2
6.3 Natural Vibration of a Rotor System 159

Fig. 6.7 Cylindrical rotor

the equation of motion for the rotor is obtained as:

I1 h  jXIp h_ kl2 h 0 6:15

The tilting angle h of the shaft (disk) can be converted to the displacement z of
the disk by multiplying by the arm length l0 : z l0 h. This conversion corresponds
to observation of the disk motion as a Lissajous orbit on XY plane traced by the
output signals of sensors placed at the X and Y directions. The equation of motion
is now written in terms of z as

I1z  jXIp z_ kl2 z 0 6:16

Therefore, the basic form of the equation of motion for the single-dof system
rotordynamics can be written as

z  jXc_z x2n z 0 6:17


p
where c Ip =I1 (gyroscopic factor) and xn kl2 =I1 (natural frequency of the
rotor at rest).
The transverse moment of inertia Id and polar moment of inertia Ip of a rotating
disk are illustrated in Fig. 6.7; Ip = 2Id for the thinnest disk. The range of the
gyroscopic factor c is therefore

0 c\2 6:18

Writing the whirling natural frequency as x in free vibration, the whirl motion is

z aejxt 6:19

where x xf [ 0 (forward natural circular frequency) or x xb \0 (backward


natural circular frequency, xb , itself is assumed to be a positive value).
160 6 Gyroscopic Effect on Rotor Vibrations

Fig. 6.8 Whirling motions of (b) Y


rotor
forward f
whirling

X
Y
(a)
Y

unidirectional
vibration n X
backward
b
whirling

= 0 ( at rest ) > 0 ( rotating state )

Substituting Eq. (6.19) into Eq. (6.17) yields the equation for the eigenvalue:

x2  x2n c Xx 6:20

The solutions are:


s
 
xf cX 2 cX
1 6:21
xn 2 xn 2 xn
s
 
xb cX 2 cX
1  6:22
xn 2 xn 2 xn

The whirling behavior of a rotor is represented as in Fig. 6.8. When the rotor is at
rest X 0, then x xn , i.e. the forward xf and backward xb natural frequencies
coincide as to be convergent to the natural circular frequency xn . When rotating, the
rotordynamics cause the separation of both the forward xf and backward xb natural
circular frequencies. The free vibration of the rotor system contains two whirling
motions, each of which has its own whirling frequency at different values.

6.3.2 Influence of the Gyroscopic Factor

The solutions of the eigenvalues Eq. (6.20) are shown graphically in Fig. 6.9. The
whirling natural circular frequency x is an intersection of a parabola with roots at
xn and a straight line with the gradient proportional to the gyroscopic factor c and
the rotational speed X. As the rotational speed increases, the intersection for the
forward natural circular frequency xf increases away from xn , while that for the
backward natural circular frequency xb moves toward zero away from xn (de-
creases in the absolute value). It is thus clear that, for whirling natural circular
frequencies during rotation, xf [ xb .
6.3 Natural Vibration of a Rotor System 161

Fig. 6.9 Solution of


eigenvalue equation

Fig. 6.10 Natural frequency 5


curves (xf and xb are

.0
=2
1.5
plotted) 4
f 1.0


3
0.5
/ n

2 0.25
0 = 2.0
1
1.0
1 2 3 4 5
0 0.5
0.25
1 0
b / n

This dependence of the whirling natural frequency on the rotational speed is


represented in Fig. 6.10 with the gyroscopic factor c as the parameter. The forward
natural frequencies appear on the positive side of the ordinate, and the backward
natural frequencies on the negative side. The former (solid curves) and the absolute
value of the latter (broken curves) are shown in Fig. 6.11, where the natural fre-
quency x is split into the upper and lower branches as the rotational speed
increases, reflecting the influence of the gyroscopic effect. A wide split means that a
large gyroscopic factor exists.
At very high rotational speeds, Eqs. (6.21) and (6.22) become xf and xb
asymptotes to the following straight lines:

xf ! c Xxn ; xb ! 0: 6:23

On the other hand, at low rotation speeds, the split of the curves is described
approximately by
c c
xf xn X; xb xn  X 6:24
2 2

Specically, for a very thin disk rotor (e.g. a hard disk drive having an eigen-
mode with a single nodal diameter) where c 2, the split is approximately
162 6 Gyroscopic Effect on Rotor Vibrations

Fig. 6.11 Natural frequency


curves (xf and xb are plotted)

xf xn X; xb xn  X 6:25

i.e., the forward natural frequency xf increases, and the backward natural frequency
xb decreases, by the rotational speed, resulting in a split width of xf  xb 2X.
A natural frequency curve without a split width means that the gyroscopic effect
is negligibly small, c
0, and therefore xn
xf
xb . This is nothing other than
representing the vibration of a structure. A long-shafting rotor with simple support
at both ends, as appropriate for a turbine rotor, is considered in this state for the rst
eigenmode. However, the gyroscopic effect may no longer be negligible for the
second eigenmode where the tilting motion of the disk at the midpoint of the shaft
must occur.

6.3.3 Calculation of the Natural Frequency of Whirling


in Multi-dof Rotor System

The equation of motion for free vibration of an undamped rotor system is derived
from Eq. (6.3):
Mz  j X G_z K Kb z 0 6:26

where Kb is the isotropic stiffness matrix of the bearings in the X and Y directions.
Substituting the vibrational solution

z /ekt 6:27

into Eq. (6.26) yields the general equation for the eigenvalue:

kBU AU 6:28
6.3 Natural Vibration of a Rotor System 163

where
 
M 0
B positive definite and symmetric: B Bt ;
0 K Kb
 
jXG K Kb
A skew-Hermitian matrix: A A  t and
K Kb 0
 
/k
U :
/

Since the eigenvalue is purely imaginary, k jx, the undamped natural fre-
quency x, which is positive for the forward whirling and negative for the backward
whirling, is obtained. The sign of the eigenvalue thus has a physical meaning in this
sense.
Note: Whirling motion of a gyroscopic system
If the equation of motion

I1 h  jXIp h_ kh h 0 6:29

has a solution

h aekt 6:30

the characteristic equation for k jx is

I1 x2  X Ip x kh 6:31

The left and right sides of this equation are shown graphically in Fig. 6.12, from
which the condition for existence of the purely imaginary root is clear.
In Eq. (6.31), a positive spring kh [ 0 to whirling of a rotor system , kh 0 to
attitude control of an articial satellite (equivalent to a top at zero gravity, ), and a
negative spring kh \0 corresponds to whirling of a spinning top in the gure.

6.4 Unbalance Vibration and Resonance

6.4.1 Condition for Unbalance Resonance


and Critical Speed

This section discusses the condition for resonance with unbalance vibration. It is
well known that resonance appears when the external forced frequency coincides
with the natural frequency of the system. For a rotor system, in addition to the
frequencies, resonance appears when the directions of whirling coincide.
164 6 Gyroscopic Effect on Rotor Vibrations

2
I1 Ip

k > 0 1 positive spring (rotor)


Ip

0 I1 zero gravity
k = 0 2
(artificial satellite)
k < 0 3 gravity (spinning top)

1 2 3

mg

Fig. 6.12 Whirling of gyroscopic system

Figure 6.13 presents the dependence of the natural frequency on the rotational
speed or other parameters. The solid line is the forward, and the broken line the
backward, natural frequency. An unbalance force can be regarded as a forward

resonance condition
f = > 0 only forward

pulse
natural frequency

pulse

rotor
unbalance force whirling trajectory
amplitude

direction X
direction Y
Letting F = X + j Y,


2 ( F speed
input : only forward force resonance : forward
whirling

Fig. 6.13 Unbalance vibration and resonance


6.4 Unbalance Vibration and Resonance 165

whirl. Force caused by unbalance magnitude Um e (eccentricity e multiplied by


mass m) or UDm r (unbalanced mass Dm xed at the radius r with respect to
the rotational axis) can be represented as a cosine function in the X direction and a
sine function in the Y direction. In the complex plane coordinate with a real X axis
and imaginary Y axis, it may be written as

m e X2 cos Xt j sin Xt m e X2 ejXt 6:32

This means that an unbalance force can thus be regarded as a forward whirl force
synchronized with the rotation. The frequency of the external force is X. As
expressed in the same manner with the natural frequency curve, the force can be
plotted as a straight line x X, especially for the solid line because it is in the
forward direction. The line is called the synchronous speed line.
Since rotor vibration resonance occurs when the frequency of the external force
coincides with the natural frequency of the system, including the whirling direc-
tions, the resonance point is indicated by the intersection of the solid line and the
solid curve, as shown in Fig. 6.13. The rotational speed at which the rotor resonates
with the unbalance vibration is called the critical speed Xc which may result in a
dangerous situation. At the critical speed, the forward natural frequency is equal to
the rotational speed. Coincidence of the backward natural frequency (the broken
line) with the rotational speed does not cause any response. Note that nothing
happens at the intersection. The resonance curve has therefore a peak at Xc , which
is somewhat higher than the natural frequency xn at rest.
The rotational speed that coincides with the natural frequency is determined by
substituting an assumed solution

z aejXt 6:33

into the equation of motion (6.17) for a single-dof rotor system, which includes the
influence of gyroscopic effect, to obtain the characteristic equation. The root of the
equation gives the critical speed Xc :
xn
X2 1  c x2n 0 ! Xc p 6:34
1c

For a thin disk rotor where 2 [ c 1, the critical speed does not exist since the
lines do not intersect.

6.4.2 Resonance Curves for Unbalance Vibration

The equation of motion for unbalance vibration without the gyroscopic effect is
presented in Sect. 2.3. The influence of the gyroscopic effect is now considered.
The basic equation for this situation is obtained by adding a term for the gyroscopic
effect to Eq. (2.31), leading to
166 6 Gyroscopic Effect on Rotor Vibrations

z  jXc_z 2fxn z_ x2n z e X2 ejXt 6:35

Using the complex amplitude A aeju and assuming the response

z AejX t  aeju ejX t aejX tu 6:36

the following formula is obtained:

e X2 e p2
A 6:37
x2n  X2 1  c j2fxn X 1  c 1  p2 2jbfp
p
where p X=Xc ; b 1= 1  c.
The amplitude and phase of unbalance vibration with the gyroscopic effect taken
into account are, therefore,

e p2
j Aj  a q
1c
1  p2 2 2fbp2
  6:38
1 2bfp
u  tan
1  p2

If the gyroscopic effect is neglected (c 0), i.e. p X=xn and b 1, the


unbalance vibration is represented by resonance curves with the damping ratio f as
the parameter, as shown in Fig. 2.17. If the gyroscopic effect is considered in the
range 0\c\1, the curve is dependent on two parameters: the damping ratio f and
gyroscopic factor c, as exemplied by Fig. 6.14. The peak in the resonance curve
corresponds to the critical speed Xc . The more prominent the gyroscopic effect is,
the higher the peak of the resonance amplitude and the longer the tail of the peak.
The ratio of the peak amplitude and eccentricity e, i.e., the Q-value, is

Fig. 6.14 Unbalance


resonance curves

c
6.4 Unbalance Vibration and Resonance 167

Fig. 6.15 Resonance curves


with gyroscopic effect

2/3

1/3
0

apeak 1 1 1 1
Q p 6:39
e 1  c 2bf 1  c 2f

In order to evaluate the exact Q-value, in addition to the damping ratio f, we


have to know the gyroscopic factor, c. For typically long-shaft rotors, which usually
have small gyroscopic factors, Q 1=2f is acceptable in the usual manner.
Figure 6.15 shows resonance curves of Eq. (6.38) including a wide range of the
gyroscopic factor. When c approaches unity in a short-shaft rotor, the forward
natural frequency and the synchronous speed are asymptotic, and the intersection is
innitely far away. Thus the resonance curve shows a large amplitude and a long
tail in the high frequency domain, and the resonance point is never passed. This
would be a very dangerous design. It is generally recognized that reduction control
of resonance amplitude is difcult for 0:5\c\1, no matter how carefully balancing
is conducted.
Very thin disk rotors with c 1 might seem favorable because of the absence of
resonance. However, this is actually difcult to attain, because the practical design
of rotor system trends to yield longer shaft, leading a situation equivalent to an
actual value of c that approaches unity (i.e. Q = innity). Particular attention must
therefore be paid to the gyroscopic effect when designing short-shafting rotors.

6.4.3 Calculation of Critical Speed of a Multi-dof Rotor


System

If the gyroscopic effect is neglected, the natural frequency xn of an undamped rotor


system at rest is given by the eigenvalue solution:

x2n M/ K Kb / 6:40

which is the same as for a structural system. Since the mass matrix M and stiffness
matrix (K + Kb) are symmetric matrices and usually positive denite, the
168 6 Gyroscopic Effect on Rotor Vibrations

eigenvalue x2n is a positive real number. Its square roots, xn , are the forward and
backward natural frequencies, respectively.
The rotational speed X that coincides with the forward natural frequency xn is
the critical speed Xc . Substituting a solution of whirling motion at the critical speed:

z / ejXc t 6:41

into Eq. (6.26) gives the eigenvalue problem for the critical speed analysis:

X2c M  G/ K Kb / 6:42

Since the matrix (M-G) is symmetric, but not necessarily positive denite, and
(K + Kb) is positive denite and symmetric, the eigenvalues X2c are real but not
necessarily positive. If positive, a critical speed Xc of the system exists; if negative,
the system has no critical speed because the gyroscopic factor c [ 1 of the corre-
sponding eigenmode.
Formally, Eq. (6.42) includes a mass smaller than that in Eq. (6.40). Therefore, a
critical speed Xc is somewhat greater than the corresponding zero speed natural
frequency xn .

6.5 Vibration and Resonance with Base Excitation

6.5.1 Resonance Conditions

This section examines the resonance conditions of a rotor on a vibrating base (as
with an earthquake, for example). As shown previously, the rotor system resonates
when the frequency of the external force coincides with the natural frequency of the
system including the direction of whirling.
Figure 6.16 shows the natural frequency curve for the rotor system, in which the
solid line represents the forward whirl natural frequency and the broken line the
backward whirl natural frequency.
The forward excitation and backward excitation are also considered separately.
As shown in the gure, the external force mat due to base excitation is the
absolute acceleration at of the base multiplied by the mass m, and the response is
the relative motion z of the vibration system as seen from the ground. Assuming
that the base excitation is unidirectional (e.g., the X direction) and the acceleration
of the base is represented by a harmonic function, and the direction of whirling is
recognized by its conversion to the exponential function:
a0 jmt
az a0 cos mt e ejmt 6:43
2
6.5 Vibration and Resonance with Base Excitation 169

resonance : twice
= f ( forward F)
resonance condition Q-value 1/2
= b ( backward B )
Y

z X

excitation frequency
z f = A f e j t
+
forward resonance
f =f
F
natural frequency

= 0 cos t =b
n
backward resonance
e j t B
= 0
2
F b z b = A b e j t
Y
e j t
+ 0 speed amplitude a
2
B X

Fig. 6.16 Vibration and resonance caused by base excitation

The unidirectional excitation is thus represented as the sum of a forward force


and a backward force whose excitation amplitude is reduced by 50 %.
Suppose now that at a certain speed the base is vibrated at a harmonic frequency
m which varies continuously as a sweep excitation. Figure 6.16 also shows the
excitation by plotting the vertical lines at a certain rotational speed. Corresponding
to the fact that the unidirectional excitation includes forward and backward lines
simultaneously, the lines consist of both solid and broken lines. As indicated by the
resonance conditions F and B in the gure, the excitation frequency lines agree with
those of the natural frequency curves including the whirl direction, as the solid line
and curve or the broken line and curve intersect in the frequency plot. Hence the
resonances appear at B and F in the vibration amplitude curve shown in the gure.
When the excitation frequency m is increased, the backward resonance appears
rst to correspond to the intersection B of the broken line and curve: the response
zb A b ejmt (the bar denote complex conjugation) becomes large at m xb .
The rotor traces a large orbit of backward whirling motion. On further increase in m,
response of the forward resonance zf Af ejmt becomes large at m xf , where
the rotor traces a large orbit of forward whirling motion. The fact that both the
forward and backward resonances appear in a vibration response for base excitation
means that a single unidirectional sinusoidal excitation merits the measurement
identication of the forward and backward whirl natural frequencies in one sweep
of the frequency.
170 6 Gyroscopic Effect on Rotor Vibrations

6.5.2 Forced Vibrational Solution for Base Excitation

The general equation of motion for rotor vibration with base excitation including
the gyroscopic effect is

z  j Xc_z 2fxn z_ x2n z


z0 t
a0 6:44
a0 cos mt  ejmt ejmt
2

where z is rotor vibration as seen from the base (i.e., relative displacement) and z0 is
the absolute acceleration of the base. The corresponding single-dof model of the
rotor is shown in Fig. 6.17. The vibration response also is written as the sum of the
forward and backward components:

 b ejmt
z Af ejmt A 6:45

By substituting Eq. (6.45) into Eq. (6.44) and equating the coefcients of the
exponential functions on both sides, the complex amplitudes Af and Ab are obtained as

a0 1
Af
2 x2n  m2 Xc m 2jfxn m
6:46
a0 1
Ab
2 x2n  m2  Xc m 2jfxn m

The forward resonance appears when the real part of the denominator of Af
approaches zero:

zr
Af ejmt m xf [ 0 6:47

at which the rotor shows a strong forward whirling motion. Similarly, the backward
resonance appears when the real part of the denominator of Ab approaches zero:

 b ejmt m xb \0
zr
A 6:48

Fig. 6.17 Base excitation Z


k
model
m

c

c k
..
Z0
6.5 Vibration and Resonance with Base Excitation 171

at which the rotor shows a strong backward whirling motion. Thus a rotor system
with the gyroscopic effect under base excitation shows a forward resonance and a
backward resonance at different excitation frequencies.
Equation (6.45), representing the whirling using complex displacements, can be
rewritten for vibration in the X and Y directions. Note that

f A
z z Af Ab jmt A b
x e ejmt
2 2 2
f  Ab 6:49
z  z Af  Ab jmt A
y e  ejmt
2j 2j 2j

Comparing Eq. (6.49) with the forms of complex amplitude for each direction
yields

Ax jmt x
A
x ReAx ejmt e ejmt
2 2 6:50
Ay y
A
y ReAy ejmt ejmt ejmt
2 2

Transformation formulae for the complex amplitudes in the X and Y directions


are obtained as

Ax Af Ab Ay jAf  Ab 6:51

The absolute values |Ax| and |Ay| of the complex amplitudes give the resonance
curves under base excitation shown in Fig. 6.18.
For X 0 (the rotor at rest), the gyroscopic factor c 0 and Af Ab A;
therefore

 b ejmt Re2Af ejmt  ReAejmt


z Af ejmt A 6:52

Fig. 6.18 Resonance curves 5


re s o n a n c e a m p litu d e (ra tio )

by base excitation = 0.1


4
= 0
3 vibration of derection X
vibration of derection Y
2
= 2/3
1

0 1 2 3
excitation frequency / n
172 6 Gyroscopic Effect on Rotor Vibrations

The rotor vibration is then unidirectional, and its complex amplitude A is in the
familiar form:
a0
A 6:53
x2n  m2 2jfxn m

6.5.3 Resonance Curves and Whirling Trajectories

As seen in Fig. 6.18, the rotor at rest shows a single resonance peak at m xn ,
corresponding to Eq. (6.53), in the X direction (direction of excitation) only, as
evident in structural dynamics.
When exciting the rotor at a high rotational speed where the gyroscopic effect is
signicant, two resonance peaks [52] are observed in accordance with Eq. (6.46).
As the exciting frequency is increased, the backward resonance appears rst, and
after a range of smaller amplitude, the forward resonance occurs. The resonance
amplitude is comparable in the X and Y directions, although excitation acts in the X
direction only.
Figure 6.19 shows the whirling trajectories in response to increasing the base
excitation frequency during rotation. It is observed that:
(a) At a low excitation frequency m less translational vibration in the X (excita-
tion) direction appears. As m increases, vibration in the Y direction ensues,
resulting in an elliptical trajectory with the long axis close to the X axis,
corresponding to backward whirling.
(b) As m increases further, the elliptical trajectory grows gradually and approaches
to a circle (i.e. the amplitudes in the X and Y directions become closer),
arriving at the backward resonance where the amplitudes are at a maximum.
(c) After the backward resonance, the trajectory is reduced down to a short
straight line close to the Y axis, where the whirling direction changes to a

Y
(a) (b) (c) (d) (e)
3 Y
Y 2 Y 2 Y
1 2 1 1
1
1 1 3 1 1 2 1 1 2
X 3 2 1 1 2 X X X 1 1 X
1 1 2 1 1

2 2

= 0.25 3 = 0.75 =1 = 1.72 = 2.5
n n n n n
backward forward
resonance resonance

Fig. 6.19 Whirling trajectory


6.5 Vibration and Resonance with Base Excitation 173

Fig. 6.20 Gyroscopic 5

resonance amplitude (ratio)


influence due to base =0 = 0.1
excitation 4
0.5
3
1
2
1.5
2
2.5
1

0 1 2 3
Excitation frequency / n

forward sense. The transition from the backward to the forward whirling
generally occurs through a straight-line trajectory.
(d) The forward whirling trajectory then grows to become nearly circular until the
forward resonance appears.
(e) The trajectory is reduced at higher excitation frequencies.
Thus, the trajectory is close to a circle at the forward or backward resonances,
which means that a peak appears in both the X and Y directions. The peaks are
smaller than the peak in the response at rest, because the acceleration of excitation
is split into the forward and backward accelerations of whirling excitation whose
amplitudes are half that of a. Similarly, the peak value of the vibration response
would, in principle, be half the peak value at rest.
Figure 6.20 shows the resonance curves for base excitation with the gyroscopic
factor as the parameter. The peak of backward resonance is higher than that of
forward resonance. Because the gyroscopic effect for the backward eigenmode
appears to increase the mass, it decreases the natural frequency, lowers the modal
damping ratio (greater Q-value) and nally increases the resonance peak. On the
other hand, for the forward eigenmode it decreases the mass, increases the natural
frequency, increases the modal damping ratio, reduces the Q-value and nally
makes the forward resonance peak lower than the backward peak.
As predicted, the average of the forward and backward peaks is approximately half
of the resonance peak amplitude of the rotor at rest. When a rotor is impulse-excited
during rotation, it responds in the most sensitive backward whirling mode.

6.5.4 Case Study: Aseismic Evaluation of a High-Speed


Rotor [52]

Aseismic evaluation test data of a long-shaft, high-speed rotor are presented in this
section. The rotor is represented schematically in Fig. 6.21. It is supported by a
rigid and a flexible bearing at the left and right end, respectively. The rst critical
speed in about 5 Hz, at which a large amplitude is exhibited at the right end.
174 6 Gyroscopic Effect on Rotor Vibrations

stopper gap
gap sensor
ki
1 mode inside
stopper
1 = 5.5 Hz
x outside k 0 ki
stopper
rigid soft
bearing earthquake bearing

(t )

Fig. 6.21 View of a high-speed long axis rotor

(a) (b)


deflection x, amplification ratio [ m / gal ]

deflection x, amplification ratio [ m / gal ]

150 150

100 100

50 50 right side
x sensor
right side right side
x sensor y sensor
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
excitation frequency [Hz] excitation frequency [Hz]
at rest ( = 0 rps ) at high-speed rotation ( = 763 rps )

Fig. 6.22 Rotor vibration caused by base harmonic excitation

The resonance curves for the rotor by the harmonic base excitation in a unidi-
rectional way are shown in Fig. 6.22a at rest and 6.22b at a high rotational speed.
While (a) shows a single resonance peak, at which the vibration direction is the
same as the excitation as seen from the Lissajous orbit, (b) shows forward and
backward resonance peaks, where the response trajectory is whirling in the X-Y
plane. The backward resonance sensitivity is higher than the forward resonance
sensitivity, which indicates a strong gyroscopic effect. Also, the fact that the
backward resonance peak in gure (b) is higher than the peak at rest in gure
(a) suggests that the bearing characteristics at high rotational speeds are consider-
ably different from those at low speeds.
6.5 Vibration and Resonance with Base Excitation 175

(a) Y

4
0

0.
earthquake wave X
0.3 (b)
1.5
1 0

0
1.
nonlinear limit due to stopper

maximum gap ( )
1
X 1.0 3
2 2
0
experiment
calculation
0
1.

1 0.5
= gap 1.0 mm
3 0 1 linear = acceleration 0.043 G
X
vibration 0 1 2 3 4 5
Lissajous orbit maximum acceleration rate ( )
Simulations Comparison of test result

Fig. 6.23 Effective stopper for prevention of excessive response due to an earthquake.
a Simulations. b Comparison of test result

A spring stopper with a small clearance gap, d, is provided at the right end to
prevent excessive vibration in an earthquake. The evaluation results of the stopper
to suppress severe vibrations are shown in Fig. 6.23, in which (a) shows simula-
tions and (b) the results of an experiment. The simulation corresponds to a weak
earthquake, which does not make the rotor collide with the stopper. The simulation
represents the rotor colliding with the stopper due to a stronger quake. Its
vibration is, however, limited by the stopper. The simulation shows a still
stronger quake, but the Lissajous orbit does not diverge, indicating that the rotor
vibration is suppressed effectively by the stopper.
Figure (b) compares the results of calculation (solid line) and experiment (cir-
cles). The maximum response displacement is limited to be approximately less than
1.1 times the specied gap.
Note: As stated in ISO 14839-4 [53] and API 617 [11] with relation to mag-
netically levitated rotor by AMB (see Table 1.1), the drop test is typically required
to verify the stable operation even against this kind of rubbing situation.
Concerning rotor-to-stationary rubbing dynamics, there are so many papers, some
of which are summarized in chapter ve of Muszynskas book [B33] from her
viewpoint. Another viewpoint is given by Keogh [54] and related information
[5558] are interesting.
176 6 Gyroscopic Effect on Rotor Vibrations

6.6 Ball Passing Vibration and Resonance Due to Ball


Bearing Defects

6.6.1 Ball Bearing Specications

Ball bearings are manufactured with high precision, and do not generate strong
vibration that may destroy a rotor shaft. However, vibration of ball bearings may
generate noise or cause wear. In devices of very high precision, a small vibration
may mean a serious disturbance in their function, such as reading/writing problems
in hard disk drives.
Defects of bearing balls (e.g. uneven size, deformation or flaws in surface)
engender excitation forces when spinning and revolving (i.e., whirling). An
example of ball bearing specications is shown in Fig. 6.24. The revolution speed
XR of a ball, i.e., revolving speed of retainer, in this bearing is:
   
X D 1 D
XR 1  cos a  b X ! b 1  cos a
0:4 6:54
2 d 2 d

where D is ball diameter, d the diameter of pitch circle, a contact angle, and X the
rotational speed of the shaft (or inner race). The ratio b of the revolution speed to
the rotational speed of the inner race is b 0:398 in this case.
In common ball bearings b
0:4; that is, the balls revolve forwards at a speed of
about 40 % of the shaft rotational speed as seen from a stationary system, and
backwards at a speed of about 60 % as seen from a rotational coordinate system.
A model is considered here in which the rolling surface has a recess or protuberance
over which the revolving balls pass to generate a unidirectional excitation force.
This kind of forced vibration is called the ball passing vibration.

6.6.2 Excitation by a Recess on Outer Race

Consider a bearing with a defect (recess or protuberance) in the outer race (Fig. 6.25,
upper left). The exciting frequency caused by the balls passing over a defect is the

Fig. 6.24 Specication of Outer Race


ball bearing Ball
D Inner Race
T D = ball diameter = 11.112 mm
T Z = number of balls = 9
d = contact angle = 12
d = 53.5 mm
NSK 6207
6.6 Ball Passing Vibration and Resonance 177

resonance condition
cos Z t
outer protuberance : Z = f
Z = b
inner protuberance : ( Z + 1 Z ) = f
(Z 1 Z ) = b
F
outer race deformation
er

)

Z out ce

+ 1 race

Z
ra

Z
er


F f forward

1
natural frequency

inn
Z

Z
(
F

(Z
Z F
n B
B resonance
( Z + 1 Z ) b
( Z 1 Z )
A speed backward
F
cos (1 ) Z t B
amplitude

Fr F B
B
resonance


speed
inner race deformation

Fig. 6.25 Ball passing vibration

angular velocity of revolution XR multiplied by the number of the balls Z, i.e., Zb X.


Every passage of a ball over the recess generates the unidirectional excitation:

F 2F0 cos Zb Xt 6:55

which is converted to the expression of whirling force:

Fz F0 ejZbXt ejZbXt 6:56

A defect in the outer race thus produces forward and backward whirling exci-
tation of the same frequency, which is 3:582X for the bearing shown in Fig. 6.24.
More than one recess may exist; for symmetrically arranged m recesses, hence the
expected excitation frequencies affecting a rotor are

mZb X 6:57

6.6.3 Excitation by a Recess on Inner Race

For a bearing model having a recess on the inner race (Fig. 6.25, lower left), the
excitation frequency is given by the difference of shaft rotational speed and ball
178 6 Gyroscopic Effect on Rotor Vibrations

revolution speed X  XR multiplied by the number of balls Z, or Z1  bX, which


gives rise to the unidirectional excitation seen from a rotational coordinate system:

Fr 2F0 cos Z1  bX t 6:58

which should be transferred to the whirling excitation Fz seen from the inertial
coordinate system:
n o
Fz Fr ejX t F0 ejZ 1ZbX t ejZ1ZbX t 6:59

A defect in the inner race thus produces forward and backward whirling exci-
tations of different frequencies, which are 6:48X and 4:412X, respectively, for
the bearing shown in Fig. 6.24. For symmetrically arranged m recesses, expected
excitation frequencies are

mZ 1  mZbX; mZ  1  mZbX 6:60

6.6.4 Resonance Conditions

The resonance conditions for a single recess on the outer race are

Zb X xf ; Zb X xb 6:61

and for the inner race,

Z 1  ZbX xf ; Z  1  ZbX xb 6:62

These conditions are represented graphically in the middle of Fig. 6.25. The
excitation frequency by the outer race given by Eq. (6.57) is shown by superim-
posing two solid and broken lines on a straight line. The inner race produces two
straight lines of the forward (solid line) and backward (broken line) frequencies
given by Eq. (6.60). The intersections of the lines correspond to the resonance
conditions, which yield the resonance curve shown in the lower middle of the
gure. In this example, as the rotational speed increases, the forward and backward
resonances due to the inner race, and the backward and forward resonances due to
the outer race appear in this order in the same mode.

6.6.5 Case Study: Hard Disk Drive (HDD) [59, VB218]

The base of an old model of hard disk drive (Fig. 6.26) was excited horizontally
and the vibration of the shaft top was measured with a static capacitance dis-
placement sensor. The results of vibration analysis are shown in Fig. 6.27.
6.6 Ball Passing Vibration and Resonance 179

head

base
disk

main shaft

disk = 14"( 357 ) 1.9 t 13 = 3 600 rpm , 600 MB

Fig. 6.26 Old type of HDD (rotary actuator head)

Fig. 6.27 Campbell diagram 600


: measured
of HDD
5f
500
5b

4
vibration frequency [Hz]

400
3 inner race
(c) 384 Hz

300
4
.42

.
+6

(a) 199 Hz outer race


2
200 1 f

.5
1 b 3

100
(b) 146 Hz

0
0 10 20 30 40 50 60 70
1.0
amplitude [m]

31.3 32.8 60.5

0.5
upside of
main shaft

0
0 10 20 30 40 50 60 70
speed [rps]
180 6 Gyroscopic Effect on Rotor Vibrations

The measured vibration magnitude (ordinate) of the rotor shaft at varied rotational
speeds (abscissa) is shown at the lower part of Fig. 6.27. Three conspicuous resonance
peaks are seen at about 30 rps and at 60 rps (rated speed). The Campbell diagram for
these peaks, in which the resonance amplitude is represented by the diameter of a
circle, is shown in the middle of Fig. 6.27. The resonance frequencies are:
(a) 199 Hz at a rotational speed of 31.3 rps
(b) 146 Hz at a rotational speed of 32.8 rps
(c) 384 Hz at a rotational speed of 60 rps
Analysis of these resonance conditions is shown in Campbell diagram at the
upper part of Fig. 6.27. The base was excited with a sinusoidal wave while the rotor
was at rest and rotating at different speeds, and resonances were observed under
each condition. All of resonance frequencies observed during the excitation were
plotted by , and in the gure. The plotted data yield natural frequencies x1b ,
x1f , x2 , x3 , x4 , x5b and x5f (order of increasing frequency) as suggested by
curves. The rst and fth natural frequencies split into forward and backward
frequencies, indicating that they are the eigenmodes of the rotating parts with
gyroscopic effect. In particular, in the rst mode x1f  x1b 2X, which indicates
the eigenmode of thin disk, where the number of the nodal diameter is 1, with a
gyroscopic factor c 2, as seen in Eq. (6.25). On the other hand, the second, third
and fourth natural frequencies, which do not split (the forward and backward
frequencies are the same), are for the eigenmodes of the stationary structure system
(e.g. the base plate and the bearing housing).
The ball bearing used in this hard disk drive as shown in Fig. 6.24 indicated
excitation frequencies and induced possible resonance conditions as follows:
Excitation by the outer race: 3:5X, no resonance
Forward excitation by the inner race: 6:42X, resonance (a) and (c)
Backward excitation by the inner race: 4:4X, resonance (b)
Considering the rated rotation speed of 60 rps, the drive needed prevention of
the resonance (c). This was achieved by increasing the stiffness of the stationary
side, i.e., strengthening the base plate and the cylindrical bearing housing, to
increase the natural frequency x3 slightly as to move the resonance (c) upward.
Since the natural frequency (c) does not split under the gyroscopic effect, rotor
modication may not be required for this troubleshooting problem.
Chapter 7
Approximate Evaluation for Eigenvalues
of Rotor-Bearing Systems

Abstract This chapter discusses an approximate evaluation method to consider the


effects of the dynamic characteristics of a bearing on the complex eigenvalues
(damping characteristics as the real part and damped natural frequency as the
imaginary part). The method consists basically of two steps:
(1) System reduction down to a single-dof system is executed based on the
orthogonality condition of modes in the conservative system, and the equation of
motion of reduced system is expressed in the complex displacement form, and
(2) Approximate analysis of the complex eigenvalues of the system is used to
ascertain the effects of the bearing parameters on the natural frequencies and
damping characteristics.
This combination provides a simple model that helps understanding the phenomena
of practical interest, such as the effects of the cross-stiffness of the bearing on the
system instability or the stabilizing effect of anisotropy in the bearing stiffness.
In addition, the shapes of resonance curves in unbalance vibration are discussed in
relation to the dynamic characteristics of the bearing.

 
Keywords Sliding bearing Anisotropic support Cross-spring effect Stability 

Jeffcott rotor Tuning of spring and damping coefcients

7.1 Equation of Motion for a Single-Degree-of-Freedom


Rotor System

For a general single-dof rigid rotor system as shown in Fig. 7.1, the effects of the
dynamic characteristics of the bearing supports on vibration characteristics (natural
frequency, critical speed, damping ratio, Q-value, etc.) are examined. Although an
actual rotor system has more than one degree of freedom, the model is practically
useful because the system can be reduced to a single-dof model.
In a sliding bearing supporting a horizontal rotor, the steady-state journal
position S is somewhat eccentric with respect to the bearing center O (Fig. 7.2)
because the rotor weight must be supported by the static levitation force of oil lm.

Springer Japan 2017 181


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_7
182 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

Fig. 7.1 1-d of model

Fig. 7.2 Dynamic vertical direction


characteristics of sliding
bearing oil-film y(V)
journal
(rotor) m

O kxx
S x(H) horizontal
cxx direction
kxy
cyx
cxy
bearing cyy ( anisotropic 8 parameters )
kyx kyy

The stiffness and damping coefcients of the bearing reaction force against a small
displacement {x, y} from the steady equilibrium position S are called the dynamic
characteristics of the bearing, which are not X-Y symmetric due to the eccentricity
of the position S, but represented by eight parameters including the stiffness
coefcients kij and viscous damping coefcients cij (i, j = x, y).
The general form of the equations of motion including the eight parameters of
the dynamic characteristics of the bearing, with the modal mass m and modal
gyroscopic effect G, are written as

mx XG_y cxx x_ cxy y_ kxx x kxy y 0


7:1
my  XG_x cyx x_ cyy y_ kyx x kyy y 0

or, using the complex displacement z x jy,

mz  jXG_z kf z cf z_ kb z cbz_ 0 7:2

where

kxx kyy kxy  kyx kxx  kyy kxy kyx


kf  kd  jkc j ; kb j
2 2 2 2
cxx cyy cxy  cyx cxx  cyy cxy cyx
cf  cd  jcc j ; cb j :
2 2 2 2
7.1 Equation of Motion for a Single-Degree-of-Freedom 183

Fig. 7.3 Dynamic y


characteristics of sliding
bearing (vertically supported)

journal
(rotor)

kd
O=S x
cd
m
oi
l-f
ilm
kc = kxy = kyx
cc = cxy = cyx
bearing cd kd ( isotropic 4 parameters )

On the other hand, the shaft center of a vertical rotor (Fig. 7.3) coincides with the
geometrical bearing center, and the reaction force against small displacement is iso-
tropic. Therefore the dynamic characteristics are reduced to the X-Y symmetric form:

kxx kyy kd ; kxy kyx kc


7:3
cxx cyy cd ; cxy cyx cc

The equation of motion therefore contains four parameters:

mz  jXG_z kd  jkc z cd  jcc _z 0 7:4

The system is referred to as an isotropically supported rotor.


In the following sections, the effects of the bearing parameters on the rotor
vibration characteristics are discussed rst for the isotropic support, and then for the
anisotropic support.

7.2 Vibration Characteristics of a Symmetrically


Supported Rotor System

For the sake of simplicity, this section uses the variable, xd , and dimensionless
parameters:
p
xd kd =m: undamped natural frequency of the rotor at rest;
c G=m: gyroscopic factor;
p X=xd : dimensionless rotational speed;
7:5
lc kc =kd : ratio of cross-stiffness coefficient to direct stiffness coefficient
p
fd cd =2= mkd : magnitude of direct damping; and
p
fc cc =2= mkd : magnitude of cross-damping:
184 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

7.2.1 Natural Frequencies of a Conservative System

A system is conservative if it is undamped (cd = cc = 0) and without cross-stiffness


(kc = 0, kxy = kyx). Then the equation of motion is, from Eq. (7.4),

mz  jXG_z kd z 0 7:6

Writing its vibrational solution as

z aejxt 7:7

the natural frequency x is obtained from the characteristic equation

mx2  XGx  kd 0 ! x2  cXx  x2d 0 7:8

That is,
q
x cX=2  xd 1 cp2 =4
s
 
x cp cp2 xf xb
)  1 , ; alternate signs in the same order
xd 2 4 xd xd
7:9

Thus positive and negative natural frequencies exist: x xf [ 0 is the natural


frequency of forward whirling generating a circular orbit, and x xb \0 that of
backward whirling generating a circular orbit.
Representing both natural frequencies as positive values, the solution can be
written as
s
 
xf xb cp2 cp cp cp2
; 1  1
xd xd 4 2 2 8
cp
1 signs in the same order 7:10
2

It is now clear that the natural frequency is split into the forward xf and the
backward xb of whirl frequencies due to the gyroscopic effect, as illustrated in
Fig. 7.4a, the former increasing, and the latter decreasing, as the rotational speed
increases. This feature is characteristic to the gyroscopic effect as evident in Fig. 6.11.
7.2 Vibration Characteristics of a Symmetrically 185

(a) (b)

f ]
Re[ f
rd
forw
ard forwa
d d Re[] / |
|
=0
back back f =
war ward b
d
Re[
b b]


natural frequency damping ratio

Fig. 7.4 Gyroscopic effect

7.2.2 Effects of Non-conservative System Parameters

The equation of motion (7.4), including the non-conservative parameters kc, cd and
cc, is rewritten using the dimensionless parameters:

z  jXc_z x2d 1  jlc z 2xd fd  jfc _z 0 7:11

With the vibrational solution

z aekt 7:12

the characteristic equation is

k2 2fd  jfc  jpc=2xd k 1  jlc x2d 0 7:13

which gives the exact solution

k cp p
fd jfc j  j D 7:14
xd 2

where D 1  jlc  fd  jfc  jpc=22  1  jlc jcpfd .


Assuming that each of the variables in D is sufciently smaller than unity, an
approximate solution is obtained, as also seen in Eq. (B.8) of Appendix B:

k  cp lc  cp
 fd 1   j 1 jfc 7:15
xd 2 2 2

More specically, the complex eigenvalue of the forward whirling is


186 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

(a) kd (b) cd
( at k c > 0 )

damping ratio = Re [ ] / | |
cc
d
forwar
natural frequency f f ]/ d kc
rd Re[
forwa
d d kc cd
kd
backwar
d b Re[ b]/
d backward
cc


natural frequency damping ratio

Fig. 7.5 Influence of bearing dynamic properties

kf cp l cp
 fd 1 c j1 fc 7:16
xd 2 2 2

The complex eigenvalue of the backward whirling is

kb cp l cp
 fd 1   c  j1  fc 7:17
xd 2 2 2

Equations (7.16) and (7.17) are useful to examine the effects of the
non-conservative parameters.
(A) Direct damping: cd cxx cyy ! fd
The effect of direct damping cd, appearing in the real part of the complex
eigenvalue, increases the stability in both forward and backward whirling solutions,
as illustrated in Fig. 7.5b.
(B) Cross-stiffness: kc kxy kyx ! lc

The effect of cross-stiffness kc, appearing in the real part of the complex
eigenvalue, decreases the stability in forward whirling and, while increasing the
stability in backward whirling, as illustrated in Fig. 7.5b.
The dynamic characteristics of a cylindrical sliding bearing are considerably
anisotropic at lower rotational speeds, but otherwise approximately isotropic
(kxy  kyx [ 0). In this situation increase of the cross-stiffness constant kc > 0 due
to oil lm reduces the stability for forward whirling. This may destabilize the
system, leading to self-excited vibration called oil whirl/whip. In contrast, the
stability for backward whirling increases. Note that a tilting pad bearing is highly
stable because the cross-stiffness coefcient kc is nominally zero.
(C) Cross-damping: cc cxy cyx ! fc
The effect of cross-damping cc appears in the imaginary part of the complex
eigenvalue, as shown in Fig. 7.5a, increasing the forward natural frequency and
7.2 Vibration Characteristics of a Symmetrically 187

decreasing the backward natural frequency. In actual slider bearings, however, this
effect is very small and practically negligible.
(D) Influence of the gyroscopic effect on the real part of the complex eigenvalue
The gyroscopic effect is addressed again here to examine its influence on the real
part of the complex eigenvalue.
As the rotational speed increases, the absolute value of the real part Rekf  of the
forward complex eigenvalue kf increases, but the real part Rekb  of the backward
eigenvalue kb decreases, as shown in Fig. 7.4b. The gyroscopic effect splits the real
parts of the both eigenvalues, as well as the natural frequency curves of Fig. 7.4a.
It might seem strange that the conservative gyroscopic effect influences the real
part of the complex eigenvalue. This is caused by apparent decrease in the mass for
the forward whirling due to the gyroscopic effect associated with rotation, thus
increasing the forward whirl natural frequency and the absolute value of the real
part of the complex eigenvalue. This is analogous to the case of a light mass. The
effect is opposite for the backward whirling case.
In terms of the damping ratio, Eqs. (7.16) and (7.17) yield
Damping ratio for forward whirling:

Rekf f 1 cpxd
ff      d fd 7:18
kf 1 cpxd

Damping ratio for backward whirling:

Rekb fd 1  cpxd
fb    fd 7:19
j kb j 1  cpxd

They are approximately equal, as shown in Fig. 7.4b.

7.2.3 Parameter Survey

Here the precision of the approximation is checked using calculation examples.


Calculations are made for the four cases listed in Table 7.1:
a system including mass and stiffness elements, but not damped;
addition of the damping elements;
addition of the gyroscopic effect; and
addition of the cross-stiffness effect.

It also shows the exact complex eigenvalues k, damping ratios f, and the
approximate complex eigenvalues ka obtained by Eqs. (7.16) and (7.17). It is clear
that the approximation is satisfactory; Eqs. (7.16) and (7.17) are therefore useful
tools for the evaluation of vibration characteristics.
188 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

Table 7.1 Parameters (xd 1; f Rek=k)

No. p d c comment complex eigenvalue damping


approximation
ratio a
conservative
1 0 0 0 system = j1 0 a = j1

2 0 0.2 0 added damping = 0.2 j 0.98 0.2 a = 0.2 j 1


gyroscopic = 0.29 + j 1.6 a = 0.3 + j 1.5
3 1 0.2 0 0.178
effect = 0.11 j 0.6 a = 0.1 j 0.5

instability by = 0.15 + j1.65 0.09 a = 0.2 + j 1.5


4 1 0.2 1
cross-stiffness = 0.55 j 0.65 0.64 a = 0.6 j 0.5

The complex eigenvalues ka for the four cases are plotted on a complex plane
(Fig. 7.6) to visualize the results in Table 7.1.
The characteristic roots for the system including mass and stiffness elements,
but not damped, are on the imaginary axis (k 1). The positive root cor-
responds to the forward whirling and the negative root to the backward
whirling.
For the system including mass, stiffness and damping elements, the real part of
the forward and backward whirl eigenvalues are both shifted to the left by
about 0.2 by the damping effect cd, indicating the increase of stability.
The gyroscopic effect moves the points to , i.e., the forward whirl
eigenvalue to the upper left and the backward eigenvalue to the upper right.
Then, the forward whirl natural frequency increases and the backward natural
frequency decreases. The absolute value of the real part of the forward
eigenvalue increases, and that of the backward whirl eigenvalue decreases.

Im
0.1
0.2
.3 2
=0
3 4
1.5
undamped 2
1 1
conservative 1
system
cd 0.5
damping 2 Re
1 0.8 0.6 0.4 0.2 0.2
gyroscopic G 3
effect 3 0.5
4
kc > 0 2 1
cross-stiffness 4 1
=0
.3 0
.2 0.1

Fig. 7.6 Example of calculation about characteristic root


7.2 Vibration Characteristics of a Symmetrically 189

This might suggest different levels of stability between them, but the corre-
sponding damping factors are both f 0:178. As mentioned in (D) in p.187,
the gyroscopic effect yields approximately the same damping ratios.
The effect of cross-stiffness shifts the points to . The forward whirling is
destabilized in this example in which the real part is positive, while the sta-
bility of the backward whirling increases.

7.3 Natural Frequencies of a Rotor Supported


by Anisotropic Bearings

For simplicity, the following variable and dimensionless parameters are included:

kb  jkb jejhb : Polar representation of the anisotropic component of stiffness


lb jkb j=kd : Ratio of the anisotropic component to isotropic component
Dx XGx=jkb j cpx=xb =lb
kf  kd  jkc kd 1  jlc
cf  cd  jcc 2fd  jfc xd m
cb : The anisotropic component of damping usually small and negligible
7:20

7.3.1 Natural Frequency of a Conservative System

A system is conservative if undamped (cf = cb = 0) and the stiffness coefcients


conform to the action-reaction law (kxy = kyx, or kf = kd, kc = 0). For this case, the
equation of motion (7.2) can be written as
mz  jXG_z kd z kb z 0 7:21

For the vibrational solution


 ejxt
z /f ejxt / 7:22
b

the characteristic formula is


 
mx2 XGx kd kb /f
0 7:23
kb mx2  XGx kd /b

and the natural frequency x is determined by the characteristic equation


190 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

2
( G )
( k d m 2 ) 2

( k d m 2 ) 2 | k b | 2

kd | kb |
{H , V } =
2 2
m
2
= d 2 ( 1 b )

H2 d 2 V2 22
12 kd
d 2 = m

Fig. 7.7 Natural frequency of anisotropic bearing system


2
kd  mx2 jkb j2 XGx2 7:24

The solutions of this equation are shown graphically in Fig. 7.7. The horizontal
natural frequency xH : and vertical natural frequency xV are associated with the
non-rotating state, and usually xH \xV . The undamped natural frequencies under an
average of the stiffness coefcients of the vertical and horizontal directions is xd and
kd  jkb j
x2H ; x2V x2d 1  lb signs in the same order 7:25
m

The gyroscopic effect appears in the rotating state, which shifts the natural
frequencies to x1 and x2 : The horizontal frequency xH decreases to x1 and the
vertical frequency xV increases to x2 . Their values are given by
q
2 2
2 2 kd  jkb j XGxi
x1 ; x2
 p
m
x2d 1  lb 1 D2 alternate signs in the same order for i 1; 2
7:26

7.3.2 Elliptical Whirling of a Conservative System

Equation (7.23) determines the eigenvector:


q p
/b mx2  kd  XGx  jkb j2 XGx2  XGx  1 D2  D

/f jkb jejhb jkb jejhb ejhb
7:27

Therefore
7.3 Natural Frequencies of a Rotor Supported 191

straight line orbit

1.0

1 + 2 +
1 0.8
Mode ratio

0.6
1 + 2 =

0.4

0.2
h=

0
0 0.5 1 1.5 2 2.5 3
= ( G / | kb | ) = p ( / d ) / b

Fig. 7.8 versus h for dening elliptical orbit

 
/b  p2  1 h2 ejhb
 1 D  D ejhb ;
/f h1 e b
jh 1 7:28
alternate signs in the same order
q
where hx 1 D2 x  Dx 0\h\1, h1 hx1 , h2 hx2 .
The natural vibration in this case represents an elliptical orbit because the for-
ward and backward whirling are superimposed. Figure 7.8 shows the dependence
of h on D.
(A) For the lower horizontal natural frequency x = x1, the mode ratio is

/f =/b h1 ejhb 7:29

Therefore
  
/f h1 ejhb 1
or 7:30
/b 1 1=h1 ejhb
 
Since the mode ratio magnitude h1 is smaller than unity, i.e., /f  j/b j, the
elliptical whirling orbit is backward.
(B) For the higher vertical natural frequency x x2 , the mode ratio is

/b =/f h2 ejhb 7:31

Therefore
192 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

 
/f 1
7:32
/b h2 ejhb
 
Since /f 
j/b j in this case, the elliptical whirling orbit is forward.
The mode ratio magnitude h thus determines the form of the ellipse, which has a
major axis of 21 h, and a minor axis of 21  h, although the tilt is unknown.
The gure shows the ellipse approaches a circle as D increases. Two ellipses with
long vertical and horizontal axes, respectively, and typically as shown in the gure.

7.3.3 Influence of Gyroscopic Effect

The lower and higher natural frequencies x1 and x2 for anisotropic bearing stiff-
ness are obtained from Eq. (7.24). The solutions are
v
s
  u
u
x1 x2 t cp2 cp4
; 1  cp2 l2b signs in the same order
xd xd 2 4
7:33

Calculation examples are shown in Fig. 7.9. The natural frequency curves splits
upward and downward from x1;2 =xd 1 as the gyroscopic effect cp increases. The
anisotropic support separates two upward/downward lines to a high value

1.8

1.6
6

4
0.

0.

1.4
d

0.
, 2/

1.2

1.0
b = 0 ( isotropic )
d
1 /

0.8

0.6
2
0.

0.4
4

0.
0.

0.2
0 0.2 0.4 0.6 0.8 1
G
p=
m d

Fig. 7.9 Gyroscopic effect (anisotropic/isotropic supports, conservative system)


7.3 Natural Frequencies of a Rotor Supported 193

x1 =xd [ 1 and a low value x2 =xd \1. At low rotational speeds, we recognize the
anisotropic support effect, where the influence of the gyroscopic effect is small. At
high rotational speeds, the splitting curves converge to the lines for symmetric
support. Similarly, D increases in Fig. 7.8 and the mode ratio magnitude h ap-
proaches zero, making the ellipse approach to a circle.

7.3.4 Shape of Elliptical Whirling Orbit

When a vibrational solution is given in a complex form

 b ejxt
z Af ejxt A 7:34

where Af af ejuf , Ab ab ejub .


It can be rewritten as
n o
z ejuf ub =2 af ejxt h ab ejxt h h uf ub =2 7:35
 
The direction of the major axis is uf  ub =2, the major axis is 2af ab  and
 
the minor axis is 2af  ab  in length.
The length and direction of the major and minor axes of the elliptical orbit are
thus determined. The orbit is shown in Fig. 7.10.
Table 7.2 summarizes calculation examples of the flatness of the shape of the
elliptical orbit under the condition kb =kd  lb \hb 0:2\  160 (the horizontal

Fig. 7.10 Direction of major Y


axis/minor axis of elliptical
whirling locus of rotor
supported by anisotropic
minor axis f b forward
bearing
2 ( | af | > | ab | )
X


2 backward
|a
f
a ( | af | < | a b | )
b|

major axis
b|
a
+
f
|a
2
194 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

Table 7.2 Value of whirling locus (Fig. 7.11)

kb / kd b b = 0.2 160

No. item 1 vibration 2 vibration

(1) p 0 0.1 0.2 0.4 0 0.1 0.2 0.4

(2) /d 0.894 0.886 0.858 0.791 1.095 1.108 1.14 1.24

(3) 0 0.44 0.86 1.58 0 0.55 1.14 2.48

| f | 1 0.65 0.46 0.29 1 1 1 1


(4)
| b | 1 1 1 1 1 0.59 0.38 0.19

short axis
(5) straight 0.21 0.37 0.55 straight 0.26 0.45 0.67
long axis

direction of 160 + 180 0 0 160


(6) = 10 = 80
long axis 2 2

forward
(7) whirling

backward

stiffness kxx is lower than the vertical stiffness kyy). Each row of the table is
described in the following statements:
(1) cp gyroscopic effect c dimensionless rotational speed p
(2) Dimensionless natural frequency x=xd calculated from Eq. (7.33)
(3) Calculation of D XGx=jkb j cpx=xd =lb
(4) The mode ratio magnitude h is obtained from D in Fig. 7.8 to determine the
forward and backward whirl eigenmodes.
(5) The ratio 1  h=1 h of the major axis to the minor axis of the elliptical
orbit is calculated.
(6) The direction of the major axis calculated by uf  ub =2 in Fig. 7.10.

For the backward natural frequency x1 , uf hb 180 160 180


20 and ub 0 are substituted into Eq. (7.30); for the forward natural frequency
x2 , uf 0 and ub hb 160 are substituted into Eq. (7.32).
(7) Drawing the general shape of the elliptical whirling orbits.
Figure 7.11 illustrates changes in the whirling orbits by increasing the rotational
speed. The backward and forward whirling orbits are approximately horizontal and
vertical, respectively. The orbit is a straight line when the system is not rotating
(cp 0); the elliptical orbit approaches a circular whirling as the rotational speed is
increased.
7.3 Natural Frequencies of a Rotor Supported 195

(b) Y

Y
(a)

2 1

2 10
X
4 2 4 X 1 1
0
0.1 80
2 0.2 0.4
p 0.2 1
=0 0.1
.4 0
1 mode p =

2
2 mode

Fig. 7.11 Whirling locus (conservative system kb =kd  lb \hb 0:2\  160 )

7.3.5 Effects of Non-conservative Parameters [60]

This section discusses the effects of non-conservative parameters kc, cf and cb on the
complex eigenvalues.
For simplicity, using the transformation

z zf zb ! zf /f est ; zb /b est 7:36

When we rewrite Eq. (7.2) by using the forward and backward displacements, zf
and zb, are separated:
     
m 0 zf G 0 zf kd kb z f
 jX
0 m zb 0 G z_ b kb kd z b
    7:37
cf cb z_ f jkc 0 zf
0
cb cf z_ b 0 jkc zb

The upper portion of the equation gives the undamped solution, which corre-
sponds to elliptical whirling in a constant orbit of a conservative system, while the
lower portion is associated with small parameters in a non-conservative system,
which causes the elliptical whirling orbit to attenuate or diverge.
The upper portion of Eq. (7.37) can be rewritten as an eigenvalue problem as
shown in Eq. (A.13) of Appendix A. Since B is a positive-denite Hermitian matrix
and A is a skew-Hermitian matrix, the eigenvalue is a purely imaginary number
k jx, which corresponds to the natural frequency x1 of Eq. (7.30) or x2 of
Eq. (7.32).
196 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

When assuming that k jx is either x1 or x2 , approximate modal equations of


motion, referring to Eq. (A.16) of Appendix A, can be obtained by the modal
transformation
 
zf t /f
gt 7:38
zb t /b

To do so, Eq. (7.38) is substituted into Eq. (7.37) and multiplied from the left by
the conjugate of the eigenvector:
         h i
m /f  j/b j2 g  jXG /f  j/b j2 g_ kd /f  j/b j2 g 2Re /
2 2 2  t kb / g
f b
        
cd /f  j/b j g_  jcc /f  j/b j g_  jkc /f  j/b j g 0
2 2 2 2 2 2

7:39

This equation can be evaluated for individual modes as shown previously.


 
/ h1 ejhb
(A) The lower horizontal natural frequency, x x1 , has mode f :
/b 1
 
1  h21 2h1 1  h21
g jXc g_ xd 1 
2
l jlc g
1 h21 1 h21 b 1 h21
  7:40
1  h21
2 fd jfc xd g_ 0
1 h21
 
/f 1
(B) The higher vertical natural frequency, x x2 , has mode :
/b h2 ejhb
 
1  h22 2h2 1  h22
g  jXc _
g x 2
1 l  jl g
1 h22 d
1 h22 b c
1 h22
  7:41
1  h22
2 fd  jfc xd g_ 0
1 h22

These equations permit evaluation of effects of individual parameters.


(A) Gyroscopic effect: G ! c
As well as the isotropic support system, the frequency splits into the lower
backward natural frequency x1 and the higher forward natural frequency x2 . As
 flat
h approaches 1, which corresponds to very
2
elliptical
 orbits, the split is small due
to a weakening gyroscopic effect of G /f  j/b j ! 0 in Eq. (7.39).
2

(B) Direct damping: cd cxx cyy =2 ! fd


This parameter represents damping in both Eqs. (7.40) and (7.41): it is a sta-
bilizing factor for either the forward or backward whirling.
7.3 Natural Frequencies of a Rotor Supported 197

(C) Cross-stiffness: kc kxy  kyx =2 ! lc


The sign of kc in Eq. (7.39) is identical to lc in Eq. (7.11) so that this parameter
destabilizes the forward whirling x2 and stabilizes the backward whirling x1
as well as the isotropic system. However, the effect is
 decreased as h approaches 1,
2
  2
i.e. the orbit becomes flat, because kc /f j/b j ! 0 in Eq. (7.39). It nally
disappears at h = 1 where the orbit is a straight line.
Considering that the elliptical whirling orbit is the sum of two circular whirling
orbits, this situation is readily understood by dividing the effect of cross-stiffness
into the forward and backward components, Eq. (7.39) indicates that the effect of
cross-stiffness can be evaluated by
  
/f ejxt /b ejxt ! kc /f  j/b j2
2
7:42

The forward whirling component with an amplitude /f decreases the stability.


On the other hand, the backward whirling component with  an amplitude /b
increases the stability. When the orbit is a straight line, /f  j/b j and the two
opposite effects cancel out.
The anisotropy can be utilized for stabilization. For example, instability called
seal whirl, which is often observed in centrifugal compressors, can be stabilized by
load-on-pad (LOP) type tilting pad bearings. They provide large anisotropic stiff-
ness coefcients [60], or introduce anisotropy in the housing of a magnetic bearing.
Readers are referred to the simulation described for illustrating stabilization in the
next Sect. 7.3.6.
(D) Cross-damping: cc cxy  cyx =2 ! fc
This effect is usually small enough to neglect, as in isotropic systems.

7.3.6 Parameter Survey

The effects on stability of direct damping fd , cross-stiffness lc and support stiffness


anisotropy lb are now examined. Dividing Eq. (7.2) by the mass gives
 
z  jcX_z x2d 1  jlc z lb ejhb z 2fd xd z_ 0; 7:43

which can be rewritten by using dimensionless time s xn t as

z00  jcpz0 1  jlc z lb ejhb z 2fd z0 0 7:44

Table 7.3 summarizes the stability analysis of a system with fd 0:05, cp 0:1
and hb 160 in dependence on cross-stiffness, lc , and stiffness anisotropy, lb .
198 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

Table 7.3 Stabilization by anisotropic support (fd 0:05; cp 0:1; hb 160 )

No. 1 horizontally dominant whirling 2 vertically dominant whirling


y y
(1)

Anisotropic
stiffness x
= 0.052 + j 1.07
b = 0.1
x stable
= 0.048 + j 0.928
3 x
2 y
Cross- 1 y 1
stiffness
10 x
5 15 10 15
c = 0 1
2 5

1
y
(2) y
= 0.011 + j 1.06
Anisotropic
stiffness
unstable
b = 0.1
x x

= 0.11 + j 0.942
x y
Cross- 2 1
stiffness y x
10 10 15
c = 0.15 5 15 5
2
1

y
(3)
y
= 0.025 + j 1.14
Anisotropic
stiffness
x stable
b = 0.3
x
= 0.075 + j 0.847
2 x
1 y
Cross-
5 10 1
stiffness
y 15 x 10 15
c = 0.15 1
5
1
7.3 Natural Frequencies of a Rotor Supported 199

(1) Cross-stiffness, lc 0, stiffness anisotropy, lb 0:1


The eigenmodes corresponding to natural frequencies x1 and x2 have almost
same stability degree.
(2) Cross-stiffness, lc 0:15, stiffness anisotropy, lb 0:1
Cross-stiffness destabilizes the forward whirl eigenmode at x2 , while improving
the stability of the backward whirl mode at x1 .
(3) Cross-stiffness, lc 0:15, stiffness anisotropy lb 0:3
The higher stiffness anisotropy flattens the elliptical orbit (h2 ! 1), meaning
that the destabilizing effect of cross-stiffness is compensated for and the stability of
the forward whirl eigenmode is improved. On the contrary, the stability of the
backward whirl mode decreases.

7.4 Vibration Characteristics of a Jeffcott Rotor

The Jeffcott rotor [61, 62], consisting of a disk m xed on a massless shaft and a
stiffness ks supported by bearings (stiffness coefcient kd /2 and viscous damping
factor cd /2) at both ends (Fig. 7.12), is often used as a simple model for the bending
vibration of a flexible rotor.
Figure 7.13 shows an M-K-C model for the translational motion of a Jeffcott
rotor, sometimes called the 1.5 dof system. The complex eigenvalues of the system
are here identied by the approximate solution for the single-dof system described
previously.

7.4.1 Equation of Motion

Using the complex displacement z of the disk of mass m and that of the bearing
journal, zd, the equation of motion for a Jeffcott rotor is written as
     
m 0 z 0 0 z_ ks ks z
0
0 0 zd 0 cd z_ d ks ks kd zd 7:45
) _
M Z CZ KZ 0

The characteristic equation for the system is


 2 
Ms Cs K  0 7:46
200 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

Fig. 7.12 Jeffcott rotor z


zd
ks
m

kd cd 48 EI kd cd
ks = 3
2 2 l 2 2

Fig. 7.13 Jeffcott rotor z = x + jy


model m disk displacement

ks
zd
journal displacement

kd cd

For the sake of generality, the following parameters are used:


kd
r : stiffness ratio of the bearing and shaft
krs
ks
xs : natural frequency for simple support
rm r
ks kd r
x0 xs : undamped natural frequency of the Jeffcott
m ks kd 1r
rotor
cd
fd p: bearing damping ratio
2 mkd
kd r
as
ks kd 1 r
cd 2fd 2f
s p p d : time constant corresponding to bearing damping.
kd kd =m rxs

7.4.2 Vibration Characteristics

The vibration characteristics of a Jeffcott rotor have been studied extensively. One
of the most important ndings is the existence of the optimal damping coefcient.
7.4 Vibration Characteristics of a Jeffcott Rotor 201

(a) Im
1
d =5
0.1
= =1

0.5
= 0.13

Re
1 0.5
root locus
(b) 1 (c)
0.1
1 =5
0.1
3
0.1
1 =1

q / s
=

5 0.5
0.01 = 0.13
= 0.1

0.001 0
0.01 0.1 1 10 0.1 0.5 1 5 10
d = cd / ( 2 mk d ) d = cd / ( 2 mk d )
damping ratio = / | | damped natural frequency q Im[ ]

Fig. 7.14 Vibration characteristics of Jeffcott rotor (accurate eigenvalue k a jq)

[63] For example, Fig. 7.14 shows the effect of the bearing damping fd or cd on
the characteristic root k a jq in the case of four kinds of the support stiffness
ratio r f 0:1 0:13 1 5 g.
Figure 7.14a illustrates the root locus, which starts from the undamped char-
acteristic root k jx0 and reaches k jxs . Each of the loci passes through an area
of maximum damping ratio indicated by the most left side edge of the peninsula.
In addition, the loci for small r go through an overdamping area indicated by the
negative part of the real axis.
The variations of the damping ratio f  Rek=jkj and damping natural fre-
quency q are shown in Fig. 7.14b, c, respectively. It is clear that the option to
increase of the bearing damping cd indenitely is not necessarily a good design
policy, since a condition exists under which the damping ratio f is at a maximum.
With regard to unbalance vibration, the effects of r kd =ks (the stiffness ratio) and
cd Xc =kd (the equivalent of the bearing damping) on
jkj
the critical speed Xc p,
1  2f2
and the resonance response factor
202 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

600
ks
400 m
kd 0.01
/ kd =
kd cd cd
200 2 2 2 2 Cd c

0.02
2

100
1

80
1

60
40 0.05
2

10
t bearing
rolling elemen
0.1
Q =

= k d / k s = 0.2

4
20

2
10 0.2
Q-value

1
6

0.4
0.5
0.1

4
0.04

1
2 5
2 oil-film bearing

1
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
actual critical speed c
= s
simple support critical speed

Fig. 7.15 Chart of simple calculation of Q-value

1
Q p 7:47
2f 1  f2

are shown in Fig. 7.15. The left end of the abscissa corresponds to flexible support,
and the right end to rigid support. The plot, called the Balda chart [64], clearly
shows that an optimal bearing damping cd exists that minimizes Q for a given
stiffness ratio r. The chart is thus convenient to illustrate general vibration char-
acteristics of a Jeffcott rotor.
Figures 7.14 and 7.15 suggest that a general design policy would be flexibly
supporting the shaft and then aiming at the optimal bearing damping.
Methods of approximate analysis of the vibration characteristics are discussed in
the following sections.

7.4.3 Real Mode Analysis

The eigenvector u of a conservative system where the bearing damping cd = 0 is


 
1 1
u 7:48
ks =ks kd 1=1 r
7.4 Vibration Characteristics of a Jeffcott Rotor 203

(a) 1 (b)
1
2
1 =5
0.1
0.1 0.8 = 1
1
=
a

qa / s
5 0.6
0.01 2 0.4 = 0.1
0.2
1
0.001 0
0.01 0.05 0.1 0.5 1 5 10 0.5 1 5 10
d = c d / ( 2 mk d ) d = cd / ( 2 mk d )
Damping ratio = / | | Damped natural frequency q

Fig. 7.16 Accuracy of approximate solution [ real model, complex model]

According to Eq. (A.5) of Appendix A for the modal model, the modal
transformation

z
uz 7:49
zd

is substituted into Eq. (7.45) and the resultant formula is multiplied by ut from the
left to obtain the equation of motion of the approximate single-dof model:
 2
ks ks kd
mz cd z_ z0 7:50
k s kd ks kd

Therefore, the approximate values (indicated by the subscript a) of the damped


natural frequency q and the damping ratio f are

qa  x0
 2
cd ks fd 7:51
fa 
2mx0 ks kd 1 r3=2

The curves in Fig. 7.16 increase uniformly as bearing damping fd increases.


They compare the precision of these approximations with the exact solution. The
damped natural frequency qa and damping ratio fa are both accurate for small fd ,
but they do not show the existence of a maximum damping ratio, because the real
eigenvector u (more specically, the second row element in Eq. (7.48)) is assumed
and that this magnitude of eigenmode is not limited, even by high bearing
damping, fd .
204 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

7.4.4 Complex Mode Analysis

For a non-conservative system containing the bearing damping, cd, the complex
mode, uc , at the possible natural frequency x is
 
1 1
uc 7:52
ks =ks kd jxcd 1=1 r jsxr

which leads to an approximate single-dof model, as well as the real mode analysis
mentioned above. Note that the magnitude of the second row element in Eq. (7.52)
is limited by zero in the case of high bearing damping cd . Then

z
uc z 7:53
zd

mz Ceq z_ Keq z 0 7:54

where
h i
cd ks r 1 1 jsx2 r
Ceq utc Cuc ; Keq utc Kuc :
1 r jsxr2 1 r jsxr2

From the second and third terms of Eq. (7.54), the dynamic stiffness Geq for the
frequency x is

1 jsx 1 rr sxr2 jsxr


Geq jx Keq jxCeq ks a ks 7:55
1 ajsx 1 r2 sxr2

For approximation of the natural frequency q and damping ratio f of the damped
system, the frequency x in the right side is assumed to be the undamped natural
frequency x0 , and therefore the dynamic stiffness is Geq jx0 . The approximate
values are
r s s
ReGeq jx0  1 rr sx0 r2 1 r2 r 2fd r2
qa xs 2 2
xs
m 1 r sx0 r 1 r3 2fd r2
7:56
p
ImGeq jx0  1 x2s sx0 r fd r 1 r
fa 7:57
2mx2a 2 x2a 1 r2 sx0 r2 r1 r2 2fd r2

cd 2fd
Note that sx0 x0 p
is used in Eq. (7.57).
kd 1r
7.4 Vibration Characteristics of a Jeffcott Rotor 205

For small bearing damping cd (small s), omitting the second or higher order
terms of fd results a solution identical to Eq. (7.51).
Curves in Fig. 7.16 compare these approximations with the exact solution.
Their precision is far higher than the real mode approximations (curves ). The
existence of the damping ratio peak is also represented. However, the precision is
still less than satisfactory.
For furthermore improvement of the approximation precision, see Fig. 8.27
in the following chapter.

7.5 Analysis of Characteristics of Unbalance Vibration

7.5.1 Equation of Motion

The general form of the equation of motion for a rotor supported by anisotropic
bearings is obtained by adding a term for unbalance force to the right side of
Eq. (7.2):

mz  jXG_z kf z cf z_ kb z cbz_ meX2 ejXt 7:58

For simplicity, it is assumed here that the cross-stiffness kc = 0 and anisotropic


damping cb = 0:

z  jXc_z x2d z 2fd xd z_ x2d lb ejhb z eX2 ejXt 7:59

The anisotropy of stiffness is further assumed to be lb 0 for an isotropically


supported rotor:

z  jXc_z x2d z 2fd xd z_ eX2 ejXt 7:60

7.5.2 Unbalance Vibration of an Isotropically Supported


Rotor System

For the solution of Eq. (7.60) for forward circular whirling

z AejXt 7:61

the complex amplitude is given by


 1
A 1  cX2 2fd xd jX x2d eX2 7:62
206 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

A peak =11.2 Q =10.1


predicted value A peak =10 Q =10
1X
1.2 15
f F =0 d
c =
natural frequency ( d )

1.0
b 1
= 0.2 B
0.8 10 1
Q =8.8 A peak
1 2 d
measured value Q =10

amplitude
= 0.2
1 Re[ ]
5 Q= ,=
2 | |
=0
( for calculation )
c
Q= ( for measurement )

0 0
0.2 0.4 0.6 0.8 1 c 1.2 1.4
p= / d

Fig. 7.17 Unbalance resonance of symmetrically supported rotor system (fd 0:05 anisotropy
p
lb 0). Note As shown in Eq. (6.37), Q-value is measured by 1=2fb 10:1= 1  0:2 9:0.
Since the value is modied by gyroscopic effect, the peakness of unbalance amplitude curve is less
sharp and the foot range becomes wider.

Since the equation has been discussed in detail in Sect. 6.4, only the brief
overview is given here. As seen in Fig. 7.17, the natural frequency curves are split
into the forward (solid line xf ) and backward (broken line xb ) by the gyroscopic
effect c. The synchronous speed 1X is also shown in the gure, which intersects the
forward natural frequency (xf ) curve at the point F where the system resonates,
because the unbalance force is forward. The rotational speed at the intersection F is
the critical speed Xc . Since no resonance occurs at the point B where the backward
whirl natural frequency (xb ) curve intersects with 1X. The resonance curve has a
single peak F as shown in the gure. If the gyroscopic effect is absent (c 0), the
natural frequency curve is constant, and resonance occurs at Xc xd .
Calculation of the complex eigenvalue at the critical speed gives the Q-value and the
resonance peak amplitude Apeak shown at the top of the gure. The Q-value estimated
from the observed peak by the half power point method (see Sect. 2.4), is also shown
beside the peak in the gure, as a good approximation to the calculated value.

7.5.3 Unbalance Vibration of a Rotor Supported


by Anisotropic Bearings

For the solution of Eq. (7.59) for elliptical whirling,

 b ejXt
z Af ejXt A 7:63

the complex amplitudes Af and Ab satisfy


7.5 Analysis of Characteristics of Unbalance Vibration 207

(a) (b)
1X 1X
y y Q=11.5 15
Q=7.7 2 Q=11.8 Q=7.8
2
1.2 2 2
natural frequency ( d )

1.0 anisotropic

amplitude ( )
isotropic
0.8 1 10
x 1 1 1
x
Q=11.6 anisotropic
ax
Q=10.5 af
5
Q=7.6 ay ab
Q=7.2
0
0 0.2 0.4 0.6
0.8 1.0 1.2 1.4 0 0.2 0.4 0.60.8 1.0 1.2 1.4
p =/ d p = / d
X,Y sensors = 0.0 F,B sensors = 0.2
d = 0.05 anisotropy b = 0.4 b = 160

Fig. 7.18 Unbalance resonance of symmetrically supported by anisotropic bearing

  
1  cX2 2fd xd jX x2d lb x2d ejhb Af 1
eX2
lb x2d ejhb 1 cX2 2fd xd jX x2d Ab 0
7:64

The natural frequencies x1 and x2 are inherently different as shown in Fig. 7.18.
Resonance occurs at the two intersections and of the natural frequency curves
and the synchronous speeds noted 1X, giving the resonance curve with two peaks.
The x1 mode is an elliptical whirl along a horizontally biased axis, and the x2
mode is one along a vertically biased
 axis.
The forward amplitude af Af  and the backward amplitude ab jAb j
obtained by solving Eq. (7.64) are shown graphically in Fig. 7.18b. They can be
converted into the horizontal (X direction) amplitude ax and vertical (Y direction)
amplitude ay by the transformation of Eq. (6.51) and the corresponding curves are
shown in Fig. 7.18a.
Figure 7.18b shows that the whirling is forward if af > ab and backward if
af < ab. Therefore, the unbalance resonance at the critical speed always appears
as forward whirling, while that at the speed can be backward whirling.
As in the previous case, the damping ratio and Q-values obtained from calcu-
lation of the complex eigenvalues at the critical speeds are shown in the gure
beside the intersections and , as well as Q-values measured by the half power
point method noted beside the resonance peaks. They agree approximately with
each other, indicating that the calculation can predict unbalance vibration fairly
accurately. As the anisotropy decreases, the backward amplitude ab approaches
zero, resulting in resonance curves with a single peak mainly corresponding to the
intersection of af , corresponding to a forward circular whirl orbit.
208 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

7.6 Case Study: Vibrations of a Flexible Rotor


with Cylindrical Bearings

For a basic understanding of the flexible rotor vibrations related to the support
by sliding bearings, Kikuchis papers [65, 66] on a three-disk rotor supported by
cylindrical bearings are cited here. The dimensions of the rotor are shown in Fig. 12.1,
and the dynamic characteristics of the slider bearings are summarized in Fig. 12.3.
The resonance curves for the unbalance vibration of this rotor are shown in
Fig. 7.19. The calculated curves (solid lines) are in good agreement with experi-
mental results (circles).
The gure shows not only the response for a typical length to diameter ratio L/
D = 0.6 and a typical clearance ratio C/R = 0.001 , but also includes that for a
larger clearance C/R = 0.003 and for a still rather large C/R = 0.01 (though
unrealistic) for comparison. For larger clearances, the XY anisotropy of the bearing
dynamic characteristics becomes greater and the peak in the resonance curve splits.
The experimental data indicate a self-exciting vibration (oil whip) around X 90
Hz and it is a stable limit of rotational speed.

7.6.1 Critical Speed Map

Figure 7.20 shows the critical speed for a simplied model, which is supported by
the two bearings at both sides having identical stiffness coefcients kb. The support
stiffness is weak on the left and rigid on the right. The plot shows the dependence of
the critical speeds on the bearing stiffness and is called the critical speed map. It is a
convenient tool to determine general features in rotor design.

= 28cp = 45cp
=27cp experiment calculated
200 horizontal 32.1 rps 41.7 45.5
vertical
120 225
self-excited 270 180

amplitude [ m]

vibration 315 135


100 50 90
0 45
amplitude [ m]

80 40 40
self-excited
amplitude [ m]

60 vibration 30 30 self-excited
vibration
40 20 20

20 10 10

0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100


speed [rps] speed [rps] speed [rps]
1 L/D=0.6, C/R=0.001 2 L/D=0.6, C/R=0.003 3 L/D=0.6, C/R=0.01

Fig. 7.19 Vibration response (measured at the center disk)


7.6 Case Study: Vibrations of a Flexible Rotor 209

H V H V
300
200
150
3
100
critical speed
2
70
50

30 V 1 1 C/R=0.001
H
20 2 C/R=0.003
3 C/R=0.010

10
10 50 100 500 1,000 5,000 10,000
kb 104 N/m

Fig. 7.20 Critical speed map

The gure also shows the curves of dynamic stiffness of the bearing in the
horizontal (H) direction and in the vertical (V) direction. Letting X be rotational
speed in the ordinate axis and the possible bearing stiffness in the abscissa axis, the
following two curves are drawn for each different clearance ratio C/R:
V directional bearing stiffness = jkxx jXcxx j
H directional bearing stiffness = kyy jXcyy 
The intersections of the natural frequency curves and the horizontal and vertical
bearing stiffness curves (marked by circles) correspond to the critical speeds at
which resonance appears. For example, for case of C/R = 0.01 (a large clearance,
hence high anisotropy), the intersections correspond to a critical speed of about
35 Hz (horizontal) and one about 45 Hz (vertical), which agree with the peaks in
Fig. 7.19 .
The critical speed map thus permits fairly precise prediction of the critical speed
for a given bearing stiffness. It is frequently used as a basic tool in rotor design.

7.6.2 Calculation of Complex Eigenvalues and Q-Values

The complex eigenvalues k a jq are calculated for each rotational speed by


substituting the bearing dynamic characteristics into (Eq. 12.1) (presented later) and
by solving for the eigenvalues. Figure 7.21 shows the damped natural frequency
q and damping ratio f a=jkj dependences on the rotational speed X. Note that
the f axis is logarithmic between 1 and 0.1 and linear under 0.1.
These eigensolutions including eight parameters of oil-lm bearing are solved by
tracking [67] the continuous path of eigenvalue with variation of the rotational
speed, which is our unique method, being different from popular ways of inter-
polating eigensolutions obtained at each speed.
210 7 Approximate Evaluation for Eigenvalues of Rotor-Bearing Systems

natural frequency q [Hz]


60 1X 1X 1X
4 3 3 4 4
40 2
2 1 2
1 3
20
1 C/R=0.001 2 C/R=0.003 3 C/R=0.010
1
0
1.0 2
damping ratio = / | |

0.1 2 1
1 3 2
3
0.05 4
3
4
0
4
c = 93.7 [s1] c = 86.8 [s1] c = 88.4 [s1]
-0.05
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
[s1] [s1] [s1]
1 C/R=0.001 2 C/R=0.003 3 C/R=0.010

Fig. 7.21 Calculation of complex eigenvalue = + jq (natural frequency and damping ratio)

The intersections ( black and white circles) between the synchronous speed
1X and the natural frequency curves in the upper graphs give the critical speeds.
The corresponding damping ratios f are obtained from the lower graphs, which
allows estimation of the Q-values from Q 1=2f. For the intersections in the k3
and k4 modes, the latter showing a lower damping ratio predicting a higher
Q-value. Then, the intersections at black circles give Q-values of 200, 25 and 12
for the clearance ratios , and , respectively.

7.6.3 Root Loci

Representation of a complex eigenvalue on the complex plane with the rotational


speed X as a parameter is called the root locus as shown in Fig. 7.22. The system is
unstable and encounters oil whip when the real part of the eigenvalue enters the
positive region.
The oil whip in case occurs in the k1 (half-speed) mode, while that in cases
and in the k3 (shaft bending) mode. Different C/R values may thus cause
instability in loci of different modes, but all the actual self-excited vibrations occur
in the shaft bending mode of this system around X = 90 rps.
7.6 Case Study: Vibrations of a Flexible Rotor 211

Im Im Im
1 C/R=0.001 2 C/R=0.003 3 C/R=0.010
400 400 500

3 300 4
400
2 300
1
4 1
3 300
2
4
200 200 2
200
1
100 100 3
100

Re Re Re
40 20 0 200 100 0 400 200 0

Fig. 7.22 Root locus. White circle X 10s1 , Black circle X 100 s1 [44]

7.6.4 Resonance Curves for Unbalance Vibration

Examples of resonance curves for unbalance vibration are shown in Fig. 7.23.
Curves for greater clearance ratios C/R have two peaks: a horizontal vibration
followed by a vertical vibration, as particularly clear in the case , while those for
smaller clearance ratios tend to have a single peak (e.g. in the case ). The half
power point method gives Q = 20 for the case and Q = 12 and 5 for the case ,
while precise interpretation is impossible due to the excessively sharp peak for case
where Q > 100. The values obtained are in good agreement with the values
estimated in Fig. 7.21.
The upward arrows in the gure indicate the onset frequencies of oil whip,
which agree approximately with the rotational speeds at which the signs of the
damping ratios change from positive to negative as seen in Fig. 7.21.
In this eld, references [6878] provide further information and details on
specic problems. The reader is referred to these to experience the range of
problems concerning with complex eigenvalues of rotor-oil bearing systems dis-
cussed in the case study of Sect. 7.6.

250 60 40
Q > 100 ? 1 C/R=0.001 2 C/R=0.003 resonance 3 C/R=0.010
200 50 resonance
amplitude [ m]

resonance self-excited 30 self-excited


self-excited 40 Q=20 vibration Q=12 vibration
150 vibration
30 20
100
20 Q=5
10
50 10

0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100


speed [rps] speed [rps] speed [rps]

Fig. 7.23 Unbalance resonance curves of a rotor supported by circular cylindrical bearings
Chapter 8
Rotor System Evaluation Using
Open-Loop Characteristics

Abstract This chapter discusses an evaluation method for rotor vibration charac-
teristics by utilizing the open-loop frequency response of the system, instead of
conventional eigenvalue analysis. The vibration characteristics of a rotor system are
represented by the (damped) natural frequency and damping ratio. They have been
estimated in the previous chapters from the viewpoint of the eigenvalue solution,
the impulse response waveform, and the resonance curve (FRA) under harmonic
excitation. The open-loop characteristics are from a concept in control engineering.
A rotor-bearing system can be conceived as a control system as shown in Fig. 8.1,
of which the open-loop characteristics are related to the vibration characteristics: the
gain cross-over frequency is an estimate of the natural frequency, and the phase
margin is an indicator for the damping ratio. Details of estimation are described
below.

  
Keywords Open loop Transfer function Phase margin Gain cross-over fre-
 
quency Stability margin Tuning of spring and damping coefcients

8.1 Open-Loop Analysis of a Single-dof System

8.1.1 Open-Loop Frequency Response of a Single-dof


System

As an example of Fig. 8.1, the equation of motion for an m-k-c system shown in
Fig. 8.2

Fig. 8.1 Typical unbalance journal


rotor-bearing system r=0 displacement
bearing rotor
+

Springer Japan 2017 213


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_8
214 8 Rotor System Evaluation Using Open-Loop Characteristics

Fig. 8.2 Simple vibration x


system
m

k c

Fig. 8.3 Block diagram u (s) x (s)


1
k + cs 2
ms

mxt ut
8:1
ut kxt c_xt

in the time domain, which can be rewritten for the s-domain using the Laplace
transformation, assuming that the initial displacement and velocity are zero, as

ms2 xs us
8:2
us k csxs

The corresponding block diagram is shown in Fig. 8.3.


The system has an open-loop transfer function:

k cs k jxc
Go s ! Go jx 8:3
ms2 mx2

Its Bode plot is shown schematically in Fig. 8.4.


The frequency at which the gain is 1 (=0 dB) is called the gain cross-over
frequency xg , and the distance of the phase curve at this frequency above 180 is
the phase margin /m .
For the gain cross-over frequency xg ,
 
k jxg c
1 8:4
mx2g
8.1 Open-Loop Analysis of a Single-dof System 215

gain [ dB ]
Fig. 8.4 Gain cross-over
frequency xg and phase ga
in
margin /m
g
0 dB
frequency

phase [ ]
phase
m
180
frequency

 4  2
xg xg
) 4f2 1 0 8:5
xn xn
p
where xn k=m; 2fxn c=m.
The solution is
r
q
xg xn 2f2 1 4f4 exact solution 8:6
q
xg  xn 1 2f2 approximation for f  0 8:7

Figure 8.5 compares the exact and approximate solutions with related variables:
Damped natural frequency:
q
q xn 1  f2 8:8

Critical speed:
q
Xc xn 1  2f2 8:9

The comparison of the magnitudes of each speed and frequencies is as follows:


Critical speed Xc  gain cross-over frequency xg (exact) as seen in Eq. (8.6)
 gain cross-over frequency xg (approximate) as seen in Eq. (8.7)

Fig. 8.5 Comparison of 2 t


c xac
various approximate types g e
frequency ( n )

1.5
g approximate
1
n
0.5 q

0
0 0.2 0.4 0.6 0.8 1
damping ratio
216 8 Rotor System Evaluation Using Open-Loop Characteristics

Fig. 8.6 Phase lead angle


Dynamic
x xg
stiffness k + j g c
jgc
m

 undamped natural frequency xn


 damped natural frequency q
In a system with small damping, all of these variables approach the undamped
natural frequency xn . In addition, the gain cross-over frequency xg is close to the
critical speed Xc , and may therefore be considered to indicate the latter, in the
practical range of f (around 00.4).
The phase margin /m is related to the damping ratio f as follows. The numerator
of the open-loop characteristic Eq. (8.3) means that the argument of the triangle
k jxc (Fig. 8.6) is the phase margin:
xg c
tan /m 8:10
k

The relationship of the phase margin /m to the damping ratio f is obtained by


rearranging Eq. (8.10) with xn c=k 2f:
r
q
xn c xg
tan /m 2f 2f2 1 4f4 exact 8:11
k xn
q
xn c xg
tan /m  2f 1 2f2 approximate for f  0 8:12
k xn

The exact relationship Eq. (8.11) and the approximate relationship Eq. (8.12) are
represented graphically by the curves and , respectively, in Fig. 8.7. Curve
is a good approximation of Curve . Further approximation is obtained by
assuming /m  0 in Eq. (8.12):

Fig. 8.7 Prediction of 1.6


simplified formula
damping ratio 1.4 by Eq.(8.13) 3
damping ratio

1.2
1.0
0.8 = 1 tan m 2
2 approximate formula
0.6 by Eq.(812)
0.4
0.2 1 exact formula
by Eq.(811)
0
0 20 40 60 80 100
phase margin m [ ]
8.1 Open-Loop Analysis of a Single-dof System 217

enlarged diagram
40 20 165
gain []
gain [ dB ]

20

phase [ ]
g 80 10 170
area to be phase m
0 100 gain
enlarged
120 0 175
20 m 140
160 phase g

40 180 10 180
0.1 0.5 1 5 10 50 100 0 0.5 1 1.5 2
frequency / n frequency / n

Fig. 8.8 Open-loop characteristics of single-dof system (m = 1, k = 1 and c = 0.2)

1
tan /m 2f , f tan /m simplified approximation for f  0 8:13
2

Equation (8.13) corresponds to Curve in Fig. 8.7. Although overestimating


the damping ratio f, it is an acceptable approximation within the practical range of
/m = 040, and will be used consistently hereinafter.
Example 8.1 Figure 8.8 shows the Bode plot for the open-loop frequency response
of the system shown in Fig. 8.2 with m = 1, k = 1 and c = 0.2. Note that the
frequency axis is converted to dimensionless form using the ratio to the undamped
natural frequency xn . The right side gure gives the enlarged plot around the gain
cross-over frequency.
(1) Read the gain cross-over frequency xg and the phase margin /m from the plot.
(2) Using these values, estimate the critical speed Xc , the damping ratio f and
Q-value.
(3) Find the characteristic root k directly obtained by solving the characteristic
equation of the system, and compare with the estimated damping ratio f.

Answer
(1) xg 1 /m 11:4 ;
(2) Xc xn 1; f 0:1 estimated by Eq. (8.13) and Q 1=2f 5
(3) k2 1 0:2k 0 gives k 0:1  j0:995. Therefore f Re k=k 0:1
agrees with the estimate.

Example 8.2 Figure 8.9 shows the Bode plot for the open-loop frequency response
of the system shown in Fig. 8.2 with m = 1, k = 1 and c = 0.8. Note that the
frequency axis is converted to dimensionless form using the ratio to the undamped
natural frequency xn . The right side gure gives the detailed plot around the gain
cross-over frequency.
218 8 Rotor System Evaluation Using Open-Loop Characteristics

enlarged diagram
40 20 -120
gain
20 []

phase [ ]
g phase
gain [dB]

10 -140
80 gain m
area to be
0 enlarged 100
120 0 -160
20 140
m phase margin 160 phase g
40 180 10 -180
0.1 0.5 1 5 10 50 100 0 0.5 1 1.5 2
frequency / n frequency / n

Fig. 8.9 Open-loop characteristics of single-dof system (m = 1, k = 1 and c = 0.8)

(1) Read the gain cross-over frequency xg and the phase margin /m from the plot.
(2) Using these values, estimate the critical speed Xc , the damping ratio f and
Q-value.
(3) Find the characteristic root k of the system directly from complex eigenvalue
calculation, and compare with the estimated damping ratio f.

Answer
(1) xg 1:17 /m 43:4 ,
(2) Xc xn 1:17; f 0:41 estimated exactly by Eq. (8.11)
or f .0:47 estimated approximately by Eq. (8.13),
p
Q 1 2f 1  f2 1:33
(3) k2 1 0:8k 0 gives k 0:4  j0:917. Therefore f 0:4 agrees fairly
well with the estimate.

Example 8.3 Consider a rotor-bearing system with a series coupling support as


shown in Fig. 8.10, called the 1.5-dof model or Jeffcott rotor model.

Fig. 8.10 1.5-dof Jeffcott


rotor model m

ks

k c
8.1 Open-Loop Analysis of a Single-dof System 219

40 20 150
enlarged diagram
gain [] 15
gain [ dB ]

20 phase

phase [ ]
14.2 dB
130
10 gain 160
0 area to be 140 5
enlarged m
g 150 0 170
20 phase 160
5 g
170
m
40 180 10 180
0.1 0.5 1 5 10 50 100 0 0.5 1 1.5 2
frequency / n frequency / n

Fig. 8.11 Open-loop characteristics of 1.5-dof model

(1) Show that the open-loop characteristics of the system are given by the transfer
function

1 ks k cs
Go s 8:14
ms2 ks k cs

(2) Figure 8.11 shows the Bode plot for the system with m = 1, ks = k = 1 and
c = 0.8. Note that the frequency axis is converted to dimensionless using ratio
of the natural frequency of the shaft supported pin-pin condition;
p
xs ks =m. The right side gure gives the detailed plots around the gain
cross-over frequency. Read the gain cross-over frequency xg and the phase
margin /m from the plots.
(3) Using these values, estimate the critical speed Xc , the damping ratio f and
Q-value.
(4) Find the characteristic root k of the system directly from the complex eigen-
value calculation, and compare with the estimated damping ratio f.
(5) Assuming that the mass m is variable, nd the mass needed to obtain the
maximum damping ratio. Correspondingly, predict the critical speed and the
expected maximum damping ratio.

Answer

(1) The equivalent dynamic stiffness Keq 1=ks 1=k cs1 gives Go s
Keq
by Eq. (8.3).
ms2
(2) xg 0:75 /m 14:5 ;
(3) Xc x
.n p0:75; f 0:13 estimated approximately by Eq. (8.13),

Q 1 2f 1  f2 3:9
(4) 1 Go s 0 gives k 0:1  j0:73. Therefore f 0:13 agrees with the
estimate.
(5) The maximum damping ratio may be given at maximum portion of the phase
lead curves as shown by the broken line in the gure. We are thus
220 8 Rotor System Evaluation Using Open-Loop Characteristics

Fig. 8.12 Comparison of 40 90


open-loop characteristics 100
20 gain 2 110
g 1 120

phase [ ]
gain [dB]
0 130
3 140
20 2 150
1
phase 160
40 170
m 3
180
0.1 0.5 1 5 10 50
frequency / n

recommended to set the gain cross-over frequency xg at the maximum phase


lead by shifting the gain curve upwards by 14.2 dB. This is achieved by
reducing the mass by 14.2 dB = 0.19 = 19 %, resulting in a critical speed
p
xg Xc 0:75 1=0:19 1:75, a maximum phase lead /m 20 , and a
maximum damping ratio f 1=2 tan 20 = 0.18.

Example 8.4 Figure 8.12 compares the Bode plots in Examples 8.18.3.
(1) Conrm the following vibration characteristics:
Example 8.1 (curve ) : xg 1 /m 11:4 f 0:1 Q 5
Example 8.2 (curve ) : xg 1:17 /m 43:4 f 0:41 Q 1:33
Example 8.3 (curve ) : xg 0:75 /m 14:5 f 0:13 Q 3:9
(2) Discuss differences of the curves , and .

Answer
(1) omitted
(2) The damping ratio increases as the damping increases in the single-dof model
and . In contrast, the 1.5-dof model has a region corresponding to the
maximum phase lead which expects a maximum damping ratio. Since actual
rotor-bearing systems correspond to Example 8.3, we need to design the
optimization of bearing parameters.

Example 8.5 The amplitudes of unbalance vibration in the systems are dened:

me X2
A for Examples 8:1 and 8:2; 8:15
ms2 k cs

and

me X2
A for Example 8:3: 8:16
ks k cs
ms2
ks k cs
8.1 Open-Loop Analysis of a Single-dof System 221

Fig. 8.13 Unbalance 6


vibration resonance curve 5 Q = 3.4 70%

amplitude
3
4
Q = 4.6
3 1
70%
2
1
2
0
0 0.5 1 1.5 2
speed / n

The resulting resonance curves are shown in Fig. 8.13. Read the Q-values by the
half power point method and compare them with the estimated values of respective
example.
Answer
Curve : Q-value by half power point method is 4.6, which agrees with the
estimation Q = 5.
Curve : The half power point method does not work (However, the Q-value is
estimated numerically as 1.33).
Curve : Q-value by half power point method is 3.4, which agrees fairly well with
the estimation Q = 3.9.

8.1.2 Measurement of Open-Loop Frequency Response

The block diagram for a single-dof system (Fig. 8.14) represents a closed loop
system. The open-loop frequency response of this system can be measured by
harmonic excitation E ejxt at one end of the system. The ratio of the frequency
response amplitudes before and after the input point gives the open-loop transfer
function Go:

( by measurement )
r =0 u 1 x
k + cs open-loop GO = V 1 / V 2
ms 2
( by signal processing )
+ V1 GO V1
closed loop G c = =
V2 1 + GO E
+
1 V2
sensitivity G s = =
jt 1 + GO E
E=e

Fig. 8.14 Measuring open-loop transfer function


222 8 Rotor System Evaluation Using Open-Loop Characteristics

V1 s V1 s
Go s   8:17
V2 s V2 s Es

Since, in fact,

ms2 xs k csxs Es 8:18

under this condition,

k cs
V1 s
xs  Es
ms2 cs k
8:19
ms2
V2 s
V1 s Es 2 Es
ms cs k

Therefore,

V1 s k cs
Go s
 8:20
V2 s ms2

which is the open-loop transfer function. A practical procedure is to calculate the


ratio of the amplitude responses V1 jx and V2 jx is to use harmonic excitation.
The closed loop transfer function Gc s and sensitivity function Gs s are obtained
by signal processing [79, 80] as shown in Fig. 8.14.

8.2 Modal Open-Loop Frequency Response

8.2.1 Modal Model

The equation of motion for vibratory systems consisting of mass M, stiffness K and
damping D in general matrix form is

KXt DXt
M Xt _ 0 8:21

or, expressed in the s domain,

Ms2 Xs U
8:22
U K DsXs

The corresponding block diagram is shown in Fig. 8.15.


The modal matrix is dened by arraying the eigenvectors of a conservative
mass-stiffness system:
8.2 Modal Open-Loop Frequency Response 223

Fig. 8.15 Physical U


M 1
X
coordinate system K + Ds s2

Fig. 8.16 Physical U X


M 1
coordinate X and modal K + Ds s2

1

coordinate g
X

U /1 /2    /n  8:23

Performing a transformation of the physical coordinate to the modal coordinate,

Xt Ugt; Xs Ugs 8:24

Substituting Eq. (8.24) into Eq. (8.22) yields

s2 MUgs U
8:25
U K DsUgs

The corresponding block diagram is shown in Fig. 8.16, which is a mix of the
physical coordinate X and the virtual modal coordinate g.
A congruent transformation is performed on Eq. (8.25) by multiplying the
transposed mode matrix from the left:

s2 Ut MUgs U
8:26
U Ut K DsUgs

Since the modal matrix U is orthogonal with respect to the mass matrix M and
stiffness matrix K, the matrix obtained by the congruent transformation is diagonal:

Ut MU diagonal m 1 m 2    m n 
8:27
Ut KU diagonal k1 k2    kn 

The diagonal elements of these matrices are called modal mass, mi*, and modal
stiffness, ki*.
The damping matrix D is not generally diagonalized by the congruent trans-
formation, but it can practically be regarded as diagonal.
224 8 Rotor System Evaluation Using Open-Loop Characteristics

Fig. 8.17 Modal block (a) 1


diagrams for each mode k 1* + d 1*s 1
m 1* s 2

2
k 2* + d 2*s m 2* s 2
3
k 3* + d 3*s m 3* s 2

(b) 1
k i* + d i*s i
m i* s 2

No. i

Ut DU  diagonal d1 d2  dn  8:28

The modal damping, di*, can thus be assumed by the diagonal elements only.
All the coefcient matrices in Eq. (8.26) are thus diagonal, and the system can be
separated into individual modes for the diagonal elements. The corresponding block
diagram is shown in Fig. 8.17a, which consists of components shown in Fig. 8.17b.
Diagram (a) can be regarded as a set of the block diagram (b) of independent
single-dof systems corresponding to individual modes:
q
Natural frequency xi ki =m i
. p 8:29
Modal damping ratio fi di 2 m i ki

can be dened here.

8.2.2 Modal Open-Loop Frequency Response

Figure 8.18a shows the block diagram of the rst mode system extracted from
Fig. 8.17. As well as the open-loop measurement of the single-dof system, we
consider the measurement method for the open-loop frequency response in the
modal coordinate for the rst mode /1 of a multi-dof system. The open-loop
transfer function Go is principally obtained as the ratio V1 =V2 of the response
amplitudes before and after inputing the harmonic excitation E ejxt :

V1 s
GO s  8:30
V2 s

This is the modal expression for the virtual coordinate. In practice, we con-
sider real measurement in the physical coordinate as illustrated in the lower right
8.2 Modal Open-Loop Frequency Response 225

1
k 1* + d 1* s 1
m 1* s 2
open-loop
(a) principle V1 + measurement
+ j t
E= e
V2

U X 1
M 1 1
K + Ds s2

(b) practice
V1
j t
V2 E= e
+

Fig. 8.18 Measurement of modal open-loop response

side of Fig. 8.18b. When an excitation force E ejxt is applied to the g1 coordinate
of the rst eigenmode,

Ms2 Xs K DsUgs /1 Es



1 8:31
) Xs  Ms2 Ds K K Ds/1 Es

The response amplitude Xjx of each mass point in the physical coordinate is
therefore calculated as

1
X jx  Mx2 K jxD K jxD/1 8:32

Considering Eq. (8.24), it is converted to the amplitude response in the modal


coordinate g1 jx for the rst mode /1 as

g1 jx Cgjx CU1 X jx 8:33

where C 1 0    0  is the output matrix.


In practice, the modal matrix U may contain less number of columns than rows
due to mode truncation. An approximate inverse matrix is used in such a case since
U is not diagonal:
1
g1 jx Cgjx C Ut U Ut X jx 8:34

Considering that V1 g1 and V2 V1 1, the open-loop frequency response


for the rst modal coordinate is obtained by

g1 jx
Go1 jx  8:35
g1 jx 1
226 8 Rotor System Evaluation Using Open-Loop Characteristics

gain

1 2 3 4
1 0 dB

open-loop characteristic
phase
m1 m2 m 4
m3
2 180

3 open-loop characteristic

1 2 3 4

1st 2nd 3rd 4th


modal frequency band

eigen mode 1 2 3 4

modal response 1 2 3 4

Fig. 8.19 Modal open-loop characteristics of multi-dof system

The modal open-loop frequency response Goi jx for the i-th mode is obtained
similarly replacing g1 by gi .
In practice, since the modal excitation and the modal response measurement is
performed in the limited frequency bands around the natural frequency xi of each
eigenmode /i , a set of Goi jx (i = 1, 2, .) is nally obtained as shown in
Fig. 8.19. This yields the open-loop frequency response (plots of gain and
phase ) for individual natural frequency bands. The gain curve intersects the line
of gain = 1 = 0 dB at the gain cross-over frequency xg , i.e., the natural frequency
of the mode. The phase margin /m , or the phase lead above 180 , can be con-
verted into the modal damping ratio f by Eq. (8.13).
Example 8.6 For the 3-dof system shown in Fig. 8.20,
(1) Find the mass matrix M, stiffness matrix K and damping matrix D.
(2) Conrm that the natural frequency xi and eigenmode /i of the undamped M-
K system are given by Fig. 8.20.
(3) Find the eigenvalues of the damped M-D-K system, and conrm that the
damping ratios given by the exact damped eigenvalues corresponding to
individual modes given in Fig. 8.20.
(4) Conrm that the modal open-loop transfer functions are given by these in
Fig. 8.21.
(5) Read the gain cross-over frequencies xg and phase margins /m of the modal
open-loop transfer functions from Fig. 8.21 and estimate approximate damp-
ing ratios. Compare the results with the exact values obtained in (3).
8.2 Modal Open-Loop Frequency Response 227

x1 x2 x3
12 12
f1 6 4 4

12 1 4 2

3 3
2
1
1 = 1
1 = 0.115
1 2
2 = 2 1
2 = 0.075
3
2
3 = 3
3
3 = 0.022
5

Fig. 8.20 3-dof system

20 140
1st mode 2nd mode 3rd mode
10 150
gain [dB]

phase [ ]

g g g
0 160

10 170
m m m
20 180
0 1 2 3 4 5
frequency

Fig. 8.21 Modal open-loop transfer function

Answer
2 3 2 3 2 3
6 0 0 24 12 0 1 0 0
(1) M 4 0 4 0 5; K 4 12 24 12 5; D 4 0 0 0 5
0 0 4 0 12 16 0 0 2
(2) The undamped eigenvalue problem x2n M/ K/ gives
the eigenvalues x2n f 1 4 9 g, hence the natural frequencies
xn f 1 2 3 g. 2 3
2 1 2
The modal matrix U 4 3 0 5 5 gives the mode shapes shown in
3 1 3
Fig. 8.20.
228 8 Rotor System Evaluation Using Open-Loop Characteristics
 
(3) The characteristic equation Ms2 Ds K  0 gives
Damped eigenvalues s f 0:116  j0:998 0:15  j1:99 0:067  j2:99 g
Exact modal damping ratios: fe Re s=jsj f 0:115 0:0754 0:0225 g
(4) The response in the modal coordinate is:
1
gs U1 Ms2 Ds K Ds K U (3 3 matrix)
The modal open-loop transfer functions are dened as
1st mode: G0 jx g1; 1=g1; 1 1 0 \ x \ 1:5 ,
2nd mode: G0 jx g2; 2=g2; 2 1 1:5\x\2:5 and
3rd mode: G0 jx g3; 3=g3; 3 1 2:5 \ x ,
A set of each G0 jx yields Bode plots of Fig. 8.21.
(5) The gain cross-over frequencies xg and phase margins /m read from Fig. 8.21
are:
xg f 1 2 3 g and

/m f 13 8:8 2:8 g ! fa 0:5tan/m f 0:116 0:077 0:0244 g


They give the estimation of damping ratios fa , which agree with fe obtained in
(3).

8.3 Open-Loop Frequency Response of a Jeffcott Rotor

As stated in Sect. 7.4, the following variable and dimensionless parameters are used
in this section.
r kd =ks : ratio of the bearing stiffness to the shaft stiffness
p
xs ks =m: natural frequency for simple support
p p
x0 ks kd =ks kd =m xs r=1 r: undamped natural frequency
cd
fd p: bearing damping ratio
2 mkd

8.3.1 Series Coupling and Phase Lead Function

In Fig. 8.22, a 1.5-dof model (b) of a Jeffcott rotor (a) supported by bearings with
direct spring kd and viscous damping cd, the shaft stiffness ks and bearing dynamic
stiffness kd + cds can be connected by a series coupling for Gr s:
cd s
1
1 ks kd cd s ks kd kd
Gr 8:36
1 1 ks kd cd s ks kd kd cd s
1
ks kd cd s ks kd kd
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 229

m m
(a) z (b) (c)
zd
ks ks
m 0 < s < 1

kd cd 48 EI kd cd 1 + s
ks = 3 kd cd G r = ks s
2 2 l 2 2 1 + s s

Fig. 8.22 Jeffcott rotor model

1 ss
) Gr s ks as 8:37
1 as ss

r cd 2fd 2f
where as , and s
p p d is the time constant corre-
1r kd kd =m rxs
sponding to the bearing damping. This means that the rotor can be considered as a
system supported by the transfer function Gr as illustrated in Fig. 8.22c. Since
0\as \1, the transfer function corresponds to a phase lead circuit as well known
for use in an electronic control system.

8.3.2 Open-Loop Frequency Response

A Jeffcott rotor is represented by a block diagram of Fig. 8.23. The open-loop


frequency response is:

G r s ks kd cd s x2 g1 s
Go s 2 2s as 8:38
ms 2 ms ks kd cd s s g2 s

where
1 s
g1 s 1 ss 1 2fd p
r xs
1 s
g2 s 1 as ss 1 as 2fd p
r xs

The characteristic equation is:

1 Go s 0 8:39
230 8 Rotor System Evaluation Using Open-Loop Characteristics

k s ( k d + cd s ) 1 + s 1 z
G r (s ) = = k s s 2
k s + k d + cd s 1 +s s ms

Fig. 8.23 Block diagram of 1.5-dof Jeffcott rotor

8.3.3 Gain Cross-Over Frequency and Phase Margin

An example of the Bode plot of the transfer function Eq. (8.36) is shown in
Fig. 8.24. If the rotor is rigid (ks 1 and Gr s kd cd s), the gain g jGr jxj
monotonically increases and the phase lead \Gr jx by up to 90, as bearing
damping coefcient cd increases. If the shaft stiffness ks is nite, the gain g
jGr jxj is saturated at ks, and the phase lead curve has a peak in a certain frequency
domain in which the maximum damping effect is expected.
The open-loop frequency response represented by Eq. (8.38) gives the same
phase plot as the phase lead function of Eq. (8.36). As the gain cross-over fre-
quency xg , at which the gain curve intersects the line of 0-dB, is the natural
frequency, it is desirable for the phase lead to be maximal at the intersection in
the gure; a design using the intersection is unfavorable because of lower phase
lead. Such design optimization is relatively easy for controllable magnetic bearings,
but not always so for oil bearings.

Fig. 8.24 Phase lead and 90



=

open-loop characteristics
Gr ( j )

ks


= phase

log

ks =
gain g = 20 log 10| Gr ( j )|

open-loop ks
open-loop

g g
0
1 2 log
g kd
ks
kd + ks
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 231

(a) (b) (c)


cd cd
1 1
kd cd 1 k kd
ks d ks ks
cd 2 2
2 cd cd

1 2 1 2
( low freq. ) ( mid freq. ) ( high freq. )

Fig. 8.25 Actual phase lead h1  h2

Fig. 8.26 Open-loop (a)


characteristics 40 100
=1
20 d = 0.1 120
0.8

phase [ ]
gain [dB]

gain
g 5
0 140
m
d = 5 0.1
20 se 160
pha
0.8
40 180
0.01 0.05 0.1 0.5 1 5 10
/s
(b)
20 150
d = 0.1
15 gain 0.8
= 5 5
10 160
phase [ ]
gain [dB]

5
0.8
e
phas
0 170

5
0.1
10 180
0 0.2 0.4 0.6 0.8 1

The existence of a peak in the phase curve for the 1.5-dof system is explained in
Fig. 8.25. The upper half of the gure shows the phase lead h1 \kd jxcd ,
which appears in the numerator of Eq. (8.36), and the lower half the phase lag
h2 \ks kd jxcd in the denominator. As h1 [ h2 geometrically, the dif-
ference h1  h2 is the actual phase lead for the damping force acting on the
system. Since h1  h2 approaches zero for x either increasing in high frequency
domain or decreasing in low frequency domain, there must be a maximum phase
lead at an intermediate frequency.
If ks is small and kd is large, the phase lead approaches zero so that damping is
not effective; conversely if ks is large and kd is small, the phase lead becomes large
232 8 Rotor System Evaluation Using Open-Loop Characteristics

and the damping effect is thus expected. This means that high shaft stiffness and
flexible support are preferable in designing for a high-damping specication.
As an example, Fig. 8.26 shows calculation results for the open-loop transfer
function Eq. (8.38) with given parameters fd = {0.1, 0.8, 5} assuming that the
stiffness ratio r = 1.
The gain cross-over frequencies, xg , at which the gain curves intersect the line
of 0 dB, are indicated by white circles; they are regarded as the natural frequencies.
The gure shows xg =xs = {0.7, 0.83, 0.98}.
The black circles indicate the phase margins, /m , which are converted to the
damping ratios f = 0.5 tan /m = {0.035, 0.18, 0.05}, which are in good agreement
with the exact values {0.035, 0.206, 0.05} obtained from the complex eigenvalues.
This example shows that, as the given fd changes, the peak in the phase curve shifts
while the gain curves remain approximately the same. The phase margin /m
coincides with the peak of the phase curve at the given fd = 0.8, but not at the given
fd = 0.1 or 5. The bearing damping parameter fd = 0.8 is therefore recommended;
other fd values are not acceptable.

8.3.4 Precision of Approximate Solutions

The transfer function Gr jx for the 1.5-dof system consisting of the phase lead
function, coincides to the dynamic stiffness Geq jx Keq jxCeq in the complex
modal analysis as stated in Eq. (7.55). The precision of the approximate solution
obtained by the replacement from Geq jx to Gr jx is thus guaranteed within the
same degree seen in Fig. 7.16. In order to improve the accuracy, the dynamic
stiffness should be assessed based on a more approximate frequency.
The gain cross-over frequency xg is recommended for this case. This frequency
is determined by
 
Go jxg  1 8:40

Applying this to Eq. (8.38) yields


 2  
xg g1 jxg 

as  
xs g2 jxg 
 4
2
2 8:41
xg r2 sxg r r2 2fd 2 xg =xs r
)
2
2
xs 1 r2 sxg r 1 r2 2fd 2 xg =xs r

Solving this equation gives xg for approximate natural frequency. The results
are shown as the curves of Fig. 8.27b, which shows good agreement with the
exact solutions except for the over-damped region trapped locally by the eigen
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 233

(a) (b)
1
3 g
1 =5
0.1

q / s , g / s
0.1 0.8 =1
1
a

= 0.6
5
0.01 3 0.4 = 0.1
0.2
0.001 0
0.01 0.05 0.1 0.5 1 5 10 0.5 1 5 10
d = c d / ( 2 mk d ) d = cd / ( 2 mk d )

damping ratio = / | | damped eigenvalue q = j


gain cross-over frequency g

Fig. 8.27 Accuracy of approximate solution ( open-loop)

frequency q = 0, indicating non-vibratory behavior. The phase margin /m is the


difference of the arguments g1 and g2:

1 sjxg 1  as sxg
/m \g1 jxg  \g2 jxg \ 8:42
1 as sjxg 1 as sxg 2

Substituting the gain cross-over frequency xg found for a given bearing damping
fd into Eq. (8.42) provides the phase margin /m , which is converted to the esti-
mated damping ratio fa :
p
1 1 sxg r fd xg =xs r
fa tan /m
2
2 8:43
2 2 r1 r sxg r r1 r 2fd 2 xg =xs r

Note that the damping ratio can also be determined directly by Eq. (8.43)
without knowing the phase margin.
Examples of this calculation of Eq. (8.43) are shown in Fig. 8.27a as curves ,
showing that the estimated values agree almost perfectly with the exact solutions
over the entire range of the given parameters fd . The precision is thus much higher
than the curves and in Fig. 7.16. This improvement is due to the gain
cross-over frequency xg used in the calculation. This example illustrates the
effectiveness of the open-loop method.
The low Q-value design method shown in Fig. 7.15 is further discussed here
referring to the ndings above. Since the critical speed is approximated by the gain
cross-over frequency xg , Eq. (8.41) is thus rewritten for the critical speed Xc using
sxg s Xc  sx0 cd x0 =kd :
 4
Xc 1 cd x0 =kd 2
a2s 8:44
xs 1 a2s cd x0 =kd 2
234 8 Rotor System Evaluation Using Open-Loop Characteristics

Likewise the damping ratio f can be obtained from Eqs. (8.42) and (8.43), and
converted to the Q-value:

1 1  as cd x0 =kd
f 8:45
2 1 as cd x0 =kd 2

1
Q p 8:46
2f 1  f2

Plotting Xc =xs on the abscissa and the Q-value on the ordinate using these
results gives the same graph as in Fig. 7.15. Thus, Eqs. (8.44)(8.46) serve as the
guidelines for achieving the optimal damping (low Q-value) in rotor design.
Example 8.7 For a Jeffcott rotor supported by tilting pad bearings for a centrifugal
compressor with m = 148 kg, ks = 43.4 MN/m, kd = 53.7 MN/m, and cd = 48.7,
(1) Approximate the critical speed Xc and Q-value.
(2) Calculate the complex eigenvalues k 32:56  j408:2 and check the pre-
cision of the approximate values.

Answer
(1) With xs = 86.3 Hz, r = 1.24, a = 0.55, x0 = 64.2 Hz, sxg cd x0 =kd
0:365 and sxg  s Xc , Eq. (8.44) gives Xc = 65.4 Hz. Equation (8.45) gives
f 0:076, therefore Q = 6.58 according to Eq. (8.46).
(2) The complex eigenvalues gives exact values with Xc jkj=2=p = 65.6 Hz,
f Rek=k = 0.08 and Q 1=2f 6:3. The approximate values (1) are
in good agreement with the exact values (2).

8.3.5 Optimal Damping

Tuning of the system to nd out the optimal damping as completing the maximum
damping ratio, is now discussed. For a phase lead function with a general form

1 ss
Gs 0\as \1 8:47
1 as ss

the optimal condition for the maximum phase lead is given by


   
1 1 p 1 1
tan /max p  as ; gain p sxopt p 8:48
2 as as as

The maximum phase lead is thus a function of as r=1 r only, and


determined automatically once the stiffness ratio r is given. The maximum possible
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 235

2.0 80 20

4 d
5 Af peak

1.5 60 15
2 max ,

( s )
1 max

1 max [ ]
4 d
1.0 40 10

peak
2 max

5 Af
3 ,
c
s

0.5 20 5

3 c / s
0 0 0
0.01 0 .0 5 0.1 0.5 1.0 5 10
= kd / ks

Fig. 8.28 Optimal condition of Jeffcott rotor

damping ratio fmax is thus dependent on the stiffness ratio r only. For example, if
r = 1, then as = 0.5,
 
1 1 1 p
fmax tan /max p  a s 0:18 8:49
2 4 as

which agrees with the peak of the approachable maximum damping ratio fmax in
Fig. 8.27. The corresponding critical speed r and the bearing damping cd (con-
verted to fd ) are given by:

ks kd 1 p Xc
X2c x2g p x2s as ! a1=4 0:84 8:50
mks kd as xs s

p
cd 1 1 1 r 1 r3=4
sxg xg p ! fd 0:84 8:51
kd as 2 xg =xs 2r1=4

Around fd = 0.8 in Fig. 8.27, we can see a corresponding peak value of fmax
0:18 in gure (a) and the critical speeds Xc xg = 0.84 in gure (b).
These optimization conditions using Eqs. (8.48)(8.51) are plotted by curves
noted by with relation to the stiffness ratio r in Fig. 8.28. This chart is
helpful for tuning for the optimal damping design of 1.5-dof models.

Table 8.1 Optimal conditions

s = /( + 1) max max c / s d
0.1 0.11 56 0.75 0.55 0.96
1 0.5 20 0.177 0.84 0.84
5 0.83 5 0.046 0.96 1.28
236 8 Rotor System Evaluation Using Open-Loop Characteristics

enlarged diagram
40 20 150
120 g
3
g
20 10 160
[dB]

3 140

[]
3
0 1 0 170
1, 2
3 160
g

20 2 10 180
1 2
180
40 1, 2 20 190
0.1 0.5 1 5 10 50 100 0 0.2 0.4 0.6 0.8 1.0
/ s / s

Fig. 8.29 Example of Jeffcott rotor (cross-stiffness effect)

Example 8.8 Determine the optimal conditions in the case of r = {0.1, 1, 5}.
Compare with the peaks in Fig. 8.27.
Answer
Figure 8.28 gives the conditions shown in Table 8.1. The values agree well with
the conditions for the peak damping ratios in Fig. 8.27.
Example 8.9 Figure 8.29 shows the open-loop frequency response for three
combinations between the shaft and bearing stiffness ratio r, given damping value
fd and cross-spring constant kc lc kd . The calculation is done by the following
open-loop transfer function:

ks kd 1  jlc cd s x2s 1 ss  jlc


Go s as 8:52
ms ks kd 1  jlc cd s s
2 2 1 as ss  jlc
Case studies:
(1) r = 1, fd = 0.1, lc 0: viscous damping only; stable
(2) r = 1, fd = 0.1, lc 0:4: cross-spring added, instability appears
(3) r = 1, fd = 0.886, lc 0:4: viscous damping tuned for the optimal stable
condition
Read the gain cross-over frequencies and phase margins from curves. Using
these reading values, predict the complex eigenvalue. Then calculate the exact
complex eigenvalues and compare the forward eigenvalue kf with the predicted
values ka , referring to Table 8.2.
Answer
The values read from the gure give an approximation of the forward complex
eigenvalue ka as shown in the upper part of Table 8.2. Note that no information on
the backward whirl eigenvalue is obtained because only the region x > 0 is excited
and shows the response. The complex eigenvalues k are given by solving the
characteristic equation 1 Go s 0 as listed in the lower part of Table. Both
agree well.
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 237

Table 8.2 Measured open-loop characteristic value and prediction of complex eigenvalue

case 1 damping only 2 cross-stiffness 3 optimal condition

measured value g 0.706 0.706 0.788


measured value m 4
measurement

7.14 18
= 0.5tan m 0.035 0.063 0.162
= g 0.025 0.044 0.128
predicted value a 0.025 + j 0.706 0.044 + j 0.706 0.128 + j 0.788 ( forward )

exact value f 0.025 + j 0.708 0.044 + j 0.716 0.151 + j 0.75 ( forward )


calculation

stable unstable stable


exact value b 0.025 j 0.708 0.088 j 0.734 0.166 j 0.872 ( backward )

8.3.6 Frequency Response

Although the optimal conditions are recommendations for a rotor to have a high
damping ratio (low Q-value) as mentioned above, the tuned flexible support
induces, in practice, a high susceptibility to external excitation. This trade-off sit-
uation is claried by examining two types of frequency responses, i.e. unbalance
and external force.
The unbalance vibration response of a model shown in Fig. 8.30 at a rotational
speed X is

me X2 eX=xs 2
Au  2 8:53
ms Gr s
2
sjX X 1 jsxs X=xs
 as
xs 1 as jsxs X=xs

Examples of calculation results are shown in Figs. 8.31 and 8.32.

Fig. 8.30 Unbalance j t


m 2 e
response model
z
m

1 + s
G r = ks s
1 + s s

unbalance vibration
238 8 Rotor System Evaluation Using Open-Loop Characteristics

Fig. 8.31 Unbalance 16


vibration 14 d = 0.1 =1

Au /
12
10

amplitude
8
6 d = 5
4 d = 0.8
2
0
0 0.5 1.0 1.5 2.0
speed / s

Fig. 8.32 Unbalance 25


vibration (comparison of

amplitude A u /
20
optimal conditions) = 10
15

10
=1
5 = 0.1 = 0.7

0
0 0.5 1.0 1.5 2.0
speed / s

The unbalance vibration responses in Fig. 8.31 are for r = 1 and fd = {0.1, 0.8, 5}.
The corresponding open-loop frequency responses are shown in Fig. 8.26, in
which fd = 0.8 gives the greatest phase lead. In Fig. 8.31, in fact, the resonance peak
at the critical speed is lowest for fd = 0.8 rather than others. Figure 8.32 shows
resonance curves for r = {0.1, 0.7, 1, 10} with the respective optimized fd = {0.96,
0.81, 0.84, 1.7} obtained from in Fig. 8.28. The resonance peak amplitude clearly
decreases as r decreases so that more and more flexible support is recommended.
For a system with external force excitation at a frequency x as shown in
Fig. 8.33a, the response is

f0 ds
Af  2 8:54
ms2 Gr s sjx x 1 jsxs x=xs
 as
xs 1 as jsxs x=xs

f0
where ds .
ks
The resonance peak is estimated by the product of the static deflection and
Q-value:

ds 1r 1
Af peak Q ds q 8:55
as r
2fmax 1  f2max
8.3 Open-Loop Frequency Response of a Jeffcott Rotor 239

(b) 25
j t
f0 e
20
(a) = 10

Af / s
z
m 15
= 0.1

amplitude
10
1 + s =1
G r = ks s
1 + s s = 0.7
5

0
0 0.5 1.0 1.5 2.0
Force excitation model excitation frequency / s

Fig. 8.33 Force excitation resonance curves (comparison of optimal conditions)

The plot of Eq. (8.55) in the dimensionless form with stiffness ratio r on the
abscissa and the maximum amplitude Af peak =ds on the ordinate is in Fig. 8.28.
The minimum in the curve indicates that a stiffness ratio r of approximately 0.7 and
the bearing damping fd 0:8 optimized with respect to the phase lead function is
the optimal design policy for a rotor which has a low Q-value and strong resistance
to force excitation.
Figure 8.33b shows the force excitation resonance curves for r = {0.1, 0.7, 1, 10}
with the respectively optimized bearing damping ratios fd . The large amplitudes in
the low frequency domain for small r reflects the fact that the flexible support is
susceptible to external forces. On the other hand, the high resonance peak for large r
appears around x=xs = 1, because a too strong spring support suppresses the
movement of damper and causes a lower damping ratio. The best result in this
example is obtained for r = 0.7. These results suggest that an appropriate bearing
stiffness is necessary to minimize each response to unbalance and external force
simultaneously.
Chapter 9
Bridge Between Inertial and Rotational
Coordinate Systems

Abstract This chapter discusses a bridge for the knowledge with respect to the
rotor-shaft vibration dened in an inertial coordinate system and the rotating
structure vibration formulated in a rotating coordinate system. The equations of
motion for rotor vibration discussed hitherto have been based on the description
concerning the absolute complex displacement z = x + jy measured in an inertial
(xed, stationary) coordinate system. This description is requested from a practical
viewpoint, because the vibrations measurement corresponds to displacement sen-
sors (or gap sensors, displacement meters) placed on a stationary part of machine.
Alternatively, this vibration can be measured by strain gauges xed at a rotational
coordinate system, as written by the displacement zr. These variables are mutually
related by: z zr ejXt (X = rotational speed) Therefore, if an eigenvalue is k in the
inertial coordinate system and kr in the rotational coordinate system, these entities
are mutually related by:

k kr jX

This chapter moves the viewpoint concerning vibration measurement from z to zr.

 
Keywords Inertial coordinate Rotating coordinate Gyroscopic effect Coriolis
 
effect Anisotropic stiffness Resonance condition

9.1 Vibration Waveforms (Displacement and Stress


Caused by Strain)

Rotordynamics involves the analysis of vibration of a rotating shaft with the natural
frequency and resonance characteristics determined with respect to the inertial
coordinate system. However, if the vibration of a rotor is measured with strain
gauges instead of displacement sensors, the measured values are thus transformed
to the rotational coordinate. As seen in Fig. 9.1, the relationship between

Springer Japan 2017 241


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_9
242 9 Bridge Between Inertial and Rotational Coordinate Systems

Fig. 9.1 Inertial coordinate Y z x jy


and rotational coordinate Yr zr xr j yr
systems

t
y

yr
Xr

xr t

x X

z = x + jy: vibrational displacement in the inertial coordinate system, measured


with displacement (gap) sensors, and
zr = xr + jyr: vibrational displacement in the rotational coordinate system,
measured with strain gauges, being proportional to shaft stress r and nally r / zr ,
is described as follows:

z = zr ejXt 9:1

The representations of unbalance vibration in the respective coordinate systems


are related to each other as follows:
Circular whirling:

z = AejXt ! zr A /r 9:2

Elliptical whirling:
 b ejXt ! zr = Af + A
z = Af ejXt A  b e2jXt /r 9:3

The steady-state unbalance vibration response of an isotropically (e.g. vertically)


supported rotor results in the circular whirling motion of Eq. (9.2). In this case, the
shaft is statically bent, and the stress is a constant and stationary value so that it
does not influence fatigue failure, hence the material strength is not signicant. The
steady-state unbalance vibration response of an anisotropically supported rotor (e.g.
a horizontal rotor supported by oil-lm bearings) is expressed by the elliptical
whirling motion of Eq. (9.3) as far as observed by displacement sensors. From the
viewpoint of the rotational coordinate system, the shaft experiences repeating
bending stresses with a frequency of 2X, which may cause fatigue failure. The shaft
stresses measured by strain gauges appear as backward wave propagation with a
frequency of 2X on the baseline of a static stress component.
Static deflection caused by gravity is represented in the respective coordinates as
9.1 Vibration Waveforms (Displacement and Stress Caused by Strain) 243

mg
z zg !zr zg ejXt / r 9:4
k

While being constant as measured by a displacement sensor, the signal measured


by a strain gauge is observed as backward propagation at the frequency X.
For self-excited whirling featured by the natural frequency xn , the variables
are related as

z = aejxn t !zr = aejxn Xt / r 9:5

This whirling means that repetitive bending stresses of the shaft occur, which
may result in fatigue failure. Self-excited vibration readily grows into a limit cycle
with large amplitude generating highly repetitive stresses. This must absolutely be
avoided. Prevention of self-excited vibration is thus a basic rule for safety design.

9.2 Natural Frequencies

Representations of the natural frequency in the two coordinate systems are related as

z aejxt ; zr aejxr t ; x xr X 9:6

where x is the natural frequency observed in the inertial coordinate system (x > 0
for forward whirling and x < 0 for backward whirling), while xr is observed in the
rotational coordinate system (xr > 0 for forward propagation and xr < 0 for
backward propagation). The natural frequency x of a single-dof rotor system with
the gyroscopic effect c observed in the inertial coordinate system is, as explained in
relation to Eqs. (6.21) and (6.22),
s
 
x cX 2 cX
 1 + 9:7
xn 2 xn 2 xn

The corresponding graphical representation is shown in Fig. 9.2. The whirling is


forward if x [ 0 and backward if x\0. xn is the natural frequency of the system
at rest. In the gure, exciting frequency lines of 2X, X and X are also indicated.
The natural frequency, xr , observed in the rotational coordinate system is
s
 
xr c X 2 c X
 1 1 9:8
xn 2 xn 2 xn

of which the graphical representation is shown in Fig. 9.3. The propagation is


forward if xr [ 0 and backward if xr \0.
244 9 Bridge Between Inertial and Rotational Coordinate Systems

Fig. 9.2 Inertial coordinate 2


system 3

2.0

1.5
g
2
1.0 0.5
0.25

/ n
1

natural frequency
0 g
2.0
1.0
0.5
1 0.25
0
g

2
0 1 2 3
p / n

Fig. 9.3 Rotational 2


coordinate system 0
c 0.5

1.0
/ n
r

0 0
natural frequency

1.5
1.7
2.

5
0

1
c
0
1.0
2
1.5
1.75
c 2 2.0
2
3
0 1 2 3
p / n

The natural frequencies in Fig. 9.3 are obtained from xr x  X, subtracting


the rotational speed from the natural frequencies shown in Fig. 9.2. Geometrically,
tilting the curves of Fig. 9.2 by 45 about (0, 1) in a clockwise sense gives the
natural frequency curves of Fig. 9.3.
9.2 Natural Frequencies 245

In a rotor with a thin disk, the natural frequency xn of the out-of-plane vibration
mode of the disk at rest (having a single nodal diameter) can be obtained from a
standard handbook. The natural frequency when rotating is obtained by substituting
c = 2 into Eq. (9.8):
s
 2
xr X
 1 ! x2r x2n X2 9:9
xn xn

The natural frequency in rotation xr is thus higher than that at rest. The dif-
ference, called the centrifugal effect, increases with X2 .

9.3 Resonance Conditions

Several resonance conditions taken into account with the gyroscopic factor are
indicated by red lled circles, blue circles and green squares in Figs. 9.2 and 9.3.
Recall that Fig. 9.2 represents the relationship between the natural frequency curves
and external excitation frequencies in the inertial coordinate system, while Fig. 9.3
represents that in the rotational coordinate system.
The natural frequencies observed in the respective systems are related by
Eq. (9.6). Unbalance force may generates resonances at red circles for +X and
blue circles for X, as shown in Eq. (9.3) and the force for secondary resonance
due to gravity and the anisotropy of rotor stiffness causes resonances at green
squares for +2X,as indicated in Eq. (11.10). These are considered in Fig. 9.2. The
straight line (X) are associated with anisotropic support stiffness as arising in a
slider bearing, which causes resonance even in the backward whirling as indicated
in Fig. 7.18. These symbols at the intersections between the natural frequency
curves and the external exciting frequencies indicate various resonance points.
The same analysis applies to Fig. 9.3 for the rotational coordinate system. The
external forces +X, X and 2X in Fig. 9.2 correspond to the abscissa (X = 0), 2X
and +X, respectively, in Fig. 9.3. The intersections of these lines and the natural
frequency curves indicate the resonance points in both gures. However, notice
must be taken of the fact that the inertial coordinate system (displacement sensor)
and the rotational coordinate system (strain gauge) give different results of fre-
quency analysis for the same vibration behavior.
The stress r observed by a strain gauge attached to the shaft or disk of a rotor
has waveforms of propagation shown in Fig. 9.4. The wave propagates forward if
the high spots (maxima and minima) of the waveform proceed to the direction of
rotation, and otherwise backward.
246 9 Bridge Between Inertial and Rotational Coordinate Systems

(b) (c)
(a)
7 7
2
3 1 6 6
5 5

0 4 4

4 3 3
2 2
7 1 1
5 0
6 t 0
t

stress waveform stress waveform


strain gauge forward waveform backward waveform

Fig. 9.4 Stress r propagation waveform measured by strain gauge

9.4 Representation of Equation of Motion

9.4.1 Gyroscopic Moment and Coriolis Force

This section discusses the effect of a coordinate transformation on the equation of


motion. The general form of the equation of motion for a symmetrically supported
single-dof system in the inertial coordinate system is

mz  jXcg m_z + kz d_z 0 9:10

where cg is the gyroscopic factor (0 < cg < 2). This is dictated by c in other
chapters, but cg is used only in this chapter for the purpose of comparing it with the
Coriolis factor cc, stated later. For a rotor mass, the vibrational displacement
z implies translational motion and a gyroscopic factor cg = 0. For a thin disk, the
vibrational displacement z implies tilting motion and a gyroscopic factor cg = 2.
The representation in the inertial coordinate system is transformed to that in the
rotational coordinate system by substituting

z zr ejXt
z_ z_ r ejXt jXzr ejXt 9:11
z zr e jXt
2jX_zr e jXt
 X zr e
2 jXt

into Eq. (9.10), resulting in

mzr jX2  cg m_zr k  mX2 mX2 cg zr d z_ r jXzr 0 9:12

The gyroscopic effect is called the Coriolis effect in the rotational coordinate
system. Equation (9.12) is rewritten using the Coriolis factor cc :
9.4 Representation of Equation of Motion 247

 
mzr jXcc m_zr k mX2 1  cc zr d z_ r jXzr 0 9:13

where the Coriolis factor cc ranges over 0 < cc < 2.


The gyroscopic factor cg and Coriolis factor cc are related to each other by the
following formula:

cg cc 2 9:14

The specic equations of motion for undamped systems are listed below:
(A) Inertial coordinate system:

mz kz 0 9:15

gyroscopic effect cg = 0.
Rotational coordinate system:

mzr 2jXm_zr k  mX2 zr 0 9:16

Coriolis effect cc = 2.
(B) Inertial coordinate system:

mz  2jXm_z kz 0 9:17

gyroscopic effect cg = 2.
Rotational coordinate system:

mzr k mX2 zr 0 9:18

Coriolis effect cc = 0.
Equation (9.15) means that translational motion of masses alone includes the
vibrational displacement z in the radial direction and the Coriolis factor is full with
cc = 2. On the other hand, Eq. (9.17) indicates tilting motion of a thin disk
exclusively, including the axial vibrational component. It has no vibrational dis-
placement in the radial direction, consequently there is no Coriolis force with
cc = 0. It is noted that the magnitudes of the gyroscopic and Coriolis effects are
reversed.
The assumption of an isotropic support for Eq. (9.10) is not merely for sim-
plicity: if the equation of motion contains a conjugate term such as kbz due to an
anisotropic support as seen in Eq. (7.2), the coefcients in the equation for the
rotational coordinate system obtained by the transformation Eq. (9.11) are no
longer constant, but time-dependent. In fact, kbz is modied by Eq. (9.11) and
kbzr e2jXt appears in Eq. (9.12). Conversely, the transformation between the rota-
tional and inertial coordinate systems is possible for isotropic supports only.
248 9 Bridge Between Inertial and Rotational Coordinate Systems

(c)
(a)
(b)

u
u

Fig. 9.5 Coriolis effect

Example 9.1 Give approximate evaluations of Coriolis effects in the rotating


structures shown in Fig. 9.5:
(a) Axial blades, e.g. in a turbine
(b) Cylindrical blades, e.g. in a sirocco fan
(c) Impeller blades, e.g. in a mixed flow pump

Answer
(a) The blades vibrate in out-of-plane mode only with no vibrational displacement
in the radial direction: no Coriolis effect.
(b) The cylindrical blades vibrate mainly in the radial direction: large Coriolis
effect.
(c) The vibrational displacement of the blades includes a certain radial component:
slight Coriolis effect.

9.4.2 Case Study: Multi-blade Fan (Sirocco Fan)


[81, VB55]

Hagiwara reports a vibration problem experienced in a multi-blade fan blower


shown in Fig. 9.6, in which
the fan is belt-driven by an electric motor (rated speed: 15 rps (=900 rpm)),
this type is called a sirocco fan (Fig. 9.7), and
the fan shaft is supported by ball bearings at both ends.
By the inspection after the vibration was noticed, many cracks were found in the
core plate of the blade wheel.
Vibration measurements during resonance gave the following ndings:
9.4 Representation of Equation of Motion 249

Fig. 9.6 Multi blade fan


belt core plate
1,420
blade

subplate

400
315
560 discharge

934 suction slip ring

motor
core plate thickness = 3.2 mm blade thickness = 2.3 mm
subplate number = 56 blade number = 56

Fig. 9.7 Fan model discharge


I p , Id

core plate
suction k tilting
stiffness

m
anisotropy
center of
gravity
G
amplitude
x y
cos2 t

discharge

(1) An accelerometer placed on the stationary side detected the predominant


frequency of twice the rotational speed.
(2) A strain gauge was attached to the core plate of the fan, of which the output
signal was transmitted to the stationary side via a slip ring. This stress
vibration detected by this strain gauge was assessed by an order analysis and a
1x vibration synchronized with the rotational speed was dominant as shown in
Fig. 9.8.
In order to nd the causes of the vibration, the natural frequencies were iden-
tied by a hammer test, where the core plate was impacted at rest and during
rotation. Frequency analysis of the vibrational stress yielded the results shown in
Fig. 9.9:
250 9 Bridge Between Inertial and Rotational Coordinate Systems

Fig. 9.8 Resonace in strain


20

vibration stress [10MPa]


gauge single
1Rotational frequency
16
component
12

0
0 300 500 700 900
speed [rpm]

Fig. 9.9 FFT of impulse (a)


response 0 rpm
Hz
FFT amplitude

15.6

94.4
(b) 68.6
200 rpm 200 rpm
FFT amplitude

13.2

Hz
18.0

71.0
69.0
37.2

94.6

(c) 400 rpm 400 rpm


11.2

Hz
FFT amplitude

20.6

37.2

95.8
66.8
73.4

frequency [Hz] 100


9.4 Representation of Equation of Motion 251

Fig. 9.10 Campbell diagram 24


of nodal diameter j = 1

3X
nodal diameter = 1
20
( backward wave )

natural frequency [Hz]


16

1X
12

nodal diameter = 1
( forward wave )
8
resonance experiment
calculation
4
1X vibration
(Fig.9 8)
0
0 4 8 12 16 20
rotational speed [Hz]

(a) The FFT spectrum at rest: peaks at {15.6, 68.6, 94.4} Hz


(b) The FFT spectrum at 200 rpm: peaks at {13.2, 18.0, 37.2, 69.0, 71.0, 94.6} Hz
(c) The FFT spectrum at 400 rpm: peaks at {11.2, 20.6, 37.2, 66.8, 73.4, 95.8} Hz
The low-frequency components among these measured frequencies are summa-
rized in relation to the rotational speed, as indicated by small circles in Fig. 9.10.
With some additional investigations taken into account, the resonance mecha-
nism was identied and solved as follows:
(1) It was found that the natural frequencies xh in the tilting mode (the node diameter
j = 1) were not uniform, when the impulse test was undertaken by hitting the
core plate along the circumferential direction. They revealed somewhat different
natural frequencies of the core plate in the X and Y directions, i.e., xhx 6 xhy ,
indicating anisotropy in the tilting stiffness of the core plate.
(2) This anisotropic stiffness causes the slight fluctuation (twice the rotational
frequency: 2X) due to vertical reciprocation of the center of gravity of the fan
during rotation, as detailed in Sect. 11.2. The acceleration of the fluctuation
acts as external excitation:

z cos 2Xt !
9:19
z / cos 2Xt / e2jXt e2jXt

or, the following frequency components are observed in the rotational


coordinate system,

zejXt / e jXt e3jXt 9:20


252 9 Bridge Between Inertial and Rotational Coordinate Systems

Fig. 9.11 Waveform of forward wave


strain gauge

2 3
5

1 4
4
0 5
strain gauge number
3

0
NO.

vib. stress propagation ( near resonance ) ( 570 rpm )

(3) The radial lines for the +X (solid) and 3X (broken) excitations are drawn in
Fig. 9.10, which were dened with respect to the rotational coordinate system.
(4) The intersection of the radial line with the natural frequency curve (indicated
by a large circle) corresponds to the resonance.
Natural frequency ! forward propagation
External force ! forward +1X = +X
(5) The peak amplitude measured with the strain gauge (Fig. 9.8) corresponds to
the above intersection seen in the resonance of Fig. 9.10. The intersection
indicates the resonance of the 1X forward propagation. Strain gauges attached
to the core plate in the circumferential direction permitted observation of the
forward propagation waveforms as shown in Fig. 9.11.
(6) The +X resonance described above has been observed in the rotational
coordinate system. It corresponds to +2X in the inertial coordinate system and
the resonance would be observed as the 2X forward whirling if the fan
vibration were measured by an displacement sensor placed at the stationary
side.
(7) For the countermeasure, it was recommended to increase the thickness of the
core plate to avoid a resonance frequency in the operational range of the
rotational speed, or to machine precisely the core plate for a more uniform
thickness and to compensate for the stiffness anisotropy in order to approach
xhx xhy .
Chapter 10
Vibration Analysis of Blade and Impeller
Systems

Abstract This chapter discusses vibrations of rotating structures such as blades in


turbines and impellers in pumps or compressors. The natural frequencies of a
rotating structure may be analyzed using the 3-D nite element method and clas-
sied by the number of the nodal diameters or circular nodal modes. These results
are represented in the rotational coordinate system. The difference between the
inertial coordinate system xed to the stationary side and the rotational coordinate
system xed with the rotor must be taken into account in analysis of:
(1) resonance caused by any static load distributed in the circumference direction
of the stationary side facing blades or impellers, and
(2) resonance caused by harmonic excitation at a certain point in the stationary side
facing blades or impellers.

Keywords Rotating structure 


Blading Cyclic matrix  Eigenvalue 
 
Eigenmode Resonance condition Nodal diameter

10.1 Natural Frequencies of Rotating Structure Systems

10.1.1 Natural Frequencies of a Thin Disk [9]

The natural circular frequencies xn of vibration of a thin disk at rest, under the
boundary conditions of xed inner circumference and free outer circumference as
shown in Table 10.1, are given by
s
h E
xn k 2 2
10:1
a 121  m2 q

The graphical representation of this equation shown in Fig. 10.1 indicates that
the coefcient k is largely dependent on the radius ratio b/a.

Springer Japan 2017 253


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_10
254 10 Vibration Analysis of Blade and Impeller Systems

Table 10.1 k of disk xed at inner cylinder (Poissons ration m = 0.3) [9]
nodal nodal nodal nodal
0 1 2 3
diameter diameter diameter diameter
2a
b/ a b/ a b/ a b/ a
2b
0.276 2.50 0.060 1.68 0.186 2.50 0.43 4.0
0.642 5.00 0.397 3.00 0.349 3.00 0.59 5.0
0.840 9.00 0.603 4.60 0.522 4.00 0.71 7.0
0.634 5.00 0.769 8.00 0.82 10.0
0.771 8.00 0.81 10.00
h
0.827 11.00

Fig. 10.1 k of disk (inner 12


side xedoutside free)
10
=0
8
=3
6

=2
4
=1
2

0
0 0.2 0.4 0.6 0.8 1.0
b/a

While Eq. (10.1) is concerned with the disk at rest, the natural frequencies xr
during rotation at a speed X are obtained by a correction using the Southwells
equation:
s
 2
xr X
1C 10:2
xn xn

3m 1 3m
where C j 2sj 2 2s  j2 (approximated from Table 10.3),
8 8
j number of nodal diameters,
s = number of circular nodes
The correction factor, C, for the rotation effect (increasing tension due to cen-
trifugal force) is called the centrifugal coefcient. For example, C = 1 for an eigen
mode with a single nodal diameter (j 1, s = 0). The natural frequency xr of
various modes observed in the rotational coordinate system increases with the
rotational speed, as shown in Fig. 10.2, which is due to the variation induced by the
centrifugal coefcient.
10.1 Natural Frequencies of Rotating Structure Systems 255

Fig. 10.2 Centrifugal effect 8


( , s) = (4,0) (1,1)
(disk)

n a t u r a l fr e q u e n c y r / n
7
(3,0)
6
= nodal diameter (2,0)
5
s = nodal circle (1,0)
4
(0,1)
3
2
(0,0)
1
0
0.5 1 1.5 2 2.5 3
speed p = / n

Note: Natural frequencies for a single nodal diameter. The natural frequency
with (j 1, s = 0) is the out-of-plane mode vibration of the thin disk as observed in
the rotational coordinate system, and it could be re-evaluated in the inertial coordinate
system as seen by the corresponding tilting mode of the disk xed on the shaft.
Hence, the effect of rotation for the tilting mode vibration is expressed by the gy-
roscopic factor c, which is 2 for the thin disk. Consequently, the forward and back-
ward natural frequencies in the inertial coordinate system are taken from Eqs. (6.21)
and (6.22):

  s
 2
xf xb X X
 1 10:3
xn xn xn xn

As explained in relation to (Eq. 9.9), subtracting the rotational speed from


Eq. (10.3) gives the natural frequencies of forward propagation and backward
propagation in the rotational coordinate system:

  s
 2
xr xf xb X X
  1 10:4
xn xn xn xn xn

Comparison with Eq. (10.2) makes it clear that the centrifugal coefcient C = 1
is justied for j s 1 0. For a thin disk rotor, therefore, the gyroscopic
factor c 2 in the inertial coordinate system and the centrifugal coefcient C = 1 in
the rotational coordinate system corresponds to the same physical phenomenon.
Example 10.1 Find the natural frequencies of a mode having a single nodal
diameter concerning a 14-inch disk of an HDD (Hard Disk Drive) at rest X 0
and during the rotation at the rated speed (X 60 rps). The specications of the
disk are:
b = 84 mm, a = 178 mm, t = 1.91 mm
made of aluminum (q 2; 670 kg /m3, E = 68.6 GPa, (Poissons ratio m 0:33)
256 10 Vibration Analysis of Blade and Impeller Systems

Compare the results with the experimental curve for the rst natural frequency
x1 in Fig. 6.27.
Answer
For b/a = 84/178 = 0.47, Table 10.1 gives k 3:35.
Natural frequency at rest:
s
1 1:91 68:6  109
xn 3:352 167 Hz
2p 178  103
2 121  0:332 2670
A displacement sensor xed with the stationary side observes this value as the
tilting natural frequency of the disk. At the rated rotational speed X 60 rps, the
gyroscopic factor c 2 is applied, and using Eqs. (6.21) and (6.22):
Forward natural frequency:
s
 
60 2
xf 167 1 60 237 Hz
167

Backward natural frequency:


s
 
60 2
xb 167 1 60 117 Hz
167

In fact, as the rotational speed increases, the natural frequency curves in Fig. 6.27
change from 167 Hz. When the rotational speed reaches X 60 rps, the solid line
increases to 237 Hz and the broken line decreases to 117 Hz. These natural fre-
quencies can be transformed to the rotational coordinate system by Eq. (9.6):

xf  X 237  60 177
xr Hz X 60 rps
xb  X 117  60 177

These are the natural frequencies of the out-of-plane vibration of the disk
observed by a strain gauge.
This value can also be obtained using the formula of Eq. (10.2) with the cen-
trifugal coefcient C = 1:
s
 
60 2
xr 167 1 177 Hz X 60 rps
167

It is thus understood that the centrifugal coefcient C = 1 (in the rotational


coordinate system) and the gyroscopic factor c 2 (in the inertial coordinate
system) describe the same phenomenon for the mode having a single nodal
diameter.
10.1 Natural Frequencies of Rotating Structure Systems 257

500

400
=3 =2

r [Hz]
300
=0
200
=1
100

0
0 50 100 150 200
speed [Hz]

Fig. 10.3 Natural frequency of 14 inch HDD

Table 10.2 l of disk xed at center [9]


2a
nodal
fixed disk at center diameter 0 1 2 3
circular
node
0 5.16 1.05 10.61 56.87
h
1 160.2 154.3 443.9 1 040
h E 2 1 349 1 222 2 624 4 285
n =
2 a2
3 5 250 5 266 8 574 13 287
[rad/s]

Example 10.2 Draw the natural frequency curves versus rotational speed for
modes with one or more nodal diameters in the 14-inch disk of the previous
example.
Answer
Figure 10.3. These are the natural frequencies in the rotational coordinate system as
would be measured by a strain gauge. The increase of the frequency reflects the
centrifugal effect.
The natural frequencies xn of a thin disk xed at the center are calculated in
Table 10.2 for a condition at rest, and the centrifugal coefcients are estimated in
Table 10.3 for non-zero rotation.

10.1.2 Natural Frequencies of Blades [9]

The natural frequencies of a blade consisting of a uniform plate as shown in


Fig. 10.4 are given by the formula for a cantilever at rest as stated previously in
Table 3.1:
258 10 Vibration Analysis of Blade and Impeller Systems

Table 10.3 Centrifugal coefcient C of disk xed at center [9]


rotating disk nodal
fixed at center diameter
circular 0 1 2 3
node
0 0< C <0.5 1 2.35 4.05
1 3.3 5.95 8.95 12.3
2 9.0 14.2 18.85 28.85
r
= 1+ C ( )
2
n n 3 19.8 25.75 32.05 38.7

Fig. 10.4 Rotating cantilever


rad/s

vibration
direction

2R
l
l

s
k2 EI
xn 2 10:5
l qA

where the rst three coefcients are k f1:875; 4:694; 7:855g.


The natural frequencies of a rotating plate are determined by Southwells
equation with appropriate centrifugal coefcients C:
s
 2
xr X
1C 10:6
xn xn

where
C = 1 + 1.45R/l (R/l = 0 10) for the rst mode,
C = 6.06 + 7.64R/l (R/l = 0 3.92) for the second mode, and
C = 15.03 + 18.91R/l (R/l = 0 1.11) for the third mode
The centrifugal coefcients C for the correction are presented graphically in
Fig. 10.5.
When the blade conguration is pre-twisted with respect to the axial direction,
the stagger angle h is dened as shown in Fig. 10.6. In this case, the centrifugal
coefcient C* is given by
10.1 Natural Frequencies of Rotating Structure Systems 259

Fig. 10.5 Centrifugal effect 8

natural frequency of only


(blade) 7 1
l =
0 R/

1
=
6 =0

l=
blade r /n

R/
5 e
od od
e
m
4 3r
d dm 1
R/l = 0
2n =
3 mode
1st
2
1
0
0.5 1 1.5 2 2.5 3
speed p = / n

Fig. 10.6 Blade with stagger


angle [9]
rad/s

r
= 1 + C
2
vibration n n
direction l C = C cos 2
2R

l
= Stagger angle

C  C  cos2 h 10:7

For the blade of Fig. 10.4, we take h 90 .

Note: Centrifugal coefcients Let us note again that the natural frequencies
during rotation are determined by the centrifugal coefcient C given by Eq. (10.2)
or Eq. (10.6), which would be the value dened in the rotational coordinate system
and measured by a strain gauge.

10.1.3 Vibration Analysis of Cyclic Symmetry Structural


Systems

Many rotating structural systems, including blades and impellers, are symmetric
with respect to the center of rotation and have cyclic structures as shown in
Fig. 10.7. The 3-D nite element analysis for rotating structures takes the cyclic
symmetry conditions into account [82, 83]. This section examines the eigenvalue
problem from the simplied viewpoint of nite element analysis.
Fundamentals of the vibration analysis of a cyclic symmetric structure are
described here using a spring-mass system at rest shown in Fig. 10.8. This model has
four masses each representing a blade (the total number of blades N = 4), the
vibrational displacement of which is in the circumference direction is denoted by
hi i 03. The equation of motion for the entire system involves the
260 10 Vibration Analysis of Blade and Impeller Systems

period
N=6
period
N=6

Fig. 10.7 Cyclic structure

Fig. 10.8 4-blade model 0


m
0
kc

3
kd m
1 3
m
1

2
m
2

circumferential displacement hi of each mass, circumferential coupling spring con-


stant kc , and the radial spring constant kd between the system center and each mass:

M h Kh 0 10:8

where

h h0 h1 h2 h3 t
2 3
m 0 0 0
60 m 0 07
6 7
M6 7
40 0 m 05
0 0 0 m
2 3
kd 2kc kc 0 kc
6 kc kd 2kc kc 7
6 0 7
K6 7
4 0 kc kd 2kc kc 5
kc 0 kc kd 2kc
10.1 Natural Frequencies of Rotating Structure Systems 261

In consideration of this example more generally, the mass matrix M and stiffness
matrix K of a cyclic structure are re-dened as more detailed cyclic matrices:
2 3 2 3
m0 m1 m2 m3 k0 k1 k2 k3
6 m3 m0 m1 m2 7 6 k3 k0 k1 k2 7
M6
4 m2
7; K6 7 10:9
m3 m0 m1 5 4 k2 k3 k0 k1 5
m1 m2 m3 m0 k1 k2 k3 k0

The example system of Fig. 10.8 is a special case in which


m0 = m, m1 = m2 = m3 = 0,
k0 = kd + 2kc, k1 = kc, k2 = 0, k3 = k1
In the cyclic matrix, the elements of the second row are the same as those of the
rst row but offset by one element: the last element of the rst row becomes the rst
in the second row. The same applies to the third and further rows throughout the
matrix.
In general, the equation of motion for a cyclic structure consists of cyclic
matrices. It is assumed hereinafter that mi and ki are matrices having a certain
dimensional length, being equal to the number of freedoms of a block in the cyclic
structure. Assuming the vibrational solution h /ejxt of the equation of motion,
the eigenvalue equation is obtained:

K/ x2 M/ 10:10

The / vectors are now converted to y vectors by the following coordinate


transformation W:

/ Wy 10:11

where
2 3 2 3 2 3
e00 e01 e02 e03 1 1 1 1 y0
6 e1 e11 e12 17 6 1 j 7 6y 7
6 e3 7 6 1 j 7 6 17
W ~
e0 ~
e1 ~
e2 ~
e3  6 02 7 6 7; y 6 7;
4 e0 e21 e22 e23 5 4 1 1 1 1 5 4 y2 5
e30 e31 e32 e33 1 j 1 j y3
2p 2p
ek ej k ej k cos k90 j sin k90 k 03:
N 4

M and K are symmetric cyclic matrices. W is called the mode matrix with
columns equal to the eigenmodes~ e0 . . .~
e3 . Each element in the matrix corresponds to
the circumferential point on the unit circle in the complex plane, the real part of
which is the actual displacement. W is symmetric. Note that the coefcient matrices
in Eq. (10.10) can be block-diagonalized by using a unitary-like transformation as
follows:
262 10 Vibration Analysis of Blade and Impeller Systems

W 
 t MW WMW M~
  10:12
~ diagonal m
M ~0 ~1
m ~2
m ~ 3  is block-diagonal and real
m

 t KW WKW
W  K~

 10:13
~
K diagonal ~k0 ~k1 ~k2 ~k3 is block-diagonal and real

In this manner, the eigenvalue problem of the full-order M-K system of


Eq. (10.10) for the entire cyclic structure is replaced by a set of N independent
eigenvalue problems of smaller order, corresponding to the m~ i  ~ki system of each
block matrix structure. In this case, the 4-dof eigenvalue problem is eventually
reduced to a set of 4 single-dof eigenvalue problems:

~ i yi ~ki yi
x2 m i 03 10:14

Since N = 4, m1 = m2 = m3 = k2 = 0 and k1 = k3 in the example case, it follows


that

X
N 1
~k N
m elk ml 4 m
l0
X
N 1 X
3
2p
~kk N elk kl 4 ej klkl 4k0 2k1 cos k90
l0 l0
4

Therefore the N = 4 set include eigenvalues obtained depending upon the


number of nodal diameter, i.e.:

x20 ~k0 =m
~0 kd =m; /0 1 1 1 1 t for j 0;
~ ~1
x 1 k 1 =m
2
kd 2kc =m; /1 1 0 1 0 t for j 1
~ ~2
x 2 k 2 =m
2
kd 4kc =m; /2 1 1 1 1 t for j 2 and
x2 ~k3 =m
3 ~3 ~k1 =m
~ 1 x21 ; /3 0 1 0 1 t for j 1:
10:15

Note that /3 is the same mode group as /1 , because of having j 1.


The number of the eigenvalues is mathematically equal to N, or the number of
degrees of freedom. However, only the eigenvalues corresponding to nodal diameters
numbering up to N/2 (for even N) or (N 1)/2 (for odd N) are unique for the solution of
this vibration problem. The corresponding mode shapes are shown in Fig. 10.9.
Each gure for j 0; 1 and 3, and 2 is the number of nodal diameters are 0, 1 and 2,
respectively. In conclusion, the vibration characteristics of the entire structure may be
obtained from the mass and stiffness matrices of each block with reduced order.
Example 10.3 Determine the vibration characteristics of a cyclic structure of N = 8
(Fig. 10.10) assuming that the mass m = 1. Hence, show that the mass matrix is:
10.1 Natural Frequencies of Rotating Structure Systems 263

=0 = 1, 3 =2
2 2 2
0 = kd /m 1 = k d + 2 k c /m 2 = kd + 4 kc /m
node node

node node

node node

Fig. 10.9 Natural frequencies and eigenmodes

Fig. 10.10 8-blade model 0


kc
m 7

1
m m
kd

6
m m
2

m m
5

3 m

M8 mE8 E8 is the 8th-order unit matrix 10:16

10:17

where kd kd 2kc .


Answer
Eigenvalues are obtained by solving x2n M8 / K8 / to give the following vibration
characteristics. The corresponding eigen modes are shown in Fig. 10.11.
264 10 Vibration Analysis of Blade and Impeller Systems

=0 = 1, 7 = 2, 6

0 0 0

1 7 1 7 1 7

2 6 2 6 2 6

3 5 3 5 3 5

4 4 4

= 3, 5 =4

0 0

1 7 1 7

2 6 2 6

3 5 3 5

4 4

Fig. 10.11 Eigenmodes

Number of nodal diameters j 0:

x20 ~k0 =m
~ 0 kd /0 f 1 1 1 1 1 1 1 1 gt

Number of nodal diameters j 1:


 
p 1 1 1 1
x21;7 ~k1 =m
~ 1 kd 2  2 kc /1 1 p 0 p 1 p 0 p
2 2 2 2

Number of nodal diameters j 2:

x22;6 ~k2 =m
~ 2 kd 2kc /2 f 1 0 1 0 1 0 1 0 gt

Number of nodal diameters j 3:


 t
p 1 1 1 1
x23;5 ~k3 =m
~ 3 kd 2 2 kc /3 1 p 0 p 1 p 0 p
2 2 2 2

Number of nodal diameters j 4:


10.1 Natural Frequencies of Rotating Structure Systems 265

x24 ~k4 =m
~ 4 kd 4kc /4 f 1 1 1 1 1 1 1 1 gt

Example 10.4 Solve the eigenvalue problem for the system of Fig. 10.10 regarded
as consisting of a four-block cyclic structure (N = 4), each containing two blades.
Conrm that the results are the same as in Example 10.3.
Answer
The mass matrix of Eq. (10.16) and stiffness matrix of Eq. (10.17) are regarded as
consisting of four blocks as seen in Eq. (10.9) by reformulation:

1 0
m0 m ; m1 m2 m3 0; 10:18
0 1
 
kd 2kc kc 0 0
k0 ; k1 ; k2 0; k3 k1t 10:19
kc kd 2kc kc 0

Transformation dened by 4  4 matrix W of Eq. (10.11) gives the set of 4


block matrices of Eqs. (10.1210.13), where each block consists of 2  2 matrices.
~ i yi ~ki yi has 2 eigensolutions, which are:
Each set x2i m

~i
m 1 0
m i 03
4 0 1
~k0 
kd 2kc 2kc kd kd 4kc
k0 k1 k3 ) x20 ; x24
4 2kc kd 2kc m m

~ 0  ~k0 system
for m

~k1 
kd 2kc amp; 1 jkc
k0 jk1  k3  ) x21
4 1  jkc amp; kd 2kc
 p
kd 2  2 kc 2
; x5 x23
m

~ 1  ~k1 system
for m

~k2 
k 2kc 0 kd 2kc 2
k0  k1  k3 d ) x22 ; x6 x22
4 0 kd 2kc m

~ 2  ~k2 system
for m
266 10 Vibration Analysis of Blade and Impeller Systems

~k3 
kd 2kc 1  jkc
k0  j k1  k3  ) x23
4 1 jkc kd 2kc
 p
kd 2 2 kc
; x27 x21
m

~ 3  ~k3 system.
for m
Solving the eigenvalue problem for these 2  2 block matrices yields the
eigenvalues indicated by arrows ) in the above. They are identical with those
obtained in Example 10.3. It is also possible to obtain the eigenvectors. For example,
eigenvalues:

x20 ; x21 ; x22 ; x23 ; x24 10:20

~ i  ~ki block system:


eigenvectors for each m
    
1 1 1 1 1
y0 ; y1 j45 ; y2 ; y3  ; y4 10:21
1 e 0 ej45 1

Converting these eigenvectors to the physical coordinates and picking up the real
parts yields the actual eigenmodes as the same as Example 10.3 mentioned
previously:
82 32 39 82 39
>
>
1 1 1 1 y0 0 0 0 y4 >
> >
>
y0 y1 y2 y3 y4 >
>
>6
< 76 0 7 >
= >
<6 y >
=
6 1 j 1 j 76 y1 0 0 07 6 0 jy1 y2 jy3 y4 7
7
Re 6 76 7 Re 6 7
>
> 4 1 1 1 1 54 0 0 y2 0 0 5>> >
>4 y0 y1 y2 y3 y4 5 >
>
>
: >
; >
: >
;
1 j 1 j 0 0 0 y3 0 y0 jy1 y2 jy3 y4
! /0 /1 /2 /3 /4 
10:22

10.1.4 General Vibration Analysis of Blades and Impellers


in a Rotational Coordinate System

Vibration analysis of rotating structures may be considered for


multi-degrees-of-freedom for deflection, twist and tilt along and around X-Y-Z axes
of each node as seen in Fig. 10.12. Taking the continuously varying stagger angle
and the centrifugal coefcient due to rotation into account, 3-D nite element
analysis may be utilized on the basis of the equation of motion dened in the
rotational coordinate system [84]:
10.1 Natural Frequencies of Rotating Structure Systems 267

Fig. 10.12 Blades

Y
Z

 
Mx 2XMc x_ K KE  X2 ME x f 10:23

where
x displacement vector (deflection, tilting motion, axial motion twist angle, etc.)
of a node
M mass matrix
K stiffness matrix
Mc Coriolis matrix characterizing the Coriolis force
KE geometric stiffness matrix characterizing the centrifugal effect
ME geometric mass matrix characterizing the centrifugal effect
f equivalent external force vector acting on each node
Since this is a cyclically symmetry structure, the matrices are cyclic and can be
block-diagonalized by the coordinate transformation of Eq. (10.11), eventually
leading to
 
M ~ ci y_ i K
~ iyi 2XM ~i K ~ Ei yi ~fi
~ Ei  X2 M i 0
N  1 10:24

The vibration analysis for a cyclic symmetric structure, such as with blades and
impellers, can thus be reduced to that of a designated block system. The eigenvalue
problem gives undamped natural frequencies as observed in the rotational coordi-
nate system. This type of analysis can be performed using a commercially available
software package, e.g., ANSYS, NASTRAN, etc.
268 10 Vibration Analysis of Blade and Impeller Systems

10.2 Vibration and Resonance of Blades and Impellers

10.2.1 Conditions for Blade-Shaft Coupled Vibration

In rotating shaft-bearing system design, blades and impellers are usually regarded
as forming a rigid body and modeled as a rotating disk; the elasticity is neglected.
On the other hand, in designing blades or impellers, the boss (center) of the rotating
structure, which is connected with the shaft, is assumed as a xed part; the elastic
vibration is calculated by Eq. (10.24) neglecting the shaft movement.
However, for eigenmodes with zero or one nodal diameter of a rotating structure,
they can be coupled with the thrust (axial), torsional or flexural vibration of the
shaft. The possibilities of such coupling effect are shown in Table 10.4.
As illustrated in the gure:
(1) The torsional vibration of the shaft causes circumferential vibration of the
blades with a circumferentially uniform mode with j 0.
(2) The thrust vibration of the shaft causes vibration of the blades in an
umbrella-like mode with j 0.
(3) The horizontal translational vibration of the shaft causes in-plane vibration of
the blades with one nodal diameter, j 1:
(4) The tilting vibration of the shaft causes out-of-plane vibration of the blades
with one nodal diameter, j 1:
The causality may be reversed: vibration of the blades in one of the modes
mentioned above causes vibration of the shaft in the corresponding mode.
Since the eigenmodes with j 0 and 1 of the blade and impeller system can
thus be coupled with the shaft vibration, analysis of the entire system is ultimately
needed. As for the modes with j 2, the analysis of the rotating structure is only
sufcient, because the modes are not coupled with the shaft vibration.

Table 10.4 Shaft-blade coupling

blade / disk vibration


rotor ( = nodal diameter )
vibration
=0 =1 =2
(1) blade: in-plane (3) blade: in-plane
translating (3) shaft: torsion shaft: translation
bending
tilting (4)

torsion (1)
axial / thrust (2) (2) blade: out-of-plane (4) blade: out-of-plane
shaft: thrust shaft: tilt
10.2 Vibration and Resonance of Blades and Impellers 269

node
2 2 2 node
3 1 3 1 3 1
+ r r r
amplitude 4 + + 0 4 + 0 4 + + 0
+ blade blade
5 7 5 7 5 7
blade
6 6 6 node

schematic
illustration + =0 + =1 + + =2

Fig. 10.13 Eigenmodes of blade (rotational coordinate)

10.2.2 Natural Vibration Modes of Blades and Blade


Wheels

Consider eigenmodes with j-th nodal diameters of blade vibration. Figure 10.13
shows examples for N = 8 and modes with j = 0, 1 and 2. The circumferential or
axial vibration component of the blade is considered here. With the phase hr in the
rotational coordinate system, the eigenmodes are

/hr cos jhr j 0; 1; 2; . . .; N=2 10:25

For the natural frequency xr [ 0, the out-of-plane vibration displacement d at


each phase of the blade is

dhr ; t cos jhr cos xr t 10:26

These are standing waves formed by both interference of the forward and the
backward waves. They depend on the distribution of the external forces as to which
wave occurs at the resonance.
Some eigenmodes for nodal diameter j 2 are shown in the inner part of
Fig. 10.14, with the mode notation +//+/ on 4 blades.

10.2.3 External Forces Acting on Blades and Impellers

The outer curves of Fig. 10.14 illustrate the static pressure distribution generated in
the stationary vanes due to non-uniformity of flows. Consider an exciting force
p that produces a static pressure distribution of the J-th harmonic function along the
circumferential direction. The gure shows examples for J = 2 and J = 6. J is often
270 10 Vibration Analysis of Blade and Impeller Systems

(a) (b) vibration mode


Y Y ( rotating coordinates )
vibration mode
node ( rotating coordinates ) node
node + +

2 Yr 2 Yr
1 1
3 r 3 r
0 X 0 X
+ ++ + + + ++ + +
4 Xr 4 Xr
blade
7 blade 7

5 5
6 6

pressure distribution + +
node ( static state ) pressure distribution
( static state )
J=2 J=6

Fig. 10.14 Resonance condition (inside blade mode, outside static pressure distribution)

the number of stationary vanes. For the phase h in the inertial coordinate system, the
excitation p is given by

ph cos Jh 10:27

In the rotational coordinate system, the spatial mode of the pressure distribution
in the stationary side rotates clockwise (opposite to the rotation), as shown in the
gure by the arrows X. The exciting force in the rotational coordinate system is
obtained by substituting h hr Xt into Eq. (10.27):

phr cos J hr Xt cosJhr JXt 10:28

10.2.4 Resonance Conditions of Blades

First, the resonance condition of a uniform shaft is reviewed in the familiar example
of unbalance vibration. In a simple model shown in Fig. 10.15, assuming that the
natural frequency is xk for mode /k , the k-th modal vibration d is:

dn; t /k n cos xk t sin kpn cos xk t k 1; 2; 3; . . . 10:29

The unbalance force F acting on the system in a uniform distribution p can be


written as
10.2 Vibration and Resonance of Blades and Impellers 271

U U U

1 ( ) 2 ( ) 3 ( )
modal exciting force 0 modal exciting force = 0 modal exciting force 0

Fig. 10.15 Resonance of rotor (unbalance and eigenmodes)

Fn; t pnejXt UX2 ejXt 10:30

The resonance conditions for unbalance vibration are:


[I] The frequency of external force = natural frequency: X xk
[II] The modal exciting force = the inner product between the eigen mode and
the distribution of external force 6 0:
Z Z
1 1
UX2
p/k dn UX2 sin kpndn f2 0 2=3 0 ...g
0 0 p 10:31
k 1; 2; 3; . . .

Thus, resonance occurs when the rotational speed coincides with the natural
frequencies in the modes of odd orders, but not in the even orders, as shown in the
lower part of Fig. 10.15.
The resonance conditions of blades in the rotational coordinate system can be
treated similarly. Comparison between the modal vibration of Eq. (10.26) with
natural frequency xr and the exciting force of Eq. (10.28) yields the rst resonance
condition with respect to frequencies:

I JX xr 10:32
The second resonance condition with respect to the inner product requires that
the integral of the modal exciting force throughout blades = (eigen mode) 
(distributed exciting force):

/hr phr cos jhr cosJhr JXt 1=2 cosJ jhr JXt
II
1=2 cosJ  jhr  JXt
10:33
which is estimated by integration in the peripheral direction to be non-zero.
Since the phases corresponding to the blades are discrete, the integral is, in practice,
the sum of the values of Eq. (10.33) on each blade. For N = 8, the distribution of
272 10 Vibration Analysis of Blade and Impeller Systems

(a) (b)

0 0
( r)
P ( )
( r) P ( )

0 0
( r ) P ( )
( r) P ( )
0 1 2 3 4 5 6 7 0 0 1 2 3 4 5 6 7 0
blade position blade position

product product

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
blade position blade position
=2,J=2 =2,J=6

Fig. 10.16 Modal exciting force (product of eigenmode and distributed exciting force)

the exciting force at Xt 0 is represented by the bar chart of the product at each
blade in Fig. 10.16.
(a) The number of nodal diameters j 2 in the eigenmodes and the exciting
modal number J = 2: The modal exciting force of the inner product is clearly
strong because the polarities of the eigenmodes and the exciting mode coin-
cide. As the bar chart shows, the product of the exciting force and the mode
exists for the blades of No. 0, 2, 4 and 6, and is zero for the blades of No. 1, 3,
5 and 7. The sum of these products is called the modal participation.
(b) j 2 and J = 6: Similarly the sum of the products exists and resonance
occurs.
Figure 10.17 shows the exciting force on a blade as a waveform during one
rotation of the rotor according to Eq. (10.33):
(a) j 2 and J = 2: Excitation as backward wave propagation
(b) j 2 and J = 6: Excitation as forward wave propagation
Therefore the resonance is observed by backward waves for j 2, J = 2, and
forward waves for j 2, J = 6.
Figure 10.18 shows the case of j 2 and J = 4. The product of the exciting
force and mode exists in the blades of No. 0 and 4 only. Resonance does not occur
because their phases are reversed and the sum is zero.
10.2 Vibration and Resonance of Blades and Impellers 273

(a) (b)
7 : blade number
7
6 6
exciting force of blade

exciting force of blade


5
5
4
4
3
3
2
2
1
1
0 : blade
number
0
t t
0 2 4 0 2 4
=2,J=2 = 2 ,J = 6

Fig. 10.17 Waveform of exciting force

( r )
P ( )

( r ) P ( )
0 1 2 3 4 5 6 7 0
blade position
=2, J=4

product

0 1 2 3 4 5 6 7

blade position

Fig. 10.18 Example of non-resonance condition

10.2.5 Criterion of Blade Resonance: Campbell Diagram

The combinations of the number of nodal diameters j in the eigenmodes and the
orders J of the excitating spatial mode are reviewed in Fig. 10.19 to nd the
distribution of the resonance condition. In this chart, 0 means that the inner product
of the eigenmode and excitating spatial mode is zero (no resonance); B (left up,
red), F (left down, blue) and S (at corners, green) mean resonance with backward
wave propagation, forward wave propagation and standing wave, respectively.
274 10 Vibration Analysis of Blade and Impeller Systems

20 19 18 17 16 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0 J/
S 0 0 0 0 0 0 0 S 0 0 0 0 0 0 0 S 0 0 0 0 4
0 B 0 0 0 0 0 F 0 B 0 0 0 0 0 F 0 B 0 0 0 3
0 0 B 0 0 0 F 0 0 0 B 0 0 0 F 0 0 0 B 0 0 2
0 0 0 B 0 F 0 0 0 0 0 B 0 F 0 0 0 0 0 B 0 1
0 0 0 0 S 0 0 0 0 0 0 0 S 0 0 0 0 0 0 0 S 0

Fig. 10.19 Resonance condition chart (N = 8)

In the system of Example 10.3, with kd = 10 and kc = 35, the natural frequencies
at rest are, in the ascending order of the number of nodal diameters j,

xj f x0 3:2 x1 5:5 x2 8:9 x3 11:4 x4 12:2 g


10:34
j 0. . .N=2 4

Lines corresponding to these frequencies are drawn in the natural frequency-


rotational speed plot in Fig. 10.20. These lines are horizontal here because the cen-
trifugal coefcient is neglected, but in reality these would be slightly slanted lines.

extended nodal diameter


(a) nodal number number
J = 13 ..... 6 5 4 3 K
4 2 4 12 4
3
natural frequency r

3 5 11 13 3
10
2 2 6 10 2

1 1 1 7 9 1
5
0 8 0
forward standing backward
wave wave wave
0 1 2 3 4 5 6
speed
J = 1~13 extended nodal diameter
nodal number number
(b) K
J = 24 ..... 16 15 14 13
4 12 12 20 4
3 11
natural frequency r

10 11 13 19 21 3
10
2 9 10 14 18 22 2
8
1 9 15 17 23 1
5
0 8 16 24 0
forward standing backward
wave wave wave
0 0.2 0.4 0.6 0.8 1.0
speed
J = 8~24

Fig. 10.20 Campbell diagram


10.2 Vibration and Resonance of Blades and Impellers 275

Figure 10.20a includes radiant lines representing exciting mode number J = 113;
higher orders are included in the Fig. 10.20b. All of the intersections between the
radiant lines and the horizontal lines are candidates of resonance condition, but
actually resonance occurs only at those which are non-zero in Fig. 10.19, which are
indicated by circles. As for the relationship between Figs. 10.20 and 10.21, white
circles for the backward propagation wave means resonances B, black circles for the
forward propagation wave resonances F, and double circles for standing wave reso-
nances S.
This plot, called a Campbell diagram [85], show the distinction between the
resonant and non-resonant intersections. This is a convenient tool to identify the
type of resonance and the rotational speed at which it occurs. In practice, blade
natural frequencies are adjusted to avoid the resonant intersections within the
possible range of rotational speed at rated operation.

Note: Resonance conditions related to modal excitation


Figures 10.19 and 10.20 were constructed according to the inner product of the
exciting modal number J and the eigenmode with the j-th nodal diameters. These
gures are intended to ease understanding of the resonance condition with the
product of non-zero values as stated in the following.
If one extends the number of nodal diameters j over the actually possible
number f 0; 1; . . .; N  1 g, the overflow f N; N 1; N 2; . . . g may
be listed in the right side of the gures. The natural frequency series according to
this extended nodal number j for the natural frequencies of Eq. (10.34) is

xj fx0 ; x1 ; x2 ; x3 ; x4 ; x3 ; x2 ; x1 ; x0 ; x1 ; x2 ; x3 ; x4 . . .g 10:35

The resonance condition [II], in which the integral of Eq. (10.33) with respect to
hr around the blade is evaluated to be non-zero simply reduces to

II J extended j K 10:36


The resonance conditions are thus summarized by Eqs. (10.32) and (10.36). In
practice, the resonant intersections are identied on the Campbell diagram of
Fig. 10.20 as the points at which an exciting modal number J coincides with an
extended number K of nodal diameters listed on the right side. The corresponding
actual numbers of nodal diameters j are shown in the rightmost column.
Example 10.5 Resonance conditions for a turbo-compressor rotor
Find the resonance conditions for the number of rotating blade Zr (e.g., turbine
blade, impeller blade, runners, vane, etc.) and stationary vanes Zg (e.g., guide vanes,
diffuser vanes, etc.).
Answer
In this case the number of the cyclic symmetry structure N is the same as the
number of rotating blades: Zr . The predominant component of the circumferential
pressure distribution in the stationary side is a harmonic function of the order equal
276 10 Vibration Analysis of Blade and Impeller Systems

to the number of stationary vanes, which is assumed to be an exciting force due to


the pressure distribution mode standing in the circumference direction. Since any
point on the impeller undergoes pressure variation every time it passes by the
stationary vanes, the order of excitation is thus J Zg and the excitation frequency
Zg X. The frequency resonance conditions corresponding to Eq. (10.32) are there-
fore given by

Zg X xj j 0. . .Zr =2 10:37

where j is the number of nodal diameters.


The resonance condition with respect to modal exciting force is here identied
by Fig. 10.21 (a modied version of Fig. 10.19). As the case of N Zr 8 and
J Zg includes three equations of the non-zero combinations, J j, J 8 j
and J 16 j as shown in the gure, the more general expression is given by the
following equation:

S 0 0 0 20 0 0 0 0
J=16

0 B 0 0 19 0 0 0 0
0 0 B 0 18 0 0 0 0
0 0 0 B 17 0 0 0 0
0 0 0 0 S 0 0 0 0
0 0 0 0 15 F 0 0 0
0 0 0 0 14 0 F 0 0
0 0 0 0 13 0 0 F 0
S 0 0 0 12 0 0 0 S
J=8

0 B 0 0 11 0 0 0 0
0 0 B 0 10 0 0 0 0
0 0 0 B 9 0 0 0 0
0 0 0 0 S 0 0 0 0
0 0 0 0 7 F 0 0 0
0 0 0 0 6 0 F 0 0
0 0 0 0 5 0 0 F 0
S 0 0 0 4 0 0 0 S
J=

0 B 0 0 3 0 0 0 0
0 0 B 0 2 0 0 0 0
0 0 0 B 1 0 0 0 0
-4 3 2 1 S0 +1 +2 +3 +4

Fig. 10.21 Resonance conditions (N = 8)


10.2 Vibration and Resonance of Blades and Impellers 277

48
32 extended nodal number K
natural frequency of nodal diameter

16
S
5 5 6 16 17 27 28 38 39 49 50
B
4 4 7 15 18 26 29 37 40 48


3 3 8 14 19 25 30


2 2 9 13 20 24 31
F
1 1 10 12 21 23 32

0 0 11 22

stress

Fig. 10.22 Example of impeller resonance condition (Zr = 11, Zg = 16)

Zg hZr j ! Zg  j hZr double signs in the same order for F=B 10:38

where h = 0, 1, 2, 3, and j 0; 1; 2; 3; . . .; N=2.


F signies the resonance of the forward wave propagation and B the resonance
of the backward wave propagation.
Let these conditions be further generalized. Signs are given to the number of
nodal diameters j; the plus and minus signs corresponding to forward and back-
ward propagation wave mode, respectively, and 0 or  N=2 to standing wave
modes. In addition, integer multiples of nZg X are also possible to be included in
excitation frequencies. Under these assumptions, the frequency resonance condition
of Eq. (10.37) is replaced by

I nZg X xj j 0. . .Zr =2; n 1; 2; 3. . . 10:39


and the modal excitation resonance condition of Eq. (10.33) becomes [86, 87]

II nZg  j hZr 10:40


where h = 0, 1, 2, 3, and j 0; 1; 2; 3; . . .; N=2.

Example 10.6 Describe the resonance conditions of an impeller of a centrifugal


compressor with rotating blades Zr = 11 and stationary vanes Zg = 16. Assume the
natural frequency curves in Fig. 10.22.
Answer
The excitation frequencies are rotational speed multiplied by the number of sta-
tionary vanes and integer multiples: the candidate order of excitation J = {Zg = 16,
2Zg = 32, 3Zg = 48}. The corresponding radiant lines are drawn in Fig. 10.22. The
right side of the gure shows the extended nodal number K (folded if necessary),
278 10 Vibration Analysis of Blade and Impeller Systems

among which the order of excitation J coincides with the number of nodal diameters
j at

fJ K; jg f16; 5@Sg; f32; 1@Fg; f48; 4@Bg; . . .


The corresponding intersections are indicated by circles. Also included in the
gure is the expected amplitude curves of the vibration stress expected to be
measured on the impeller blade.
Different solution: Considering 16 5 = 11, 2  16 + 1 = 3  11 and
3  16 4 = 4  11 in Eq. (10.40), {n, j, h} = {1, 5 @S, 1}, {2, +1 @F, 3},
{3, 4 @B, 4}

10.2.6 Case Study: Resonance in Impeller Blades of


Centrifugal Compressor [VB958]

Resonance of an impeller blade of a centrifugal compressor with rotating blades


(Zr = 11) and stationary vanes (Zg = 27) is analyzed according to Fukushima
[VB958]. The calculated and observed natural frequency curves are shown in
Fig. 10.23, along with the line for the order of excitation J = Zg = 27 and the
extended number K of nodal diameters. The only resonance point where K J is
marked with S (standing wave) at a rotation speed of 242.6 rps (the resonance
frequency is 242.6  27 = 6,550 Hz, called blade/vane passing frequency). An
amplitude curve that would be observed by measuring the vibration stress of blades
is included in the gure.
The FFT results of the vibration stress measured by strain gauges at two speeds
of around 243 and 248 rps are shown in Fig. 10.24. Plot (b), corresponding to
resonance, shows a high stress peak at 6,550 Hz, being equal to J = 27 times the
rotational speed. At a slightly different rotational speed, the resonance disappears as

extended nodal number K


8000 calculation ( measured values at rated speed )
S 27
natural frequency [Hz]

6355 (6525)Hz
5 5 6 16 17 27 28
6000
4970 (5125)Hz
4 4 7 15
4000 3319 (3425)Hz
3 3 8 14

1569 (1600)Hz
2000 1312 ( - ) Hz
2 2 9 13
0 0 11
1 1 10 12
847 ( - ) Hz
0
0 50 100 150 200 250 300
speed [rps]

Fig. 10.23 Campbell diagram of test impeller (Zr = 11, Zg = 27)


10.2 Vibration and Resonance of Blades and Impellers 279

vibration stress discharge pressure pulsation


(a) (c)
2 5

pressure Pv [kpaG]
4 11
stress v [%]

1.5 22
non-resonance

3
1
2
0.5
1
0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
frequency [Hz] frequency [Hz]
14,850 min-1=248s-1 14,850 min-1=248s-1

(b) (d)
100 5

pressure Pv [kpaG]
27 22
11
stress v [%]

80 4
6550Hz
resonance

60 3
40 2
20 1
0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
frequency [Hz] frequency [Hz]
14,550 min-1=243s-1 14,550 min-1=243s-1

Impeller blade Z r = 11 , Diffuser blade Z g = 27

Fig. 10.24 Stress and pressure pulsation at resonance and non-resonance conditions

evidenced in the corresponding plot (a) indicating a number of low peaks. High
peak resonance results from a high Q-factor, which characterizes the resonance
vibration of impeller blades with the occurrence at a very narrow frequency band.
The FFT results of delivery pulsation pressure measured in rotating tests are
shown in (c) and (d) in the gure. Peaks appear at 243  11 2; 670 Hz and twice
5,340 Hz because the pressure pulsation has components of (rotational speed) 
(number of rotating blades Zr) and its integer multiples, called blade-passing fre-
quency and multiples. This pressure pulsation always acts upon the stationary part,
regardless of the rotating blade vibration. Therefore, note that the bearing supports
or other stationary-side in sub-structures of the equipment should be avoided from
mechanical resonances caused by this pulsation.
The FFT results of the vibration stress of the rotor measured in the decelerating
process in the same tests by the peak-holding mode are shown in Fig. 10.25, which
show peaks generated by passage over the resonance condition
fJ K; jg f27; 5 Sg. The eigenmode for j 5 is illustrated in Fig. 10.26.
Example 10.7 Consider a rotating shaft system with three blades shown in
Fig. 10.27. The natural frequencies for the number of nodal diameter j 0 and 1
are written as x0 and x1 , respectively. Air is blown from nozzles onto the two
locations on the periphery of the blading plane. Construct a Campbell diagram for
resonance conditions assuming x0 [ x1 :
280 10 Vibration Analysis of Blade and Impeller Systems

Fig. 10.25 Peak hold 100


spectrum strain gauge No.4 =5
80

stress v
60

40
=3 =4
20

0
0 2000 4000 6000 8000 10000
frequency [Hz]

Fig. 10.26 Analytica result


of natural frequency of blade
mode of nodal diameter j = 5
(6,550 Hz)

Fig. 10.27 Rotating shaft nozzle


system with 3 blades blade

shaft

nozzle nozzle

Answer
The nozzle injection at two locations produce a hydrostatic distribution shown in
Fig. 10.28 corresponding to Zg 2 and then its integer multiples are the excitation
frequencies. Since the orders of the exciting modal numbers are given over Zg 2 as
J nZg 2; 4; 6; 8, the Campbell diagram is shown in Fig. 10.29, with inclusion
of the extended nodal numbers K over Zr 3 as K = 0, 1, 2, 3, 4  8. These
intersections are the resonance conditions which satisfy the formula of Eq. (10.40).
10.3 Blade/Impeller Vibrations Excited at Stationary Side 281

Fig. 10.28 Hydrostatic


pressure distribution

J = Zg = 2

Fig. 10.29 Campbell J = n Zg


diagram (Zr = 3, Zg = 2) 8 7 6 5
r 4
extended
3 nodal number
K
2
0 0 3 6
1
1 1 2 4 5 7 8

10.3 Blade/Impeller Vibrations Excited at Stationary Side

10.3.1 Difference in Excitation Methods and Resonance


Conditions

The previous section discussed the resonance conditions of a blade and an impeller
excited by a static circumferential pressure distribution given by the stationary side
of the equipment. This section deals with resonance of a rotating blade caused by
sinusoidal excitation at a location in the stationary side, for example, by harmonic
excitation from an electromagnet located at a stationary stand. The difference from
the previous section in excitation method should not be overlooked. The natural
frequencies of a blade as observed from the stationary side are examined rst, and
then it is stated what excitation in such stationary frequency domain causes
resonances.

10.3.2 Representation of Vibration of Blades and Impellers


in an Inertial Coordinate System

In a layout shown in Fig. 10.30, the vibration displacement (deformation) of a blade


and an impeller in an out-of-plane mode is proportional to the stress rb . The general
282 10 Vibration Analysis of Blade and Impeller Systems

Fig. 10.30 Measurement of 2 stress gap sensor


stress waveform 3 1 a
b

4 0


5 7
6

formula for the stress/strain db with the natural frequency xj with a number of
nodal diameter j and a radial mode Ar is given at a blade phase angle h:

db 2Ar cos jh cos xj t 10:41a

db Arcosxj t jh cosxj t  jh 10:41b

The rst Eq. (10.41a) represents a standing wave and the rst and second terms
of the second Eq. (10.41b), respectively, a backward and forward propagation
wave. The stress ri of the i-th blade listed in the upper part of Table 10.5 is
obtained by substituting the phase angle hi of the i-th blade into Eq. (10.41a).
The lower part of the table shows the waveforms of out-of-plane vibration
observed with a gap sensor in the stationary side (see Fig. 10.30), which are
obtained by substituting the blade phase angle hi Xt into Eq. (10.41a):

da 2Ar cos jXt cos xj t


10:42
Arcosxj t jXt cosxj t  jXt

This means that the frequency xs seen from the inertial coordinate system,
which is obtained by looking at each component of blade vibration from the sta-
tionary side, is

xs xj  jX 10:43

Table 10.5 Observation of blade vibration

blade vibration forward wave backward wave standing wave


1 i cos( t i ) cos( t + i ) cos i cos t
frequency
observation in cos( t + t ) cos( t t ) cos t cos t
2 the fixed frame
frequency + , +
10.3 Blade/Impeller Vibrations Excited at Stationary Side 283

Fig. 10.31 Blade vibration in s


inertial coordinates ave

natural frequency at
ard w
forw

static coordinate

+ +





back
ward
wave
speed

This relationship is represented graphically in Fig. 10.31. In practice, the line xk


may be slightly curved due to the centrifugal coefcient influence.
Example 10.8 Table 10.6 shows the stress vibration waveforms ri for the i-th
blade with a natural frequency xj 3Hz and at a rotational speed X 1Hz for
numbers of nodal diameters j 1, 2 and 3, as measured by strain gauges attached
to the blades at a pitch of 45 . At the same time, we can see the corresponding
waveforms da as measured by a gap sensor in the stationary side. Conrm the
frequency relationships shown in the table.
Answer
According to Table 10.5, the frequency of waveforms ri = {forward xj, backward
xj, standing xj } Hz@X\,Hz is converted to the frequency of waveforms ra
fxj jX; xj  jX; xj  jXgHz as follows:
for Table 10.6 j = 1, ri = {3, 3, 3}Hz @1 Hz ! ra = ri +{1, 1, 1} = {4,
2, 4 and 2} @1 Hz,
for Table 10.6 j = 2, ri = {3, 3, 3}Hz @1 Hz ! ra = ri + {2, 2, 2} = {5,
1, 5 and 1} @1 Hz and
for Table 10.6 j = 3, ri = {3, 3, 3}Hz @1 Hz ! ra = ri + {3, 3, 3} = {6,
0, 6 and 0} @1 Hz.
These converted frequencies are measured in the corresponding waveforms.

10.3.3 Resonance Condition 1

When the out-of-plane vibration of a disk is excited with a frequency m at a single


location from the stationary side of the equipment, the resonance condition is:

m xj  jX 10:44
284 10 Vibration Analysis of Blade and Impeller Systems

Table 10.6 Examples of stress ri t waveform and gap sensor da t waveform X 1 Hz

nodal
diameter forward wave ( 3Hz ) backward wave ( 3Hz ) standing wave ( 3Hz )
7 7 7
6 6 6
i 5
4
5
4
5
4
3 3Hz 3 3Hz 3 3Hz
2 2 2
(1) = 1 1 1 1
0 0 0
0 1 2 0 1 2 0 1 2
a 4Hz
2Hz
2Hz
0 1 2 0 1 2 0 1 2 4Hz
7 7 7
6 6 6
5 5 5
i 4
3
3Hz 4
3 3Hz 4
3 3Hz
2 2 2
(2) = 2 1
0
1
0
1
0
0 1 2 0 1 2 0 1 2
5Hz 1Hz
a 1Hz
0 1 2 0 1 2 0 1 2 5Hz
7 7 7
6 6 6
5 5 5
i 4
3
3Hz 4 3Hz 4
3 3Hz
3
2 2 2
(3) = 3 1
0
1
0
1
0
0 1 2 0 1 2 0 1 2
a 6Hz 0Hz 0Hz
0 1 2 0 1 2 0 1 2 6Hz

where + and means forward and backward blade resonance, respectively.


Harmonic excitation at a constant rotational speed X induces resonance at two
frequencies with the same number of nodal diameters, as shown in Fig. 10.32:

m1 xj  jX
10:45
m2 xj jX

This is applied for identifying the natural frequency xj and the number of nodal
diameter j as follows:

Fig. 10.32 Resonance on e0


condition 1

natural frequency measured


+
on the inertia frame

2

1

speed
10.3 Blade/Impeller Vibrations Excited at Stationary Side 285

xj m1 m2 =2
10:46
j m2  m1 =2X

Resonance with any number of nodal diameters j can be thus induced by


sweeping the excitation frequency m at any rotational speed X. This relationship is
utilized to search for natural frequencies with the number of the nodal diameter.

10.3.4 Resonance Condition 2

When the excitation frequency m is proportional to the rotational speed X, i.e.,


m pX, the resonance condition is:

pX xj  jX 10:47

where p is a proportional constant, and the j means forward and backward prop-
agation wave of blade resonances, respectively. This situation is shown graphically in
Fig. 10.33. The blades resonate at two rotational speeds X1 , and X2 which corre-
spond, respectively, to resonances of backward and forward propagation waves.
Note: In this eld, references [8898] provide further information and details on
specic problems. The reader is referred to these to experience the range of
problems that relate to the techniques discussed in this chapter.

Fig. 10.33 Resonance on e 0


condition 2

natural frequency measured

p
in the inertia frame

1 2
speed
Chapter 11
Stability Problems in Rotor Systems

Abstract This chapter discusses three typical topics of rotor dynamics problems:
internal/external damping effects, vibration due to non-symmetrical shaft stiffness
and thermal unbalance behavior. Though a rotor should rotate in a stable manner in
a rotation test, problems are encountered in some cases. Most of the problems are
related to unbalance, against which the countermeasure is balancing. However,
more serious problems may occur that cannot be solved by balancing. In such cases
other solutions must be sought. This chapter discusses the following three problems
that may be encountered:
(1) Internal damping: Loose ttings on the shaft cause damping due to sliding
friction. It might seem that any damping is welcome, but this type of damping is
rather a destabilizing factor at high speeds of rotation.
(2) Asymmetric section of the rotor: Asymmetry in shaft stiffness, e.g. due to a key
slot on the shaft often generates troublesome vibration.
(3) Vibration due to thermal bow: The unbalance vibration vector of a rotor can be
monitored during operation by a Nyquist plot. While the vector point normally
remains unchanged during steady state operation, thermal deformation of the rotor,
e.g., due to rubbing will move it. The mechanism of this phenomenon is described.
For simplicity, a single-degree-of-freedom model is used in the following discussion.


Keywords System stability External damping  Internal damping  Asymmetric
 
stiffness Thermal unbalance Hot spot

11.1 Unstable Vibration Due to Internal Damping


of a Rotor [B5]

11.1.1 Equation of Motion

A rubber coupling of the rotor (Fig. 11.1) provides very strong damping effect in an
impulse test at rest. Loose t due to wear between the shaft and the blade disk
causes friction damping. Wear or poor lubrication of a gear coupling tooth gives
Springer Japan 2017 287
O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_11
288 11 Stability Problems in Rotor Systems

Fig. 11.1 Internal damping cr and external damping ce

rise to sliding friction damping. Fiber-reinforced plastic (FRP) used in the rotating
part allows material damping. These types of damping are inherent to the rotating
rotor, and collectively called internal damping, cr . Conversely, the damping dis-
cussed so far which acts on the shaft from the ground is called external damping, ce .
The internal damping, cr , of a rotor is dened in the rotational coordinate
system. The reaction force Fr is represented in the complex form

Fr cr x_ r jcr y_ r cr z_ r 11:1

where zr  xr jyr is the complex displacement measured in the rotational coor-


dinate system. This is transformed to the reaction force F in the inertial coordinate
system:
d
F Fr ejXt cr z_ r ejXt cr zejXt ejXt
dt 11:2
cr z_  jXcr z

where z zr ejXt is the complex displacement measured in the inertial coordinate


system.
The equation of motion for the single-degree-of-freedom system corresponding
to the rotor is, including the internal and external damping,

mz  jXcg z_ kz ce z_ cr z_  jXcr z 0 11:3a

mz  jXcg z_ kz ce cr z_  jkXz 0 11:3b

where k cr =ce cr
11.1 Unstable Vibration Due to Internal Damping of a Rotor 289

11.1.2 Stability Condition

The damping ratio of the rotor at rest ( = 0) is high because the damping is the
sum of the internal and external damping ce cr . During the rotation, however,
the internal damping Xcr acts in the same way as the cross-spring kc that decreases
the damping ratio for forward whirling mode and increase that for backward
whirling mode. With internal damping, kc Xcr is increased proportionally with
rotational speed, so that the stability of forward whirling decreases as the rotational
speed increases, eventually resulting in unstable vibration.
The rst of Eq. (B.5) in Appendix B, with the gyroscopic effect neglected and
Xcr and ce cr substituted for kc and cd , respectively, gives
 
da 1 Xcr
 ce cr  a 11:4
dt 2m xn

which leads to the stable condition

X ce
\1 , kX\xn 11:5
xn cr

The graphical representation in Fig. 11.2 shows the rotational speed limit for stable
operation. In particular, a system consisting of a rotor coupled by a rubber coupling
(with a high cr ) and supported by ball bearings (with a low ce  0) often shows
unstable vibration immediately after the rotational speed has passed through the
natural frequency (critical speed), as shown by the dotted line.

Fig. 11.2 Instability onset


speed
290 11 Stability Problems in Rotor Systems

11.1.3 Stability Analysis

Figure 11.3 explains graphically the instability due to the internal damping. The
damping force is given by

F ce z_ cr z_  jXcr z
11:6
 ce cr jxaejxt jXcr aejxt

where z aejxt is the natural vibration in a circular whirling mode.


In the forward whirling (x xf , (a) in the gure), the downward damping
(ce cr ) counteracts whirling, while the upward Xcr promotes whirling.
Consequently, the system is stable as far as the counteracting force is totally greater
than promoting force;

ce cr xf [ Xcr 11:7

which is identical with Eq. (11.5).


In contrast, the backward whirling (x xb , (b) in the gure) is stable even in
high rotational speeds because it is counteracted by both ce cr , and Xcr .
This analysis thus suggests some design policies, such as
Connecting the rotor tightly without looseness and /or using soft materials
Providing damping forces solely by dampers xed at the ground
Paying attention to frictional damping due to wear in couplings and gears

+ j b ce
y
(a) (b) y
f + j b cr
j cr
b

j cr


z x
j f cr z x

j f ce

forward = f backward = b

Fig. 11.3 Instability due to internal damping


11.1 Unstable Vibration Due to Internal Damping of a Rotor 291

1. Case study: Wear in gear tooth coupling [VB38]


A centrifugal compressor driven by an electric motor via a step-up gear and gear
coupling (Fig. 11.4) showed excessive vibration. The FFT results are shown in
Fig. 11.5. The frequency of the predominant component at 72 Hz was estimated to
be about 1/3 of the compressor rotational speed, about twice the motor rotational
speed, or as such non-synchronous with these rotational speeds. It was initially
impossible to determine the causes of this vibration problem, whether due to some
non-linear resonance or self-excited vibration. Detailed study of the dominant
frequency was performed in rotation tests at a xed rotational speed and varied
delivery pressure Pd . The dominant frequency was found to shift slightly with Pd
although the rotational speed was kept constant, indicating that the vibration was
self-excited.
Further investigation revealed that the instability was caused by damping due to
sliding friction of the worn teeth of the gear coupling. This is the internal damping
discussed previously.

Fig. 11.4 Centrifugal compressor

Fig. 11.5 FFT frequency


analysis
292 11 Stability Problems in Rotor Systems

Fig. 11.6 Diaphragm


coupling

expensive, but suitable for high speed and low load

The system was re-stabilized by using a diaphragm coupling (Fig. 11.6, no


internal damping), which was expensive but slip-free, rather than trying other
means to increase the external damping.
It is a general trend to extend the operational life of equipment, often through
continued operation, despite having worn gears in the rotor system. In fact, some
cases of geared rotors have been reported in which unstable vibrations suddenly
occurred after long-term operation with stable vibration.
2. Case study: Silicon crystal manufacturing equipment [100, VB268]
The silicon crystal manufacturing equipment shown in Fig. 11.7 is used to
manufacture single-crystal silicon by taking up the growing crystal from a molten
melt, while rotating the wafer. When the conventional long rotating take-up rod ((a)
in the gure) was replaced by a wire rope ((b) in the gure) in order to reduce the
size of the equipment, slow forward whirling motions appeared, which was visually
observable.
This was a slow self-excited whirling with the natural frequency of the pendulum
mode involving a gradual increase of the amplitude. By comparison with conven-
tional equipment, it was found that the instability was caused by the internal damping

Fig. 11.7 Silicon crystal


manufacturing system (a) (b)
11.1 Unstable Vibration Due to Internal Damping of a Rotor 293

Fig. 11.8 Magnetic damper

due to the friction in the take-up wire, which was fabricated from yarns. Stability was
improved by using a wire rope twisting together fewer number of thick yarns.
Example 11.1 A magnetic damper (Fig. 11.8) consisting of an aluminum plate
between magnets, which can be used to reduce vibration of a rotor. Which
arrangement in the gure is better, and why?
Answer:

Note: Damping comes from eddy currents generated in the aluminum plates. Then
the aluminum plates should be placed at the stationary side and the magnets at the
rotor side to produce external damping , therefore to avoid internal damping .

11.2 Unstable Vibration of an Asymmetric Rotor System

11.2.1 Equation of Motion

The cross-section of the rotating shaft should be circular in principle, but sometimes
it is asymmetric, for example, by the presence of a spline or key way (Fig. 11.9a).
In this case, the symmetry is achieved by adding dummy splines (Fig. 11.9b).
However, the shaft of a two-pole generator (Fig. 11.10) has to remain asymmetric
due to structural reasons. This chapter discusses the vibration characteristics of such
a rotor with asymmetric flexural rigidity [B5].
For an asymmetric shaft shown in Fig. 11.11, it is assumed that the flexural
rigidity in the Xr direction is greater than that in the Yr direction as observed in a
rotational coordinate system. Considering this as follows:

 Fxr  k Dkxr
 Fyr  k  Dkyr

The relationship of the displacement zr xr jyr and the reaction force Fr


Fxr jFyr in complex form can be written as
294 11 Stability Problems in Rotor Systems

Yr Yr
(a) keyway (b) keyway

Xr Xr
dummy
dummy

mg mg

asymmetric k > 0 symmetric k = 0

Fig. 11.9 Key slot

Fig. 11.10 Rotor of 2-pole cross slot


coil slot
generator
lead slot
lead hole

end coil and ventilating slot

Fig. 11.11 Coordinates of Yr Y


asymmetric shaft
k
+k

yr Xr

xr
t
g
X
11.2 Unstable Vibration of an Asymmetric Rotor System 295

Fr  k Dkxr jk  Dkyr kzr Dkzr 11:8

Considering the following equation:

F Fr ejXt ; z zr ejXt 11:9

the relationship of the displacement z and the reaction force F measured in the
inertial coordinate system is written as

F Fr ejXt kzr ejXt Dkzr ejXt kz Dke2jXtz 11:10

Therefore, the equation of motion for a single-dof model of the rotor system,
including the terms for unbalance me and gravity g, is represented as

mz  jXcg z_ c_z kz Dke2jXtz meX2 ejXt mg 11:11

For simplicity, the variables

cg c k Dk
c ; 2fxn ; x2n ; 2l D 11:12
m m m k

are introduced to rewrite Eq. (11.11) as

z  jcX_z 21xn z_ x2n z 2lx2n e2jXtz eX2 ejXt g 11:13

The following discussion is based on this equation.

11.2.2 Overview of Vibration in an Asymmetric Rotating


Shaft

Asymmetry of the rotating shaft can be evaluated by impulse tests at different phase
angles on the circumferential of the stationary rotor to check the uniformity of the
natural frequency. The uniformity is estimated from the measured values as:
xx  xy  
l , xx ; xy xn 1  l double signs in the same order 11:14
2xn

where xx and xy = the higher and lower natural frequencies, respectively;

xn xx xy =2 average natural frequency:

A rotor shaft having stiffness difference in the two perpendicular directions


generates peculiar vibration phenomena as shown in Fig. 11.12. Three features [A],
[B] and [C] indicated in the gure are discussed here:
296 11 Stability Problems in Rotor Systems


2
2nd
[A] [B]
2
[C]
1st
[A] [B] 2
x
1
y

y x
1 +2
n
1 = n 2 2
2
1st mid 2nd
resonance
due to gravity
unstable
shaft vib.

2X vib. zone
circular orbit


p=
0 0.5 1.0 n
rotational speed

Fig. 11.12 Vibration characterstics of asymmetric shaft

[A] Resonance at the secondary critical speed: X xn =2


[B] Stability of the resonance mode: X  xn
[C] Stability between the resonance modes: X x1 x2 =2

[A] Resonance at the secondary critical speed: X  xn =2


Assuming that the absolute displacement z consists of the vibration displacement
w from the bending static position zg caused by gravity,

z w zg 11:15

where zg g=x2n .
Even if the rotor is perfectly balanced (unbalance = 0) and the gyroscopic
effect is neglected, approximation of Eq. (11.13):

21xn w_ x2n w 2lx2n 


w w zg e2jXt  2lge2jXt 11:16
11.2 Unstable Vibration of an Asymmetric Rotor System 297

(a) (b) core


plate

x=y
mg

2
cross-section up and down
frequency
mg sirocco fan
k x = k y

2
shaft deflection ( 1rev. ) up and down
frequency mg
shaft thin circular plate

Fig. 11.13 Shaft stiffness and tilting stiffness of thin plate

shows the existence of a forward excitation force corresponding to twice the


rotational speed as shown in Fig. 11.13a. This originates from the slight
up-and-down movement of the shaft deflection curve twice per revolution due to the
gravity and the asymmetric stiffness.
The Campbell diagram in the upper part of Fig. 11.12 indicates that a subor-
dinate resonance synchronizing with a frequency at twice the rotational speed to the
forward whirling appears at a rotational speed at half the natural frequency:

X  xn =2 11:17

if the gyroscopic effect is absent. This rotational speed is called the secondary
critical speed of the asymmetric shaft.
Since the Q factor is 1=2f, the peak amplitude at the secondary critical speed is
estimated by
 
2lg 1 l g
apeak 11:18
x2n 2f f x2n

Figure 11.13a indicates the secondary resonance due to the deflection of the shaft
caused by the anisotropy of the shaft stiffness combined with gravity (because a
horizontal rotor is considered here).
On the other hand, a rotor with a thin circular plate, called a core plate in a
sirocco fan as shown in the gure (b), often has asymmetric bending stiffness due to
the circumferential non-uniformity of thickness of the core plate. In this case the
core plate is statically tilted by gravity g acting on the fan. The tilted part experi-
ences up-and-down small movement twice in a revolution, causing the secondary
vibration resonance in the tilting vibration mode of the plate to have one nodal
diameter. Refer to Sect. 9.4.2 for details.
298 11 Stability Problems in Rotor Systems

Fig. 11.14 Secondary n f


critical speed

2
=

1


p=
0 1 n

Since the gyroscopic effect is signicantly higher in the tilting mode, the sec-
ondary critical speed is, as shown in Fig. 11.14, the intersection between the
excitation force 2X and the forward natural frequency xf (the solid line), marked
by a circle. No particular phenomenon appears at the intersection between the force
2X and the backward natural frequency (the broken line xb ), marked by a star.
Furthermore, a rotor with a thin blade as shown in Fig. 11.15 often has asym-
metric stiffness of the blade system due to the non-uniform thickness of each blade.
In addition, the rotor is subject to a constant moment if the hydrostatic pressure Dp
acting on blades is not uniform. Since the static load Dp is equivalent to the gravity
g in this tilting mode, the secondary resonance as shown in Fig. 11.14 appears also
in this case.
Equation (6.23) shows that, with the gyroscopic effect , the asymptote of the
forward natural frequency xf is cX. In the tilting vibration mode, therefore, even if
an unbalance induced resonance does not appear due to the gyroscopic effect > 1,
a secondary resonance may occur at higher rotational speeds. In the extreme case of
= 2, however, no secondary resonance occurs because the natural frequency curve
has no intersection with the 2 line as seen in Fig. 11.14.

(a) (b) (c)


p

2 p
tilting motion
frequency

impeller tilting mode non-uniform pressure distribution

Fig. 11.15 Tilting vibration of impeller (nodal dia. = 1)


11.2 Unstable Vibration of an Asymmetric Rotor System 299

[B] Stability of the resonance mode

r r
k  Dk k Dk
instability range: xy  \X\xx  11:19
m m

Even if the unbalance and gravity can be neglected, Eq. (11.13) becomes:

z 21xn z_ x2n z 2lx2n e2jXtz 0 11:20

The gyroscopic effect is neglected for simplicity. This is called a linear system,
involving parametric excitation at a frequency of 2, which shows parametric
resonance when the natural frequency xn is close to half the exciting frequency 2,
being approximately xn  2X=2 X. In fact, the difference between the x and
y directional natural frequencies becomes the instability range as indicated by
Eq. (11.19): The reason is given as follows.
If vibration synchronous with the rotation is assumed to be caused by unbalance:

z AejXt 11:21

a new exciting force synchronous to the rotation

2lx2 e2jXtz 2lx2 e2jXt AejXt  2lx2 ejXt A

is added to Eq. (11.20) as a new another force, which induces an additional reso-
nance vibration because of xn  X. Continuing in this manner, the amplitude
A develops. In this situation, the system becomes destabilized and its amplitude
increases.
The stability of the parametric resonance can be analyzed by applying the
method of averaging the amplitude and phase of Eq. (11.21) to obtain the solution
of Eq. (11.20), assuming that the complex amplitude At changes only gradually
with time. If the damping ratio and anisotropy are assumed to be innitesimally
small quantities of order ,

_ e; OrderA
OrderA e2 ; Orderx2  X2  e 11:22
n

The derivatives can therefore be written to the rst order of small

_ jXt
z_ AjXejXt Ae
11:23
_
z AX2 ejXt 2AjXe jXt jXt  AX2 ejXt 2AjXe
Ae _ jXt

Consequently, for the rst order of , Eq. (11.20) is replaced by


300 11 Stability Problems in Rotor Systems

 
_ 2lx2 A 0
x2n  X2 2fxn jX A 2AjX 11:24
n

For the solution of this equation

A A1 est A2 est 11:25

the characteristic equation is



x2n  X2 2fxn jX 2jXs 2lx2n A1


0 11:26
2lxn2
xn  X  2fxn jX  2jXs
2 2
A2

For the undamped case ( = 0), the equation is


 2
4X2 s2 x2n  X2 4l2 x4n 0 11:27

Since the characteristic root s must be purely imaginary for a substantially stable
system, the stability condition requires
 2
x2n  X2 [ 4l2 x4n ! r [ l; 11:28

jX  xn j X
where r 1  is the separation margin.
xn xn
This formula indicates the range of stable separation, supporting Eq. (11.19).
The damping ratio required for stable system is obtained from the characteristic
root s of the equation derived from Eq. (11.26) for the worst case of X xn :

4X2 s fxn 2 4l2 x4n 0


11:29
) s2 2fxn s f2  l2 x2n 0

Thus, the damping ratio needed for stabilization is:

f[l 11:30

[C] Stability between two resonance modes

x1 x2
instability range: X 11:31
2

In addition to the instability at rotational speeds around the rst or second natural
frequencies, x1 or x2 , described in [B], a multi-degree-of-freedom system can
show a third instability at a rotational speed between x1 and x2 . This is the coupled
instability problem [C] between the two modes, which is discussed below.
11.2 Unstable Vibration of an Asymmetric Rotor System 301

The equation of motion for a multi-degree-of-freedom system involving a mass


matrix and asymmetric stiffness matrix (neglecting the gyroscopic effect and
damping) is written as follows by analogy with Eq. (11.11):

M Z K Z DK Ze
 2jXt 0 11:32

where K  is the average shaft stiffness and DK  is the small asymmetry in stiffness.
Equation (11.32) can be converted to the equation of motion represented in the
modal coordinates g1 and g2 by modal analysis with respect to the rst and second
modes /1 and /2 :





m1 0 g1 k 0 g1 D11 k1 D12 k1 


g1 2jXt
1 e 0 11:33
0 m2 g2 0 k2 g2 D21 k2 D22 k2 
g2

where mi  /ti M /i is the modal mass, ki  /ti K /i the modal stiffness (i = 1, 2),
Dij ki  /ti DK /j the anisotropy of the modal stiffness (i, j = 1, 2).
D11 and D22 in Eq. (11.33) are the orthogonal elements of the modal stiffness, of
which the effect was already included in the above discussed in [B]. Here the
coupling elements D12 and D21 are converted to the modal parameters l12 and l21 ,
and the following equations of motion (including damping) are examined instead of
Eq. (11.33):

g1 2f1 x1 g_ 1 x21 g1 2l12 x21 


g2 e2jXt 0
11:34
g2 2f2 x2 g_ 2 x22 g2 2l21 x22 
g1 e2jXt 0

When vibration of the rst mode g1 Aejx1 t arises due to some source, a new
external force

2l21 x22 g1 e2jXt 2l21 x22 Aej2Xx1 t  2l21 x22 Aejx2 t

induces another resonance in the second mode g2 in the second of Eq. (11.34).
Conversely, a vibration of the second mode causes resonance in the rst mode.
These two modes are thus coupled to produce another resonance, the amplitude of
which increases so as to nally lead to instability.
In order to examine the stability condition in this situation, the vibrational
solutions in the two modes are assumed:

g1 Aejx1 t
11:35
g2 Bej2Xx1 t

with the amplitudes gradually changing with time. By analogy with Eq. (11.23),
substituting the following derivatives evaluated by approximation in the rst small
order :
302 11 Stability Problems in Rotor Systems

_ jx1 t
g_ 1 Ajx1 ejx1 t Ae
_ j2Xx1 t
g_ 2 Bj2X  x1 ej2Xx1 t Be
11:36
_ 1 ejx1 t Ae
g1 Ax2 ejx1 t 2Ajx jx1 t  Ax2 ejx1 t 2Ajx
_ 1 ejx1 t
1 1
_
g2  B2X  x1 2 ej2Xx1 t 2Bj2X  x1 ej2Xx1 t

into Eq. (11.34) and ignoring terms higher than rst order,
 
2jx1 A_ f1 x1 A 2l12 x21 B 0
h i   11:37
x22  2X  x1 2 B  2j2X  x1 B_ f2 x2 B 2l21 x22 A 0

are obtained, from which the characteristic formula in the s domain is obtained:

2jx1 s f1 x1 2l12 x21 A


2 0
2l21 x22 x2  2X  x1  2j2X  x1 s f2 x2
2 B
11:38

Considering Eq. (11.31) for an undamped case (i.e. damping ratio i is zero) and
approximating the (2, 2) element of the above characteristic matrix as
x1 x2
x22  2X  x1 2  2j2X  x1 s 4 X  x2  2jx2 s
2

the following characteristic equation and characteristic root are obtained:

s2  2jX  x1 x2 =2s  l12 l21 x1 x2 0


q 11:39
) s jX  x1 x2 =2  l12 l21 x1 x2  X  x1 x2 =22

The condition of substantial stability requires the expression under the square root
to be negative, which yields the following inequality, called the stable separation:
x1 x2 p

X  [ l12 l21 x1 x2 11:40
2

It corresponds to zone [C] in Fig. 11.12.


As for the damping ratio needed for stabilization, Eq. (11.38) gives the char-
acteristic equation for the worst case situation corresponding to 2X x1 x2 :

s f1 x1 s f2 x2  l12 l21 x1 x2 0 11:41

For stability, therefore, the damping ratio must satisfy


11.2 Unstable Vibration of an Asymmetric Rotor System 303

f1 f2 [ l12 l21 11:42

As described so far, asymmetry of the rotating shaft stiffness gives rise to the
secondary resonance [A] and the unstable zone [B, C] illustrated in Fig. 11.12,
making stable operation of the rotor difcult to achieve. The flexural rigidity of the
shaft and the tilting stiffness of the disk (number of nodal diameter j 1) should
therefore be uniform in the circumferential direction. Lessons learned from the
analysis are:
The rotor section should be a true circle,
The disk thickness should be uniform, and
Precise cancellation design is needed for a low damping rotor having asym-
metric stiffness and mass.

11.2.3 Simulation of Vibration of Asymmetric Rotor

This section visualizes the phenomena discussed above by simulation waveforms.


Calculations are based on the equation of motion for the two-mass system shown in
Fig. 11.16:

M Z DZ_ KZ DKZe2jXt UX2 ejXt Fg 11:43



z 1 0
where the complex displacement Z 1 , mass matrix M ,stiffness
z2 0 1

2:5 1:5 1
matrix K , unbalance U e,
1:5 2:5 j

1.0 j

Complex eigenvalues
12 1.5 1
1 1 1st
damped eigenvalue = 0.025 j 1
0.05 z1 y z2 0.05
damping ratio 1 = 0.025
0.71 0.71
Q-value = 20

2nd
m1* = 1 21 = 1 1= 1 1
damped eigenvalue = 0.025 j 2
0.71
damping ratio 1 = 0.0125
Q-value = 40
0.71
m 2* = 1 22 = 4 2= 2 2

Fig. 11.16 An example model of asymmetric rotating shaft system


304 11 Stability Problems in Rotor Systems

2D 0 1
asymmetric stiffness matrix DK , and gravity Fg Mg .
0 D 1
As indicated in Fig. 11.16, the real eigenvalues of the M-K system give the
natural frequencies and the corresponding eigenmodes, which are normalized by the
mass matrix. The damping ratio and Q value are also shown, obtained from the
complex eigenvalues of the M-D-K system.

1 1
Using the modal matrix, U /1 ; /2  0:71 , the modal asymmetric
1 1
stiffness, gravity and unbalance are,

" 2 #
1:5 0:5 2x1 l11 2x21 l12
U DKU D
t

0:5 1:5 2x22 l21 2x22 l22
 p t 11:44
Ut Fg 2 0 g
 t
Ut U ej45 ej45 e

The equations of motion in the modal coordinate system gi are


p
g1 2f1 x1 g_ 1 x21 g1 2x21 l11 g1 l12 g2 e2jXt eej45 X2 ejXt 2g
11:45
g2 2f2 x2 g_ 2 x22 g2 2x22 l21 g1 l22 g2 e2jXt eej45 X2 ejXt

l11 l12 0:75 0:25


where D .
l21 l22 0:0625 0:1875
The amplitude jAjXj of the unbalance vibration of the normal (i.e., symmetric)
system is measured by a displacement sensor located at the left mass as follows:
 1
As 1 0  Ms2 Ds K UX2 11:46

and used to draw the resonance curve in Fig. 11.17. The procedures and input data
to identify the resonance peaks are placed on the right side.
The time history response curves according to Eq. (11.45) and FFT results for
the cases [A], [B] and [C] described above are shown in Fig. 11.18.
[A] Secondary critical speed = 0.5:
A time history response curve is shown in [A] for D 0:1;
l l11 0:75D;g=x2n g=x21 3e.
The peak amplitude according to Eq. (11.18) is 2
Q1 lg=x2n Bin Cout 2

p
Q1
0:75
0:1
3e 9e (output coefcient Cout 0:71 1= 2, input coef-
  p
cient = Bin Ut Fg 1; 1 2 and Q1 20). x1 1 is predominant in the
spectrum.
[B1] First critical speed = 1.0:
A time history response curve is shown in [B1].
11.2 Unstable Vibration of an Asymmetric Rotor System 305

[ B2]
30
critical speed

25 1st [ B 1 ] 2nd [ B 2 ]

modal mass eccentricity * =


amplitude [ ]

20
[ B1 ]
45 45
15
[A] modal peak amp. peak = *Q
10 20 40
[C] sensor peak amp.
5
20 0.71 40 0.71
0 = 14 = 28
0 0.5 1 1.5 2 2.5
speed 1

Fig. 11.17 Unbalance vibration resonance curve

T = 50 1
amplitude ( ) 15 20
0

0
0
20
]
[A 40
15 0 1 2 3 4
T = 100
0.5

15 20
1
0
[dB]

] 0
[B 1
20
15 40
1

( )

0 1 2 3 4
displ.

vib. spectrum

T = 100 1+ 2
10 20 1
0 2 2
[C] 0
1.5

20
15 40
T = 100 0 1 2 3 4
100 20
[ B2 ] 2
0
2

0
20
freq.

100 40
rotational pulse 0 1 2 3 4
sampling time T [s] freq. [rad/s]
1

Fig. 11.18 Vibration waves and FFT display

Since = 0.04 is assumed here, l11 0:75


0:04 0:03 [ f1 0:025, the
system is unstable.
The amplitude gradually increases and x1 1 is predominant in the spectrum.
[C] The rotational speed between the rst and second critical speeds = 1.5:
306 11 Stability Problems in Rotor Systems

A time history response curve is shown in [C].


Since = 0.15 here, l12 l21 0:25
0:0625D2 0:00035 [ f1 f2
0:025
0:0125 0:000314, the system is unstable. The amplitude gradually
increases. In addition to the unbalance vibration at = 1.5, predominant peaks of
the unstable vibrations are found at x1 1 and x2 2.
[B2] Second critical speed = 2.0:
A time history response curve is shown in [B2].
Since = 0.04 here, l22 0:1875
0:04 0:0075\f2 0:0125, the system
is stable.
The amplitude is steady state and the only predominant peak is seen at x2 2.

11.3 Vibration Due to Thermal-Bow


by Contact Friction

11.3.1 Thermal-Bow

The deflection of a rotor shaft in a circular whirling motion due to unbalance is


observed as a vibration waveform measured by a displacement sensor placed at the
stationary side. It is, however, a constant value meaning static bending deformation,
if observed as a shaft stress in the rotational coordinate system. The orientation of the
cross-section of the rotor is therefore invariable under a constant rotational speed: the
outermost part of the shaft remains the same. When this part is in contact with a part
of the stationary side, friction generates heat which causes partial deformation of the
shaft, generating another eccentricity. Such a heat source is called the hot spot, and
the unbalance vibration due to the hot spot is often referred to as thermal bow
vibration. This is of course an unwanted effect and needs a countermeasure.
The thermal bow vibration in which the hot spot on the shaft surface is due to
friction is called the Newkirk effect [101103]. On the other hand, that due to the
hot spot on the bearing journal is called the Morton effect [104,105], which appears
as a result of a temperature difference in the circumferential direction on the journal
surface generated by the shear flow of the oil lm in the bearing. The temperature
difference in the shear flow is found between the thicker side and the thinner side
compared with average thickness of the oil lm, because the journal whirls around
the eccentric equilibrium point in the oil bearing. This chapter deals only with the
Newkirk effect.
The contact between the rotor and the stationary side is illustrated in Fig. 11.19
along with the location of the hot spot (HS) on the rotor cross section.
The thermal bow unbalance induced by the hot spot appears in either of two
ways as shown in Fig. 11.19: if the disk is between the bearings (gure ), the
thermal bow simply produces unbalance in the same phase as the hot spot; in the
overhang arrangement (gure ), the effect of thermal bow generates unbalance in
11.3 Vibration Due to Therma-Bow by Contact Friction 307

rub
rub

induced unbalance
HS
HS
in-phase
H >0
opposite phase
H<0

induced unbalance
1 2

Fig. 11.19 Hot spot (HS) induced unbalance

the opposite phase to that of the hot spot. These phase relations are stated in a
simplied way. They are also affected by the vibration modes. The generation
factor H of the thermal bow induced unbalance is introduced here, which indicates
in-phase unbalance when positive and out-of-phase unbalance when negative. The
dynamics of the unbalance generation can be succinctly described using a block
diagram of a rst-order delay system.

11.3.2 Thermal-Bow Model

Figure 11.20 illustrates a spinning rotor in contact with the elastic wall due to
unbalance vibration in a certain mode. The corresponding single-dof model is

mz c_z kz jlka z meX2 ejXt 11:47

where m is the modal mass, c modal damping, k is the modal stiffness plus the
stiffness of the elastic wall, ka is the stiffness of the elastic wall (clearance being

Fig. 11.20 Rotor rubbing Y


model
ka
ka HS j ka z
S

z

ka
ka

flexible casing
308 11 Stability Problems in Rotor Systems

neglected), is the friction coefcient, and is the eccentricity of the center of


gravity.
Substituting the vibrational solution

z AejXt aeju ejXt 11:48

into Eq. (11.47) gives

k jXc jlka  mX2 A meX2 11:49

This relationship is represented by the upper right part of Fig. 11.21. Converting
Eq. (11.49) to a formula for the equilibrium of forces in the rotational coordinate
system:

k jXc jlka  mX2 A meX2 0 11:50

it permits the graphical representation of the components of forces acting on the


rotor section (Fig. 11.22). The phase of the vibration displacement A lags behind
the unbalance by the phase . On the shaft center S, the unbalance force meX2 ,
spring reaction force kA, inertial force mz mAX2 , damping force jXcA
and contact friction force jlka A due to rubbing are balanced.

m 2 1 A

+ + k + j c m 2

+ thermal bend
j

H
k a
fh s + 1 fb

Fig. 11.21 Rubbing induced thermal bent model

Fig. 11.22 Force balance 2

and hot spot (HS)


m

( k m 2 ) A

S
A HS
ka
j cA

j ka A
11.3 Vibration Due to Therma-Bow by Contact Friction 309

The contact point is the hot spot where thermal bow gives rise of unbalance
force, being proportional to the contact friction force lka A. The force appears in a
rst-order delay system as a newly excited unbalance force fh :

H
fh lka A 11:51
ss 1

where H is a dimensionless constant of proportion, which is to be added to the right


side of Eq. (11.49):

H
k jXc jlka  mX2 A meX2 lka A 11:52
ss 1

Correspondingly, a thermal bow component is added to the block diagram


(Fig. 11.21).

11.3.3 Stability Analysis

Equation (11.52) gives the characteristic equation

ss 1k jXc jlka  mX2  Hlka 0 11:53

In order to apply the Routh-Hurwitz criterion, Eq. (11.53), containing the complex
coefcient, is converted by the transformation s js0 to:

A0 jB0 s0 A1 jB1 0 11:54

This gives the stability condition


 
A0 A1 slka cX k  mX2  Hlka

B0 B1 \0 ! s k  mX2  lka cX \0 11:55

k
which can be rewritten, by introducing the following parameters, x2n ,
m
c ka X
2fxn , aa , p = dimensionless rotational speed and for simplicity, as
m k xn

slaa 2fp 1  p2  Hlaa
\0 11:56
s1  p2 laa 2fp

This gives the stability condition


310 11 Stability Problems in Rotor Systems

Fig. 11.23 Stability map

in-phase
0.3
laa 0:5 30
unstable 0.2
20 0.1
=0
10
stable

H
0

10

reverse phase
stable =0
20 0.1
0.2 unstable
30
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
speed P = / n

1  p2 2 laa 2fp2
H\ p\1
laa 1  p2
11:57
1  p2 2 laa 2fp2
H[  p [ 1
laa p2  1

Calculation examples are shown in Fig. 11.23.


A large stable zone for the out-of-phase unbalance (H < 0) is seen in the rota-
tional speed range under the critical speed (p < 1); the reverse is the case in the
rotational speed range over the critical speed (p > 1). The stable zone is very
narrow around the critical speed, readily leading to instability due to contact
friction.
Introduction of damping could extend the stable zone. In order to achieve the
stability over the entire rotational speed range, it is necessary to provide appropriate
damping and small H (=thermal input/heat radiation) to lower the sensitivity of the
system to thermal bow.

11.3.4 Physical Interpretation of Stability

Physical interpretation of stability is provided by Fig. 11.24. The phase angle of the
eccentricity of the center of gravity of the rotor is led by (in the same direction
as rotation) with respect to the displacement A to the right direction of the rotor. The
angle is less than 90 for rotational speeds slower than the critical speed, and
greater than 90 for rotational speeds faster than the critical speed.
The hot spot is located at the phase of 0 (the right extreme of the rotor
cross-section). Consequently, the thermal bow induced unbalance is right-pointing
11.3 Vibration Due to Therma-Bow by Contact Friction 311

(a) + 2
(b) + 1
+1 +2

2 1 A 2 1 A
S HS S HS
H <0 H >0 H <0 H >0

low speed < n high speed n <

Fig. 11.24 Description of stability and instability

(e1 ) if in-phase and left-pointing (e2 ) if out-of-phase. Subsequent progress of the


thermal bow is as follows.
(A) Rotation slower than the critical speed (Fig. 11.24a)
In-phase unbalance (H > 0) is unstable:
e ! e e1 (unbalance increased) instability induced
this increasing unbalance moves against the rotation direction once the rotor
becomes unstable.
polar plot diverges spirally in the lagging direction.
Out-of-phase unbalance (H < 0) is stable:
e ! e e2 (unbalance decreased) stable under any condition
this decreasing unbalance moves in the rotation direction. polar plot
settles to a point in the leading direction.
(B) Rotation faster than the critical speed (Fig. 11.24b)
In-phase unbalance (H > 0) is stable:
e ! e e1 (unbalance decreased) stable under any condition
the unbalance moves against the rotation direction. polar plot settles to a
point in the lagging direction.
Out-of-phase unbalance (H < 0) is unstable:
e ! e e2 (unbalance increased) instability induced
the unbalance moves in the rotation direction once the rotor becomes
unstable.
polar plot diverges spirally in the leading direction.
These descriptions correspond with the results in Fig. 11.23.
312 11 Stability Problems in Rotor Systems

11.3.5 Simulation of Thermal-Bow Induced Vibration

The block diagram in Fig. 11.21 is a representation in the s domain after Laplace
transformation. The corresponding differential equation in the time domain is, using
dimensionless parameters,

dfh
s fh Hfb 11:58
dt

fb  fkb l kka A laa A; fh  fkh

ep2 f  jf 
where A 1p2 h 2jfpb .
Simulation results of the complex amplitude At Ax t jAy t for each case
in Table 11.1 are shown in Figs. 11.25 and 11.26.

Table 11.1 Parameters for simulation

figures speed thermal bend stability selected stability


limit value
p = / n a H H

1 Fig.1125 0.9 0.05 9.37 10 unstable


2 Fig.1125 0.9 0.05 9.37 7 stable
3 Fig.1126 1.1 0.05 11.14 8 stable

4 Fig.1126 1.1 0.05 11.14 12 unstable

note : damping ratio = 0.1 , = 1 , = 10

Fig. 11.25 Simulation Ay


(clockwise) 10
t = n phase lag

3 4.
..
. Ax
10 2 A0 10
n=0
2 A2
3 2 1
1
1
10

p = / n = 0.9 a = 0.05 = 0.1


11.3 Vibration Due to Therma-Bow by Contact Friction 313

Fig. 11.26 Simulation Ay


(counterclockwise) 10
phase lead
t = n

.4 Ax
.
10 . 10
. A0 3 2
n=0
A2 3
1 2
1 4

10
p = / n = 1.1 a = 0.05 = 0.1

(A) Fig. 11.25 shows the calculation results at the rotation lower than the critical
speed (p = 0.9). The exact stability limit of the in-phase unbalance (H > 0) is
H = 9.37 as shown in the table. Therefore, H = 10 is unstable so that the phase of
unbalance moves against the rotation direction, and the spiral locus of the polar
plot diverges clockwise. H = 7 is stable so that the spiral locus of the polar plot
settles to a point.
(B) Fig. 11.26 shows the calculation results at the rotation faster than the critical
speed (p = 1.1). The exact stability limit of the out-of-phase unbalance (H < 0) is
H = 11.14 as shown in the table. Therefore, H = 8 is stable so that the spiral
locus of the polar plot settles to a point. H = 12 is unstable so that the phase of
unbalance moves in the rotation direction, and the spiral locus of the polar plot
diverges counterclockwise.
The vibration amplitude before rotor-wall contact is, since fh fb 0,
Just before thermal-bow:

ep2
A0 11:59
1 2jfp  p2

After the contact, the polar plot settles to a point if the parameters permit stability.
Thus, the terminal amplitude approaches to the equilibrium point in thermal-bow,
since fn Hfb

ep2
A 2 A 1 11:60
1 j2fp laa  p2  Hlaa

Example 11.2 Find the equilibrium point in thermal bow for the stable cases
and using the parameters shown in Table 11.1. Conrm that the result is in
agreement with the simulation diagrams.
Answer Substitute the parameters in Table 11.1 into Eqs. (11.59) and (11.60).
fA0 ; A2 g f3:1\  43 ; 2:9\  124 g
fA0 ; A2 g f4\  137 ; 3:7\  55 g
314 11 Stability Problems in Rotor Systems

11.4 Thermal-Bow Induced Vibration of an Active


Magnetic Bearing Equipped Rotor

11.4.1 Thermal-Bow Model

The dynamics of this system is described here according to Takahashi [106].


Consider a single-dof modal model m-k for a certain eigenmode. When unbalance
force due to the modal eccentricity of the center of gravity and the force Fb due to
the active magnetic bearing (AMB) are in action, the equation of motion is for a
certain eigenmode:

mz kz /Fb meX2 ejXt 11:61

The coefcient is a constant reflecting the contribution of the force due to the
AMB, which is determined usually by the displacement in the eigenmode at the
bearing. It is assumed that a PD control law applies where the control force, Fb, of
the AMB corresponds to spring and viscous damping force:

Fb c_z kd z jkc z 11:62

where kd is the direct spring constant, kc is the cross spring constant and c is the
viscous damping factor.
Using the variables in the rotational coordination system dened by

z AejXt ; Fb fb ejXt

the unbalance vibration of the system can be written as

k  mX2 A /jXc kd jkc A meX2 11:63

The corresponding part appears in the upper right part of the block diagram shown
in Fig. 11.27. Equation (11.63) represents static balance of the forces as observed
in the rotational coordinate system. The vector components are illustrated on a rotor
cross-section in Fig. 11.28.
Since the bearing reaction force fb of the magnetic bearing is generated by the
attractive force of an electromagnet, the rotor surface facing the magnet is heated by
eddy current loss. Consequently the hot spot occurs in the fb direction; it moves under

Fig. 11.27 Block diagram m 2 A


1
g1 =
+ + k m 2
thermal bend

H
g3 = g 2 = j c + k d + jk c
fh s + 1 fb
11.4 Thermal-Bow Induced Vibration 315

Fig. 11.28 Force balance


and hot spot (HS)
k
m k
2

kd A m

( k m 2)A
HS2
S A
j c A

fb
j kc A
fa fb

HS
AMB
HS1

the particular control of the magnetic bearing, e.g. to HS1 by cross spring control
(kc [ 0; kd c 0) or to HS2 by direct spring control (kd [ 0; kc c 0).
The thermal bow induced unbalance fh is proportional to the magnetic bearing
reaction force fb, which is here assumed to be represented by the rst-order delay
system

H H
fh fb jXc kd jkc A 11:64
ss 1 ss 1

Here s is the time constant and H a constant representing the sensitivity to thermal
bow, is added to the right side of Eq. (11.63), as a negative term as well as the
bearing reaction force. The block diagram (Fig. 11.27) shows a corresponding part,
which represents the unusual state in the thermal bow induced vibration.

11.4.2 Stability Analysis

The open-loop characteristic of the thermal bow system (Fig. 11.27) is


g1
Go s g2 g3 s 11:65
1 /g1 g2

The characteristic equation is, therefore, from 1 G0 s 0,

k  mX2 1 ss /1 ss Hkd jkc jXc 0 11:66

By deriving the characteristic equation as in the previous section, and introducing


for simplicity the following parameters
316 11 Stability Problems in Rotor Systems

k c kd kc X
x2n , 2fxn , ad , ac , p (dimensionless rotational speed),
m m k k xn
the stability condition is obtained:

s1 /ad  p2 H /ac 2fp
\0 11:67
/sac 2fp 1  p2  H /ad

[1] Cross-spring control


With the damping factor c left unchanged and kd = 0, the stability condition is
obtained from Eq. (11.67):

1  p2 2 /2 ac 2fp2
H[  11:68
/ac 2fp2

This means that, as shown in Fig. 11.29a, the system is always stable with the
in-phase unbalance (H > 0), and stable above the curve with the out-of-phase
unbalance (H < 0). Since the damping is in the same tangential direction as the
cross spring c, increased damping makes the stable zone narrower. In other words,
increasing damping has negative effects on the stability in this case.
Since the cross-spring control is used around the critical speed, the stable zone
can be approximated as

H[ / for p  1 11:69

Three measures for stability improvement are possible:


H > 0: In-phase unbalance mode
H < 0: Decreasing jH j reduces sensitivity to thermal bowing

(a) (b)
6
in-phase

4 stable stable unstable


0.2
= 0.1 2
2 3 = 0
2 = 1
1 c = d = 0 , = 0.4
H

0
stable
2 d = 0.4
opposite phase

.4 3 1 =0
4 =0
c
2 2 = 0.1
unstable unstable
1 3 = 0.2
6

0 0.5 1.0 1.5 0 0.5 1.0 1.5
speed p = / n speed p = / n

Cross-spring control ( = 1 ) Direct spring control ( = 1 )

Fig. 11.29 Direct/cross-spring control


11.4 Thermal-Bow Induced Vibration 317

is recommended to be increased: It might be possible to place the magnetic


bearing at a location where the amplitude is large in the eigenmode
[2] Direct spring control
The stability condition obtained with c left as it is and kc = 0 is illustrated in
Fig. 11.29b. Addition of the direct spring kd results in a natural frequency
p
x/ xn 1 /ad . Stability features are different between the left and right side of
this frequency. Since the hot spot HS2 in Fig. 11.22 is the opposite of HS in
Fig. 11.22, the corresponding stability chart of Fig. 11.29b is the reverse of
Fig. 11.23.

11.4.3 Physical Interpretation of Stability

The physical interpretation of the situation is summarized in Table 11.2.


In the case of the cross-spring control, the hot spot HS1 in Fig. 11.28 appears in
the bottom of the rotor cross-section for a rotor displacement A located in the right

Table 11.2 Stable and unstable phenomenon

control crossing-spring direct spring

+2

+1
(1)
2
2 +
low
H<0 +1
speed A
HS
< n A 1 2
H>0 1 H>0 H<0 k d only

k c only
HS
+ 2

+2
+
(2) 1
2 H<0
high

A A
speed +1 HS
n < H>0 1 2
1 k d only H>0 H<0

k c only
HS
318 11 Stability Problems in Rotor Systems

direction. The in-phase (H > 0) and out-of-phase (H < 0) unbalance occur, there-
fore, in the lower 1 and upper half 2 of the rotor section, respectively. The system
is inherently stable on the side in which e e1 is located. The side of e e2 is
unstable.
On the other hand, in the case of direct spring control, the hot spot HS2 appears
on the left of the rotor cross-section for a rotor displacement A located in the right
direction. The in-phase (H > 0) and out-of-phase (H < 0) unbalance occur therefore
in the left 1 and right half 2 of the rotor section, respectively. The system is
likewise inherently stable on the side of e e1 in low speed rotation, and also stable
on the side of e e2 in high speed rotation.

11.4.4 Simulation of Thermal Bow Induced Vibrations

The dimensionless form of the representation in the time domain converted from
the s domain block diagram Fig. 11.27 is:

Fig. 11.30 Control by cross-spring and direct spring


11.4 Thermal-Bow Induced Vibration 319

Table 11.3 Parameter for simulating thermal bend of AMB rotor

speed control stability selected stability


figures limit value
p = / n c / d H

Fig.1130 1 2 0.9 cross-spring 0.4/0 1.22 1.1/1.3 stable/unstable


Fig. 11 30 3 4 1.3 cross-spring 0.4/0 3.97 3/4.5 stable/unstable
Fig. 11 30 5 6 0.9 direct spring 0/0.4 1.475 0.1/1.5 stable/unstable
Fig. 11 30 7 8 1.3 direct spring 0/0.4 0.725 0.8/0.5 unstable/stable

note : damping ratio =0, =1, =1, =1

dfh
s fh Hfb 11:70
dt

where fb  fb =k 2jfp ad jac A, fh  fh =k,


 
A ep2  fh = 1  p2 /2jfp ad jac :

The polar plots in Fig. 11.30 represent the complex amplitude At Ax jAy for
each case in Table 11.3.
[1] Cross-spring control: The polar plot trajectory has a spiral form.
[2] Direct spring control: Since the bearing force consists of spring force only,
the unbalance and vibration vector are in the in-phase or in the out-of-phase,
and the unbalance added by the hot spot is in the same phase as the existing
unbalance. Therefore the polar plot trajectory representing the thermal bow
moves only in the radial direction, and either settles or diverges, but is not
spiral in nature.
Note: With the relation to this chapter, further information and discussion will
be provided by references, e.g., [107112] for internal damping, [113118] for
asymmetric rotor system, and [119124] for thermal instability, though these are a
very small subset of the entire knowledge.
Chapter 12
Rotor Vibration Analysis Program:
MyROT

Abstract This chapter describes MyROT, a versatile software package for the
rotor vibration analysis of rotating machinery. The calculation is based on a
combination of beam elements according to nite element modeling , termed
1D-FEM, which is discretized by dening the mass, stiffness and damping matrices
according to the actual geometry of the rotating shaft. These matrices are applied to
calculations of natural frequencies, complex eigenvalues for stability analysis,
unbalance vibration by frequency response analysis (FRA) and so on. Specically,
this chapter describes the theoretical background for formulation and presents
outputs of typical rotor dynamic calculations in order to show how this program is
convenient in analyzing rotordynamics. A trial version of the package is available at
http://www.nda.ac.jp/cc/mech/member/osami.html. An input data manual and other
documents related to the package may also be downloaded from the site.


Keywords Software MyROT 1D-FEM Discretization   Rotor vibration anal-
 
ysis Guyan reduction Mode synthesis modeling

12.1 Data on Rotor Systems

12.1.1 Rotor Drawing and Discretization

To obtain the rotor shaft data, the rotor drawing consisting of a shaft and disks
should be prepared. First, the rotor is sub-divided into two parts contributing mainly
to the stiffness (called the shaft) and contributing to the mass (called the disks). It is
benecial to distinguish these two parts by colors, e.g., yellow and green as shown
in Fig. 12.1. It shows a rotor divided into the beam elements (yellow) of the rotating
shaft between nodal numbers of the left and right sides and the disk elements
(green) attached to the shaft at a nodal point. The former contributes to the stiffness
and mass matrices and the latter only to the mass matrix through added masses.
Node numbers are then located, from left to right, on points where the shaft section
changes. The upper part of the table in Fig. 12.1 summarizes the lengths and

Springer Japan 2017 321


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1_12
322 12 Rotor Vibration Analysis Program: MyROT

Fig. 12.1 Shaft system and sliding bearing

diameters of individual beam elements and the middle part the thicknesses and
diameters of individual added inertia disks. It is strongly recommended to tabulate
the data for individual shaft and added mass elements as shown in the gure.
Nodes and in the gure are the positions of the sliding bearings, which are
modeled using spring and the viscous damping coefcients dependent on the
rotational speed. These coefcients are listed as kij and cij (i, j = x, y) in the lower
part of the table.
Although it might seem tedious, the enumeration of input data combined with
the rotor drawing, as seen in Fig. 12.1, is very useful and important to effectively
12.1 Data on Rotor Systems 323

Fig. 12.2 Description of 4 6 8 10 11 12


node numbering

1 2 3 5 7 9 good example

6
1 2 3 4 5 poor example

7 8 9 10 11 12

complete the calculation. It is noted that there have been many cases in which
negligence of the enumeration at this stage has led to much loss of time.
Since the bending theory of beams assumes that the shaft length is larger than the
diameter, sections of shaft that are almost uniform, e.g., with diameter changes at
corners, edges and small grooves in the shaft, need not be divided into too small
sections. Although the software includes the theory of shear deformation to com-
pensate the input of very short beams, too many nodes will lead to large matrices
and therefore excessive calculation times.
The matrices in MyROT have a banded nature for the shafting part. Therefore, as
seen in Fig. 12.2, the node numbers at both ends of a shaft element should
preferably be as close as possible. The node numbers of a connecting spring ele-
ment are also recommended to be neighbors, otherwise the bandwidth of the
stiffness matrix may become too large in some cases. For example, the two-shaft
rotor of Fig. 12.2 should result in minimal bandwidth of the matrices.

12.1.2 Data Organization of a Rotor System

The input data groups for MyROT are organized as follows:


(1) Total number card: The total numbers of entry card in each (2)(8) group below.
(2) Material card: Data on used materials, such as mass density q, Youngs
modulus E, and modulus of transverse elasticity G.
(3) Shaft element card: Data on the shaft (beam) elements, such as length, outer
diameter and inner diameter.
(4) Boundary condition card: Designation of conditions, such as simple (pinned)
support, roller end and xed end.
(5) Node station card: Designation of coordinates of individual node station;
automatically assigned if not specied.
(6) Added inertia card: Designation of geometry (thickness, diameter, etc.) of
disks or other added inertia elements.
(7) Spring element card: Bearings spring coefcient kb independent of the rota-
tional speed.
324 12 Rotor Vibration Analysis Program: MyROT

(8) Damping element card: Bearings viscous damping coefcient cb independent


of the rotational speed.
(9) Sliding bearing card: Denition of bearings dynamic characteristic constants,
kij X and cij X; i; j x; y as functions of the rotational speed .
An example of the input data for the rotor of Fig. 12.1 is shown in Fig. 12.3. All
the data on individual cards are listed except the boundary condition card (4), which

Fig. 12.3 Input data


12.1 Data on Rotor Systems 325

1 ball bearing 2 sliding (oil) bearing 3 magnetic bearing (AMB)


oil sensor
electric magnet

1 2 3

G ( s)

large small

k Cb k () C b () k ( ) C b ( )
b b
b
7
small 8
large 6
small
8 10 N/m 4 10 N/m 10 N/m

(constant) (rotational speed dependency) (frequency dependency)

Fig. 12.4 Bearing modeling

is absent because no node in this system is constrained. For the input format on
each card, the reader is referred to the HELP section of the software package.
The total number card requires designation of the bandwidth of the matrices. It
can be estimated using the pair of the greatest difference of node numbers N1, N2
for the ends of shaft elements by the formula

Bandwidth Abs N1  N2 1  2:

Another way is to start with an appropriately large value of the bandwidth. Once the
input data have been loaded, the exactly calculated value of the bandwidth is
printed, which can be used subsequently.
A bearing may be represented by a spring-damper system as shown in Fig. 12.4.
The ball bearing is usually regarded as a constant spring coefcient at the input
data group (7). The sliding (oil lm) bearing is represented by the spring and
damping coefcients dependent on the rotational speed at the input data group (9),
prepared as kij X and cij X; i; j x; y. As a special option, the magnetic
bearing data are required to be input as a transfer function G(s) only before
executing AMB related calculations. For the purpose of approximate evaluation of
rotor vibration, bearings or may be input as the predetermined approximate
values of the spring and damping coefcients at bearing portions, being indepen-
dent from the rotational speed and dened at the input data groups (7) and (8).
326 12 Rotor Vibration Analysis Program: MyROT

12.2 Matrices

12.2.1 Matrices in the Original System

The following matrices are constructed using input data such as shaft dimensions
and boundary conditions:
Mass matrix: lumped mass matrix M
Stiffness matrix: Ks for the shafting, Kij (i, j = x, y) for the bearing spring
Damping matrix: Ds for material damping in non-rotating structural components, Dr
for material damping in rotating parts, and Dij (i, j = x, y) for damping coefcients
at bearings
Gyroscopic matrix: G
These matrices are either diagonal, as the case of the mass matrix, or banded
matrices as with the shaft stiffness matrix.
The equations of motion are, in terms of the displacement vectors x, y in the X, Y
directions,

Mx XG_y D_x XDr y Ks x Kxx x Kxy y Dxx x_ Dxy y_ 0


12:1
My  XG_x D_y  XDr x Ks y Kyx x Kyy y Dyx x_ Dyy y_ 0

where D = Ds + Dr.
As explained in Sect. 7.1, these can be rewritten using the complex displacement
vector z = x + jy as

Mz  jXG_z D_z  jXDr z Ks z Kf z Kbz Df z_ Dbz_ 0 12:2

where

Kf Kxx Kyy = 2  jKxy  Kyx = 2; Kb Kxx  Kyy = 2 jKxy  Kyx = 2;


Df Dxx Dyy = 2  jDxy  Dyx = 2 and Db Dxx  Dyy = 2 jDxy  DKyx = 2:

12.2.2 Reduced Matrices in the Guyan Method

Denition of the geometry of a vibrational structure to be analyzed by the nite


element method generally requires ne discretization of the object. In practice,
however, the number of the eigenvalues needed for the rotor design is often much
less than the total number of nodes. In other words, the nodes are excessively
numerous compared with the necessary node number to capture the main features of
the mode shapes of the eigenvectors employed for the rotor design. It is then
12.2 Matrices 327

practical to reduce the number of nodes meshed for vibration calculations. The
Guyan reduction method [15] is one such procedure.
MyROT permits a user to assign most important nodes in the node coordinate
card group (5) by indexing the variables D (translational Displacement of the node),
A (tilting Angle of the node) and B (Both of translational displacement and tilting
angle of the node). The indexed nodes are called the master coordinates and uns-
elected nodes the slave coordinates. Usually the displacement (D) of the nodes
provides sufcient information for the master/slave selection; A or B are used in
special models only.
Figure 12.5 shows a rotor divided into 36 nodes (72 dimensions including
translating and tilting motions). This is reduced to dimension 13 by assigning the
node numbers correspond to the bearing and disk mounts, by indexing nodes by the
shaded (red) circles, with respect to the variable D. The same procedure chooses the
shaded circled nodes for the master coordinates to reduce the size of matrices from
10 nodes 2 = 20 dimensions to dimension 7 in Fig. 12.1.
The bearing journal locations must always be chosen for the master coordinates
in MyROT.
After choosing the master and slave coordinates, z1 and z2, respectively, the
procedure described in Sect. 4.1.2 is employed: The slave coordinates are inter-
polated by the master coordinates according to
      
z1 Ks11 Ks12 z1 Keq
Ks  t z1  12:3
z2 Ks12 Ks22 z2 0

Using the transformation matrix for the interpolation:


1 t
z2 Ks22 Ks12 z1  d1 d2    z1  dz1 12:4

The entire coordinates including the slave coordinates are then dened by the
master coordinates:
   
z1 E
z   T g z1  12:5
z2 d 1

Thus, the matrices for the reduced system are


1 t
Tgt MTg ! Mg ; Tgt Ks Tg Ks11  Ks12 Ks22 Ks12  Keq ! Kg ; Tgt GTg ! Gg
12:6

As shown in Fig. 12.6, the diagonal or banded matrices are transformed to fully
populated matrices having the dimensions of the master coordinates, the process
being termed as Guyan reduction. MyROT uses this reduction method as necessary
for alleviating the burden of full order calculations.
328 12 Rotor Vibration Analysis Program: MyROT

Fig. 12.5 Finite element model (1-D FEM)


12.2 Matrices 329

coordinate
mass matrix M

master Mg
n1
..
di full n1 z1
Tg
ag
slave
on
n2
al

shaft stiffness matrix K s


Kg
ba

n1
nd

full n1 z1
n2
Tg *

possible to substitute bearing stiffness

Fig. 12.6 Guyan reduction model

The bearing nodes are chosen for the master coordinates and remain after the
transformation; the bearing stiffness or damping factor can be directly superim-
posed on the reduced matrices arising from the Guyan reduction.

12.2.3 Matrices in the Mode Synthesis Models

MyROT frequently uses the mode-synthesis model [16, 17] as a more efcient
method to generate reduced order models. The procedure is described here in a
more general manner using matrices in Sect. 4.1.3. The method is similar to the
substructure method or Craig-Bampton method in 3D structural vibration analysis.
In a rotor system, however, only the 1D shafting is treated, and therefore the
method is less complicated.
In this section, the master node is assigned more strictly. The master coordinates
z1 should always include the rotor displacement at the journal bearing portions.
Other important nodes may be included, but usually only the bearing portions are
enough. Then strictly selected master coordinates, i.e., the boundary coordinates,
might result in a reduced number of the master coordinates than in the Guyan
method of the previous Sect. 12.2.1. Now, it is assumed that displacement of all the
nodes except the boundary coordinates z1 (the bearing journals, etc.) belong to the
slave coordinates z2. Neglecting the gyroscopic effect, the equation of motion of the
conservative M-K system is considered:
       
M1 0 z1 Ks11 Ks12 Kbrg 0 z1
0 12:7
0 M2 z2 t
Ks12 Ks22 0 0 z2
330 12 Rotor Vibration Analysis Program: MyROT

where Kbrg contains the bearing stiffness elements.


The modes for mode synthesis transformation use the forced deflection modes d
and the eigen modes / of the internal system. Since the meaning of the forced
deflection mode is the same as in the Guyan method, the number of the forced
deflection modes is equal to that of the boundary coordinates. It will be the number of
the bearings m, only if the bearing portions are selected as the boundary coordinates:
1 t
d Ks22 Ks12
1 t
12:8
Keq Ks11  Ks12 Ks22 Ks12

Usually m = 2 (two straight lines for the rigid body modes), Keq = 0. The eigen
modes / of the internal system are obtained by solving the eigenvalue problem for
the slave system with the journals constrained:

x2q M2 / Ks22 / 12:9

where
 
M2 /t M2 / Diagonal    mi /ti M2 /i    ;
h i
K2 /t Ks22 / Diagonal    mi x2qi    :

The lowest n eigen modes are chosen for calculation.


These two types of mode, m + n in the total number, are used for the coordinate
transformation
      
z1 E 0 z1 z1
 Tq 12:10
z2 d / g g

The mode synthesis model is thus obtained by applying the transformation


matrix Tq. The displacement vector obtained in this model consists of the journal
displacement z1 and the modal coordinates g. The matrices in the mode synthesis
model are the congruent transform of those in the original system with the trans-
formation matrix Tq.
" #
M1 dt M2 d dt M2 /
Tqt MTq ! Mq off - diagonal matrix with non - zero elements at edges
sym: M2
 
Keq 0
Tqt Ks Tq ! Kq 
diagonal matrix or similar
0 K2
Tqt GTq ! Gq ; Tqt DTq ! Dq
12:11
12.2 Matrices 331

m
mass matrix M Mq
boundary (bearing) ..
m z1
0 number of
slave ..
n2 Tq eigenvector n
0

diagonal matrix with edge

shaft stiffness matrix K s possible to substitute Kq


bearing stiffness
m m
0 z1
0
slave
n2 Tq 0 n
0

Fig. 12.7 Mode synthesis model

As shown in Fig. 12.7, since the bearing journal displacement z1 (dened as the
boundary coordinates) remains after the transformation, the direct superposition of
bearing information, e.g., kij and cij, is then possible.

12.2.4 Discretization of Beam Elements [125]

The displacement of the nodes on an elastic round shaft element (Fig. 12.8) is
represented by four variables: deflections d1 ; d2 and tilting angles, h1 and h2 .
Representing them as complex displacements, the discrete equation of motion for
the transversal vibration of the beam element is
2 3 2 3 2 3 2 3
d1 d_ 1 d1 V1
6 h1 7 6 h_ 1 7 6 h1 7 6 M1 7
M6 7 6 7 6 7 6 7
4 d2 5  jXG4 d_ 2 5 K 4 d2 5 4 V2 5 12:12
h2 h_ 2 h2 M2

m = Al = d 2l
Fig. 12.8 Shaft element 1 2
M1 M2 4
, l, d, EI m 2
V1 V2 Ip = d
8
1 1 2 2 d2 l 2
Id = m +
16 12
332 12 Rotor Vibration Analysis Program: MyROT

where

M diagonal Dm=2 DId =2 Dm=2 DId =2 ;


   
G diagonal 0 DIp 2 0 DIp 2 ;
2 3
12 6l 12 6l
6
EI 6 6l 4l2 6l 2l2 7
7
K 36 7:
l 4 12 6l 12 6l 5
6l 2l2 6l 4l2

12.3 Analyses Corresponding to Job Commands

12.3.1 Analysis Menu

MyROT performs various shaft vibration analyses using the input data. The anal-
yses are batch commands according to the items specied through the analysis
menu. The analysis menu window in the MS-DOS and Windows/Visual Basic
versions are shown in Figs. 12.9 and 12.10, respectively.
Frequently used job commands include the following:
(1) RWDATA (Read & Write DATA): Reads the rotor data and information
needed for analysis;
Construction of matrices used in 1-D nite element (beam) model and Guyan
model or mode synthesis reduction model are prepared for other menu. This
command must be executed before any other commands.
(2) ROTPLT (ROTor conguration PLoT): Plotting the rough drawing of the
rotor from the input data.

Fig. 12.9 Menu of MyRot on DOS


12.3 Analyses Corresponding to Job Commands 333

MyROT ver 2008/12

MyROT VB ver 2008.12 O.Matsushita

c: [IBM_PRELOAD]
2 : ROTPLT 6 : DELCTG

C:
Exec .out plot Exec .out plot
H9BADATA
MyROT Kato01SIdat
DATA KIKU001.DAT 3 : FRESR1 7 : BRGEIG
DEBUG KIKU001g.DAT
KIKU001gSIDAT
KIKU001SIDAT Exec .out plot Exec .out plot

C:MyROTDATADEBUGH9BAT.DAT 4 : FRESRG 8 : UNBLNCa


0 : Edit Input Data Exec .out plot Exec .out plot
MS-EXCEL Editor
5 : CRTMAP SM : SubMenu
1 : RWDATA Exec .out plot
Torsional Free Vib.
Quasi-Modal Model
VB Exec ms Model.out others

DOS Exec Mathematica 1m END HELP

Fig. 12.10 Menu-windows of MyROT

(3) FRESR1 (FREe vibration of Symmetric Rotor with isotropic support, type =
1): Natural frequency analysis of non-rotating system, i.e., without gyro-
scopic effects and with symmetric bearings supports Kxx = Kyy from the
viewpoint of the averaging bearing stiffness values with no damping:

x2n M/ Ks Kxx /

(4) FRESRG (FREe vibration of Symmetric Rotor with General support form,
type = 6): Complex eigenvalue analysis on an entire system dened by the
input data (1)(9) at a specied rotational speed using the Guyan method,
considering all the parameters including the gyroscopic effect and the eight
constants of the oil-lm bearings without the approximation.
(5) CRTMAP (CRiTical speed MAP): Showing the behavior of the critical speed
when the bearing spring constant changes as the parameter.
(6) DFLCTG (DeFLeCTion generated by Gravity): Showing the deflection curve
of the shaft by its own weight. The baseline is the lines between both bearing
centers.
(7) BRGEIG (oil BeaRinG parameters + EIGenvalue analysis): Continuous
analysis of complex eigenvalues in a certain range of rotational speed with
consideration of an entire input condition related to oil bearing data and AMB
data. Mode synthesis model and the tracking method for the complex eigen-
values are applied.
(8) UNBLINC (UNBaLaNCe response analysis): Plotting unbalance resonance
curves for the specied unbalance distribution.
See HELP for other analysis commands.
334 12 Rotor Vibration Analysis Program: MyROT

12.3.2 Analysis Examples

This section presents screenshots of several types of analyses frequently used.


(1) Command: ROTPLT (Fig. 12.11): Rotor drawing
A rotor drawing is prepared using the input data. This serves also as a check of
the input data. The data are simultaneously output in a text format, and can be
exported to spreadsheets, e.g., Excel, to allow a user to draw own plots.
(2) Command: FRESR1 (Fig. 12.12): Natural frequencies and eigenmodes for an
undamped M-K system
The natural frequencies and eigenmodes at rest (0 rpm) are drawn to show the
basic characteristics of the rotor.
(3) Command: FRESRG (Fig. 12.13): Natural frequencies and eigenmodes of a
damped system
The complex eigenvalues of the rotor-shaft system at the specied rotational
speed are calculated to evaluate the stability and stability margin of the system. This
is a general solution that takes the bearing dynamic characteristics and other all
parameters for bearings, but does not support analysis involving magnetic bearings.
The animation of damped mode is provided in 3D form.
(4) Command: CRTMAP (Fig. 12.14): Critical speed map for an undamped
system

ROTPLT/D Drawing

ROTPLT
100
Shaft diameter (mm)

80
60
40
20
0
-20
-40
-60
-80
-100
0 200 400 600 800 1000
Shaft Length (mm)

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.11 ROTPLT (drawing of rotor shafting)


12.3 Analyses Corresponding to Job Commands 335

FRESR1/D Drawing

FRESR1
1.0

0.5
1
Deflection

0
5 3
-0.5 4
2
-1.0
0 200 400 600 800 1000
Shaft Length (mm)

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.12 FRESR1 (eigenmode of M-K system)

FRESRG/D Drawing
Speed = 10 Hz
FRESRG Eigenvalues
Y No.1 -0.499 +j 5.33 Hz
2
No.2 -0.498 +j 5.33 Hz
1
No.3 -0.0028 +j -46.8 Hz
X No.4 -0.00352 +j 47.3 Hz
-2 -1 0 1 2
-1 No.5 -0.00162 +j -177 Hz
4 No.6 -0.00169 +j 179 Hz
-2
No.7 -0.000355 +j -366 Hz
500
No.8 -0.000329 +j 370 Hz

3
1000
Animation Length(mm)

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.13 FRESRG (complex eigenvalue analysis)

This command gives basic information for rotor design. The map along with the
bearing stiffness as the abscissa axis permits approximations of the critical speed as
the axis of the ordinate. The left side is close to free-free and the right pinned-pined
as the boundary condition at the bearing portions.
336 12 Rotor Vibration Analysis Program: MyROT

CRTMAP/D Drawing

CRTMAP
1000
Natural Frequency (Hz)

100

10

1
100 1000 10000 100000
Spring Constant ( N/mm = KN/m )
1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.14 CRTMAP (critical speed map)

UNBLNCa/D Drawing

UNBLNCa
50 0
90

(deg.)
40 180
Amplitude (m)

270
30 360

20

10

0
0 10 20 30 40 50 60 70 80 90 100
Rotational Speed (rps)

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.15 Bode plot of UNBLNC (unbalance response)

(5) Command: UNBLNC: Unbalance response


The Bode plot (resonance curve, Fig. 12.15) or Nyquist plot (polar plot,
Fig. 12.16) of an unbalance vibration response is drawn for a given unbalance
distribution. The plot can be selected in the EDIT screen.
12.3 Analyses Corresponding to Job Commands 337

UNBLNCa/D Drawing

UNBLNCa
90

180 0 10 20 30 40 50
Amplitude

270

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.16 Polar plot of UNBLNC (unbalance response)

BRGEIG/D Drawing

BRGEIG XY if flag=5, and vs speed


60 1.0
q3 q2
Frequency (Hz)

Damping Ratio
0.5
40 q4 q1
2 0.1

20 1
3 0.05
4
0
0
0 10 20 30 40 50 60 70 80 90 100
Rotational Speed (rps) Gauss

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.17 BRGEIG (damping ratio and natural frequency versus speed)

(6) Command: BRGEIG: Calculation of complex eigenvalues


The complex eigenvalues are calculated in a specied rotational speed range
with all the bearing parameters taken into account, including the dynamic charac-
teristics of oil-lm bearing or the transfer function of a magnetic bearing.
Figure 12.17 shows the dependence of complex eigenvalues on the rotational speed
and Fig. 12.18 presents the root locus.
338 12 Rotor Vibration Analysis Program: MyROT

BRGEIG/D Drawing

BRGEIG if flag=5 Im
400
3 1
300
4
2
200

100

Re
-50 -40 -30 -20 -10 0 10 XY

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.18 BRGEIG (root locus with parameter of speed)

DFLCTG/D Drawing

DFLCTG
100
Shaft (mm)

-100
0 200 400 600 800 1000
Deflection(m)

200
100
0
-100
-200
0 200 400 600 800 1000
Shaft Length (mm)

1:File 2:Edit 3:Plot 4:Print 5:Default 6:Quit

Fig. 12.19 DFLCTG (deflection due to gravity)

(7) Command: DFLCTG (Fig. 12.19): Static deflection by gravity


Deflection curve of the rotor by its own weight is drawn as seen from the bearing
center line.
12.3 Analyses Corresponding to Job Commands 339

ROTPLT/E Editing rotor model


Length Division pitch Chart total length
(mm) (mm) (cm)
X axis : 0 1062 100 20
X title : Shaft Length (mm)
Length Division pitch Chart total length
(mm) (mm) (cm)
Y axis : dflctg -0.2 0.2 0.1 6
Y title : Deflection (mm)
Y axis : rotplt 1. Both
-100 100 100 6
2. Upper
Y title : Shaft Length (mm) 1
NOTE! : mm : Physical length cm : Chart length
Title : DFLCTG O.K
File comment : KIKUCH 2-0.6-0.010 35.4Hz 88 RPS Cancel

Fig. 12.20 EDIT (editing the plotting parameters)

12.3.3 EDIT Screen

All scales of the abscissa and ordinate in the screens as shown above can be
adjusted by selecting parameters in the EDIT screen as shown in Fig. 12.20.
Note: As rotordynamics software, references [126129] are commercially
available as seen in each Website. A very limited number of references [130135] are
here listed as the software application provides further information and discussion.
Appendices

Appendix A: Approximate Modal Equation of Motion

The equation of motion for a multi-degree-of-freedom rotor system is generally


written in a matrix form:

M Z  jXGZ_ DZ_ KZ eFZ; Z


_ 0 A:1

where M is the mass matrix, D the damping matrix, K the stiffness matrix, G the
gyroscopic matrix, X the rotational speed, e F nonlinear force, etc. (where << 1 is
small factor) and Z the displacement vector.

Appendix A.1: Non-rotating System (XG 0)


with Small Damping D

In the eigenvalue problem for the basic conservative M-K structural system:

x2n M/ K/ A:2

the coefcient matrices M and K are positive denite and symmetric matrices.
Therefore, the solution can be written as

Eigenvalues : x2n ! x21 x22 . . . positive real numbers


Eigenvectors : / ! /1 /2 . . . real vectors

The eigenvectors are orthogonal through the coefcient matrices as follows


diagonal:

Ut MU ! diagonal; Ut KU ! diagonal A:3

where U /1 /2 . . .  is the mode matrix.

Springer Japan 2017 341


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1
342 Appendices

With the transformation to the modal coordinate g that corresponds to the natural
frequency xn and eigen mode /:

Z /g; A:4

substituting Eq. (A.4) into Eq. (A.1) and multiplying by t from the left approxi-
mately yields the modal equation of motion.

/t M/g /t D/g_ /t K/g /t eF/g; jxn /g 0 A:5

Appendix A.2: Non-rotating System (XG 0)


with Damping D That Is Not Small

Equation (A.1) can be rewritten as the state equation with variables Zd =


Z (displacement) and Zv = dZ/dt (velocity):
       
M 0 Z_ v D K Zv eFZd ; Zv
A:6
0 K Z_ d K 0 Zd 0

Regarding the basic system as an M-D-K system, the eigenvalue problem is:
     
M 0 /v D K /v
k
0 K /d K 0 /d A:7
B A : coefficient matrices

The solutions are

Eigenvalues : k ! k1 k2 . . . complex numbers


     
/v k1 /1 k2 / 2
Eigenvectors : ! . . . complex vectors
/d /1 /2

Since the coefcient matrices B and A are real symmetric matrices, the eigen-
vectors are orthogonal through the coefcient matrix as follows:

Ut BU ! diagonal; Ut AU ! diagonal A:8


    
k1 /1 k2 / 2
where U     is the mode matrix.
/1 /2
With the transformation to the modal coordinate g that corresponds to the natural
frequency k and eigen mode displacement /,
Appendices 343

  
Zv k/
g; A:9
Zd /

substituting Eq. (A.9) into Eq. (A.6) and multiplying by k/t /t  from the left:

k2 /t M/g_ /t M/g_ k2 /t D/g 2k/t K/g /t eF/g; k/g

Considering that

g_ kg; A:10

the approximate modal equation of motion is:

/t M/g /t D/g_ /t K/g /t eF/g; k/g 0 A:11

Thus, the equations of motion (A.5) and (A.11) in the modal coordinate system
are identical regardless of damping. It should be noted, however, that the modal
parameter values (modal mass, modal damping, modal stiffness) are real in the
former case and complex in the latter.

Appendix A.3: Rotating System with Small Damping D

As well as Eq. (A.6), Eq. (A.1) is rewritten as the state equation, with the basic
conservative M-G-K rotating system:
       
M 0 Z_ v jXG K Zv DZv eFZd ; Zv
 A:12
0 K Z_ d K 0 Zd 0

The eigenvalue problem for the basic system is:


     
M 0 /v jXG K /v
k
0 K /d K 0 /d A:13
B Bt A At : coefficient matrices

Since B is positive denite symmetric matrix and A is skew Hermitian matrix,


the solutions are:

Eigenvalues : k jx ! jx1 jx2 . . . purely imaginary numbers


     
/v jx1 /1 jx2 /2
Eigenvectors : ! . . . complex vectors
/d /1 /2
344 Appendices

The eigenvectors are orthogonal through the coefcient matrix as follows:

 t BU ! diagonal; U
U  t AU ! diagonal A:14
    
jx1 /1 jx2 /2
where U     is the mode matrix.
/1 /2
With the transformation to the modal coordinate g that corresponds to the natural
frequency jxn and eigen mode displacement /
   
Zv jxn /
g; A:15
Zd /
 
t /
substituting Eq. (A.15) into Eq. (A.12) and multiplying by jxn /  t from the
left, and manipulating as in Eq. (A.10) yields the approximate modal equation of
motion.

 t M/g  jX/
/  t G/g_ /
 t D/g_ /
 t K/g /
 t eF/g; jxn /g 0 A:16

Appendix B: Effects of Non-conservative Parameters

Effects of the non-conservative parameters kc, cd and cc on vibration characteristics


can be evaluated by approximation using the asymptotic expansion.
The equation of motion is divided into the conservative and non-conservative
parts:
mz  jXG_z kd z cd  jcc _z jkc z B:1

The right side is assumed to be small quantitative parameters. To evaluate its


effects, the asymptotic solution is written as
z aeju B:2

More specically, assuming that the amplitude and phase are slowly varying
variables characterized with << 1.

da du
eAa; x eBa
dt dt

where x xd 1  cp=2, then eA and eB can be determined. The forward and
backward natural frequencies obtained by Eq. (7.8) are thus used as the basic
natural frequency x. The velocity and acceleration are obtained by differentiating
the displacement, and retaining the small e by the rst order for simplicity,
Appendices 345

z_ jxaeju eA jeBaeju B:3

z x2 aeju 2jxeA jeBaeju B:4

Substituting Eqs. (B.2)(B.4) into Eq. (B.1) and equating the real and imagi-
nary parts of the coefcient of eju in the left and right sides, with the dimensionless
parameters given in Eq. (7.5), yields
  
x kc xd l xd 
a_ eA cd a fd c a
2mx  GX x 1  cp=2xd =x 2 x
 cp xd  l xd 
xd 1 fd c a
2 x 2 x
xcc xd
u_ x eB x x f
2mx  GX 1  cp=2xd =x c
B:5

These equations are applied to the forward whirling mode if x xf  xd [ 0


and the backward whirling mode if x xb  xd \0. They are converted to
the complex eigenvalue k a jq:

a_ eA  cp xd  l xd
Real part : a    xd fd 1 xd c
a a 2 x 2 x B:6
Imaginary part : q  u_  x eB x xd fc

Further, the dimensionless complex eigenvalue /d is obtained in the manner of


Eq. (7.10):
a  cp xd  lc xd  cp lc
Real part :  fd 1  fd 1  
xd 2 x 2 x 2 2
q x cc  cp  B:7
Imaginary part :  1 fc
xd xd 2m 2

k  cp lc  cp
Therefore; fd 1   j 1 jfc B:8
xd 2 2 2

Equation (B.8) agrees with the approximate solution Eq. (7.15).


References

R1E References (JSME v_BASE Data Book)1

VB117. Shaft vibration of rotary compressors equipped for air conditioners


VB245. Modal balancing of AMB equipped flexible rotors by feedforward excitation
VB218. Vibrations due to ball pass frequencies that occurred in the spindle of a hard disc drive
VB055. Vibration due to the non-axisymmetric stiffness of a rotating thin disk in a fan
VB958. Vibration resonance of impellers in a centrifugal compressor
VB038. Self-excited vibration induced by gear coupling
VB268. Self-excited vibration of wire-winding-type growing equipment for silicon single crystal
manufacture

R1E References (Papers)

1. Kanemitsu Y, Azuma T, Takahashi N, Fukushima Y, Matsushita O (2006) Magnetic


bearing guidebook for rotating machine designers. Touka-Shobo, ISBN: 4-88757-123-2
2. Shiraki K (1972) Troubleshooting of vibration problems in the eld. J JSME 75(639):507
524 (in Japanese)
3. JSME v_BASE Committee, v_BASE Databook (1994), v_BASE Database in CD 2nd edn
(2002), v_BASE Database in CD 3rd edn (2011)
4. Tanaka M (1998) Rotordynamic problems experienced in actual machinescase studies
collected by JSME v-Base committee. In: Proceedings of the 5th IFToMM international
conference on rotor dynamics, Darmstadt, pp 6882
5. Tanaka M (2012) JSME v_BASE activities and some troubleshooting case studies of
rotating machinery, Proceedings of IMechE VIRM10, London, C1326-026
6. Tanaka M (2014) Designing and operating rotating machinery free from vibration troubles
based on case studies. In: Proceedings of the 10th IFToMM international conference on
rotor dynamics, Milan
7. ISO 2041:2009 Mechanical vibration, shock and condition monitoringvocabulary
8. Ishida Y (ed) (2013) Special issue: reviews of Japans rotordynamics. J System Design and
Dynamics (JSDD) JSME 7(2)
9. The Japan Society of Mechanical Engineer (2014) Mechanical engineers handbook
DVD-ROM
10. ISO 10814:1996 Mechanical vibrationsusceptibility and sensitivity of machines to
unbalance
11. American Petroleum Institute (API) Recommended practice 684 (2005) Lateral critical
speeds, unbalance response, stability, train torsionals, and rotor balancing, 2nd edn; API
Standard 617 (2002) Axial and centrifugal compressors and expander-compressors for
petroleum, chemical and gas industry services, 7th edn. API, Washington D.C.
12. Seto K, Mitsuta S (1992) A new method for making a reduced-order model of flexible
structures using unobservability and uncontrollability and its application to vibration
control, 1st MOVIC, Yokohama, pp 152158
13. Nagamatsu A, Ookuma M (1981) Analysis of vibration by component mode synthesis
method : Part 1 natural frequency and natural mode (I). Bull JSME 24(194):14481453

1
http://www.jsme.or.jp/dmc/Links/vbase/data_english.html

Springer Japan 2017 347


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1
348 References

14. Nagamatsu A (1995) Modal analysis, Baifukan (in Japanese)


15. Guyan RJ (1965) Reduction of stiffness and mass matrices. AIAA J 3(2):380
16. Kaga M, Kikuchi K, Matsushita O, Takagi M, Furudono M (1982) Quasi-modal nonlinear
analysis for seismic response in pump rotor, In: IFToMM International conference, Rome,
Italy, pp 8394
17. Matsushita O, Bleuler H, Sugaya T (1988) Modeling fro flexible mechanical systems. In:
Proceedings of IMACS symposium on system modeling and simulation, pp 3338
18. Craig RR Jr, Bampton MCC (1968) Coupling of substructure for dynamic analysis. AIAA J
6(7):13131319
19. Craig RRJr (1977) Methods of component mode synthesis. Shock Vib Digest 9:310
20. Nelson HD, Meacham WL, Fleming DP, Kascak AF (1982) Nonlinear analysis of rotor-
bearing systems using component mode synthesis In: Proceedings of ASME, 82-GT-303
21. Nelson HD et al (1988) Nonlinear analysis of rotor-bearing systems using component mode
synthesis. Trans ASME J Eng Power 105(3):606614
22. Matsushita O, Takagi M, Takahashi R (1983) Vibration analysis of rotor-bearing systems
by quasi-modal transformation (analysis of complex eigenvalue and response history).
Bull JSME 26(213):414423
23. Matsushita O, Ida M, Takahashi R (1984) Application of quasi-modal concept to rotational
ratio response analysis and new balancing. IMechE C319(84):427437
24. Fan GW, Nelson HD, Mignolet MP (1993) Optimal output feedback control of asymmetric
systems using complex modes. Trans ASME J Eng. Gas Turbines Power 115(2):307313
25. Kobayashi M, Aoyama S (2011) Transient response analysis of a high speed rotor using
constraint real mode synthesis. J Environ Eng JSME 523541
26. ISO 1940-1:2003 Mechanical vibrationbalance quality requirements for rotors in a
constant (rigid) statePart 1: specication and verication of balance tolerances
27. ISO 11342:1998 Mechanical vibrationmethods and criteria for the mechanical balancing
of flexible rotors
28. Kellenberger W (1972) Should a flexible rotor be balanced in N or (N+2) planes?
Trans ASME J Eng Ind 94(2):548559
29. Kobayashi M, Saito S, Yamauchi S (1991) Nonlinear steady-state rotordynamic analysis
using transfer coefcient method. Rotating machinery and vehicle dynamics ASME,
DE-vol 35, pp 5158
30. Matsushita O, Yoshida T (1993) Open loop balancing to universal balancing specied by
the bearing condition-free method. In: Proceedings of Asia-Pacic vibration conference,
pp 499504
31. Goodman TP (1964) A least-squares method for computing balance corrections.
Trans ASME J Eng Ind 86(3):273279
32. Federn K (1968) Multi-plane balancing of elastic rotors: fundamental theories and practical
application. General electric technical information series. No. 85GL121, GE Corporation,
Faireld, CT, USA
33. Bishop RED, Parkinson AG (1972) On the use of balancing machines for flexible rotors.
J Eng Ind 94(2):561
34. Nelson HD, McVaugh JM (1976) The dynamics of rotor-bearing system using nite
elements. ASME J Eng Ind 98:593600
35. Saito S, Azuma T (1983) Balancing of flexible rotors by the complex modal method. J Vib
Acoust 105(1):94100. doi:10.1115/1.3269075
36. Kanemitsu Y, Ohsawa M, Watanabe K (1990) Real time balancing of a flexible rotor
supported by magnetic bearing. IFToMM, pp 263268
37. Parkinson AG (1991) Balancing of rotating machinery. J Mech Eng Sci 205(C1):5366
38. Tan SG, Wang XX (1993) A theoretical introduction to low speed balancing of flexible
rotors: unication and development of modal balancing and influence coefcient
techniques. J Sound Vib 168(3):385394
References 349

39. Lum K-Y, Coppola VT, Bernstein DS (1998) Adaptive virtual auto balancing for a rigid
rotor with unknown mass imbalance supported by magnetic bearings. Trans ASME J Vib
Acoust 120(2):557570
40. Matsushita O, Hisanaga Y, Saitoh S (1996) Evaluation of unbalance resonance vibration
control methods for active magnetic bearing equipped flexible rotor. In: IMechE VIRM6
C500/024/96, pp 461470
41. Kreuzinger-Janik T, Irretier H (2000) Experimental modal analysis a tool for unbalance
identication of rotating machines. Int J Rotating Mach 6(1):1118
42. Chen PY, Feng N, Hahn EJ (2001) Flexible rotor balancing taking bearing non-linearity
into account. In: Proceedings of design engineering technical conference and computer
information in engineering conference ASME, DETC /VIB-21380
43. Sinha JK, Friswell MI, Lees AW (2002) The identication of the unbalance and the
foundation model of a flexible rotating machine from a single rundown. Mech Syst Signal
Proc 16:255271
44. El-Shafei A, El-Kabbany AS, Younan AA (2004) Rotor balancing without trial weights.
Trans ASME 126:604609
45. Horvath R, Flowers GT, Fausz J (2008) Passive balancing of rotor systems using pendulum
balancers. Trans ASME J Vib Acoust 130:041011-111
46. DeSmidt HA (2009) Imbalance vibration suppression of a supercritical shaft via an
automatic balancing device. Trans ASME J Vib Acoust 131:0410011041001-13
47. Pennacchi P, Vania A, Chatterton S (2009) Application of robust regression methods to
rotor balancing using high breakdown point and bounded-influence estimators. In:
Proceedings of ASME, IDETC/CIE 2009-86475, pp 10331042
48. Murthy R, Tomei JC, Wang XQ, Mignolet M, El-Shafei A (2013) Nonparametric stochastic
modeling of structural uncertainty in rotordynamics: unbalance and balancing aspects. In:
Proceedings of ASME turbo expo 2013: turbine technical conference and exposition,
Texas, USA, GT2013-95779
49. Fang J, Wang Y, Han B, Zheng S (2013) Field balancing of magnetically levitated rotors
without trial weights. Sensors 13(12):1600016022. doi: 10.3390/s131216000
50. Kanki H, Adachi K (2013) Reviews of Japans rotordynamics development balancing.
JSDD, JSME 7(2):170177
51. Yu J (2014) The necessity of a third balance plane for generator rotor eld balancing. In:
Proceedings of ASME turbo expo, Dsseldorf, Germany, GT2014-25706
52. Matsushita O, Takagi M, Kikuchi K (1984) Analysis of rotor vibration excited by seismic
wave. Bull JSME 27(224):278288
53. ISO 14839-4:2012 Mechanical vibrationvibration of rotating machinery equipped with
active magnetic bearingspart 4: technical guidelines
54. Keogh PS (2012) Contact dynamic phenomena in rotating machine: active/passive
considerations. Mech Syst Signal Process 29:1933
55. Cade IS, Sahinkaya MN, Burrow CR, Keogh PS (2008) On the design of an active auxiliary
bearing for rotor/magnetic bearing systems. In: Proceedings of 11th international
symposium on magnetic bearings, Nara, Japan
56. Ulbrich H, Ginzinger L (2006) Stabilization of a rubbing rotor using a robust feedback
control, Paper-ID:306. In: 7th IFToMM conference rotor dynamics, Viena, Autria
57. Wegener G, Markert R, Pothmann K (1998) Steady-state analysis of multi-disk and
continuous rotors in retainer bearings. In: Proceedings of 5th IFToMM conference on rotor
dynamics, Darmstadt, Germany, pp 816826
58. Yanabe S, Kaneko S, Kanemitsu Y, Tomi N, Sugiyama K, (1998) Rotor vibration due to
collision with annular guard during passage through critical speed. ASME J Vib Acoust
120:544
59. Matsushita O, Sonoda T, Ohta H, Naruse J, Inoue Y, Koromegawa I (1986) Vibration
analysis of disk-spindle due to ball bearing excitation. Bull JSME 29(256):35103519
350 References

60. Kurohashi M, Iwatsubo T, Kawai R, Fujikawa T (1983) Energy evaluation of stability in


rotor-bearing system (Consideration of each element to logarithmic decrement). Bull JSME
26:276282
61. Jeffcott H (1919) The lateral vibration of loaded shafts in the neighborhood of a whirling
speed-the effect of want of balance. Phil Mag 37(6):304314
62. Toward R (2008) Theory and practice of rotor dynamics (http://nptel.ac.in/courses/
112103024/3), and A breif history and state of the art of rotor dynamics (http://www.iitg.
ernet.in/engfac/rtiwari/resume/rtiwari01.pdf#search='A+Brief+History+and+state+of+the
+Art+of+Rotor+Dynamics'), NPTEL, IIT Guwahati
63. Saito S, Someya T (1984) Study of damped critical speeds and damping ratios of flexible
rotors. J Vib Acoust 106(1):6271. doi:10.1115/1.3269157
64. Balda M (1974) Dynamic properties of turboset rotors. Dynamics of rotors, IUTAM,
Springer, p 30
65. Kikuchi K (1970) Analysis of unbalance vibration of rotating shaft system with many
bearings and disks. Bull JSME 13(61):864
66. Kikuchi K, Kobayashi S (1977) Stability analysis of a rotating shaft system with many
bearings and disks. Bull JSME 20(150):1592
67. Matsushita O, Fujiwara H, Ito M (2004) A Tracking solver using sliding mode control to
obtain eigensolutions of rotor-bearing systems. In: IMechE VIRM8, pp 111120
68. Kirk RG, Gergman LA (2003) Special issue: the contributions of Jorgen W. Lund to rotor
dynamics. ASME J Vib Acoust 125:14
69. Newkirk BL, Taylor HD (1925) Shaft whipping due to oil action in journal bearings.
General Electric Review
70. Burrows CR, Sahinkaya MN (1982) Parameter estimation of multi-mode rotor-bearing
systems. Proc R Soc Lond A Math Phys Sci 379:367387
71. Guinzburg A, Brennew CE, Acosta AJ, Caughey TK (1992) The effect of inlet swirl on the
rotrodynamics shroud forces in a centrifugal pump. Trans ASME, 92-GT-126, 8 pp
72. Holmes R, Skyes JE (1996) The vibration of an aero-engine rotor incorporating two
squeeze-lm damper. Int Mech Eng part G, 210(1), pp 3951
73. Kirk RG (2000) Design guidelines for improved rotating machinery stability. In: IFToMM
conference on rotor dyanmics, Sydney, pp 319
74. Siew CC, Hill M, Holmes R, Brennan MJ (2002) Reduced-order modelling methods to
assess the overall vibration response of a flexible rotorsqueeze lm damper bearing
assembly. Jl Mech Eng Sci 216(C):10811097
75. Ertas BH, Vance JM (2006) Rotordynamic force coefcients for a new damper seal design.
In: 7th IFToMM conference rotor dynamics, and ASME J Tribol 129:365374, April 2007
76. Hunter CC, Maslen EH, Barrett LE (2008) Damping ratio estimation techniques for
rotordynamic stability measurements. ASME J Eng Gas Turbines Power 131:012504-111
77. Tonoli A, Amati N, Silvagni M (2010) Vibration control1. Electromechanical dampers
for vibration control of structures and rotors. Intech, pp 132
78. El-Shafei A, Dimitri AS, Saqr T, El-Hakim M (2014) Test rig characterization and dynamic
testing of a smart, electro-magnetic actuator journal integrated bearing. In: 9th IFToMM
conference on rotor dynamics, Milan, pp 11451156
79. ISO 14839-3:2006 Mechanical vibrationvibration of rotating machinery equipped with
active magnetic bearingspart 3: evaluation of stability margin
80. ONO SOKKI (1993) Portable dual channel FFT analyzer CF-350/360, Instruction manual
81. Hagiwara N, Kikuchi K, Shiinoki K (1981) Resonance of forwards and backwards
vibration mode of impeller with gyroscopic effect, 81-0524-CP, AIAA/ASME/ASCE/AHS,
22nd structures. In: Structural dynamics and materials conference, Atlanta, Georgia
82. Michimura S, Nagamatsu A (1981) A vibration analysis of an impeller: use of the cyclic
symmetry method. Theoretical and Applied Mechanics, University of Tokyo Press, vol 30,
pp 313321
References 351

83. Michimura S, Nagamatsu A, Hagiwara N, Kikuchi K (1984) Vibration of impellers (7th


report, vibration analysis at moving coordinates xed in rotating impellers). Bull JSME 27
(231):19962001
84. Nikolic M, Ewins DJ, Petrov EP, Maio DD (2006) The effect of Coriolis forces on vibration
properties of bladed discs. In: 7th IFToMM conference on rotor dynamics
85. Campbell W (1924) Protection of steam turbine disk wheels from axial vibration.
Trans ASME 46:31160
86. Kubota Y, Suzuki T, Tomita H, Nagafugi T, Okamura C (1983) Vibration of rotating
bladed disc excited by stationary distributed forces. Bull JSME 26(221):19521957
87. Ohashi H (1994) Case study of pump failure due to rotor-stator interaction. Int J Rotating
Mach 1(1):5360
88. Chivens DR, Nelson HD (1975) The natural frequencies and critical speeds of a rotating,
flexible shaft-disk system. Trans ASME J Eng Ind 97:881886
89. Bladh R, Castanier MP, Pierre C (2001) Component-mode-based reduced order modeling
tecniques for mistuned bladed diskspart I: theoretical models. Trans ASME J Eng Gas
Turbines Power 123:8999
90. Al-Bedoor BO (2001) Reduced-order nonlinear dynamic model of coupled shaft-torsional
and blade-bending vibrations in rotors. Trans ASME J Eng Gas Turbines Power 123:8288
91. Rieger NF (2002) Progress with the solution of vibration problems of steam turbine blades.
In: 6th IFToMM conference on rotor dynamics, Keynote pp 2737
92. Christensen RH, Santos IF (2004) A study of active rotor-blade vibration control using
electro-magnetic actuationpart 1: theory. In: Proceedings of ASME turbo expo 2004,
GT2004-53509
93. Petrov EP (2004) A method for use of cyclic symmetry properties in analysis of nonlinear
multiharmonic vibrations of bladed disks. Trans ASME J Turbomach 126:175183
94. Genta G, Silvagni M (2005) Some considerations on cyclic symmetry in rotordynamics. In:
ISCORMA-3, Cleveland, Sept 2005, Paper No. 509
95. Adams ML Jr. (2006) On the use of rotor vibration analysis and measurement tools to cure
power plant machinery vibration. In: 7th IFToMM conference rotor dynamics
96. Castro HF, Cavalca KL, Nordmann R (2006) Rotor-bearing system instabilities considering
a non-linear hydrodynamic model. In: 7th IFToMM conference rotor dynamics, Paper
No. 136
97. Batailly A, Legrand M, Cartraud P, Pierre C, Lombard J-P (2007) Study of component
mode synthesis methods in a rotor-stator interaction case. In: Proceedings of ASME,
DETC-34781, pp 12351242
98. Kaneko Yasutomo (2013) Progress of technology on mechanical design of steam turbine
blade in Japan. JSDD, JSME 7(2):8094
99. Childs DW, McLean JE, Zhang M, Arthur SP (2014) Rotordynamic performance of a
negative-swirl brake for a tooth-on-stator labyrinth seal. In: Proceedings of ASME turbo
expo 2014, Dsseldorf, Germany, GT2014-25577
100. Ota H, Mizutani K, Fujita T (l989) Self-excited vibrations of wire-winding-type growing
apparatus for silicon single crystals (experimental results and considerations relevant to
self-excited vibrations). JSME Int J, Series III 32(4):554559
101. Newkirk BL (1926) Shaft rubbing. Mech Eng 48(8):830
102. Dimarogonas AD (1973) Newkirk effect: thermally induced dynamic instability of
high-speed rotors. ASME paper 73-GT-26
103. Kellenberger W (1980) Spiral vibrations due to the seal rings in turbogenerators thermally
induced interaction between rotor and stator. J Mech Des 102(1980):177184
104. Keogh PS, Morton PG (1993) Journal bearing differential heating evaluation with influence
on rotor dynamic behavior. In: Proceedings of Royal Society London, Series A, vol 441,
p 527
352 References

105. Keogh PS, Morton PG (1994) The dynamic nature of rotor thermal bending due to unsteady
lubricant shearing within bearing. In: Proceedings of Royal Society London, Series A, vol
445, pp 273290
106. Takahashi N, Kaneko S (2013) Thermal instability in a magnetically levitated doubly
overhung rotor. J Sound Vib 332(5), pp 11831203
107. Nagaya K, Kojima H, Karube Y, Kibayashi H (1984) Braking forces and damping
coefcients of eddy-current brakes consisting of cylindrical magnets and plate conductors
of arbitrary shape. IEEE Trans Magn 20(6):21362145
108. Morii S, Nagai N, Katayama K, Sueoka A (1996) Stability analysis of rotors operating in
magnetic eld. In: IMechE VIRM6, London, pp 331340
109. Kligerman Y, Grushkevich A, Darlow MS (1998) Analytical and experimental evaluation
of instability in rotordynamics system with electromagnetic eddy-current damper.
Trans ASME J Vib Acoust 120(1):272278
110. Olsson K-O (2003) Some unusual cases of rotor instability. ASME J Vib Acoust 125:477
481
111. Ishida Y, Inoue T (2005) Vibration suppression of a rotor system by a magnetic damper. In:
ISCORMA-3, Cleveland, Ohio, Paper No. 136
112. Scholz A, Liebich R, Paysan G, Blutke R (2010) Dynamics of flexible rotors supported on
elastomer bearings. In: 8th IFToMM conference on rotor dynamics, pp 207213
113. Crandall SH, Brosens PJ (1961) On the stability of rotation of a rotor with rotationally
unsymmetric inertia and stiffness properties. J Appl Mech ASME, 28(4), 567570
114. Yamamoto T, Ota H (1963) On the vibrations of the shaft carrying an asymmetrical rotating
body. Bull JSME 6(21):2936
115. Iwatsubo T, Kanki H, Kawai R (1972) Vibration of asymmetric rotor through critical speed
with limited power supply. J Mech Eng Sci 14(3):184194
116. Forrai L (1976) Vibration of a rotating asymmetric shaft carrying two disks supported in
asymmetric bearing. In: VIRM, I MechE, pp. 4348
117. Mayes LW, Davis WGR (1980) A method of calculating the vibrational behaviors of
coupled rotating shaft containing transverse crack. In: Proceedings of VIRM, I Mech E,
pp 1727
118. Lee C-W, Suh J-H, Hong S-W (2004) Model analysis of rotor system using modulated
co-ordinates for asymmetric rotor system with anisotropic stator. In: IMech E, VIRM8,
C623/051/2004, p 575
119. Liebich R, Gasch R (1996) Spiral vibrationsmodal treatment of a rotor-rub problem
based on coupled structural/thermal equations. IMech E, VIRM6 C500(042):405413
120. Balbahadur AC, Kirk RG (2004) Part 1theoretical model for a synchronous thermal
instability operating in overhung rotors. Int J Rotating Mach 10(6):469475
121. Eckert L, Schmied J, Ziegler A (2006) Case history and analysis of the spiral vibration of a
large turbogenerator using three different heat input models. In: 7th IFToMM conference
rotor dynamics paper 161
122. Grigoriev BS, Fedorov A, Schmied J (2010) Asymmetrical heating in a tilting pad journal
bearing causing shaft bending. In: Proceedings of 8th IFToMM conference rotor dynamics,
pp 101108
123. Kirk RG, Guo Z (2013) Design tool for prediction of thermal synchronous instability. In:
Proceedings of ASME 2013 international design engineering technical conference and
computer and information in engineering conference IDETC/CIE DETC2013-12966
124. Golebiowski M, Nordmann R, Knopf E (2014) Rotordynamic investigation of spiral
vibrations: thermal mode equation development and implementation to combined-cycle
power train. In: Proceedings of ASME turbo expo, Dsseldorf, Germany, GT2014-25430
125. Matsushita O, Kikuchi K, Kobayashi S, Furudono M (1980) Solution method for
eigenvalue problem of rotor-bearing systems (part 1, undamped system with isotropic
supports). Bull JSME 23(185):18721878
126. Murphy BT Rotordynamics software XLrotor. http://www.xlrotor.com/
References 353

127. RBTS, Inc. Rotordynamics software ARMD. http://www.rbts.com/


128. Schmied J, (DELTA JS AG), rotordyanmics software MADYN. http://www.delta-js.ch
129. Smalley AJ, Almstead LG, Lund JW, Koch ES (1974) Users manualMTI cadense
program CAD-25dynamic stability of a flexible rotor. Mechanical Technology Inc
130. Nelson HD, McVaugh JM (1976) The dynamics of rotor-bearing systems using nite
elements. ASME J Eng Ind 593600
131. Hawkins LA, Murphy BT, Kajs J (2000) Analysis and testing of a magnetic bearing energy
storage flywheel with gain-scheduled, mimo control. ASME turbo expo, 2000-GT-405
132. Memmott EA (2003) Usage of the Lund rotordynamic programs in the analysis of
centrifugal compressors (Jorgen Lund Special Issue). ASME J Vib Acoust 125, pp 500506
133. Moore JJ, Camatti M, Smalley AJ, Vannini G, Vermin LL (2006) Investigation of a
rotordynamic instability in a high pressure centrifugal compressor due to damper seal
clearance divergence. In: 7th IFToMM conference on rotor dynamics, Paper ID-130
134. Wagner MB, Younan A, Allaire P, Cogill R (2010) Model reduction methods for rotor
dynamic analysis: a survey and review. Int J Rotating Mach 2010, Article ID 273716, 17 pp
135. de Noronha RF, Miranda MA, Memmott EA, Cavalca KL, Ramesh K (2015) Stability
testing of CO2 compressors, engineers notebook, insights issue, Dresser Land, pp 1629

R1E References (Books)

B1. Den Hartog JP (1956) Mechanical vibrations, McGraw-Hill


B2. Bogoliubov NN, Mitropolsky YA (1961) Asymptotic methods in the theory of non-linear
oscillations, Gordon and Breach
B3. Tondl A (1966) Some problems of rotor dynamics, Chapman and Hall
B4. Przemieniecki JS (1968) Theory of matrix structural analysis, Dover Civil and Mech Eng;
Bathe KJ (1982) Finite Element Procedures in Engineering Analysis, Prentice-Hall Civil
and Mech Eng
B5. Gasch R, Pftzner H (1975) Rotordynamik eine Einfhrung, Springer
B6. Levy S, Wilkinson JPD (1976) The component element method in dynamics, McGraw-Hill,
Inc
B7. Rieger NF, Crofoot JF (1977) Vibrations of rotating machinery, Vibration Institute
B8. Craig RRJ (1981) Structural dynamicsan introduction to computer methods, Wiley, New
York
B9. Bathe KJ (1982) Finite element procedures in engineering analysis, Prentice Hall
B10. Rao JS (1983) Rotor dynamics, Wiley Eastern Limited
B11. Vance JM (1988) Rotordynamics of turbomachinery, Wiley
B12. Rieger NF (ed) (1988) Rotordynamics 2 problems in turbomachinery, Springer
B13. Someya T (ed) (1988) Journal-bearing databook, Springer, Tokyo
B14. Goodwin MJ (1989) Dynamics of rotor bearing systems, Thomson Learning
B15. Darlow MS (1989) Balancing of high-speed machinery, Springer
B16. M. Lalanne, G. Ferraris (1990), Rotordynamics Prediction in Engineering, John Wiley and
Sons
B17. Ehrich FF (ed) (1991) Handbook of rotordynamics, Chapter 3 Balancing of rigid and
flexible rotors, McGraw Hill
B18. Lee CW (1993) Vibration analysis of rotors, Kluwer Academic Publishers
B19. Childs D (1993) Turbomachinery rotordynamics: phenomena, modeling, and analysis, John
Wiley & Sons (1993), ISBN-13: 978-0471538400
B20. Kramer E (1993) Dynamics of rotors and foundations, Springer
B21. Lalanne M, Ferraris G (1997) Rotordynamics prediction in engineering, 2nd edn, Wiley,
266 pp. ISBN: 978-0-471-97288-4
B22. Eshleman RL (1999) Basic machinery vibrations: an introduction to machine testing.
analysis, and monitoring, VIPress, Incorporated. ISBN: 0-9669500-0-3
354 References

B23. Genta G (1999) Vibration of structures and machines, 3rd edn, Springer
B24. Ewins DJ (2000) Modal testing, theory, practice, and application, John Wiley & Sons.
ISBN-13: 978-0863802188
B25. Adams ML Jr. (2001) Rotating machinery vibration, from analysis to troubleshooting,
Marcel Dekker
B26. Yamamoto T, Ishida Y (2001) Linear and nonlinear rotordynamics: a modern treatment
with applications, Wiley
B27. Chen WJ, Gunter EJ (2001) Introduction to dynamics of rotor-bearing systems,
self-published
B28. Gasch R, Nordmann R, Pfutzner H (2002) Rotordynamik, Springer
B29. Bently DE, Hatch CT (2002) Fundamentals of rotating machinery diagnostics, Bently
Pressurized Bearing Press
B30. Harris CM, Piersol AG (2002) Shock and vibration handbook, 5th edn, McGraw-Hill
B31. Logan E Jr. (2003) Handbook of turbomachinery, CRC Press
B32. Ehrich FF (2004) Handbook of rotordynamics, 3rd edn, Krieger Publishing Co
B33. Muszynska A (2005) Rotordynamics, Taylor & Francis
B34. Eshleman RL (2005) Rotor dynamics and balancing, Vibration Institute
B35. Genta G (2005) Dynamics of rotating systems, Springer
B36. Kicinski J (2006) Rotor dynamics, IFFM Publishers Gdansk
B37. Schweitzer G, Maslen EH (eds) (2009) Magnetic bearing, Springer
B38. Friswell MI, Penny JET, Garvey SD, Lees AW (2010) Dynamics of rotating machines,
Cambridge University Press
B39. Adams ML (2010) Rotating machinery vibration: from analysis to troubleshooting, CRC
Press
B40. Bachschmid N, Pennacchi P, Tanzi E (2010) Cracked rotors, a survey on static and
dynamic behaviour including modelling and diagnosis, Springer
B41. Vance JM, Zeidan FY, Murphy B (2010) Machinery vibration and rotordynamics, Wiley
B42. Randall RB (2011) Vibration-based condition monitoring, Wiley, Chichester, WestSussex,
UK. ISBN: 978-0-470-74785-8
B43. Gasch R, Knothe K, Liebich R (2010) Strukturdynamik. Diskrete Systeme und Kontinua,
Springer International Publishing AG
B44. Ishida Y, Yamamoto T (2012) Linear and nonlinear rotordynamics: a modern treatment
with applications, 2nd edn, Wiley, 474 pp. ISBN: 978-3-527-40942-6
B45. Vollan A, Komzsik L (2012) Computational techniques of rotor dynamics with the nite
element method, CRC Press
B46. Schneider H (2013) Auswuchttechnik (VDI-Buch), 8th edn, German, VDI
B47. Timoshenko SP (2013) Vibration problems in engineering, 5th edn, Wiley
B48. Childs DW (2013) Turbomachinery rotordynamics with case studies, Minter Spring
Publishing
B49. Fundamentals of balancing manual3rd edn, Schenck USA. https://sales.schenck-usa.
com/Manuals.html
Index

A Bearing, 4
Acceleration, 36 Bearing center, 181
Active magnetic bearing (AMB), 136, 314 Bearing reaction force, 31
Added inertia, 322 Bending vibration, 8
Added mass, 15, 322 Blade, 257
Added mass of spring, 16 Blade phase angle, 282
Air conditioner, 2 Blade resonance, 273
AMB, 141, 175, 325 Blade-shaft coupled vibration, 268
Amplitude, 27 Block-diagonal, 267
Amplitude envelope, 20 Block diagram, 24, 214
Anisotropic bearing, 189, 206 Bode plot, 27, 214, 336
Anisotropic stiffness, 251 Boundary coordinate, 81
ANSYS, 267 Boundary force, 82
Anti-resonant (or zero) frequencies, 100 Boundary spring constant, 87
API 617, 175
API standard, 33 C
Approximate evaluation, 181 Campbell diagram, 180, 273
Approximate solutions, 232 Cantilever, 82
Aseismic evaluation, 173 Centrifugal coefcient, 254, 256259
Asymmetric rotor, 293 Centrifugal compressor, 141, 278
Asymptotic solution, 344 Centroid, 26, 105
Characteristic equation, 185, 217
B Characteristic root, 19
Backward natural circular frequency, 160 Children, 92
Backward natural frequency, 256 Circular node, 254
Backward propagation, 243 Circular whirling, 242
Backward resonance, 169 Clearance gap, 175
Backward wave, 273 Closed loop, 221
Backward wave propagation, 272 Closed loop transfer function, 222
Backward whirl, 155 Complex amplitude, 26
Balance quality grade, 109 Complex displacement, 26, 326
Balda chart, 202 Complex eigenvalue, 19, 181, 209, 218, 321,
Ball bearing, 4, 325 334, 337
Ball bearing defect, 176 Congruence transformation, 46, 85
Ball diameter, 176 Congruent transform, 330
Banded matrices, 326 Conservative, 15
Bandwidth, 325 Conservative system, 184, 189
Base excitation, 168 Contact angle, 176
Beam element, 321, 331 Contact friction, 308

Springer Japan 2017 355


O. Matsushita et al., Vibrations of Rotating Machinery,
Mathematics for Industry 16, DOI 10.1007/978-4-431-55456-1
356 Index

Continua, 53 Drop test, 175


Control force, 24 Dunkerleys formula, 67
Controllable, 49 Dynamic characteristics, 182
Controller transfer function, 24
Coriolis factor, 247 E
Coriolis force, 246 Eccentricity, 105, 165
Coriolis matrix, 267 Eddy currents, 293
Correction mass, 115 Effective mass, 82, 86
Coupled mass, 86 Effective vector, 115
Coupled spring constant, 91 Eigenmode, 46, 334
Couple unbalance, 107 Eigen solution, 46
Craig-Bampton method, 329 Eigenvalue, 241, 262, 342
Critical speed, 9, 28, 163, 215 Eigenvalue analysis, 46
Critical speed map, 208, 334 Eigenvalue equation, 261
Critical viscous damping coefcient, 18 Eigenvector, 46, 334
Cross-damping, 186, 197 Eight parameters, 182
Cross-spring, 236 Elliptical orbit, 193
Cross-spring control, 316 Elliptical trajectory, 172
Cross-stiffness, 186, 197 Elliptical whirling, 190, 242
Cyclic matrix, 261 Energy, 15
Cyclic symmetry structural system, 259 Equation of motion, 9, 26, 264, 326
Cylindrical bearing, 208 Equations of motion of a top, 157
Equilibrium position, 182
D Equivalent mass, 16, 61, 62
1.5-dof, 218 Equivalent mass matrix, 84
1.5-dof model, 228, 235 Equivalent mass of the internal system, 91
1D-FEM, 12 Equivalent Q-value, 37
3D-FEM, 12 Equivalent shaft stiffness, 81
3D nite element analysis, 266 Equivalent stiffness, 84
Damped natural angular frequency, 19 Excitation method, 281
Damped natural frequency, 215 Excited at stationary side, 281
Damped system, 334 Exciting frequency, 9
Damped vibration waveform, 18 External damping, 288
Damping, 9 External force, 9, 43, 267, 269
Damping coefcient, 18
Damping matrix, 43, 326 F
Damping ratio, 19, 20, 216, 233 Feedback system (FB/CNTR), 24, 141
Deflection curve, 17 Feed-Forward (FF) Excitation, 136
Deflection mode, 81, 89 FEM, 12
Deformation, 176 FFT, 23
Diagonal, 47, 341 Finite element method (analysis), 12
Diagonal mass matrix, 91 Finite element model, 321
Diaphragm coupling, 292 Flaw, 176
Difference, 28 Force balance, 31
Diffuser vane, 275 Forward natural circular frequency, 160
Direct damping, 186, 197 Forward natural frequency, 256
Direct spring control, 317 Forward propagation, 243, 252
Discretization, 321 Forward resonance, 169
Disk element, 321 Forward wave, 273
Displacement, 242 Forward wave propagation, 272
Displacement sensor, 178 Forward whirl, 155, 243
Displacement vector, 43, 267 Four run method, 143
Index 357

Free vibration, 10 Iterative Method (Power Method), 68


Frequency response, 237
Frequency response analysis (FRA), 321 J
Japan Society of Mechanical Engineers
G (JSME), 7
Gain cross-over frequency, 214, 215, 226, 230 Jeffcott rotor, 198, 228
Gap sensor, 282 Job command, 332
Gas turbine, 2 Journal position, 181
Gear coupling, 291
Geometric mass matrix, 267 K
Geometric stiffness matrix, 267 Key way, 293
Gravity, 105 Kinetic energy, 15
Gravity center, 26
Guide vane, 275 L
Guyan method, 326 Laplace transformation, 57
Guyan reduction, 80 Linear relationship, 114
Gyroscopic effect, 153, 167, 182, 192, 196 Lissajous orbit, 159, 174
Gyroscopic factor, 159, 160, 255, 256 Logarithmic decrement, 20
Gyroscopic matrix, 155, 326 Longitudinal vibration, 8
Gyroscopic moment, 156, 246
Gyroscopic system, 163 M
Magnetic bearings, 4
H Magnetic damper, 293
Half power point method, 33 Mass, 9, 13, 321
Hard-bearing balancing machine, 108 Mass eccentricity, 27, 57
Hard disk, 4 Mass effect, 15
Hard Disk Drive (HDD), 178, 255 Mass matrix, 43, 71, 155, 326
Harmonic base excitation, 174 Master, 80, 327
Hot spot, 309 Maximum, 33
Maximum phase lead, 231, 234
I Mechanical model, 90
Impeller, 268 Modal analysis, 41
Impulse response, 18 Modal balance, 116
Inertial coordinate, 241 Modal coordinate, 47, 55
Influence coefcient, 65 Modal damping, 47
Influence coefcient matrix, 78 Modal damping ratio, 226
Influence coefcient method, 117 Modal eccentricity, 57
Inner product, 271 Modal equation of motion, 342
Inner race, 176, 177 Modal excitation, 226
In-phase, 149 Modal external force, 86
Input data, 323 Modal input coefcient, 47
Input matrix, 43 Modal mass, 53, 55, 60, 86
Instability range, 299, 300 Modal mass matrix, 46
Internal coordinate, 81 Modal matrix, 46, 55
Internal damping, 288 Modal model, 47, 55, 222
Internal system, 85 Modal open-Loop, 224
ISO 10814, 32 Modal output coefcient, 47
ISO 1940-1, 109 Modal response, 226
ISO CD14839-4, 175 Modal stiffness, 86
Isotropically supported, 183 Modal stiffness matrix, 46
358 Index

Modal unbalance, 53 Overhung, 158


Mode function, 53
Mode synthesis, 79, 84 P
Mode synthesis model, 329 3-plane, 127
Morton effect, 306 Parallel mode, 149
Mother, 92 Parametrically excited oscillation, 10
Multi-blade fan (Sirocco Fan), 248 Passing through a critical speed, 39
Multi-degree-of-freedom (multi-dof), 41, 43, Peak-hold, 279
53 Permissible unbalance, 109
MyROT, 12, 323 Phase, 28
Phase difference, 24
N Phase lag angle, 26
(n + 2)-plane, 126, 131 Phase lead circuit, 229
(n + 3)-plane, 132 Phase lead function, 228
NASTRAN, 267 Phase lead/lag, 24
Natural angular (circular) frequency, 13 Phase margin, 214, 216, 226, 230, 233
Natural frequency, 13, 15, 256, 321, 334 Physical coordinate, 47
Natural frequency map, 88 Pitch circle, 176
Negative damping, 25 Plane, 107
Newkirk effect, 306 Plant transfer function, 24, 99
Nodal diameter, 254, 268, 282 Polar moment of inertia, 156
Node number, 321 Polar plot, 28
Non-conservative, 18 Pole frequencies, 101
Non-conservative parameter, 344 Positive denite, 46
Non-conservative system, 185 Potential energy, 15
Nonlinear vibration, 10 Power generator, 6
Non-zero edge element, 87 Principal axis of inertia, 106
Nozzle injection, 280
N-Plane Balancing, 126 Q
Number of dimension, 43 Quasi-modal model, 90, 97
Number of the balls, 177 Q-value criterion, 32
Nyquist (polar) plot, 28, 34, 111, 336 Q-value (Q-factor), 13, 28, 33

O R
Observable, 49 Rapid acceleration, 35
Off-diagonal matrix, 330 Rayleigh damping, 47
Oil-lm bearing, 337 Rayleigh quotient, 63
Oil whip, 211 Real symmetric, 342
Onset frequencies, 211 Recess, 177
Open-loop, 213, 221 Regular phase pitch, 146
Open-loop frequency response, 213, 217, 221 Relative coordinate, 90
Open-loop transfer function, 214, 222, 236 Residual permissible unbalance, 108, 109
Optimal damping, 234 Resonance, 9
Optimal damping design, 235 Resonance condition, 178, 245, 270, 274, 277,
Optimal design policy, 239 281
Orthogonality, 46, 341 Resonance curves, 27, 165
Orthogonality condition, 53 Resonance frequency, 114, 252, 278
Outer race, 176 Resonance sensitivity, 123, 174
Out-of-phase, 107 Rigid rotor, 108
Out-of-phase balancing, 149 Rod, 292
Out-of-plane mode, 255 Roller bearing, 4
Output matrix, 43 Root Loci, 210
Index 359

Rotating blade, 275 Steam turbine-generator set, 2


Rotating structure, 253 Stiffness, 9, 321
Rotating system, 343 Stiffness anisotropy, 198
Rotational angular velocity, 26 Stiffness coefcient, 182
Rotational coordinate, 241 Stiffness matrix, 43, 71, 155
Rotational coordinate system, 178, 266 Stiffness matrix method, 70
Rotational pulse, 111 Stopper, 175
Rotational speed, 176 Strain energy, 15
Rotor drawing, 334 Strain gauge, 278
Rotordynamics, 6, 154 Stress, 278
Rubbing, 308 Structural dynamics, 153
Structural system, 341
S Sweep excitation, 169
s-domain, 57 Synchronous speed line, 165
Secondary critical speed, 296
Self-excited vibration, 10, 25 T
Self-excited whirling, 292 Thermal bow, 306
Sensitivity function, 222 Thermal-bow induced vibration, 312
Series coupling, 228 Thermal bow model, 307, 314
Shaft stress, 242 Thin disk, 253
Shear deformation, 323 Thrust vibration, 268
Silicon crystal manufacturing equipment, 292 Tilting/conical mode, 149
Single-degree-of-freedom (single-dof, 1-dof), Tilting vibration, 268
13, 26, 41, 48, 181 Time constant, 229
Single-dof system, 213 Top, 155
Single nodal diameter, 255 Torsional vibration, 8, 268
Single-plane balancing (Modal Balancing), 111 Transfer function, 26, 337
Skew Hermitian, 343 Transfer matrix method, 73
Slave, 327 Transformation matrix, 330
Slave coordinate, 80 Translational vibration, 268
Sliding (oil lm) bearing, 4, 181, 325 Transmissibility, 32
Soft-bearing balancing machine, 109 Transverse moment of inertia, 156
Southwells equation, 254 Trial mass, 116, 146
Spinning, 156 Truncation, 47
Spiral locus, 313 Turbo-compressor, 275
Spline, 293 Two-dof, 42
Split, 161 Two-plane balancing, 147
Spring constant, 13, 14 Two-pole generator, 293
Stability analysis, 331
Stability condition, 289, 316 U
Stable, 19 Unbalance, 105
Stable separation, 302 Unbalanced mass, 165
Stagger angle, 258 Unbalance force, 31, 164
Standing wave, 269, 273 Unbalance magnitude, 165
Static deflection, 338 Unbalance resonance, 163
Static deflection mode, 81 Unbalance response, 336
Static equilibrium point, 15 Unbalance vibration, 25, 26, 163
Static pressure distribution, 269 Undamped free vibration, 13
Static unbalance, 106 Undamped M-K system, 334
Stationary vane, 275 Universal balancing, 131
Steady-state, 181 Unstable, 19
360 Index

V W
v_BASE, 7 Washing machine, 3
Vector locus, 28 Wear, 291
Vector monitor, 35, 111 Whirling, 29, 156
Vibration characteristics, 181 Whirling motion, 155, 163
Vibration Engineering Database Committee, 7 Whirling trajectories, 172
Virtual absolute coordinate, 90 Wire, 292
Viscous damping coefcient, 18, 182

You might also like