Sei sulla pagina 1di 250

Foreword

The identification of vibrating structures and machines through the use


of experimental data is a prevalent task for those interested in obtaining a
rational and reliable design of mechanical systems.
Typically, the description of a dynamical system by a mathematical
model has been extensively used in such fields as aerospace and automotive
structures, rail transportation systems, machine tools, machine foundations,
buildings, etc.
The early studies were limited to the determination of natural
frequencies and mode shapes, as well as to the measurement of damping and
dynamic stiffness properties. The development of the analog and digital
computers and the new digital real-time measuring equipment have
contributed to the implementation of new computer oriented system
identification techniques.
Perhaps one of the most powerful application is the optimization of
the design of a structure in respect to its dynamic behavior by altering the
mass and stiffness properties of the corresponding analytical model. This
avoids making physical modifications on a prototype, which would be an
inefficient, time-consuming process, made simply on a trial and error basis.
The purpose of this book is twofold. First, to present a critical survey
of the classical works on parameter estimation of mechanical structures and
to provide the research workers, design and production engineers with useful
identification schemes, assisting them to formulate logically based
mathematical models of the actual physical systems. Second, methods
recently developed by the author are described, which were successfully
applied to the determination of the dynamic properties of materials and
structures.
The basic concepts are defined in the introductory chapter.
Chapter two is concerned with the parameter estimation of single-
degree-of-freedom systems. Separate treatment is given to the systems with
hysteretic and viscous damping in order to avoid misuse of the approximate
formulae to heavily damped systems. The most frequently used data-
reduction techniques are discussed, particularly the graphical analysis of the
frequency response curves. Both grounded and ungrounded systems are
considered, excited by harmonic forces of constant amplitude or due to

5
rotating unbalance. Identification methods using excitation with forces in
quadrature are described in detail.
Chapter three contains some elements from the theory of vibration of
lumped-parameter multi-degree-of-freedom systems having proportional
and nonproportional damping. The harmonic response of such discrete
systems is expressed in terms of the real (classical or forced) or complex
modes of vibration.
In chapter four comments are made on the derivation of the necessary
experimental data and on the structure of the analytical models used in
system identification. Difficulties raised by ill conditioned matrices and
truncated models are presented.
The following chapters are concerned with identification methods
using single-point (chapter 5) and multi-point (chapter 6) excitation. Both
direct and iterative parameter optimization methods are discussed.
Chapter seven provides an introduction to the parameter estimation of
single-degree-of-freedom weakly non-linear systems, with linear hysteretic
damping and cubic stiffness.
Aspects connected to the instrumentation used and the practical
implementation of the presented methods are treated in reference [19].

June 25, 1977 Mircea Radeş

6
Table of Contents

Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1. The Dynamic Characteristics of Mechanical Systems and Their Identification . . 14

1.1. Nature and effects of vibrations . . . . . . . . . . . . . . . . . . . . 14


1.2. Steps of a dynamic study . . . . . . . . . . . . . . . . . . . . . . 16
1.3. Analysis of the dynamic response of a mechanical system . . . . . . . . . 17
1.4. Identification of vibrating systems . . . . . . . . . . . . . . . . . . 18

2. Parameter Identification of Single-Degree of-Freedom Linear Systems . . . . 22

2.1. Damping models. . . . . . . . . . . . . . . . . . . . . . . . . . . 22


2.1.1. The linear viscous damping . . . . . . . . . . . . . . . . . . . 22
2.1.2. The hysteretic damping . . . . . . . . . . . . . . . . . . . . . 24
2.1.3. The hereditary damping .. .. . . . . . . . . . . . . . . . . . . . 25
2.1.4. Other damping mechanisms . . . . . . . . . . . . . . . . . . . . 27
2.2. Excitation with harmonic forces applied to the mass . . . . . . . . . . . . 28
2.2.1. Identification of parameters of the system with hysteretic damping . . . . . 29
2.2.1.1. The frequency response functions . . . . . . . . . . . . . . 29
2.2.1.2. Identification based on the diagrams of receptance . . . . . . . . 29
2.2.1.3. Identification based on the diagrams of mobility . . . . . . . . 35
2.2.1.4. Identification based on the diagrams of inertance . . . . . . . . . 39
2.2.1.5. Identification based on the diagrams of dynamic stiffness. . . . . . 41
2.2.1.6. Identification based on the diagrams of mechanical impedance . . . 42
2.2.1.7. Identification based on the diagrams of apparent mass. . . . . . . 44
2.2.1.8. Identification based on the diagram of force transmissibility . . . . 45
2.2.2. Identification of parameters of the system with linear viscous damping . . . . 46
2.2.2.1. The frequency response functions . . . . . . . . . . . . . . . 46
2.2.2.2. Identification based on the diagrams of receptance . . . . . . . . 46
2.2.2.3. Identification based on the diagrams of mobility . . . . . . . . 51
2.2.2.4. Identification based on the diagrams of inertance . . . . . . . . . 56
2.2.2.5. Identification based on the diagrams of dynamic stiffness. . . . . . 58
2.2.2.6. Identification based on the diagrams of mechanical impedance . . . 58
2.2.2.7. Identification based on the diagrams of apparent mass. . . . . . . 59
2.2.2.8. Identification based on the diagram of force transmissibility . . . . 60
2.3. Excitation using harmonic forces in quadrature. . . . . . . . . . . . . . . 61
2.3.1. Parameter identification of a system with hysteretic damping . . . . . . . 62
2.3.1.1. Identification based on the diagrams of receptance . . . . . . . . 62
2.3.1.2. Identification based on the diagrams of dynamic stiffness. . . . . . 64
2.3.2. Parameter identification of systems with viscous damping . . . . . . . . . 65

7
2.4. Excitation with rotating unbalance forces . . . . . . . . . . . . . . . . 68
2.4.1. Parameter identification of a system with hysteretic damping . . . . . . . 68
2.4.2. Parameter identification of systems with viscous damping . . . . . . . . . 70
2.5. Harmonic excitation of ungrounded systems . . . . . . . . . . . . . . . 72
2.5.1. Harmonic excitation through the resilient element . . . . . . . . . . . . 73
2.5.1.1. Identification of systems with hysteretic damping . . . . . . . . 73
2.5.1.2. Identification of systems with viscous damping . . . . . . . . 74
2.5.2. Harmonic excitation of the mass-damped spring-mass system. . . . . . . 77
2.5.2.1. Identification of systems with hysteretic damping . . . . . . . . 77
2.5.2.2. Identification of systems with viscous damping . . . . . . . . 80
2.6. Identification methods without frequency sweep . . . . . . . . . . . . . 83
2.6.1. Method of “displaced” frequencies. . . . . . . . . . . . . . . . . . 83
2.6.1.1. Method of additional masses . . . . . . . . . . . . . . . . 83
2.6.1.2. Method of additional stiffnesses . . . . . . . . . . . . . . . 85
2.6.2. Method of introduced work . . . . . . . . . . . . . . . . . . . . 86
2.6.3. Method of transient response. . . . . . . . . . . . . . . . . . . . 87
2.6.3.1. Starting transient . . . . . . . . . . . . . . . . . . . . 87
2.6.3.2. Decaying transient . . . . . . . . . . . . . . . . . . . . 89
2.6.4. Method of self-excited vibrations . . . . . . . . . . . . . . . . . . 91

3. Elements of the Theory of Linear Lumped Parameter Systems . . . . . . . . 93

3.1. Vibration of linear undamped systems. . . . . . . . . . . . . . . . . 93


3.1.1. Free vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.1.2. Forced vibrations . . . . . . . . . . . . . . . . . . . . . . . . 95
3.2. Vibration of damped systems. . . . . . . . . . . . . . . . . . . . . 97
3.2.1. Assumptions on damping . . . . . . . . . . . . . . . . . . . . . 97
3.2.2. Real “classical” modes of vibration. . . . . . . . . . . . . . . . . . 98
3.2.2.1. Viscous damping . . . . . . . . . . . . . . . . . . . . . 98
3.2.2.2. Hysteretic damping . . . . . . . . . . . . . . . . . . . . . 99
3.2.3. Real forced modes of vibration . . . . . . . . . . . . . . . . . . . 103
3.2.4. Complex modes of vibration . . . . . . . . . . . . . . . . . . . . 110
3.2.4.1. Viscous damping . . . . . . . . . . . . . . . . . . . . . 111
3.2.4.2. Hysteretic damping . . . . . . . . . . . . . . . . . . . . . 115

4. Principles of the Identification of Vibrating Systems . . . . . . . . . . . . 118

4.1. Scheme of the dynamic identification processes . . . . . . . . . . . . . 118


4.2. Experimental determination of the dynamic characteristics of actual systems . . 120
4.2.1. Imposed excitation identification methods . . . . . . . . . . . . . . 122
4.2.2. Identification methods using excitation from normal operation . . . . . . 126
4.2.3. Identification with adjustable models . . . . . . . . . . . . . . . . 127
4.2.4. Nodal pattern analysis . . . . . . . . . . . . . . . . . . . . . . 127
4.3. Structure of the analytical models . . . . . . . . . . . . . . . . . . . 130
4.3.1. Discrete models . . . . . . . . . . . . . . . . . . . . . . . . 130
4.3.2. Modal models . . . . . . . . . . . . . . . . . . . . . . . . . 130
4.3.3. Nature of analytical models . . . . . . . . . . . . . . . . . . . . 132
4.3.4. Transformation from the modal model to the spatial model. . . . . . . . 133

8
4.3.4.1. Computational difficulties . . . . . . . . . . . . . . . . . 133
4.3.4.2. Non-uniqueness of the proportional damping matrix . . . . . . 134
4.3.5. Complete versus incomplete models . . . . . . . . . . . . . . . . 137
4.3.5.1. Complete modal information . . . . . . . . . . . . . . . . 137
4.3.5.2. Incomplete modal information . . . . . . . . . . . . . . . . 138
4.3.6. Computation of pseudoinverse matrices . . . . . . . . . . . . . . . 139

5. Identification of Linear Systems Using Single Point Excitation. . . . . . . . 141

5.1. Principle of identification methods using single point harmonic excitation . . . 141
5.2. Identification of undamped systems . . . . . . . . . . . . . . . . . . 142
5.2.1. The shape of the frequency response curves . . . . . . . . . . . . . 142
5.2.2. Method of “skeleton” diagrams . . . . . . . . . . . . . . . . . 145
5.2.3. Modal model method . . . . . . . . . . . . . . . . . . . . . 147
5.3. Identification of systems with proportional damping . . . . . . . . . . . . 148
5.3.1. Closely spaced modes . . . . . . . . . . . . . . . . . . . . . . 148
5.3.2. Identification methods based on the diagram of response amplitude . . . . 152
5.3.3. Calculation of damping from the phase angle plot . . . . . . . . . . . 156
5.3.4. Identification methods based on the diagrams of response vector components 156
5.3.4.1. The quadrature response method . . . . . . . . . . . . . . 157
5.3.4.2. The maximum quadrature component method . . . . . . . . . 157
5.3.4.3. The in-phase response method . . . . . . . . . . . . . . . 158
5.3.4.4. Phase separation method . . . . . . . . . . . . . . . . . 159
5.3.4.5. Mode separation by regression . . . . . . . . . . . . . . 161
5.3.4.6. Iterative mode separation method . . . . . . . . . . . . . . 164
5.3.4.7. Modal matrix elimination method . . . . . . . . . . . . . . 166
5.3.5. Identification methods based on the polar plot of the response . . . . . 168
5.3.6. “Modal path” method . . . . . . . . . . . . . . . . . . . . . . 172
5.4. Identification of systems with nonproportional damping. . . . . . . . . . . 175
5.4.1. Systems with hysteretic damping . . . . . . . . . . . . . . . . . . 175
5.4.1.1. Polar plot method . . . . . . . . . . . . . . . . . . . 175
5.4.1.2. Modal matrix elimination method . . . . . . . . . . . . . . 179
5.4.2. Systems with viscous damping . . . . . . . . . . . . . . . . . . 181
5.4.2.1. Polar plot method . . . . . . . . . . . . . . . . . . . 181
5.4.2.2. Modal matrix elimination method . . . . . . . . . . . . . . 187
5.5. Identification of heavily damped systems. . . . . . . . . . . . . . . 188
5.5.1. Direct identification methods . . . . . . . . . . . . . . . . . . 189
5.5.1.1. Analytical methods . . . . . . . . . . . . . . . . . . . 189
5.5.1.2. Graphical methods . . . . . . . . . . . . . . . . . . . . 190
5.5.2. Identification methods with error minimization . . . . . . . . . . . . 194

6. Identification of Linear Systems Using Multi- Point Harmonic Excitation . . 197

6.1. Principle of simultaneous multi-point excitation methods . . . . . . . . 197


6.2. Phase resonance method . . . . . . . . . . . . . . . . . . . . . 202
6.2.1. Apportioning of excitation forces . . . . . . . . . . . . . . . . 202
6.2.1.1. Pure experimental methods . . . . . . . . . . . . . . . . 202
6.2.1.2. Semi-experimental methods . . . . . . . . . . . . . . . . 203
6.2.1.3. Method of “reinjection” . . . . . . . . . . . . . . . . . 208

9
6.2.2. Determination of modal parameters . . . . . . . . . . . . . . . 209
6.2.2.1. Method of “work introduced” . . . . . . . . . . . . . . . 209
6.2.2.2. Method of additional masses . . . . . . . . . . . . . . . . 210
6.2.2.3. Method of forces in quadrature . . . . . .. . . . . . . . . . 214
6.2.2.4. Method of polar plots . . . . . . . . . . . . . . . . 216
6.2.2.5. Method of complex power supplied to the structure . . . . . . . 217
6.3. Multi-point successive excitation methods . . . . . . . . . . . . . . 221
6.3.1. Method of independent forces . . . . . . . . . . . . . . . . . . 221
6.3.2. Method of independent additional loads . . . . . . . . . . . . . . 225

7. Parameter Identification of Single-Degree-of-Freedom Nonlinear Systems . . 227

7.1. Harmonic excitation methods . . . . . . . . . . . . . . . . . . . 227


7.1.1. Polar plot of the frequency response . . . . . . . . . . . . . . . . 227
7.1.2. Jump phenomena . . . . . . . . . . . . . . . . . . . . . . . 230
7.1.3. Parameter estimation . . . . . . . . . . . . . . . . . . . . 233
7.2. Harmonic excitation with forces in quadrature . . . . . . . . . . . . . 235
7.2.1. Polar diagrams of the frequency response . . . . . . . . . . . . . . 235
7.2.2. Jump phenomena . . . . . . . . . . . . . . . . . . . . . . . 238
7.2.3. Parameter identification . . . . . . . . . . . . . . . . . . . . 238
7.2.3.1. Method of the three circles . . . . . . . . . . . . . . . 238
7.2.3.2. Method of the two circles . . . . . . . . . . . . . . . . 239

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

10
List of Symbols

𝑎𝑟 , 𝑏𝑟 ̶ abbreviations, see eqs. (3.72)


𝐴𝑟 , 𝐵𝑟 ̶ abbreviations, see eqs. (3.73)
[𝐴], [𝐵] ̶ matrices defined by eqs. (5.68) and (5.48)
c ̶ coefficient of viscous damping
𝑐𝑟 ̶ modal damping coefficient, element of the matrix [ˋ𝑐ˏ]
[𝐶] ̶ viscous damping matrix
(𝑟)
𝐶𝑙𝑗 ̶ damping coefficient of the modal path ℓ– 𝑗
e = 2.718 ̶ basis of natural logarithms
E ̶ error
f ̶ matrix defined by eq. (6.108)
(𝑟)
{ℱ } ̶ principal mode of excitation
g ̶ hysteretic damping factor
[𝐺] ̶ matrix defined by eq. (6.108)
ℎ(𝑡) ̶ weighting function
ℎ ̶ coefficient of hysteretic damping
[𝐻] ̶ hysteretic damping matrix
𝐻(𝑖⍵) ̶ frequency response function
𝑖 = √−1 ̶ imaginary unit
[𝐼] ̶ unity matrix
𝑗 ̶ current location of response measurement
[𝐽] ̶ matrix defined by eq. (6.115)
𝑘 ̶ stiffness
𝑘𝑟 ̶ modal stiffness, element of the matrix [ˋ𝑘ˏ]
[𝐾] ̶ stiffness matrix
(𝑟)
𝐾𝑙𝑗 ̶ stiffness of the modal path ℓ– 𝑗
ℓ ̶ current location of force application
𝑚 ̶ lumped mass
𝑚𝑟 ̶ modal mass, element of the matrix [ˋ𝑚ˏ]
[𝑀] ̶ mass matrix

11
(𝑟)
𝑀𝑙𝑗 ̶ mass of the modal path ℓ– 𝑗
ℳ ̶ mobility (ratio velocity/force)
𝑁 ̶ number of degrees of freedom
𝑁𝑓 ̶ number of excitation frequencies
𝑝𝑟 ̶ principal undamped coordinate
𝑃𝑟 ̶ coefficient in the transmittance function (5.187)
𝒫 ̶ complex power
𝑞𝑟 ̶ independent generalized coordinate
(𝑟)
{𝑞 } ̶ complex mode of vibration (viscous damping)
𝑄𝑟 ̶ coefficient in the transmittance function (5.187)
𝑟 ̶ eccentricity; modal index
{𝑅} ̶ vector defined by eq. (3.171)
𝑠 ̶ arc of curve; modal index
{𝑆} ̶ vector defined by eq. (3.151)
𝑡 ̶ time
𝑇 ̶ period of vibration
𝒯 ̶ transmissibility
𝑢𝑟 , 𝑣𝑟 ̶ modal parameters, see eqs. (3.160)
[𝑈], [𝑉] ̶ matrices defined by eq. 3.150)
(𝑟)
{𝑤 } ̶ complex mode of vibration (hysteretic damping)
𝑊𝑐 , 𝑊𝑝 , 𝑊𝑑 ̶ kinetic energy, potential energy, dissipated energy
𝑊𝑅 , 𝑊𝐼 ̶ active energy, reactive energy
𝑥 ̶ linear displacement
{𝑥} ̶ vector defined by eq. (3.151)
𝑋(𝑖⍵) ̶ Fourier transform of 𝑥(𝑡)
𝑦 ̶ linear displacement
𝑌(𝑖⍵) ̶ Fourier transform of 𝑦(𝑡)
𝑧 ̶ relative linear displacement
𝑧𝑟 ̶ complex principal coordinate
𝑍 ̶ mechanical impedance (ratio force/velocity)
𝛼 ̶ receptance (ratio displacement/force)
𝛽 ̶ dynamic stiffness (ratio force/displacement)
𝛾 ̶ mass ratio

12
𝛾𝑠 ̶ coefficients in eq. (6.36)
(𝑟)
𝛿𝑗𝑙 ̶ modal coefficient, see eq. (3.75)
𝜁 ̶ damping ratio
𝜂 ̶ inertance (ratio acceleration/force)
θ ̶ angle
(𝑟)
𝜘𝑗𝑙 ̶ modal coefficient, see eq. (3.75)
𝜆 ̶ ratio of components at the excitation with forces in quadrature
𝜆𝑟 ̶ coefficients in eq. (3.120)
𝜇 ̶ coefficient of nonlinearity; apparent mass (ratio force/acceleration)
𝜈 ̶ frequency (Hz)
𝜈𝑟 ̶ damped principal coordinate
(𝑟)
{𝜉 } ̶ eigenvector (nonproportional viscous damping)
𝜎𝑟 ̶ complex circular frequency, see (3.163)
𝜏𝑟 ̶ notation, see eq. (5.88)
𝜑 ̶ phase angle
(𝑟)
{𝛷 } ̶ real forced mode of vibration
(𝑟)
𝜒𝑗𝑙 ̶ modal coefficient, see eq. (3.74)
ϕ ̶ phase angle
(𝑟)
{𝛹 } ̶ “classical” real mode of vibration
𝜔 ̶ circular frequency (rad/sec)
𝜔𝑛 ̶ natural frequency (single-degree-of-freedom systems)
𝜔𝑟 ̶ natural frequency of the r-th mode)
𝜔
̅𝑓 ̶ excitation frequency (rad/sec)
Ω ̶ dimensionless frequency

This list contains the symbols most frequently used in the book. All notations are
defined in the text at the appropriate location. When the same symbol was used for different
quantities, care had been taken to avoid confusion.

13
CHAPTER 1

The Dynamic Characteristics of Mechanical Systems


and Their Identification

1.1. NATURE AND EFECTS OF VIBRATIONS

Oscillations – of mechanical, thermal or electromagnetic nature – are


dynamic phenomena, characterized by the time fluctuation of a system state
variable, usually near the value corresponding to equilibrium.
Vibrations are oscillations of elastic systems, i.e. movements of mechanical
systems due to elastic restoring forces.
When an elastic system is disturbed from the stable equilibrium position
(or configuration), under the action of a static load, it stores potential/strain energy.
If the system is released, without any energy input, it performs free vibrations.
Displaced from a position of stable equilibrium and released, the system
comes back under the action of elastic restoring forces. Due to the mass of the
system, these forces produce accelerations and the system passes through the
equilibrium position with some velocity, hence it has kinetic energy.
The displacement continues until the elastic restoring forces, which now
produce a deceleration, stop the system. In the quiescent position, the kinetic energy
possessed at the equilibrium position transforms into potential/strain energy, and
the process repeats itself. In the presence of friction forces, the mechanical energy
is dissipated and the vibration dies out, for small damping, after a certain number
of cycles.
2. When the displacement takes place in the presence of a source of energy,
the system can undergo self-excited vibrations. In this case, the motion amplitude
increases continuously, until it is limited by nonlinear effects or by damping. The
motion is controlled by a periodic force, created or determined by the motion of the
system itself, at one of its natural frequencies, though the energy is uniformly
supplied by the external source.
3. Forced vibrations are produced by excitation forces which are
independent of the motion. Generally, forces and displacements are applied
dynamically, with certain velocity, hence they are time-variable. Such excitations
imply an energy transfer from the perturbing source to the system. If the energy
transfer is periodic, the same during each cycle, the forced vibration is stationary,
of constant amplitude. If the transfer is non-uniform, the vibration is a transient,
with variable amplitude, until reaching a steady-state regime or dying out.

14
4. The sudden application of a perturbation gives rise to shocks or impacts.
A shock is a suddenly applied non-periodic perturbation transmitting kinetic energy
to the system in a relatively short time compared to its natural period of vibration.
Starting from the moment of action removal, the response to a shock is a free
vibration. Transients may last for several periods of vibration of the system.
5. The periodic and transient vibrations are deterministic phenomena. In
most practical applications, the vibrations are random, non-deterministic, such that
the instantaneous values of the quantities defining the motion are no more
predictable. Using the probability calculus, statistical quantities and average values
are used which, in the case of stationary ergodic processes, with Gaussian
distribution, are predictable.
6. When a linear time-invariant system is subjected to a perturbation, the
resulting motion is the sum of two distinct components: the forced vibration,
described by a function resembling the excitation, and the natural vibration,
dependent solely upon the dynamic characteristics of the system, whose time
function is usually a combination of a sinusoid and an exponential.
7. In the case of a stationary harmonic or random perturbation, the natural
vibration dies away soon after the start of the motion and remains only the forced
vibration, which, under certain conditions, can produce resonance.
8. Resonance takes place at the frequencies where the sum of the two
“reactive” recoverable – potential and kinetic – energies vanishes. A resonance
vibration builds up an amplitude which reaches a level where the energy input
equals the damping losses. It arises when the excitation time function resembles the
time function of the system natural vibration or when the excitation spectrum spans
a frequency range encompassing the system natural frequencies.
Resonance implies large motion amplitudes accompanied by large loads
and stresses in some parts of the system, or considerable relative motions which
can lead to fatigue cracks, improper operation, wear, chatter, hence noise and
distress.
9. Avoidance of dangerous vibrations, especially near resonance, can be
done by:
- identification of vibration sources and their attenuation or isolation;
- modification of the structure or the parameters (mass, stiffness) of the
excited system to avoid resonances;
- use of damping when resonances cannot be avoided;
- use of dynamic vibration absorbers.
In all cases, it is necessary to understand how the structural modifications
influence the system dynamic response to various excitations. On this basis, the
design engineer will be able to develop techniques and to use models able to predict
the dynamic response of actual systems, the test engineer will correctly interpret
the test results and together they will take sensible decisions of how to adjust the
system structure and/or parameters in the right direction.

15
1.2. STEPS OF A DYNAMIC STUDY

The aim of a dynamic analysis of a mechanical system is either the


minimization of the adverse effects of vibration, or, where the vibration is a useful
process, the optimization of the operating parameters. This requires the knowledge
and sometimes the prediction of the system dynamic response, as well as an
understanding of how the modifications of the parameters of either the system or
the perturbing source determine its dynamic behavior.
Generally, the analytical study of a vibration problem entails four steps:
Step 1. Definition of the studied problem, respectively of the analyzed
system.
Step 2. Modeling the relevant characteristics of the observed dynamic
phenomena. Usually this step is split into two stages; first, the physical modeling,
second the mathematical modeling. Sometimes one goes directly to the
mathematical modeling.
Physical modeling implies formulation of a “physical model” whose
behavior approximates as good as possible the behavior of the actual system. The
physical model has the same basic characteristics as the actual system, but it is
simpler, hence most prone to analysis.
The component parts of a machine or structure can be modeled as beams,
plates, shells, discs, tubes, massive bodies, etc. The reciprocal action of two bodies
can be schematized by point forces, couples, distributed loads, etc.
In many cases, the structural dynamic response can be described by a
“lumped parameter model”, consisting of masses, springs and dashpots. When the
structure is complex and its response is of interest over a large frequency range, the
number of these elements can be large. However, for simple systems or for complex
structures within a limited frequency range, the response can be quite accurately
represented using a physical model consisting of only a few discrete elements.
The approximations used in building physical models include: neglecting
secondary effects, neglecting some interactions with the environment, replacement
of “distributed” characteristics by similar “concentrated” parameters, linearization
of excitation-response relations between physical variables, neglecting the time
variation of some parameters, neglecting the randomness of some phenomena, etc.
As the model is improved and the problem at hand is more accurately defined, some
of these approximations are removed. For instance, at high frequencies inertia
effects cannot be neglected, the Bernoulli-Euler beam model is replaced by the
Timoshenko beam. For large amplitude vibrations, geometrical non-linearities are
taken into account, usually neglected at small amplitudes.
Mathematical modeling implies constructing a “mathematical model” able
to represent the physical model, i.e. to derive the differential (integral, finite
difference) equations of motion of the physical system.
In order to derive the mathematical model from the physical model, four
successive steps are necessary:

16
a – to select the variables which describe the state of the system at a given
time;
b – to derive the equilibrium equations for the system as a whole or for
each component or subsystem;
c – to write the compatibility equations which describe the relationships
between the motions of the interconnected subsystems;
d – to derive the physical laws, i.e. the constitutive relationships for each
component.
Step 3. Study of the dynamic behavior of the mathematical model by
solving the equations of motion and deriving the relations between the modal
parameters and the experimentally measurable quantities. In other words – use of
the model to discover the relevant behavior of the real system.
Step 4. Validation of the model, by correlating its behavior to that of the
system being tested.
As progress is made, one comes back to the first stage, following the
sequence as often as necessary. The design implies the selection of the system
physical parameters and the constructive solutions, or their modification in order to
obtain the desirable behavior.
The steps involving the definition of the problem and the transposition of
the results in a project are beyond the aim of this book; the necessary experience is
gained in the design process. Also, no detailed treatment is presented of either the
solution of the equations of motion – covered in works on the dynamics of elastic
systems [20] or of the measurement of the dynamic response of actual systems –
treated in dedicated works [19].
In this book the emphasis is on the determination of the relationships
between the parameters of the models frequently used in vibration theory and the
experimentally measurable quantities.

1.3. ANALYSIS OF THE DYNAMIC RESPONSE OF A


MECHANICAL SYSTEM

Mechanical vibrations are produced by time-variable external loads or


impressed displacements, generally referred to as excitations (or perturbations).
The mechanical effects of these perturbations are the movements of the
system points and the dynamic stresses in its components, generally referred to as
responses.
The response depends on both the excitation parameters and the system
dynamic characteristics. The excitation-response relationship is system dependent.
The solution of a vibration problem consists of the derivation of the
relationships between excitation, response and the system dynamic characteristics,
as shown in the block diagram of Fig.1.1.
The study of the dynamic response of a mechanical system entails two main
problems: direct and inverse [8].

17
If the equations describing the dynamic behavior of a linear time-invariant
system are known, than the direct problem is that of finding the response to a
specific excitation. For example, given the equations describing the dynamics of an
airplane, predict its response to a wind gust.

In the inverse problem, the response to a given excitation is known, but


either the equations of motion or the structure or some system parameters are
unknown and must be determined, or the system is known and the excitation must
be predicted.
The inverse problem can be subdivided into three sub-problems [8]:
a) – The design or synthesis problem. Given the excitation and the response
find a physically realizable system which fits the excitation-response relation as
close as possible. Generally there is no unique solution, different degrees of
approximation can result in systems with increased complexity.
The design implies a compromise between several different requirements;
the realization of the practical system implies combining science and art, the latter
referring to the instinct for correct proportions.
b) – The control problem. Given the system description and its response,
find the forces which cause the response. Examples are the identification of
excitation forces, the measurements using instruments with known transfer function
or calibration curve.
c) – The system identification problem. Given the excitation and the
response, find an analytical model of the system. The excitation-response
relationships are determined by measurements. The most often used test data are
the values of the complex frequency response functions obtained by excitation with
“test signals” (harmonic, non-periodic, random). Based on their curves, it is
possible to estimate natural frequencies and mode shapes, as well as specific
dynamic properties – damping, dynamic stiffness, etc.
The “identification in the frequency domain” will be treated with priority
in this book.

1.4. IDENTIFICATION OF VIBRATING SYSTEMS

The solution of a vibration problem involves an iterative process


combining theoretical analysis and testing. In this frame, an essential element is the
knowledge of the dynamic characteristics of materials and structures. Their most
advanced expression is the mathematical model of the analyzed system. This allows

18
a quantitative study of vibration phenomena, being very useful in design and
optimization problems.
System identification may be defined as the process of determining the
differential (or finite difference) equations which describe the system behavior, in
accordance with some predetermined performance criterion, based on
relationships between the quantities that characterize the excitation and response
[105].
The dynamic identification entails the derivation of the equations of motion
and their coefficients, hence the determination of the system dynamic
characteristics.
Generally, an identification process entails three phases:
a) – Formulation of the model structure. The differential equations are
selected, based on previous experience. Comparing and correlating the system
measured frequency response curves with the analytical curves of known models it
is possible to estimate the number of significant degrees of freedom, the type of
damping, the opportunity of introducing non-linear elements, etc. and make a pre-
structuring of the system. The unknown parameters of this structure – masses,
stiffnesses, damping constants – are determined afterwards.
b) – Selection of the correlation criterion of the model and the real
structure. A mathematical criterion is specified which has to be optimized to
accomplish the identification.
In the simplest form, the frequency response curves of the structure and the
model are directly compared, usually in a few points near resonances.
In more advanced methods, the agreement of the model and actual structure
properties is expressed by a criterion function, for instance, the integral of the
square error (here the error is the difference between the instantaneous responses
of the structure and the model). Optimal model parameters are determined by
finding the minimum of a multivariate function. In the regression method, the mean
square error criterion is adopted.
In other cases, the error is defined as the difference between the frequency
responses of the structure and model. When the loops of the polar diagrams are
approximated by circles (see § 5.3.5), the “best circle” can be traced based on the
condition of minimum mean square error between the coordinates of the test points
and points of the theoretical curve representing the model response.
The experimental identification of the coefficients of a transfer function
can be analogously performed [142].
c) – Determination of model parameters. An algorithm for adjustment of
the unknown parameters is selected and used to identify the parameters in such a
way that the system identification criterion is minimized. For many practical cases,
a fairly good accuracy is obtained using “direct identification” methods
(considering zero error), without statistical processing of the results or optimization
by regression, maximum likelihood method, non-linear filters, Bayesian method,
etc.

19
The three stages are not compulsory. Usually, the structure and some
parameters of the mathematical model are known and the identification problem
reduces to the estimation of some unknown parameters (“grey box” identification).
In other cases, the system description is totally unknown, and we have the “black
box” identification problem, as in the determination of the transfer functions of
system components in order to predict the response of the assembly [64].
In the modern meaning, the identification of mechanical systems has a
broader sense, encompassing either the partial or the total realization of the
following steps:
a – experimental derivation of excitation-response relations, often as
directional frequency response functions;
b – measurement of the natural modes of vibration, or of “modal maps”
consisting of equal vibration amplitude lines;
c – construction of a modal model or a physical model made of elastic,
inertial and dissipative elements, able to describe a physical phenomenon of interest
for the design or production engineer; sometimes, the derivation of algebraic
expressions of the system transfer functions, based on frequency response functions
given in graphical or numerical form;
d – use of the model for the system analysis and optimization.
Large interest applications of the identification of mechanical systems
include:
a – identification of the dynamic behavior of a mechanical system as part
of an adaptive or logical control system;
b – identification of the transfer characteristics of the driving and control
systems of machines;
c – identification of the directional frequency response characteristics of a
structure made of several subassemblies;
d –calculation of the response to other excitations or to several
simultaneous excitations, based on data obtained from test data obtained with a
known excitation;
e – prediction of the effects of structural modifications made to obtain the
required constructive solution, without testing several prototypes.
An exhaustive bibliography on the general system identification problems
can be found in works by Nicolai [82], Eykhoff [32], Ǻström and Eykhoff [3],
Iserman [59, 60], Sage and Melsa [105] and in the papers presented at I.F.A.C.
symposia [56-58]. State of art summaries on the identification methods of
mechanical systems are published by Young and On [138], Natke [79-81] and
Radeş [93]. The monograph published in 1972 (W.D. Pilkey and R. Cohen, eds.) at
the A.S.M.E. Symposium on “System Identification of Vibrating Structures –
Mathematical Modelling Based on Experimental Data” [115] is a synthesis of the
recent contributions in this field. The emphasis is on new methods of statistical
analysis of test data, parameter estimation by modern techniques of error

20
minimization such as the gradient method, maximum likelihood (Fischer),
conditional probability (Bayes), non-linear filters (Kalman), etc.
In this book, only methods applied in mechanical engineering are
discussed: the determination of the dynamic characteristics of machine tool
structures, chassis and vehicle frames, railway equipment and earth moving
machines, the ground and flight testing of airplanes and rockets, the dynamic
properties of soils and materials used in vibration isolation, etc. The book is
intended to establish a bridge between the classical works on the determination of
the dynamic characteristics of mechanical systems and the modern works on system
identification.

21
CHAPTER 2

Parameter Identification of Single-Degree-of-Freedom


Linear Systems

This chapter deals with the simplest “direct identification” methods, based
on the analysis of the dynamic response of the single-degree-of-freedom model.
In the following, only linear time-invariant systems are considered.
First, several damping representations are presented, currently used in
vibration studies. Second, methods of graphical analysis of the frequency response
curves are discussed, obtained by a frequency sweep near resonance, called
“resonance methods”. They are followed by techniques based on the modification
of system impedance, on the calculation of the dissipated energy or the analysis of
transient vibrations.

2.1. DAMPING MODELS

The main damping mechanisms of the vibrations of a deformable structure


are the “internal damping” due to energy losses within the material, the “structural
damping” due to the friction at joints and in bearings and the “external damping”
due to resistances in the ambient medium. Because the physical nature of these
damping mechanisms is so different, several mathematical models have been
developed for their description. The most often used damping models are shortly
described in the following.

2.1.1. The Linear Viscous Damping

The simplest mechanical model, which apart from storing potential (strain)
energy describes the energy loss, is the Kelvin-Voigt model (Fig.2.1). It is
comprised of an ideal elastic element, represented by a spring of stiffness 𝑘, in
parallel with a dashpot of damping coefficient 𝑐.
The elastic force of the spring is proportional to the relative displacement
|𝑓𝑒 | = 𝑘(𝑥 − 𝑦) = 𝑘𝑧; the damping force developed in the dashpot is proportional
to the relative velocity |𝑓𝑑 | = 𝑐(𝑥̇ − 𝑦̇ ) = 𝑐𝑧̇ . The “force-displacement”
relationship is
𝑓 = 𝑘𝑧 + 𝑐𝑧̇ , (2.1)
where a dot denotes a derivative with respect to time.
Though as a rheological model it has several drawbacks [68], the Kelvin-
Voigt model is extensively used in the calculation of damped vibrations due to the
simple and linear form of equation (2.1).

22
When the force is known as a function of time, the solution of the
differential equation (2.1) is
𝜏
1 𝑡 −
𝑧(𝑡) = 𝑧(0) + ∫0 𝑒 𝜏𝑜 𝑓(𝑡 − 𝜏)𝑑𝜏, (2.2)
𝑐
𝑐
where 𝜏𝑜 = is the “time constant” of the unit.
𝑘
When a harmonic displacement
𝑧 = 𝑧̂𝑐𝑜𝑠𝜔𝑡 (2.3)
is imposed on the model, it reacts with a force
𝑓 = 𝑧̂ (𝑘 𝑐𝑜𝑠𝜔𝑡 − 𝜔𝑐 𝑠𝑖𝑛𝜔𝑡). (2.4)
The elimination of time between equations (2.3) and (2.4) yields the
hysteresis curves shown in Fig.2.2,

of equation

𝑓 𝑧 𝜔𝑐 2
= ± √1 − (𝑧) . (2.5)
𝑘𝑧̂ 𝑧̂ 𝑘 𝑧̂
𝜔𝑐
They are ellipses of parameter .
𝑘
The energy dissipated during a cycle of vibration is proportional to the area
enclosed by the ellipse
2𝜋
𝑊𝑑 = ∮ 𝑓𝑑𝑧 = ∫0𝜔 𝑓𝑧̇ 𝑑𝑡 = 𝜋𝑧̂ 2 𝑐𝜔. (2.6)

For a given amplitude of vibration 𝑧̂ , 𝑊𝑑 increases proportionally to the


frequency ω.

23
Using the complex algebra, for a harmonic displacement
𝑧 = 𝑧̂𝑒 𝑖𝜔𝑡 (2.7)
the force is
𝑓 = (𝑘 + 𝑖𝜔𝑐)𝑧̂ 𝑒 𝑖𝜔𝑡 = 𝑘̅ 𝑧 (2.8)
where
𝑘̅ = 𝑘 + 𝑖𝜔𝑐 (2.9)
is a complex stiffness [11].
When a harmonic force 𝑓 = 𝑓̂ 𝑒 𝑖𝜔𝑡 is applied to the model, the
𝑓̂ 𝑐𝜔
displacement is 𝑧 = 𝑧̂𝑒 𝑖(𝜔𝑡+𝜑) , where 𝑧̂ = and 𝜑 = tan−1 .
√𝑘 2 +𝑐 2 𝜔2 𝑘
Equation (2.6) yields
𝑐𝜔
𝑓̂2 𝜋
𝑘
𝑊𝑑 = 𝑐𝜔 2
.
𝑘 1+( )
𝑘
.
2.1.2. The Hysteretic Damping

Practical observations [62] have shown that in many materials the energy
dissipated per cycle of vibration is proportional to the square of the displacement
amplitude but independent of frequency, hence the linear viscous damping model
does not describe correctly their behavior.
This can be modeled with a damper (Fig.2.3) in which the damping
coefficient 𝑐 varies with the inverse of frequency [12]

𝑐=𝜔∙ (2.10)

Inserting equation (2.10) into (2.6) we have


𝑊𝑑 = 𝜋𝑧̂ 2 ℎ (2.11)
where ℎ is the coefficient of hysteretic damping.
In the case of harmonic vibrations, the damping
force

𝑓𝑑 = 𝜔 𝑧̇ = 𝑖ℎ𝑧 (2.12)
is proportional to the relative displacement 𝑧 = 𝑥 − 𝑦, but
phase shifted 90𝑜 before it, hence in phase with the relative
velocity 𝑧̇ = 𝑖𝜔𝑧.
The complex stiffness (2.9) is independent of frequency
𝑘̅ = 𝑘 + 𝑖ℎ = 𝑘(1 + 𝑖𝑔), (2.13)

24

where the damping factor 𝑔 = is constant. This was observed experimentally at
𝑘
some building materials [109] and elastomers [111].
Remember that the model of hysteretic damping (also referred to as
“constructive” or “structural” damping) was postulated and is valid for harmonic
vibrations only. It does not represent a physically acceptable energy dissipation
mechanism because it leads to a physically inacceptable transient behavior. The
force exerted by the model from Fig 2.3
𝑘𝑔 ∞ 𝑑𝜏
𝑓(𝑡) = 𝑘 ∙ 𝑧(𝑡) + ∫−∞ 𝑧(𝜏) 𝜏−𝑡
𝜋
does not only depend on the past history of the motion but on its future as well
[122].
However, in harmonic regime and in some definite frequency bands, the
hysteretic damping is an acceptable approximation. Originally introduced for
describing the internal damping of solid materials [62] it has been extended to
represent the structural damping [27].

2.1.3. The Hereditary Damping

If a linear viscous dashpot 𝑐 is introduced in the “three parameter model”


from figure 2.4 (corresponding to the rheological model of the “linear standard
solid” [68]) where 𝑘1 and 𝑘2 are elastic springs, we obtain the Maxwell model
whose behavior is described by two equations
𝑓 = 𝑘1 𝑧1 + 𝑐(𝑧̇1 − 𝑧̇2 ), (2.14)
0= 𝑘2 𝑧2 − 𝑐(𝑧̇1 − 𝑧̇2 ), (2.15)
where 𝑧1 = 𝑦1 − 𝑥, 𝑧2 = 𝑦2 − 𝑥.
Since we are not interested in the internal degree of
freedom 𝑧2 , we solve equation (2.15) for 𝑧2 and substitute
the result in equation (2.14). Assuming that for 𝑡 = 0 the
model in unstrained, we obtain
𝑡
𝑓 = 𝑘1 𝑧1 + ∫0 𝐺(𝑡 − 𝜏)𝑧̇1 (𝜏)𝑑𝜏, (2.16)
where
𝑘2
𝐺(𝑡) = 𝑘2 𝑒 − 𝑐 𝑡 . (2.17)
In equation (2.1) the damping term is directly proportional to the
instantaneous relative velocity. In equation (2.16) it depends on the past history of
this velocity. For this reason it is called “hereditary damping”.
When the force is known as a function of time, the solution of equations
(2.14) and (2.15) is [122]

25
2 𝜏
𝑓(𝑡) 1 𝑘2 𝑡 −
𝑧1 (𝑡) = 𝑘 + 𝑐 (𝑘 ) ∫0 𝑒 𝜏2 𝑓(𝑡 − 𝜏)𝑑𝜏, (2.18)
1 +𝑘2 1 +𝑘2
where
1 1
𝜏2 = 𝑐 (𝑘 + 𝑘 ).
1 2

The first term in the right hand side describes the instantaneous response
of the model from Fig.2.4, a behavior often observed at many materials.
For harmonic motion, using the complex algebra, substituting in equations
(2.14) and (2.15) harmonic solutions
𝑧1 = 𝑧̂1 𝑒 𝑖𝜔𝑡 , 𝑧2 = 𝑧̂ 2 𝑒𝑖𝜔𝑡 (2.19)
and eliminating the internal coordinate 𝑧2 , we obtain
𝑓 = 𝑘̅ 𝑧1 ,
where the complex stiffness 𝑘̅ is
𝜔𝑐𝑘2
𝑘̅ = 𝑘1 + 𝑖 . (2.20)
𝑘2 +𝑖𝜔𝑐

Equation (2.20) may be written in a form similar to (2.9) obtained in the


case of simple viscous damping, namely
𝑘̅ = 𝑘𝑒 + 𝑖𝜔𝑐𝑒 , (2.21)
in which
𝜔2 𝑐 2 𝑘2
𝑘𝑒 = 𝑘1 + 𝑘2 𝑘 2 +𝜔2𝑐 2, 𝑐𝑒 = 𝑐 𝑘 2 +𝜔2 2𝑐 2. (2.22)
2 2

The behavior of the hereditary damping model is thus reduced to that of


the Kelvin-Voigt model with frequency dependent characteristics.
The energy dissipated per cycle is
𝑘 2 𝜔𝑐
𝑊𝑑 = 𝜋𝑧̂12 𝜔𝑐𝑒 = 𝜋𝑧̂12 𝑘 2 +𝜔
2
2𝑐 2 ∙
(2.23)
2
𝑘2
It is zero at 𝜔 = 0 and 𝜔 = ∞, having a minimum value at 𝜔𝑜 =
𝑐
𝑐
where 𝑐𝑒 = 2. Accordingly, at 𝜔 = 𝜔𝑜 we have an ellipse of maximum area, and
at 𝜔 = 0 and 𝜔 = ∞ the ellipse degenerates into straight lines of pure spring
behavior.
Figure 2.6 shows the frequency dependence of the equivalent parameters
𝑘𝑒 and 𝑐𝑒 . The model shown in Fig.2.4 can be improved, connecting in series or in
parallel several similar units with different time constants, to obtain a better
correlation with the test data. The same result can be obtained using a suitable
hereditary function 𝐺(𝑡) in equation (2.16). However, the complexity of these
models has limited their use both analytically (in dynamic analyses) and
experimentally (in characterizing materials). For engineering applications, it is

26
customary to use simple models, yet present the behavior of real materials with
sufficient accuracy [11].

2.1.4. Other Damping Mechanisms

The Coulomb damping or dry friction damping is a non-linear damping


mechanism encountered in many structural systems with sliding parts. The
Coulomb damping force is of constant amplitude, independent of displacement and
opposite to velocity
𝑧̇
𝑓𝑑 = 𝑅 = 𝑅 sgn(𝑧̇ ).
|𝑧̇ |

The energy dissipated per cycle by a Coulomb damper, in which a force 𝑓𝑑


acts through a relative displacement 𝑧 = 𝑧̂𝑐𝑜𝑠𝜔𝑡, is given by [104]
𝑊𝑑 = 4𝑅𝑧̂ .
It is independent of frequency.
The quadratic damping or velocity squared damping is a non-linear
damping mechanism which can be associated with the turbulent flow of a fluid
through an orifice. It practically occurs at relatively high flow velocities. The
damping force is proportional to the square of the relative velocity across the
damper

27
𝑓𝑑 = 𝑐2 𝑧̇ 2 sgn(𝑧̇ ) .
The energy dissipated per cycle by a quadratic damper experiencing a
harmonic relative displacement 𝑧 = 𝑧̂𝑐𝑜𝑠𝜔𝑡 is given by
8
𝑊𝑑 = 3 𝑐2 𝜔2 𝑧̂ 3 . (2.24)

It is dependent on both the frequency and amplitude of vibration.


Usually, using the concept of equivalent viscous damping [104], the non-
linear damping force is replaced by a linear viscous force, so that the energy
dissipated per cycle of vibration by the non-linear damping element be equal to the
energy dissipated by an equivalent viscous damper experiencing the same harmonic
relative displacement.
The equivalent coefficient of viscous damping is in this case
𝑊 8
𝑐𝑒𝑞 = 𝜋𝜔𝑧̂𝑑2 = 3𝜋 𝑐2 𝜔𝑧̂ . (2.25)
Generalizing the concept of equivalent damping, the analysis of damped
vibrations is simplified using two damping models – viscous or hysteretic. The
energy dissipated per cycle of vibration by friction and other resistances, including
the radiation damping (by waves in infinite continua), is considered the same as the
energy dissipated by a single mechanism – viscous or hysteretic – in vibrations of
the same amplitude. We obtain either an equivalent viscous damping coefficient or
an equivalent hysteretic damping coefficient [74], quantities which are functions of
frequency and vibration amplitude. The range of applicability of this hypothesis has
to be determined experimentally.

2.2. EXCITATION WITH HARMONIC FORCES APPLIED TO THE MASS

The single-degree-of-freedom system with fixed base, acted upon by a time


variable force applied to the mass (Fig.2.7) is a suitable model to describe the
response of a complex structure in a principal mode of
vibration. It can also represent the behavior of the simplest
experimental set up used for measuring the characteristics
of a resilient pad used in vibration isolation.
For this reason, it was considered useful to present
in detail the exact relationships used to calculate the
system parameters (valid also for highly damped
structures) for both hysteretic and linear viscous damping.
For simplification, the same symbols are used to express
the response for both damping models.
In the following, frequency response curves are not presented in
dimensionless coordinates. Experimentally measurable quantities are preferred as
coordinates, as they are directly used in the formulae used in practice.

28
2.2.1. Identification of Systems with Hysteretic Damping

2.2.1.1. Frequency response functions

Consider the system with hysteretic damping from Fig.2.7. The equation
of (harmonic) motion of the mass can be written as

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘𝑥 = 𝑓̂ 𝑒 𝑖𝜔𝑡 (2.26)
where 𝑚 is the mass, 𝑘 – stiffness, ℎ - coefficient of equivalent hysteretic damping,
𝑥 – mass displacement, 𝜔 – forcing frequency, 𝑡 – time, 𝑓̂ – amplitude of the
excitation force.
The steady-state response (measured after the transient disappeared) can be
expressed in terms of displacement, velocity or acceleration
𝑥(𝑡) = 𝑥̂ 𝑒 𝑖 (𝜔𝑡+𝜑𝑑) = 𝑥̃𝑒 𝑖𝜔𝑡 ,
𝑥̇ (𝑡) = 𝑥̇̂ 𝑒 𝑖 (𝜔𝑡+𝜑𝑣 ) = 𝑥̇̃ 𝑒 𝑖𝜔𝑡 , (2.27)
𝑥̈ (𝑡) = 𝑥̈̂ 𝑒 𝑖 (𝜔𝑡+𝜑𝑎) = 𝑥̈̃ 𝑒 𝑖𝜔𝑡 ,
where 𝜑𝑑 , 𝜑𝑣 and 𝜑𝑎 are phase shifts with respect to the force.
The quantities from equations (2.27) are related by the following
relationships
𝑥̈̃ = 𝑖𝜔 𝑥̇̃ = −𝜔2 𝑥̃,
𝑥̈̂ = 𝜔𝑥̇̂ = 𝜔2 𝑥̂, (2.28)
𝜋
𝜑𝑎 = 𝜑𝑣 + 2 = 𝜑𝑑 + 𝜋.
The complex frequency response functions are presented in Table 2.1 as
𝜔 𝑘
functions of the dimensionless frequency 𝛺 = (where 𝜔𝑛 = √𝑚 is the natural
𝜔𝑛

frequency of the associated conservative system) and 𝑔 = is the equivalent
𝑘
hysteretic damping factor.

2.2.1.2. Identification based on the diagrams of receptance

Peak amplitude method

The variation with frequency of the modulus of the complex receptance is


illustrated by the resonance curve in Fig.2.8. Usually, this is plotted for 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡.
as a diagram of the displacement amplitude versus frequency.
𝑘
Amplitude resonance occurs at the frequency 𝜔𝑟𝑒𝑠 = 𝜔𝑛 , where 𝜔𝑛 = √𝑚
1
is the undamped natural frequency and the receptance has the peak value 𝛼𝑟𝑒𝑠 = ℎ.

29
Table 2.1
Frequency
response Definition Symbol Complex function Modulus Phase angle
function
Receptance 𝑑𝑖𝑠𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 𝑥̃ 1 1 𝑥̂ 1⁄𝑘 −g
(admittance, ∝ 𝛼̅ = = 𝛼= = 𝑡𝑎𝑛 𝜑d =
compliance)
𝑓𝑜𝑟𝑐𝑒 𝑓̂ 𝑘 1 − 𝛺2 + 𝑖𝑔 𝑓̂ √(1 − 𝛺2 )2 + 𝑔2 1 − Ω2

𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦 𝑥̇̃ 1 𝛺 𝑥̇̂ 1 𝛺 1 − 𝛺2


Mobility ℳ ̅ =
ℳ = ℳ= = 𝑡𝑎𝑛 𝜑𝑣 =
𝑓𝑜𝑟𝑐𝑒 𝑓̂ √𝑘𝑚 𝑔 − 𝑖(1 − 𝛺2 ) 𝑓̂ √𝑘𝑚 √(1 − 𝛺2 )2 + 𝑔2 𝑔

Inertance 𝑎𝑐𝑐𝑒𝑙𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑥̈̃ 1 𝛺2 𝑥̈̂ 1 𝛺2


η 𝜂̅ = =− 𝜂= = 𝜑𝑎 = 𝜑𝑑 + 𝜋
(accelerance) 𝑓𝑜𝑟𝑐𝑒 𝑓̂ 𝑚 1 − 𝛺2 + 𝑖𝑔 𝑓̂ 𝑚 √(1 − 𝛺2 )2 + 𝑔2

Dynamic 𝑓𝑜𝑟𝑐𝑒 𝑓̃ 𝑓̂ 𝑔
β 𝛽= = 𝑘(1 − 𝛺2 + 𝑖𝑔) 𝛽= = 𝑘√(1 − 𝛺2 )2 + 𝑔2 𝑡𝑎𝑛 𝜓𝑑 =
stiffness 𝑑𝑖𝑠𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 𝑥̂ 𝑥̂ 1 − 𝛺2

Mechanical 𝑓𝑜𝑟𝑐𝑒 𝑓̃ 𝑔 + 𝑖(𝛺2 − 1) 𝑓̂ √(1 − 𝛺2 )2 + 𝑔2 𝛺2 − 1


impedance 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦
𝑍 𝑍̅ = = √𝑘𝑚 𝑍= = √𝑘𝑚 𝑡𝑎𝑛 𝜓𝑣 =
𝑥̇̂ 𝛺 𝑥̇̂ 𝛺 𝑔

𝑓𝑜𝑟𝑐𝑒 𝑓̃ 1 − 𝛺2 + 𝑖𝑔 𝑓̂ √(1 − 𝛺2 )2 + 𝑔2
Apparent mass μ µ̅ = = −𝑚 µ= =𝑚 𝜓𝑎 = 𝜓𝑑 + 𝜋
𝑎𝑐𝑐𝑒𝑙𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑥̈̂ 𝛺2 𝑥̈̂ 𝛺2

Force 𝑡𝑟𝑎𝑛𝑠𝑚𝑖𝑡𝑡𝑒𝑑 𝑓𝑜𝑟𝑐𝑒 𝑓̃𝑇 1 + 𝑖𝑔 𝑓̂𝑇 1 + 𝑔2 −𝑔𝛺2


𝒯 𝒯̅ = = 𝒯= =√ 𝑡𝑎𝑛 𝜑 𝑇 =
transmissibility 𝑎𝑝𝑝𝑙𝑖𝑒𝑑 𝑓𝑜𝑟𝑐𝑒 𝑓̂ 1 − 𝛺2 + 𝑖𝑔 𝑓̂ (1 − 𝛺2 )2 + 𝑔2 1 − 𝛺 2 + 𝑔2

30
The crossing points of the resonance curve with horizontal lines of
1
ordinates 𝜀 𝛼𝑟𝑒𝑠 have frequencies 𝜔1𝜀 and 𝜔2𝜀 (1 < 𝜀 < √2). The damping factor
𝑔 is given by [46]
𝜔22𝜀 −𝜔21𝜀
𝑔= . (2.29)
2𝜔2𝑛 √𝜀2 −1
For lightly damped systems 𝜔1𝜀 ≅ 𝜔2𝜀 ≅ 𝜔𝑛 and equation (2.29) becomes
𝜔2𝜀 −𝜔1𝜀 ∆𝜔𝜀
𝑔≅ = . (2.29, a)
𝜔𝑛 √𝜀2 −1 𝜔𝑛 √𝜀2 −1
∆𝜔𝜀
Plotting the ratio 𝜔𝑛
versus √𝜀 2 − 1, a straight line can be drawn through
the data points, of slope 𝑔 (Fig.2.9).

For 𝜀 = √2, points 𝐵𝜀 and 𝐶𝜀 coincide with the half power points, of
frequencies
𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔,
in which the power dissipated by damping is half the power dissipated at resonance.
The damping factor is given by the familiar formula of Broadbent and
Hartley [18]
𝜔22 −𝜔21 𝜔2 −𝜔2
𝑔=
2𝜔2𝑛
= 𝜔22+𝜔12. (2.30)
2 1
For lightly damped systems, equation (2.30) becomes
𝜔2 −𝜔1
𝑔≅
𝜔𝑛
= ∆𝜔
𝜔
. (2.31)
𝑛

It is possible to compute the stiffness


1
𝑘 = 𝑔𝛼 (2.32)
𝑟𝑒𝑠
then the mass
𝑘
𝑚 = 𝜔2 . (2.33)
𝑛

31
Sometimes, it is advantageous to plot the receptance curve in log-log scales
(Fig.2.10).
When the curve is plotted for a relatively wide frequency range outside
1
resonance, the “quality factor” 𝑄 = 𝑔 , which is a measure of the sharpness of
resonance, is given by the vertical distance between the ordinate at resonance and
the ordinate at 𝜔 = 0. If the low frequency information is insufficient, the same
method can be used observing that the ordinate at the frequency 𝜔𝑠 = √2𝜔𝑛 is
equal to the ordinate at 𝜔 = 0.
When only a small part of the curve is available near the resonance, the
half-power points are located at 3 dB below the level at resonance and the damping
is given by equation (2.31).
It is recommended to use annotated plots, including constant mass lines of
slope −12 𝑑𝐵/𝑜𝑐𝑡𝑎𝑣𝑒 and horizontal constant stiffness lines. The mass 𝑚 and
stiffness 𝑘 of the single-degree-of-freedom system are deduced from the asymptotic
behavior of its response at off-resonance frequencies.

Phase angle method

The variation of the phase angle with frequency is shown in Fig.2.11. The
half power points 𝐵 and 𝐶, defined by frequencies 𝜔1 and 𝜔2 , correspond to phase
angles of −45° and −135°, respectively; thus, the damping factor 𝑔 can be
evaluated using equation (2.30).

The damping factor can be calculated from [85]


2
𝑔= 𝑑𝜑 (2.34)
𝜔𝑛 (− 𝑑 )
𝑑𝜔 𝜔=𝜔𝑛
𝑑𝜑𝑑
where ( ) is the slope of the line tangent to the phase angle diagram at the
𝑑𝜔 𝜔=𝜔𝑛
inflection point 𝑀, of frequency 𝜔𝑛 .
The phase angle diagram yields no information for determining 𝑘 and 𝑚.

32
In-phase response method

Figure 2.12 shows the variation with 𝜔 of the in-phase (real) component of
𝑥̂ 𝑑𝛼𝑅
receptance 𝛼𝑅 = 𝑓̂ 𝑐𝑜𝑠𝜑𝑑 . Resonance is located at point 𝑀 where 𝛼𝑅 = 0 and
𝑑𝜔
is a maximum. The half power points 𝐵 and 𝐶 correspond to |𝛼𝑅 |𝑚𝑎𝑥 so that
equation (2.30) can again be used for evaluating the damping factor.
Best results are sometimes obtained by locating the resonance at the
𝑑2 𝛼 𝑅
inflection point, where = 0, and calculating the damping factor from the
𝑑𝜔 2
slope of the line tangent at that point [93]
2𝛼𝑟𝑒𝑠
𝑔= 𝑑𝛼𝑅 . (2.35)
𝜔𝑛 (− )
𝑑𝜔 𝜔=𝜔𝑛

In equation (2.35) 𝛼𝑟𝑒𝑠 is measured as the distance between the horizontal


lines tangent at the points 𝐵 and 𝐶.

Out-of-phase response method

Figure 2.13 shows the diagram of the out-of-phase (imaginary) component


𝑥̂
𝛼𝐼 = 𝑓̂
𝑠𝑖𝑛𝜑𝑑plotted against the frequency 𝜔. Resonance is located at the point 𝑀
𝑑𝛼
𝐼
where 𝛼𝐼 has a minimum value and = 0. It can be seen that |𝛼𝐼 | peaks more
𝑑𝜔
sharply than the total response 𝛼 and that the value at resonance is equal to the total
response 𝛼𝑟𝑒𝑠 since the in-phase response is zero.
The half power points 𝐵 and 𝐶 correspond to frequencies 𝜔1 and 𝜔2 at
which the quadrature response is half the peak amplitude; thus equations (2.30),
(2.32) and (2.33) can once again be used for evaluating the damping factor. Using
the diagrams of both in-phase and quadrature components of receptance, we obtain

33
2(𝛼𝐼 )𝜔=𝜔𝑛
𝑔= 𝑑𝛼𝑅 . (2.36)
𝜔𝑛 (− )
𝑑𝜔 𝜔=𝜔𝑛

Polar plot method

By eliminating 𝜔 between the expressions of 𝛼𝑅 and 𝛼𝐼 , we obtain a circle


(Fig.2.14,a) of equation
1 2 1 2
(𝛼𝐼 + 2ℎ) + 𝛼𝑅2 = (2ℎ) , (2.37)

which is the locus of the tip of receptance vector in the complex plane, called vector
diagram, polar plot, Nyquist plot or transfer locus.
Each point along the circle corresponds to a given frequency 𝜔. At 𝜔 = 0,
1 −𝑔
the tip of the vector is at the point 𝑃𝑜 [ ,
(1+𝑔 )𝑘 (1+𝑔2 )𝑘
2
]. With increasing
frequency 𝜔, point 𝑃 moves along the circle clockwise, and for 𝜔 → ∞ it coincides
with the origin 𝑂.
Both amplitude and phase resonance occur at the point 𝑀, of frequency 𝜔𝑛 ,
at the intersection of the circle with the negative imaginary semiaxis. The circle
1
diameter is ̅̅̅̅̅
𝑂𝑀 = . ℎ

The half power points, which correspond to frequencies 𝜔1 and 𝜔2 , are


the ends of the diameter 𝐵𝐶 perpendicular to 𝑂𝑀. The damping factor can be
determined from equations (2.30).
Kennedy and Pancu [61] have shown that if the system from Fig.2.7 is
subjected to a harmonic force of constant amplitude and if the displacement
response is plotted point by point at equal frequency increments ∆𝜔 and a smooth
curve is drawn through these points, the length of the arc of circle ∆s between two
successive points is a maximum at resonance (Fig.2.14, b).
This criterion of resonance location is based on the fact that the derivative

34
𝑑𝑠 1 𝑑𝜑𝑑 1
𝜔2
=− 𝜔2
= 2 = 𝑘𝛼 2 (2.38)
𝑑( 2 ) 𝑔𝑘 𝑑( 2 ) 𝜔2
𝜔𝑛 𝜔𝑛 𝑘[(1− 2 ) +𝑔2 ]
𝜔𝑛
is a maximum at 𝜔 = 𝜔𝑛 . In equations (2.38) 𝑑𝑠 is the arc length corresponding to
a variation 𝑑𝜑𝑑 of the phase angle (Fig.2.15):
1 𝑑𝜑𝑑
𝑑𝑠 = (𝑟𝑎𝑑𝑖𝑢𝑠) × (𝑐𝑒𝑛𝑡𝑟𝑎𝑙 𝑎𝑛𝑔𝑙𝑒) = 2𝑔𝑘 (2𝑑𝜑𝑑 ) = 𝑔𝑘
.
̅̅̅̅ =
In the diagram shown in Fig.2.16, 𝜃 is the angle between the vectors 𝑂𝑃
̅̅̅̅̅
𝛼𝑃 and 𝑂𝑀 = 𝛼𝑟𝑒𝑠 , of frequencies 𝜔𝑃 and 𝜔𝑛 .
The damping factor is given by the equation [143]
𝜔𝑝 2
𝑔 = |1 − (𝜔 ) | cot𝜃. (2.39)
𝑛

Alternative expressions have been used by Natke [77]


1
2 −
𝜔𝑝 2 𝛼𝑟𝑒𝑠
2
𝑔 = |1 − (𝜔 ) | (| 𝛼 | − 1) (2.40)
𝑛 𝑝

by Wolfe and Kirkby [136] for light damping


∆𝜔 ∆𝜔
𝑔 ≅ 𝜔 (2 + 𝜔 ) cot𝜃 (2.41)
𝑛 𝑛

where ∆𝜔 = 𝜔𝑝 − 𝜔𝑛 , and by Kennedy and Pancu [61], also for light damping
1

𝛼𝑝 ∆𝜔 𝛼𝑝 2 2
𝑔 ≅ 2𝛼 |𝜔 | (1 − |𝛼 | ) . (2.42)
𝑟𝑒𝑠 𝑛 𝑟𝑒𝑠

In all expressions of the damping factor only dimensionless ratios are used,
requiring no calibration of the measurement system. This is necessary only for the
estimation of the stiffness 𝑘 and mass 𝑚.
Other identification methods, based on the graphical analysis of the polar
plots of the frequency response functions, are described in § 5.3.5.

2.2.1.3. Identification based on the diagrams of mobility

Peak velocity method

The variation with frequency of the modulus of the complex mobility (see
Table 2.1) is illustrated in Fig.2.17. The peak value
1 1
ℳ𝑚𝑎𝑥 =
√𝑘𝑚 √2(√1+𝑔2
−1)

̅̅̅̅̅̅̅̅̅
occurs at the frequency 𝜔𝑟𝑣 = 𝜔𝑛 ∜1 + 𝑔2 .

35
The frequency 𝜔𝑛 corresponds to the point 𝑇 in which a straight line
passing through the origin is tangent to the resonance curve, and the mobility is
1 1
ℳ𝑇 = .
√𝑘𝑚 𝑔

Let 𝑃 and 𝑄 be the crossing points of the mobility curve from Fig.2.17 with
the constant velocity line (∆) of equation ℳ = 𝑎𝜔. Their frequencies 𝜔𝑝 and 𝜔𝑞
satisfy the equations
𝜔𝑞2 + 𝜔𝑝2 = 2𝜔𝑛2 ,
(2.43)
1
𝜔𝑞2 − 𝜔𝑝2 = 2𝜔𝑛2 √ − 𝑔2 .
𝑎2 𝑘 2

Let ℳ𝑞 be the mobility at the frequency 𝜔𝑞 . Because


ℳ𝑞 1 ℳ𝑇
𝑎= , = ,
𝜔𝑞 𝑘𝑔 𝜔𝑛
denoting
ℳ𝑇 𝜔𝑞
𝑒= (2.44)
ℳ𝑞 𝜔𝑛
the damping factor is given by the expression
𝜔𝑞2 −𝜔𝑝
2
1
𝑔= . (2.45)
𝜔𝑞2 +𝜔𝑝
2 √𝑒 2
−1

The line 𝑂𝑆, tangent at the origin to the resonance curve, has the slope
1
and intersects the mobility curve at the point 𝑆, of frequency 𝜔𝑠 = √2𝜔𝑛 .
𝑘 √1+𝑔2
In this case, formula (2.45) becomes
1
𝑔= (2.46)
√2𝑒𝑜2 −1

36
ℳ𝑇
where 𝑒𝑜 = is the ratio of mobilities at the points 𝑇 and 𝑆.
ℳ𝑠
The stiffness is
𝜔
𝑘 = 𝑔ℳ𝑛 . (2.47)
𝑇

Good results are obtained for systems with low natural frequency and high
damping, for which the origin of axes is not far from the resonance region. A log-
log plot is recommended (Fig.2.18). In this case, though the origin of frequencies
is not localized, the line 𝑂𝑆 coincides with the low frequency asymptote. A fairly
good accuracy is obtained by locating the point 𝑇 by the line tangent to the mobility
curve parallel to 𝑂𝑆. However, in most cases point 𝑇 cannot be distinguished from
point 𝑅 of maximum mobility.

Phase angle method

The variation with frequency of the phase angle 𝜑𝑣 between the force and
the velocity of the mass is shown in Fig.2.19. Phase resonance occurs at 𝜔 = 𝜔𝑛
where 𝜑𝑣 = 0. The resonance frequency can be located at the crossing of the phase
angle diagram with the straight line 𝜑𝑣 = 0 at the point 𝑀.
The half power points 𝐵 and 𝐶, defined by frequencies 𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔,
correspond to phase angles 𝜑𝑣 = ±45°; thus, the damping factor 𝑔 can be
calculated using equations (2.30).
If only a small part of the phase angle diagram is available near resonance,
the damping factor can be calculated replacing 𝜑1 by 𝜑𝑣 in equation (2.34).

In-phase response method

Figure 2.20 shows the variation of the real component of mobility ℳ𝑅 =


ℳ𝑐𝑜𝑠𝜑𝑣 with 𝜔 (for 𝑔 = 𝑐𝑜𝑛𝑠𝑡. ).
The tangent line 𝑂𝑇 locates the point 𝑇, of frequency 𝜔𝑛 , where the real
1
mobility is ℳ𝑇 = . The line (∆) intersects the curve at the points 𝑃 and 𝑄,
𝑔√𝑘𝑚
of frequencies 𝜔𝑝 and 𝜔𝑞 .
The damping factor can be calculated from
𝜔𝑞2 −𝜔𝑝
2
1
𝑔= (2.48)
𝜔𝑞2 +𝜔𝑝
2
√𝑒𝑅 −1
in which
ℳ𝑇 𝜔𝑞
𝑒𝑅 =ℳ 𝜔
, (2.49)
𝑅𝑞 𝑛

where ℳ𝑅𝑞 is the ordinate at 𝜔𝑞 .

37
Using the tangent line 𝑂𝑆 at the origin, point 𝑆 has an abscissa 𝜔𝑠 = √2𝜔𝑛
and formula (3.48) becomes
1
𝑔= , (2.50)

√√2 ℳ 𝑇 −1
𝑅𝑠
where ℳ𝑅𝑠 is the ordinate of point 𝑆 and ℳ𝑇 is the ordinate of 𝑇.
The stiffness is calculated from (2.47).

Out-of-phase response method

Figure 2.21 shows the graph of the imaginary component of complex


mobility ℳ𝐼 = ℳ𝑠𝑖𝑛𝜑𝑣 against 𝜔 (for 𝑔 = 𝑐𝑜𝑛𝑠𝑡. ). The curve intersects the
frequency axis at 𝜔𝑛 . The slope of the line tangent at the crossing point is
𝑑ℳ 2
( 𝑑𝜔𝐼) =−
𝑘𝑔2
. (2.51)
𝜔=𝜔𝑛
Equation (2.51) cannot be used for the evaluation of damping because 𝑘 is
unknown.
Using the diagram of the in-phase component (Fig.2.20), we obtain the
following formula

38
2 (ℳ𝑅 )𝜔=𝜔𝑛
𝑔= 𝑑ℳ𝐼 . (2.52)
𝜔𝑛 (− )
𝑑𝜔 𝜔=𝜔𝑛

Polar plot method

By eliminating 𝜔 between ℳ𝑅 and ℳ𝐼 , we obtain a curve of equation


ℳ𝑅2 ℳ𝑅 ℳ𝐼
(ℳ𝑅2 + ℳ𝐼2 )2 − + =0 (2.53)
𝑔2 𝑘𝑚 𝑔𝑘𝑚

which can be approximated with a circle only for light damping.


Lines 𝑂𝐵 and 𝑂𝐶, inclined ±45° with respect to the real axis, locate the
frequencies 𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔, which can be used to calculate the damping factor
from the second equation (2.30)
𝜔22 −𝜔21
𝑔=
𝜔22 +𝜔21
.
The polar plot crosses the real axis at the point 𝑀, of frequency 𝜔𝑛 .
̅̅̅̅̅, we can calculate 𝑘 = 𝜔𝑛 , then 𝑚 = 𝑘2 .
Measuring the segment 𝑂𝑀 ̅̅̅̅̅
𝑔𝑂𝑀 𝜔𝑛
For highly damped systems, locating on the diagram the point 𝑅 of
maximum response we can determine the frequency 𝜔𝑟𝑣 .
The damping factor is
𝜔 4
𝑔 = √( 𝜔𝑟𝑣 ) − 1 . (2.54)
𝑛

2.2.1.4. Identification based on the diagrams of inertance

Peak acceleration method

The variation with frequency of the modulus of the complex inertance (see
Table 2.1) is illustrated in Fig.2.23. In practice, it is customary to use the graph of
the acceleration versus frequency, for excitation with a harmonic force of constant
amplitude.
Acceleration resonance occurs at the frequency 𝜔𝑟𝑎 = 𝜔𝑛 √1 + 𝑔2 , where
the inertance has the peak value
1 1
𝜂𝑚𝑎𝑥 = √ +1. (2.55)
𝑚 𝑔2
1
The frequencies of the points 𝐵′ and 𝐶′, in which the inertance is 𝜂𝑚𝑎𝑥 ,
√2
are respectively
1+𝑔2 1+𝑔2
𝜔ˈ = 𝜔𝑛 √ 1+𝑔 , 𝜔" = 𝜔𝑛 √ 1−𝑔 , (2.56)

The damping can be calculated using a formula similar to (2.30)

39
2 2
𝜔" −𝜔ˈ
𝑔= 2 2 , (2.57)
𝜔" +𝜔ˈ
the mass from (2.55) and the stiffness from
2 𝑚
𝑘 = 1+𝑔2 𝜔𝑟𝑎 . (2.58)

1
The asymptote line 𝜂∞ = intersects the curve at the point 𝑆, of
𝑚
𝜔𝑟𝑎
frequency 𝜔𝑠 =
√2
. The damping factor can also be calculated as
1
𝑔= (2.59)
√𝜀 2 −1
𝜂𝑚𝑎𝑥
where 𝜀 = 𝜂∞
. Formula (2.59) is useful especially when the mobility curve from
Fig.2.23 is plotted in logarithmic coordinates.

Phase angle method

The variation of the inertance phase angle with frequency is represented


graphically in Fig.2.24. The horizontal line 𝜑𝑎 = 90° intersects the curve at the
point 𝑀, of frequency 𝜔𝑛 . The lines 𝜑𝑎 = 135° and 𝜑𝑎 = 45° cross the phase
angle curve at the points 𝐵 and 𝐶, of frequencies 𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔 from which
we obtain the damping factor given by (2.30)
𝜔22 −𝜔12
𝑔= . (2.60)
𝜔22 +𝜔12

Polar plot method

The polar plot of the complex inertance (Fig.2.25) is a circle of equation


1 2 1 2 1+𝑔2
(𝜂𝑅 − 2𝑚) + (𝜂𝐼 − 2𝑔𝑚) = 4𝑔2 𝑚2 . (2.61)

40
It is obtained by eliminating the frequency between the expressions of the
vector components of the inertance 𝜂𝑅 = 𝜂 𝑐𝑜𝑠𝜑𝑎 and 𝜂𝐼 = 𝜂 𝑠𝑖𝑛𝜑𝑎 .
Phase resonance occurs at the frequency 𝜔𝑛 , at the point 𝑀 of intersection
with the imaginary axis. The diameter OR, equal to the peak inertance, is inclined
1
an angle 𝜃 = 𝑡𝑎𝑛−1 𝑔 with respect to the chord 𝑂𝑀 = 𝑚𝑔 . Point 𝑅 can be localized
using the Kennedy-Pancu criterion, where for regular increments of frequency, the
arc of circle between successive points has a maximum length.
The diameter 𝐵′𝐶′, perpendicular to 𝑂𝑅 , locates on the circle the
frequencies 𝜔′ and ω" given by equations (2.56) and used in formula (2.57) to
calculate the damping factor.
The mass is given by
1
𝑚 = ̅̅̅̅̅ (2.62)
𝑔 𝑂𝑀
then the stiffness by 𝑘 = 𝑚 𝜔𝑛2 .

A different method is based on the construction from Fig.2.26. The chords


𝑂𝐶 and 𝑂𝐵, inclined 45° and 135°, respectively, with the positive real semiaxis,
determine on the circle the frequencies 𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔 and the damping factor
can be calculated from equation (2.60).

2.2.1.5. Identification based on the diagrams of dynamic stiffness

Minimum force method

Figure 2.27 illustrates the variation with frequency of the dynamic stiffness
modulus (see Table 2.1). This diagram is obtained using harmonic excitation with
constant displacement amplitude and variable force amplitude.

41
At resonance, the force required to maintain a constant displacement
amplitude has a minimum and 𝛽𝑚𝑖𝑛 = 𝑘𝑔 = ℎ. Point 𝑀 has a frequency 𝜔𝑛 . The
horizontal line 𝛽 = √2𝛽𝑚𝑖𝑛 intersects the curve at the points 𝐵 and 𝐶, of abscissae
𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔. The damping factor is given by equation (2.60), the stiffness by
𝛽𝑚𝑖𝑛 𝑘
𝑘= 𝑔
and the mass by 𝑚 = 𝜔2 .
𝑛

Polar plot method

The vector diagram of the complex dynamic stiffness (Fig.2.28) is a


straight line of equation 𝛽𝐼 = ℎ , parallel to the real axis. The line intersects the
imaginary axis at the point 𝑀 of frequency 𝜔𝑛 . The frequencies 𝜔1 and 𝜔2 of the
half power points are determined at the crossing points with the lines 𝑂𝐵 and 𝑂𝐶,
inclined 45° and 135°, respectively, with the positive real semiaxis. They are used
to calculate the damping factor (2.60).

The arc of circle of radius ̅̅̅̅


𝑂𝑁 = 𝜀 ̅̅̅̅̅
𝑂𝑀 intersects the dynamic stiffness line
at the frequencies 𝜔𝑝 and 𝜔𝑞 which can be used to estimate the damping factor
𝜔𝑞2 −𝜔𝑝
2
1
𝑔 = 𝜔2 +𝜔2 √𝜀2
−1
. (2.63)
𝑞 𝑝

For better accuracy, it is recommended to calculate 𝑔 for different values


of the ratio 𝜀 and take an average value.
Any deviation of the dynamic stiffness diagram from the shape shown in
Fig.2.28 invalidates the selected model, especially the damping type.

2.2.1.6. Identification based on the diagrams of mechanical impedance

Minimum force method

The variation with frequency of the modulus of the complex mechanical


impedance (see Table 2.1) is shown in Fig.2.29. Usually, the graph illustrates the

42
variation of the excitation force during a frequency sweep with constant velocity
amplitude.
The minimum value

𝑍𝑚𝑖𝑛 = √𝑘𝑚√2(√1 + 𝑔2 − 1)

occurs at the point 𝑅 of frequency 𝜔𝑟𝑣 = 𝜔𝑛 ∜1 ̅̅̅̅̅̅̅̅̅


+ 𝑔2 .
The tangent line 𝑂𝑇 locates the point 𝑇 of frequency 𝜔𝑡 = 𝜔𝑛 √1 + 𝑔2 and
ordinate 𝑍𝑡 = 𝑔√𝑘𝑚. The asymptote 𝑂𝑆 intersects the curve in 𝑆, of frequency
𝜔𝑡 1+𝑔2
𝜔𝑠 = and ordinate 𝑍𝑠 = √𝑘𝑚√ 2
. The damping factor is given by
√2
1
𝑔= (2.64)
√2𝜀2 −1
𝑍𝑠
where 𝜀 =
𝑍𝑡
.

𝑍
The mass 𝑚 is equal to the slope of the line 𝑂𝑆, 𝑚 = 𝜔𝑠 . Good results are
𝑠
obtained using logarithmic scales.

Polar plot method

The polar diagram of the complex mechanical impedance is a curve of


equation (Fig.2.30)
𝑍𝑅2 + 𝑔𝑍𝑅 𝑍𝐼 − 𝑔2 𝑘𝑚 = 0 . (2.65)
Equation (2.65) is obtained by eliminating the frequency between the
expressions of the vector components 𝑍𝑅 = 𝑍 𝑐𝑜𝑠𝜓𝑣 and 𝑍𝐼 = 𝑍 𝑠𝑖𝑛𝜓𝑣 .
Phase resonance occurs at the intersection of the mechanical impedance
plot with the horizontal axis in 𝑀. Amplitude resonance occurs at 𝑅, of frequency

43
𝜔𝑟𝑣 , the point of tangency with a circle of radius 𝑂𝑅 and center at the origin. The
frequencies 𝜔1 and 𝜔2 of the half power points 𝐵 and 𝐶 are located by the lines
𝑂𝐵 and 𝑂𝐶 inclined ±45° with respect to the positive real semi-axis. They are used
in equation (2.60) to calculate the damping factor.
̅̅̅̅̅
𝑂𝑀
The mass is given by 𝑚 = and the stiffness by 𝑘 = 𝑚𝜔𝑛2.
𝑔𝜔𝑛

2.2.1.7. Identification based on the diagrams of apparent mass

Minimum force method

The variation with frequency of the modulus of the complex apparent mass
(see Table 2.1) is sketched in Fig.2.31. Usually, the graph illustrates the variation
of the excitation force during a frequency sweep with constant acceleration
amplitude.

The damping can be calculated using the half power points formula.
Acceleration resonance occurs at point 𝑅 of frequency 𝜔𝑟𝑎 = 𝜔𝑛 √1 + 𝑔2 and
𝑔 𝜇
minimum ordinate 𝜇𝑚𝑖𝑛 = 𝑚 2
. At the intersection with the line 𝜇 = 𝑚𝑖𝑛 ,
√1+𝑔 √2
points 𝐵′ and 𝐶′ locate the frequencies 𝜔′ and ω" (2.56) which can be used in (2.57)
√1+𝑔2
to calculate the damping factor. The mass is given by 𝑚 = 𝜇𝑚𝑖𝑛 and the
𝑔
𝑚 2
stiffness by 𝑘 = 𝜔𝑟𝑎 .
1+𝑔2

Polar plot method

Eliminating the frequency between the real and imaginary components of


the complex apparent mass 𝜇𝑅 = 𝜇 𝑐𝑜𝑠𝜓𝑎 and 𝜇𝐼 = 𝜇 𝑠𝑖𝑛𝜓𝑎 , we obtain the
following equation of a straight line [16]

44
1
𝜇𝑅 − 𝑔 𝜇𝐼 − 𝑚 = 0 (2.66)

represented graphically in Fig.2.32.


The natural frequency 𝜔𝑛 is obtained at the point of intersection with the
negative imaginary semiaxis. The circle with the center in the origin, tangent to the
plot, locates the point 𝑅 of frequency 𝜔𝑟𝑎 . The circle centered at 𝑅 and radius 𝑂𝑅
intersects the plot at points 𝐵′ and 𝐶′, of frequencies 𝜔′ and 𝜔", which can be used
in (2.57) to determine the damping factor.
Lines 𝑂𝐶 and 𝑂𝐵 drawn at inclinations 45° and 135° , respectively, with
the positive real semiaxis, locate the frequencies 𝜔1 and 𝜔2 which can be used in
equation (2.60) to calculate the damping factor.
̅̅̅̅̅
𝑂𝑀
The mass is 𝑚 = and the stiffness is 𝑘 = 𝑚 𝜔𝑛2 .
𝑔

2.2.1.8. Identification based on the diagram of force transmissibility

Figure 2.33 is a sketch of the polar plot of the complex force


transmissibility, for a given value of 𝑔. It is a circle of equation [90]
1 2 1 2 1+𝑔2
(𝒯𝑅 − 2) + (𝒯𝐼 + 2𝑔) = . (2.67)
4𝑔2

̅̅̅̅̅ = √ 12 + 1 corresponds to 𝜔𝑛 which can be located


The peak value 𝑂𝑀 𝑔
using the Kennedy-Pancu criterion.

The diameter 𝑂𝑀 is inclined an angle 𝜃 = 𝑡𝑎𝑛−1 𝑔 with the negative


imaginary semiaxis. The diameter 𝐵𝐶 , perpendicular to 𝑂𝑀, locates on the circle
the frequencies 𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔. The damping factor can be calculated from
(2.60). Other parameters cannot be determined in this case.

45
2.2.2. Identification of Systems with Linear Viscous Damping

2.2.2.1. Frequency response functions

Consider the system with linear viscous damping from Fig.2.34. The
equation of motion of the mass can be written under the form
𝑚𝑥̈ + 𝑐𝑥̇ + 𝑘𝑥 = 𝑓(𝑡), (2.68)
where 𝑚 is the mass, 𝑘 – stiffness, ℎ - coefficient of equivalent viscous damping.
For harmonic excitation 𝑓(𝑡) = 𝑓̂ 𝑒 𝑖𝜔𝑡 , the steady-state response can be
expressed as displacement, velocity or acceleration (2.27).
The complex frequency response functions presented in Table 2.2 are
𝜔
expressed in terms of the dimensionless frequency 𝛺 = 𝜔 and the damping ratio
𝑛
𝑐 𝑐 𝑐
𝜁= = = . (2.69)
𝑐𝑐𝑟 2√𝑘𝑚 2𝑚𝜔𝑛

2.2.2.2. Identification based on the diagrams of receptance

Peak amplitude method

The resonance curve from Fig.2.35 illustrates the variation with frequency
of the modulus of the complex receptance (see Table 2.2). Amplitude resonance
𝑘
occurs at the frequency 𝜔𝑟𝑑 = 𝜔𝑛 √1 − 2𝜁 2, where 𝜔𝑛 = √ is the undamped
𝑚
1 1
natural frequency, and point 𝑅 has the ordinate 𝛼𝑚𝑎𝑥 = . Phase
𝑘 2𝜁√1−𝜁 2
1 1
resonance (𝜑𝑑 = −90°) occurs at 𝜔𝑛 and point 𝑀 has the ordinate 𝛼𝑀 = . In
𝑘 2𝜁
most practical applications, for lightly damped systems, the two points cannot be
distinguished from one another and 𝛼𝑚𝑎𝑥 ≅ 𝛼𝑀 . For 𝜁 < 0.4 the error is less than
10% [104].
1
Crossing the receptance curve with the line 𝛼 = 𝛼𝑚𝑎𝑥 we obtain the half
√2
power points 𝐵′ and 𝐶′, of frequencies
2
𝜔′1,2 = 𝜔𝑛2 (1 − 2𝜁 2 ∓ 2𝜁√1 − 𝜁 2 ) . (2.70)
For light damping
2
𝜔′1,2 ≅ 𝜔𝑛2 (1 ∓ 2𝜁)
and the damping ratio is given by
2 2 2 2
𝜔ˈ2 −𝜔ˈ1
2𝜁 ≅ 2 2 = 𝜔ˈ2𝜔
2 −𝜔ˈ1
2 = 𝜔ˈ2𝜔−𝜔ˈ1 𝜔ˈ2𝜔
2 +𝜔ˈ1
.
𝜔ˈ2 +𝜔ˈ1 𝑛 𝑛 𝑛

46
Table 2.2
Frequency
response Symbol Complex function Modulus Phase angle
function

𝑥 𝑥̃ 1 1 𝑥̂ 1⁄ 𝑘 −2ζΩ
Receptance 𝛼̅ = = = 𝑡𝑎𝑛 𝜑d =
𝑓 𝑓̂ 2
𝑘 1 − 𝛺 + 𝑖2𝜁𝛺 𝑓̂ √(1 − 𝛺2 )2 + 4𝜁 𝛺 2 2 1 − Ω2

𝑥̃̇ 1 1 𝑥̂̇ 1 𝛺
𝑥̇
= = 1 − 𝛺2
Mobility ̅ =
ℳ 𝑓̂ 𝑡𝑎𝑛 𝜑𝑣 =
𝑓 𝑓̂ √𝑘𝑚 2𝜁 + 𝑖 (𝛺 − 1 ) √𝑘𝑚
√(1 − 𝛺2 )2 + 4𝜁2 𝛺2 2𝜁𝛺
𝛺
𝑥̂̈ 1 𝛺2
𝑥̈ 𝑥̃̈ 1 𝛺2 =
Inertance 𝜂̅ = =− 𝑓̂ 𝑚 𝜑𝑎 = 𝜑𝑑 + 𝜋
𝑓 𝑓̂ 𝑚 1 − 𝛺2 + 𝑖2𝜁𝛺 √(1 − 𝛺2 )2 + 4𝜁2 𝛺2

Dynamic 𝑓 𝑓̃ 𝑓̂ 2
2𝜁𝛺
stiffness 𝛽̅ = = 𝑘(1 − 𝛺2 + 𝑖 2𝜁𝛺) = 𝑘√(1 − 𝛺2 )2 + 4𝜁2 𝛺 𝑡𝑎𝑛 𝜓𝑑 =
𝑥 𝑥̂ 𝑥̂ 1 − 𝛺2

Mechanical 𝑓 𝑓̃ 2𝜁𝛺 + 𝑖(𝛺2 − 1) 𝑓̂ √(1 − 𝛺2 )2 + 4𝜁2 𝛺2 𝛺2 − 1


impedance 𝑍̅ = = √𝑘𝑚 = √𝑘𝑚 𝑡𝑎𝑛 𝜓𝑣 =
𝑥̇ 𝑥̂̇ 𝛺 𝑥̂̇ 𝛺 2𝜁𝛺

𝑓 𝑓̃ 1 − 𝛺2 + 𝑖2𝜁𝛺 √(1 − 𝛺2 )2 + 4𝜁2 𝛺2


Apparent mass µ̅ = = −𝑚 𝑓̂ 𝜓𝑎 = 𝜓𝑑 + 𝜋
𝑥̈ 𝑥̂̈ 𝛺2 =𝑚
𝑥̂̈ 𝛺 2

Force 𝑓𝑇 𝑓̃ 𝑇 1 + 𝑖2𝜁𝛺 𝑓̂ 𝑇 1 + 4𝜁2 𝛺


2 −2𝜁𝛺3
𝒯̅ = = =√ 𝑡𝑎𝑛 𝜑𝑇 = 2
transmissibility 𝑓 𝑓̂ 1 − 𝛺2 + 𝑖2𝜁𝛺 𝑓̂ (1 − 𝛺2 )2 + 4𝜁2 𝛺2 1 − 𝛺2 + 4𝜁2 𝛺

47
Because
𝜔ˈ2 +𝜔ˈ1
≅ 1,
2𝜔𝑛
it is customary to use the following approximate relationship
𝜔ˈ2 −𝜔ˈ1
𝜁≅ 2𝜔𝑛
. (2.71)

It is possible to avoid these approximations in the calculation of damping.


1 1
At 𝜔 = 0 the static receptance is 𝛼𝑠 = 𝑘 . The horizontal line 𝛼 = 𝑘 intersects the
curve at the point 𝑆 of frequency 𝜔𝑠 = √2𝜔𝑟𝑑 (Fig.2.36).

The horizontal line 𝛼 = 𝜀𝛼𝑠 intersects the curve at the points 𝑃 and 𝑄 of
frequencies 𝜔𝑝 and 𝜔𝑞 , respectively. The damping ratio 𝜁 is given by the following
relation [72]
2 +𝜔2
𝜔𝑝
1 𝑞 1
2
𝜁 = − √1 − , (2.72)
2 4𝜔𝑝 𝜔𝑞 𝜀2

taking the average calculated for different values of 𝜀.


Good results are obtained when the resonance curve is plotted in
logarithmic scales. In this case, the parameters 𝑘 and 𝑚 are determined using the
asymptotes to the curve [106], plotted on special annotated logarithmic paper.
A variant of equation (2.72) is the following [79]
1 𝜔2𝑝 +𝜔2𝑞
𝜁2 = 2 − , (2.73)
4𝜔2𝑛
but the frequency 𝜔𝑛 can be determined only approximately.
The stiffness is given by
1 1
𝑘 ≅ 2𝜁 𝛼 .
𝑚𝑎𝑥

48
Phase angle method

The variation of the phase angle with frequency is shown in Fig.2.37.


Phase resonance (𝜑𝑑 = −90°) occurs at the point 𝑀 of frequency 𝜔𝑛 . The points
𝐵 and 𝐶, defined by frequencies 𝜔1,2 = 𝜔𝑛 (√𝜁 2 + 1 ∓ 𝜁) correspond to phase
angles of −45° and −135°, respectively.
The damping ratio 𝜁 can be evaluated using the exact expression
𝜔2 −𝜔1
𝜁= . (2.74)
2𝜔𝑛

𝑑𝜑𝑑
The slope of the line tangent at the point 𝑀 is 𝑡𝑎𝑛𝜃 = − ( ) . The
𝑑𝜔 𝜔=𝜔𝑛
damping ratio is
1
𝜁= 𝑑𝜑
(2.75)
𝜔𝑛 (− 𝑑𝜔𝑑)
𝜔=𝜔𝑛
which is similar to (2.34).

In-phase response method

The graph shown in Fig.2.38 can be obtained by plotting the variation with
frequency of the displacement in phase with the excitation force, during a frequency
sweep with 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡.
The curve intersects the horizontal axis at the point 𝑀, of frequency 𝜔𝑛 , at
phase resonance. Points 𝐵1 and 𝐶1 , where 𝛼𝑅 has extreme values, correspond to
frequencies 𝜔′, 𝜔" = 𝜔𝑛 √1 ∓ 2𝜁 .
The damping ratio is given by the following exact expressions
2 2 2 2
1 𝜔" −𝜔ˈ
𝜁=2 2 2 = 𝜔"4𝜔−𝜔ˈ
2 . (2.76)
𝜔" +𝜔ˈ 𝑛

49
The stiffness is estimated based on the slope of the tangent in 𝑀
1
𝑘= 2 𝑑𝛼 . (2.77)
2𝜁 𝜔𝑛 (− 𝑑𝜔𝑅)
𝜔=𝜔𝑛

As the vertical distance between points 𝐵1 and 𝐶1 is 𝛼𝑚𝑎𝑥 , the following


formula can be used
1
𝑘= . (2.78)
2𝜁√1−𝜁2 𝛼𝑚𝑎𝑥

Out-of-phase response method

The diagram shown in Fig.2.39 can be obtained by plotting the variation


with frequency of the displacement in quadrature with the excitation force, during
a frequency sweep with 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡.
The peak value of the imaginary component 𝛼𝐼 occurs at a frequency lower
than 𝜔𝑛 , but this can be neglected in practical calculations.
1
At the frequency 𝜔𝑛 , the ordinate is (𝛼𝐼 )𝜔=𝜔𝑛 = − . The damping
2𝜁𝑘
ratio can be estimated using the diagrams from figures 2.38 and 2.39
|𝛼𝐼 |𝜔=𝜔𝑛
𝜁= 𝑑𝛼𝑅 . (2.79)
𝜔𝑛 (− )
𝑑𝜔 𝜔=𝜔𝑛

The tangent line 𝑂𝑇 locates the point 𝑇 of frequency 𝜔𝑟𝑑 . The tangent line
𝑂𝑆 in origin intersects the curve at the point 𝑆, of frequency 𝜔𝑠 = √2𝜔𝑟𝑑 and
2𝜁 2𝜁 𝜔𝑠
ordinate 𝛼𝐼𝑠 = − √2(1 − 2𝜁 2 ) = − . The line (∆) intersects the curve
𝑘 𝑘 𝜔𝑛
at points 𝑃 and 𝑄 of frequencies 𝜔𝑝 and 𝜔𝑞 respectively, which can be used to

50
calculate 𝜁 from eq. (2.73) in which 𝜔𝑛 is determined approximately as the
frequency of peak response amplitude.

Polar plot method

Eliminating the frequency between the real and imaginary components


𝛼𝑅 = 𝛼 𝑐𝑜𝑠𝜑𝑑 and 𝛼𝐼 = 𝛼 𝑠𝑖𝑛𝜑𝑑 of the complex receptance, we obtain a curve
(Fig.2.40) of equation
𝛼𝑅 1
(𝛼𝑅2 + 𝛼𝐼2 )2 − (𝛼𝑅2 + 𝛼𝐼2 ) − 𝛼𝐼2 = 0 . (2.80)
𝑘 4𝜁 2 𝑘 2

For viscous damping, the vector diagram of the receptance is not a circle,
as for hysteretic damping. However, for light damping it can be approximated by a
circle.
The curve intersects the imaginary axis at the point 𝑀, of frequency 𝜔𝑛 .
The lines 𝑂𝐵 and 𝑂𝐶, inclined ±45° with respect to the negative imaginary
semiaxis, locate on the curve the frequencies 𝜔1,2 = 𝜔𝑛 (√𝜁 2 + 1 ∓ 𝜁) which can
be used to calculate the damping ratio using the exact expression (2.74).
The stiffness is obtained from
1
𝑘= ̅̅̅̅̅̅ (2.81)
2𝜁𝑂𝑀
𝑘
̅̅̅̅̅ = |𝛼𝐼 |𝜔=𝜔 and the mass from 𝑚 = 2 .
where 𝑂𝑀 𝑛 𝜔 𝑛

2.2.2.3. Identification based on the diagrams of mobility

Peak velocity method

The graph from Fig.2.41 illustrates the variation with frequency of the
modulus of the complex mobility (see Table 2.2). Resonance of velocity amplitude
𝑘 1 1 1
occurs at the point 𝑅, of frequency 𝜔𝑛 = √ and ordinate ℳ𝑚𝑎𝑥 = = .
𝑚 √𝑘𝑚 2𝜁 𝑐
Points 𝐵𝜀 and 𝐶𝜀 , located by crossing the curve with the horizontal line
1
ℳ = 𝜀 ℳ𝑚𝑎𝑥 have frequencies
𝜔1𝜀 ,2𝜀 = 𝜔𝑛 (√𝜁 2 (𝜀 2 − 1) + 1 ∓ 𝜁√𝜀 2 − 1) .

The damping is calculated using the exact expression


𝜔2𝜀 −𝜔1𝜀
𝜁= . (2.82)
2𝜔𝑛 √𝜀 2 −1
Because 𝜔1𝜀 𝜔2𝜀 = 𝜔𝑛2 , denoting ∆𝜔𝜀 = 𝜔2𝜀 − 𝜔1𝜀 , equation (2.82) can
be written

51
∆𝜔𝜀
𝜁= . (2.82, a)
2√𝜔1𝜀 𝜔2𝜀 (𝜀 2 −1)
Repeating the construction for several values 𝜀 , we can plot the ratio
∆𝜔𝜀
against √𝜀 2 − 1, then draw a straight line through the data points. The
2√𝜔1𝜀 𝜔2𝜀
damping ratio is equal to the slope of this line.

1
For 𝜀 = √2, the line ℳ = ℳ𝑚𝑎𝑥 intersects the curve in points 𝐵 and 𝐶
√2
of frequencies
𝜔1,2 = 𝜔𝑛 (√𝜁 2 + 1 ∓ 𝜁). (2.83)
The damping ratio can be calculated using the exact expression [134]
𝜔 −𝜔 ∆𝜔
2
𝜁 = 2𝜔 1
= 2𝜔 . (2.84)
𝑛 𝑛

Because 𝜔1 𝜔2 = 𝜔𝑛2 , the expression above can be written


𝜔2 −𝜔1
𝜁=
2√𝜔1 𝜔2
= 12 (√𝜔
𝜔
2
− √𝜔
𝜔
1
). (2.85)
1 2

The mass can be calculated as


1 𝑐
𝑚=
2𝜁𝜔𝑛 ℳ𝑚𝑎𝑥
=𝜔 . (2.86)
2 −𝜔1
This method can lead to errors when the resonance curves are too flat (high
damping) or too sharp (light damping), requiring selection of adequate scales for
the graphical display.
The curve from Fig.2.41 becomes symmetrical when sketched using
logarithmic scales (Fig.2.43, a). In this case it is helpful to use annotated diagrams
with special grids [33] including constant stiffness lines of slope 6 𝑑𝐵⁄octave and
constant mass lines of slope −6 dB⁄𝑜𝑐𝑡𝑎𝑣𝑒 . The mass 𝑚 is determined from the
asymptotic behavior at high frequencies, and the stiffness 𝑘 – from the asymptote
at low frequencies.

52
The damping ratio is estimated from equation (2.84) where ∆𝜔 = 𝜔2 − 𝜔1
is the bandwidth at 3𝑑𝐵 below the peak. The same value is obtained calculating the
1
“quality factor” 𝑄 = 2𝜁 as the vertical distance between the crossing point of the
asymptotes at low and high frequencies and the point 𝑅 of maximum amplitude
(Fig.2.43, b).

Phase angle method

On the graph shown in Fig.2.44, the points 𝐵, 𝑀 and 𝐶 locate the


frequencies 𝜔1 , 𝜔𝑛 and 𝜔2 from which the damping ratio is obtained using eq.
(2.84).
We can establish the expression
1
𝜁= 𝑑𝜑 (2.87)
𝜔𝑛 (− 𝑣 )
𝑑𝜔 𝜔=𝜔𝑛

which is helpful when only a small part of the phase diagram is available in the
neighborhood of 𝑀.
It was found [86] that expression (2.87) yields better estimates for the
damping ratio than the peak amplitude method.

In-phase response method

Figure 2.45 shows the variation with frequency of the real component of
1 1
mobility ℳ𝑅 = ℳ 𝑐𝑜𝑠𝜑𝑣 . The peak value ℳ𝑅𝑚𝑎𝑥 = occurs at the natural
√𝑘𝑚 2𝜁
1
frequency 𝜔𝑛 . The line ℳ𝑅 = ℳ𝑅𝑚𝑎𝑥 intersects the curve at the points 𝐵𝑒 and
𝑒
𝐶𝑒 , of frequencies 𝜔1𝑒 and 𝜔2𝑒 , respectively, which can be used to obtain the
following exact expression of the damping ratio

53
𝜔2𝑒 −𝜔1𝑒
𝜁= . (2.88)
2𝜔𝑛 √𝑒−1

1
For 𝜀 = √2, the line ℳ𝑅 = 2 ℳ𝑅𝑚𝑎𝑥 intersects the curve at the points 𝐵
and 𝐶 of frequencies 𝜔1 and 𝜔2 , used in formula (2.84) for the evaluation of
damping.

Out-of-phase response method

Figure 2.46 shows the variation with frequency of the imaginary


component of mobility ℳ𝐼 = ℳ 𝑠𝑖𝑛𝜑𝑣 . The curve crosses the frequency axis at
point 𝑀, of abscissa 𝜔𝑛 , which is an inflection point.
1
The extreme values ℳ𝐼𝑚𝑎𝑥,𝑚𝑖𝑛 = ± occur at the points 𝐵 and 𝐶,
4𝜁√𝑘𝑚
of abscissae 𝜔1,2 = 𝜔𝑛 (√𝜁 2 + 1 ∓ 𝜁). The damping ratio is given by formula
(2.84).
1 1 1
The distance between the tangent lines in 𝐵 and 𝐶 is ℳ𝑚𝑎𝑥 = =
√𝑘𝑚 2𝜁 𝑐
so that the mass can be estimated from equation (2.85).
𝑑ℳ𝐼 1
The line tangent in 𝑀 has a slope (− ) = . This yields
𝑑𝜔 𝜔=𝜔𝑛 2𝜁 2 𝑘
2 (ℳ𝐼 )𝜔=𝜔1
𝜁= 𝑑ℳ𝐼 . (2.89)
𝜔𝑛 (− )
𝑑𝜔 𝜔=𝜔𝑛

1
The line 𝐵𝐶 has a slope (− 4𝜁 2𝑘) . It is useful to use a different
construction. Locating the points 𝐵ˈ and 𝐶ˈ at twice the ordinates of points 𝐵 and 𝐶,
the line 𝐵ˈ𝐶ˈ is tangent in 𝑀 at the curve [127].

54
Polar plot method

The polar plot of the complex mobility (Fig.2.47) is a circle of equation


1 2 1 2
(ℳ𝑅 − 4𝜁√𝑘𝑚) + ℳ𝐼2 = (4𝜁√𝑘𝑚) . (2.90)

The intercept of the circle with the real axis locates the frequency 𝜔𝑛 . The
1
diameter ̅̅̅̅̅
𝑂𝑀 = ℳ𝑚𝑎𝑥 = 𝑐 . The diameter 𝐵𝐶, perpendicular to 𝑂𝑀, intersects the
circle at frequencies 𝜔1 and 𝜔2 (2.83). The damping ratio 𝜁 can be obtained from
(2.84).
Denoting 𝑑𝑠 the arc of circle which subtends an angle 𝜑𝑣 , we can write
1
𝑑𝑠 = 𝑑𝜑𝑣 . The derivative
𝑐
𝑑𝑠 1 𝑑𝜑𝑣 1 𝑑 𝛺2 −1 1 𝛺2 +1
= −𝑐 = 𝑐 𝑑𝛺 (𝑡𝑎𝑛−1 )= (2.91)
𝑑𝛺 𝑑𝛺 2𝜁𝛺 √𝑘𝑚 (1−𝛺2 )2 +4𝜁 2 𝛺2

has a maximum value


𝑑𝑠 1 1
(𝑑𝛺) = (2.92)
𝑚𝑎𝑥 4√𝑘𝑚 √1−𝜁 2 −(1−𝜁 2 )
at the dimensionless frequency

𝛺𝑜 = √2√1 − 𝜁 2 − 1 (2.93)

which is lower than the resonance frequency 𝛺 = 1 .


For usual values of the damping ratio 𝜁, 𝛺𝑜 ≅ 1 and
𝑑𝑠 1 1 1
(𝑑𝛺) ≅ =
𝑚𝑎𝑥 2𝜁 2 √𝑘𝑚 𝜁𝑐
so that the Kennedy-Pancu method can be used to locate the resonance [131].

55
If only a small part of the polar plot near resonance is available, expression
(2.84) cannot be used. In this case, first, the point 𝑀 defined by 𝜔𝑛 is marked where
𝑑𝑠
is a maximum. Then, the “best circle” is constructed through the data points
𝑑𝜔
in the neighborhood of 𝑀 [129]. The diameter 𝑂𝑀 locates the displaced origin 𝑂.
Two data points 𝑃 and 𝑄, of known frequencies 𝜔𝑝 and 𝜔𝑞 , are selected close to 𝑀
and the subtended angles 𝜃𝑝 and 𝜃𝑞 are measured (Fig.2.48). Expression (2.87)
yields
𝑑𝜑 𝜃𝑞 −𝜃𝑝 1
( 𝑑𝜔𝑣) ≅
𝜔𝑞 −𝜔𝑝
=−
𝜁𝜔𝑛
𝜔=𝜔𝑛
wherefrom [74]
𝜔𝑞 −𝜔𝑝 1
𝜁= . (2.94)
𝜔𝑛 𝜃𝑝 +|𝜃𝑞 |
When |𝜃𝑞 | = 𝜃𝑝 = 𝜃
𝜔𝑞 −𝜔𝑝
𝜁= cot 𝜃 . (2.95)
2𝜔𝑛

2.2.2.4. Identification based on the diagrams of inertance

Peak acceleration method

Figure 2.49 shows the variation with frequency of the modulus of the
complex inertance (see Table 2.2). Acceleration resonance occurs at the frequency
𝜔𝑛 1 1
𝜔𝑟𝑎 =
√1−2𝜁 2
, where the point 𝑅 has the ordinate 𝜂𝑚𝑎𝑥 =
𝑚 2𝜁√1−𝜁 2
.

𝜔𝑟𝑎
At the frequency 𝜔𝑠 = is located the point 𝑆 where the asymptote
√2
1 𝜀
𝜂= 𝑚
intersects the curve. The line 𝜂 = 𝑚 locates the points 𝑃 and 𝑄, of
frequencies 𝜔𝑝 and 𝜔𝑞 , from which the damping ratio is given by (2.72)

56
1 𝜔2𝑝 +𝜔2𝑞 1
𝜁 2 = 2 − 4𝜔 𝜔 √1 − 2 .
𝑝 𝑞 𝜀
The tangent line 𝑂𝑇 locates the point 𝑇 of frequency 𝜔𝑛 and ordinate
1 1
𝜂𝑇 = 𝑚 2𝜁
. The line (∆) intersects the curve in 𝑈 and 𝑉, of frequencies 𝜔𝑢 and 𝜔𝑣
so that 𝜔𝑛 = √𝜔𝑢 𝜔𝑣 and [72]
1 (𝜔2 2 2
𝑝 +𝜔𝑞 )𝜔𝑛
𝜁2 = 2 − . (2.96)
4𝜔𝑝 𝜔2𝑞
2

The mass can be calculated from the ordinate of 𝑆 (Fig.2.49) and 𝑘 = 𝑚𝜔𝑛2 .
Better accuracy is obtained when the resonance curve is plotted using
logarithmic scales.

Polar plot method

Eliminating the frequency between the expressions of the real and


imaginary components of the complex inertance, 𝜂𝑅 = 𝜂 𝑐𝑜𝑠𝜑𝑎 and 𝜂𝐼 = 𝜂 𝑠𝑖𝑛𝜑𝑎 ,
we obtain a curve of equation
𝜂𝑅 1
(𝜂𝑅2 + 𝜂𝐼2 )2 − (𝜂𝑅2 + 𝜂𝐼2 ) − 𝜂𝐼2 = 0 . (2.97)
𝑚 4𝜁 2 𝑚2

This shows that for viscous damping the vector diagram of the inertance is
not a circle.
The curve intersects the imaginary axis at the point 𝑀 of frequency 𝜔𝑛
(phase resonance). The chords 𝑂𝐵 and 𝑂𝐶, subtending angles of ±45° with the
positive imaginary semiaxis, locate on the diagram the frequencies 𝜔1,2 =
𝜔𝑛 (√𝜁 2 + 1 ∓ 𝜁) which can be used to calculate the damping ratio using the exact
formula (2.74).
1
The mass is obtained from 𝑚 = 2𝜁𝑂𝑀̅̅̅̅̅
, where ̅̅̅̅̅
𝑂𝑀 = (𝜂𝐼 )𝜔=𝜔𝑛 .

57
2.2.2.5. Identification based on the diagrams of dynamic stiffness

Minimum force method

Figure 2.52 depicts the variation with frequency of the dynamic stiffness
modulus (see Table 2.2). At resonance 𝜔𝑟𝑑 = 𝜔𝑛 √1 − 2𝜁 2 and 𝛽𝑚𝑖𝑛 =
𝑘 2𝜁√1 − 𝜁 2 . Crossing the curve with the line 𝛽 = √2𝛽𝑚𝑖𝑛 we obtain the points
𝐵ˈ and 𝐶ˈ, of frequencies 𝜔ˈ1 and 𝜔ˈ2 (2.70). For light damping, expression (2.70)
simplifies to (2.71).

Polar plot method

The polar diagram of the complex dynamic stiffness is a parabola


(Fig.2.53) of equation
1 1
𝛽𝐼2 = 4𝜁 2 (1 − 𝑘 𝛽𝑅 ) (2.98)
𝑘2
sometimes referred to as Runge’s parabola [108].
The curve intersects the imaginary axis at the point 𝑀 of frequency 𝜔𝑛 .
The lines 𝑂𝐵 and 𝑂𝐶, oriented ±45° with respect to the positive imaginary
semiaxis, locate the frequencies 𝜔1 and 𝜔2 (2.83). The damping ratio is given by
the exact expression
𝜔2 −𝜔1
𝜁= .
2𝜔𝑛

Estimates for the mass and stiffness parameters are obtained from
1
𝑚= and 𝑘 = 𝑚𝜔𝑛2 .
2𝜁(𝛽𝐼 )𝜔=𝜔𝑛

2.2.2.6. Identification based on the diagrams of mechanical impedance

Minimum force method

Figure 2.54 is a magnitude-frequency graph of the mechanical impedance


of the system shown in Fig.2.34.

58
The damping ratio 𝜁 is given by the exact formula (2.84), in which 𝜔1 and
𝜔2 are the frequencies of the crossing points with the line 𝑍 = √2𝑍𝑚𝑖𝑛 . The mass
is given by
𝑐 𝑍𝑚𝑖𝑛
𝑚=𝜔 = 2𝜁𝜔 . (2.99)
2 −𝜔1 𝑛

If the graph of the mechanical impedance is sketched in logarithmic scales


(Fig.2.55), the mass and stiffness parameters are determined from the asymptotic
behavior away from resonance, and the damping ratio 𝜁 – from the “quality factor”
1
𝑄 = 2𝜁 measured as the vertical distance between the crossing point 𝑆 of the
asymptotes to the curve and the point 𝑅 of minimum magnitude.
The points of intersection of the curve with the horizontal line drawn at
3 𝑑𝐵 above the minimum value 𝑍𝑚𝑖𝑛 have frequencies 𝜔1 and 𝜔2 , so the damping
ratio can be calculated from (2.84).

Polar plot method

Because the real component 𝑍𝑅 = 2𝜁√𝑘𝑚 is independent of frequency, the


vector diagram of the mechanical impedance is a straight line parallel to the
imaginary axis (Fig.2.56). The line intercepts the real axis at the point 𝑀, of
frequency 𝜔𝑛 . The lines 𝑂𝐵 and 𝑂𝐶, oriented at angles ±45° to the positive real
semiaxis, locate the frequencies 𝜔1 and 𝜔2 . The system parameters are estimated
from (2.84) and (2.99).

2.2.2.7. Identification based on the diagrams of apparent mass

Minimum force method

Figure 2.57 is a graphical display of the variation with frequency of the


modulus of the apparent mass (see Table 2.2). Usually the graph is plotted keeping
constant the acceleration amplitude and measuring the variation of the amplitude
of the excitation force during a frequency sweep.

59
The minimum value µ𝑚𝑖𝑛 = 𝑚2𝜁√1 − 𝜁 2 occurs at the frequency
𝜔𝑛
𝜔𝑟𝑎 = 2
√1−2𝜁
. The horizontal asymptote µ = 𝑚 intersects the curve at the point
𝜔𝑟𝑎 𝑚
𝑆, of frequency 𝜔𝑠 = . The ratio 𝜀 = µ is measured as the distance between
√2 𝑚𝑖𝑛
the line tangent to the curve in 𝑅 and the horizontal asymptote. The damping ratio
is given by the exact formula
√𝜀+1−√𝜀−1
𝜁= . (2.100)
2 √𝜀
The mass and stiffness parameters are determined from the asymptotic
response away the resonance.

Polar plot method

The polar plot of the complex apparent mass is a parabola (Fig.2.58) of


equation
µ2𝐼 µ𝑅
= 1− . (2.101)
4𝜁 2 𝑚2 𝑚
The curve intercepts the imaginary axis at the point 𝑀, of frequency 𝜔𝑛 ,
the phase resonance. The lines 𝑂𝐵 and 𝑂𝐶, drawn at angles ±45° to the negative
imaginary semiaxis, locate the frequencies 𝜔1 and 𝜔2 . The damping ratio is
calculated using (2.84) and the mass – from
1
𝑚 = 2𝜁 (−µ𝐼 )𝜔=𝜔𝑛 . (2.102)

2.2.2.8. Identification based on the diagram of force transmissibility

The polar plot of the force transmissibility is sketched in Fig.2.59. The


maximum value

60
4𝜁 2
𝒯𝑚𝑎𝑥 =
√16𝜁 4 −8𝜁 2 −2+2√1+8𝜁 2
occurs at the frequency
𝜔𝑛
𝜔𝑟 = √√1 + 8𝜁 2 − 1 .
2𝜁

The frequency 𝜔𝑛 can be obtained crossing the diagram with the line
𝒯𝑅 = 1. The damping ratio has the expression
1
𝜁= . (2.103)
2(−𝒯𝐼 )𝜔=𝜔𝑛

The curve intersects the imaginary axis at the point 𝐿, of frequency 𝜔ℓ =


𝜔𝑛
. This yields another expression for the damping ratio
√1−4𝜁 2

1 𝜔 2
𝜁 = 2 √1 − ( 𝜔𝑛) . (2.104)

2.3. EXCITATION USING HARMONIC FORCES IN QUADRATURE

The method of forces in quadrature, first used by G. de Vries [131], is a


variant of the method of displaced frequencies (see § 2.6.1). The method has been
developed by the present author, especially for the identification of systems with
hysteretic damping [90, 93]. Although its application requires a specialized
equipment, its use leads to more accurate results than the classical Kennedy-Pancu
method, especially for very lightly damped systems.
The method is based on the modification of the frequency response
diagram when, apart from the active excitation force 𝑓̂𝑐𝑜𝑠𝜔𝑡, the system is excited
by a reactive force 𝑓̂𝑠𝑖𝑛𝜔𝑡, phase shifted 90° (the quadrature component). In
complex algebra, the force 𝑓(𝑡) = 𝑓̂𝑒 𝑖𝜔𝑡 is replaced by
𝑓(𝑡) = (𝑓̂ + 𝑖𝑓ˈ
̂ )𝑒 𝑖𝜔𝑡 , (2.105)
̂ = 𝜆𝑓̂,
or, denoting 𝑓ˈ
𝑓(𝑡) = 𝑓̂(1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 . (2.106)
Analytically, the expressions of the complex frequency response functions
having a force at the denominator (receptance, mobility, inertance) are multiplied
by the complex quantity (1 + 𝑖𝜆), what is tantamount to a magnification by
√1 + 𝜆2 and a counterclockwise rotation (for 𝜆 > 0) with an angle 𝑡𝑎𝑛−1 𝜆.

61
2.3.1. Identification of Systems with Hysteretic Damping

2.3.1.1 Identification based on the diagram of receptance

Method of the two circles

If the system from Fig.2.7 is excited by a force (2.106), the equation of


motion can be written

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘𝑥 = 𝑓̂ (1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 . (2.107)
For a steady-state harmonic displacement
𝑥 = 𝑥̃𝑒 𝑖𝜔𝑡 (2.108)
the complex receptance is
𝑥̃
𝛼̅ = = 1𝑘 1+𝑖𝜆
(2.109)
𝑓̂ 1−𝛺2 +𝑖𝑔
where
ℎ 𝜔 𝑘
𝑔=𝑘, 𝛺=𝜔 , 𝜔𝑛 = √𝑚 . (2.110)
𝑛

Expression (2.109) can be written under the form


1−𝛺2 +𝜆𝑔 𝜆(1−𝛺2 )−𝑔
𝛼̅ = 𝛼𝑅 + 𝑖𝛼𝐼 = 𝑘[(1−𝛺2 )2 +𝑔2 ] + 𝑖 . (2.111)
𝑘[(1−𝛺2 )2 +𝑔2 ]

Eliminating 𝛺 between 𝛼𝑅 and 𝛼𝐼 , we obtain the vector diagram of the


complex receptance
𝜆 2 1 2 1+𝜆2
(𝛼𝑅 − 2ℎ) + (𝛼𝐼 + 2ℎ) = . (2.112)
4ℎ2
This is sketched in Fig.2.60 by the circle drawn with solid line (𝜆 ≠ 0).

62
On the same graph is drawn with dashed line the circle of equation (2.37)
and parameter 𝜆 = 0. The two circles cross each other on the imaginary axis at the
point 𝑀 of frequency 𝜔𝑛 on the circle 𝜆 = 0 and 𝜔ˈ = 𝜔𝑛 √1 + 𝜆𝑔 on the circle
𝜆 ≠ 0. This can be checked noting that the diameter 𝑂𝑀ˈ is oriented at an angle
𝑡𝑎𝑛−1 𝜆 to 𝑂𝑀. We can obtain an expression for the damping factor
1 𝜔ˈ 2
𝑔 = 𝜆 [(𝜔 ) − 1]. (2.113)
𝑛

̅̅̅̅̅.
The stiffness is given by equation (2.32) where 𝛼𝑟𝑒𝑠 = 𝑂𝑀

Method of the three circles

If three circles of parameters 𝜆1 , 𝜆2 and 𝜆 = 0 are drawn on the same graph


(Fig.2.61), they all cross the negative imaginary semiaxis at the resonance point 𝑀.
The damping factor is
1 𝜔ˈ2 −𝜔"2 𝜔ˈ2 −𝜔"2
𝑔= 2 = (2.114)
𝜆1 −𝜆2 𝜔𝑛 𝜆1 𝜔"2 −𝜆2 𝜔ˈ2
where 𝜔𝑛 , 𝜔ˈ and 𝜔" are the frequencies of the point 𝑀 on the circles 𝜆 = 0, 𝜆1 and
𝜆2 , respectively.
If 𝜆1 = +1 and 𝜆2 = −1, then ˈ = 𝜔2 = 𝜔𝑛 √1 + 𝑔 , 𝜔" = 𝜔1 =
𝜔𝑛 √1 − 𝑔 and equations (2.114) simplify to (2.30).
The measurement can be carried out at constant displacement amplitude,
which is useful for the analysis of slightly nonlinear systems. Resonance location
on the circle 𝜆 = 0 is independent of the coordinate system. This is helpful at
systems with several degrees of freedom and relatively close natural frequencies.

Method of constant frequency lines

Eliminating 𝜆 between the expressions of


𝛼𝑅 and 𝛼𝐼 (2.111) we obtain the equation of a
pencil of constant frequency lines (Fig.2.62)
1
passing through the point 𝑀 (0, − )

1−𝛺2 1
𝛼𝐼 = 𝛼𝑅 − ℎ . (2.115)
𝑔

The location of resonance on the circle


𝜆 = 0 can be done crossing the circle with two
straight lines (2.115) of frequencies 𝜔1 = 𝑐𝑜𝑛𝑠𝑡.
and 𝜔2 = 𝑐𝑜𝑛𝑠𝑡., respectively.

63
Determining the value of λ corresponding to the crossing point of the line
𝜔1 = 𝑐𝑜𝑛𝑠𝑡. with the circle, the damping factor is given by
1 𝜔2
𝑔 = 𝜆 (𝜔21 − 1). (2.116)
𝑛

The stiffness and mass parameters are


1 𝑘
𝑘 = 𝑔𝛼 and 𝑚 = 2 .
𝑚𝑎𝑥 𝜔𝑛

2.3.1.2 Identification based on the diagrams of dynamic stiffness

Method of the lines 𝝀 = 𝒄𝒐𝒏𝒔𝒕.

If the measurements are carried out at constant amplitude displacement and


variable amplitude of the harmonic excitation force, the following function is of
interest
1 1−𝛺2 +𝑖𝑔
𝛽̅ = 𝛼̅ = 𝑘 (2.117)
1+𝑖𝜆
or
1−𝛺2 +𝑔𝜆 𝑔−𝜆 (1−𝛺2 )
𝛽̅ = 𝛽𝑅 + 𝑖𝛽𝐼 = 𝑘 +𝑖𝑘 (2.118)
1+𝜆2 1+𝜆2
Eliminating 𝛺 between the components 𝛽𝑅 and 𝛽𝐼 , for 𝜆 = 𝑐𝑜𝑛𝑠𝑡. the tip
of the vector (2.117) moves along lines of equation (Fig.2.63)
𝛽𝐼 = −𝜆𝛽𝑅 + ℎ. (2.119)

For 𝜆 = 0 we obtain the graph from Fig.2.28. Any line 𝜆 = 𝑐𝑜𝑛𝑠𝑡.


intersects the line 𝜆 = 0 at the point 𝑀, of frequency 𝜔𝑛 . On a line of parameter 𝜆1
point 𝑀 corresponds to a frequency 𝜔ˈ = 𝜔𝑛 √1 + 𝑔𝜆1 , and on a line of parameter

64
𝜆2 – to a frequency 𝜔" = 𝜔𝑛 √1 + 𝑔𝜆2 . The damping factor is calculated from the
expression (2.114).

Method of constant frequency circles

Due to the inversion (2.117), eliminating 𝜆 between the expressions of 𝛽𝑅


and 𝛽𝐼 in (2.118), for 𝜔 = 𝑐𝑜𝑛𝑠𝑡., the polar diagram of the complex dynamic
stiffness is a circle of equation
2 2
1−𝛺2 ℎ 2 ℎ2 +𝑘 2 (1−𝛺2 )
(𝛽𝑅 − 𝑘 ) + (𝛽𝐼 − 2) = . (2.120)
2 4
All circles 𝜔 = 𝑐𝑜𝑛𝑠𝑡. intersect the circle 𝜔 = 𝜔𝑛 on the imaginary axis
at the resonance point 𝑀 (Fig.2.64). On two circles drawn at constant frequencies
1 𝜔ˈ2
𝜔ˈ and 𝜔", respectively, the point 𝑀 corresponds to 𝜆1 = (1 − 𝜔2 ) and
𝑔 𝑛
1 𝜔"2
𝜆2 = 𝑔 (1 − 2 ), so that 𝑔 can be calculated from expression (2.114).
𝜔𝑛

2.3.2. Identification of Systems with Viscous Damping

This section is concerned with the presentation of three identification


methods based on the analysis of the polar plots of the complex mobility. Other
methods, based on the analysis of the variation with the excitation force of the real
and imaginary components of the velocity, as well as the analysis of the diagram of
the phase angle between force and velocity can be found in [89].

Method of the two circles

The equation of motion of the system shown in Fig.2.34, under the action
of an excitation force (2.106), is
𝑚𝑥̈ + 𝑐𝑥̇ + 𝑘𝑥 = 𝑓̂ (1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 . (2.121)
For steady-state harmonic motion, the complex mobility is
̃̇
̅
ℳ = 𝑓𝑥̂ = √ 1 1+𝑖𝜆
(2.122)
𝑘𝑚 2𝜁+𝑖(𝛺−𝛺1 )
where
𝑐 𝜔 𝑘
𝜁 = 2√𝑘𝑚 , 𝛺=𝜔 , 𝜔𝑛 = √𝑚 . (2.123)
𝑛

Expression (2.122) can be written under the form

65
1 1
2𝜁 + 𝜆 (𝛺 − )
1 1 2𝜁𝜆 − (𝛺 − 𝛺 )
̅ = ℳ𝑅 + 𝑖ℳ𝐼 =
ℳ 𝛺
2+𝑖 2
√𝑘𝑚 4𝜁 2 + (𝛺 − 1 ) √𝑘𝑚 4𝜁 2 + (𝛺 − 1 )
𝛺 𝛺
(2.124)
Eliminating 𝛺 between ℳ𝑅 and ℳ𝐼 , we obtain the equation of a circle
1 2 𝜆 2 1+𝜆2
(ℳ𝑅 − 4𝜁√𝑘𝑚) + (ℳ𝐼 − 4𝜁√𝑘𝑚) = 16𝜁 2 𝑘𝑚 (2.125)

represented graphically in Fig.2.65 with solid line (𝜆 ≠ 0). The circle 𝜆 = 0 is


drawn with dashed line on the same figure.
For 𝜆 > 0, the center 𝑂ˈ of the circle is located in the imaginary positive
halfplane, and for 𝜆 < 0 – in the imaginary negative halfplane. At amplitude
1 1+𝜆2
resonance, the diameter is 𝑂𝑀ˈ = √ and subtends an angle 𝑡𝑎𝑛−1 𝜆 with
2𝜁 𝑘𝑚
the positive real semiaxis.
The two circles cross each other on the real axis at the point 𝑀, which on
the circle 𝜆 = 0 corresponds to the resonance frequency 𝜔𝑛 and on the circle 𝜆 ≠ 0
– to the frequency 𝜔ˈ = 𝜔𝑛 (√𝜁 2 𝜆2 + 1 + 𝜁𝜆).
The damping ratio is given by [131]
1 𝜔ˈ 𝜔
𝜁 = 2𝜆 (𝜔 − 𝜔ˈ𝑛 ). (2.126)
𝑛
The mass is calculated from (2.85).

Method of the three circles

A better location of the resonance frequency 𝜔𝑛 on the circle 𝜆 = 0 is


obtained crossing the circle with two other circles of parameters 𝜆1 and 𝜆2 ,
respectively (Fig.2.66). If 𝜔ˈ and 𝜔" are the frequencies of the point 𝑀 on these
circles, the damping ratio is given by
2 2
𝜔ˈ −𝜔"
𝜁 = 2𝜔 𝜆 𝜔ˈ−𝜆 𝜔" . (2.127)
𝑛( 1 2 )

Usually 𝜆1 > 0 and 𝜆2 < 0.


If 𝜆1 = 𝜆 and 𝜆2 = −𝜆 , equation (2.127) becomes [75]
1 𝜔ˈ−𝜔" 1 𝜔ˈ−𝜔"
𝜁 = 2𝜆 𝜔 = 2𝜆 . (2.128)
𝑛 √ 𝜔ˈ𝜔"
If 𝜆1 = 1 and 𝜆2 = −1, then 𝜔ˈ = 𝜔2 and 𝜔" = 𝜔1 , and equation (2.128)
becomes (2.84).
Generally, for the estimation of damping, only a few points are required on
the three circles, located near the crossing point.

66
Method of the constant frequency lines

Vector diagrams of the complex mobility (2.124) are plotted in Fig.2.67 for
different values of 𝜆 = 𝑓ˈ ̂ ⁄𝑓̂. All circles cross the vector diagram 𝜆 = 0 at the point
𝑀, which on this circle defines the resonance frequency 𝜔𝑛 . Any line ∆ passing
through 𝑀 is a line of constant excitation frequency, crossing the circles at points
𝑃𝑗 (𝑗 = 1, 2, 3 … ) having 𝜔𝑗 = 𝑐𝑜𝑛𝑠𝑡.
The equation of these lines is obtained eliminating the parameter 𝜆 between
the components ℳ𝑅 and ℳ𝐼 from equation (2.124)
2𝜁 1
ℳ𝐼 = 1 (ℳ𝑅 − 2𝜁√𝑘𝑚). (2.129)
𝛺−
𝛺

If 𝛺 is the dimensionless frequency at the point 𝑃 on the circle 𝜆 = 0, the


vector 𝑂𝑃 has a slope
1
−Ω
tan θ = Ω2ζ . (2.130)

The line ∆ of parameter 𝛺 is perpendicular to 𝑂𝑃, hence is inclined an angle


𝜃 with respect to the tangent at 𝑀 to the circle 𝜆 = 0.
The value of 𝜆 which corresponds to the point 𝑀 on a line ∆ of frequency
𝜔1 is
𝜔𝑛 𝜔1 1

𝜔1 𝜔𝑛 𝛺1
−𝛺1
𝜆𝑀 = 2𝜁
= 2𝜁
= tan θ. (2.131)

67
The resonance frequency is determined at the intersection of the line
𝜔1 = 𝑐𝑜𝑛𝑠𝑡. with the circle 𝜆 = 0. The damping ratio can be calculated from the
following equation [75]
1 𝜔 𝜔
𝜁 = ( 𝑛 − 1 ) cot𝜃. (2.132)
2 𝜔 𝜔 1 𝑛

The value of the diameter 𝑂𝑀 of the circle 𝜆 = 0 is used to determine the


other system parameters.

2.4. EXCITATION WITH ROTATING UNBALANCE FORCES

In many practical applications, the perturbing forces are due to rotating out-
of-balance masses. Also, vibrators used for determining the dynamic characteristics
of soils and antivibration mountings generate unbalanced forces proportional to the
square of the angular speed. In the following we consider forces of the form
𝑓(𝑡) = 𝑚𝑜 𝑟𝜔2 𝑐𝑜𝑠 𝜔𝑡 (2.133)
𝑚
produced by two eccentric masses 2𝑜 , symmetrically located with respect to the
median vertical plane, turning in opposite directions with angular speeds 𝜔 along
circles of radius 𝑟 (Fig.2.68).

2.4.1. Identification of Systems with Hysteretic Damping

The equation of motion of the system from Fig.2.68



𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘𝑥 = 𝑚𝑜 𝑟𝜔2 𝑒 𝑖𝜔𝑡 . (2.134)

is obtained substituting 𝑓̂ = 𝑚𝑜 𝑟𝜔2 in equation (2.26).


For a steady-state harmonic displacement
𝑥 = 𝑥̃𝑒 𝑖𝜔𝑡 , (2.135)

68
using notations (2.110), we obtain the complex displacement amplitude
𝛺2
𝑥̃ = 𝑚𝑚𝑜𝑟 . (2.136)
1−𝛺2 +𝑖𝑔
Expression (2.136) is similar to the frequency response function of the
complex inertance (see Table 2.1). Thus we can use in this case the identification
methods discussed in § 2.2.1.4.

Method of forces in quadrature

The method of forces in quadrature can be used adding on each shaft of the
𝑚ˈ
vibrator an eccentric mass 2𝑜 at a radius 𝑟ˈ, shifted ±90° with respect to the radius
𝑚𝑜
𝑟 of the mass (Fig.2.69). The equation of motion of the mass 𝑚 becomes
2

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘𝑥 = 𝑚𝑜 𝑟𝜔2 𝑐𝑜𝑠 𝜔𝑡 − 𝑚ˈ𝑜 𝑟ˈ𝜔2 𝑠𝑖𝑛 𝜔𝑡. (2.137)
Using complex algebra and inserting
𝑚𝑜′ 𝑟ˈ
𝜆= 𝑚𝑜 𝑟
(2.138)
in equation (2.137) we obtain

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘𝑥 = 𝑚𝑜 𝑟𝜔2 (1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 . (2.139)
For a steady-state solution (2.135), the displacement complex amplitude is
2
𝑥̃ = 𝑚𝑚𝑜𝑟 𝛺 (1+𝑖𝜆)
2 . (2.140)
1−𝛺 +𝑖𝑔
The vector diagram of 𝑥̃ is a circle, plotted with solid line in Fig.2.70. On
the same graph, the polar plot for excitation without forces in quadrature (𝜆 = 0)
is drawn with dashed line.
The two circles cross each other at the point 𝑅, which on the circle 𝜆 = 0
corresponds to a frequency 𝜔𝑟 = 𝜔𝑛 √1 + 𝑔2 , and on the circle 𝜆 ≠ 0 - to a
1+𝑔2
frequency 𝜔∗ = 𝜔𝑛 √1−𝜆𝑔 .
The damping factor is given by [93]
1 𝑟 𝜔2
𝑔 = 𝜆 (1 − 𝜔∗2 ). (2.141)

It should be noted that the crossing point 𝑅, which corresponds to the


amplitude resonance (on the circle 𝜆 = 0), is different from the point 𝑀, which
defines the phase resonance. As the location of 𝑅 is independent of the coordinate
system, the method has a fairy good accuracy.

69
Point 𝑅 can also be located on the circle 𝜆 = 0, using the Kennedy-Pancu
criterion, in the region where the distance between two successive data points
plotted at regular frequency intervals is a maximum. The diameter at resonance is
𝑚 𝑟 1
̅̅̅̅
𝑂𝑅 = 𝑚𝑜 √𝑔2 + 1 . (2.142)

Crossing the circle 𝜆 = 0 with two circles, plotted using two masses
located at ±90°, allows a more accurate location of the resonance. The damping
factor expressions are [93]
𝜔∗2 −𝜔∗∗2 1 𝜔2𝑟 𝜔2𝑟
𝑔=
𝜆1 𝜔∗2 −𝜆2 𝜔∗∗2
=𝜆 −𝜆
( ∗∗2 − 𝜔∗2
) (2.143)
1 2 𝜔

where 𝜔𝑟 , 𝜔∗ and 𝜔∗∗ are the frequencies of the point 𝑅 on the three circles of
parameters 𝜆 = 0, 𝜆1 and 𝜆2 , respectively (Fig.2.71).

2.4.2. Identification of Systems with Viscous Damping

When the system from Fig.2.34 is excited by a force (2.133), the equation
of motion becomes
𝑚𝑥̈ + 𝑐𝑥̇ + 𝑘𝑥 = 𝑚𝑜 𝑟𝜔2 𝑒 𝑖𝜔𝑡 . (2.144)
For a steady-state solution, the complex amplitude of the displacement is
𝛺2
𝑥̃ = 𝑚𝑚𝑜𝑟 = 𝑥̂ 𝑒𝑖𝜑𝛼 . (2.145)
1−𝛺2 +𝑖2𝜁𝛺
Equation (2.145) is similar to the expression of the complex inertance so
that the identification methods used in § 2.2.2.4 can be applied.

70
Method of displacement amplitude

Figure 2.72 illustrates graphically the variation with frequency of the


modulus of the function (2.145) for 𝜁 = 𝑐𝑜𝑛𝑠𝑡. A line through the origin intersects
the curve at the points 𝐾 and 𝐿, of frequencies 𝜔𝑘 and 𝜔ℓ , from which we can
obtain an expression for the natural frequency
𝜔𝑛 = √𝜔𝑘 𝜔ℓ . (2.146)

A horizontal line crosses the curve at the points 𝑃 and 𝑄, of frequencies 𝜔𝑝


and 𝜔𝑞 . The damping factor is given by
1 𝜔2 𝜔2
2𝜁 2 = 1 − ( 𝑛2 − 𝑛2 ). (2.147)
2 𝜔 𝑝𝜔 𝑞

The mass 𝑚 is determined from the ordinate of the asymptote at 𝜔 → ∞.

Method of the vector components of the dynamic stiffness

Measuring the displacement amplitude 𝑥̂ and phase angle 𝜑𝛼 at different


frequencies 𝜔, we can draw through the data points the graphs of the following
quantities
𝑚𝑜 𝑟𝜔2 𝑚𝑜 𝑟𝜔2
𝑐𝑜𝑠 𝜑𝛼 = 𝑘 − 𝑚𝜔2 , 𝑠𝑖𝑛 𝜑𝛼 = 𝑐, (2.148)
𝑥̂ 𝑥̂

as functions of 𝜔2 (Fig.2.73). They are straight lines.


The method, first used by Lorenz [71], has been used with good results by
several authors [53].

Method of the diagram of velocity

The variation with frequency of the amplitude of velocity of the mass 𝑚 is


illustrated in Fig.2.74 by the curve of equation

71
𝑚 𝑟 𝑘 𝛺2
𝑥̇̂ = 𝜔𝑥̂ = 𝑚𝑜 √𝑚 2
. (2.149)
√4𝜁 2 +(𝛺− 1 )
𝛺

The points of extrema 𝑅 and 𝐴 have abscissae


2
𝜔𝑅,𝐴 = 𝜔𝑛2 [2(1 − 2𝜁 2 ) ∓ √4(1 − 2𝜁 2 )2 − 3]. (2.150)

The natural frequency 𝜔𝑛 is given by


𝜔𝑅 𝜔 𝐴
𝜔𝑛2 = . (2.151)
√3
The damping ratio can be expressed as
2 2
𝜔𝑅 +𝜔𝐴
2𝜁 2 = 1 − 2 . (2.152)
4𝜔𝑛
𝑚𝑜 𝑟
The asymptote to 𝜔 → ∞ starts from the origin and has a slope tan𝜃 = 𝑚
𝜔𝑛
The line intersects the curve at the point 𝑆, of frequency 𝜔𝑠 = and
√2(1−2𝜁 2 )
ordinate
𝑚 𝑟 𝜔𝑛
𝑥̇̂𝑠 = 𝑜 . (2.153)
𝑚 √2(1−2𝜁 2 )
The mass 𝑚 can be calculated from (2.153).
Other identification methods are presented in [72], based on the
measurement of the power delivered by the vibrator or of the force transmitted at
the supporting point of the resilient element.

2.5. HARMONIC EXCITATION OF UNGROUNDED SYSTEMS

For the measurement of the dynamic characteristics of vibration isolator


pads, the resilient element is introduced in a mounting whose configuration can be
easily modeled by a lumped parameter mechanical system.

72
When an unyielding supporting surface is not available, the experimental
arrangement is ungrounded. This section deals with methods for the identification
of base-excited systems.

2.5.1. Harmonic Excitation through the Resilient Element

In the following we analyze single-degree-of-freedom systems excited by


either a harmonic force acting through the spring and damper, or a prescribed
harmonic displacement imposed to the base.

2.5.1.1. Identification of systems with hysteretic damping

Consider the base-excited system from Fig.2.75, acted upon by a force


𝑓(𝑡) = 𝑓̂𝑒 𝑖𝜔𝑡 . Denoting 𝑦(𝑡) = 𝑦̃𝑒 𝑖𝜔𝑡 the base displacement and 𝑥(𝑡) = 𝑥̃𝑒 𝑖𝜔𝑡 the
displacement of the mass 𝑚, we obtain
𝑦̃ 1 1−𝛺2 +𝑖𝑔 1 1 1 𝑔
=− = [( − ) − 𝑖 1+𝑔2], (2.154)
𝑓̂ 𝑘𝛺2 1+𝑖𝑔 𝑘 1+𝑔2 𝛺2
𝑥̃−𝑦̃ 𝛺2
= , (2.155)
𝑦̃ 1−𝛺2 +𝑖𝑔
𝑥̃ 1+𝑖𝑔
= . (2.156)
𝑦̃ 1−𝛺2 +𝑖𝑔
The vector diagram of the function (2.154) is a straight line parallel to the
real axis (Fig.2.76). The line intersects the imaginary axis at the antiresonance point
𝐴, of frequency 𝜔𝐴 = 𝜔𝑛 √1 + 𝑔2 [90]. The lines 𝑂𝐵ˈ and 𝑂𝐶ˈ, inclined ±45° to
the imaginary negative semiaxis, locate on the diagram the frequencies
𝜔
𝜔ˈ, 𝜔" = 𝐴 . The damping factor is given by
√1±𝑔
2 2
𝜔" −𝜔ˈ
𝑔= 2 2 . (2.157)
𝜔" +𝜔ˈ
The stiffness is
𝑔
𝑘=
(1+𝑔2 )𝑂𝐴
̅̅̅̅̅ , (2.158)

̅̅̅̅ is the response at antiresonance.


where 𝑂𝐴
The polar diagram of the inverse of the function (2.154) is a circle. It can
be obtained measuring the variation with frequency of the amplitude 𝑓̂ of the force
required to maintain the displacement amplitude 𝑦̂ constant.
The damping is evaluated from expression (2.157) where 𝜔ˈ and 𝜔" are the
frequencies of the half power points. The stiffness is calculated from the value of
the diameter of the circle, equal to the dynamic stiffness at the frequency of
antiresonance 𝜔𝐴 .

73
The expression of the relative transmissibility (2.155) is similar to the
expression of the inertance of the system from Fig.2.7. In this case, the system
parameters have expressions similar to those presented in § 2.2.1.4 [102].
In practice it is customary to use the polar diagram of the absolute
transmissibility (2.156), depicted in Fig.2.33. In this case, only the damping factor
can be calculated using the procedure from § 2.2.1.8.
Expressions (2.155) and (2.156) can also be obtained for a given kinematic
base excitation of the system from Fig.2.75.

2.5.1.2. Identification of systems with viscous damping

Consider the system from Fig.2.77, excited by a known force 𝑓(𝑡) =


̂
𝑓 𝑒 . Let 𝑦̇ = 𝑦̇̃ 𝑒 𝑖𝜔𝑡 be the base velocity and 𝑥̇ = 𝑥̇̃ 𝑒 𝑖𝜔𝑡 the velocity of the mass.
𝑖𝜔𝑡

We can establish the following frequency response functions (mobilities):


̃ 2𝜁𝛺2 (1−4𝜁 2 )𝛺2 −1
̅ 𝑦 = 𝑦̇ =

1
+𝑖
1
, (2.159)
̂ 𝑓 √𝑘𝑚 1+4𝜁 2 𝛺 2 √𝑘𝑚 𝛺(1+4𝜁 2 𝛺2 )
̃
̅ 𝑥 = 𝑥̇ =

1 −𝑖
, (2.160)
̂ 𝑓 √𝑘𝑚 𝛺
̃ ̃ 2𝜁𝛺2
̅ 𝑥𝑦 = 𝑦̇ −𝑥̇ =

1
+𝑖
1 𝛺
, (2.161)
̂ 𝑓 √𝑘𝑚 1+4𝜁 2 𝛺2 √𝑘𝑚 1+4𝜁 2 𝛺2
𝑥̇ ̃ 1+𝑖2𝜁𝛺
𝒯̅𝑎𝑏𝑠 = ̂ = , (2.162)
𝑦̇ 1−𝛺2 +𝑖2𝜁𝛺
̃ ̃
𝑥̇ −𝑦̇ 𝛺2
𝒯̅𝑟𝑒𝑙 = ̂ = . (2.163)
𝑦̇ 1−𝛺2 +𝑖2𝜁𝛺
Denoting
̅ 𝑦 = ℳ𝑦𝑅 + 𝑖ℳ𝑦𝐼
ℳ (2.159, a)
and eliminating the frequency between the expressions of the two vector
components of the function (2.159), we obtain a curve of equation
2𝜁 2 1 2
(ℳ𝑦𝑅 −
√𝑘𝑚
) (2𝜁√𝑘𝑚ℳ − 1) = ℳ𝑦𝐼 , (2.164)
𝑦𝑅

represented graphically in Fig.2.78.


The frequency 𝜔𝑛 is determined approximately at the antiresonance point
𝐴. If the mass 𝑚 is known, we can calculate 𝑘. The curve intersects the real axis at
2𝜁
the point 𝑀, of abscissa ̅̅̅̅̅
𝑂𝑀 = , wherefrom we can determine the damping
√𝑘𝑚
ratio 𝜁.
The polar diagram of the function (2.160) is a straight line which overlaps
the negative imaginary semiaxis. The polar diagram of the function (2.161) is a

74
semicircle (Fig.2.79). None of them gives enough information for the identification
of system parameters.

Function (2.162) has been examined in § 2.2.2.8. Function (2.163) is


similar to the complex inertance of the mass driven system (Fig.2.34) analyzed in
§ 2.2.2.4.
Better results are obtained from the graphical analysis of the polar diagram
of the direct impedance, i.e. the inverse of the function (2.159)
𝑓̃ 2𝜁𝛺 4 2
𝛺(1−𝛺 +4𝜁 𝛺 ) 2 2
𝑍̅𝑦 = 𝑦̇̂ = √𝑘𝑚 (1−𝛺2 )2 +4𝜁 2𝛺2 + 𝑖√𝑘𝑚 (1−𝛺2 )2 +4𝜁 2𝛺2 . (2.165)

In figure 2.80, the graph of the dimensionless function 𝑍𝑦̅ ⁄√𝑘𝑚 is plotted
with solid line, for 𝜁 = 0.1. The curve emanates from the origin at 𝛺 = 0, reaches
the peak value at the antiresonance frequency 𝛺1 ≅ 1, intersects the real axis at
1
𝛺𝑀 = 2
and ends on the real axis at the point 𝑁 (2𝜁, 0) for 𝛺 → ∞. Point 𝐴
√1−4𝜁
can be located using the Kennedy-Pancu criterion, where the arc of circle between
two successive data points plotted at regular frequency intervals is a maximum.
In figure 2.80, with dashed lines are drawn the polar diagrams of the
function 𝑍̅𝑦 ⁄√𝑘𝑚 for excitation with forces in quadrature 𝑓(𝑡) = 𝑓̂(1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 ,
for 𝜆 = ±0.5 and 𝜆 = ±1. The curve 𝜆 = 0 intersects these diagrams at points
defined approximately by the frequency 𝛺𝐴 = 1.
Over a very narrow band of frequencies in the immediate vicinity of
antiresonance, the curve 𝜆 = 0 can be approximated by a circle with the center at
the middle of the diameter 𝑂𝐴, of equation
𝑍̅𝑦 1+𝑖2𝜁
= = 𝑍𝑦𝑅 + 𝑖𝑍𝑦𝐼 . (2.166)
√𝑘𝑚 2𝜁−𝑖(1−𝛺2 )

75
Analogously, at frequencies near the intersection with the curve 𝜆 = 0, the
diagrams obtained using excitation with forces in quadrature can be approximated
𝑍𝑦𝑅 +𝑖𝑍𝑦𝐼
by circles of equation .
1+𝑖𝜆

The circle of parameter 𝜆 intersects the circle 𝜆 = 0 at the point of


frequency 𝛺 = √1 + 2𝜁𝜆 . Drawing two circles, of parameters 𝜆1 and 𝜆2 , the
frequencies 𝛺1 and 𝛺2 of the crossing points with the circle 𝜆 = 0 can be used to
calculate the damping ratio
𝛺22 −𝛺12 1 𝜔22 −𝜔12
𝜁= = 2 . (2.167)
2(𝜆2 −𝜆1 ) 𝜆2 −𝜆1 2𝜔𝐴
Expression (2.167) yields a good approximation of the value of 𝜁 as well
as for the diagrams from Fig.2.80. If the mass is known, the stiffness is 𝑘 = 𝑚𝜔𝐴2 .

76
2.5.2. Harmonic Excitation of the Mass-Damped Spring-Mass System

The construction of a practical set-up as depicted in Fig.2.75 is difficult,


because the force cannot be applied directly to the resilient pad, but through a rigid
element which has inherent mass. The compensation of this mass requires
additional analysis or special equipment.
The simple model from Fig.2.75 is replaced by a system comprising two
masses, referred to as the mass - damped spring - mass system (Fig.2.81). If the
selected coordinate is the relative displacement of the two masses, this system has
a single vibratory degree of freedom. The relative displacement is proportional to
the dynamic strain in the resilient material, fact that allows a correct definition of
the resonance.

2.5.2.1. Identification of systems with hysteretic damping

Consider the system from Fig.2.81 with a harmonic force


𝑓(𝑡) = 𝑓̂𝑒 𝑖𝜔𝑡 applied on the mass 𝑚1 . Let 𝑦 = 𝑦̃𝑒 𝑖𝜔𝑡 be the
displacement of the mass 𝑚1 and 𝑥 = 𝑥̃𝑒 𝑖𝜔𝑡 the displacement
of the sprung mass 𝑚2 .
Denoting
𝑘 𝜔 𝑚 ℎ
𝑝 = √𝑚 , 𝛺 = 𝑝
, 𝛾 = 𝑚1 , 𝑔 = 𝑘 (2.168)
2 2

we obtain the following frequency response functions


𝑦̃ 1 1−𝛺2 +𝑖𝑔
𝛼̅𝑦 = =− , (2.169)
𝑓̂ 𝑘(1+𝛾)𝛺2 1−1+𝛾
𝛾 2
𝛺 +𝑖𝑔

𝑥̃ 1 1+𝑖𝑔
𝛼̅𝑥 =
𝑓̂
=− 2 𝛾 2 , (2.170)
𝑘(1+𝛾)𝛺 1−1+𝛾𝛺 +𝑖𝑔

𝑦̃−𝑥̃ 1 1
𝛼̅𝑥𝑦 = = 𝛾 . (2.171)
𝑓̂ 𝑘(1+𝛾) 1− 𝛺2 +𝑖𝑔
1+𝛾

Relative receptance method

The polar diagram of the function (2.171), sketched in Fig.2.82 with solid
line, is a circle, tangent at the origin to the real axis. The circle intersects the
negative imaginary semiaxis at the point 𝑀 of frequency
1 1
𝜔𝑀 = √𝑘 (𝑚 + 𝑚 ) (2.172)
1 2

which coincides with the undamped natural frequency.

77
During a frequency sweep with a force of constant amplitude, the relative
displacement of the two masses is a maximum at the frequency 𝜔𝑀 .
Point 𝑀 can be localized on the circle 𝜆 = 0 using the Kennedy-Pancu
∆𝑠
procedure, where the ratio ∆𝜔 is a maximum (∆𝑠 is the arc of circle corresponding
to an increase ∆ω of the frequency). It can also be localized crossing it with two
circles, drawn with dashed lines, obtained using excitation with forces in
quadrature.

The diameter at resonance is


̅̅̅̅̅ =
𝑂𝑀
1
ℎ(1+𝛾)
= 𝑘𝑔(𝑚𝑚2+𝑚 ) = 𝑔𝑚1𝜔2 . (2.173)
1 2 1 𝑀

The diameter BC, perpendicular to 𝑂𝑀, locates the frequencies


2 2 (1
𝜔1,2 = 𝜔𝑀 ∓ 𝑔) (2.174)
which can be used to calculate the damping factor
𝜔22 −𝜔21 𝜔22 −𝜔21
𝑔=
𝜔22 +𝜔21
= 2𝜔2𝑀
. (2.175)

From equations (2.172) and (2.173) we determine 𝑘 and one mass, when
the other mass is known.

Direct receptance method

The polar diagram of the function (2.169) is plotted in Fig.2.83 with solid
line. Resonance, defined by the frequency 𝜔𝑀 , at which “a minimum of force
produces a maximum of relative displacement” of the two masses, occurs at the
point 𝑀1 , where the imaginary component of the direct receptance has an extreme
value.

78
The abscissa of point 𝑀1 is independent of the damping. Its modulus is
𝛾 1 1
̅̅̅̅̅̅̅
𝑀1 𝐿1 =
(1+𝛾)2
= (𝑚 2 . (2.176)
𝑘 1 +𝑚2 )𝜔𝑀

The ordinate of the point 𝑀1 (absolute value) is


1 𝑘 𝑚2
̅̅̅̅̅̅̅
𝐾1 𝑀1 = = 4 = 2 . (2.177)
ℎ(1+𝛾)2 𝑔𝑚12 𝜔𝑀 𝑔𝑚1 (𝑚1 +𝑚2 )𝜔𝑀
The frequency 𝜔𝑀 and the coordinates of the point 𝑀1 can be determined
experimentally.
If one mass is known, then equation (2.176) yields the other mass, then,
from equation (2.172) we can determine 𝑘 and from (2.177) – the damping factor.
If both masses are unknown, we can make an approximate identification,
𝑘
considering that the antiresonance frequency is 𝜔𝐴 ≈ 𝑝 = √ .
𝑚2
Two curves of the direct receptance, obtained for excitation with forces in
quadrature 𝑓(𝑡) = 𝑓̂(1 + 𝑖𝜆)𝑒 𝑖𝜔𝑡 , are plotted in Fig.2.83 with dashed lines. The
point 𝑅1 , located on the diagram 𝜆 = 0 at the middle of the distance between the
points of intersection with the diagrams = ±0.5 , coincides approximately with the
point of maximum response amplitude. It follows that neither the peak amplitude
criterion nor the method of the three diagrams can be used for the location of
resonance.

Transfer receptance method

The polar diagram of the function (2.170) is plotted in Fig.2.84 with solid
line. Two other curves for excitation with forces in quadrature (for 𝜆 = ±0.5) are
plotted with dashed lines. These curves cross the diagram 𝜆 = 0 at point 𝑅2 of
maximum amplitude and frequency

79
2
1+𝛾 𝜔𝑀
𝜔𝑅2 = 𝑝2 (3 + √1 − 8𝑔2 ) = (3 + √1 − 8𝑔2 ). (2.178)
4𝛾 4

Point𝑅2 is different from the point 𝑀, of frequency 𝜔𝑀 , in which the


imaginary component of the receptance has an extreme value.
The abscissa of point 𝑀2 is independent of damping
1
̅̅̅̅̅̅̅
𝑀2 𝐿2 = 2 . (2.179)
(𝑚1 +𝑚2 )𝜔𝑀
and the ordinate is
𝛾 1
̅̅̅̅̅̅̅
𝐾2 𝑀2 = = 𝑔(𝑚 2 . (2.180)
ℎ(1+𝛾)2 1 +𝑚2 )𝜔𝑀

The curve 𝜆 = 0 intercepts the imaginary axis at point 𝐹2 of frequency


𝜔𝐹 = 𝜔𝑀 √1 + 𝑔2 . (2.181)
If one mass is known, e.g. 𝑚1 , from (2.172), (2.179) and (2.180) we can
calculate 𝑚2 , 𝑘 and 𝑔.
The graph from Fig.2.82 can be plotted point by point, combining the
diagrams from Fig.2.83 and Fig.2.84 as in Fig.2.85.

2.5.2.2. Identification of systems with viscous damping

Consider the system from Fig.2.86 with a harmonic force 𝑓(𝑡) = 𝑓̂𝑒 𝑖𝜔𝑡
applied on the mass 𝑚1 . Let 𝑦̇ = 𝑦̇̃ 𝑒 𝑖𝜔𝑡 be the velocity of the mass 𝑚1 and
𝑥̇ = 𝑥̇̃ 𝑒 𝑖𝜔𝑡 the velocity of the mass 𝑚2 .
Denoting
𝑘 𝜔 𝑚 𝑐
𝑝 = √𝑚 , 𝛺 = 𝑝
, 𝛾 = 𝑚1 , 𝜁 = (2.182)
2 2 2√𝑘𝑚2

we obtain the following frequency response functions (mobilities)


̃ 2𝜁𝛺−𝑖(1−𝛺2 )
̅ 𝑦 = 𝑦̇
ℳ = 1 1
, (2.183)
𝑓̂ √𝑘𝑚2
1+𝛾 𝛺(1− 𝛾 𝛺2 )+𝑖2𝜁𝛺2
1+𝛾

̃
̅ 𝑥 = 𝑥̇ =

1 1 2𝜁𝛺−𝑖
, (2.184)
̂ 𝑓 √𝑘𝑚2
1+𝛾 𝛺(1−1+𝛾𝛺 )+𝑖2𝜁𝛺2
𝛾 2

̃ ̃
̅ 𝑥𝑦 = 𝑦̇ −𝑥̇ =

1 1 1
. (2.185)
̂ 𝑓 1+𝛾 1 𝛾 2
√𝑘𝑚2 2𝜁−𝑖 𝛺
(1− 1+𝛾
𝛺 )

Relative mobility method

In the complex plane, the locus of the affixes of the function (2.185) is a
circle, tangent at the origin to the imaginary axis (Fig.2.87).

80
The circle intersects the real axis at the point 𝑀, of frequency
1 1 1+𝛾
𝜔𝑀 = √𝑘 ( + ) = 𝑝√ . (2.186)
𝑚1 𝑚2 𝛾

The diameter at resonance is


1 1
̅̅̅̅̅
𝑂𝑀 = 2𝜁(1+𝛾)
= 2𝜁𝑝(𝑚1 +𝑚 ) = 2𝜁𝑚1 𝜔 √1+𝛾 .
𝛾
(2.187)
√𝑘𝑚2 1 2 1 𝑀

The diameter 𝐵𝐶, perpendicular to 𝑂𝑀, locates on the circle the


frequencies 𝜔ˈ and 𝜔", which satisfy the following equations
2
𝜔ˈ𝜔" = 𝜔𝑀 , (2.188)
2
1+𝛾 𝜔𝑀 1+𝛾
𝜔" − 𝜔ˈ = 2𝜁 𝛾
𝑝 = 2𝜁 𝑝
= 2𝜁𝜔𝑀 √ 𝛾
. (2.189)

If one mass is known, e.g. 𝑚1 , from (2.187) and (2.189) we obtain 𝜁 and
𝛾, hence the mass 𝑚2 . Then we get 𝑘 from (2.186).

Direct mobility method

The polar diagram of the function (2.183) is plotted in Fig.2.88 with solid
line. Resonance occurs at the point 𝑀1 , of frequency 𝜔𝑀 , where the real component
of the direct mobility is a maximum. The coordinates of the point 𝑀1 (absolute
values) are
1 1
̅̅̅̅̅̅̅
𝐾1 𝑀1 =
2𝜁 (1+𝛾)2
= 2𝜁𝜔1 𝑚 𝛾
√(1+𝛾)3 , (2.190)
√𝑘𝑚2 𝑀 1

81
1 1 𝛾 1 1
̅̅̅̅̅̅̅
𝐿1 𝑀1 = 1+𝛾 √1+𝛾 =𝑚 = (𝑚 . (2.191)
√𝑘𝑚2 2 𝜔𝑀 (1+𝛾) 1 +𝑚2 )𝜔𝑀

The ordinate ̅̅̅̅̅̅̅


𝐿1 𝑀1 is independent of the damping level.
If the mass 𝑚1 is known, we can calculate 𝜁 and γ from (2.190) and (2.191),
then from (2.186) we obtain 𝑘.
If both masses are unknown, then we can consider that antiresonance (point
𝐴) occurs at the frequency
𝑘
𝜔𝐴 ≈ 𝑝 = √𝑚 . (2.192)
2
The four parameters 𝜁, 𝑘, 𝑚1 and 𝑚2 can be determined from equations
(2.186), (2.190), (2.191) and (2.192).
The point 𝑀1 cannot be localized using excitation with forces in
quadrature. The curves 𝜆 = ±0.5 do not intersect the curve 𝜆 = 0 at the same point.
In this case we can use the point 𝑅1 of maximum amplitude [91].

Transfer mobility method

The polar diagram of the function (2.184) is plotted with solid line in
Fig.2.89 for 𝜁 = 𝑐𝑜𝑛𝑠𝑡.

The point 𝑀2 , of frequency 𝜔𝑀 (2.186), corresponds to the extreme value


of the real component of the transfer mobility and has coordinates
1 𝛾 3
̅̅̅̅̅̅̅
𝐾2 𝑀2 = 2𝜁𝜔 𝑚 √( , (2.193)
𝑀 1 1+𝛾 )3

82
1 1
𝐿2 𝑀2 = 𝜔 𝑚 (1+𝛾) = 𝜔 𝑚 +𝑚
̅̅̅̅̅̅̅ . (2.194)
𝑀 2 𝑀 ( 1 2)

The ordinate is independent of the level of damping and is equal to ̅̅̅̅̅̅̅


𝐿1 𝑀1
(Fig.2.88).
If the mass 𝑚1 is known, from (2.193) and (2.194) we obtain 𝜁 and 𝑚2 ,
and from (2.186) – the stiffness 𝑘.
The method of forces in quadrature locates the point 𝑅2 , of maximum
response amplitude, which is different from 𝑀2 .
The frequency of the point 𝐹, in which the curve 𝜆 = 0 intersects the real
axis, is
𝜔𝑀
𝜔𝐹 = 1+𝛾
.
√1− 𝛾 4𝜁 2

2.6. IDENTIFICATION METHODS WITHOUT FREQUENCY SWEEP

In this section we discuss several identification methods which are not


based on the graphical analysis of frequency response curves plotted with a
frequency sweep.

2.6.1. Method of “Displaced” Frequencies

The natural frequency of the single degree of freedom system is a function


of the mass 𝑚 and the stiffness 𝑘. Modifying these parameters we can produce a
“displacement” (variation, shift) of the resonance frequency used to estimate the
parameters of the nonmodified system.

2.6.1.1. Method of additional masses

The complex receptance of the system from Fig.2.7 is given by


1 1
𝛼̅ = 𝑘 𝑚 2 . (2.195)
1− 𝑘 𝜔 +𝑖𝑔

On the corresponding polar diagram (Fig.2.90, a), the point 𝑀, of


𝑘
frequency 𝜔𝑛 = √ is located using the resonance criterion discussed in § 2.2.
𝑚
If a mass ∆𝑚 is added to the initial mass 𝑚, so that
∆𝑚
𝛾= (2.196)
𝑚
the complex receptance of the modified system can be written as

𝛼̅ = 1𝑘 1−𝑚𝜔2(11+𝛾)+𝑖𝑔 . (2.197)
𝑘

83
The polar diagram of the function (2.197) is a circle of the same diameter
1
(Fig.2.90, b) but with a different distribution of the parameter ω.
𝑘𝑔

The resonance frequency of the modified system, corresponding to the


point 𝑅, is
𝜔
𝜔∗ = 𝑛 . (2.198)
√1+𝛾

The point 𝑀1 , of frequency 𝜔𝑛 , is the tip of the vector ̅̅̅̅̅̅


𝑂𝑀1 , which is
oriented at an angle
𝛾
𝜃 = 𝑡𝑎𝑛−1 𝑔 (2.199)
to the diameter 𝑂𝑅.
The mass of the initial system is determined from (2.198)
∆𝑚
𝑚= 𝜔 2
. (2.200)
( 𝑛∗ ) −1
𝜔
and the stiffness is
𝑘 = 𝜔𝑛2 𝑚. (2.201)
The damping factor is obtained from (2.199)
∆𝑚
𝑔= 𝑚
𝑐𝑜𝑡𝜃. (2.202)
If the measurement system is calibrated, the diameter 𝑂𝑀 is known and 𝑔
is given by the equation
1
𝑔 = ̅̅̅̅̅ . (2.203)
𝑘 𝑂𝑀

In practical applications, an approximate formula has been used [126]


which gives satisfactory results when the added masses are small in comparison
with the initial mass. Denoting the variation of the resonance frequency
∆𝜔𝑛 = 𝜔𝑛 − 𝜔∗, (2.204)
the stiffness 𝑘 can be expressed in terms of the quantities before and after the
addition of the mass 𝑚
𝑘 = 𝑚𝜔𝑛2 = (𝑚 + ∆𝑚)(𝜔𝑛 − ∆𝜔𝑛 )2.

84
Neglecting infinitesimals of higher order, the mass 𝑚 can be expressed as
𝜔𝑛 ∆𝑚
𝑚≅ . (2.205)
2 ∆𝜔𝑛
∆𝜔𝑛
Expression (2.205) shows that plotting the ratio as a function of ∆𝑚
𝜔𝑛
(Fig.2.91) we obtain a straight line whose inclination provides the mass 𝑚.

2.6.1.2. Method of additional stiffnesses

If the stiffness of the system from Fig.2.7 is modified by adding a stiffness


∆𝑘 so that
∆𝑘
𝜘= , (2.206)
𝑘
the complex receptance (2.195) becomes
1 1
𝛼̅ = 𝑘(1+𝜘 ) 1− 𝑚
𝜔2 +𝑖𝑔
. (2.207)
𝑘(1+𝜘)

The polar plots of the functions (2.195) and (2.207) are presented in
Fig.2.92.
The resonance frequency of the modified system is 𝜔∗∗ = 𝜔𝑛 √1 + 𝜘 .
The initial stiffness can be expressed as
∆𝑘
𝑘= 𝜔∗∗
2 (2.208)
( ) −1
𝜔𝑛
and the mass
𝑘
𝑚 = 𝜔2 . (2.209)
𝑛

On the polar plot of the modified system let denote


by 𝜃 the angle subtended by the line 𝑂𝑀1 (the response at 𝜔𝑛 )
and the diameter 𝑂𝑅 (the negative imaginary semiaxis). The
damping factor is given by
cot𝜃
𝑔= 𝑘 (2.210)
1+
∆𝑘

which does not require a calibration of the measurement


system.
The stiffness can be calculated from the ratio of the
diameters of the two circles
̅̅̅̅̅
𝑂𝑀 ∆𝑘
̅̅̅̅
𝑂𝑅
=1+ 𝑘
. (2.211)
In practice, however, altering the stiffness is less easy than altering the
mass, hence the method is not implemented as presented above. It was applied by
G. de Vries [130] using “electrical stiffnesses” which requires special equipment.

85
The method of forces in quadrature presented in § 2.3 can be considered as
a variant of the method of additional stiffnesses [131].
Equation (2.120) can be written under the form
𝑘
(𝑖𝑚𝜔 + 𝑐 + ) 𝑥̇̃ = 𝑓̂ + 𝑖𝜆𝑓̂ . (2.212)
𝑖𝜔

Equating the real and imaginary parts, for 𝜔 = 𝜔ˈ, we obtain (Fig.2.65)
𝑐𝑥̇̂ = 𝑓̂, (2.213)
𝑘
(𝑖𝑚𝜔ˈ + 𝑖𝜔ˈ) 𝑥̇̂ = 𝑖𝜆𝑓̂ , (2.214)
or
𝜆𝜔ˈ𝑓̂
𝑚𝜔ˈ2 − (𝑘 + 𝑥̇̂
) = 0. (2.215)
The term
𝜆𝜔ˈ𝑓̂ 𝑓̂
= 𝜆 ̂ = 𝑘ˈ (2.216)
𝑥̇̂ 𝑥
plays the role of an additional stiffness.
We can write
𝑚𝜔𝑛2 = 𝑘 ,
𝑚𝜔ˈ2 = 𝑘 + 𝑘ˈ,
which yields
𝑘ˈ 𝜔ˈ 𝜆𝑓̂
𝑚= 2 = 2 ̂ . (2.217)
𝜔ˈ −𝜔2𝑛 𝜔ˈ −𝜔2𝑛 𝑥̇
Using equations (2.213) and (2.217), we obtain formula (2.126).

2.6.2. Method of Introduced Work

In the case of stationary forced vibrations, the energy supplied per cycle by
the excitation (equal to the work of the excitation force) has the expression
𝑡+𝑇 𝑡+𝑇 𝑡+𝑇
𝑑
∫ 𝑓(𝑡)𝑑𝑥 = ∫ 𝑓(𝑡)𝑥̇ (𝑡)𝑑𝑡 = ∫ 𝑓̂𝑐𝑜𝑠𝜔𝑡 [𝑥̂𝑐𝑜𝑠(𝜔𝑡 + 𝜑)] 𝑑𝑡 = 𝜋𝑓̂𝑥̂𝑠𝑖𝑛𝜑
𝑑𝑥
𝑡 𝑡 𝑡
At resonance, it is equal to the energy dissipated by damping [125].
For 𝜔 = 𝜔𝑛
𝑊𝑑 = 𝜋𝑓̂ 𝑥̂𝑜 (2.218)
hence 𝑊𝑑 can be determined using a wattmeter, in the case of excitation with
electromagnetic vibrators, or measuring 𝑓̂ and 𝑥̂𝑜 .
For the system with viscous damping
𝑓̂ = 𝑐𝜔𝑛 𝑥̂𝑜 = 2𝜁𝑚𝜔𝑛2 𝑥̂𝑜 (2.219)
whence

86
𝑊𝑑 = 𝜋𝑐𝜔𝑛 𝑥̂𝑜2 = 2𝜋𝜁𝑚𝜔𝑛2 𝑥̂𝑜2. (2.220)
The maximum kinetic energy of the system is
1
𝑊𝑐 = 2 𝑚𝜔𝑛2 𝑥̂𝑜2 . (2.221)
From equations (2.220) and (2.221) we obtain the known expression of the
damping ratio [144]
1 𝑊
𝜁 = 4𝜋 𝑊𝑑 . (2.222)
𝑐
The above formula can be used only when the mass 𝑚 is known.
It comes out that the energy method provides values for the product
𝑓̂ 𝑊
𝜁𝑚 = 2 𝑥̂ = 2𝜋𝜔2𝑑𝑥̂ 2 (2.223)
2𝜔𝑛 𝑜 𝑛 𝑜

so that one of the two parameters has to be determined using another method.
𝑓̂
Sometimes, the ratio ̂ from equation (2.223) is determined from the slope
𝑥𝑜
̂
of the line plotted with 𝑓 as ordinate and 𝑥̂𝑜 as abscissa, where 𝑥̂𝑜 is measured as
the displacement amplitude at resonance.

2.6.3. Method of Transient Response

The differential equation of the forced vibrations of the single-degree-of-


freedom system with viscous damping
𝑚𝑥̈ + 𝑐𝑥̇ + 𝑘𝑥 = 𝑓̂ 𝑐𝑜𝑠𝜔𝑡 (2.224)
has a general solution of the form

𝑥(𝑡) = 𝑒 −𝜁𝜔𝑛 𝑡 [𝐴 𝑐𝑜𝑠(√1 − 𝜁 2 𝜔𝑛 𝑡) + 𝐵 𝑠𝑖𝑛(√1 − 𝜁 2 𝜔𝑛 𝑡)] + 𝑥̂ 𝑐𝑜𝑠(𝜔𝑡 + 𝜑)


(2.225)
in which
𝑘 𝑐 𝑓̂ −𝑐𝜔
𝜔𝑛 = √𝑚 , 𝜁 = , 𝑥̂ = , 𝜑 = 𝑡𝑎𝑛−1 𝑘−𝑚𝜔 2 (2.226)
2𝑚𝜔𝑛 √(𝑘−𝑚𝜔 2 )2 +(𝑐𝜔)2

𝐴 and 𝐵 being constants of integration dependent on the initial conditions.


The first term in the right hand side of expression (2.225) is the solution of
equation (2.224) for 𝑓̂ = 0. The second term is the particular solution for the
harmonic excitation with a force of constant amplitude.

2.6.3.1. Starting transient

Consider the transient motion of the system from Fig.2.7, starting from rest,
acted at the time 𝑡 = 0 by a force 𝑓(𝑡) = 𝑓̂ 𝑐𝑜𝑠𝜔𝑛 𝑡, of frequency equal to the
natural frequency of the undamped system.

87
𝜋 𝑓̂ 𝑓̂
For 𝜔 = 𝜔𝑛 in equations (2.226), 𝜑 = − 2 , 𝑥̂ = 𝑐𝜔 = 2𝜁𝑘 and because at
𝑛
𝑡 = 0, 𝑥 = 0 and 𝑥̇ = 0, the general solution (2.225) becomes
𝑓̂ 𝑒 −𝜁𝜔𝑛𝑡
𝑥(𝑡) = 2𝜁𝑘 [𝑠𝑖𝑛𝜔𝑛 𝑡 − 𝑠𝑖𝑛(√1 − 𝜁 2 𝜔𝑛 𝑡)]. (2.227)
√1−𝜁 2

In the following we shall only be concerned with the transient response of


lightly damped systems. In practice, for many structures, 𝜁 = 0.01 … . .0.04, hence
the damped natural frequency 𝜔𝑑 = √1 − 𝜁 2 𝜔𝑛 almost coincides with the
undamped natural frequency 𝜔𝑛 , and √1 − 𝜁 2 ≈ 1.
In this case we obtain
𝑓̂
𝑥(𝑡) = 2𝜁𝑘 (1 − 𝑒 −𝜁𝜔𝑛 𝑡 ) 𝑠𝑖𝑛 𝜔𝑛 𝑡. (2.228)

a graphical representation of which is shown in Fig.2.93.

The curve has two envelopes


𝑓̂
𝑦 = ± 2𝜁𝑘 (1 − 𝑒 −𝜁𝜔𝑛 𝑡 ) (2.229)

which tend asymptotically towards the amplitude of the steady-state vibrations at


the phase resonance
𝑓̂
𝑥̂∞ = . (2.230)
2𝜁𝑘
Taking only positive values of 𝑦, we can write
𝑥̂∞ − 𝑦 = 𝑥̂∞ 𝑒 −𝜁𝜔𝑛 𝑡
or
𝑙𝑛(𝑥̂∞ − 𝑦) = 𝑙𝑛 𝑥̂∞ − 𝜁𝜔𝑛 𝑡. (2.231)
The function 𝑙𝑛(𝑥̂∞ − 𝑦) is expressed as a function of time by a straight
line of slope 𝜁𝜔𝑛 . Hence, the value of ζ is obtained knowing 𝜔𝑛 [46].
A simpler method is based on the construction from Fig.2.94. The tangent
in the origin at the envelope intersects the asymptote at the point 𝐾 of abscissa

88
1
𝑡𝑘 = 𝜁𝜔 . (2.232)
𝑛
2𝜋
The time constant 𝑡𝑘 can be expressed as a function of the period 𝑇 = 𝜔
𝑛
by equation 𝑡𝑘 = 𝑛𝑇. This yields the damping ratio
1
𝜁 = 2𝜋𝑛 (2.233)
then, if the record is long enough, the ordinate of the asymptote yields the stiffness.
A different method is based on the evaluation of the hatched area
∞ 𝑥̂∞
𝑥̂∞ ∫0 𝑒 −𝜁𝜔𝑛 𝑡 𝑑𝑡 =
𝜁𝜔𝑛

wherefrom we obtain 𝜁.
The velocity is the derivative of the function (2.228)
𝑑𝑥 𝑓̂
𝑑𝑡
= 𝜔𝑛 2𝜁𝑘 [(1 − 𝑒 −𝜁𝜔𝑛 𝑡 )𝑐𝑜𝑠 𝜔𝑛 𝑡 + 𝜁𝑒 −𝜁𝜔𝑛 𝑡 𝑠𝑖𝑛𝜔𝑛 𝑡] . (2.234)

Neglecting the second term which contains 𝜁 we obtain


𝑓̂𝜔𝑛
𝑥̇ (𝑡) ≅ 2𝜁𝑘
(1 − 𝑒 −𝜁𝜔𝑛 𝑡 ) cos 𝜔𝑛 𝑡. (2.235)

It is possible to use the construction from Fig.2.94 on a time record of the


𝜔𝑛 𝑓̂ 𝑓̂
velocity [6]. In this case the asymptote has an ordinate 2𝜁𝑘
=𝑐.

2.6.3.2. Decaying transient

Consider now the transient motion of the system from Fig.2.7, which
initially performs forced stationary vibrations of frequency 𝜔𝑛 and amplitude
(2.230) , then the force is suddenly removed .
The general solution of the damped free vibrations is given by the first term
in the right hand of expression (2.225).
If at 𝑡 = 0 , 𝑥 = 𝑥̂𝑜 and 𝑥̇ = 0,
𝜁
𝑥(𝑡) = 𝑥̂𝑜 𝑒 −𝜁𝜔𝑛 𝑡 [𝑐𝑜𝑠(√1 − 𝜁 2 𝜔𝑛 𝑡) + 𝑠𝑖𝑛(√1 − 𝜁 2 𝜔𝑛 𝑡)]. (2.236)
√1−𝜁 2

If, for light damping, 𝜁 2 is neglected with respect to 1, then


𝑥(𝑡) ≅ 𝑥̂𝑜 𝑒 −𝜁𝜔𝑛 𝑡 (𝑐𝑜𝑠𝜔𝑛 𝑡 + 𝜁 𝑠𝑖𝑛𝜔𝑛 𝑡), (2.237)
𝑥̇ (𝑡) ≅ 𝜔𝑛 𝑥̂𝑜 𝑒 −𝜁𝜔𝑛 𝑡 𝑠𝑖𝑛𝜔𝑛 𝑡. (2.238)
It can be considered that the envelopes (Fig.2.95)
𝑦1 = 𝑥̂𝑜 𝑒 −𝜁𝜔𝑛 𝑡 , (2.239)
𝑦2 = −𝑥̂𝑜 𝑒 −𝜁𝜔𝑛 𝑡 (2.240)

89
are tangent to the curve (2.237) at the points of zero velocity (or maximum
displacement), hence at 𝜔𝑛 𝑡 = 𝑗𝜋, where 𝑗 is an integer.
Expression (2.239) can be written
𝑙𝑛 𝑦1 = 𝑙𝑛 𝑥̂𝑜 − 𝜁𝜔𝑛 𝑡 (2.241)
which plotted in semilogarithmic scales (Fig.2.96) is a straight line of slope
(−𝜁𝜔𝑛 ). Knowing 𝜔𝑛 the value of 𝜁 can be obtained [79].
An analogous method can be developed based on expression (2.238) if we
measure the velocity of the response [6].

The tangent to the curve (2.239) at the point 𝐿 intersects the time axis
1
(Fig.2.97) at the point 𝐾, of abscissa 𝑡𝑘 = which can be expressed as a function
𝜁𝜔𝑛
1
of the period of vibration 𝑇, 𝑡𝑘 = 𝑛𝑇, hence 𝜁 = .
2𝜋𝑛
If the waveform of the record of function (2.236) is clear enough so as to
enable the measurement of two successive amplitudes of the same sign, 𝑥̂1 and 𝑥̂2 ,
it is possible to calculate the logarithmic decrement
𝑥̂1 2𝜋𝜁
𝛿 = 𝑙𝑛 = 𝜁𝜔𝑛 𝑇 = ≅ 2𝜋𝜁 . (2.242)
𝑥̂2 √1−𝜁 2

Measuring 𝑥̂𝑜 , then the amplitude 𝑥̂𝑛 after 𝑛 cycles of vibration, the
logarithmic decrement is
1 𝑥̂
𝛿 = 𝑛 𝑙𝑛 𝑥̂ 1 . (2.243)
𝑛

The damping ratio is given by


𝛿
𝜁 = √𝛿2 (2.244)
+4𝜋 2

90
Equation (2.241) has a more general interpretation, based on energy
considerations [144].
Let 𝑥̂𝑛 and 𝑥̂𝑛+1 be two successive amplitudes with the same sign, of the
free vibration. During the time elapsed between them, the approximate variation of
the kinetic energy is
1 𝑥̂𝑛 +𝑥̂𝑛+1 2
𝑊𝑐 = 𝜔2 𝑚 ( ) . (2.245)
2 2

The dissipated energy is equal to the variation of the kinetic energy


1 2 ).
𝑊𝑑 = 2 𝜔2 𝑚(𝑥̂𝑛2 − 𝑥̂𝑛+1 (2.246)
Equation (2.222) can then be written
1 2 2 ) 1 𝑥̂𝑛 +𝑥̂𝑛+1 2
𝜔 𝑚(𝑥̂𝑛2 − 𝑥̂𝑛+1 = 4𝜋𝜁 ( 𝜔2 𝑚) ( )
2 2 2

from which we obtain


1−𝜋𝜁
𝑥̂𝑛+1 = 𝑥̂𝑛 1+𝜋𝜁 (2.247)
and analogously
1−𝜋𝜁 𝑛
𝑥̂𝑛 = 𝑥̂𝑜 (1+𝜋𝜁 ) . (2.248)

Using logarithms
1−𝜋𝜁
𝑙𝑜𝑔𝑥̂𝑛 = 𝑙𝑜𝑔𝑥̂𝑜 + 𝑛 𝑙𝑜𝑔 1+𝜋𝜁 . (2.249)

Expanding in series, for small values of 𝜁, we can write


1−𝜋𝜁
𝑙𝑜𝑔 ≅ −2𝜋𝜁 .
1+𝜋𝜁

With a fair degree of approximation, equation (2.249) becomes


𝑙𝑜𝑔𝑥̂𝑛 = 𝑙𝑜𝑔𝑥̂𝑜 − 2𝜋𝑛 𝜁 (2.250)
which is identical to equation (2.241), and can be represented graphically by a
straight line even when the damping is not exclusively viscous.
In this case 𝜁 is an equivalent viscous damping ratio, as defined in § 2.1.4.

2.6.4. Method of Self-Excited Vibrations

When the system from Fig.2.34 is excited using an electromagnetic


vibrator, it is possible to use a reaction signal feeded in the vibrator coil through a
power amplifier. The signal is obtained from a velocity pickup attached to the
system at the driving point. If the electronic equipment which controls the vibrator
is free of phase distortions, it is possible to produce selfexcited vibrations at the
frequency 𝜔𝑛 by a proper adjustment of the amplifier gain.

91
Let insert in equation (2.68) the force delivered by the electromagnetic
shaker
𝑓(𝑡) = 𝛤𝑣 𝑖, (2.251)
where 𝑖 is the current through the coil and 𝛤𝑣 is the force-current constant.
The voltage across the excitation coil is proportional to the velocity at the
driving point
𝑒 = 𝐶𝑥̇ . (2.252)
The electrical circuit equation is
𝑒 = 𝑖 𝑍𝑣 + 𝛤𝑐 𝑥̇ (2.253)
where 𝑍𝑣 is the coil impedance and 𝛤𝑐 𝑥̇ is the back electromotive force induced in
the coil.
Equations (2.252) and (2.253) can be combined to give the current
(𝐶−𝛤𝑐 )𝑥̇
𝑖= 𝑍𝑣

which inserted in equation (2.251) yields


𝐶−𝛤𝑐
𝑓(𝑡) = 𝛤𝑣 𝑍𝑣
𝑥̇ = (𝐶ˈ − 𝐶")𝑥̇ (2.254)

where the constants 𝐶ˈ and 𝐶" are real if 𝑍𝑣 is real.


Substituting expression (2.254) in the equation of motion (2.68) we obtain
𝑚𝑥̈ + (𝑐 + 𝐶" − 𝐶ˈ)𝑥̇ + 𝑘𝑥 = 0. (2.255)
The constant 𝐶ˈ is proportional to the amplifier gain. If
𝐶ˈ = 𝑐 + 𝐶" (2.256)
the system damping is cancelled. The system performs undamped vibrations with
the frequency 𝜔𝑛 .
The equivalent viscous damping coefficient is obtained from (2.256)
𝑐 = 𝐶ˈ − 𝐶",
where 𝐶ˈ and 𝐶" depend on the arrangement circuit and the amplifier setting.
The practical application of this method is difficult due to the instability of
motion when 𝐶ˈ becomes incidentally larger than 𝑐 + 𝐶". In order to stabilize the
motion amplitude, a slightly nonlinear electronic component is introduced in the
amplifier, which makes the condition (2.256) realizable for a given value of the
speed 𝑥̇ . When 𝑥̇ decreases below this value, 𝐶ˈ increases and vice versa.

92
CHAPTER 3

Elements of the Theory of Vibration of Linear Lumped


Parameter Systems

This chapter is a succinct presentation of the main results of the theory of


vibration of multi-degree-of-freedom systems obtained by modal analysis and
spectral analysis. Only systems with finite number of degrees of freedom are treated
because most of the identification methods of elastic structures are based on
discrete models of the real distributed parameter structures.

3.1. VIBRATION OF LINEAR UNDAMPED SYSTEMS

The study of conservative systems is of interest for the analysis of


structures with negligible damping (car chassis, frames, machine tool structures,
etc.) as well as for structures with “proportional” damping, whose response can be
expressed in terms of the natural modes of vibration of the associate undamped
system.

3.1.1. Free Vibrations

The equations of free motion for an 𝑁-degree of freedom undamped


system, expressed in terms of the independent coordinates 𝑞1 , 𝑞2 , . . . . , 𝑞𝑁 , can be
written as
[𝑀] {𝑞̈ } + [𝐾] {𝑞} = {0} . (3.1)
The mass matrix [𝑀] and the stiffness matrix [𝐾] are positive definite,
nonsingular and usually symmetric (when the coordinates measure the
displacement from a fixed position in space) and {𝑞} is the column vector of
generalized coordinates.
We seek whether there exists a solution of the form
{𝑞} = {𝛹} 𝑐𝑜𝑠 𝜔𝑡 , (3.2)
which represents a harmonic motion in which the displacements corresponding to
the 𝑁 coordinates are synchronous and in-phase.
Substituting the solution (3.2) into equation (3.1), we get a set of
homogeneous linear equations
([𝐾] − 𝜔2 [𝑀]) {𝛹} = {0} . (3.3)
Therefore we have to solve the generalized eigenvalue problem associated
with the 𝑁𝑥𝑁 square matrices [𝑀] and [𝐾].

93
Equations (3.3) have nontrivial solutions only if the determinant of the
coefficient matrix is zero
𝑑𝑒𝑡 ([𝐾] − 𝜔2 [𝑀]) = 0 . (3.4)
2
This is an algebraic equation of degree 𝑁 in 𝜔 , referred to as the frequency
equation. Its roots 𝜔12 , 𝜔22 , . . . , 𝜔𝑁
2
(considered distinct) called eigenvalues are all
real and positive. The quantities 𝜔1 , 𝜔2 , . . . , 𝜔𝑁 are termed undamped natural
frequencies.
Associated with each eigenvalue 𝜔𝑟2 there is an eigenvector {𝛹 (𝑟) } with
(𝑟)
real elements 𝛹𝑗 which satisfies the equation
([𝐾] − 𝜔𝑟2 [𝑀]) {𝛹 (𝑟) } = {0} (3.5)
and defines the shape of a undamped principal mode of vibration.
(𝑟)
These modal vectors are unique, the ratio of any two elements 𝛹𝑗 and
(𝑟)
𝛹ℓ being constant (while their individual values are arbitrary). The process of
“adjustment” of the elements of eigenvectors to render their amplitude unique is
called normalization. The resulting vectors define the shape of the undamped
(𝑟)
normal modes of vibration. This can be accomplished taking either 𝛹𝑚𝑎𝑥 = 1 or
𝑇
{𝛹 (𝑟) } [𝑀]{𝛹 (𝑟) } = 1 , where 𝑇 denotes transposition. Other normalizations of
the modal vectors are presented in Chapter 5.
The eigenvalue problem (3.5) can be written in the standard form
[𝑀]−1 [𝐾]{𝛹 (𝑟) } = 𝜔𝑟2 {𝛹 (𝑟) } , (3.5,a)
1
[𝐾]−1 [𝑀]{𝛹 (𝑟) } = {𝛹 (𝑟) } . (3.5,b)
𝜔𝑟2

The right eigenvectors of the matrices [𝑀]−1 [𝐾] and [𝐾]−1 [𝑀] are
identical, while the corresponding eigenvalues are inverse to each other (since [𝑀]
and [𝐾] are symmetric).
The eigenvectors {𝛹 (𝑟) } satisfy the following orthogonality conditions
𝑇
{𝛹 (𝑠) } [𝑀] {𝛹 (𝑟) } = 0 , (𝑟 ≠ 𝑠) (3.6)
𝑇
{𝛹 (𝑠) } [𝐾] {𝛹 (𝑟) } = 0 , (𝑟 ≠ 𝑠) (3.7)
in which [𝑀] and [𝐾] are weighting matrices.
Denoting
𝑇
𝑚𝑟 = {𝛹 (𝑟) } [𝑀]{𝛹 (𝑟) } , (3.8)
𝑇
𝑘𝑟 = {𝛹 (𝑟) } [𝐾]{𝛹 (𝑟) } , (3.9)
equation (3.5) becomes
𝑘𝑟 − 𝜔𝑟2 𝑚𝑟 = 0 . (3.10)
The square of a natural frequency is expressed as the ratio

94
𝑇
𝑘𝑟 {𝛹 (𝑟) } [𝐾]{𝛹 (𝑟) }
𝜔𝑟2 = = 𝑇 , (3.11)
𝑚𝑟 {𝛹 (𝑟) } [𝑀]{𝛹 (𝑟) }
where 𝑘𝑟 is a generalized stiffness (modal stiffness) and 𝑚𝑟 is a generalized mass
(modal mass).
Let introduce the 𝑁𝑥𝑁 matrices of eigenvectors and eigenvalues. The
modal matrix [𝛹] has the eigenvectors as columns
[𝛹] = [ {𝛹 (1) } {𝛹 (2) } . . . . {𝛹 (𝑟) } . . . {𝛹 (𝑁) }] (3.12)
and the diagonal spectral matrix has the eigenvalues along the main diagonal
[ˋ𝜔𝑟2 ˏ] = 𝑑𝑖𝑎𝑔 [𝜔𝑟2 ] . (3.13)
The equations (3.5) become
[𝐾] [𝛹] − [𝑀] [𝛹] [ˋ𝜔𝑟2 ˏ] = [0]. (3.14)

3.1.2. Forced Vibrations

The equations for the forced motion of an 𝑁-degree of freedom undamped


system, can be written as
[𝑀] {𝑞̈ } + [𝐾] {𝑞} = {𝑓} , (3.15)
where {𝑓} is the forcing vector.
In order to uncouple equations (3.15), in which matrices [𝑀] and [𝐾] are
not simultaneously diagonal, one can use the linear coordinate transformation
{𝑞} = [𝛹] {𝑝} (3.16)
which can also be written as
{𝑞} = ∑𝑁 (𝑟)
𝑟=1{𝛹 } 𝑝𝑟 . (3.16,a)
Equation (3.16,a) introduces the time-dependent undamped principal
coordinates 𝑝𝑟 and develops the vector {𝑞} as a linear combination of the
eigenvectors {𝛹 (𝑟) } .
Inserting (3.16) into (3.15) and premultiplying both sides by [𝛹]𝑇 we
obtain
[𝛹]𝑇 [𝑀][𝛹]{𝑝̈ } + [𝛹]𝑇 [𝐾][𝛹] {𝑝} = [𝛹]𝑇 {𝑓} . (3.17)
The 𝑟-th equation is obtained substituting (3.16,a) into (3.15) and using the
orthogonality conditions (3.6) and (3.7) :
𝑇 𝑇 𝑇
{𝛹 (𝑟) } [𝑀]{𝛹 (𝑟) }𝑝̈𝑟 + {𝛹 (𝑟) } [𝐾]{𝛹 (𝑟) } 𝑝𝑟 = {𝛹 (𝑟) } {𝑓} . (3.18)
In view of (3.8) and (3.9), equation (3.18) can be rewritten
𝑚𝑟 𝑝̈𝑟 + 𝑘𝑟 𝑝𝑟 = 𝐹𝑟 , (𝑟 = 1, 2, . . , 𝑁) (3.19)

95
where 𝐹𝑟 is the generalized force (modal force) corresponding to the r-𝑡ℎ mode
𝑇
𝐹𝑟 = {𝛹 (𝑟) } {𝑓} . (3.20)
Introducing the diagonal matrix of modal masses
[ˋ𝑚ˏ] = [𝛹]𝑇 [𝑀] [𝛹] = 𝑑𝑖𝑎𝑔[𝑚𝑟 ], (3.21)
the diagonal matrix of modal stiffnesses
[ˋ𝑘ˏ] = [𝛹]𝑇 [𝐾] [𝛹] = 𝑑𝑖𝑎𝑔[𝑘𝑟 ], (3.22)
and the column vector of modal forces
{𝐹} = [𝛹]𝑇 {𝑓} (3.23)
equations (3.17) can be rewritten in the form
[ˋ𝑚ˏ] {𝑝̈ } + [ˋ𝑘ˏ] {𝑝} = {𝐹} . (3.17,a)
This is a set of 𝑁 uncoupled equations (3.19), which can be solved
individually like the equation of motion of the undamped single-degree-of-freedom
system.
For harmonic excitation {𝑓} = {𝑓̂} 𝑒 𝑖𝜔𝑡 , the steady-state solution has the
form {𝑞} = {𝑞̂} 𝑒 𝑖𝜔𝑡 , and the transformation (3.16) becomes
{𝑞̂} = [𝛹] {𝑝̂ } = ∑𝑁 (𝑟)
𝑟=1{𝛹 } 𝑝̂𝑟 . (3.16,b)
The principal coordinate is
(𝑟)𝑇
{𝛹 ̂}
} {𝑓
𝑝̂𝑟 = 2
𝑘𝑟 −𝜔 𝑚𝑟
which, substituted back into (3.16,b), gives the vector of displacement amplitudes
𝑟 𝑇
̂ } {𝛹 𝑟 }
{𝛹 } {𝑓
( ) ( )

{𝑞̂} = ∑𝑁
𝑟=1 𝑚 (𝜔2 −𝜔2 ) . (3.24)
𝑟 𝑟

Although the natural modes of vibration {𝛹 (𝑟) } of an undamped system


exist in the absence of any external forcing, it is possible to attach to each of them
a natural mode of excitation, referred to as a principal mode of excitation or
excitation modal vector.
If the force {𝑓} = {ℱ̂ (𝑟) }𝑒 𝑖𝜔𝑡 excites the response {𝑞} = {𝛹 (𝑟) }𝑒 𝑖𝜔𝑡 , from
(3.16,a) it follows that {𝑝} = {𝐼}𝑟 𝑒 𝑖𝜔𝑡 , where {𝐼}𝑟 is the 𝑟-th column of the
identity matrix [ 𝐼 ], and equation (3.15) becomes
{ℱ̂ (𝑟) } = [𝐾 − 𝜔2 𝑀] {𝛹 (𝑟) } (3.25)
which defines the 𝑟-th principal mode of excitation, as the forcing required if the
system is to vibrate in the 𝑟-th undamped mode at a frequency 𝜔 .

96
𝑇
Premultiplying in (3.25) by {𝛹 (𝑠) } , and using (3.6) and (3.7), we obtain
𝑇
{𝛹 (𝑠) } {ℱ̂ (𝑟) } = 0,
hence the work done by the forces of an excitation modal vector on the
displacements of other modes of vibration is zero.
Equations (3.8) and (3.9) yield
𝑇 𝜔 2
{𝛹 (𝑟) } {ℱ̂ (𝑟) } = 𝑘𝑟 − 𝜔2 𝑚𝑟 = 𝑘𝑟 (1 − 𝜔2)
𝑟

which for 𝜔 ≠ 𝜔𝑟 is different from zero.

3.2. VIBRATIONS OF DAMPED SYSTEMS

3.2.1. Assumptions on Damping

The equation governing the forced vibrations of a system with linear


viscous damping is
[𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} , (3.26)
where [𝐶] is the matrix of the coefficients of viscous damping, referred to as the
damping matrix. It is considered to be real, symmetric and positive definite.
The coordinate transformation (3.16) and premultiplication by [𝛹]𝑇 yields
[𝛹]𝑇 [𝑀][𝛹]{𝑝̈ } + [𝛹]𝑇 [𝐶][𝛹] {𝑝̇ } + [𝛹]𝑇 [𝐾][𝛹] {𝑝} = [𝛹]𝑇 {𝑓} . (3.27)
Using notations (3.21) to (3.23) and
[𝑐] = [𝛹]𝑇 [𝐶][𝛹] (3.28)
equation (3.27) becomes
[ˋ𝑚ˏ] {𝑝̈ } + [𝑐] {𝑝̇ } + [ˋ𝑘ˏ] {𝑝} = {𝐹} . (3.29)
If the matrix [𝑐] is nondiagonal, the viscous damping is referred to as
nonproportional (nonclassical).
The first differential equation in (3.29) can be written
𝑚1 𝑝̈1 + 𝑐11 𝑝̇1 + ∑𝑁𝑟=2 𝑐1𝑟 𝑝̇𝑟 + 𝑘1 𝑝1 = 𝐹1 . (3.30)
The third term on the left hand side is due to the coupling of the modes of
vibration by damping. In this case, there are complex natural frequencies and
complex mode shapes, as shown in § 3.2.4.
When the matrix [𝑐] is diagonal, the viscous damping is called
proportional (or classical).
In this case, denoting
[ˋ𝑐ˏ] = 𝑑𝑖𝑎𝑔 [𝑐𝑟 ] , (3.31)
the equations (3.29) become

97
[ˋ𝑚ˏ] {𝑝̈ } + [ˋ𝑐ˏ] {𝑝̇ } + [ˋ𝑘ˏ] {𝑝} = {𝐹} , (3.32)
therefore the modes of vibration are uncoupled; the first equation (3.32) is
𝑚1 𝑝̈1 + 𝑐1 𝑝̇1 + 𝑘1 𝑝1 = 𝐹1 , (3.33)
and can be solved independently of the others [5].
The name proportional damping comes from an observation of Rayleigh
[100] which, in matrix form, can be expressed as follows: If
[𝐶] = 𝜎 [𝑀] + 𝜏 [𝐾] , (3.34)
where σ and τ are proportionality constants, then the modal damping matrix
[𝛹]𝑇 [𝐶][𝛹] is diagonal. Indeed
[𝛹]𝑇 [𝐶][𝛹] = 𝜎[𝛹]𝑇 [𝑀][𝛹] + 𝜏[𝛹]𝑇 [𝐾][𝛹] = 𝜎[ˋ𝑚ˏ] + 𝜏[ˋ𝑘ˏ] = [ˋ𝑐ˏ].
Proportional damping should not be limited to a linear combination of the
mass and stiffness matrices. It was shown [21], [22], [37] that if the damping matrix
can be expressed as a polynomial function of [𝑀] and/or [𝐾], then the modal matrix
[𝛹] diagonalizes the matrix [𝐶] like the matrices of [𝑀] and [𝐾], provided it is
orthogonal [𝛹]𝑇 [𝛹] = [𝐼].
The (necessary and sufficient) condition for uncoupling the equations
(3.26) by the transformation (3.16) is [22] that [𝑀]−1 [𝐶] commutes with [𝑀]−1 [𝐾]
[𝐶] [𝑀]−1 [𝐾] = [𝐾] [𝑀]−1 [𝐶] .
In practice, the use of proportional damping is not based on the fulfilment
of such a complicated condition, but on simply neglecting the off-diagonal elements
of the modal damping matrix, i.e. neglecting the modal couplings due to damping.

3.2.2. Real “Classical” Modes of Vibration

When the damping matrix is diagonalized by the same coordinate


transformation which uncouples the equations (3.1), the response of the damped
system can be expressed in terms of the modes of vibration {𝛹 (𝑟) } of the associated
undamped system, referred to as the real “classical” modes of vibration.

3.2.2.1. Viscous damping

In the case of proportional viscous damping, the following orthogonality


condition can be established
𝑇
{𝛹 (𝑠) } [𝐶] {𝛹 (𝑟) } = 0 . (𝑟 ≠ 𝑠) (3.35)
The modal viscous damping coefficients are defined by
𝑇
𝑐𝑟 = {𝛹 (𝑟) } [𝐶] {𝛹 (𝑟) } , (3.36)
For harmonic excitation and steady-state response

98
{𝑓} = {𝑓̂}𝑒 𝑖𝜔𝑡 , {𝑞} = {𝑞̃}𝑒 𝑖𝜔𝑡 , (3.37)
equation (3.26) becomes
(−𝜔2 [𝑀] + 𝑖𝜔[𝐶] + [𝐾]) {𝑞̃} = {𝑓̂} . (3.38)
The coordinate transformation
{𝑞̃} = ∑𝑁 (𝑟)
𝑟=1{𝛹 } 𝑝̃𝑟 (3.39)
and use of equations (3.6) to (3.10), (3.35) and (3.36) yields the complex principal
coordinates
(𝑟) 𝑇
{𝛹 ̂}
} {𝑓
𝑝̃𝑟 =𝑚 2 2 (3.40)
𝑟 (−𝜔 +𝑖2𝜁𝑟 𝜔𝜔𝑟 +𝜔𝑟 )
where
𝑐𝑟
𝜁𝑟 = (3.41)
2𝑚𝑟 𝜔𝑟
is the modal damping ratio. Substituting (3.40) into (3.39) we obtain the column
vector of complex displacement amplitudes
𝑟 𝑇
{𝛹 } {𝑓̂ } {𝛹 𝑟 }
( ) ( )

{𝑞̃} = ∑𝑁
𝑟=1 𝑚 (−𝜔2 +𝑖 2𝜁 𝜔𝜔 +𝜔2 ) . (3.42)
𝑟 𝑟 𝑟 𝑟

3.2.2.2. Hysteretic damping

The equations of forced vibrations of a linear hysteretically damped system


are
1
[𝑀] {𝑞̈ } + [𝐻] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} , (3.43)
𝜔

where [𝐻] is the (real, symmetric and positive definite) matrix of the coefficients
of hysteretic damping.
For proportional hysteretic damping, the following orthogonality condition
can be established
𝑇
{𝛹 (𝑠) } [𝐻] {𝛹 (𝑟) } = 0 . (𝑟 ≠ 𝑠) (3.44)
The modal hysteretic damping coefficients are defined by
𝑇
ℎ𝑟 = {𝛹 (𝑟) } [𝐻] {𝛹 (𝑟) } . (3.45)
For a harmonic excitation
{𝑓} = {𝑓̂} 𝑒 𝑖𝜔𝑡 , (3.46)
the steady-state displacement is
{𝑞} = {𝑞̃} 𝑒 𝑖𝜔𝑡 , (3.47)

99
hence equations (3.43) become
(−𝜔2 [𝑀] + 𝑖 [𝐻] + [𝐾]) {𝑞̃} = {𝑓̂} . (3.48)
The coordinate transformation (3.39)
{𝑞̃} = [𝛹] {𝑝̃} = ∑𝑁 (𝑟)
𝑟=1{𝛹 } 𝑝̃𝑟
uncouples equations (3.48), which become
(−𝜔2 [ˋ𝑚ˏ] + 𝑖[ˋℎˏ] + [ˋ𝑘ˏ]) {𝑝̃} = [𝛹]𝑇 {𝑓̂} = {𝐹̂ } (3.49)
where
[ˋℎˏ] = 𝑑𝑖𝑎𝑔 [ℎ𝑟 ] . (3.50)
The 𝑟-th equation (3.49) can be written
(𝑘𝑟 − 𝜔2 𝑚𝑟 + 𝑖 ℎ𝑟 )𝑝̃𝑟 = 𝐹̂𝑟 (3.51)
hence the complex modal coordinate 𝑝̃𝑟 is
(𝑟) 𝑇
{𝛹 ̂}
} {𝑓
𝑝̃𝑟 =𝑘 2 . (3.52)
𝑟 −𝜔 𝑚𝑟 +𝑖 ℎ𝑟

This yields the vector of complex displacement amplitudes


(𝑟) 𝑇 (𝑟)
{𝛹 } {𝑓} {𝛹 } ̂
{𝑞̃} = ∑𝑁
𝑟=1 2 (3.53)
𝑘𝑟 (1−𝜔2+𝑖𝑔𝑟 )
𝜔𝑟

in which
𝑔𝑟 = ℎ𝑘𝑟 (𝑟 = 1, 2, … , 𝑁) (3.54)
𝑟
are the hysteretic damping factors.
For the identification methods using single point excitation, of interest is
the response at 𝑞𝑗 produced by a harmonic force applied at 𝑞ℓ . From expression
(3.53) we obtain
(𝑟) (𝑟)
𝜕𝑞̃ ℓ ̂ 𝜕𝑞̃ 𝑗
𝛹ℓ 𝑓̂ ℓ 𝛹𝑗 𝑓 ̃
𝜕𝑝̃
𝑟 ℓ 𝜕𝑝
𝑞̃𝑗 = ∑𝑁
𝑟=1 2 = ∑𝑁
𝑟=1 2
𝑟
. (3.55)
𝑘𝑟 (1−𝜔2+𝑖𝑔𝑟 ) 𝑘𝑟 (1−𝜔2+𝑖𝑔𝑟)
𝜔𝑟 𝜔𝑟
The complex (cross or transfer) receptance is
(𝑟) (𝑟) (𝑟) (𝑟)
𝑞̃𝑗 𝛹ℓ 𝛹𝑗 𝛹ℓ 𝛹𝑗
𝛼̅𝑗ℓ = = ∑𝑁
𝑟=1 𝜔2
= ∑𝑁
𝑟=1 𝑘 𝑔 𝑠𝑖𝑛𝜓𝑟 𝑒−𝑖𝜓𝑟 (3.56)
𝑓̂ℓ 𝑘𝑟 (1− 2+𝑖𝑔𝑟 ) 𝑟 𝑟
𝜔𝑟
where

100
𝑔𝑟
𝜓𝑟 = 𝑡𝑎𝑛−1 𝜔2
. (3.57)
1− 2
𝜔 𝑟

The same result is obtained using the frequency response method. Equation
(3.48) can be written
[𝛽̅ ] {𝑞̃} = {𝑓̂} (3.48,a)
where
[𝛽̅ ] = [−𝜔2 [𝑀] + 𝑖 [𝐻] + [𝐾]] (3.58)
is the matrix of complex dynamic stiffnesses.
The matrix of complex receptances (dynamic influence coefficients) is
−1 −1
[𝛼̅] = [𝛽̅ ] = [−𝜔2 [𝑀] + 𝑖 [𝐻] + [𝐾]] (3.59)
with elements of the form
𝜕𝑞
̃
𝑗
𝛼̅𝑗ℓ =
𝜕𝑓̂ ℓ
. (3.60)

Using the coordinate transformation (3.39), equation (3.48,a) becomes


[𝛽̅ ] [𝛹] {𝑝̃} = {𝑓̂} . (3.61)
Premultiplying by [𝛹]𝑇 we obtain the uncoupled equations
[ˋ𝛽𝑟̅ ˏ] {𝑝̃} = [𝛹]𝑇 {𝑓̂} , (3.62)
in which
[ˋ𝛽𝑟̅ ˏ] = [𝛹]𝑇 [𝛽̅ ] [𝛹] = −𝜔2 [ˋ𝑚ˏ] + 𝑖[ˋℎˏ] + [ˋ𝑘ˏ] = 𝑑𝑖𝑎𝑔 [𝛽𝑟̅ ] (3.63)
and the complex modal dynamic stiffness
𝛽𝑟̅ = 𝑘𝑟 − 𝜔2 𝑚𝑟 + 𝑖 ℎ𝑟 . (3.64)
The complex modal receptance is defined as
1 1 𝑠𝑖𝑛 𝜓𝑟 𝑒−𝑖𝜓𝑟
𝛼̅𝑟 = ̅
𝛽
= 2 = ℎ𝑟
(3.65)
𝑟 𝑘𝑟 (1−𝜔2+𝑖𝑔𝑟 )
𝜔𝑟

and the corresponding diagonal matrix is


−1
[ˋ𝛼̅𝑟 ˏ] = 𝑑𝑖𝑎𝑔 [𝛼̅𝑟 ] = [ˋ𝛽𝑟̅ ˏ] . (3.66)
Inverting (3.63) we get
−1 −1
[ˋ𝛽𝑟̅ ˏ] = [𝛹]−1 [𝛽̅] [𝛹]−𝑇 (3.67)
or, using notations (3.59) and (3.66),
[ˋ𝛼̅𝑟 ˏ] = [𝛹]−1 [𝛼̅] [𝛹]−𝑇 . (3.68)

101
Premultiplying by [𝛹] and postmultiplying by [𝛹]𝑇 equation (3.68)
becomes
[𝛹] [ˋ𝛼̅𝑟 ˏ] [𝛹]𝑇 = [𝛼̅] . (3.69)
The ℓ-th column of the matrix [𝛼̅] is the response vector to a force applied
at coordinate ℓ
{𝛼̅}ℓ = [𝛹] [ˋ𝛼̅𝑟 ˏ] {𝛹}ℓ = [𝛹] {𝛼̅𝑟 𝛹ℓ } , (3.70)
where {𝛹}ℓ is a vector containing the ℓ-th element of each natural mode
(𝑟) (𝑟)
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 𝛹ℓ 𝛹𝑗 𝛼̅𝑟 (3.71)
which is identical to (3.56).
In order to simplify the presentation of the identification methods exposed
in Chapter 5, it is useful to introduce the following notations
2
1−𝜔2
𝜔𝑟 −𝑔𝑟
𝑎𝑟 (𝜔) = 2 , 𝑏𝑟 (𝜔) = 2 , (3.72)
𝜔2 𝜔2
(1− 2) +𝑔2
𝑟 (1− 2) +𝑔2
𝑟
𝜔𝑟 𝜔𝑟

𝐴𝑟 (𝜔) = 𝑔𝑟 𝑎𝑟 (𝜔) , 𝐵𝑟 (𝜔) = 𝑔𝑟 𝑏𝑟 (𝜔) , (3.73)


(𝑟) (𝑟)
(𝑟) 𝛹ℓ 𝛹𝑗
𝜒𝑗ℓ = 𝑘 𝑔 , (3.74)
𝑟 𝑟
(𝑟) (𝑟)
(𝑟) 𝛹ℓ 𝛹𝑗
𝜘𝑗ℓ = . (3.75)
𝑘𝑟
Separating the real and imaginary parts of the complex receptances,
expression (3.56) can be written
(𝑟) (𝑟)
𝛹ℓ 𝛹𝑗
𝛼̅𝑗ℓ (𝜔) = 𝛼𝑗ℓ𝑅 (𝜔) + 𝑖 𝛼𝑗ℓ𝐼 (𝜔) = ∑𝑁
𝑟=1 [𝑎𝑟 (𝜔) + 𝑖 𝑏𝑟 (𝜔)] (3.76)
𝑘𝑟

therefore
𝑟 ( ) 𝑟 ( )
𝛼𝑗ℓ𝑅 (𝜔) = ∑𝑁 𝑁
𝑟=1 𝜘𝑗ℓ 𝑎𝑟 (𝜔) = ∑𝑟=1 𝜒𝑗ℓ 𝐴𝑟 (𝜔) , (3.77)
(𝑟) (𝑟)
𝛼𝑗ℓ𝐼 (𝜔) = ∑𝑁 𝑁
𝑟=1 𝜘𝑗ℓ 𝑏𝑟 (𝜔) = ∑𝑟=1 𝜒𝑗ℓ 𝐵𝑟 (𝜔) . (3.78)
We introduce the column vectors
(𝑟)
{𝛼ℓ𝑅 (𝜔)} = ∑𝑁
𝑟=1 {𝜘ℓ } 𝑎𝑟 (𝜔) = [𝜘] {𝑎(𝜔)} = [𝜒] {𝐴(𝜔)} , (3.79)
(𝑟)
{𝛼ℓ𝐼 (𝜔)} = ∑𝑁
𝑟=1 {𝜘ℓ } 𝑏𝑟 (𝜔) = [𝜘] {𝑏(𝜔)} = [𝜒] {𝐵(𝜔)} , (3.80)
in which

102
(1) (2) (𝑁)
𝛹ℓ 𝛹ℓ 𝛹ℓ
[𝜘] = [{𝜘ℓ(1) } {𝜘ℓ(2) } . . . . {𝜘ℓ(𝑁) } ] = [{𝛹 (1) } {𝛹 (2) } . . . . {𝛹 (𝑁) } ],
𝑘1 𝑘2 𝑘𝑁
(3.81)
{𝑎(𝜔)} = {𝑎1 (𝜔) 𝑎2 (𝜔) . . . . . 𝑎𝑁 (𝜔)}𝑇 , (3.82,a)
{𝑏(𝜔)} = {𝑏1 (𝜔) 𝑏2 (𝜔) . . . . . 𝑏𝑁 (𝜔)}𝑇 , (3.82,b)
𝑇
{𝛼ℓ𝑅 (𝜔)} = {𝛼1ℓ𝑅 𝛼2ℓ𝑅 . . 𝛼𝑗ℓ𝑅 . . . 𝛼𝑁1ℓ𝑅 } , (3.83,a)
𝑇
{𝛼ℓ𝐼 (𝜔)} = {𝛼1ℓ𝐼 𝛼2ℓ𝐼 . . 𝛼𝑗ℓ𝐼 . . . 𝛼𝑁1ℓ𝐼 } , (3.83,b)

[𝜒] = [{𝜒ℓ(1) } {𝜒ℓ(2) } . . .. {𝜒ℓ(𝑁) }] , (3.84)


{𝐴(𝜔)} = {𝐴1 (𝜔) 𝐴2 (𝜔) . . . . . 𝐴𝑁 (𝜔)}𝑇 , (3.85,a)
{𝐵(𝜔)} = {𝐵1 (𝜔) 𝐵2 (𝜔) . . . . . 𝐵𝑁 (𝜔)}𝑇 . (3.85,b)

3.2.3. Real Forced Modes of Vibration

The equations for the forced vibrations of a system with viscous and
hysteretic damping are
1
[𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐻] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} . (3.86)
𝜔

In the case of harmonic excitation


{𝑓} = {𝑓̂}𝑒 𝑖𝜔𝑡 , {𝑞} = {𝑞̃}𝑒 𝑖𝜔𝑡 , (3.87)
equations (3.86) become
[[𝐾 − 𝜔2 𝑀] + 𝑖 [𝐻 + 𝜔𝐶]] {𝑞̃} = {𝑓̂}. (3.86,a)
Following Bishop and Gladwell [13], it is enquired whether there is a real
column vector {𝑓̂} which defines an excitation vector with in-phase forces such that
at any frequency 𝜔 the displacements 𝑞𝑗 (𝑗 = 1, 2, . . . , 𝑁) are all in phase, though
not necessarily in phase with the forces. The vector {𝑞̃} will be of the form
{𝑞̃} = {𝑞̂}𝑒 −𝑖𝜑 , (3.88)
where 𝜑 is the unknown phase lag angle between forces and displacements, and
{𝑞̂} is a unknown vector of real displacement amplitudes.
Substituting (3.87) and (3.88) into equation (3.86) we get
([𝐾 − 𝜔2 𝑀] + 𝑖 [𝐻 + 𝜔𝐶]) (𝑐𝑜𝑠 𝜑 − 𝑖 𝑠𝑖𝑛 𝜑) {𝑞̂} = {𝑓̂} . (3.89)
Separating the real and imaginary parts, we obtain
([𝐻 + 𝜔𝐶] 𝑐𝑜𝑠 𝜑 − [𝐾 − 𝜔2 𝑀] 𝑠𝑖𝑛 𝜑 ) {𝑞̂} = {0} , (3.90)

103
([𝐻 + 𝜔𝐶]𝑠𝑖𝑛 𝜑 + [𝐾 − 𝜔2 𝑀] 𝑐𝑜𝑠 𝜑 ) {𝑞̂} = {𝑓̂} . (3.91)
Provided that 𝑐𝑜𝑠 𝜑 ≠ 0, equation (3.90) becomes
([𝐻 + 𝜔𝐶] − [𝐾 − 𝜔2 𝑀] 𝑡𝑎𝑛 𝜑 ) {𝑞̂} = {0} , (3.92)
which is a set of homogeneous equations which has nontrivial solutions if
𝑑𝑒𝑡 ( 𝑡𝑎𝑛 𝜑 [𝐾 − 𝜔2 𝑀] − [𝐻 + 𝜔𝐶] ) = 0 . (3.93)
This is a problem of latent roots of a matrix pencil [66], in which the latent
roots are 𝑡𝑎𝑛 𝜑. Relation (3.93) is an algebraic equation of order 𝑁 in 𝑡𝑎𝑛 𝜑. For
each root 𝑡𝑎𝑛 𝜑𝑟 , there is an associated modal vector {𝛷 (𝑟) } satisfying the
equation
( 𝑡𝑎𝑛 𝜑𝑟 [𝐾 − 𝜔2 𝑀] − [𝐻 + 𝜔𝐶] ){𝛷 (𝑟) } = 0 . (3.94)
Both 𝑡𝑎𝑛 𝜑𝑟 and {𝛷 (𝑟) } are real and frequency dependent.
The corresponding force vector {ℱ̂ (𝑟) } required to produce {𝛷 (𝑟) } is
obtained from (3.91)
{ℱ̂ (𝑟) } = 𝑐𝑜𝑠 𝜑𝑟 [𝐾 − 𝜔2 𝑀]{𝛷 (𝑟) } + 𝑠𝑖𝑛 𝜑𝑟 [𝐻 + 𝜔𝐶] {𝛷 (𝑟) } (3.95)
or from one of the following relations
[𝐾 − 𝜔2 𝑀]{𝛷 (𝑟) } = 𝑐𝑜𝑠 𝜑𝑟 {ℱ̂ (𝑟) } , (3.96)
[𝐻 + 𝜔𝐶] {𝛷 (𝑟) } = 𝑠𝑖𝑛 𝜑𝑟 {ℱ̂ (𝑟) } . (3.97)
It comes out that the solution (3.88) is possible, as first pointed out by
Fraeijs de Veubeke [120]. There exist 𝑁 real modes of vibration {𝛷 (𝑟) } called
modes of distortion, normal frequency response modes or monophase response
modes [120]. They are different from the principal modes of vibration {𝛹 (𝑟) } of the
undamped system; their shape and the phase lag angle 𝜑𝑟 vary with frequency.
They are postulated for the forced vibrations of a system with nonproportional
damping, while the modes {𝛹 (𝑟) } are defined for the free vibrations of the associate
undamped system (or a system with proportional damping).
The corresponding monophase excitation vectors {ℱ̂ (𝑟) } are termed normal
excitation modes [120].
It follows then from (3.94) that
𝑇
(𝑟) (𝑟)
{𝛷 } [𝐻+𝜔𝐶] {𝛷 }
𝑡𝑎𝑛 𝜑𝑟 =
(𝑟) 𝑇 (𝑟)
. (3.98)
{𝛷 } [𝐾−𝜔2 𝑀]{𝛷 }

The phase lag angles 𝜑𝑟 have a continuous variation (with the frequency
𝜔 ) from a small positive value (at 𝜔 = 0) to 180𝑜 for high frequencies [121].
When 𝜑𝑟 = 90𝑜 , 𝑐𝑜𝑠 𝜑𝑟 = 0 , and equation (3.90) becomes
([𝐾] − 𝜔2 [𝑀]) {𝑞̂} = {0}, (3.99)

104
which is identical to (3.3), the eigenvalue
problem of the associate undamped system
([𝐾] − 𝜔𝑟2 [𝑀]) {𝛹 (𝑟) } = {0}.
(3.99,a)
If 𝜔 = 𝜔𝑟 , then 𝜑𝑟 = 90𝑜 , and the
monophase response mode {𝛷 (𝑟) } may be
identified with the 𝑟-th principal mode of the
undamped system.
The remaining (𝑁 − 1) response modes, associated to the other (𝑁 − 1)
roots 𝑡𝑎𝑛 𝜑𝑟 of equation (3.93), behave in a similar way. For instance, the shape
of the mode {𝛷 (𝑠) } varies with ω . When 𝜔 = 𝜔𝑠 , we get 𝑐𝑜𝑠 𝜑𝑠 = 0, 𝜑𝑠 = 90𝑜
𝜋
and {𝛷 (𝑠) } ≡ {𝛹 (𝑠) } . Hence 𝜑𝑠 (𝜔) is the characteristic phase lag which is when
2
𝜔 = 𝜔𝑠 .
Figure 3.1 shows the variation of characteristic phase lags with frequency.
The monophase response modal vectors {𝛷 (𝑟) } satisfy the orthogonality
conditions
𝑇
{𝛷 (𝑟) } [𝐾 − 𝜔2 𝑀] {𝛷 (𝑠) } = 0, (𝑟 ≠ 𝑠) (3.100)
𝑇
{𝛷 (𝑟) } [𝐻 + 𝜔 𝐶] {𝛷 (𝑠) } = 0 . (𝑟 ≠ 𝑠) (3.101)
(𝑠) 𝑇
Premultiplying equation (3.95) by {𝛷 } and using the above
orthogonality conditions we obtain
𝑇 𝑇
{𝛷 (𝑠) } {ℱ̂ (𝑟) } = {𝛷 (𝑟) } {ℱ̂ (𝑠) } = 0 , (𝑟 ≠ 𝑠) (3.102)
hence the energy input from the monophase excitation mode {ℱ̂ (𝑟) } goes only into
its associated monophase response mode {𝛷 (𝑟) }. The energy dissipated per cycle is
(𝑟) 𝑇 ̂ (𝑟) 𝑇
𝑊𝑟 = 𝜋 𝑠𝑖𝑛 𝜑𝑟 {𝛷 } {ℱ } = 𝜋 {𝛷 (𝑟) } [𝐻 + 𝜔 𝐶] {𝛷 (𝑟) }. (3.103)
Let introduce a square matrix having the response modes as columns
[𝛷] = [{𝛷 (1) } {𝛷 (2) } . . . {𝛷 (𝑁) }] (3.104)
and the coordinate transformation
{𝑞} = [𝛷] {𝜈} = ∑𝑁 (𝑟)
𝑟=1{𝛷 } 𝜈𝑟 (3.105)
in which the multipliers 𝜈𝑟 are the damped principal coordinates [13].
Inserting (3.105) into (3.86) and premultiplying by [𝛷]𝑇 we obtain
1
[𝑚 ̅ ] {𝜈̇ } + [𝑘̅ ] {𝜈} = [𝛷]𝑇 {𝑓}
̅ ] {𝜈̈ } + 𝜔 [ˋ𝑑ˏ (3.106)
where
̅ ] = [𝛷]𝑇 [𝑀] [𝛷] ,
[𝑚 (3.107,a)

105
1 1
̅]
[ˋ𝑑ˏ = [𝛷]𝑇 [𝐶 + 𝜔 𝐻] [𝛷] , (3.107,b)
𝜔

[𝑘̅ ] = [𝛷]𝑇 [𝐾] [𝛷] . (3.107,c)


Generally, matrices [𝑚 ̅ ] and [𝑘̅ ] are not diagonal, but according to the
orthogonality conditions (3.100) and (3.101), the following matrices are diagonal
[ˋ [𝑘̅] − 𝜔2 [𝑚
̅ ] ˏ] = 𝑑𝑖𝑎𝑔 [𝑘̅𝑟 ] , (3.108)
̅ ] = 𝑑𝑖𝑎𝑔[𝑑̅𝑟 ] .
[ˋ𝑑ˏ (3.109)
The equation associated with the coordinate 𝜈𝑟 is
𝑇 1 𝑇
{𝛷 (𝑟) } [𝑀] {𝛷 (𝑟) } 𝜈̈𝑟 + 𝜔 {𝛷 (𝑟) } [𝜔𝐶 + 𝐻] {𝛷 (𝑟) } 𝜈̇𝑟 +
𝑇 𝑇
+{𝛷 (𝑟) } [𝐾] {𝛷 (𝑟) } 𝜈𝑟 = {𝛷 (𝑟) } {𝑓} . (3.110)
For steady-state harmonic conditions
{𝑓} = {𝑓̂}𝑒 𝑖𝜔𝑡 , {𝜈} = {𝜈̃}𝑒 𝑖𝜔𝑡 , (3.111)
which substituted into equation (3.110) yield
𝑇
{𝛷 (𝑟) } {𝑓̂}
𝜈̃𝑟 = 𝑇 𝑇 (3.112)
{𝛷(𝑟) } [𝐾−𝜔2 𝑀] {𝛷 (𝑟) } +𝑖 {𝛷 (𝑟) } [𝜔𝐶+𝐻] {𝛷 (𝑟) }
or, using equations (3.108) and (3.109)
𝑇
{𝛷(𝑟) } {𝑓̂ }
𝜈̃𝑟 = ̅ 𝑟 +𝑖 𝑑̅𝑟
(3.113)
𝑘
Hence, equation (3.86) has the solution
𝑟 𝑇
̂ } {𝛷 𝑟 }
{𝛷 } {𝑓
( ) ( )

{𝑞̃} = ∑𝑁
𝑟=1 ̅ +𝑖 𝑑̅ . (3.114)
𝑘 𝑟 𝑟
Equation (3.98) can be written
𝑑̅
𝑡𝑎𝑛 𝜑𝑟 = 𝑘̅𝑟 . (3.115)
𝑟

In this case, the complex modal receptance has the expression


1 1 1 𝑠𝑖𝑛 𝜑𝑟 𝑒 −𝑖𝜑𝑟
𝛼̅𝑟 = ̅ 𝑟 +𝑖 𝑑̅𝑟
= ̅ 1 = . (3.116)
𝑘 𝑑𝑟 +𝑖 𝑑̅𝑟
𝑡𝑎𝑛 𝜑𝑟

Using notation (3.116), the expression (3.114) can be written


𝑇
{𝑞̃} = ∑𝑁 ̅𝑟 {𝛷 (𝑟) } {𝛷 (𝑟) } {𝑓̂}
𝑟=1 𝛼 (3.117)

106
𝑇
where the dyadic products {𝛷 (𝑟) } {𝛷 (𝑟) } are square matrices of order 𝑁 and
rank 1, of the form
(𝑟) (𝑟) (𝑟) (𝑟) (𝑟) (𝑟)
𝛷1 𝛷1 𝛷1 𝛷2 ⋯ 𝛷1 𝛷𝑁
(𝑟) 𝑇
{𝛷 (𝑟) } {𝛷 } =[ ⋯ ⋯ ⋯ ⋯ ]. (3.118)
(𝑟) (𝑟) (𝑟) (𝑟) (𝑟) (𝑟)
𝛷𝑁 𝛷1 𝛷𝑁 𝛷2 ⋯ 𝛷𝑁 𝛷𝑁
The same result is obtained expressing the excitation vector {𝑓̂} as a linear
combination of the monophase vectors {ℱ̂ (𝑟) }, which are linearly independent [13].
Because the excitation {𝑓̂} = {ℱ ̂ (𝑟) } produces a response {𝑞̃} =
{𝛷 (𝑟) }𝑒 −𝑖 𝜑𝑟 , an excitation of the form
{𝑓̂} = ∑𝑁 ̂ (𝑟) }
𝑟=1 𝜆𝑟 {ℱ (3.119)
produces a response
{𝑞̃} = ∑𝑁 (𝑟) −𝑖 𝜑𝑟
𝑟=1 𝜆𝑟 {𝛷 }𝑒 (3.120)
where 𝜆𝑟 are real multipliers.
𝑇
Premultiplying in (3.119) by {𝛷 (𝑟) } and using (3.102) we obtain
𝑇
(𝑟)
{𝛷 ̂}
} {𝑓
𝜆𝑟 = (3.121)
(𝑟) 𝑇 (𝑟)
{ 𝛷 } {ℱ ̂ }

or, using (3.97)


(𝑟) 𝑇
{𝛷 ̂ } 𝑠𝑖𝑛 𝜑
} {𝑓 𝑟
𝜆𝑟 =
(𝑟) 𝑇 (𝑟)
. (3.122)
{𝛷 } [𝐻+𝜔𝐶 ]{𝛷 }

Inserting (3.122) into (3.120) we obtain expression (3.117)


𝑟 ( ) 𝑟 ( ) 𝑇
{𝛷 } {𝛷 } 𝑠𝑖𝑛 𝜑𝑟 𝑒−𝑖 𝜑𝑟 ̂
{𝑞̃} = ∑𝑁
𝑟=1 {𝑓} . (3.123)
(𝑟) 𝑇 (𝑟)
{𝛷 } [𝐻+𝜔𝐶 ]{𝛷 }

The complex receptance 𝛼̅𝑗ℓ , defined as the response at 𝑞𝑗 due to a


harmonic force at 𝑞ℓ , is
(𝑟) (𝑟)
𝑞̃ 𝛷ℓ 𝛷𝑗 𝑠𝑖𝑛 𝜑𝑟 𝑒−𝑖 𝜑𝑟
𝛼̅𝑗ℓ = 𝑓̂𝑗 = ∑𝑁
𝑟=1 (𝑟) 𝑇 (𝑟)
= (3.124)

{𝛷 } [𝐻+𝜔𝐶 ]{𝛷 }

(𝑟) (𝑟)
𝛷ℓ 𝛷𝑗 𝑒−𝑖 𝜑𝑟
= ∑𝑁
𝑟=1 2 2
.
𝑇 𝑇
√({𝛷(𝑟) } [𝐻+𝜔𝐶 ] {𝛷(𝑟) }) +({𝛷(𝑟) } [𝐾−𝜔2 𝑀 ] {𝛷(𝑟) })

It is helpful to introduce the notations

107
(𝑟)𝑇 (𝑟)
{𝛷 } [𝐻+𝜔𝐶 ] {𝛷 }
= 𝛾𝑟 , (3.125)
(𝑟) 𝑇 (𝑟)
{𝛷 } [𝐾 ] {𝛷 }

𝑇
(𝑟) (𝑟)
{𝛷 } [𝐾 ] {𝛷 }
= 𝛺𝑟2 , (3.126)
(𝑟) 𝑇 (𝑟)
{𝛷 } [𝑀 ] {𝛷 }

and to normalize the monophase response modes by assigning to each the same
“modal mass”
𝑇
{𝛷 (𝑟) } [𝑀 ] {𝛷 (𝑟) } = 𝜇 . (3.127)
Separating the real and imaginary components of the response
{𝑞̃} = {𝑞𝑅 } + 𝑖 {𝑞𝐼 } = [𝛼̅] {𝑓̂} = [𝛼𝑅 + 𝑖 𝛼𝐼 ] {𝑓̂} , (3.128)
in view of (3.120), (3.122) and (3.125) to (3.127), we obtain the in-phase and
quadrature responses
1 𝑠𝑖𝑛 𝜑𝑟 𝑐𝑜𝑠 𝜑𝑟 𝑇
{𝑞𝑅 } = ∑𝑁
𝑟=1 𝜆𝑟 𝑐𝑜𝑠 𝜑𝑟 {𝛷
(𝑟)
} = ∑𝑁
𝑟=1
𝜇 𝛾𝑟 𝛺𝑟2
{𝛷 (𝑟) } {𝑓̂}{𝛷 (𝑟) }, (3.129)

(𝑟) 1 𝑠𝑖𝑛2 𝜑𝑟 (𝑟) 𝑇 ̂


−{𝑞𝐼 } = ∑𝑁
𝑟=1 𝜆𝑟 𝑠𝑖𝑛 𝜑𝑟 {𝛷 } = 𝜇 ∑𝑁
𝑟=1 2
𝛾𝑟 𝛺𝑟
{𝛷 } {𝑓}{𝛷 (𝑟) }, (3.130)

1 𝑠𝑖𝑛 𝜑𝑟 𝑐𝑜𝑠 𝜑𝑟 𝑇
[𝛼𝑅 ] = ∑𝑁
𝑟=1 {𝛷 (𝑟) }{𝛷 (𝑟) } (3.131)
𝜇 𝛾𝑟 𝛺𝑟2
1 𝑠𝑖𝑛2 𝜑𝑟 𝑇
−[𝛼𝐼 ] = ∑𝑁
𝑟=1 {𝛷 (𝑟) }{𝛷 (𝑟) } (3.132)
𝜇 𝛾𝑟 𝛺𝑟2
In this way we have actually inverted relation (3.86,a) and found the
reciprocal
−1
[[𝐾 − 𝜔2 𝑀] + 𝑖 [𝐻 + 𝜔𝐶]] = [𝛼𝑅 ] + 𝑖 [𝛼𝐼 ] . (3.133)
We can calculate now the components of the complex energy input per
cycle.
The active energy
𝑇 (𝑟) 𝑇 [𝐻
𝑊𝑅 = 𝜋{𝑓̂} {−𝑞𝐼 } = 𝜋{−𝑞𝐼 }𝑇 {𝑓̂} = 𝜋 ∑𝑁 2
𝑟=1 𝜆𝑟 {𝛷 } + 𝜔𝐶]{𝛷 (𝑟) } (3.134)
is a positive definite quadratic form, which is consistent with its character of
dissipated energy.
The reactive energy
𝑇 (𝑟) 𝑇 [𝐻
𝑊𝐼 = 𝜋{𝑓̂} {𝑞𝑅 } = 𝜋{𝑞𝑅 }𝑇 {𝑓̂} = 𝜋 ∑𝑁 𝑟=1 𝑐𝑜𝑡𝑎𝑛𝜑𝑟 {𝛷 } + 𝜔𝐶]{𝛷 (𝑟) }
(3.135)
is generally not positive definite. Each time a phase resonance is passed, one of the
𝑐𝑜𝑡𝑎𝑛 𝜑𝑟 coefficients passes from a positive to a negative value, changing the sign
of the expression.

108
These characteristics represent the experimental basis of the complex
power method presented in § 6.2.2.5.
Fraeijs de Veubeke [122] has demonstrated that, if the response is in a
natural mode of vibration, the reactive energy 𝑊𝐼 is stationary with respect to
arbitrary small variations of the generalized forces. The condition
𝜕𝑊𝐼 𝑇
𝜕𝜆𝑟
= 2𝜋𝜆𝑟 𝑐𝑜𝑡𝑎𝑛𝜑𝑟 {𝛷 (𝑟) } [𝐻 + 𝜔𝐶]{𝛷 (𝑟) } = 0 (3.136)
is satisfied for all values of 𝑟 if all 𝜆𝑟 ′s are zero except one
𝜆1 = 𝜆2 = . . . = 𝜆ℓ−1 = 𝜆ℓ+1 = . . . = 𝜆𝑁 = 0 , 𝜆ℓ ≠ 0
and
𝜋
𝑐𝑜𝑡𝑎𝑛 𝜑ℓ = 0 or 𝜑ℓ = 2
.
These conditions are similar to the following:
a) 𝜔 = 𝜔ℓ is a natural frequency;
b) {𝑞̃} = −𝑖 𝜆ℓ {𝛹 (ℓ) } is a natural mode in quadrature with the excitation;
c) {𝑓̂} = 𝜆ℓ [𝐻 + 𝜔ℓ 𝐶] {𝛹 (ℓ) } is a pure mode excitation.
Based on equations (3.128), the reactive energy (3.135) may be written
under the form
𝑇
𝑊𝐼 = 𝜋{𝑓̂} [𝛼𝑅 ] {𝑓̂} . (3.137)
Because for arbitrary variations {𝛿𝑓̂} of the excitation forces
̂ }𝑇 [𝛼𝑅 ] {𝑓̂} + {𝑓̂}𝑇 [𝛼𝑅 ] {𝛿𝑓
𝛿𝑊𝐼 = 𝜋 ({𝛿𝑓 ̂ }𝑇 [𝛼𝑅 ] {𝑓̂} = 0, (3.138)
̂ } ) = 2𝜋 {𝛿𝑓

it results that the vanishing of the response in phase


{𝑞𝑅 } = [𝛼𝑅 ] {𝑓̂} = {0} (3.139)
implies the existence of a natural response mode in quadrature with the forces,
which can only occur at a resonance frequency.
The natural frequencies are therefore also solutions of the equation
𝑑𝑒𝑡 [𝛼𝑅 (𝜔)] = 0 (3.140)
and the pure excitation modes are the corresponding eigensolutions of equation
(3.139).
Substituting (3.139) into (3.128) we obtain
{𝑞𝐼 } = [𝛼𝐼 ] {𝑓̂} (3.141)
which yields the shapes of the undamped natural modes, if {𝑓̂} is replaced by the
expression obtained from (3.139).
These observations form the basis of the excitation of pure modes,
presented in Chapter 6.

109
3.2.4. Complex Modes of Vibration

The dynamic response of a linear structure with nonproportional damping,


described in terms of the real monophase vectors, can be replaced by a modal
representation in terms of complex modes of vibration.
A general formulation of the problem is based on the use of the so-called
“lambda matrices” [66].
In the case of the free vibrations of a system with pure viscous damping,
substituting in the equations of motion (3.26) a solution of the form
{𝑞(𝑡)} = {𝑦} 𝑒 𝜆 𝑡 , (3.142)
we obtain a set of 𝑁 homogeneous algebraic equations, which represents the
quadratic eigenvalue problem
(𝜆2 [𝑀] + 𝜆 [𝐶] + [𝐾]) {𝑦} = {0} . (3.143)
These equations have a nontrivial solution only if the determinant of the
lambda matrix vanishes
𝑑𝑒𝑡 (𝜆2 [𝑀] + 𝜆 [𝐶] + [𝐾]) = 0 . (3.144)
The characteristic equation (3.144) is an algebraic homogeneous equation
of order 2𝑁 in 𝜆 . Its roots can be real, pure imaginary or complex. The negative
real roots correspond to an overdamped system which exhibits a decaying aperiodic
motion. The pure imaginary roots are conjugate and correspond to undamped
systems. These two cases are not considered in the following. Of interest is the
special case of complex conjugate roots with negative real part (stable systems)
which correspond to underdamped systems having a decaying periodic motion.
Each eigenvalue 𝜆𝑟 has an associate modal vector {𝑦 (𝑟) } with complex
elements, defining a complex mode of vibration. As the eigenvalues are complex
conjugate, so are the modal vectors. A pair of complex conjugate modal vectors,
multiplied by the time dependent exponential functions, can be combined to yield
a real quantity describing a damped oscillatory motion.
The general solution of the homogeneous equation (3.26) with zero right
hand side can be expressed as a linear combination of solutions of the form
{𝑦 (𝑟) }𝑒 𝜆𝑟 𝑡 multiplied by arbitrary constants 𝛾𝑟
{𝑞(𝑡)} = [𝑦] {𝛾 𝑒 𝜆 𝑡 } (3.145)
where the modal matrix
[𝑦] = [ {𝑦 (1) } {𝑦 (2) } . . . {𝑦 (2𝑁) } ] (3.146)
and {𝛾 𝑒 𝜆 𝑡 } is a column vector of elements which play the role of damped principal
coordinates.
In the case of hysteretic damping, the modal representation consists of only
𝑁 modal vectors, which simplifies the calculations.

110
A complex mode of vibration has the character of a travelling wave (unlike
the stationary wave character of the “classical” real modes), because each complex
element of a modal vector has a different phase lag angle, hence the corresponding
coordinate reaches its maximum excursion at a different time than the others.
Complex modes are more difficult to measure by test because the motion, though
synchronous, is not monophase. The vibration “nodes” continuously change their
position during a cycle, but during the next cycle the pattern repeats itself, so that
the coordinates are in the same position as at the beginning of the previous cycle.
In the case of free vibrations, the maximum excursions decay exponentially
from cycle to cycle, unlike the free motion of undamped systems, in which the
nodes are stationary and the maximum elongations do not decay in time.
Due to these difficulties of ”recognition” and measurement of the shape of
a complex mode of vibration, it is customary to use combined
analytical/experimental identification methods to calculate the modal vectors based
on test data.

3.2.4.1. Viscous damping

Although the general theory of vibration of systems with nonproportional


damping can be derived using the Laplace transform [101] and lambda matrices,
the following presentation is based on the method introduced by Frazer, Duncan
and Collar [40].
The equation of motion of a system with viscous damping is
[𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} , (3.147)
where [𝐶] is symmetric.
Adjoining the identity
[𝑀] {𝑞̇ } − [𝑀] {𝑞̇ } = {0} (3.148)
we obtain a set of 2𝑁 first order differential equations
[𝑈] {𝑥̇ } + [𝑉] {𝑥} = {𝑆} (3.149)
in which
[0] [𝑀] −[𝑀] [0]
[𝑈] = [ ], [𝑉] = [ ] (3.150)
[𝑀] [𝐶] [0] [𝐾]
are real, symmetric, square matrices of order 2𝑁 and
{𝑞̇ } {0}
{𝑥} = { } , {𝑆} = { } (3.151)
{𝑞} {𝑓}
are column vectors with 2𝑁 elements.
The solution of equation (3.149) by modal analysis follows closely the
procedure used for undamped systems.

111
a) Free vibrations
Consider the homogeneous equation
[𝑈] {𝑥̇ } + [𝑉] {𝑥} = {0} (3.152)
and seek a solution of the form
{𝑥} = {𝜉} 𝑒 𝜎𝑡 (3.153)
in which {𝜉} is a column vector consisting of 2𝑁 constant elements.
Equation (3.153), when introduced into (3.152), leads to the eigenvalue
problem
𝜎 [𝑈] {𝜉} + [𝑉] {𝜉} = {0} . (3.154)
If the stiffness matrix [𝐾] is nonsingular, equation (3.154) can be
premultiplied by [𝑉]−1 which yields
1
([𝐷] − 𝜎 [𝐼]) {𝜉} = {0}, (3.155)

in which [𝐼] is the identity matrix of order 2𝑁 and


[0] [𝐼]
[𝐷] = −[𝑉]−1 [𝑈] = [ −1 [𝑀] ], (3.156)
−[𝐾] −[𝐾]−1 [𝐶]
plays the role of a dynamic matrix.
Equations (3.155) have nontrivial solutions if
1
𝑑𝑒𝑡 ([𝐷] − 𝜎 [𝐼]) = 0 . (3.157)
1
The solutions of this characteristic equation of order 2𝑁 in are the
𝜎
1
eigenvalues (𝑟 = 1, 2, . . . , 2𝑁). In the following they are considered distinct.
𝜎𝑟
1
Corresponding to each eigenvalue there is an eigenvector {𝜉 (𝑟) } having 2𝑁
𝜎𝑟
components which satisfies the equation
1
([𝐷] − 𝜎 [𝐼]) {𝜉 (𝑟) } = {0}
𝑟
or
(𝑟)
(𝜎𝑟 [𝑈] + [𝑉]) {𝜉 } = {0} . (3.158)

The eigenvectors {𝜉 (𝑟) } satisfy the following orthogonality conditions


𝑇
{𝜉 (𝑠) } [𝑈]{𝜉 (𝑟) } = 0 , (𝑟 ≠ 𝑠)
(3.159)
(𝑠) 𝑇 [𝑉] (𝑟)
{𝜉 } {𝜉 }=0. (𝑟 ≠ 𝑠)
Denoting

112
𝑇
𝑢𝑟 = {𝜉 (𝑟) } [𝑈]{𝜉 (𝑟) } ,
(3.160)
(𝑟) 𝑇 [𝑉] (𝑟)
𝑣𝑟 = {𝜉 } {𝜉 },
equation (3.158) becomes
𝜎𝑟 𝑢𝑟 + 𝑣𝑟 = 0 . (3.161)
The eigenvectors have the form

(𝑟) 𝜎𝑟 {𝑞(𝑟) }
{𝜉 } = { (𝑟) } (3.162)
{𝑞 }
where {𝑞(𝑟) } is a vector having 𝑁 elements and represents the lower half of the
vector {𝜉 (𝑟) }.
For underdamped stable systems, the frequency 𝜎𝑟 is complex and is
usually expressed under the form
𝜎𝑟 = −𝑛𝑟 + 𝑖 𝜇𝑟 , (3.163)
where 𝑛𝑟 is a damping factor (decay rate) and 𝜇𝑟 is the damped natural frequency.
Other related and commonly used terms are the damping ratio
2 𝑛𝑟 2 𝑛𝑟
𝜏𝑟 = = (3.164)
𝜔𝑟 √𝑛𝑟2 +𝜇𝑟2
and the resonant frequency

𝜔𝑟 = √𝑛𝑟2 + 𝜇𝑟2 . (3.165)

b) Forced vibrations

Equations (3.149) can be uncoupled by the linear transformation


{𝑥} = [𝜉] {𝑧} = ∑2𝑁 (𝑟)
𝑟=1 {𝜉 } 𝑧𝑟 , (3.166)
where 𝑧𝑟 are principal coordinates, and the modal matrix is
[𝜉] = [{𝜉 (1) } {𝜉 (2) } . . . {𝜉 (2𝑁) }] (3.167)
Substituting (3.166) into (3.149) and premultiplying by [𝜉]𝑇 we obtain the
uncoupled equations
[ˋ𝑢ˏ] {𝑧̇ } + [ˋ𝑣ˏ] {𝑧} = {𝑅} (3.168)
in which
[𝜉]𝑇 [𝑈] [𝜉] = [ˋ𝑢ˏ] = 𝑑𝑖𝑎𝑔 [𝑢𝑟 ] , (3.169)
[𝜉]𝑇 [𝑉] [𝜉] = [ˋ𝑣ˏ] = 𝑑𝑖𝑎𝑔 [𝑣𝑟 ] , (3.170)

113
[𝜉]𝑇 {𝑆} = {𝑅}. (3.171)
The 𝑟-th equation in (3.168) is
𝑢𝑟 𝑧̇𝑟 + 𝑣𝑟 𝑧𝑟 = 𝑅𝑟 (3.172)
or, based on equation (3.161),
𝑅
𝑧̇𝑟 − 𝜎𝑟 𝑧𝑟 = 𝑢𝑟 . (3.173)
𝑟

Assuming zero initial conditions, the solution of equation (3.173) can be


written under the form of the convolution integral
1 𝑡
𝑧𝑟 (𝑡) = 𝑢 ∫0 𝑅𝑟 (𝜏) 𝑒 𝜎𝑟 (𝑡−𝜏) 𝑑𝜏.
𝑟

For harmonic steady-state conditions


{𝑓(𝑡)} = {𝑓̂}𝑒 𝑖𝜔𝑡 , {𝑞(𝑡)} = {𝑞̃}𝑒 𝑖𝜔𝑡 , (3.174)
we can denote
{𝑥(𝑡)} = {𝑥̃}𝑒 𝑖𝜔𝑡 , {𝑆(𝑡)} = {𝑆̂}𝑒 𝑖𝜔𝑡 , (3.175)
and
{𝑧(𝑡)} = {𝑧̃ }𝑒 𝑖𝜔𝑡 , {𝑅(𝑡)} = {𝑅̃}𝑒 𝑖𝜔𝑡 . (3.176)
Substituting solutions of the form (3.176) into (3.173) we obtain
𝑅̃𝑟
𝑧̃𝑟 = (3.177)
𝑢𝑟 (𝑖𝜔−𝜎𝑟 )
or, using (3.171),
𝑇
{𝜉 (𝑟) } {𝑆̂}
𝑧̃𝑟 = . (3.178)
𝑢𝑟 (𝑖𝜔−𝜎𝑟 )
Inserting (3.178) into (3.166) and using (3.175) we get
(𝑟) (𝑟) 𝑇
{𝜉 } {𝜉 }
{𝑥̃} = ∑2𝑁 ̂
𝑟=1 𝑢𝑟 (𝑖𝜔−𝜎𝑟 ) {𝑆} (3.179)

From (3.151) and based on (3.174)


𝑇
𝑖𝜔{𝑞̃} {𝜉 (𝑟) } {𝜉 (𝑟) } {0}
{ } = ∑2𝑁
𝑟=1 𝑢 (𝑖𝜔−𝜎 ) { ̂ } (3.180)
{𝑞̃} 𝑟 𝑟 {𝑓}
or
𝑇
𝑖𝜔{𝑞̃} {𝜉 (𝑟) } {0} (𝑟)
{ } = ∑2𝑁
𝑟=1 𝑢 (𝑖𝜔−𝜎 ) { ̂ } {𝜉 } (3.181)
{𝑞̃} 𝑟 𝑟 {𝑓}
Based on (3.162)

114
𝑇
{𝑞 (𝑟) } {𝑓̂ }
{𝑞̃} = ∑2𝑁
𝑟=1 𝑢 {𝑞 (𝑟) } . (3.182)
𝑟 (𝑖𝜔−𝜎𝑟 )

For single point harmonic excitation at ℓ and response at 𝑗 , the receptance


is
(𝑟) (𝑟)
𝑞̃𝑗 𝑞ℓ 𝑞𝑗
𝛼̅𝑗ℓ = = ∑2𝑁
𝑟=1 𝑢 (𝑖𝜔−𝜎 ) (3.183)
𝑓̂ℓ 𝑟 𝑟
or
(𝑟 )
𝛿𝑗ℓ
𝛼̅𝑗ℓ = ∑2𝑁
𝑟=1 𝑖𝜔−𝜎 , (3.184)
𝑟
where
(𝑟)
𝑞(ℓ𝑟) 𝑞(𝑗𝑟)
𝛿𝑗ℓ =
𝑢𝑟
. (3.185)

Because for underdamped systems the eigenvalues 𝜎𝑟 and the eigenvectors


{𝑞𝑟 } occur in complex conjugate pairs, expression (3.183) can be written as a sum
of 𝑁 terms
(𝑟 ) (𝑟 )∗
𝛿𝑗ℓ 𝛿𝑗ℓ
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 ( + ) (3.186)
𝑖𝜔−𝜎𝑟 𝑖𝜔−𝜎𝑟∗

where 𝜎𝑟∗ and 𝛿𝑗ℓ


(𝑟)∗ (𝑟)
are the complex conjugates of 𝜎𝑟 and 𝛿𝑗ℓ , respectively.

3.2.4.2. Hysteretic damping

Consider the equations of motion (3.48) for a hysteretically damped system


with harmonic excitation
(−𝜔2 [𝑀] + 𝑖 [𝐻] + [𝐾]) {𝑞̃} = {𝑓̂}, (3.187)
where matrices [𝑀] and [𝐾 + 𝑖𝐻] are symmetric, of order 𝑁, {𝑞̃} is the column
vector of complex displacement amplitudes and {𝑓̂} – the real forcing vector.
Let 𝜔2 = 𝜆 and consider the homogeneous equation
([𝐾 + 𝑖𝐻] − 𝜆 [𝑀]) {𝑞̃} = {0}. (3.188)
It has nontrivial solutions if
𝑑𝑒𝑡 ([𝐾 + 𝑖𝐻] − 𝜆 [𝑀]) = 0 . (3.189)
It is demonstrated [63] that there exist a set of 𝑁 complex eigenvalues 𝜆̅𝑟 ,
solutions of equation (3.189), and 𝑁 associated complex eigenvectors {𝑤 (𝑟) } which
satisfy the homogeneous equation
([𝐾 + 𝑖𝐻] − 𝜆̅𝑟 [𝑀]) {𝑤 (𝑟) } = {0} . (3.190)

115
In the following it is assumed that the 𝑁 eigenvalues are distinct, and the
associated modal vectors are linearly independent.
The modal vectors satisfy the orthogonality conditions
𝑇
{𝑤 (𝑠) } [𝑀]{𝑤 (𝑟) } = 0 , (𝑟 ≠ 𝑠) (3.191)
𝑇
{𝑤 (𝑠) } [𝐾 + 𝑖𝐻]{𝑤 (𝑟) } = 0 (𝑟 ≠ 𝑠) (3.192)
The complex modal mass and complex modal stiffness are defined as
̅𝑟 = {𝑤 (𝑟) }𝑇 [𝑀] {𝑤 (𝑟) } ,
𝑀 (3.193)
̅𝑟 = {𝑤 (𝑟) }𝑇 [𝐾 + 𝑖𝐻] {𝑤 (𝑟) } .
𝐾 (3.194)
Because the 𝑁 eigenvectors {𝑤 (𝑟) } are linearly independent, any vector {𝑞̃}
in the 𝑁 space can be expressed as a linear combination of these eigenvectors
{𝑞̃} = ∑𝑁 (𝑟)
𝑟=1{𝑤 } 𝑝̅𝑟 = [𝑤] {𝑝̅ }. (3.195)
Substituting (3.195) into equation (3.187) and using equations (3.191) to
(3.194) we obtain the damped principal coordinates
𝑇 ̂
{𝑤(𝑟) } {𝑓 }
𝑝̅𝑟 = ̅ 2 ̅̅̅ (3.196)
𝐾𝑟 −𝜔 𝑀𝑟
The solution of equation (3.187) is
𝑇
̂ } {𝑤(𝑟) }
{𝑤(𝑟) } {𝑓
{𝑞̃} = ∑𝑁
𝑟=1 ̅̅̅ (3.197)
𝑀𝑟 (𝜆̅ 𝑟 −𝜔2 )
where
𝐾̅
𝜆̅𝑟 = 𝑀̅𝑟 = 𝜔𝑟2 (1 + 𝑖𝑔𝑟 ) (3.198)
𝑟

in which 𝜔𝑟 is the frequency of resonance and 𝑔𝑟 – the hysteretic damping factor


of the 𝑟-th mode.
For single point excitation, the complex receptance is
𝑞
̃ 𝑤ℓ(𝑟) 𝑤(𝑗𝑟)
𝛼̅𝑗ℓ =
𝑗
𝑓̂ ℓ
= ∑𝑁
𝑟=1 𝜔2
. (3.199)
̅̅̅
𝑀 2
𝑟 𝜔𝑟 (1− 2+𝑖𝑔𝑟 )
𝜔𝑟
Denoting
(𝑟) (𝑟)
𝑤ℓ 𝑤𝑗 (𝑟) (𝑟)
̅𝑟 𝜔𝑟2
= 𝑥𝑗 + 𝑖𝑦𝑗 (3.200)
𝑀

expression (3.199) becomes

116
(𝑟 ) (𝑟 ) (𝑟 ) (𝑟)
𝑥𝑗 +𝑖𝑦𝑗 (𝑥𝑗 +𝑖𝑦𝑗 ) 𝑒−𝑖𝜓𝑟
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 𝜔2
= ∑𝑁
𝑟=1 (3.201)
1− 2 +𝑖𝑔𝑟 2 2
𝜔 𝑟 √(1−𝜔2) +𝑔𝑟2
𝜔𝑟

in which
𝑔𝑟
𝜓𝑟 = 𝑡𝑎𝑛−1 𝜔2
. (3.202)
1− 2
𝜔 𝑟

117
CHAPTER 4

Principles of the Identification of Vibrating


Systems

4.1. SCHEME OF THE DYNAMIC IDENTIFICATION PROCESSES

A large class of identification methods of mechanical systems are based on


the simplified scheme from Fig.4.1. The response of the actual system is compared
to the response of the analytical model to the same excitation and adjustments are
made in the parameters to be identified until its response matches that of the system
to some prescribed accuracy.

The actual system is subjected to known excitation – periodic, transient or


random – and the response is measured at the points of interest. The measurement
is carried out either in the “time domain” – using vibrograms and correlograms,
which describe the time history of the system, or in the “frequency domain” –
obtaining frequency response functions or spectrograms (Fig.4.2).

118
The vibrograms or the graphs of the frequency response functions provide
quantitative information only about the particular test configuration, being strictly
limited to the tested structure, to the applied loading and to the coordinates of
response measurement. It is often of interest to predict the response at the measured
coordinates under different loading conditions, or due to changes in the structure
configuration or supporting conditions. The effect of such modifications on the
dynamic response is estimated using a model of the structure.

The pure analytical modeling has been developed in recent years,


especially due to the use of digital computers, which allow an efficient analysis of
a large number of solutions of the same problem. However, pure theoretical models
are largely dependent on the analist’s intuition and experience, having in general a
qualitative character, unable to predict precisely the results of actual tests.
The aim of system identification is to construct models based on test data
obtained by measurements on the actual structure. Assumptions are also made on
the analytical representation of the structure. Part of these are shown in Fig.4.3,
underlining those adopted in the following chapters of this book.
For a large range of vibration regimes, the behavior of elastic structures can
be adequately described by a set of ordinary linear differential equations of the form
(3.26) or (3.43). These equations represent the analytical model. Solving them, for
loading conditions identical to those used during the testing on the actual system,

119
we can obtain the model response, for example, under the form of frequency
response functions.
The model parameters are estimated comparing the response of the actual
system to the model response (Fig.4.1). This can be done either by direct
verification, considering zero error, or by statistical processing. In the last case, the
difference between the various identification methods consists of the error
minimization criterion between test and estimated values.

120
Table 4.1
Classification of excitations used in structural identification
Electromagnetic or
electrohydraulic exciters
Constant amplitude harmonic forces
Mechanical exciters
acting through elastic
Harmonic element
excitation
Mechanical exciters with
Harmonic forces with amplitude rotating eccentric masses
proportional to frequency squared
Inertial exciters with
masses in translation

Constant amplitude harmonic Vibrating tables with


displacements control of displacement
Imposed Loading by winch drawn
excitation Initial displacement and snapback cables or hydraulic jacks,
then sudden release
Impulsive loads by
hammer or free falling
Initial velocity followed by free masses, explosive
Transient
vibration charges and impacting
excitation
or shock devices, use of small
rockets, etc.
Electromagnetic exciters
Harmonic force of constant amplitude controlled by signal
and variable frequency generators with rapid
frequency sweep
Harmonic force with amplitude Exciters with eccentric
proportional to frequency squared and masses, brought to full
variable frequency speed then running down

Wide bandwidth random force Electromagnetic exciters


Random controlled by random
excitation signals from signal
Narrow bandwidth random force
generators

Periodic Gas and liquid pulsations in compressor and pump pipes.


steady-state Forces due to unbalanced rotors, misaligned shafts, couplings, gears,
Ambient excitation bearings and constant speed shafts
excitation
Transient Microseisms, wind loading
excitation

Random Random forces due to the roughness of the road surface, air
excitation turbulence, lack of homogeneity and evenness of workpieces
fabricated on machine tools, normal operation of engines, etc.

121
4.2. EXPERIMENTAL DETERMINATION OF THE DYNAMIC
CHARACTERISTICS OF ACTUAL SYSTEMS

According to the above presentation, the identification of vibrating systems


is based on the mathematical processing of some experimental results, determined
to provide maximum information on the dynamic behavior of the analyzed system.
The structural excitation methods can be classified as in Table 4.1.

4.2.1. Imposed Excitation Identification Methods

In most identification methods, the system is subjected to a known external


excitation, sometimes referred to as a “test signal” which facilitates the
interpretation of results.
During testing it is important to minimize (even eliminate) the influence of
other perturbing sources. Also, the equipment used to excite the structure and to
measure its response should not modify considerably its parameters.
Such methods are used for grounded structures, stationary vehicles (e.g.
ground testing of airplanes), non-operating machines and equipment (e.g. machine
tools while the tool does actually not remove the metal). The characteristics of
structures tested in these conditions may differ from those determined in normal
operating conditions, especially for nonlinear structures. However, selection of
adequate excitation levels, preloading and operation of some subassemblies yields
satisfactory results.
The dynamic characteristics of a single input 𝑥(𝑡) single output 𝑦(𝑡) linear
system (Fig.4.4) can be defined in the time domain by the impulse response function
ℎ(𝑡) and in the frequency domain by the frequency response function 𝐻(𝑖𝜔) (or
the transfer function) which form a pair of Fourier transforms

𝐻(𝑖𝜔) = ∫−∞ ℎ(𝑡)𝑒 −𝑖𝜔𝑡 𝑑𝑡. (𝑡 > 0) (4.1)
In the case of deterministic excitation, the response is given by the
convolution integral
𝑡
𝑦(𝑡) = ∫0 ℎ(𝑡 ′ ) 𝑥(𝑡 − 𝑡 ′ ) 𝑑𝑡 ′ = 𝑥(𝑡) ∗ ℎ(𝑡) , (4.2)
so that the weighting function ℎ(𝑡) represents the response to a unit impulse
excitation 𝑥(𝑡) = 𝛿𝑡 (Dirac’s 𝛿 function). Practically, however, this excitation is
unrealizable and very short impulses have not enough energy to produce a
measurable response.
The frequency response function has the expression
𝑌(𝑖𝜔)
𝐻(𝑖𝜔) = , (4.3)
𝑋(𝑖𝜔)

where 𝑋(𝑖𝜔) and 𝑌(𝑖𝜔) are the excitation and response Fourier transforms
∞ ∞
𝑋(𝑖𝜔) = ∫−∞ 𝑥(𝑡)𝑒 −𝑖𝜔𝑡 𝑑𝑡 , 𝑌(𝑖𝜔) = ∫−∞ 𝑦(𝑡)𝑒 −𝑖𝜔𝑡 𝑑𝑡 , (4.4)

122
in which the lower limit is zero for actual systems.
a) For harmonic excitation, substituting functions of the form
𝑥(𝑡) = 𝑥̂𝑒 𝑖𝜔𝑡 , 𝑦(𝑡) = 𝑦̂𝑒 𝑖(𝜔𝑡+𝜑) (4.5)
into expressions (4.4) and (4.3) we obtain
𝑦(𝑡) 𝑦̂
𝐻(𝑖𝜔) = 𝑥(𝑡) = 𝑥̂ 𝑒 𝑖𝜑 . (4.6)

Therefore, the modulus of the frequency response function |𝐻(𝑖𝜔)| can be


obtained from the magnitude-versus-frequency characteristic (𝑦̂⁄𝑥̂) − 𝜔 , and the
phase angle of this function 𝜑 - from the phase-versus-frequency characteristic
𝜑 − 𝜔 . These diagrams can be experimentally determined, using excitation with
constant amplitude and variable frequency. Both graphs can be drawn either point
by point, performing measurements at discrete frequencies in steady-state
conditions, or continuously, in quasi steady-state conditions, using a frequency
sweep slow enough to allow the response to reach the steady-state at each
frequency, especially near lightly damped resonances. Figure 4.5 shows a typical
example of such diagrams for a system with two resonances in the frequency range
of measurements.

Modern measurement systems use a transfer function analyzer. The signals


𝑥(𝑡) = 𝑥̂ 𝑠𝑖𝑛 𝜔𝑡, 𝑦(𝑡) = 𝑦̂ 𝑠𝑖𝑛 (𝜔𝑡 + 𝜑) , and a reference signal 𝑧(𝑡) = 𝑧̂ 𝑐𝑜𝑠 𝜔𝑡
are fed at the input. The multiplier of the analyzer performs the products
2 𝑦̂ 𝑦̂ 𝑦̂
𝑥(𝑡)𝑦(𝑡) = 2 𝑠𝑖𝑛 𝜔𝑡 𝑠𝑖𝑛 (𝜔𝑡 + 𝜑) = 𝑐𝑜𝑠𝜑 − 𝑐𝑜𝑠(2𝜔𝑡 + 𝜑),
𝑥̂ 2 𝑥̂ 𝑥̂ 𝑥̂
2 𝑦̂ 𝑦̂ 𝑦̂
𝑥̂𝑧̂
𝑧(𝑡)𝑦(𝑡) = 𝑥̂ 2 𝑐𝑜𝑠 𝜔𝑡 𝑠𝑖𝑛 (𝜔𝑡 + 𝜑) = 𝑥̂ 𝑠𝑖𝑛𝜑 + 𝑥̂ 𝑠𝑖𝑛(2𝜔𝑡 + 𝜑).
The constant terms in these products are the real and imaginary parts of the
frequency response function (4.6) whose diagrams are shown in Fig.4.6.
The variable terms are eliminated with low-pass filters, averaging the
products over an integer number of excitation cycles. The signals of the two
components are fed to the inputs of an 𝑋-𝑌 recorder, and polar (Nyquist) plots of
the response (Fig.4.7) are obtained using a frequency sweep.

123
It is appreciated that these methods offer the highest precision, requiring
the less expensive equipment, but the measurements and the assembly of the
instrumentation set up take a long time, requiring frequent calibrations. Because at
present the methods of experimental identification with harmonic excitation are the
most widely used, their presentation will cover almost the entire book.
b) With the fast Fourier transform and digital computers, the frequency
response functions can be computed based on equation (4.3) using excitation with
transient signals or impulses.
The simplest form are the rectangular, triangular, trapezoidal, or semi-
sinusoidal impulses which can be produced by electronic signal generators coupled
with electrodynamic exciters via power amplifiers. Because the amplitude spectra
of these impulses have zeroes at certain frequencies (Fig.4.8) and theoretically span
an infinite frequency range, in the case of systems with many degrees of freedom
some resonances may not be excited, while resonances outside the range of interest
may be excited.

These drawbacks are eliminated using excitation with sinusoidal signals of


constant amplitude and rapid frequency sweeps [135]. A signal of the form
𝑥(𝑡) = 𝑥𝑜 𝑠𝑖𝑛 (𝑎𝑡 2 + 𝑏𝑡), (0 < 𝑡 < 𝑇)
where 𝑇 is the time duration over which the signal frequency varies continuously
from an initial value 𝑓1 to a final value 𝑓2,

124
𝜋
𝑎 = 𝑇 (𝑓2 − 𝑓1 ), 𝑏 = 2𝜋𝑓1,
has an essentially uniform frequency spectrum over the limited frequency band
(𝑓2 − 𝑓1 ), being adequate for the identification of elastic structures (Fig.4.9).
In studies on reduced scale models, the excitation with a special hammer
equipped with a force transducer and an accelerometer is used to obtain the signals
𝑥(𝑡) and 𝑦(𝑡). The frequency response function is determined with a real time
Fourier analyzer, by division of the Fourier transforms of the response and
excitation time histories [67].
Transient test techniques require a minimum measurement time. Even
when using exciters, the assembly of the instrumentation set up is simplified [135].
On the contrary, the measurement equipment is much more expensive. The
dynamic range of the instrumentation must be larger than that required for harmonic
testing. During signal recording, the amplifiers and analog-to-digital converters
must not be overloaded. Generally, at the impact tests the signal-to-noise ratios are
lower than those obtained using other types of excitation.
c) Random excitation techniques are often used in system identification.
Wide-band signal generators are used to this purpose, which connected to
electrodynamic exciters generate an excitation with a uniform frequency spectrum
over the whole band of frequencies of interest. Thus, all spectral components are
simultaneously excited and a real time spectral analysis is performed displaying the
results simultaneously on an oscilloscope.
The frequency response function is calculated as
𝑆𝑥𝑦 (𝑖𝜔)
𝐻(𝑖𝜔) = 𝑆𝑥𝑥 (𝜔)
, (4.7)

where 𝑆𝑥𝑥 (𝜔) is the power spectral density of the excitation (stationary and
ergodic) and 𝑆𝑥𝑦 (𝑖𝜔) is the cross spectral density function between the excitation
and response.
If 𝑆𝑥𝑥 (𝜔) = 𝑆𝑜 = 𝑐𝑜𝑛𝑠𝑡. through the whole range of frequencies of
interest (“white noise” technically realizable), the power cross spectral density is
proportional to the frequency response function 𝑆𝑥𝑦 (𝑖𝜔) = 𝑆𝑜 𝐻(𝑖𝜔) and the cross-
correlation function of the excitation and response

𝑅𝑥𝑦 (𝜏) = ∫−∞ 𝑆𝑥𝑦 (𝑖𝜔)𝑒 𝑖𝜔𝜏 𝑑𝜔 (4.8)
is proportional to the impulse response function ℎ(𝑡).
When the information on the variation of phase angle with frequency is of
no interest, we can use the relation
𝑆𝑦𝑦 (𝜔)
|𝐻(𝑖𝜔)|2 = , (4.9)
𝑆𝑥𝑥 (𝜔)

where 𝑆𝑦𝑦 (𝜔) is the power spectral density of the response.


Another method is based on the use of excitation with pseudo-random
signals. In this case, the signals are either produced by a specially designed

125
electronic generator or synthesized with a fast Fourier analyzer which performs the
inverse transformation of a given spectral function.
Tests with random input signals have the advantage of exciting all spectral
components within the frequency range of interest, which permits the identification
of structures with time-variable parameters. Using an exponential averaging at the
calculation of 𝑆𝑥𝑥 (𝜔) and 𝑆𝑥𝑦 (𝑖𝜔), we can perform a real time analysis, hence an
“on-line” observation of these modifications. The procedure is particularly useful
for the optimization of a structural response by modifications of the distribution of
masses, stiffnesses and damping.
Generally, the equipment used in this type of measurements is more
expensive than that used in harmonic testing, the signal-to-noise ratio is lower (but
can be improved increasing the averaging time) and the attainable response level is
limited due to the insufficient energy density.

4.2.2. Identification Methods Using Excitation from Normal Operation

In some cases it is preferable to use identification in normal operating


conditions or with ambient excitation. For example, the damping of an airplane in
flight or during tests in the wind tunnel is different from the damping during the
ground testing, both in the physical nature and distribution. The lateral vibrations
of flexible whirling rotors are influenced by gyroscopic effects which do not occur
for a non-rotating rotor. Also, structural nonlinearities (due to clearances, couplings
and slides) act differently at different levels of the forces arising during the normal
operation.
The “active” excitation (using
test signals), especially the harmonic
excitation, increases the level of structural
response beyond the admissible limits,
with the danger of fracture or damage
during testing. In other cases, the active
excitation is not sufficient, as for bridges
and large buildings, or massive
foundations, when the seismic or wind
excitation produce responses of
measurable levels of the entire structure.
Good results are obtained
calculating the frequency response function from expression (4.7). Using the Fast
Fourier Transform on digital computers, the functions 𝑆𝑥𝑦 (𝑖𝜔) and 𝑆𝑥𝑥 (𝜔) are
obtained from the cross-correlation function
1 𝑇 ⁄2
𝑅𝑥𝑦 (𝜏) = lim 𝑇 ∫−𝑇⁄2 𝑥(𝑡) 𝑦(𝑡 + 𝜏) 𝑑𝑡
𝑇→∞
and the autocorrelation function

126
1 𝑇 ⁄2
𝑅𝑥𝑥 (𝜏) = lim 𝑇 ∫−𝑇⁄2 𝑥(𝑡) 𝑥(𝑡 + 𝜏) 𝑑𝑡 ,
𝑇→∞

from the signals 𝑥(𝑡) and 𝑦(𝑡) measured at two points on the structure (Fig.4.10).
The method is based on the relations (4.8) and (4.10)

𝑆𝑥𝑥 (𝜔) = ∫−∞ 𝑅𝑥𝑥 (𝜏) 𝑒 −𝑖𝜔𝜏 𝑑𝜏. (4.10)
The same operations can be performed using real time spectrum analyzers,
based on the Fast Fourier Transform.
The extent of the frequency response range and the amplitude of the
nonlinear response can be evaluated using the coherence function
2 1⁄2
|𝑆𝑥𝑦 (𝑖𝜔)|
(𝑆 ) .
𝑥𝑥 (𝜔) 𝑆𝑦𝑦 (𝜔)

Apart from the above presented methods, some ad-hoc excitation methods
are being used in practice, especially when the equipment includes variable speed
motors. The modification of the operating parameters, especially the speed, allows
the application of the identification methods based on the analysis of the response
in the neighborhood of structural resonances.

4.2.3. Identification Using Adjustable Models

The identification methods using adjustable models, though widely used to


the identification of automatic processes, still have a reduced application to
mechanical systems. In principle, a physical model of the system is used – e.g.: a
transfer function generator – and a signal identical to the excitation signal of the
studied structure is fed at its input. The output signal of the model is compared with
the response signal of the structure and their difference is used to adjust the model.
Usually, the structure of the model is fixed and only the parameters are
modified – manually or automatically – until the minimization of an error
functional, ensuring a rapid convergence of the process. If the adjustment of model
parameters is performed automatically – it is called an adaptive model. If the model
is provided with a memory, so that, based on the information on the time
modification of the structural parameters, it adopts an adequate strategy of
parameter adjustment, it is called a learning model.
These methods have the advantage of application in normal operating
conditions, allowing the direct building of a high accuracy mathematical model.
Their large scale use is delayed by the still high cost of the analog and digital
equipment necessary for practical application.

4.2.4. Nodal Pattern Analysis

A systematic study of the natural modes of vibration of plate-like structural


elements, especially of impeller blades, bladed discs, turbine and axial compressor
blades, aircraft and marine propeller blades, is carried out by the analysis of nodal

127
patterns, referred to as the Chladni figures, after the name of their originator [23]
in 1787.

For example, if a free circular plate (or a rotor disc) is excited by a harmonic
force and fine sand (or semolina) is sprinkled on the horizontal surface, at each
resonance the sand will align along the nodal lines (lines with zero amplitude of
vibration) which, according to the theory of vibration of thin plates, are diameters
and circles, as shown in Fig.4.11.
A systematic representation of the resonance frequencies can be done on a
graph as in Fig.4.12, in which they are grouped in families of modes having the
same number of nodal circles.
For the experimental determination of natural frequencies, first, points are
plotted at the value of the resonance frequency (in ordinate) and the number of
nodal diameters (in abscissae). Continuous lines are then drawn through the points
corresponding to the same number of nodal circles. If one of the resonance
frequencies has been omitted, it can be determined at the intersection of a curve
with the vertical line corresponding to a given number of nodal diameters.
As shown in Fig.4.2, it is possible that different modes of vibration have
(approximately) the same resonance frequency (linked by dashed lines). In this
case, the two independent modes of vibration are excited simultaneously resulting
a compounded mode whose nodal lines are no more circles or diameters.
For example, the natural mode of vibration with 1 circle and 2 diameters,
labelled symbolically “mode 1/2” and the mode with 0 circles and 5 diameters,
labelled “mode 0/5” have approximately the same natural frequency, giving rise to
a “compounded mode 1/2+0/5”, shown in Fig.4.13,a.

128
The manner in which these lines are formed may be demonstrated using
the following graphical construction. Figures 4.13,b and c show the modes 1/2 and
0/5 in which the shaded areas are considered to move ‘down’ and the unshaded area
move ‘up’. Overlapping the two figures (Fig.4.13,d), the doubly shaded are
supposed to move ‘down’ while the unshaded areas move ‘up’. The single shaded
areas contain the points where the two motions cancel.

The nodal lines pass between the doubly ‘down’ and doubly ‘up’ areas and
will occur only in the single shaded areas. Their exact courses in these areas will
depend on the relative amplitude of the vibration in the two constituent modes of
vibration and on the uniformity of the plate.
According to the phase relation of the two modes, there are always two
possible forms of combination of the modal lines of a compounded mode. If the
motion in the mode 1/2 is as shown in Fig.4.13,e (in opposite phase to that shown
in Fig.4.13,b), then the combination with the mode 0/5 (Fig.4.13,c) gives rise to the
compounded mode from Fig.4.13,f , identical with the mode of Fig.4.13,d but
rotated 180°.
The systematic study of the simple and
compounded modes of vibration of cantilever flat
plates enabled a understanding of the compounded
modes of vibration of turbine blades and a correct
interpretation of their nodal lines.
Thus, denoting 𝑚 - the number of nodal lines
chordwise (along the blade width), 𝑛 - the number of
nodal lines lengthwise, labelling the modes of
vibration by the ratio 𝑚⁄𝑛, we obtain the
compounded modes from Fig.4.14 determined
experimentally.
Diagrams of the natural frequency versus the
number of diameters, with curves of constant number
of nodal circles (Fig.4.12) allow the construction “on
the paper” of the modes undiscovered in experiments,
which can be afterwards put into evidence by repeated
tests near resonance.
Resolving the compounded modes into the
constituent modes allows a useful comparison

129
between the dynamic behavior of a blade with modified shape and the blade with
the initial design. Otherwise, due to the highly modified pattern of the nodal lines,
it would be impossible to make non-erroneous comparisons between the different
compounded modes.
A presentation of the equipment used in the identification of elastic systems
is given in [19].

4.3. STRUCTURE OF THE ANALYTICAL MODELS

4.3.1. Discrete Models

Real structures are generally continuous systems, with distributed


parameters. Their configuration should be defined by an infinite number of
generalized coordinates, hence by an infinite number of degrees of freedom.
From a practical point of view, usually, of interest is the motion of only a
few significant points of the structure, defined by the variation of a finite number
of generalized coordinates. Hence the necessity and utility of the models with a
finite number of degrees of freedom, as well as the justification of using a set of
equations of motion of the form (3.26) or (3.43).
In general, the characteristics of a model are primarily dependent on the
way the studied structure is loaded. It is never required that the structure behaves
from all points of view identically with the actual structure. In each practical
application, of interest is to model the response in a limited number of points on the
structure and within a limited frequency range of the excitation forces.
The measurement points should be representative for the structure
response. For example, in the modal synthesis, they must include the connecting
points between different subsystems. In structural optimization they must include
the points where there mass or stiffness modifications are possible. Generally, the
response and excitation points should be selected to obtain a comprehensive image
of the dynamic deformed shape of the structure. However, selection of too many
points brings difficulties, especially at the transformation from the modal model to
the spatial model.

4.3.2. Modal Models

Most identification methods described in this book aim to define a modal


model. Summarizing the analysis of systems with proportional viscous damping
presented in Chapter 3, from a mathematical point of view, a modal model consists
of three types of equations:
a) A differential equation for the motion in each mode
𝑚𝑟 𝑝̈𝑟 + 𝑐𝑟 𝑝̇𝑟 + 𝑘𝑟 𝑝𝑟 = 𝐹𝑟 , (𝑟 = 1, 2, . . . , 𝑁) (4.11)
which shows that the system responds in each mode as a completely decoupled
single degree of freedom system.

130
b) An equation of coordinate transformation
(𝑟)
𝑞𝑗 = ∑𝑁
𝑟=1 𝛹𝑗 𝑝𝑟 , (𝑗 = 1, 2, . . . , 𝑃)) (4.12)
which shows that the motion at the points of interest of the actual system can be
computed by summing up the motions in all modes, multiplied by the associated
modal shape coefficient.
c) An equation of force transformation
(𝑟)
𝐹𝑟 = ∑𝑃ℓ=1 𝛹ℓ 𝑓ℓ , (𝑟 = 1, 2, . . . , 𝑁)) (4.13)
which shows that the force “acting on each mode” is equal to the sum of external
forces acting on the structure, multiplied by the associated shape coefficients.
Similar complex quantities are defined for nonproportional damping (see
Chapter 3).
Within the limitations to linear non-gyroscopic systems, with time
invariable parameters, the damping level is important for the structure of the
analytical model. The following four cases may be considered:
a) For very lightly damped structures, for which the damping ratio (fraction
of critical damping) is less than 0.05, the damping can be neglected, and the motion
is described by equations (3.15).
The harmonic response can be expressed in terms of the modal parameters
and the real “classical” modes of vibration, under the form (3.24)
𝑟 𝑇
̂ } {𝛹 𝑟 }
{𝛹 } {𝑓
( ) ( )

{𝑞} = ∑𝑁
𝑟=1 𝑚 (𝜔2 −𝜔2 ) 𝑒 𝑖𝜔𝑡 , (4.14)
𝑟 𝑟

where 𝜔𝑟 are the natural frequencies, 𝑚𝑟 – modal masses, {𝛹 (𝑟) } – the modal
vectors.
b) For lightly damped structures, the off-diagonal elements of the matrix
[𝛹]𝑇 [𝐶][𝛹] are neglected, and the response can be expressed in terms of the
undamped modes of vibration (3.42)
𝑟 𝑇
̂ } {𝛹 𝑟 }
{𝛹 } {𝑓
( ) ( )

{𝑞} = ∑𝑁
𝑟=1 𝑚 (𝜔2 −𝜔2 +𝑖 2𝜁 𝜔𝜔 ) 𝑒 𝑖𝜔𝑡 , (4.15)
𝑟 𝑟 𝑟 𝑟
𝑐𝑟
where 𝜁𝑟 = is the modal damping ratio.
2𝑚𝑟 𝜔𝑟
c) For moderately damped structures or substructures, the damping
coupling cannot be neglected, the damping is nonproportional, and the response
can be expressed either in terms of the real forced modes of vibration (3114)
(𝑟) 𝑇
{𝛷 ̂ }{𝛷(𝑟) }
} {𝑓
{𝑞} = 𝑇 𝑒𝑖𝜔𝑡 (4.16)
(𝑟) (𝑟) (𝑟) 𝑇 (𝑟)
{𝛷 } [𝐾−𝜔2 𝑀] {𝛷 } +𝑖 {𝛷 } [𝜔𝐶+𝐻] {𝛷 }

or in terms of the complex modes of vibration (3.182)

131
𝑇
{𝑞 (𝑟) } {𝑓̂ }{𝑞 (𝑟) }
{𝑞} = ∑2𝑁
𝑟=1 𝑒 𝑖𝜔𝑡 (4.17)
𝑢𝑟 (𝑖𝜔−𝜎𝑟 )

where
1 1
𝜎𝑟 = 𝜔𝑟 (− 2 𝜏𝑟 + 𝑖 √1 − 4 𝜏𝑟2 ),

and 𝜏𝑟 and 𝑢𝑟 are defined by (3.164) and (3.160).


d) For highly damped components, for which individual modes of vibration
cannot be distinguished, the frequency response is expressed as a rational fraction
of two polynomials
𝑞̃𝑗 ∑𝑚 𝑃𝑟 (𝑖𝜔)𝑟
𝛼̅𝑗ℓ = = 1+∑𝑟=0
𝑛 (4.18)
𝑓̂ℓ 𝑟=1 𝑄𝑟 (𝑖𝜔)
𝑟

in which the coefficients 𝑃𝑟 and 𝑄𝑟 are determined using iterative methods.

4.3.3. Nature of Analytical Models

The parameters which define the response of a continuous structure can be


classified as: measurable (observable) and abstract (intuitive).
Measurable parameters can be determined (at least in principle) by direct
measurements on the actual structure. They are independent of the structure of the
analytical model.
For example, the elements of the influence coefficient matrix [𝛿] = [𝐾]−1
are measurable. By definition, the element 𝛿𝑖𝑗 of the matrix [𝛿] represents the
displacement of coordinate 𝑖 due to a unit force applied at coordinate . This
quantity, which is measurable, is not dependent on the form of the analytical model.
In the same category can be included the modal vectors {𝛹 (𝑟) }, the modal
parameters 𝑚𝑟 , 𝑘𝑟 and 𝑐𝑟 , hence the frequency response functions like the
receptance, mobility, inertance, etc.
The measurable parameters can be expressed in terms of the modal vectors.
Thus, [ˋ𝑘ˏ] = [𝛹]𝑇 [𝐾] [𝛹]. Their analytical determination involves the solution of a
differential equation.
Abstract parameters are obtained directly from the material properties,
hence are referred to as intuitive. They cannot be determined by measurements on
the actual structure; their values depend on the specific formulation of the analytical
model.
For example, the elements of the matrices [𝑀], [𝐶] and [𝐾] cannot be
measured. The element 𝑘𝑖𝑗 of the matrix [𝐾] is defined as the internal force at
coordinate 𝑖 due to a unit displacement of coordinate 𝑗 when all coordinates except
𝑗 are constrained from motion. Such constraints are impossible to duplicate in the
laboratory and thus 𝑘𝑖𝑗 is not directly measurable. The values of the elements 𝑘𝑖𝑗
depend, for example, on the number and location of the points selected as
coordinates at the model definition.

132
Analytically, the intuitive parameters can be expressed in terms of the
inverse of a modal matrix. Thus, [𝐾] = [𝛹]−𝑇 [ˋ𝑘ˏ] [𝛹]−1 . They can be obtained from
the mathematical analysis of the model and have an intuitive physical meaning.
Their transformation into measurable quantities requires a matrix inversion or the
solution of a differential equation.
In conclusion, ”one cannot measure the parameters that can be modeled
and one cannot directly model what can be measured” [10].

4.3.4. Transformation from the Modal to the Spatial Model

As already mentioned, many identification methods are limited to the


determination of natural frequencies, modal vectors and modal parameters – hence
of a modal model. Methods which derive a spatial model from the modal model or
directly from a response model encounter difficulties arising from the inversion of
numerically ill-conditioned matrices and from the non-uniqueness of the matrices
of the intuitive physical parameters.

4.3.4.1. Computational difficulties

The identification of vibrating systems would be much simplified if no


complications existed at the inversion of experimentally determined poorly
conditioned matrices. We can illustrate this with a simple example.
The equations of the harmonic vibrations of an undamped system are
([𝐾] − 𝜔2 [𝑀]) {𝑞̂} = {𝑓̂} . (4.19)
Consider the receptance matrix [𝛼] measured at two frequencies 𝜔1 ≠ 𝜔2
−1
[[𝐾] − 𝜔12 [𝑀] ] = [𝛼(𝜔1 )] , (4.20)
−1
[[𝐾] − 𝜔22 [𝑀] ] = [𝛼(𝜔2 )] . (4.21)

Inverting the two measured matrices [𝛼(𝜔1 )] and [𝛼(𝜔2 )] we obtain


[𝐾] − 𝜔12 [𝑀] = [𝛼(𝜔1 )]−1 , (4.22,a)
[𝐾] − 𝜔22 [𝑀] = [𝛼(𝜔2 )]−1 , (4.22,b)
wherefrom we can determine matrices [𝐾] and [𝑀] .
Except for some particular cases (systems with a small number of degrees
of freedom), the method cannot be used because the receptance matrices cannot be
inverted, especially for a large number of coordinates. In this case, the range of
natural frequencies is so wide that the requirements regarding the accuracy of
measurements cannot be met.
Indeed, from the spectral development of a matrix it is known that the
smallest eigenvalue of a matrix becomes the largest eigenvalue of the inverse
matrix.

133
Using the notation [𝐾]−1 = [𝛿], equations (3.5) can be written
[𝑀]−1 [𝐾]{𝛹 (𝑟) } = 𝜔𝑟2 {𝛹 (𝑟) } , (4.23,a)
1
[𝛿] [𝑀]{𝛹 (𝑟) } = {𝛹 (𝑟) } . (4.23,b)
𝜔𝑟2

The right eigenvectors of the matrices [𝑀]−1 [𝐾] and [𝛿] [𝑀] are identical.
The dominant mode of the matrix [𝑀]−1 [𝐾] is the mode with the highest natural
frequency 𝜔𝑟 , while the dominant mode of the matrix [𝛿] [𝑀] is the mode with the
lowest frequency 𝜔𝑟 (since the contribution of each mode is inversely proportional
to the square of its natural frequency).
Using the modal matrix (3.12), equations (4.23) become
[𝑀]−1 [𝐾] [𝛹] = [𝛹] [ˋ𝜔𝑟2 ˏ] , (4.24,a)
1
[𝛿] [𝑀] [𝛹] = [𝛹] [ˋ 2 ˏ] . (4.24,b)
𝜔 𝑟

Using equation (3.21) we obtain


𝜔2 𝜔2
(𝑟) 𝑇
[𝐾] = [𝑀] [𝛹] [ˋ 𝑟 ˏ] [𝛹]𝑇 [𝑀] = ∑𝑁
𝑚
𝑟
𝑟=1 𝑚 [𝑀] {𝛹 } {𝛹 (𝑟) } [𝑀], (4.25)
𝑟 𝑟

1 (𝑟) 1 𝑇
[𝛿] = [𝛹] [ˋ 2 ˏ] [𝛹]𝑇 = ∑𝑁
𝜔 𝑚 𝑟=1 𝜔2 𝑚 {𝛹 } {𝛹 (𝑟) } . (4.26)
𝑟 𝑟 𝑟 𝑟

The dominant terms in the [𝐾] matrix come from the high frequency modes
(since 𝜔𝑟 is in the numerator), while the dominant terms in the [𝛿] matrix come
from the low frequency modes (since it is inversely proportional to 𝜔𝑟2 ).
If the measurable matrix [𝛿] is ill conditioned and does not contain useful
information about the high frequency modes, its inverse [𝐾] also cannot, and thus
the elements of the [𝐾] matrix will have non-meaningful values. Berman [10] has
𝜔
shown that if the ratio of the high and low natural frequencies is 𝜔𝑁 = 100 , then
1
𝜔 2 4
( 𝑁) = 10 and the precision of the test derived values 𝛿𝑖𝑗 must be of the order
𝜔1
4
1: 10 in order to have a meaningfully invertible matrix [𝛿].

4.3.4.2. Non-uniqueness of the proportional damping matrix

In the case of proportional viscous damping, based on equation


[𝑐] = [𝛹]−𝑇 [ˋ2𝜁𝑟 𝜔𝑟 𝑚𝑟 ˏ] [𝛹]−1 (4.27)
obtained from (3.28), (3.36) and (3.41), it is theoretically possible to determine the
elements of the matrix [𝑐] if the modal matrix [𝛹], the resonance frequencies 𝜔𝑟 ,
the modal damping ratios 𝜁𝑟 and the modal masses 𝑚𝑟 are known.
To simplify the following presentation, it will be considered that the modal
vectors are normalized to unit modal mass 𝑚𝑟 = 1 (𝑟 = 1, 2, . . , 𝑁) so that equation
(4.27) becomes

134
[𝑐] = [𝛹]−𝑇 [ˋ2𝜁𝑟 𝜔𝑟 ˏ] [𝛹]−1 . (4.28)
1
However, because an 𝑁𝑥𝑁 symmetric matrix can have 2 𝑁(𝑁 + 1) distinct
elements, but only 𝑁 equations (4.28) can be derived, any calculation method of
the matrix [𝐶] based on the values 𝜁𝑟 and 𝜔𝑟 will not provide a sufficient number
of equations to uniquely determine the elements of the damping matrix. There is
only one exception to this, when [𝐶] is a diagonal matrix, hence when the dashpots
which model the energy dissipation (in the spatial model) are connected between
each mass and a fixed reference, and there is no internal damping.
The simplest form of proportional viscous damping is described by
equation (3.34) in which the matrix [𝐶] has the form
[𝐶] = 𝜎 [𝑀] + 𝜏 [𝐾] . (4.29)
Equations (4.28) and (4.29) yield
[ˋ2𝜁𝑟 𝜔𝑟 ˏ] = 𝜎 [𝐼] + 𝜏 [ˋ𝜔𝑟2 ˏ] (4.30)
where [𝐼] is the identity matrix.
If the damping ratios 𝜁1 and 𝜁2 of two modes of vibration with natural
frequencies 𝜔1 and 𝜔2 are determined experimentally, then the constants 𝜎 and 𝜏
can be calculated as [50]
2𝜔1 𝜔2 (𝜔1 𝜁2 −𝜔2 𝜁1 )
𝜎= , (4.31)
𝜔12 −𝜔22
2 (𝜔1 𝜁1 −𝜔2 𝜁2 )
𝜏= . (4.32)
𝜔12 −𝜔22

With this approach, the damping rations of the higher modes (𝑟 > 2) are
forced to be
𝜎+𝜏 𝜔𝑟2
𝜁𝑟 = (4.33)
2 𝜔𝑟

which varies in an approximately a linear manner with frequency.


The drawback is that, in this method the values (4.33) are imposed to the
higher modes, so that negative damping ratios can occur in some modes of
vibration.
In a more general form, the matrix [𝐶] is obtained by summing powers of
the mass and stiffness matrices [50].
First, the mass matrix is partitioned into
[𝑀] = [𝐴]𝑇 [𝐴] (4.34)
where [𝐴] is a triangular matrix. Then, using the coordinate transformation
{𝑞} = [𝐴]−1 {𝜉} (4.35)
and substituting (4.35) into (3.26), after premultiplication by [𝐴]−𝑇 we obtain
[𝐼] {𝜉̈ } + [𝐴]−𝑇 [𝐶] [𝐴]−1 {𝜉̇ } + [𝐴]−𝑇 [𝐾] [𝐴]−1 {𝜉} = [𝐴]−𝑇 {𝑓} (4.36)

135
or
[𝐼] {𝜉̈ } + [𝑐 𝑜 ] {𝜉̇ } + [𝑘 𝑜 ] {𝜉} = {𝑓 𝑜 } , (4.37)
where
[𝑐 𝑜 ] = [𝐴]−𝑇 [𝐶] [𝐴]−1,
[𝑘 𝑜 ] = [𝐴]−𝑇 [𝐾] [𝐴]−1 , (4.38)
{𝑓 𝑜 } = [𝐴]−𝑇 {𝑓} .
The matrix [𝑐 𝑜 ] can be expressed in the matrix power series form
[𝑐 𝑜 ] = ∑𝑁−1 𝑜 𝑟
𝑟=𝑜 𝛾𝑟 [𝑘 ] . (4.39)
Equation (4.37) yields 𝑁 algebraic equations of the form
2(𝑁−1)
𝛾0 + 𝛾1 𝜔𝑟2 + 𝛾2 𝜔𝑟4 + . . . +𝛾𝑁−1 𝜔𝑟 = 2 𝜁𝑟 𝜔𝑟 , (𝑟 = 1, 2, . . , 𝑁)
which can be written in matrix form as

1 𝜔12 𝜔14 ⋯ 𝜔12(𝑁−1) 𝛾0 2𝜁1 𝜔1


1 𝜔22 𝜔24 ⋯ 𝜔22(𝑁−1) { 𝛾1 } = { 2𝜁2 𝜔2 } (4.40)
⋮ ⋮ ⋮ ⋱ ⋮ ⋮

1 𝜔𝑁2
𝜔𝑁4 ⋯ 2(𝑁−1) 𝛾𝑁−1 2𝜁𝑁 𝜔𝑁
[ 𝜔𝑁 ]
or
[𝜔𝑜 ] {𝛾} = {2𝜁𝑟 𝜔𝑟 }. (4.41)
From (4.41) we obtain the values 𝛾𝑟 which satisfy equation (4.39)
{𝛾} = [𝜔𝑜 ]−1 {2𝜁𝑟 𝜔𝑟 }. (4.42)
Using the first equation (4.38) and (4.39) we obtain
[𝐶] = [𝐴]𝑇 [𝑐 𝑜 ] [𝐴] = ∑𝑁−1 𝑇 −𝑇 −1 𝑟
𝑟=1 𝛾𝑟 [𝐴] ([𝐴] [𝐾] [𝐴] ) [𝐴] (4.43)
which is uniquely determined based on the values 𝛾𝑟 calculated from (4.42).
However, this is not the only proportional damping matrix that can be obtained
from the test data.
It must be added that, because for the modes with natural frequencies
higher than 𝜔𝑁 it was considered that 𝛾𝑠 = 0 (𝑠 = 𝑁, 𝑁 + 1, . . . , 𝑃 − 1), we have
forced the analytical model to have a damping value in all these modes, given by
1
𝜁𝑠 = 2𝜔 ∑𝑁−1 2 𝑟
𝑟=0 𝛾𝑟 (𝜔𝑠 ) . (4.44)
𝑠

Like in the first case, this can result in unrealistic damping coefficients for
the modes higher than those input.
The application of the method is difficult due to the inversion of the matrix
[𝜔𝑜 ] which is usually ill-conditioned, and to the presence of high powers of the
matrix [𝐾].

136
A method that eliminates the shortcomings of the previous methods has
been suggested by Hart and Collins [50]. In this method, the damping in all modes
above the 𝑟-th is assumed to be zero.
The orthogonality relations of the normalized modal vectors can be written
𝑇 ⌈ 1 , 𝑟=𝑠
{𝛹 (𝑠) } [𝑀] {𝛹 (𝑟) } = . (4.45)
⌊ 0 , 𝑟≠𝑠
Using the coordinate transformation
{𝑄} = [𝑀] {𝑞}, (4.46)
in the space of the {𝑄} coordinates, the eigenvectors are
{𝜌(𝑟) } = [𝑀] {𝛹 (𝑟) } . (4.47)
It is then possible to define a damping matrix which is diagonalized by the
𝑟-th eigenvector and represents the damping in the 𝑟-th mode. This matrix has the
form
[𝐶] = ∑𝑁𝑟=1[𝐶]𝑟 (4.48)
and is composed by matrices
𝑇
[𝐶]𝑟 = 2𝜁𝑟 𝜔𝑟 {𝜌(𝑟) } {𝜌(𝑟) } . (4.49)
It is important to note that, in constructing the damping matrix, the damping
is constrained to be zero in all modes not included in the summation (4.48). In this
way, the possibility to have negative damping coefficients in the higher
(unmeasured) modes of the system is avoided.

4.3.5. Complete versus Incomplete Models

Consider a spatial model consisting of 𝑁 lumped masses, interconnected


by elastic and dissipative elements, so that the movement of each mass represents
the movement of a significant point of the structure. The motion is thus defined by
𝑁 independent coordinates, the model has 𝑁 degrees of freedom, hence 𝑁 natural
frequencies.
Generally, the number of coordinates 𝑁 is different from the number of
resonances 𝑃 in the frequency range covered by measurements. Thus, the
identification methods in the frequency domain (Fig.4.2) can be classified into two
large categories:

4.3.5.1. Complete modal information

If the number of coordinates is small, it is possible to determine by test the


modal vectors and the modal parameters for a number 𝑃 of modes of vibration,
equal to the number of independent coordinates 𝑁 of the spatial model. Therefore,

137
if 𝑃 = 𝑁, the complete determination of the spatial model is theoretically possible,
based on the expressions defined for proportional damping
[𝑀] = [𝛹]−𝑇 [ˋ𝑚ˏ] [𝛹]−1, (4.50)
[𝐾] = [𝛹]−𝑇 [ˋ𝑘ˏ] [𝛹]−1 = [𝛹]−𝑇 [ˋ𝜔𝑟2 𝑚𝑟 ˏ] [𝛹]−1 , (4.51)
[𝐶] = [𝛹]−𝑇 [ˋ𝑐ˏ] [𝛹]−1 = [𝛹]−𝑇 [ˋ2𝜁𝑟 𝜔𝑟 𝑚𝑟 ˏ] [𝛹]−1 . (4.52)
Indeed, in this case the modal matrix [𝛹] is a squared 𝑁𝑥𝑁 matrix, with
elements measured in the resonance regions, so that the inversion can be performed
without problems.
For the identification of a unique model, it is necessary to take
measurements at the 𝑁 coordinates in the neighborhood of the 𝑁 resonance
frequencies, so that all natural modes be represented in the test data. Otherwise, as
already shown, we obtain a set of ill conditioned equations, whose solution is very
sensitive to small measurement errors.
Even in these conditions, the determination of the mass matrix [ˋ𝑚ˏ] is
generally done after pre-structuring the spatial model, hence knowing the mass
matrix [𝑀]. From expressions (4.51) and (4.52) we obtain the elements of the
stiffness and damping matrices.
If the elements of the matrices [𝑀] and [𝐾 ] can be computed directly, from
the drawings of the actual structure, then it is possible to measure the modes of
vibration, the modal masses and dampings, so that the matrix [𝐶] can be computed
based on (4.52).

4.3.5.2. Incomplete modal information

Usually, the number of degrees of freedom 𝑁 is large, and the frequency


range required to measure the response at all 𝑁 resonances is larger than the range
of frequencies of interest. When only 𝑃 < 𝑁 resonances are measured, there is an
infinite number of analytical models which describe the measured response, within
the usual limits of the experimental errors.
For this case, Berman and Flanelly [9] have developed the theory of
incomplete models of dynamic structures. Its detailed presentation is beyond the
aim of this book. In short, it is about truncated models, having 𝑁 independent
coordinates but only 𝑃 modes of vibration.
Expression (4.25) can be written under the form
𝜔𝑟2 𝑇
[𝐾] = ∑𝑁
𝑟=1 𝑚𝑟
[𝑀] {𝛹 (𝑟) } {𝛹 (𝑟) } [𝑀] .

If 𝑁 = 𝑃, the model is complete and the identification unique. If 𝑃 < 𝑁


𝜔2 𝑇
[𝐾]𝑖𝑛𝑐 = ∑𝑃𝑟=1 𝑟 [𝑀] {𝛹 (𝑟) } {𝛹 (𝑟) } [𝑀]
𝑚 𝑟

which can be computed if the matrix [𝑀] is known.

138
Because the terms containing high values of the frequency 𝜔𝑟 are not
included in the above summation, the dominant terms of the matrix [𝐾] are missing,
hence the matrix [𝐾]𝑖𝑛𝑐 will be different from [𝐾], its elements being devoid of
physical meaning.
The matrix [𝐾]𝑖𝑛𝑐 is singular (having order 𝑁 but rank 𝑃), hence cannot be
obtained by inversion and should be constructed separately, summing 𝑃 < 𝑁 terms.
Using the pseudoinverse matrices [36] we obtain
[𝐾]𝑖𝑛𝑐 = [𝛹]+𝑇 +
𝑖𝑛𝑐 [ˋ𝑘ˏ] [𝛹]𝑖𝑛𝑐 ,

[𝑀]𝑖𝑛𝑐 = [𝛹]+𝑇 +
𝑖𝑛𝑐 [ˋ𝑚ˏ] [𝛹]𝑖𝑛𝑐 , (4.53)
[𝐶]𝑖𝑛𝑐 = [𝛹]+𝑇 +
𝑖𝑛𝑐 [ˋ𝑐ˏ] [𝛹]𝑖𝑛𝑐 ,
where
−1
[𝛹]+ 𝑇
𝑖𝑛𝑐 = ([𝛹]𝑖𝑛𝑐 [𝛹]𝑖𝑛𝑐 ) [𝛹]𝑇𝑖𝑛𝑐
and [𝛹] is an 𝑁𝑥𝑃 rectangular modal matrix.
It is recommended to select 𝑃 equal to the number of effective degrees of
freedom of the structure. This is equal to the limited number of natural modes of
vibration, with frequencies within the range covered by tests and with dominant
contribution to the response of the structure.

4.3.6. Computation of Pseudoinverse Matrices

A problem giving rise to computation difficulties at the use of identification


methods is the inversion of rectangular matrices.
For example, if equation (3.78) is written for 𝑁𝑓 excitation frequencies
𝜔
̅1 , 𝜔
̅2 , . . . , 𝜔
̅𝑁𝑓 , we obtain the following set of linear equations

{𝛼𝑗ℓ𝐼 } = [ℬ] {𝜒𝑗ℓ } , (4.54)


where
𝑇
{𝛼𝑗ℓ𝐼 } = {𝛼𝑗ℓ𝐼 (𝜔
̅1 ) 𝛼𝑗ℓ𝐼 (𝜔
̅2 ) . . . 𝛼𝑗ℓ𝐼 (𝜔
̅ 𝑁𝑓 ) } , (4.55)

(1) (2) (𝑁) 𝑇


{𝜒𝑗ℓ } = { 𝜒𝑗ℓ 𝜒𝑗ℓ . . . 𝜒𝑗ℓ } , (4.56)
and
[ℬ] = [{𝐵1 (𝜔
̅𝑓 )} {𝐵2 (𝜔
̅𝑓 )} . . . {𝐵𝑁 (𝜔
̅𝑓 )} ] (4.57)
is a rectangular 𝑁𝑓 𝑥𝑁 matrix. Matrices {𝛼𝑗ℓ𝐼 } and [ℬ] are known (see § 5.3.4.4)
and we seek the vector {𝜒𝑗ℓ }.
If 𝑁𝑓 > 𝑁, equations (4.56) are incompatible. The approximate solution
which minimizes the Euclidean norm of the error vector {𝑒} = {𝛼𝑗ℓ𝐼 } − [ℬ] {𝜒𝑗ℓ }
is
{𝜒𝑗ℓ } = [ℬ]+ {𝛼𝑗ℓ𝐼 } (4.58)

139
where
−1
[ℬ]+ = [[ℬ]𝑇 [ℬ]] [ℬ]𝑇 (4.59)
is the Moore-Penrose pseudoinverse of the matrix [ℬ] (see § 5.3.4.5).
In many practical applications, the matrix [ℬ]𝑇 [ℬ] is numerically ill-
conditioned, hence [ℬ]+ cannot be computed from relation (4.59). In these cases
the method suggested by Golub and Reinsch [48] can be used. It is based on the
singular value decomposition of the matrix [ℬ]
[ℬ] = [𝑋] [ˋ𝑆ˏ] [𝑌]𝑇 , (4.60)
where [𝑋] is an 𝑁𝑓 𝑥𝑁 modal matrix having as columns 𝑁 orthonormalized natural
vectors associated to the highest 𝑁 eigenvalues of the matrix [ℬ] [ℬ]+ , and [𝑌] is
the 𝑁𝑥𝑁 modal matrix of the matrix [ℬ]𝑇 [ℬ].
Matrices [𝑋] and [𝑌] are orthonormalized
[𝑋]𝑇 [𝑋] = [𝑌]𝑇 [𝑌] = [𝑌] [𝑌]𝑇 = [𝐼], (4.61)
where [𝐼] is the identity matrix.
The elements of the diagonal matrix [ˋ𝑆ˏ] are the nonnegative square roots
of the eigenvalues of the matrix [ℬ]𝑇 [ℬ] and are referred to as singular values
[ˋ𝑆ˏ] = 𝑑𝑖𝑎𝑔(𝑠1 ; 𝑠2 ; . . . 𝑠𝑃 ; 𝑠𝑃+1; . . . 𝑠𝑁 ) (4.62)
where 𝑃 is the rank of [ℬ] and [ℬ]𝑇 [ℬ] .
If 𝑃 < 𝑁, then 𝑠𝑃+1 = 𝑠𝑃+2 = . . . = 𝑠𝑁 = 0.
Using transformations of the form
{𝜒𝑗ℓ } = [𝑌]{𝜒 𝑜 }, {𝛼𝑗ℓ𝐼 } = [𝑋] {𝛼 𝑜 } (4.63)
equations (4.58), (4.59) and (4.61) yield
[ℬ]+ = [𝑌] [ˋ𝑆ˏ]+ [𝑋]𝑇 (4.64)
where
[ˋ𝑆ˏ]+ = 𝑑𝑖𝑎𝑔 (𝑠𝑖+ ) (4.65)
and
⌈ 𝑠𝑖−1 𝑓𝑜𝑟 𝑠𝑖 > 0 ,
𝑠𝑖+ = (4.66)
⌊ 0 𝑓𝑜𝑟 𝑠𝑖 = 0 .
A detailed treatment can be found in the books by Boullion and Odell
[145], Rao and Mitra [146] and Albert [147]. A numerical method for the
computation of pseudoinverse matrices is described by Ionescu and Lupaş [148].

140
CHAPTER 5

Identification of Linear Systems Using Single Point


Excitation

5.1. PRINCIPLE OF IDENTIFICATION METHODS USING


SINGLE POINT EXCITATION

The complex receptance (3.56) of a hysteretically damped 𝑁-degree-of-


freedom system with proportional damping can be expressed as a sum of partial
fractions, each representing the response in one mode of vibration
(𝑟) (𝑟)
𝑞
̃ 𝛹ℓ 𝛹𝑗
𝛼̅𝑗ℓ =
𝑗
𝑓̂ ℓ
= ∑𝑁
𝑟=1 𝑘𝑟 −𝜔2 𝑚𝑟 +𝑖 ℎ𝑟
(5.1)

To obtain it experimentally, the system is excited by a force 𝑓ℓ = 𝑓̂ℓ 𝑒 𝑖𝜔𝑡


applied in ℓ (corresponding to the generalized coordinate 𝑞ℓ ) and the response
𝑞𝑗 = 𝑞̃𝑗 𝑒 𝑖𝜔𝑡 is measured in 𝑗 (corresponding to the generalized coordinate 𝑞𝑗 ), at
different excitation frequencies ω.
The receptance
𝑞̃𝑗
𝛼̅𝑗ℓ = = 𝛼𝑗ℓ 𝑒 𝑖𝜑 = 𝛼𝑗ℓ𝑅 + 𝑖 𝛼𝑗ℓ𝐼 (5.2)
𝑓̂ℓ

can be represented graphically in one of the following three formats:


a) magnitude-frequency (𝛼𝑗ℓ ­𝜔) and phase angle-frequency (𝜑­𝜔)
diagrams;
b) diagrams of the in-phase component (𝛼𝑗ℓ𝑅 ­𝜔) and out-of-phase
component (𝛼𝑗ℓ𝐼 ­𝜔) versus frequency;
c) polar plots, representing the locus of the tip of the vector 𝛼̅𝑗ℓ in the
complex plane, plotting the real and imaginary parts of the complex receptance
along the horizontal and vertical axes, and having ω as parameter.
During the frequency sweep, when the excitation frequency ω is close to
𝑘
𝜔𝑟 = √𝑚𝑟 , the contribution of the 𝑟-th term in the expression of 𝛼̅𝑗ℓ becomes
𝑟
dominant. The frequency variation of this term corresponds to the near resonance
behavior of an equivalent single-degree-of-freedom system, with parameters 𝑚𝑟 ,
𝑘𝑟 and ℎ𝑟 .
The response in the 𝑟-th mode is superimposed on the response described
by the other terms of the summation (5.1). The methods of frequency response

141
analysis differ by the assumptions made on the contribution of the “off-resonant”
modes of vibration to the total response near each resonance and by the separation
procedures of the modes of vibration with closely spaced natural frequencies.
Generally, the experimental modal analysis is used for the determination
of the following quantities: a) the resonance frequencies, considered equal to the
natural frequencies 𝜔𝑟 of the associate undamped system; b) the modal parameters
(𝑟)
𝜁𝑟 or 𝑔𝑟 , 𝑚𝑟 and 𝑘𝑟 ; c) the modal displacements 𝛹𝑗 , hence the modal matrix [𝛹].
Then, the modal model is used to determine the spatial model defined by the
matrices [𝑀], [𝐾], [𝐶] or [𝐻].
In all cases, it is necessary to repeat the measurements with the exciter
located at different points of the structure, to ensure the excitation of all important
modes of vibration, and to select the response measurements points with the largest
values of the signal/noise ratio.
Similar conclusions result from the analysis of systems with
nonproportional damping. In this case, the difficulties encountered in the
measurement of the monophase forced modes or the complex modes of vibration
imposes their analytical determination, hence the use of combined
analytical/experimental techniques and computers.

5.2. IDENTIFICATION OF UNDAMPED SYSTEMS

This section is devoted to the presentation of the identification methods


used for systems with negligible damping (chassis, girders, lathe beds) or for
damped systems, based on measurements away of resonance.

5.2.1. The Shape of Frequency Response Curves

The steady-state solution to the equations of motion of a undamped 𝑁-


degree of freedom system, excited by harmonic in-phase forces, has the form (3.24)
𝑟 𝑇
{𝛹 } {𝑓̂ } {𝛹 𝑟 }
( ) ( )

{𝑞̂} = ∑𝑁
𝑟=1 2 . (5.3)
𝑘𝑟 (1−𝜔2)
𝜔𝑟

In the case of single point excitation, the drive point receptance is


(𝑟) (𝑟) 𝑟)
𝑞
̂ 𝛹ℓ 𝛹ℓ 𝜘(ℓℓ
𝛼ℓℓ = 𝑓̂ ℓ = ∑𝑁
𝑟=1 2 = ∑𝑁
𝑟=1 2 (5.4)
ℓ 𝑘𝑟 (1−𝜔2) 1−𝜔2
𝜔𝑟 𝜔𝑟

and the transfer (or cross) receptance is


𝑞
̂
𝑗 𝑁
(𝑟) (𝑟)
𝛹ℓ 𝛹𝑗 𝑁 𝜘(𝑗ℓ𝑟)
𝛼𝑗ℓ = 𝑓̂ = ∑𝑟=1 2 = ∑𝑟=1 2 . (5.5)
ℓ 𝑘𝑟 (1−𝜔2) 1−𝜔2
𝜔𝑟 𝜔𝑟

142
(𝑟) (𝑟)
(𝑟) 𝛹ℓ 𝛹ℓ (𝑟)
The quantity 𝜘ℓℓ = is always positive. The quantity 𝜘𝑗ℓ =
𝑘𝑟
(𝑟) (𝑟)
𝛹ℓ 𝛹𝑗
is positive when the points 𝑗 and ℓ move in phase, and negative – when
𝑘𝑟
these points move out-of-phase.
For reasons explained later in the text, it is convenient to use the graphs of
mobility and mechanical impedance.
The expression of the drive point mobility is
𝑞̇̂ (𝑟) 𝜔
ℳℓℓ = 𝑓̂ℓ = 𝜔 𝛼ℓℓ = ∑𝑁
𝑟=1 𝜘ℓℓ 𝜔2
(5.6)
ℓ 1− 2
𝜔 𝑟

and the transfer mobility is


𝑞̇̂ (𝑟) 𝜔
ℳ𝑗ℓ = 𝑓̂𝑗 = 𝜔 𝛼𝑗ℓ = ∑𝑁
𝑟=1 𝜘𝑗ℓ 𝜔2
. (5.7)
ℓ 1− 2
𝜔 𝑟

The derivative with respect to 𝜔 of the general term of expression (5.7)


is
𝜔2 2𝜔 𝜔2
𝜔 𝜘𝑗ℓ (𝑟) (1− 2 )−𝜔 (− 2) 1+ 2
𝑑 (𝑟) 𝜔𝑟 𝜔𝑟 (𝑟) 𝜔 𝑟
( 2) =𝜘𝑗ℓ 2 = 𝜘𝑗ℓ 2 .
𝑑𝜔
1−𝜔2 𝜔2 𝜔2
𝜔𝑟 (1− 2) (1− 2)
𝜔𝑟 𝜔𝑟

(𝑟)
The sign of the derivative is given by the sign of 𝜘𝑗ℓ .

The slope of the diagram of drive point mobility (Fig.5.1,a) is always


positive. The curve has vertical asymptotes at the frequencies of resonance
(Foster’s theorem) [38].
Between two frequencies of resonance there is always a frequency of
antiresonance and vice versa.
The curve of the transfer mobility (Fig.5.1,b) has a different shape due to
(𝑟)
the negative terms 𝜘𝑗ℓ in expression (5.7). Because the slope of the diagram can
be negative, a minimum different from zero can occur between two frequencies of
resonance.

143
In order to explain these features of the mobility diagrams, let consider only
two terms in the summations (5.6) and (5.7), denoted
𝜘1 𝜔 𝜘2 𝜔
ℳ1 = 2 , ℳ2 = 𝜔2
(5.8)
1−𝜔2 1− 2
𝜔1 𝜔 2
where 𝜘1 > 0 and 𝜘2 > 0.
Figure 5.2,a shows the variation with frequency of ℳ1 (solid line) and ℳ2
(dashed line). Figure 5.2,b shows the diagrams of the functions ℳ1 (solid line) and
(−ℳ2 ) (dashed line). Figure 5.2,c shows the graph of (ℳ1 + ℳ2 ) and Fig.5.2,d -
the graph of (ℳ1 − ℳ2 ). Figures 5.2,e and f depict the graphs of the magnitudes
of these functions, and Figs.5.2,g and h – the same diagrams in log-log scaling.

This shows the advantage of representing magnitude-frequency diagrams


in logarithmic (at least semi-logarithmic) scales, which makes the anti-resonances
more evident.

Outside the resonance and anti-resonance regions, the response curve can
be approximated with a “skeleton diagram” which consists of straight lines defining

144
the response of a pure mass or a pure spring (like the asymptotes of Bode’s
diagrams).
Of the three representations of the drive-point response, as diagrams of the
drive-point receptance (Fig.5.3,a), drive-point mobility (Fig.5.3,b) and drive-point
inertance (Fig.5.3,c), the diagram of the mobility is preferred due to its “symmetry”
(about a vertical line at the frequency of resonance). It consists of lines with slopes
of ± 20 𝑑𝐵/𝑜𝑐𝑡𝑎𝑣𝑒 (slopes of ±1 in the case of equal scales) so that the “skeleton”
is constructed easier.
The same applies to the drive-point impedance curve (Fig.5.3,d) where
anti-resonances appear as maxima, resonances as minima, and inevitably alternate.

5.2.2. Method of “Skeleton” Diagrams

Plotting the curve of mobility or mechanical impedance on a log-log graph


with annotated grids including constant mass and constant stiffness lines, a skeleton
diagram is superimposed which has arms to which appropriate values of 𝑚′ and 𝑘′
can be attached. At the intersection of the 𝑘′ and 𝑚′ lines, the frequency of
𝑘’
resonance 𝜔𝑅 or the frequency of anti-resonance 𝜔𝐴 is equal to √ .
𝑚’

For the skeleton diagram from Fig.5.4, the following relations can be
established
2
𝜔𝐴3 𝜔𝐴3 𝜔𝐴2 2 ′ 𝜔𝐴3 𝜔𝐴2 2 ′
𝑘3′ = 𝜔𝐴3
2
𝑚2′ = 2 𝑘 ′
2 = ( ) 𝑚1 = ( ) 𝑘1 =
𝜔𝑅2 𝜔𝑅2 𝜔𝑅2 𝜔𝑅1
𝜔𝐴3 𝜔𝐴2 𝜔𝐴1 2
=( ) 𝑚0′ . (5.9)
𝜔𝑅2 𝜔𝑅1

For any two successive resonances or anti-resonances (Fig.5.5)


𝜔𝑅2 ℳ𝐼 𝜔𝐴2 ℳ𝑅
= , = (5.10)
𝜔𝑅1 ℳ𝐴 𝜔𝐴1 ℳ𝐼
where ℳ𝐼 is the ordinate of the intersection point 𝐼 obtained extending the opposite
arms of two adjacent links.

145
146
For several skeleton diagrams with simple configuration, Table 5.1
presents the corresponding lumped mass models and the expressions of their mass
and stiffness parameters. As shown in the table, the identification is not unique,
each diagram corresponds to two different models.
The following observations are useful at the model construction and the
derivation of the expressions of their parameters:
a – The frequencies of resonance are independent of the stiffness of a spring
attached between the driving point and the remainder of the system.
𝑘
Thus, for system 3, the frequency 𝜔𝑅1 = √𝑚1 is independent of 𝑘2 , and
1

1 1
for system 7, the frequency 𝜔𝑅1 = √𝑘1 (𝑚 + ) is independent of 𝑘2 .
1 𝑚2
b – The frequencies of anti-resonance are independent of the elements
connected in parallel to sub-systems whose drive-point impedance becomes infinite
at these frequencies.
𝑘1
For system 4, 𝜔𝐴1 = √ and the stiffness 𝑘2 is inactive. For system 5,
𝑚1
𝑘 𝑘
𝜔𝐴1 = √𝑚1 because the mass 𝑚2 is inactive. At system 8, 𝜔𝐴1 = √𝑚1 , the
1 1
1
elements 𝑚2 and 𝑘2 being inactive. At system 9, 𝜔𝐴1 = √𝑚 (𝑘1 + 𝑘2 ) because
1
𝑚2 is inactive.
c – The very low frequency response of a grounded system is springlike,
and of a ungrounded system – masslike.
𝑘 𝑘
Thus, for system 9, 𝑘1′ = 1 2 and for system 11, 𝑚0′ = 𝑚1 + 𝑚2 + 𝑚3
𝑘1 +𝑘2
is the total mass of the system.
d – If there is any mass element directly attached to the drive point, the
very high frequency response is mass-like, and spring-like if there is no mass
element directly attached to the drive point.
Thus 𝑚2′ = 𝑚2 for systems 8, 9 and 11, 𝑚2′ = 𝑚3 for system 10, and 𝑘2′ =
𝑘2 for systems 3 and 7.
In the book by Salter [106] a trial and error method is presented for
structuring lumped-mass models which describe the response of continuous
structures over limited frequency ranges.

5.2.3. Modal Model Method

If, during testing, the structure is supported freely on a very soft suspension
with negligible damping, equation (5.4) can be written
(𝑟) (𝑟)
𝑞
̂ 𝛹ℓ 𝛹ℓ 1
𝛼ℓℓ = 𝑓̂ ℓ = ∑𝑁
𝑟=1 𝑚 2 2 − 2
𝜔 𝑀ℓℓ
. (5.11)
ℓ 𝑟 (𝜔𝑟 −𝜔 )

147
The additional term in the right hand side is due to the rigid body motion
of the structure on the elastic suspension. The constant parameter 𝑀ℓℓ depends on
the total mass of the structure, its mass moment of inertia and the location of point
ℓ with respect to the center of mass.
The frequencies of resonance 𝜔𝑟 and anti-resonance 𝜔𝐴𝑟 are obtained from
the frequency response diagram 𝛼ℓℓ ­𝜔 determined experimentally.
Davis [28] recommends the normalization of the modes of vibration by
setting all modal masses equal to the total mass of the structure. In this case,
evaluating expression (5.11) at each frequency of anti-resonance, where 𝛼ℓℓ = 0 ,
the following set of equations is obtained
(𝑟) (𝑟)
𝛹ℓ 𝛹ℓ 𝑚𝑟
∑𝑁
𝑟=1 2 = 2 (5.12)
𝜔𝑟2 −𝜔𝐴𝑟 𝜔𝐴𝑟 𝑀ℓℓ
(in which 𝜔𝐴𝑟 is the frequency of anti-resonance just under the frequency of
(𝑟)
resonance 𝜔𝑟 ) wherefrom 𝛹ℓ is determined.
𝑚
The quantity 𝑀 𝑟 in the right hand side is calculated weighing the structure
ℓℓ
(if possible) and measuring its free response, when supported on calibrated springs.
With the modal coefficients measured at the driving point, the remaining
coefficients are determined vibrating the structure at each resonance, so that the
response be dominated by the associated mode of vibration. When testing at
resonance is considered dangerous for the structure integrity, methods based on
near-resonance measurements can be used, like those described in § 5.3.

5.3. IDENTIFICATION OF SYSTEMS WITH PROPORTIONAL


DAMPING

5.3.1. Closely Spaced Modes

In order to better understand the difficulties occurring at the graphical


analysis of the frequency response curves of systems with several degrees of
freedom, let consider the summation of two modes of vibration with relatively close
natural frequencies.
Denoting 𝛼̅𝑗ℓ = 𝛼̅, the complex transfer receptance (5.1) can be written
𝜘1 𝜘2
𝛼̅ = 𝛼̅1 + 𝛼̅2 = 𝜔2
+ 𝜔2
(5.13)
1− 2 +𝑖𝑔1 1− 2 +𝑖𝑔2
𝜔1 𝜔2
where
(1) (1) (2) (2)
𝛹ℓ 𝛹𝑗 𝜕𝑞 𝜕𝑞 𝛹ℓ 𝛹𝑗 𝜕𝑞 𝜕𝑞
𝜘1 = 𝑘1
= 𝑘1 𝜕𝑝ℓ 𝜕𝑝 𝑗 , 𝜘2 = 𝑘2
= 𝑘1 𝜕𝑝ℓ 𝜕𝑝 𝑗 (5.14)
1 1 1 2 2 2

are real quantities.


The complex receptance (5.13) can be expressed in one of the following
two forms

148
𝛼̅ = 𝛼𝑒 𝑖𝜑 = 𝛼𝑅 + 𝑖 𝛼𝐼 , (5.15)
where
𝜔2 𝜔2
1− 2 1− 2
𝜔 1 𝜔2
𝛼𝑅 = 𝜘1 2 + 𝜘2 2 = 𝛼𝑅1 + 𝛼𝑅2 ,
𝜔2 𝜔2
(1− 2 ) +𝑔12 (1− 2 ) +𝑔22
𝜔1 𝜔2
(5.16)
−𝑔1 −𝑔2
𝛼𝐼 = 𝜘1 2 + 𝜘2 2 = 𝛼𝐼1 + 𝛼𝐼2 ,
𝜔2 𝜔2
(1− 2 ) +𝑔12 (1− 2 ) +𝑔22
𝜔 1 𝜔 2
𝛼𝐼
𝛼 = √𝛼𝑅2 + 𝛼𝐼2 , 𝜑 = 𝑡𝑎𝑛−1 , (5.17)
𝛼𝑅

and also
2 2 2 2
𝛼1 = √𝛼𝑅1 + 𝛼𝐼1 , 𝛼2 = √𝛼𝑅2 + 𝛼𝐼2 ,
(5.18)
𝛼 𝛼
𝜑1 = 𝑡𝑎𝑛−1 𝛼 𝐼1 , 𝜑2 = 𝑡𝑎𝑛−1 𝛼 𝐼2 .
𝑅1 𝑅2

Figures 5.6 to 5.10 display the corresponding frequency response curves,


𝑚 𝑚
calculated for 𝜘1 = 0.2 , 𝜘2 = 0.3 , 𝑔1 = 0.03, 𝑔2 = 0.05, 𝜔1 = 23 𝑟𝑎𝑑⁄𝑠,
𝑁 𝑁
𝜔2 = 25 𝑟𝑎𝑑⁄𝑠. The individual diagrams for each mode are drawn with dashed
lines, and the resultant diagram – with solid line [39].

It is seen that the peaks of the magnitude plot (Fig.5.6) as well as the points
of zero real component (Fig.5.8) and the points of ­90° phase angle (Fig.5.7) have
abscissae which do not correspond to the natural frequencies of the two modes. On

149
the contrary, the peaks of the plot of the imaginary component of receptance
(Fig.5.9) and to a less extent the inflection points of the diagrams of the real
component (Fig.5.8) and the phase angle (Fig.5.7) have abscissae very close to the
frequencies 𝜔1 and 𝜔2 . Their location in figure 5.9 is facilitated by the sharp peak
of the curve 𝛼𝐼 ­𝜔 at resonance, where the bandwidth is smaller than in the diagram
𝛼­𝜔. It can also be seen that at resonance, the peak value of the total imaginary
component is very close to the peak value of the imaginary component of the
resonant mode.

The polar plot (Fig.5.10) has two almost circular bulges in the
neighborhood of the frequencies 𝜔1 and 𝜔2 , connected by a loop corresponding to
the anti-resonance. The frequencies of resonance are no more at the crossing with
the negative imaginary axis. They correspond to the points where the imaginary
component 𝛼𝐼 has maximum values and where, for regular increments of frequency
𝛥𝜔 , the lengths of the arc of curve between two successive points is a maximum.

150
Figure 5.11 shows the polar plot of the complex receptance given by
equation (5.13), calculated for two close natural frequencies, 𝜔1 = 23 𝑟𝑎𝑑⁄𝑠 and
𝜔2 = 23.6 𝑟𝑎𝑑⁄𝑠 , the other parameters being the same as for the diagrams from
Figs. 5.6-5.10 [39]. The diagram has two almost circular parts but the connecting
loop has become merely a bulge with two points of inflection on the curve. In this
case too, the frequencies of resonance can be located where for equal increments
∆𝜔 = 0.1 𝑟𝑎𝑑⁄𝑠 the distance between successive points is a maximum.
For comparison, Fig.5.12 shows the polar diagram of the complex
receptance (5.13) plotted for two values of the hysteretic damping factor 𝑔1 = 0.01
and 𝑔2 = 0.05 , the other parameters being the same as for Fig.5.10. The difference
in the level of damping of the two modes generates two loops which approximate
circles with different diameters.
For an elastically supported rigid body, vibrating in a single plane, Fig.5.13
shows the shape of the two modes of vibration. Figure 5.14 depicts the polar plots
measured in four points of the body.
Figure 5.14,a shows that at point ① the movements in the two modes are
out-of-phase so that the polar plot exhibits a loop in the positive imaginary region.
Because point ② is selected as a node of the mode Ⅱ, the vibration in this point
takes place only in mode Ⅰ, and the polar plot is a circle (Fig.5.14,b). At points ③
and ④ the displacements in the two modes are in-phase, the contribution of each
mode to the overall response is different, so that the diagrams from Figs. 5.14,c and
d have a loop for each mode of vibration, but different shape.

A special case occurs when three modes of vibration have resonant


frequencies close to one another but one is “hidden”. Kennedy and Pancu [61]
considered the function
𝜘𝑗
𝛼̅ = ∑3𝑗=1 (5.19)
𝜔2
1− 2+𝑖𝑔𝑗
𝜔𝑗
𝑚 𝑚 𝑚
where 𝜘1 = 0.2 𝑁 , 𝜘2 = −0.075 𝑁
, 𝜘3 = 0.3 𝑁
, 𝑔1 = 𝑔2 = 0.03, 𝑔3 = 0.05,
𝜔1 = 23.33 𝑟𝑎𝑑⁄𝑠, 𝜔2 = 24.16 𝑟𝑎𝑑⁄𝑠, 𝜔3 = 25 𝑟𝑎𝑑⁄𝑠.

151
The polar plot of the complex receptance (5.19) is shown in Fig.5.15. The
diagram resembles the plot from Fig.5.10 (the difference is the increase in the size
of the inward loop) hence it might be considered as resulting from the superposition
of only two modes with natural frequencies close together.

The magnitude versus frequency diagram (Fig.5.16) has only two maxima
like the diagram of the imaginary component (Fig.5.17). On the other hand, the plot
of the amplitude of the derivative 𝑑𝑠⁄𝑑𝜔 , where 𝑑𝑠 is the arc length corresponding
to a frequency increment 𝑑𝜔, has the shape depicted in Fig.5.18 with three maxima
which allows the location of three frequencies of resonance, hence the identification
of three close modes of vibration.

5.3.2. Identification Methods Based on the Diagram of Response


Amplitude

For most practical structures, the magnitude-frequency diagram exhibits a


peak value at each natural frequency.
Resonance location is done using the criterion of maximum amplitude. It
is considered that the abscissae of the points of maximum amplitude of the total
response are equal to the undamped natural frequencies 𝜔𝑟 (Fig.5.19).
The calculation of modal parameters and mode-shape vectors is performed
differently depending on the relative spacing of natural frequencies.
a) For systems with well-spaced natural frequencies it is considered that in
the neighborhood of resonance the vibration takes place exclusively in the resonant
mode shape, so that the contribution from the off-resonant modes is neglected. The
near resonance portions of the response curve are assimilated to the response curve
of a single-degree-of-freedom system (see Ch.3). The damping is given by the
Broadbent and Hartley’s formula (2.30)
𝜔𝑟′′2 −𝜔𝑟′2 𝜔 ′′2 −𝜔′2
𝑔𝑟 = = 𝜔𝑟′′2 +𝜔𝑟′2 (5.20)
2𝜔𝑟2 𝑟 𝑟

where 𝜔𝑟′ and 𝜔𝑟′′ are the frequencies of the half-power points, where the amplitude
is 1⁄√2 from the peak value at resonance (Fig.5.19).
The modal masses are

152
1
𝑚𝑟 = . (5.21)
𝜔2𝑟 𝑔𝑟 𝛼𝑗ℓ (𝜔𝑟 )

The shapes of the modal vectors are calculated from the ratios of the
amplitudes at various points when the structure is being driven at a natural
frequency
(𝑟)
𝛹𝑗 𝛼𝑗ℓ (𝜔𝑟 )
(𝑟) = . (5.22)
𝛹𝑗,𝑚𝑎𝑥 𝛼𝑗ℓ (𝜔𝑟 )𝑚𝑎𝑥

Rusu and Rozsa [149] have successfully applied the principle of this
method to the structural identification of a turning and boring lathe.
b) For lightly damped systems with close natural frequencies better
accuracy is obtained assuming that in some frequency range around the resonance
the contribution to the amplitude of the total response from the off-resonant modes
is constant.
Application of Gladwell’s method [47] requires the coordinates of three
points on each resonance peak (Fig.5.20).
In the vicinity of the natural frequency 𝜔𝑟 , the equation of the response
curve can be written under the form
1
𝛼𝑗ℓ = + 𝜉𝑟 , (5.23)
2 2
𝑘𝑟 √(1−𝜔2) +𝑔2𝑟
𝜔𝑟

where 𝜉𝑟 is the constant contribution from the off-resonant modes.

Denoting
𝛼𝑗ℓ (𝜔𝐴 ) = 𝛼𝐴 , 𝛼𝑗ℓ (𝜔𝐵 ) = 𝛼𝐵 , 𝛼𝑗ℓ (𝜔𝑟 ) = 𝛼𝑅 ,
2
𝛼𝑅 −𝛼𝐴 𝜔𝐵 −𝜔𝑟2 𝛼𝐵 −𝜉𝑟
= 𝜎, 2 −𝜔2 = 𝜏, = 𝛿, (5.24)
𝛼𝐴 −𝛼𝐵 𝜔𝐴 𝑟 𝛼𝐴 −𝛼𝐵
we obtain
[ 2𝜎 (𝜏 2 − 1) − 2] 𝛿 3 + [ (𝜎 2 + 2𝜎) (𝜏 2 − 1) − 4𝜎 − 5 ] 𝛿 2 −
−2 (𝜎 + 1) (𝜎 + 2) 𝛿 − (𝜎 + 1)2 = 0 . (5.25)

153
𝛼𝑅 +𝛼𝐵
Choosing 𝜎 = 1 , i.e. 𝛼𝐴 = , equation (5.25) simplifies to
2
(2𝜏 2 − 4) 𝛿 3 + 3 (𝜏 2 − 4) 𝛿 2 − 12𝛿 − 4 = 0 . (5.26)
Equation (5.25) has just one positive root 𝛿 provided that
𝜎+1
𝜎 >0, 𝜏2 > 𝜎
.
When the root of equation (5.25) is known, the damping factor 𝑔𝑟 may be
calculated as
2 2 2 2
𝜔2𝐴 𝜏 𝛿 −(𝛿+1)
𝑔𝑟2 = (1 − ) . (5.27)
𝜔2𝑟 2𝛿+1
The “true” resonant peak amplitude is given by
𝛼𝑗ℓ (𝜔𝑟 ) − 𝜉𝑟 = (1 + 𝜎 + 𝛿) (𝛼𝐴 − 𝛼𝐵 ),
(𝑟)
which is proportional to the element 𝛹𝑗 of the modal vector.
It is customary to calculate modal vectors by a direct separation method
from the experimentally measured overall response.
Using equations (3.65) and (3.11), the vector of complex receptances (3.70)
becomes
(𝑟)
𝛹ℓ {𝛹 𝑟 } 𝑒−𝑖 𝜓𝑟
( )

{𝛼̅}ℓ = ∑𝑁
𝑟=1 . (5.28)
2 2
𝑚𝑟 𝜔2𝑟 √(1−𝜔2) +𝑔2𝑟
𝜔𝑟

Because in the neighborhood of resonances cos 𝜓𝑟 → 0 in the dominant


terms of the summation (5.28), which is equivalent to approximate the total
response by the sum of out-of-phase responses in each mode, the vector of real
receptances can be written under the form
(𝑟)
𝛹ℓ {𝛹 𝑟 }
( )
−𝑠𝑖𝑛 𝜓𝑟
{𝛼ℓ (𝜔)} = ∑𝑁
𝑟=1 (5.29)
𝑚𝑟 2 2
𝜔2𝑟 √(1−𝜔2) +𝑔2𝑟
𝜔𝑟
or
(𝑟)
𝛹ℓ {𝛹 𝑟 }
( )

{𝛼ℓ (𝜔)} = ∑𝑁
𝑟=1 𝑄𝑟 (𝜔), (5.30)
𝑚𝑟
where
−𝑠𝑖𝑛 𝜓𝑟 −𝑔𝑟
𝑄𝑟 (𝜔) = = = 𝑏𝑟𝜔(𝜔)
2 . (5.31)
2 2 2 2 𝑟
𝜔2𝑟 √(1−𝜔2) +𝑔2𝑟 𝜔2𝑟 [(1−𝜔2) +𝑔2𝑟 ]
𝜔𝑟 𝜔𝑟
Defining a square matrix
(1) (2) (𝑁)
𝛹ℓ 𝛹ℓ 𝛹ℓ
[𝛹ℓ 𝛹] = [ {𝛹 (1) } {𝛹 (2) } . . . {𝛹 (𝑁) } ] , (5.32)
𝑚1 𝑚2 𝑚𝑁

154
in which the modal vectors can be normalized such as [`𝑚ˏ] = [𝐼] , expression
(5.30) becomes
{𝛼ℓ (𝜔)} = [𝛹ℓ 𝛹] {𝑄(𝜔)} . (5.33)
Equations (5.33) and (5.31) are evaluated at the 𝑁 frequencies of resonance
𝜔1 , 𝜔2 , . . . . , 𝜔𝑁 determined as abscissae of the points of maximum total response.
The results are used to construct two square matrices
[𝛼] = [ {𝛼ℓ (𝜔1 )} {𝛼ℓ (𝜔2 )} . . . . . {𝛼ℓ (𝜔𝑁 )} ] , (5.34)
where 𝛼𝑗ℓ (𝜔𝑟 ) are the peak receptances determined from the magnitude-frequency
diagram, and
[𝑄] = [ {𝑄(𝜔1 )} {𝑄(𝜔2 )} . . . . . {𝑄(𝜔𝑁 )} ] (5.35)
so that equations (5.33) yield
[𝛼] = [𝛹ℓ 𝛹] [𝑄] (5.36)
which can be written
(1) (1) (2) (2) (𝑁) (𝑁)
𝛼1 (𝜔1 ) 𝛼1 (𝜔2 ) ⋯ 𝛼1 (𝜔𝑁 ) 𝛹ℓ 𝛹1 𝛹ℓ 𝛹1 ⋯ 𝛹ℓ 𝛹1
(1) (1) (2) (2) (𝑁) (𝑁)
𝛼 (𝜔 ) 𝛼2 (𝜔2 ) ⋯ 𝛼2 (𝜔𝑁 )
[ 2 1 ] = 𝛹ℓ 𝛹2 𝛹ℓ 𝛹2 ⋯ 𝛹ℓ 𝛹2 ·
⋮ ⋮ ⋮ ⋮ ⋮ ⋮
𝛼𝑁 (𝜔1 ) 𝛼𝑁 (𝜔2 ) ⋯ 𝛼𝑁 (𝜔𝑁 ) (1) (1) (2) (2) (𝑁) (𝑁)
[𝛹ℓ 𝛹𝑁 𝛹ℓ 𝛹𝑁 ⋯ 𝛹ℓ 𝛹𝑁 ]
(𝜔
𝑄1 1 ) (𝜔
𝑄1 2 ) ⋯ 𝑄1 𝑁 (𝜔 )
𝑄2 (𝜔1 ) 𝑄2 (𝜔2 ) ⋯ 𝑄2 (𝜔𝑁 )
·[ ].
⋮ ⋮ ⋮
𝑄𝑁 (𝜔1 ) 𝑄𝑁 (𝜔2 ) ⋯ 𝑄𝑁 (𝜔𝑁 )
From equation (5.36) we obtain
[𝛹ℓ 𝛹] = [𝛼] [𝑄]−1 . (5.37)
The modal displacements are given by
(𝑟) (𝑟) (𝑟)
𝛹ℓ = √𝛹ℓ 𝛹ℓ , (𝑟 = 1, 2, . . , 𝑁), (5.38)
(𝑟) (𝑟)
(𝑟) 𝛹ℓ 𝛹𝑗 (𝑟 = 1, 2, … , 𝑁),
𝛹𝑗 = (𝑟)
𝛹ℓ (𝑗 = 1, 2, . . , ℓ − 1, ℓ + 1, . . . , 𝑁).

The method has been used by Thoren [116] in a test of the Saturn V vehicle,
S-Ⅱ stage, LOX tank-engine support structure.
c) For systems with close natural frequencies and heavy damping, the
maximum amplitude methods yield erroneous results. The peak amplitude does not
occur at the natural frequency of the resonant mode. The response curve is

155
asymmetric and modified with respect to the curve of a single-degree-of-freedom
system, so that the damping estimation is not accurate.
When a heavily damped mode has the natural frequency close to that of a
lightly damped mode (Fig.5.21), the former can be completely obscured. It is
recommended to use other identification methods, presented in the following.

5.3.3. Calculation of Damping from the Phase Angle Diagram

For the modes of vibration with relatively spaced natural frequencies, the
phase resonance corresponds to phase lags of 90° or 270° (Fig.5.22) between force
and displacement.
For modes of vibration with relatively close natural frequencies, this
criterion cannot be used. Good results are obtained if the natural frequencies are
determined as abscissae of the
inflection points of the phase angle
diagram close to 90° and 270°,
respectively, on the portions with
positive slope.
The damping factor is
calculated based on the value of the
slope of tangent to the curve at the
inflection points

2
𝑔𝑟 = 𝑑𝜑 . (5.40)
𝜔𝑟 |𝑑𝜔 |
𝜔=𝜔𝑟

Obviously, the mode shapes, the modal stiffnesses 𝑘𝑟 and the modal
masses 𝑚𝑟 cannot be determined from the phase angle diagram alone.
It is considered [85] that this method yields more accurate damping values
than the peak amplitude method.

156
5.3.4. Identification Methods Based on the Diagrams of Response
Vector Components

The representation of the vibration response by the in-phase and out-of-


phase components for the parameter identification of an elastic system was first
suggested by G. de Vries [124]. The analysis techniques have been developed by
Kennedy and Pancu [61], Stahle and Forlifer [113], Klosterman [63], Berman [10]
and others. In the following, the methods are exemplified using receptance
diagrams.

5.3.4.1. The quadrature response method

Under this name [86] the simplest identification method has been applied
only to systems with well-spaced natural frequencies. Resonances are located at the
frequencies at which the in-phase component rather than the absolute amplitude is
zero. The damping factor is calculated using the formula of half-power points
(5.20), which correspond to the peak in-phase response near a resonance
(Fig.5.23,a).
The modal stiffness 𝑘𝑟 is given by
1
𝑘𝑟 = (5.41)
𝑔𝑟 |𝛼𝑗ℓ (𝜔𝑟 )|
𝐼
The mode shapes are calculated from the amplitudes of the quadrature
component at various points when the structure is being driven at a natural
(𝑟)
frequency (Fig.5.23,b). Therefore 𝛹𝑗 = 𝛼𝑗ℓ𝐼 (𝜔𝑟 ).

5.3.4.2 The maximum quadrature component method

As shown in Fig.5.5, for systems with relatively close natural frequencies,


the in-phase response is not zero at the natural frequencies of the individual modes
of vibration.
In the “maximum quadrature component method” [86] resonances are
defined by the frequencies at which the quadrature component of the response is a
maximum. The advantage is that the quadrature response peaks more sharply than
the total response.
It is assumed that the 𝑟-th mode is in quadrature with the force at 𝜔𝑟 and
that the quadrature components of the response in all other modes are negligible.
This implies that the difference between the total response at 𝜔𝑟 and the quadrature
component of the 𝑟-th mode at this frequency is due exclusively to the in-phase
components of the non-resonant modes (Fig.5.24).
The damping factor is calculated using the formula of half-power points
(5.20) where the amplitude of the quadrature component is half the peak amplitude
at resonance (Fig.2.13).

157
The shapes of the principal modes are calculated based on the maximum
values of the quadrature component measured at various significant points, at each
frequency 𝜔𝑟 .

The present author has used a variant of this method for the structural
identification of a milling machine [150].

5.3.4.3. The in-phase response method

If only the diagram of the in-phase response is available, the resonance is


more accurately located by the abscissa of an inflection point located between two
points of extrema (Fig.5.25).
The damping factor is given by

2 𝜎𝑟
𝑔𝑟 = 𝑑𝛼𝑗ℓ (5.42)
𝜔𝑟 | 𝑑𝜔𝑅 |
𝜔=𝜔𝑟

where 𝜎𝑟 is the distance between the


tangents at the points of extrema and
𝑑𝛼𝑗ℓ𝑅
| | = 𝑡𝑎𝑛 𝜃
𝑑𝜔 𝜔=𝜔𝑟
is the absolute value of the slope of the line
tangent to the curve at the inflection point.

158
The modal stiffnesses are
1
𝑘𝑟 = 𝑔 𝜎 (5.43)
𝑟 𝑟

and the modal masses


𝑘𝑟
𝑚𝑟 = . (5.44)
𝜔2𝑟
The shapes of the principal modes are calculated based on the values 𝜎𝑟
(measured at various points of the structure) proportional to the peak response in
the resonant mode, hence to the elements of the modal vector {𝛹 (𝑟) }.
For systems with close natural frequencies, the methods presented above
produce distorted mode shapes, dependent on the particular excitation points
selected. Several identification methods, used for the accurate experimental
determination of normal vibration modes from the total response, are presented in
the following.

5.3.4.4. Phase separation method

Under this name, Stahle and Forlifer [113] suggested the first technique for
the separation of modes from the measured total quadrature response.
Starting from the observation that the diagram of the quadrature response
is the least influenced by the effects of interaction between modes, they suggest the
calculation of 𝜔𝑟 and 𝑔𝑟 as in the maximum quadrature component method
(§5.3.4.2).
For the calculation of the mode shapes, they consider the equation (3.80)

{𝛼ℓ𝐼 (𝜔)} = [𝜒] {𝐵(𝜔)} (5.45)

which shows the frequency dependence of the quadrature response at coordinates


𝑞𝑗 (𝑗 = 1, 2, . . . , 𝑁) due to a harmonic force applied at coordinate 𝑞ℓ .
Equation (5.45) is evaluated at 𝑁𝑓 excitation frequencies 𝜔
̅1 , 𝜔
̅2 , . . . , 𝜔
̅ 𝑁𝑓
obtaining the following set of equations

[𝛼ℓ𝐼 ] = [𝜒] [𝐵] (5.46)


where
[𝛼ℓ𝐼 ] = [ {𝛼ℓ𝐼 (𝜔
̅1 )} {𝛼ℓ𝐼 (𝜔
̅2 )} . . . .. {𝛼ℓ𝐼 (𝜔
̅𝑁𝑓 )} ] (5.47)

[𝐵] = [ {𝐵(𝜔
̅1 )} {𝐵(𝜔
̅2 )} . . . .. {𝐵 (𝜔
̅𝑁𝑓 )} ]. (5.48)
The equations (5.46) can be written

159
𝛼1ℓ𝐼 (𝜔
̅1 ) 𝛼1ℓ𝐼 (𝜔
̅2 ) ⋯ 𝛼1ℓ𝐼 (𝜔
̅𝑁𝑓 ) (1)
𝜒1ℓ 𝜒1ℓ
(2)
⋯ 𝜒1ℓ
(𝑁)

(1) (2) (𝑁)


𝛼2ℓ𝐼 (𝜔
̅1 ) 𝛼2ℓ𝐼 (𝜔
̅2 )
⋯ 𝛼2ℓ𝐼 (𝜔
̅𝑁𝑓 ) = 𝜒2ℓ 𝜒2ℓ ⋯ 𝜒2ℓ ·
⋮ ⋮ ⋮ ⋮ ⋮ ⋮
(1) (2) (𝑁)
[𝛼𝑁ℓ𝐼 (𝜔
̅1 ) 𝛼𝑁ℓ𝐼 (𝜔 ̅𝑁𝑓 )] [𝜒𝑁ℓ
̅2 ) ⋯ 𝛼𝑁ℓ𝐼 (𝜔 𝜒𝑁ℓ ⋯ 𝜒𝑁ℓ ]
−𝑔12
−𝑔12 −𝑔12
⋯ 2
̅2
𝜔 𝑁
2 2 𝑓
̅2
𝜔 ̅2
𝜔 (1− ) +𝑔12
(1− 12 ) +𝑔12 (1− 22 ) +𝑔12 𝜔2
1
𝜔1 𝜔1

−𝑔22
−𝑔22 −𝑔22
⋯ 2
̅2
𝜔 𝑁
2 2 𝑓
· ̅2
𝜔 ̅2
𝜔 (1− ) +𝑔22 .
(1− 12 ) +𝑔22 (1− 22 ) +𝑔22 𝜔2
2
𝜔2 𝜔2

⋮ ⋮ ⋮
2
2 2 −𝑔𝑁
−𝑔𝑁 −𝑔𝑁 ⋯ 2
2 2 ̅2
𝜔
̅2
𝜔 ̅2
𝜔 𝑁
𝑓
(1− 21 ) 2
+𝑔𝑁 (1− 22 ) 2
+𝑔𝑁 (1− ) 2
+𝑔𝑁
𝜔𝑁 𝜔𝑁 𝜔2
[ 𝑁 ]
If 𝑁𝑓 = 𝑁 and the matrix [𝐵] is nonsingular, equation (5.46) yields the
matrix which contains the elements of the mode shapes (3.74)
[𝜒] = [𝛼ℓ𝐼 ] [𝐵]−1. (5.49)
The elements of the matrix [𝐵]−1 are calculated from the values of 𝜔𝑟 and
𝑔𝑟 determined experimentally and the values 𝜔̅𝑓 of the excitation frequencies. The
elements of the matrix [𝛼ℓ𝐼 ] are measured as ordinates of the diagrams of the
quadrature component of receptance in 𝑁 distinct points at 𝑁𝑓 = 𝑁 frequencies 𝜔̅𝑓 .
We obtain terms of the form (3.74)
(𝑟) (𝑟)
(𝑟) 𝛹ℓ 𝛹𝑗
𝜒𝑗ℓ = 𝑔 𝑟 𝑘𝑟
(5.50)

hence the elements of the modal vector {𝛹 (𝑟) } are proportional to the elements of
the vector {𝜒 (𝑟) }. After the normalization of the modal vectors, from equation
𝑘𝑟
(5.50) we obtain the modal stiffnesses 𝑘𝑟 , then the modal masses 𝑚𝑟 = .
𝜔2𝑟
In the original implementation, Stahle and Forlifer selected the excitation
frequencies 𝜔 ̅𝑓 equal to the undamped natural frequencies 𝜔𝑟 . They have shown
that in equation (5.46) “it is only necessary to include those modes which are
sufficiently close to cause modal interaction in the measured quadrature response.
It will seldom be necessary to include more than four modes in the equation” [113].
This method of separation is not tenable when two different modes have
equal or very close natural frequencies and have similar damping factors 𝑔𝑟 . For

160
two natural frequencies to be approximately equal, it is generally necessary that the
modes be of different types (e.g.: bending and torsion) or that the modes be
associated with distinctly different parts of the vibrating structure. In these
instances, their separation is done selecting an appropriate excitation or adopting
an identification method with multi-point excitation (see Ch.6).
If 𝑁𝑓 ≠ 𝑁, a least-squares procedure is used, obtaining the approximate
−1
solution [𝜒] = [𝛼ℓ𝐼 ] [𝐵]# where [𝐵]# = [𝐵]𝑇 [[𝐵] [𝐵]𝑇 ] .

5.3.4.5. Mode separation by regression

A first improvement of the method of Stahle and Forlifer has been


developed by Klosterman [63] and Davis [28]. In their approach, initial estimates
of the natural frequencies 𝜔𝑟 and damping factors 𝑔𝑟 are determined as in the
previous method.
Using the ordinates of the diagrams of the in-phase and quadrature
components of the response at point 𝑗, calculated at 𝑁𝑓 > 𝑁 excitation frequencies
𝜔
̅1 , 𝜔
̅2 , . . . , 𝜔
̅𝑁𝑓 of the harmonic force applied in ℓ, two sets of 𝑁 equations can be
set up, each equation having the form (3.77) or (3.78):
(𝑟)
̅ 𝑓 ) = ∑𝑁
𝛼𝑗ℓ𝑅 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑎𝑟 (𝜔
̅ 𝑓 ),
(𝑟)
(5.51)
̅ 𝑓 ) = ∑𝑁
𝛼𝑗ℓ𝐼 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑏𝑟 (𝜔
̅ 𝑓 ).

The differences between the measured vector components of


receptances 𝛼𝑗𝑅 (𝜔
̅𝑓 ) and 𝛼𝑗𝐼 (𝜔̅𝑓 ) and the values given by the analytical
expressions (5.51) are the errors 𝜀𝑅 (𝜔
̅𝑓 ) and 𝜀𝐼 (𝜔
̅𝑓 ) defined by

𝑟 ( )
𝜀𝑅 (𝜔 ̅𝑓 ) − ∑𝑁
̅𝑓 ) = 𝛼𝑗𝑅 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑎𝑟 (𝜔
̅𝑓 ),
𝑟 ( )
(5.52)
𝜀𝐼 (𝜔 ̅𝑓 ) − ∑𝑁
̅𝑓 ) = 𝛼𝑗𝐼 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑏𝑟 (𝜔
̅𝑓 )

Using a least squares procedure, the error squared is summed up over 𝑁𝑓


excitation frequencies to obtain

𝑁 𝑁 (𝑟) 2
𝑓
𝐸𝑅 = ∑𝑓=1 𝜀𝑅2 (𝜔 𝑓
̅ 𝑓 ) = ∑𝑓=1 ̅ 𝑓 ) − ∑𝑁
[𝛼𝑗𝑅 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑎𝑟 (𝜔
̅ 𝑓 )] ,
2 (5.53)
𝑁 𝑁 (𝑟)
𝑓
𝐸𝐼 = ∑𝑓=1 𝜀𝐼2 (𝜔 𝑓
̅ 𝑓 ) = ∑𝑓=1 ̅ 𝑓 ) − ∑𝑁
[𝛼𝑗𝐼 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑏𝑟 (𝜔
̅ 𝑓 )] ,

(𝑟)
and the coefficients 𝜘𝑗ℓ are determined to minimize 𝐸𝑅 and 𝐸𝐼 .
(𝑠)
Differentiating with respect to 𝜘𝑗ℓ and setting the result to zero we obtain

161
𝜕𝐸𝑅 𝑁 𝑁𝑓 (𝑟)
(𝑠)
𝑓
= −2 ∑𝑓=1 𝛼𝑗𝑅 (𝜔
̅ 𝑓 )𝑎𝑠 (𝜔 ̅ 𝑓 ) ∑𝑁
̅ 𝑓 ) + 2 ∑𝑓=1 𝑎𝑠 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑎𝑟 (𝜔
̅ 𝑓) = 0 ,
𝜕𝜘𝑗ℓ
(5.54)
𝜕𝐸𝐼 𝑁𝑓 𝑁𝑓 (𝑟)
(𝑠) = −2 ∑𝑓=1 𝛼𝑗𝐼 (𝜔
̅ 𝑓 )𝑏𝑠 (𝜔
̅ 𝑓) + 2 ∑𝑓=1 ̅ 𝑓 ) ∑𝑁
𝑏𝑠 (𝜔 𝑟=1 𝜘𝑗ℓ 𝑏𝑟 (𝜔
̅ 𝑓) =0.
𝜕𝜘𝑗ℓ

Equations (5.54) can be written under the form


𝑇 𝑇
{𝑎𝑠 (𝜔
̅ 𝑓 )} {𝛼𝑗 (𝜔
𝑅
̅ 𝑓 )} = {𝑎𝑠 (𝜔
̅ 𝑓 )} [𝑎] {𝜘𝑗ℓ }
𝑇 𝑇 (5.55)
{𝑏𝑠 (𝜔
̅ 𝑓 )} {𝛼𝑗 (𝜔
𝐼
̅ 𝑓 )} = {𝑏𝑠 (𝜔
̅ 𝑓 )} [𝑏] {𝜘𝑗ℓ },

where
𝑇
̅𝑓 )} = { 𝑎𝑠 (𝜔
{𝑎𝑠 (𝜔 ̅1 ) 𝑎𝑠 (𝜔
̅2 ) . . . 𝑎𝑠 (𝜔
̅ 𝑁𝑓 ) } ,
𝑇 (5.56)
̅𝑓 )} = { 𝑏𝑠 (𝜔
{𝑏𝑠 (𝜔 ̅1 ) 𝑏𝑠 (𝜔
̅2 ) . . . 𝑏𝑠 (𝜔
̅ 𝑁𝑓 ) } ,

(1) (2) (𝑁) 𝑇


{𝜘𝑗ℓ } = { 𝜘𝑗ℓ 𝜘𝑗ℓ . . . . . 𝜘𝑗ℓ } , (5.57)

[𝑎] = [ {𝑎1 (𝜔
̅𝑓 )} {𝑎2 (𝜔
̅𝑓 )} . . . .. {𝑎𝑁 (𝜔
̅𝑓 )} ], (5.58)
[𝑏] = [ {𝑏1 (𝜔
̅𝑓 )} {𝑏2 (𝜔
̅𝑓 )} . . . .. {𝑏𝑁 (𝜔
̅𝑓 )} ], (5.59)
𝑇
̅𝑓 )} = {𝛼𝑗𝑅 (𝜔
{𝛼𝑗𝑅 (𝜔 ̅1 ) 𝛼𝑗𝑅 (𝜔
̅2 ) . . . 𝛼𝑗𝑅 (𝜔
̅𝑁𝑓 ) } ,
𝑇 (5.60)
̅𝑓 )} = {𝛼𝑗𝐼 (𝜔
{𝛼𝑗𝐼 (𝜔 ̅1 ) 𝛼𝑗𝐼 (𝜔
̅2 ) . . . 𝛼𝑗𝐼 (𝜔
̅ 𝑁𝑓 ) } .
(𝑠)
Repeating the differentiation for the remaining values 𝜘𝑗ℓ , two sets of 𝑁
equations are obtained, known as the normal equations of the least squares method
[𝑎]𝑇 {𝛼𝑗 (𝜔
̅ 𝑓 )} = [𝑎]𝑇 [𝑎] {𝜘𝑗ℓ },
𝑅
(5.61)
[𝑏]𝑇 {𝛼𝑗 (𝜔
̅ 𝑓 )} = [𝑏]𝑇 [𝑏] {𝜘𝑗ℓ }.
𝐼

Because [𝑎] and [𝑏] are rectangular 𝑁𝑓 𝑥𝑁 matrices, the products [𝑎]𝑇 [𝑎]
and [𝑏]𝑇 [𝑏]
are square 𝑁𝑥𝑁 matrices. Premultiplying to the inverses of these
matrices, we obtain
{𝜘𝑗ℓ } = [𝑎]+ {𝛼𝑗𝑅 (𝜔
̅ 𝑓 )} (5.62)
{𝜘𝑗ℓ } = [𝑏]+ {𝛼𝑗𝐼 (𝜔
̅ 𝑓 )} (5.63)
where
−1
[𝑎]+ = [[𝑎]𝑇 [𝑎]] [𝑎]𝑇 , [𝑎]+ [𝑎] = [𝐼], (5.64)
−1
[𝑏]+ = [[𝑏]𝑇 [𝑏]] [𝑏]𝑇 , [𝑏]+ [𝑏] = [𝐼], (5.65)

162
have the form of pseudoinverse matrices, which can be calculated as shown in
§4.3.6.
The problem has been treated in the general case by Penrose [88] and is
known as the “minimum norm solution of an incompatible set of algebraic
equations” or the generalized inverse solution of equations (5.61).
Repeating measurements and calculations for all 𝑗 = 1, . . . , 𝑁 points, the
matrix [𝜘] defined by (3.81) is completely determined. Its columns are proportional
to the modal vectors {𝛹 (𝑟) }.
A similar method can be used to determine the elements of the matrix [𝜒]
defined by (3.84). However, a different procedure is presented in the following.
Equations (5.52) can be written under the form
𝑟 ( )
𝜀𝑅 (𝜔 ̅𝑓 ) − ∑𝑁
̅𝑓 ) = 𝛼𝑗𝑅 (𝜔 𝑟=1 𝜒𝑗ℓ 𝐴𝑟 (𝜔
̅𝑓 ),
𝑟 ( )
𝜀𝐼 (𝜔 ̅𝑓 ) − ∑𝑁
̅𝑓 ) = 𝛼𝑗𝐼 (𝜔 𝑟=1 𝜒𝑗ℓ 𝐵𝑟 (𝜔
̅𝑓 ).
Evaluating these errors at 𝑁𝑓 excitation frequencies, two sets of 𝑁𝑓
equations are set up
{𝜀𝑅 } = {𝛼𝑗 (𝜔
̅ 𝑓 )} − [𝒜] {𝜒𝑗ℓ },
𝑅
(5.66)
{𝜀𝐼 } = {𝛼𝑗 (𝜔
̅ 𝑓 )} − [ℬ] {𝜒𝑗ℓ },
𝐼

where
𝑇
{𝜀𝑅 } = {𝜀𝑅 (𝜔
̅1 ) 𝜀𝑅 (𝜔
̅2 ) . . . 𝜀𝑅 (𝜔
̅ 𝑁𝑓 ) } ,
𝑇
{𝜀𝐼 } = {𝜀𝐼 (𝜔
̅1 ) 𝜀𝐼 (𝜔
̅2 ) . . . 𝜀𝐼 (𝜔
̅ 𝑁𝑓 ) } ,
[𝒜] = [𝐴]𝑇 , [ℬ] = [𝐵]𝑇 , (5.67)
[𝐴] = [{𝐴(𝜔
̅1 )} {𝐴(𝜔
̅2 )} . . . {𝐴 (𝜔
̅𝑁𝑓 )}]. (5.68)

The matrix [𝐵] is defined by (5.48) and the vector {𝜒𝑗ℓ } - by (4.56).
Using the least squares method, we first calculate the sum of the errors
squared
𝑇
𝐸𝑅 = {𝜀𝑅 }𝑇 {𝜀𝑅 } = {{𝛼𝑗𝑅 (𝜔
̅ 𝑓 )} − [𝒜] {𝜒𝑗ℓ }} {{𝛼𝑗 (𝜔
𝑅
̅ 𝑓 )} − [𝒜] {𝜒𝑗ℓ }} =
𝑇 𝑇 𝑇
= {𝛼𝑗𝑅 (𝜔 ̅𝑓 )} − 2{𝜒𝑗ℓ } [𝒜]𝑇 {𝛼𝑗𝑅 (𝜔
̅𝑓 )} {𝛼𝑗𝑅 (𝜔 ̅𝑓 )} + {𝜒𝑗ℓ } [𝒜]𝑇 [𝒜]{𝜒𝑗ℓ },
(𝑟)
then differentiating with respect to 𝜒𝑗ℓ and setting the result to zero
𝜕𝐸𝑅
(𝑟) = −2{𝐼}𝑇𝑟 [𝒜]𝑇 {𝛼𝑗𝑅 (𝜔
̅ 𝑓 )} + 2{𝐼}𝑇𝑟 [𝒜]𝑇 [𝒜] {𝜒𝑗ℓ } = 0 ,
𝜕𝜒𝑗ℓ

where {𝐼}𝑟 is the 𝑟-th column of the identity matrix [𝐼].

163
The result is
𝜕𝐸𝑅
= −2 ([𝒜]𝑇 {𝛼𝑗𝑅 (𝜔
̅ 𝑓 )} − [𝒜]𝑇 [𝒜] {𝜒𝑗ℓ }) = {0} ,
𝜕{𝜒𝑗ℓ }

hence the normal equations are


[𝒜]𝑇 [𝒜] {𝜒𝑗ℓ } = [𝒜]𝑇 {𝛼𝑗 (𝜔
̅ 𝑓 )}.
𝑅

The same procedure can be set up with the imaginary part of the receptance,
obtaining
[ℬ]𝑇 [ℬ] {𝜒𝑗ℓ } = [ℬ]𝑇 {𝛼𝑗 (𝜔 ̅ 𝑓 )}.
𝐼

Therefore
{𝜒𝑗ℓ } = [𝒜]+ {𝛼𝑗𝑅 (𝜔
̅𝑓 )}, (5.69)

{𝜒𝑗ℓ } = [ℬ]+ {𝛼𝑗𝐼 (𝜔


̅𝑓 )}, (5.70)
where
−1 −1
[𝒜]+ = [[𝒜]𝑇 [𝒜]] [𝒜]𝑇 = [[𝐴][𝐴]𝑇 ] [𝐴]
−1 −1 (5.71)
[ℬ]+ = [[ℬ]𝑇 [ℬ]] [ℬ]𝑇 = [[𝐵][𝐵]𝑇 ] [𝐵]
Using the same procedure for all response points, the matrix [𝜒] is
completely determined and then the modal vectors.

5.3.4.6. Iterative mode separation method

Klosterman [63] suggested an iterative method for the separation of the


modes of vibration, based on equations

{𝜘𝑗ℓ } = [𝑎]+ {𝛼𝑗𝑅 (𝜔


̅ 𝑓 )}, (5.62)

{𝜒𝑗ℓ } = [ℬ]+ {𝛼𝑗𝐼 (𝜔


̅𝑓 )}. (5.70)
The elements of matrices [𝑎] and [ℬ] are calculated based on initial
estimations of 𝜔𝑟 and 𝑔𝑟 , from the diagrams of the quadrature component of
receptances (eventually both diagrams 𝛼ℓ𝑅 (𝜔) and 𝛼ℓ𝐼 (𝜔)) measured in a number
of response points larger than the number of effective degrees of freedom, within
the considered frequency range, where the summation of modes requires the use of
an analytical separation method. The elements of the vectors {𝛼𝑗𝑅 (𝜔 ̅𝑓 )} and
{𝛼𝑗𝐼 (𝜔
̅𝑓 )} are measured from the same diagrams.
(𝑟) (𝑟)
Calculating 𝜘𝑗ℓ and 𝜒𝑗ℓ from equations (5.62) and (5.70), a new value is
obtained for the damping factor

164
𝜘(𝑗ℓ𝑟)
𝑔𝑟 = (5.72)
𝜒(𝑗ℓ𝑟)
(𝑟)
which is used in an iterative process converging to the true values of 𝜘𝑗ℓ and 𝑔𝑟 .
The modal vectors {𝛹 (𝑟) } can be determined repeating the procedure for
various response points.
A technique very similar to the above has been used for machine tool
applications by Morse, Shapton and Wood [76]. They report that the accuracy of
results depends largely on the accuracy of the initial assumptions on the undamped
natural frequencies 𝜔𝑟 .
Flanelly, Berman and Giansante [36] have successfully applied an iterative
method for the identification of incomplete models based solely on the diagrams of
the quadrature response.
Evaluating expression (3.80) at 𝑁𝑓 > 𝑁 excitation frequencies
𝜔̅1 , 𝜔
̅2 , . . . , 𝜔
̅𝑁𝑓 , they obtained the following set of equations
(𝑟)
`𝛹
[𝛼ℓ𝐼 ] = [𝜘] [𝑏]𝑇 = [𝛹] [ 𝑘ℓ 𝑏𝑟𝑚𝑎𝑥 ˏ] [𝑏0 ] (5.73)
𝑟

in which [𝛹] is a rectangular (𝑁𝑚 𝑥𝑁) matrix. The testing conditions imposed by
this method require that the number of response measurement points (hence the
number of rows of the matrix[𝛼ℓ𝐼 ]) be larger than the effective number of degrees
of freedom (𝑁𝑚 > 𝑁). The matrix [𝜘] is given by expression (3.81). The matrix
[𝑏0 ] is obtained from [𝑏]𝑇 normalizing each row by division to the maximum value
𝑏𝑟𝑚𝑎𝑥 .
The elements of the matrix [𝛼ℓ𝐼 ] are measured at the 𝑁𝑓 excitation
frequencies. Matrices [𝛹] and [𝑏0 ] are unknown. The method of the pseudoinverse
matrix is used in a double iteration as follows:
1) A first value of [𝑏0 ] is calculated based on approximate values 𝜔𝑟 and
𝑔𝑟 determined experimentally by the methods exposed above.
2) The matrix [𝛹] is calculated from.

[𝛹] = [𝛼ℓ𝐼 ][𝑏0 ]# [`


𝑘𝑟
(𝑟)
𝛹ℓ 𝑏𝑟𝑚𝑎𝑥
ˏ] (5.74)

where
−1
[𝑏0 ]# = [𝑏0 ]𝑇 [[𝑏0 ][𝑏0 ]𝑇 ] , [𝑏0 ][𝑏0 ]# = [𝐼] . (5.75)

3) An improved value of the matrix [𝑏0 ] is determined from

[𝑏0 ] = [`
𝑘𝑟 +
(𝑟)
𝛹ℓ 𝑏𝑟𝑚𝑎𝑥
ˏ] [𝛹] [𝛼ℓ𝐼 ] , (5.76)

165
in which
−1
[𝛹]+ = [[𝛹]𝑇 [𝛹]] [𝛹]𝑇 , [𝛹]+ [𝛹] = [𝐼], (5.77)
then the procedure is repeated from point 2, normalizing each solution until
convergence is obtained.
The calculation can be simplified choosing
` 𝑘𝑟
[ (𝑟)
𝛹ℓ 𝑏𝑟𝑚𝑎𝑥
ˏ] = [𝐼] .
From (5.76) we then obtain [𝑏]𝑇 , hence the elements
1
𝑏 (𝜔
𝑘𝑟 𝑟
̅𝑓 ) = ℐ𝑚 𝛼̅𝑟 (𝜔 ̅𝑓 ).
̅𝑓 ) = 𝛼𝑟𝐼 (𝜔

From an analogous relationship, based on the matrix [𝛼ℓ𝑟 ], we calculate


ℛℯ 𝛼
̅ 𝑟 (𝜔
̅ 𝑓 ) = 𝛼𝑟𝑅 (𝜔
̅ 𝑓 ) , then
𝛼𝑟𝑅 (𝜔
̅ 𝑓) ̅2
𝜔
̅ (𝜔
ℛℯ 𝛽 ̅ 𝑓 ) = 𝛽𝑟 (𝜔
̅ 𝑓) = 2 = 𝑘𝑟 (1 − 𝜔𝑓2) .
𝑟 𝑅 𝛼 ̅ 𝑓 )+𝛼2𝑟 (𝜔
𝑟𝑅 (𝜔 ̅ 𝑓) 𝑟
𝐼

Drawing the graph of the function 𝛽𝑟𝑅 (𝜔 ̅𝑓 ) for several values 𝜔


̅𝑓 , the
undamped natural frequencies 𝜔𝑟 are determined as the abscissae of the points
where 𝛽𝑟𝑅 = 0. From the above relation we obtain 𝑘𝑟 , then
̅ 𝑓2
𝜔 𝛼𝑟𝐼 𝑘𝑟
𝑔𝑟 = (𝜔2 − 1) 𝛼 𝑟𝑅
, 𝑚𝑟 =
𝜔2𝑟
.
𝑟

5.3.4.7. Modal matrix elimination method

Natke [77] generalized the phase separation method for the case when the
undamped natural frequencies 𝜔𝑟 and the damping factors 𝑔𝑟 are unknown.
The procedure starts from equations (3.79) and (3.80), defined for 𝑁𝑚 > 𝑁
response measurement points
{𝛼ℓ𝑅 (𝜔)} = [𝜒] {𝐴(𝜔)} , {𝛼ℓ𝐼 (𝜔)} = [𝜒] {𝐵(𝜔)}.
These expressions are evaluated at 𝑁𝑓 = 𝑁 excitation frequencies
𝜔
̅1 , 𝜔
̅2 , . . . , 𝜔
̅𝑁𝑓 , and the following sets of equations are set up

[𝛼ℓ𝑅 ] = [𝜒] [𝐴] , (5.78)


[𝛼ℓ𝐼 ] = [𝜒] [𝐵] , (5.79)
where [𝐴] and [𝐵] are 𝑁𝑥𝑁 square matrices, and [𝛼ℓ𝑅 ], [𝛼ℓ𝐼 ] and [𝜒] are
rectangular (𝑁𝑚 𝑥𝑁) matrices.

166
In equations (5.78) and (5.79) the notations (3.84), (5.47), (5.48) and
(5.68) have been used, as well as
[𝛼ℓ𝑅 ] = [ {𝛼ℓ𝑅 (𝜔
̅1 )} {𝛼ℓ𝑅 (𝜔
̅2 )} . . . .. {𝛼ℓ𝑅 (𝜔
̅𝑁𝑓 )} ]. (5.80)

The matrix [𝜒] is eliminated between equations (5.78) and (5.79).


Postmultiplying (5.79) by the inverse of [𝐵] we obtain
[𝛼ℓ𝐼 ][𝐵]−1 = [𝜒] , (5.81)
which substituted into equation (5.78) yields
[𝛼ℓ𝑅 ] = [𝛼ℓ𝐼 ][𝐵]−1 [𝐴] . (5.82)
𝑇
Premultiplying by [𝛼ℓ𝐼 ]
𝑇 𝑇
[𝛼ℓ𝐼 ] [𝛼ℓ𝑅 ] = [𝛼ℓ𝐼 ] [𝛼ℓ𝐼 ][𝐵]−1 [𝐴]
−1
𝑇
then by [[𝛼ℓ𝐼 ] [𝛼ℓ𝐼 ]] (because [𝛼ℓ𝐼 ] is rectangular) equation (5.82) becomes
−1
𝑇 𝑇
[[𝛼ℓ𝐼 ] [𝛼ℓ𝐼 ]] [𝛼ℓ𝐼 ] [𝛼ℓ𝑅 ] = [𝐵]−1 [𝐴].

Premultiplying by [𝐵] we obtain


[𝐴] = [𝐵] [𝒞] , (5.83)
where
+
[𝒞] = [𝛼ℓ𝐼 ] [𝛼ℓ𝑅 ] (5.84)
in which
−1
+ 𝑇 𝑇
[𝛼ℓ𝐼 ] = [[𝛼ℓ𝐼 ] [𝛼ℓ𝐼 ]] [ 𝛼ℓ 𝐼 ] .

From (3.72) and (3.73) the following relationship can be established


between the elements of the matrices [𝐴] and [𝐵]

𝐴2𝑟 (𝜔
̅𝑓 ) = −𝐵𝑟 (𝜔
̅𝑓 ) [1 + 𝐵𝑟 (𝜔
̅𝑓 )]. (5.85)
From (5.83) we obtain
̅𝑓 ) = ∑𝑁
𝐴𝑟 (𝜔 𝑒=1 𝐵𝑟 (𝜔
̅𝑒 ) 𝒞𝑒𝑓 (5.86)
where 𝒞𝑒𝑓 are the elements of the experimentally measured matrix [𝒞] = [𝒞𝑒𝑓 ].
Substituting (5.86) into (5.85), 𝑁 sets of 𝑁 equations in 𝐵𝑟 (𝜔̅𝑓 ) are
obtained, hence 𝑁 2 equations
2
𝐵𝑟 (𝜔 ̅𝑓 )] + (∑𝑁
̅𝑓 ) [1 + 𝐵𝑟 (𝜔 𝑒=1 𝐵𝑟 (𝜔
̅𝑒 ) 𝒞𝑒𝑓 ) = 0, (𝑟, 𝑓 = 1, . . , 𝑁) (5.87)

167
which yield the 𝑁 2 elements of the matrix [𝐵].
The elements 𝐵𝑟 (𝜔
̅𝑓 ) depend solely on the undamped natural frequencies
𝜔𝑟 and the damping factors 𝑔𝑟 .
Denoting
2
1−𝜔2
𝜔𝑟
𝜏𝑟 (𝜔) = , (5.88)
𝑔𝑟
we can write
𝐵𝑟 (𝜔) = − 𝜏2(𝜔1)+1 (5.89)
𝑟
hence we can calculate
1
𝜏𝑟 (𝜔) = ±√− [1 + 𝐵 (𝜔)]. (5.90)
𝑟

Expression (5.88) is evaluated at two different excitation frequencies, 𝜔


̅𝑒
and 𝜔
̅𝑓 , to obtain
̅2
𝜔 ̅2
𝑓 −𝜔𝑒
𝑔𝑟 = , (5.91)
̅2
𝜔𝑓 𝜏𝑟 (𝜔 ̅2
̅ 𝑒 )−𝜔𝑒 𝜏𝑟 (𝜔
̅ 𝑓)

̅2
𝜔𝑓 𝜏𝑟 (𝜔 ̅2
̅ 𝑒 )−𝜔𝑒 𝜏𝑟 (𝜔
̅ 𝑓)
𝜔𝑟2 = . (5.92)
𝜏𝑟 (𝜔
̅ 𝑒 ) − 𝜏𝑟 ( 𝜔
̅ 𝑓)

Based on these values, we construct the matrix [𝐵]. If it is non-singular,


the modal matrix is obtained from equation (5.79)
[𝜒] = [𝛼ℓ𝐼 ] [𝐵]−1 , (5.93)
The modal vectors are then normalized setting the element of maximum
absolute value equal to unity
{𝜒(𝑟) }
{𝛹 (𝑟) } = 𝑟) . (5.94)
𝜒(𝑗,𝑚𝑎𝑥
The modal stiffnesses are
(𝑟)
𝛹ℓ
𝑘𝑟 = 𝑟) (5.95)
𝑔𝑟 𝜒(𝑗,𝑚𝑎𝑥
and the modal masses
𝑘𝑟
𝑚𝑟 = .
𝜔2𝑟

5.3.5. Identification Methods Based on the Polar Plot of the Response

A convenient form of representation of the structural response are the polar


plots drawn on the Argand plane as loci of affixes of a complex frequency response
function. The graphical analysis methods of these Nyquist plots by procedures
(circle-fit, maximum frequency spacing) suggested by Kennedy and Pancu [6]

168
proved to be the most accurate for determining the dynamic parameters and modal
shapes of a complex structure, especially in the presence of close natural
frequencies [85]. The method has been improved by using excitation with forces in
quadrature [90] or by the numerical analysis of the experimentally measured plots
[42, 43].
Several frequently used graphical analysis methods are presented in the
following.
The polar plot of the complex receptance of a single-degree-of-freedom
system with hysteretic damping is a circle (Fig.2.14) which passes through the
origin of coordinate axes and has the center on the negative imaginary semiaxis.
The graphical analysis of these diagrams is presented in § 2.2.1.2.
The polar plots of multi-
degree-of-freedom systems
consist of several loops, generally
one loop for each mode of
vibration. Figure 5.26 shows the
polar plot measured on a milling
machine, using vertical excitation
between the tool holder and the
working piece.
The graphical analysis of
these diagrams is done fitting a
circle to each loop and using the
expressions of modal parameters
established for single-degree-of-
freedom systems. Generally, apart
from the assumption of
proportional damping, it is
considered that in the immediate vicinity of a resonance, the contribution of the off-
resonant modes is negligible or constant (independent of frequency) and the
equation of response curve has the form (5.23).
If specific points are marked on each loop of the polar plot at equal
increments ∆𝜔 of the excitation frequency, the resonance is located where the
length ∆𝑠 of the arc between two successive points is a maximum, hence where the
vector radius sweeps out a maximum angle for a given increase ∆𝜔 of the frequency
(Fig.5.27). The best circle is then fitted through the points in the neighborhood of
each resonance.
Let 𝑀 be the point which defines the frequency of resonance 𝜔𝑟 on the
loop corresponding to the 𝑟-th mode. The diameter at resonance is
𝜕𝑞𝑗
̅̅̅̅̅̅ = 1 𝜕𝑞ℓ
𝑂′𝑀 =
1 (𝑟) (𝑟)
𝛹ℓ 𝛹𝑗 . (5.96)
ℎ𝑟 𝜕𝑝𝑟 𝜕𝑝𝑟 ℎ𝑟
Point 𝑂′ is the “displaced origin” of the vector 𝑂′𝑃, which represents the
displacement in the 𝑟-th mode at a frequency 𝜔. The vector 𝑂𝑂′ represents the

169
contribution of the non-resonant modes to the total response 𝑂𝑃 at the resonance in
the 𝑟-th mode. Vector 𝑂′𝑀 represents the peak amplitude of vibration in the 𝑟-th
mode. The mode shape {𝛹 (𝑟) } is obtained from the ratios of the diameters 𝑂′𝑀 of
the appropriate circles drawn on the plots at various coordinates 𝑞𝑗 , due to a given
excitation at a fixed station 𝑞ℓ .
∆𝑠
The resonance location where the ratio ∆𝜔 is a maximum is independent of
the coordinate axes and is valid for nonproportional damping too (see § 5.4)
Because for proportional damping and constant contribution of the non-
resonant modes to the total response, the resonance diameter 𝑂′𝑀 remains parallel
to the imaginary axis, the location of frequency 𝜔𝑟 can be done using the criterion
of maximum quadrature component of receptance. This criterion is implicitly
related to the correct positioning of coordinate axes and must be used with care in
the case of antiresonances.

The modal damping can be found by drawing the diameter 𝐵𝐶


perpendicular to 𝑂′𝑀. It intersects the circle at the half power points 𝐵 and 𝐶, of
frequencies 𝜔𝑟′ and 𝜔𝑟′′ . The damping factor is given by expression (5.20).
For polar plots of the form shown in Fig.5.28, where the frequencies 𝜔𝑟′
′′
and 𝜔𝑟 cannot be determined, the damping factor is calculated from
𝜔2𝑄 −𝜔2𝑃 1
𝑔𝑟 = (5.97)
𝜔2𝑟 𝜃2 −𝜃1
where 𝜔𝑄 and 𝜔𝑃 are the frequencies of the points 𝑄 and 𝑃 in the neighborhood of
resonance, whose vector radii subtend angles 𝜃2 and 𝜃1 , respectively, with the
positive real semiaxis.
̂ = 𝑀𝑂
If the angles 𝑃𝑂′𝑀 ̂ ′ 𝑄 = 𝜃, then

170
𝜔2𝑄 −𝜔2𝑃
𝑔𝑟 = 𝑐𝑜𝑡𝑎𝑛 𝜃. (5.98)
2 𝜔2𝑟
Other formulae for the calculation of the damping factor can be found in
§ 2.2.1.2.
The fact that the measurements are carried out at approximately constant
displacement amplitude is useful at the identification of nonlinear systems.
Modal masses are calculated using relations of the form
1
𝑚𝑟 = (5.99)
𝜔2𝑟 𝑔𝑟 ̅̅̅̅̅̅̅
𝑂′𝑀

after a “normalization” of the diameters ̅̅̅̅̅̅


𝑂′𝑀.
The analysis of the polar plots of mobility or inertance is carried out
analogously.
Figure 5.29 is part of the polar plot of the vertical displacement of the tool
holder of a milling machine, for harmonic excitation with a vertical force 𝑓̂ = 36𝑁
Using the criterion of maximum frequency spacing, the first frequency of resonance
is located at 𝜈1 = 161.9 Hz. The best circle is drawn through the points around the
resonance. The diameter 𝑂1′ 𝑀1 corresponds to a displacement of 8.25 𝜇𝑚.The
diameter 𝐵1 𝐶1 , perpendicular to 𝑂1′ 𝑀1, locates the frequencies 𝜈1′ = 157.2 Hz and
𝜈1′′ = 166.8 Hz. This yields the damping factor
𝜈′′1 −𝜈′1
𝑔1 = 𝜈1
= 166.8−157.2
161.9
= 0.0593 ,

the modal mass


𝑓̂ 36
𝑚1 = ̅̅̅̅̅̅̅̅̅̅ = = 71.1 𝑘𝑔 ,
𝜔21 𝑔1 𝑂′1 𝑀1 4·𝜋2 ·161.92 ·0.0593·8.25·10−6

and the modal stiffness


𝑁
𝑘1 = 𝑚1 𝜔12 = 71.1 · 4 · 𝜋 2 · 161.92 = 73.6 · 106 .
𝑚
The same procedure is used for the second loop of the diagram. The
frequency of resonance is 𝜈2 = 250 Hz. The diameter 𝑂2′ 𝑀2 corresponds to a
displacement of 4.79 𝜇𝑚.The diameter 𝐵2 𝐶2, locates the frequencies 𝜈2′ =
242.6 Hz and 𝜈2′′ = 256.7 Hz. This yields
𝜈′′2 −𝜈′2
𝑔2 = 𝜈2
= 256.7−242.6
250
, = 0.0564 ,

𝑓̂ 36
𝑚2 = ̅̅̅̅̅̅̅̅̅̅ = 2 = 54.1 𝑘𝑔 ,
𝜔22 𝑔2 𝑂′2 𝑀2 4·𝜋2 ·250 ·0.0564·4.79·10−6

𝑁
𝑘2 = 𝑚2 𝜔22 = 54.1 · 4 · 𝜋 2 · 2502 = 133.3 · 106 .
𝑚

171
5.3.6. “Modal Path” Method

Under this name, Raney and Howlett suggested a direct identification


method, applicable to “structures having light damping and modes of vibration
sufficiently uncoupled and frequency-separated to be represented by single-degree-
of-freedom analysis” [98]. The method is based on a set of uncoupled differential
equations, which define a “force-displacement” transfer relationship between two
specific points of the structure, so that the mathematical model is valid only for the
two points of excitation-response considered.
In § 3.2.1 it was shown how, starting from the coupled equations of motion
of a structure with viscous damping
[𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} ,
using the transformation of coordinates
{𝑞} = [𝛹] {𝑝} (5.100)
and neglecting the off-diagonal damping terms, we obtain uncoupled equations of
motion
[ˋ𝑚ˏ] {𝑝̈ } + [ˋ𝑐ˏ] {𝑝̇ } + [ˋ𝑘ˏ] {𝑝} = [𝛹]𝑇 {𝑓}.
If an external force 𝑓ℓ (𝑡) is applied at point ℓ , the equation of motion for
the 𝑟-th principal coordinate is
(𝑟)
𝑚𝑟 𝑝̈𝑟 + 𝑐𝑟 𝑝̇2 + 𝑘𝑟 𝑝𝑟 = 𝛹ℓ 𝑓ℓ (𝑡). (5.101)
From (5.100) it comes out that the response in the 𝑟-th mode is
{𝑞 (𝑟) } = {𝛹 (𝑟) } 𝑝𝑟 ,
hence the displacement at coordinate 𝑗 in the 𝑟-th mode is
(𝑟) (𝑟)
𝑞𝑗 = 𝛹𝑗 𝑝𝑟 .
Solving for 𝑝𝑟 we obtain
𝑞(𝑗𝑟)
𝑝𝑟 = (𝑟) . (5.102)
𝛹𝑗
(𝑟)
Substituting (5.102) into (5.111) and dividing by 𝛹ℓ we obtain
(𝑟) (𝑟) (𝑟) (𝑟) (𝑟) (𝑟)
𝑀ℓ𝑗 𝑞̈ 𝑗 + 𝐶ℓ𝑗 𝑞̇ 𝑗 + 𝐾ℓ𝑗 𝑞𝑗 = 𝑓ℓ (𝑡) (5.103)
where the modal path mass, damping and stiffness are
(𝑟) 𝑚𝑟 (𝑟) 𝑐𝑟 (𝑟) 𝑘𝑟
𝑀ℓ𝑗 = (𝑟) (𝑟) , 𝐶ℓ𝑗 = (𝑟) (𝑟) , 𝐾ℓ𝑗 = (𝑟) (𝑟) . (5.104)
𝛹ℓ 𝛹𝑗 𝛹ℓ 𝛹𝑗 𝛹ℓ 𝛹𝑗

172
Equation (5.103) represents the response at position 𝑗, in the 𝑟-th, due to a
force at position ℓ. Solving the 𝑁 equations (5.103) corresponding to the 𝑁 modes
of vibration, and summing up the solutions, we obtain the total response at 𝑗 due to
the force applied at ℓ
(𝑟)
𝑞𝑗 (𝑡) = ∑𝑁
𝑟=1 𝑞𝑗 (𝑡). (5.105)
Equations (5.103) define the so-called “modal paths”, because between
points ℓ and 𝑗 (Fig.5.30) we can establish 𝑁 modal paths, corresponding to the 𝑁
(𝑟)
modes of vibration. Each path has a response characteristic 𝐻ℓ𝑗 (𝜔) which can be
determined from the associated equation (5.103). Similar paths can be established
between point ℓ and other points of interest of the structure.

The equations of the modal paths have the following characteristics:


a – There is one equation for each path and mode.
b – All equations are decoupled.
c – All equations contain the same excitation force that can have any time
dependence.
(𝑟) (𝑟) (𝑟) (𝑟) (𝑟)
d – Modal path parameters 𝑀ℓ𝑗 , 𝐶ℓ𝑗 , 𝐾ℓ𝑗 are positive when 𝛹ℓ and 𝛹𝑗
are in-phase, and negative when they are out-of-phase.
e – If the modal path parameters are known, the response can be predicted
for arbitrary values of the input force 𝑓ℓ (𝑡).
The evaluation of the modal path parameters (5.104) can be accomplished
using harmonic excitation. The structure is excited at a point by a force
𝑓ℓ (𝑡) = 𝑓̂ℓ 𝑠𝑖𝑛 𝜔𝑡 (5.106)

173
and the response is measured at the points of interest. Sweeping the frequency so
that the range of measurements entails all important resonances of the structure,
frequency response curves are drawn like that from Fig.5.31.
If the frequency 𝜔 is near a resonance frequency 𝜔𝑟 , then one mode is
dominant and the total response will be
(𝑟)
𝑞𝑗 = 𝑞̂𝑗 𝑠𝑖𝑛(𝜔𝑡 − 𝜑ℓ𝑗 ). (5.107)
Substituting (5.106) and (5.107) into (5.103), the following results are
obtained
(𝑟) (𝑟) 𝑓̂
𝐾ℓ𝑗 − 𝜔2 𝑀ℓ𝑗 = 𝑞̂ℓ 𝑐𝑜𝑠 𝜑ℓ𝑗 , (5.108)
𝑗

(𝑟) 𝑓̂
𝐶ℓ𝑗 = 𝜔𝑞̂ℓ 𝑠𝑖𝑛 𝜑ℓ𝑗 . (5.109)
𝑗

𝑞̂𝑗
Measuring | ̂ | and 𝜑ℓ𝑗 at two frequencies near each resonance, the
𝑓ℓ
three parameters (5.104) can be calculated from (5.108) and (5.109).
If only a single amplitude-frequency curve is available (Fig.5.31,a), the
modal path equation to be solved is
2
(𝑟) (𝑟) 2 (𝑟) 2 ̂
𝑓
(𝐾ℓ𝑗 − 𝜔2 𝑀ℓ𝑗 ) + (𝜔 𝐶ℓ𝑗 ) = ( 𝑞̂ ) . ℓ
(5.110)
𝑗

The values of the modal path parameters (5.104) are determined measuring
𝑞̂𝑗
the receptance | ̂ | at three frequencies near each resonance.
𝑓ℓ
The method has been successfully applied to the dynamic analysis of the
reduced scale models of the space complex Apollo-Saturn 5 [98], and afterwards
by Hillyer [54] and Ibáñez [55].
If the identification is applied to the construction of a lumped-mass model
of the tested item, the following steps have to be taken:
The modal mass, modal stiffness and modal damping coefficient are first
computed in terms of the parameters of the modal paths from expressions
(𝑟) (𝑟) (𝑟)
𝑚𝑟 = 𝑀ℓ𝑗 𝛹ℓℓ 𝛹ℓ𝑗 ,
(𝑟) (𝑟) (𝑟)
𝑘𝑟 = 𝐾ℓ𝑗 𝛹ℓℓ 𝛹ℓ𝑗 , (5.111)
(𝑟) (𝑟) (𝑟)
𝑐𝑟 = 𝐶ℓ𝑗 𝛹ℓℓ 𝛹ℓ𝑗 ,
(𝑟)
where 𝛹ℓ𝑗 are normalized displacements, calculated at resonance frequencies,
using a formula of the type
(𝑟) 𝑅𝑒𝑐𝑒𝑝𝑡𝑎𝑛𝑐𝑒 𝑎𝑡 𝑗, 𝑓𝑜𝑟 𝜔=𝜔
𝛹ℓ𝑗 = 𝑅𝑒𝑐𝑒𝑝𝑡𝑎𝑛𝑐𝑒 𝑎𝑡 ℓ, 𝑓𝑜𝑟 𝜔=𝜔𝑟 .
𝑟

174
Hence
(𝑞̂𝑗 ⁄𝑓̂ℓ )
(𝑟) (𝑟) 𝜔=𝜔𝑟
𝛹ℓℓ = 1, 𝛹ℓ𝑗 = (𝑞̂ , (5.112)
ℓ ⁄𝑓̂ℓ )𝜔=𝜔
𝑟

𝑞̂𝑗
where the ratio ̂ should be negative if the motion at 𝑗 is out-of-phase with the
𝑓ℓ
force applied at ℓ.
Constructing the modal matrix
[𝛹ℓ ] = [ {𝛹ℓ(1) } {𝛹ℓ(2) } . . . . {𝛹ℓ(𝑁) } ] (5.113)
where
(𝑟) (𝑟) (𝑟) (𝑟) (𝑟) 𝑇
{𝛹ℓ } = { 𝛹ℓ1 𝛹ℓ2 . .. 𝛹ℓ𝑗 . . . 𝛹ℓℓ } ,
the matrices of the spatial model can be calculated from the following expressions
[𝑀] = [𝛹ℓ ]−𝑇 [`𝑚ˏ] [𝛹ℓ ]−1,
[𝐾] = [𝛹ℓ ]−𝑇 [`𝑘ˏ] [𝛹ℓ ]−1 , (5.114)
[𝐶] = [𝛹ℓ ]−𝑇 [`𝑐ˏ] [𝛹ℓ ]−1.

5.4. IDENTIFICATION OF SYSTEMS WITH NONPROPORTIONAL


DAMPING

5.4.1. Systems with Hysteretic Damping

5.4.1.1. Polar plot method

For single point excitation, the harmonic response of a linear system with
nonproportional hysteretic damping (3.201) can be written under the form
(𝑟) (𝑟)
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 (𝑥𝑗 + 𝑖 𝑦𝑗 ) 𝛼̅𝑟 (5.115)
where
1
𝛼̅𝑟 = 2 (5.116)
1−𝜔2+𝑖 𝑔𝑟
𝜔𝑟

is the complex modal receptance.


Considering one term in the summation (5.115), of interest is the type of
plot which is obtained from the above type of expression.
The response at point 𝑗, in the 𝑟-th mode, has the expression
(𝑟) (𝑟) (𝑟)
𝛼̅𝑗ℓ = (𝑥𝑗 + 𝑖 𝑦𝑗 ) 𝛼̅𝑟 . (5.117)

175
If only the expression (5.116) is represented in the Argand plane, we obtain
the polar plot from Fig.5.32,a.
Figure 5.32,b shows the effect of the multiplication of expression (5.116)
(𝑟) (𝑟)
by the complex number 𝑥𝑗 + 𝑖 𝑦𝑗 . The polar plot of the receptance (5.117) is
1 (𝑟)2 (𝑟)2
still a circle, but the diameter ̅̅̅̅̅
𝑂𝑀 is equal to 𝑔 √𝑥𝑗 + 𝑦𝑗 and is inclined an
𝑟
(𝑟)
(𝑟) 𝑦𝑗
angle 𝜃𝑗 = 𝑡𝑎𝑛−1 (𝑟) with respect to the negative imaginary semiaxis.
𝑥𝑗
Multiplication of the receptance 𝛼̅𝑟 by a complex number results in a rotation by
(𝑟)
an angle 𝜃𝑗 and expansion or contraction by an amount √𝑥𝑗(𝑟)2 + 𝑦𝑗(𝑟)2 .
This has been first reported by Woodcock [37] and studied in detail by
Klosterman [63]. The same effect is produced by the excitation with forces in
quadrature (§ 2.3).

Because the frequency distribution along the curve is not modified, the
criterion of resonance location at the maximum frequency spacing can be applied
to systems with nonproportional damping and widely spaced natural frequencies.
For systems with many degrees of freedom, the polar plot of the receptance
has the form shown in Fig.5.33. It results by the summation of several diagrams of
the type shown in Fig.5.32. Each loop can be approximated by a circle, which
describes the motion in a given mode of vibration.
The presence of several modes of vibration produces the “displacement of
the origin” of each circle. Like for systems with proportional damping, the vector
of displaced origin is a measure for the contribution of nonresonant modes to the
total response in the frequency range around the resonance in a mode.
First, the resonance is located using the criterion of Kennedy and Pancu
(see § 2.2.1.2) and the “best circle” is fitted through the points near the resonance.
Then, the diameter ̅̅̅̅̅̅
𝑂′𝑀 = 2𝑅 is drawn, which is a measure for the response in the
associated mode.
Provided correct measurements are made, the inclination of this diameter
is a proof of the presence of nonproportional damping (As shown in § 5.3.5 for

176
systems with proportional damping and well-spaced resonances, the diameters of
the modal circles are parallel to the imaginary axis).
The damping factor 𝑔𝑟 can be calculated using one of the formulae (5.20),
(5.97) or (5.98).
If the experimental equipment does not ensure a good frequency resolution,
the resonance location by the method of Kennedy and Pancu cannot be accurate,
particularly when the mode of interest is lightly damped. In such cases it is
recommended to use the maximum component method [63].
For a mode of vibration sufficiently separated from the others, on the
associated polar plot it is possible to determine the frequencies 𝜔𝐼 and 𝜔𝐼𝐼𝐼 of
maximum in-phase response, and the frequency 𝜔𝐼𝐼 of maximum out-of-phase
response (Fig.5.34). It is assumed that
the off-resonant terms are constant
over a range of frequencies in the
vicinity of 𝜔𝑟 , including the range
𝜔Ⅰ ­ 𝜔Ⅲ .
The angles of the vector radii
corresponding to the frequencies 𝜔𝐼 ,
𝜔Ⅱ and 𝜔Ⅲ satisfy the equations [96]
𝜋
𝜑Ⅰ = 𝜑Ⅱ − , (5.118)
4
𝜋
𝜑Ⅲ = 𝜑Ⅱ + 4 . (5.119)
Equation (5.118) can be
written
(𝑟) (𝑟) 𝜋
𝜑Ⅰ + 𝜃𝑗 = 𝜑Ⅱ + 𝜃𝑗 −
4

hence
(𝑟 )
(𝑟) (𝑟) 𝜋 𝑡𝑎𝑛(𝜑Ⅱ +𝜃𝑗 )−1
𝑡𝑎𝑛 (𝜑Ⅰ + 𝜃𝑗 ) = 𝑡𝑎𝑛 [(𝜑Ⅱ + 𝜃𝑗 ) − ] = (𝑟 ) (5.120)
4 𝑡𝑎𝑛(𝜑Ⅱ +𝜃𝑗 )+1
or, using expression (3.202),
𝜔 2
𝑔𝑟 −(1− Ⅱ2)
𝑔𝑟 𝜔𝑟
𝜔2
= 𝜔2
(5.121)
1− Ⅰ2 𝑔𝑟 +(1− Ⅱ )
𝜔𝑟 𝜔2
𝑟

because the frequency distribution is the same as for the non-rotated circle.
Equation (5.121) can be written under the form
𝜔 2
𝜔Ⅰ 2 𝜔𝑟 2 1+𝑔𝑟 −( Ⅱ )
𝜔𝑟
(𝜔 ) = (𝜔 ) [1 + 𝑔𝑟 𝜔 2
]. (5.122)
Ⅱ Ⅱ 1−𝑔𝑟 −( Ⅱ )
𝜔𝑟

177
Analogously, equation (5.119) yields

𝜔 2
𝜔Ⅲ 2 𝜔𝑟 2 1−𝑔𝑟 −( Ⅱ )
𝜔𝑟
( 𝜔 ) = (𝜔 ) [1 − 𝑔𝑟 𝜔Ⅱ 2
]. (5.123)
Ⅱ Ⅱ 1+𝑔𝑟 −( )
𝜔𝑟

A graph (Fig.5.35) can be set up from equations (5.122) and (5.123) and
used to find 𝜔𝑟 and 𝑔𝑟 , knowing the frequencies 𝜔Ⅰ , 𝜔Ⅱ and 𝜔Ⅲ . First, the ratios
𝜔Ⅲ ⁄𝜔Ⅱ and 𝜔Ⅰ ⁄𝜔Ⅱ are calculated for various values of 𝑔𝑟 and 𝜔𝑟 ⁄𝜔𝐼𝐼 . For each
pair of values, a point is plotted on the graph from Fig.5.35. Then the points with
𝑔𝑟 = 𝑐𝑜𝑛𝑠𝑡. and those with 𝜔Ⅱ ⁄𝜔𝑟 = 𝑐𝑜𝑛𝑠𝑡. are connected by smooth lines.
(𝑟)
Because equations (5.122) and (5.123) are independent of the angle 𝜃𝑗 ,
the method can also be applied when the diagram is rotated due to some constant
phase shifts introduced by the measurement system.
If 𝜔𝑟 is known, the point 𝑀 can be accurately located on the loop of the
(𝑟)
polar plot. Drawing the diameter 𝑂′𝑀, we can measure the angle 𝜃𝑗 .
If there are too little points on the diagram, the point 𝑀 cannot be located
(𝑟)
with accuracy and the angle 𝜃𝑗 can be calculated with the formula

178
2 𝑔𝑟
𝜔2
1− Ⅱ
(𝑟) 𝜔2𝑟
𝜃𝑗 = 𝑡𝑎𝑛−1 (5.124)
𝑔2
𝑟
1− 2
𝜔2
(1− Ⅱ )
𝜔2 𝑟
obtained from the condition
𝜕 (𝑟)
𝜕𝜔
(ℐ𝑚 𝛼̅𝑗ℓ ) = 0 at 𝜔 = 𝜔Ⅱ .
Evaluating expression (5.121) at the frequency 𝜔 = 𝜔𝑟 we obtain
(𝑟) (𝑟) (𝑟) (𝑟)
(𝑟) 𝑥𝑗 +𝑖 𝑦𝑗 𝑦𝑗 𝑥𝑗
𝛼̅𝑗ℓ |𝜔=𝜔𝑟 = = −𝑖 . (5.125)
𝑖 𝑔𝑟 𝑔𝑟 𝑔𝑟

(𝑟)
Measuring the diameter 2𝑅 and the angle 𝜃𝑗 , from Fig.5.34 we obtain
𝑦𝑗 𝑟
( )
(𝑟)
= 2𝑅 𝑠𝑖𝑛 𝜃𝑗 , (5.126)
𝑔𝑟

𝑥𝑗 𝑟
( )
(𝑟)
𝑔𝑟
= 2𝑅 𝑐𝑜𝑠 𝜃𝑗 . (5.127)
̅𝑟 = 1, equation (3.200) yields
Considering 𝑀
(𝑟) (𝑟) (𝑟) (𝑟)
𝑤ℓ 𝑤𝑗 = 𝜔𝑟2 (𝑥𝑗 + 𝑖 𝑦𝑗 ). (5.128)

Measuring the direct receptance at the excitation point, we obtain


(𝑟) (𝑟) (𝑟) (𝑟)
𝑤ℓ 𝑤ℓ = 𝜔𝑟2 (𝑥ℓ + 𝑖 𝑦ℓ ). (5.129)
(𝑟)
From equation (5.129) we calculate 𝑤ℓ , then, from equations (5.128) for
(𝑟)
various points 𝑗, we determine 𝑤𝑗 (𝑗 = 1, 2, . . . , 𝑗, … , 𝑁, 𝑗 ≠ ℓ), hence the modal
vector {𝑤 (𝑟) }. The modal matrix [𝑤] is constructed with the vectors {𝑤 (𝑟) } as
columns. If it is square and nonsingular, from (3.193) and (3.194) we obtain
̅𝑟 ˏ] [𝑤]−1,
[𝑀] = [𝑤]−𝑇 [`𝑀

[𝐾 + 𝑖𝐻] = [𝑤]−𝑇 [`𝜆̅𝑟 𝑀


̅𝑟 ˏ] [𝑤]−1.

5.4.1.2. Modal matrix elimination method

An analytical/experimental method has been suggested by Natke [78] as a


generalization of the method presented in § 5.3.4.7 for systems with proportional
hysteretic damping.
The complex transfer receptance (3.199) can be written

179
̅ 𝑗𝑟
( )
𝜘
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 ̅ (5.130)
𝜆𝑟 −𝜔2
where
(𝑟) (𝑟)
(𝑟) 𝑤ℓ 𝑤𝑗
𝜘̅𝑗 = ̅̅̅ . (5.131)
𝑀𝑟
The column vector of complex receptances (for excitation in ℓ ) consists of
𝑁𝑚 elements, corresponding to the 𝑁𝑚 > 𝑁 response measurement points
𝑇 ̅ (𝑟) }
{𝜘
{𝛼̅ℓ } = { 𝛼̅1ℓ 𝛼̅2ℓ . . . 𝛼̅𝑁𝑚 ℓ } = ∑𝑁
𝑟=1 𝜆
̅ 2
, (5.132)
𝑟 −𝜔
where
(𝑟)
(𝑟) 𝑤ℓ
{𝜘̅ }= ̅𝑟
{𝑤 (𝑟) } , 𝜆̅𝑟 = 𝜔𝑟2 (1 + 𝑖 𝑔𝑟 ). (5.133)
𝑀

The complex receptances (5.130) are measured at two sets of excitation


frequencies (𝜔
̅1 , . . . , 𝜔
̅𝑁𝑓 ) and (𝜔
̅𝑁𝑓+1 , . . . , 𝜔
̅2𝑁𝑓 ), where 𝑁𝑓 = 𝑁.
Constructing the following matrices
[𝛼̅ 𝐼 ] = [ {𝛼̅ℓ (𝜔
̅1 )} {𝛼̅ℓ (𝜔
̅2 )} . . . . {𝛼̅ℓ (𝜔
̅𝑁𝑓 )} ] (5.134)

[𝛼̅ 𝐼𝐼 ] = [ {𝛼̅ℓ (𝜔
̅𝑁𝑓+1 )} {𝛼̅ℓ (𝜔
̅𝑁𝑓+2 )} . . . {𝛼̅ℓ (𝜔
̅2𝑁𝑓 )} ] (5.135)

[𝜘̅ ] = [ {𝜘̅ (1) } {𝜘̅ (2) } . . . {𝜘̅ (𝑁𝑓) } ], (5.136)

equation (5.132) yields


[𝛼̅ 𝐼 ] = [𝜘̅ ] [𝛬𝐼 ] , (5.137)
[𝛼̅ 𝐼𝐼 ] = [𝜘̅ ] [𝛬𝐼𝐼 ] , (5.138)
where
1
[𝛬𝐼 ] = [𝛬𝑟𝑠 ], 𝛬𝑟𝑠 = , (𝑟, 𝑠 = 1, … , 𝑁) , (5.139)
𝜆̅ 𝑟 −𝜔̅2
𝑠
1
[𝛬𝐼𝐼 ] = [𝛬𝑟,𝑁𝑓+𝑠 ], 𝛬𝑟,𝑁𝑓+𝑠 = ̅ ̅𝑁2 (𝑟, 𝑠 = 1, … , 𝑁). (5.140)
𝜆𝑟 −𝜔
𝑓 +𝑠
The modal matrix (5.136) is eliminated between (5.137) and (5.138).
Because, generally, [𝛼̅ 𝐼 ] [𝛼̅ 𝐼𝐼 ] and [𝜘̅ ] are not square matrices, the same steps are
followed as earlier in § 5.3.4.7.
Equation (5.137) yields
[𝛼̅ 𝐼 ] [𝛬𝐼 ]−1 = [𝜘̅ ] . (5.141)
Substituting (5.141) into (5.138) we obtain
[𝛼̅ 𝐼𝐼 ] = [𝛼̅ 𝐼 ] [𝛬𝐼 ]−1 [𝛬𝐼𝐼 ] .

180
Premultiplying by the transpose of the matrix [𝛼̅ 𝐼 ]

[𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼𝐼 ] = [𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼 ] [𝛬𝐼 ]−1 [𝛬𝐼𝐼 ] .


−1
then by [[𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼 ]] we obtain
−1
[[𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼 ]] [𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼𝐼 ] = [𝛬𝐼 ]−1 [𝛬𝐼𝐼 ] .

Premultiplying by [𝛬𝐼 ] we obtain

[𝛬𝐼𝐼 ] = [𝛬𝐼 ] [𝑧̅] (5.142)


where
−1
[𝑧̅] = [𝑧̅𝑘𝑠 ] = [[𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼 ]] [𝛼̅ 𝐼 ]𝑇 [𝛼̅ 𝐼𝐼 ] (5.143)
is a matrix with known complex elements 𝑧̅𝑘𝑠 .
Equation (5.142) represents a set of 𝑁 2 equations with 𝑁 unknowns 𝜆̅𝑟
[78].
The element (𝑟, 𝑠) of the matrix [𝛬𝐼𝐼 ] is
1 𝑧̅𝑘𝑠
̅ ̅𝑁2 = ∑𝑁
𝑘=1 ̅ ̅ 𝑘2
. (𝑟, 𝑠 = 1, … , 𝑁) (5.144)
𝜆𝑟 −𝜔 𝜆𝑟 −𝜔

For each index 𝑠 we obtain an equation of degree 𝑁 in 𝜆̅𝑟 with complex


coefficients, which can be written

∏𝑁 ̅ ̅𝑘2 ) − (𝜆̅𝑟 − 𝜔
𝑘=1(𝜆𝑟 − 𝜔
2
̅𝑁+𝑠 ) ∑𝑁 𝑁 ̅ ̅𝜎2 ) 𝑧̅𝑘𝑠 = 0. (5.145)
𝑘=1 ∏ 𝜎=1(𝜆𝑟 − 𝜔
𝜎≠𝑘

The solutions of equations (5.145) are the eigenvalues 𝜆̅𝑟 , which give the
natural frequencies 𝜔𝑟 and the damping factors 𝑔𝑟 (3.198).
The values 𝜆̅𝑟 are generally dependent on 𝑠, which requires a statistical
treatment of data. If a single value 𝑠 is considered, this requires measurements at
𝑁 + 1 excitation frequencies. The modal matrix is then calculated from (5.141).

5.4.2. Systems with Viscous Damping

5.4.2.1. Polar plot method

Denoting
(𝑟) (𝑟) (𝑟)
𝛿𝑗ℓ = 𝑁𝑗ℓ + 𝑖 𝑃𝑗ℓ (5.146)
and using notation (3.163), the transfer receptance (3.186) of a linear system with
nonproportional viscous damping, excited at a single point, can be written under
the form

181
(𝑟) (𝑟) (𝑟) (𝑟)
𝑁𝑗ℓ +𝑖 𝑃𝑗ℓ 𝑁𝑗ℓ −𝑖 𝑃𝑗ℓ
𝛼̅𝑗ℓ = ∑𝑁 (
𝑟=1 𝑖𝜔+𝑛 −𝑖 𝜇 + ) (5.147)
𝑟 𝑟 𝑖𝜔+𝑛𝑟 +𝑖 𝜇𝑟
or
̅𝑟
̅ 𝑟 𝑒 −𝑖𝜑
𝑠𝑖𝑛 𝜑 1 (𝑟) (𝑟) (𝑟)
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 [ 𝜔 (𝑁𝑗ℓ 𝑛𝑟 − 𝑃𝑗ℓ 𝜇𝑟 ) + 𝑖𝑁𝑗ℓ ] (5.148)
𝑛𝑟

where
𝜏 𝛺 𝜔 2𝑛𝑟
𝜑̅𝑟 = 𝑡𝑎𝑛−1 1−𝛺
𝑟 𝑟
2 , 𝛺𝑟 = 𝜔 , 𝜏𝑟 = . (5.149)
𝑟 𝑟 𝜔𝑟

It is assumed that in the neighborhood of the frequency 𝜔 = 𝜔𝑟 all terms


in the summation (5.148) are constant, except for the 𝑟-th term. Let (𝑥0 + 𝑖 𝑦0 ) be
the contribution of nonresonant modes. Denoting

̅ 𝑟 𝑒−𝑖𝜑
𝑠𝑖𝑛 𝜑 ̅𝑟
1 (𝑟) (𝑟) ( 𝑟)
𝑥 (𝑟) + 𝑖𝑦(𝑟) = [𝜔 (𝑁𝑗ℓ 𝑛𝑟 − 𝑃𝑗ℓ 𝜇𝑟 ) + 𝑖𝑁𝑗ℓ ] (5.150)
𝑛𝑟

we obtain
𝛼̅𝑗ℓ = 𝑥(𝑟) + 𝑖 𝑦 (𝑟) + 𝑥0 + 𝑖 𝑦0 . (5.151)
At the frequency 𝜔 = 𝜔𝑟
(𝑟) (𝑟) (𝑟)
𝑁𝑗ℓ 𝑁𝑗ℓ 𝑛𝑟 −𝑃𝑗ℓ 𝜇𝑟
(𝑥 (𝑟)
+𝑖𝑦 (𝑟)
)𝜔=𝜔 = 𝑛 −𝑖 𝜔𝑟 𝑛𝑟
(5.152)
𝑟 𝑟
hence

[ℛ𝑒(𝛼̅𝑗ℓ ) − 𝑥0 ] + 𝑖 [ℐ𝑚(𝛼̅𝑗ℓ ) − 𝑦0 ] = 𝑥 (𝑟) + 𝑖 𝑦 (𝑟) =


1
̅ 𝑟 𝑒−𝑖𝜑̅𝑟 [ (𝑦 (𝑟) )
= −𝑠𝑖𝑛 𝜑 − 𝑖 (𝑥 (𝑟) )𝜔=𝜔 ]. (5.153)
𝛺 𝜔=𝜔 𝑟 𝑟 𝑟

Denoting

tan 𝜑
̅𝑟 = 𝑡𝑟 , (5.154)
expression (5.153) becomes
𝑡𝑟 1
𝑥 (𝑟) + 𝑖 𝑦 (𝑟) = − (1 − 𝑖 𝑡𝑟 ) [ (𝑦 (𝑟) )𝜔=𝜔 − 𝑖 (𝑥 (𝑟) )𝜔=𝜔 ]
1+𝑡𝑟2 𝛺 𝑟 𝑟 𝑟

(5.155)
hence
𝑡 1
𝑥 (𝑟) = 1+𝑡𝑟 2 [ 𝑡𝑟 (𝑥 (𝑟) )𝜔=𝜔 − 𝛺 (𝑦 (𝑟) )𝜔=𝜔 ] , (5.156)
𝑟 𝑟 𝑟 𝑟

𝑡 𝑡
𝑦 (𝑟) = 1+𝑡𝑟 2 [(𝑥 (𝑟) )𝜔=𝜔 + 𝛺𝑟 (𝑦 (𝑟) )𝜔=𝜔 ]. (5.157)
𝑟 𝑟 𝑟 𝑟

182
If expression (5.153) is graphically
represented in the Argand plane, we obtain the polar
plot from Fig.5.36 which is very close to a circle.
The identification method suggested by
Woodcock [137] uses the frequencies where the vector
components ℛ𝑒(𝛼̅𝑗ℓ ) and ℐ𝑚(𝛼̅𝑗ℓ ) have maximum or
minimum values. They are equal to the frequencies
where the vector components 𝑥 (𝑟) ≡ ℛ𝑒(𝛼̅𝑗ℓ ) − 𝑥0
and 𝑦 (𝑟) ≡ ℐ𝑚(𝛼̅𝑗ℓ ) − 𝑦0 have extreme values.
𝑑𝑥 (𝑟)
The condition = 0 yields
𝑑𝛺𝑟
2
𝑥(𝑟) 𝑥(𝑟)
1∓𝜏𝑟 √ 1+( (𝑟) ) +𝜏𝑟 ( (𝑟) )
𝑦 𝜔=𝜔𝑟 𝑦 𝜔=𝜔𝑟
𝛺𝑟2𝐼,𝐼𝐼𝐼 = 𝑥(𝑟)
(5.158)
1+𝜏𝑟 ( (𝑟) )
𝑦 𝜔=𝜔𝑟
in which
𝜔 𝜔𝐼𝐼𝐼
𝛺𝑟𝐼 = 𝜔 𝐼 , 𝛺𝑟𝐼𝐼𝐼 = , (5.159)
𝑟 𝜔𝑟

where 𝜔𝐼 and 𝜔𝐼𝐼𝐼 are the frequencies at the minimum and maximum of the in-
phase response.
𝑑𝑦 (𝑟)
The condition 𝑑𝛺𝑟
= 0 yields

𝑥 (𝑟) 𝑡𝑟2 [𝑡𝑟22 −1−𝛺𝑟2𝐼𝐼 (𝑡𝑟22 +3)]


(𝑦(𝑟)) = (5.160)
𝜔=𝜔𝑟 𝛺𝑟𝐼𝐼 (1+𝛺𝑟2𝐼𝐼 )(1−𝑡𝑟22 )

in which
𝜏𝑟 𝛺𝑟 𝜔𝐼𝐼
𝑡𝑟2 = 1−𝛺2𝐼𝐼 , 𝛺𝑟𝐼𝐼 = , (5.161)
𝑟𝐼𝐼 𝜔𝑟

where 𝜔𝐼𝐼 is the frequency of maximum quadrature response.


A graph (Fig.5.37) can be set up from equations (5.158) and (5.160)
analogous to Fig.5.35 drawn for systems with hysteretic damping.
Using various values 𝜏𝑟 and 𝛺𝑟𝐼𝐼 , equation (5.160) is used to calculate
𝑥 (𝑟) 𝛺𝑟𝐼𝐼𝐼
( ) , then equation (5.158) yields 𝛺𝑟𝐼 and 𝛺𝑟𝐼𝐼𝐼 . For each pair of values
𝑦 (𝑟) 𝜔=𝜔 𝛺𝑟𝐼𝐼
𝑟
𝛺𝑟𝐼
and 𝛺 a point is plotted in Fig.5.37, then continuous lines are drawn through the
𝑟𝐼𝐼
points with 𝜏𝑟 = 𝑐𝑜𝑛𝑠𝑡. and 𝛺𝑟𝐼𝐼 = 𝑐𝑜𝑛𝑠𝑡., respectively.

183
Using the graph from Fig.5.37, the values 𝜏𝑟 and 𝜔𝑟 can be determined
based on experimentally determined frequencies 𝜔𝐼 , 𝜔𝐼𝐼 and 𝜔𝐼𝐼𝐼 . Then, from
equations (3.164) and (3.165) we obtain 𝑛𝑟 and 𝜇𝑟 , hence 𝜎𝑟 .

Measuring the curvature radius 𝜌𝜔𝑟 at the point of frequency 𝜔𝑟 from


Fig.5.36, we can calculate (𝑥 (𝑟) + 𝑖𝑦(𝑟) )𝜔=𝜔 .
𝑟
Indeed, we can establish the equation
2 3
2
𝑑𝑦(𝑟) 𝜏 𝑥(𝑟)
1+( (𝑟) ) 2 {1+[ 𝑟 +( (𝑟) ) ] }
𝑑𝑥 𝜔=𝜔𝑟
(𝑦 (𝑟) )𝜔=𝜔 2 𝑦 𝜔=𝜔𝑟
2 𝑟
𝜌𝜔 𝑟
= 2 = 2 2
(5.162)
2
𝑑 𝑦(𝑟) 4 𝜏2
( (𝑟)2 ) 𝜏 𝑥(𝑟)
𝑑𝑥 𝜔=𝜔𝑟
{1− 𝑟 +[ 𝑟 +( (𝑟) ) ] }
8 2 𝑦 𝜔=𝜔𝑟

𝑥 (𝑟)
wherefrom we obtain (𝑦 (𝑟) )𝜔=𝜔 , then (𝑥 (𝑟) )𝜔=𝜔 , because (𝑦 (𝑟)) is known.
𝑟 𝑟 𝜔=𝜔𝑟
Equation (5.152) yields
(𝑟) (𝑟) 𝜔𝑟 𝑛𝑟
𝑁𝑗ℓ = 𝑛𝑟 (𝑥 (𝑟) )𝜔=𝜔 , 𝑃𝑗ℓ = 𝜇𝑟
(𝑦 (𝑟) )𝜔=𝜔 + 𝑛𝑟2 (𝑥 (𝑟) )𝜔=𝜔 (5.163)
𝑟 𝑟 𝑟

hence
(𝑟) (𝑟) (𝑟)
𝛿𝑗ℓ = 𝑁𝑗ℓ + 𝑖 𝑃𝑗ℓ .

184
Repeating the measurements at various points 𝑗 of the structure, from the
graphical analysis of the corresponding polar plots we can calculate the vector
𝑇
((𝑟)) ((𝑟))
{𝛿 (𝑟) } = { 𝛿1ℓ . . . .. 𝛿𝑁ℓ } which, in view of equation (3.185), is
proportional to the 𝑟-th modal vector {𝑞(𝑟) }.
Equations (3.150) and (3.160) yield
𝑇
{𝑞 (𝑟) } ( [𝐶] + 2 𝜎𝑟 [𝑀]) {𝑞(𝑟) } = 𝑢𝑟 ,
𝑇
{𝑞 (𝑟) } ( [𝐾] − 𝜎𝑟2 [𝑀]) {𝑞 (𝑟) } = 𝑣𝑟 .
Substituting the above expressions into equation (3.161) we obtain

([𝑀]𝜎𝑟2 + [𝐶]𝜎𝑟 + [𝐾]) {𝑞 (𝑟) } = {0}. (5.164)

Combining equations (5.164) for all values 𝑟, we obtain the matrix


equations
[𝑀][𝑞][`𝛴ˏ]2 + [𝐶] [𝑞] [`𝛴ˏ] + [𝐾] [𝑞] = [0],
(5.165)
[𝑀][𝑞 ∗ ][`𝛴∗ ˏ]2 + [𝐶] [𝑞 ∗ ] [`𝛴 ∗ ˏ] + [𝐾] [𝑞 ∗ ] = [0],
where
[𝑞] = [ {𝑞(1) } {𝑞(2) } . . . . {𝑞 (𝑁) } ], [`𝛴ˏ] = 𝑑𝑖𝑎𝑔[𝜎𝑟 ].
The superscript ∗ denotes the complex conjugate. The values 𝜎𝑟 are
arranged such as each complex conjugate pair contains one value from the first 𝑁
eigenvalues and one value from the next 𝑁 eigenvalues.
Because the matrix [𝑞] is non-singular, we can denote
[𝑞] [`𝛴ˏ] [𝑞]−1 = [𝑋] + 𝑖 [𝑌] ,
(5.166)
[𝑞 ∗ ] [`𝛴 ∗ ˏ] [𝑞 ∗ ]−1 = [𝑋] − 𝑖 [𝑌] ,
where [𝑋] and [𝑌] are real matrices.
Because
[𝑞] [`𝛴ˏ] 2 [𝑞]−1 = ([𝑋] + 𝑖 [𝑌])2 ,
[𝑞 ∗ ] [`𝛴 ∗ ˏ]2 [𝑞∗ ]−1 = ([𝑋] − 𝑖 [𝑌])2 ,
equations (5.165) yield
[𝐶] = −[𝑀] ([𝑋] + [𝑌] [𝑋] [𝑌]−1 ) ,
(5.167)
[𝐾] = [𝑀] ([𝑌]2 + [𝑌] [𝑋] [𝑌]−1 [𝑋]),
hence the matrices [𝐶] and [𝐾] can be determined if the matrix [𝑀], the eigenvalues
𝜎𝑟 and the eigenvectors {𝑞(𝑟) } are known.

185
To determine the eigenvectors, the polar plots have to be measured in at
least 𝑁 points of the structure.
The method can be applied for the identification of systems with widely
spaced natural frequencies regardless the damping level.
From expression (5.158) we can obtain exact formulae for 𝜏𝑟 which is a
measure of damping
2 2
𝜔𝐼𝐼𝐼 −𝜔𝐼2 𝑥 (𝑟) 𝑥 (𝑟)
2 +𝜔 2 = 𝜏𝑟 √ 1 + (𝑦 (𝑟) ) − 𝜏𝑟 (𝑦 (𝑟)) , (5.168)
𝜔𝐼𝐼𝐼 𝐼 𝜔=𝜔𝑟 𝜔=𝜔𝑟
and
2
𝜔𝐼𝐼𝐼 +𝜔𝐼2 1
= 𝑥(𝑟)
. (5.169)
2 𝜔𝑟2 1−𝜏𝑟 ( (𝑟) )
𝑦 𝜔=𝜔𝑟

From expression (5.156), differentiating with respect to ω, we obtain

2 (𝑦 (𝑟) )𝜔=𝜔
𝑟
𝜏𝑟 = . (5.170)
𝜔𝑟 (−𝑑𝑥(𝑟) )
𝑑𝜔 𝜔=𝜔𝑟

For systems with well-spaced natural frequencies, the contribution of non-


resonant modes to the total response in the vicinity of a resonance is sometimes
neglected, considering
𝑥 (𝑟) ≡ ℛ𝑒(𝛼̅𝑗ℓ ), 𝑦 (𝑟) ≡ ℐ𝑚(𝛼̅𝑗ℓ ). (5.171)
(𝑟)
𝑥
For 𝜏𝑟 << ( ) from equation (5.168) Zimmerman [141] obtained
𝑦(𝑟) 𝜔=𝜔
𝑟

2 2
𝜔𝐼𝐼𝐼 −𝜔𝐼2 ̅ )
ℛ𝑒(𝛼
2 +𝜔 2 = 𝜏𝑟 √1 + (ℐ𝑚(𝛼̅𝑗ℓ )) , (5.172)
𝜔𝐼𝐼𝐼 𝐼 𝑗ℓ 𝜔=𝜔𝑟

which, for proportional damping (when ℛ𝑒(𝛼̅𝑗ℓ )𝜔=𝜔 = 0) is identical to (2.76).


𝑟
Using notations (5.171), formula (5.170) becomes [141]
2 ̅ 𝑗ℓ )]
[ℐ𝑚(𝛼
𝜔=𝜔𝑟
𝜏𝑟 = 𝜔 𝑑[ℛ𝑒(𝛼
̅ 𝑗ℓ )]
(5.173)
𝑟
(− 𝑑𝜔 )
𝜔=𝜔𝑟

which is identical to expression (2.79) derived for systems with proportional


damping.
Gaukroger, Skingle and Heron [42] suggested an iterative procedure for the
analysis of polar plots, based on the minimization of the mean square error of
response.

186
5.4.2.2. Modal matrix elimination method

For structures with nonproportional damping and close natural frequencies,


the contribution of the non-resonant modes to the response in the vicinity of a
resonance is no more constant and the above presented method yields inaccurate
results. The difficulty may be overcome using the modal matrix elimination
method.
From equations (5.148), (5.150) and (5.155) we obtain the following
expression of the complex transfer receptance
𝑡𝑟 (𝑖 𝑡𝑟 −1) 1
𝛼̅𝑗ℓ = ∑𝑁
𝑟=1 [ 𝛺 (𝑦 (𝑟) )𝜔=𝜔 − 𝑖 (𝑥 (𝑟) )𝜔=𝜔 ]. (5.174)
1+𝑡𝑟2 𝑟 𝑟 𝑟

The vector components can be written under the form

ℛ𝑒(𝛼̅𝑗ℓ ) = 𝛼𝑗ℓ𝑅 = ∑𝑁
𝑟=1 [𝑎
̿ 𝑟 (𝑦(𝑟) )
̿𝑟 (𝑥(𝑟) )𝜔=𝜔 + 𝐵 ], (5.175)
𝑟 𝜔=𝜔 𝑟

ℐ𝑚(𝛼̅𝑗ℓ ) = 𝛼𝑗ℓ𝐼 = ∑𝑁 ̿ (𝑟) ̿ (𝑟)


𝑟=1 [𝑏𝑟 (𝑥 )𝜔=𝜔 + 𝐴𝑟 (𝑦 )𝜔=𝜔 ], (5.176)
𝑟 𝑟

where
𝑡2𝑟 1 𝑡2𝑟
𝑎̿𝑟 = 2 , 𝐴̿𝑟 = 𝑎̿𝑟 = ,
1+𝑡𝑟 𝛺𝑟 𝛺𝑟 (1+𝑡𝑟2 )
(5.177)
𝑡𝑟 1 𝑡𝑟
𝑏̿𝑟 = , 𝐵̿𝑟 = 𝑏 ̿ = .
1+𝑡𝑟2 𝛺𝑟 𝑟 𝛺𝑟 (1+𝑡𝑟2 )
Evaluating expressions (5.175) and (5.176) at 𝑓 = 𝑁 excitation
frequencies, we obtain
̿ (𝜔
𝑟=1 ({𝑎 ̅𝑓 )} (𝑥(𝑟) )𝜔 + {𝐵 (𝑟)
̅𝑓 )} = ∑𝑁
{𝛼𝑗ℓ𝑅 (𝜔 ̿ 𝑟 (𝜔 𝑟 ̅𝑓 )} (𝑦 )𝜔𝑟 ) =
𝑟
= [𝑎
(𝑟)
̿ 𝑓𝑟 ] {𝑥𝜔 } + [𝐵̿ ] {𝑦 (𝑟) } (5.178)
𝑟 𝑓𝑟 𝜔𝑟

̿ ̅ )} (𝑥(𝑟) ) + {𝐴̿ (𝜔
𝑟=1 ({𝑏𝑟 (𝜔
(𝑟)
̅𝑓 )} = ∑𝑁
{𝛼𝑗ℓ𝐼 (𝜔 𝑓 𝜔𝑟 𝑟 ̅𝑓 )} (𝑦 )𝜔𝑟 ) =
= [𝑏̿ ] {𝑥 (𝑟) } + [𝐴
̿ ] {𝑦 (𝑟) } (5.179)
𝑓𝑟 𝜔𝑟 𝑓𝑟 𝜔𝑟
or
[𝑎 ̿ ] {𝑥 (𝑟) }
̿ 𝑓𝑟 ] [𝐵
̅𝑓 )}
{𝛼𝑗ℓ𝑅 (𝜔 𝑓𝑟 𝜔𝑟
{ }=[ ] { (𝑟) }. (5.180)
̅𝑓 )}
{𝛼𝑗ℓ𝐼 (𝜔 ̿ ] [𝐴
[𝑏 ̿ ] {𝑦 }
𝑓𝑟 𝑓𝑟 𝜔𝑟

In (5.180) the square matrices have the form


[𝑎̿𝑓𝑟 ] = [{𝑎̿1 (𝜔
̅𝑓 )} {𝑎̿2 (𝜔
̅𝑓 )} . . . . {𝑎̿𝑁 (𝜔
̅𝑓 )} ] (5.181)
and the column vectors have the form
𝑇
̅𝑓 )} = {𝛼𝑗ℓ𝑅 (𝜔
{𝛼𝑗ℓ𝑅 (𝜔 ̅1 ) 𝛼𝑗ℓ𝑅 (𝜔
̅2 ) . . . . . 𝛼𝑗ℓ𝑅 (𝜔
̅𝑁 ) } (5.182)

187
(𝑟) (1) (2) (𝑁) 𝑇
{𝑥𝜔𝑟 } = {𝑥𝜔1 𝑥𝜔2 . . . . . 𝑥𝜔𝑁 } . (5.183)

Evaluating expressions (5.175) and (5.176) at another set of 𝑁 excitation


frequencies (𝑓 = 𝑁 + 1, . . . . , 2𝑁) we obtain

[𝑎
̿ 𝑁+𝑓,𝑟 ] [𝐵̿ (𝑟)
̅𝑁+𝑓 )}
{𝛼𝑗ℓ𝑅 (𝜔 𝑁+𝑓,𝑟 ] {𝑥𝜔𝑟 }
{ }=[ ] { (𝑟) }. (5.184)
̅𝑁+𝑓 )}
{𝛼𝑗ℓ𝐼 (𝜔 ̿
[𝑏 ] [ ̿
𝐴 ] {𝑦𝜔𝑟 }
𝑁+𝑓,𝑟 𝑁+𝑓,𝑟
Eliminating the modal matrix between equations (5.180) and (5.184) we
obtain
−1
[𝑎 ̿ ]
̿ 𝑓𝑟 ] [𝐵 [𝑎̿ 𝑁+𝑓,𝑟 ] [𝐵̿
̅𝑓 )}
{𝛼𝑗ℓ𝑅 (𝜔 𝑓𝑟 𝑁+𝑓,𝑟 ] ̅𝑁+𝑓 )}
{𝛼𝑗ℓ𝑅 (𝜔
{ }=[ ]·[ ] ·{ }
̅𝑓 )}
{𝛼𝑗ℓ𝐼 (𝜔 ̿ ] [𝐴
[𝑏 ̿ ] [ ̿
𝑏 ] [ ̿
𝐴 ] ̅𝑁+𝑓 )}
{𝛼𝑗ℓ𝐼 (𝜔
𝑓𝑟 𝑓𝑟 𝑁+𝑓,𝑟 𝑁+𝑓,𝑟
(5.185)
where the column vectors contain experimentally measured elements.
The matrix equation (5.185) contains 2𝑁 nonlinear algebraic equations
with 2𝑁 unknowns 𝜏𝑟 and 𝜔𝑟 (𝑟 = 1, . . . , 𝑁) which can be solved by a Newton-
Raphson procedure. Calculations are simplified taking into account the relationship
between the elements of the square matrices from (5.185) obtained directly from
notations (5.177).
With known 𝜏𝑟 and 𝜔𝑟 , the elements of the modal vectors are obtained
from (5.180). The ratios of the receptances measured at various coordinates 𝑞𝑗 , due
to excitation at a given coordinate 𝑞ℓ , will give the shape of the 𝑟-th mode.

5.5. IDENTIFICATION OF HEAVILY DAMPED SYSTEMS

For heavily damped systems or subsystems, for which the modes of


vibration cannot be distinguished, the identification is used to determine the
coefficients of the transfer function, expressed by a ratio of two frequency-
dependent polynomials.
The differential equation which describes the behavior of a linear system
can be written under the form
(𝑄𝑛 𝑠 𝑛 + 𝑄𝑛−1 𝑠 𝑛−1 + . . . +𝑄1 𝑠 + 1) 𝑦 =
= (𝑃𝑚 𝑠 𝑚 + 𝑃𝑚−1 𝑠 𝑚−1 + . . . +𝑃1 𝑠 + 𝑃0 ) 𝑥,
where 𝑥 and 𝑦 are the input and output, respectively, 𝑄𝑟 and 𝑃𝑟 are constant (time
𝑑
invariant) coefficients, 𝑠 is the operator 𝑑𝑡, and 𝑛 > 𝑚.
The system transfer function (transmittance) can be written
𝑃0 +𝑃1 𝑠+ ...+𝑃𝑚−1 𝑠𝑚−1 +𝑃𝑚 𝑠𝑚 ∑𝑚 𝑃𝑟 𝑠𝑟
𝐻(𝑠) =
1+𝑄1 𝑠+ ...+𝑄𝑛−1 𝑠𝑛−1 +𝑄𝑛 𝑠𝑛
= 1+𝑟=0
∑𝑛 𝑟. (5.186)
𝑟=1 𝑄𝑟 𝑠

188
The isochronous frequency response function, corresponding to the
permanent harmonic regime, is obtained from (5.186) by substituting 𝑠 = 𝑖𝜔:
𝑃0 +𝑃1 𝑖𝜔+𝑃2 (𝑖𝜔)2 + ....+𝑃𝑚 (𝑖𝜔)𝑚
𝐻(𝑖𝜔) = . (5.187)
1+𝑄1 𝑖𝜔+𝑄2 (𝑖𝜔)2 + ....+𝑄𝑛 (𝑖𝜔)𝑛

5.5.1. Direct Identification Methods

5.5.1.1. Analytical methods

Separating the real and imaginary parts of the frequency response function
(5.187)
(𝑃0 −𝑃2 𝜔2 +𝑃4 𝜔4 ....) +𝑖 (𝑃1 𝜔−𝑃3 𝜔3 +𝑃5 𝜔5 . . .)
𝐻(𝑖𝜔) = =
(1−𝑄2 𝜔2 +𝑄4 𝜔4 ....)+𝑖 (𝑄1 𝜔−𝑄3 𝜔3 +𝑄5 𝜔5 . . .)
𝑃𝑅 (𝜔)+𝑖𝑃𝐼 (𝜔)
= = ℛ𝑒(𝜔) + 𝑖 ℐ𝑚(𝜔) (5.188)
𝑄𝑅 (𝜔)+𝑖𝑄𝐼 (𝜔)
the following general equations are obtained
ℛ𝑒(𝜔)𝑄𝑅 (𝜔) − ℐ𝑚(𝜔) 𝑄𝐼 (𝜔) = 𝑃𝑅 (𝜔),
(5.189)
ℛ𝑒(𝜔)𝑄𝐼 (𝜔) + ℐ𝑚(𝜔) 𝑄𝑅 (𝜔) = 𝑃𝐼 (𝜔) .
At a given frequency 𝜔𝑓 , equations (5.189) can be written
ℛ𝑒(𝜔𝑓 ) (1 − 𝑄2 𝜔𝑓2 + 𝑄4 𝜔𝑓4 . . . ) − ℐ𝑚(𝜔𝑓 ) (𝑄1 𝜔𝑓 − 𝑄3 𝜔𝑓3 + 𝑄5 𝜔𝑓5 . . . ) =
= (𝑃0 − 𝑃2 𝜔𝑓2 + 𝑃4 𝜔𝑓4 . . . ) (5.190)
ℛ𝑒(𝜔𝑓 ) (𝑄1 𝜔𝑓 − 𝑄3 𝜔𝑓3 + 𝑄5 𝜔𝑓5 . . . ) + ℐ𝑚(𝜔𝑓 )(1 − 𝑄2 𝜔𝑓2 + 𝑄4 𝜔𝑓4 . . . ) =
= (𝑃1 𝜔𝑓 − 𝑃3 𝜔𝑓3 + 𝑃5 𝜔𝑓5 . . . ).
Equations (5.190) have 𝑛 unknowns 𝑄𝑟 and (𝑚 + 1) unknowns 𝑃𝑟 , a total
of (𝑛 + 𝑚 + 1) unknowns. Hence (𝑛 + 𝑚 + 1) equations are required for the
calculation of the coefficients 𝑄𝑟 and 𝑃𝑟 . They are obtained evaluating relations
𝑛+𝑚+1
(5.190) at 2
distinct frequencies 𝜔𝑓 . For this, it is necessary either to plot the
diagrams of the vector components of the frequency response function 𝐻(𝑖𝜔), or
𝑛+𝑚+1
to measure pairs of values of the components ℛ𝑒(𝜔) and ℐ𝑚(𝜔) at 2
excitation frequencies [112].
If only the diagram of the absolute value of the function 𝐻(𝑖𝜔) is available,
it is possible to determine the coefficients of the function
𝑃2𝑅 (𝜔)+𝑃2𝐼 (𝜔) 𝑝0 +𝑝1 𝜔2 + ...+𝑝𝑚 𝜔2𝑚
|𝐻(𝑖𝜔)|2 = 𝐻(𝑖𝜔) · 𝐻(−𝑖𝜔) = = ,
𝑄2𝑅 (𝜔)+𝑄2𝐼 (𝜔) 1+𝑞1 𝜔2 + ...+𝑞𝑛 𝜔2𝑛

189
evaluated at (𝑛 + 𝑚 + 1) frequencies 𝜔.
In both cases, before starting the computations, it is necessary to estimate
the degrees of the denominator and numerator, i.e. to prestructure the model.

5.5.1.2. Graphical methods

Identification in the frequency domain

A frequency domain identification method, not requiring a beforehand


estimation of the structure of the analyzed system, has been suggested by Dudnikov
[31].
Dividing the denominator and numerator of expression (5.187) by (𝑖𝜔)𝑛 ,
we obtain
−𝑛 1−𝑛 𝑚−1−𝑛 𝑚−𝑛
𝑃0 (𝑖𝜔) +𝑃1 (𝑖𝜔) + ...+𝑃𝑚−1 (𝑖𝜔) + 𝑃𝑚(𝑖𝜔)
𝐻(𝑖𝜔) = −𝑛 1−𝑛 −1 ,
(𝑖𝜔) +𝑄1 (𝑖𝜔) + ...+𝑄𝑛−1 (𝑖𝜔) +𝑄𝑛
(5.191)
expression which is denoted 𝐻0 (𝑖𝜔) and is written under the form
𝐻0 (𝑖𝜔) = 𝐴0 + 𝐺1 (𝑖𝜔) (5.192)
where 𝐴0 is a constant.
Consider then the function
1
𝐻1 (𝑖𝜔) = 𝐺 (𝑖𝜔)
(5.193)
1

which is written under the form


𝐻1 (𝑖𝜔) = 𝐴1 + 𝐵0 (𝑖𝜔)−1 + 𝐺2 (𝑖𝜔) (5.194)
where 𝐴1 is a constant.
The function 𝐺2 (𝑖𝜔) has a numerator of degree (2 − 𝑛) and a denominator
of degree (1 − 𝑛). Its inverse, 𝐻2 (𝑖𝜔) can be written
1
𝐻2 (𝑖𝜔) = = 𝐴2 + 𝐵1 (𝑖𝜔)−1 + 𝐺3 (𝑖𝜔) (5.195)
𝐺2 (𝑖𝜔)

and so on.
These operations lead to the representation of the initial transmittance
(5.191) as a continuous fraction
1
𝐻0 (𝑖𝜔) = 𝐴0 + 1 . (5.196)
𝐵0 (𝑖𝜔)−1 +𝐴 1 +𝐵1(𝑖𝜔)−1 +𝐴2 + .
.
.
1
+
𝐵𝑛−1 (𝑖𝜔)−1 +𝐴𝑛

190
The constant parameters 𝐴0 , 𝐵0 , 𝐴1 , 𝐵1 , . . . , 𝐵𝑛−1, 𝐴𝑛 are determined as
follows:
First, the functions 𝐻𝑟 (𝑖𝜔) are separated into the real and imaginary parts
𝐻0 (𝑖𝜔) = 𝐻0𝑅 (𝜔) + 𝑖 𝐻0𝐼 (𝜔),
𝐻1 (𝑖𝜔) = 𝐻1𝑅 (𝜔) + 𝑖 𝐻1𝐼 (𝜔),
. . . . . . . . . . . . .
𝐻𝑛 (𝑖𝜔) = 𝐻𝑛𝑅 (𝜔) + 𝑖 𝐻𝑛𝐼 (𝜔).
Then, we let 𝜔 to tend to zero. In this case |𝐻0 (𝑖𝜔)| tends asymptotically
to |𝐻0𝑅 (𝜔)|which, in the virtue of equation (5.192), tends to |𝐴0 | = |𝑃0 |;
|𝐻1 (𝑖𝜔)| tends asymptotically to |𝐻1𝐼 (𝜔)|which, in the virtue of equation
(5.194), tends to |𝐵0 (𝑖𝜔)−1 |;
|𝐻1 (𝑖𝜔) − 𝐵0 (𝑖𝜔)−1 | tends asymptotically to |𝐻1𝑅 (𝜔)|, hence to |𝐴1 |;
|𝐻2 (𝑖𝜔)| tends to |𝐻2𝐼 (𝜔)|, hence to |𝐵1 (𝑖𝜔)−1 |;
|𝐻2 (𝑖𝜔) − 𝐵1 (𝑖𝜔)−1 | tends to |𝐻2𝑅 (𝜔)|, hence to |𝐴2 |;
. . . . . . . . . . . . .
|𝐻𝑛 (𝑖𝜔)| tends to |𝐻𝑛𝐼 (𝜔)|, hence to |𝐵𝑛−1 (𝑖𝜔)−1 |;
|𝐻𝑛 (𝑖𝜔) − 𝐵𝑛−1 (𝑖𝜔)−1 | tends to |𝐻𝑛𝑅 (𝜔)|, hence to |𝐴𝑛 |.
The procedure starts with the determination of the constant 𝐴0 , whose
absolute value is equal to the ordinate at 𝜔 → 0 of the asymptote to the curve
|𝐻0𝑅 (𝜔)| (Fig.5.38). It is used to calculate
1 1
𝐻1 (𝑖𝜔) = 𝐺 (𝑖𝜔) = 𝐻 (𝑖𝜔)−𝐴 .
1 0 0

The function |𝐻1𝐼 (𝜔)| is represented graphically using logarithmic scales


and the asymptote to this curve for 𝜔 → 0 is drawn as in Fig.5.39. The asymptote
has a slope (−1). The ordinate of the point of abscissa 𝜔 = 1 is 𝑙𝑜𝑔|𝐵0 |.

191
Next |𝐴1 | is determined from the graph of the function |𝐻1𝑅 (𝜔)|. The
constant 𝐴1 is used to calculate the function 𝐺2 (𝑖𝜔) which is used to sketch the
curves of the absolute values of the imaginary and real parts, 𝐻2𝐼 (𝜔) and 𝐻2𝑅 (𝜔),
of its inverse 𝐻2 (𝑖𝜔). These graphs are used to determine the values of |𝐵1 | and
|𝐴2 |, and so on.
𝐵0 𝐵
, 𝐴1 , . . . , 𝑖 𝑛−1
The signs of 𝐴0 , 𝑖 , 𝐴𝑛 are determined from the
𝜔 𝜔
signs of the imaginary and real parts of the functions used for their
calculation.
The computation ends when the function 𝐺𝑛+1 (𝑖𝜔) is zero. Once all
coefficients are known, the function 𝐻0 (𝑖𝜔) is written under the form (5.196). The
final analytical expression of the transmittance is obtained after reduction to
common denominator.

Identification in the time domain

An often used time domain identification technique is the multiple


integration method [110, 114].
Based on expression (5.186) the following relations can be established
𝐽0 = lim 𝐻(𝑠) = 𝑃0 ,
𝑠→0
(1) 1
𝐽1 = lim 𝐻 (𝑠) = lim [ 𝐽0 − 𝐻(𝑠)] = −𝑃1 + 𝐽0 𝑄1,
𝑠→0 𝑠→0 𝑠
1
𝐽2 = lim 𝐻 (2) (𝑠) = lim 𝑠 [ 𝐽1 − 𝐻 (1) (𝑠)] = 𝑃2 + 𝐽1 𝑄1 − 𝐽0 𝑄2 ,
𝑠→0 𝑠→0
. . . . . . . . . . . . .
1
𝐽𝑟 = lim 𝐻 (𝑟) (𝑠) = lim [ 𝐽𝑟−1 − 𝐻 (𝑟−1) (𝑠)] =
𝑠→0 𝑠 𝑠→0
= (−1)𝑟 𝑃1 + 𝐽𝑟−1 𝑄1 − 𝐽𝑟−2 𝑄2 + . . . +(−1)𝑟−1 𝐽0 𝑄𝑟 ,
where 𝑃𝑟 = 0 for 𝑟 > 𝑚, and 𝑃𝑟 = 𝑄𝑟 = 0 for 𝑟 > 𝑛.
When 𝑟 = 𝑚 + 𝑛, there exist (𝑛 + 𝑚 + 1) equations for the calculation of
the (𝑛 + 𝑚 + 1) unknowns 𝑄1 , . . . , 𝑄𝑛 , 𝑃0 , . . . . , 𝑃𝑚 .
The coefficients 𝐽0 , . . . , 𝐽𝑟 can be determined by the repeated integration
of a time response function of the system, e.g. the weighting function ℎ(𝑡).
Because
1 𝑡
lim 𝐻(𝑠) = lim ℎ(𝑡), lim 𝑠 𝐻(𝑠) = lim ∫0 ℎ(𝜏) 𝑑𝜏,
𝑠→0 𝑡→∞ 𝑠→0 𝑡→∞

it follows that
𝐽0 = lim 𝐻(𝑠) = lim ℎ(𝑡),
𝑠→0 𝑡→∞
1 𝑡
𝐽1 = lim 𝑠 [ 𝐽0 − 𝐻(𝑠)] = lim ∫0 [ 𝐽0 − ℎ(𝜏)] 𝑑𝜏 .
𝑠→0 𝑡→∞

192
The coefficient 𝐽1 is equal to the hatched area from Fig.5.40.

Constructing the function


𝑡
ℎ1 (𝑡) = ∫0 [ 𝐽0 − ℎ(𝜏)] 𝑑𝜏,
it follows that
1 𝑡
𝐽2 = lim 𝑠 [ 𝐽1 − 𝐻 (1) (𝑠)] = lim ∫0 [ 𝐽1 − ℎ1 (𝜏)] 𝑑𝜏 ,
𝑠→0 𝑡→∞

therefore 𝐽2 is equal to the hatched area from Fig.5.41.

Generally, the coefficients 𝐽𝑟 are equal to the areas enclosed between the
curves
𝑡
𝑔(𝑡) = ℎ𝑟−1 (𝑡) = ∫0 [ 𝐽𝑟−2 − ℎ𝑟−2 (𝜏)] 𝑑𝜏
and the horizontal lines
𝑔(𝑡) = 𝐽𝑟−1,
in view of equation
1 𝑡
𝐽𝑟 = lim 𝑠 [ 𝐽𝑟−1 − 𝐻 (𝑟−1) (𝑠)] = lim ∫0 [ 𝐽𝑟−1 − ℎ𝑟−1 (𝜏)] 𝑑𝜏 .
𝑠→0 𝑡→∞

The method yields good results for systems with 𝑛 ≤ 3.

Another time domain identification technique is the method of moments


[4].
The function (5.186) can be developed in a series of the form
𝑠2 𝑠3
𝐻(𝑠) = ℎ∞ − 𝑠 𝑀0 + 1! 𝑀1 − 2! 𝑀2 + . . . (5.197)
where

𝑀𝜈 = ∫0 𝑡 𝜈 [ℎ∞ − ℎ(𝑡)] 𝑑𝑡
is the 𝜈-th order moment of the function [ℎ∞ − ℎ(𝑡)].

193
Substituting (5.197) into (5.186) we obtain
𝑀1 𝑀2
(ℎ∞ − 𝑀0 𝑠 + 1!
𝑠2 − 2!
𝑠 3 + . . ) (1 + 𝑄1 𝑠 + 𝑄2 𝑠 2 + . . . +𝑄𝑛 𝑠 𝑛 ) =

= 𝑃0 + 𝑃1 𝑠+ . . . +𝑃2 𝑠 2 + . . . +𝑃𝑚 𝑠 𝑚 . (5.198)

Equating the coefficients of the terms of the same degree from the two sides
of equation (5.198), we obtain the coefficients 𝑄𝑟 and 𝑃𝑟 .

For the particular case when 𝑚 = 1 and 𝑛 = 3 we obtain [59]

ℎ∞ 0 0 0
−𝑀0 ℎ∞ 0 0 1 𝑃0
𝑀1 𝑄1 𝑃1
−𝑀0 ℎ∞ 0 𝑄2 = 0 .
1
− 𝑀2 𝑀1 −𝑀0 ℎ∞ 𝑄3 0
2
1
1
𝑀 − 2 𝑀2 𝑀1 −𝑀0 { 0 } {0}
[ 6 3 ]

The drawback of the method is the requirement of apriori selection of the


values of 𝑚 and 𝑛.

5.5.2. Identification Methods with Error Minimization

Denoting by 𝐻 𝑜 (𝑖𝜔) the experimental frequency response data, the error


at the frequency 𝜔𝑓 is defined by
𝑃(𝑖𝜔 )
𝜀𝑓 = 𝜀(𝑖𝜔𝑓 ) = 𝐻 𝑜 (𝑖𝜔𝑓 ) − 𝐻(𝑖𝜔𝑓 ) = 𝐻 𝑜 (𝑖𝜔𝑓 ) − 𝑄(𝑖𝜔𝑓 .
𝑓)

Multiplying by 𝑄(𝑖𝜔𝑓 ), the weighted error at point 𝑓 is


𝜀𝑓′ = 𝜀𝑓 · 𝑄(𝑖𝜔𝑓 ) = 𝑄(𝑖𝜔𝑓 ) 𝐻 𝑜 (𝑖𝜔𝑓 ) − 𝑃(𝑖𝜔𝑓 ). (5.199)
The polynomial coefficients of expression (5.186) can be evaluated solving
the set of linear simultaneous algebraic equations obtained by minimizing the sum
𝑁 2
𝑓
∑𝑓=1|𝜀𝑓′ | at all 𝑓 = 𝑁𝑓 experimental points [69].
Sanathanan and Koerner [107] have shown that, if the transfer function has
to be determined for frequencies extending several decades, the matrix of the
coefficients is ill conditioned and, especially at low frequencies, a good fit cannot
be obtained.
They suggested an iterative procedure which effectively eliminates the
above weighting. Equation (5.199) is modified by writing

194
𝜀𝑓 𝑄(𝑖𝜔𝑓 ) 𝐻 𝑜 (𝑖𝜔𝑓 ) 𝑄(𝑖𝜔𝑓 ) − 𝑃(𝑖𝜔𝑓 )
𝜀𝑓′′ =
𝑄(𝑖𝜔𝑓 )
𝐿
= 𝑄(𝑖𝜔𝑓 )
𝐿 𝐿
(5.200)
𝐿−1 𝐿−1

where the subscript 𝐿 corresponds to the iteration number.


From (5.200) we obtain
2
2 |𝐻 𝑜 (𝑖𝜔𝑓 )𝑄(𝑖𝜔𝑓 ) −𝑃(𝑖𝜔𝑓 ) | 2
|𝜀𝑓′′ | = 𝐿
2
𝐿
= |𝜀𝑓′ | 𝑊𝑓𝐿 (5.201)
| 𝑄(𝑖𝜔𝑓 ) |
𝐿−1

where
1
𝑊𝑓𝐿 = 2 . (5.202)
| 𝑄(𝑖𝜔𝑓 ) |
𝐿−1
Summing up for all 𝑓-s
𝑁𝑓 2 𝑁𝑓 2
𝐸 = ∑𝑓=1|𝜀′′𝑓 | =∑𝑓=1|𝜀𝑓′ | 𝑊𝑓𝐿 . (5.203)

Equation (5.203) is partially differentiated with respect to each of the


polynomial coefficients, and equated to zero to evaluate the coefficients. This
yields the following matrix equation

𝑉0 0 −𝑉2 0 𝑉4 𝑇1 𝑆2 −𝑇3 𝑆4 𝑃0 𝑆0
−𝑉4 ⋯ ⋯ 𝑃1
0 𝑉2 0 0 −𝑆2 𝑇3 𝑆4 −𝑇5 𝑇1
−𝑉4 𝑉6 ⋯ ⋯ 𝑃2
𝑉2 0 0 ⋯ 𝑇3 𝑆4 −𝑇5 −𝑆6 ⋯ 𝑆2
0 𝑉4 0 −𝑉6 0 ⋯ −𝑆4 𝑇5 𝑆6 −𝑇7 ⋯ 𝑃3 𝑇3
⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮
𝑇1 −𝑇3 𝑇5 = ⋮
−𝑆2 𝑆4 ⋯ 𝑈2 0 −𝑈4 0 ⋯ 𝑄1 0
𝑆2 𝑇3 −𝑆4 −𝑇5 𝑆6 ⋯ 0 𝑈4 0 𝑈6 ⋯ 𝑄2 𝑈2
𝑇3 −𝑆4 −𝑇5 𝑆6 𝑇7 ⋯ 𝑈4 0 −𝑈6 0 ⋯ 𝑄3 0
⋯ ⋯
𝑆4 𝑇5 −𝑆6 −𝑇7 𝑆8 ⋯ 0 𝑈6 0 −𝑈8 𝑄4 𝑈4
[⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ] { ⋮} {⋮}
(5.204)
where
𝑁 𝑁
𝑓
𝑉𝑟 = ∑𝑓=1 𝜔𝑓𝑟 𝑊𝑓𝐿 , 𝑓
𝑆𝑟 = ∑𝑓=1 𝜔𝑓𝑟 𝐻𝑅𝑜 (𝜔𝑓 ) 𝑊𝑓𝐿 ,
𝑁 𝑁 2 2
𝑓
𝑇𝑟 = ∑𝑓=1 𝜔𝑓𝑟 𝐻𝐼𝑜 (𝜔𝑓 ) 𝑊𝑓𝐿 , 𝑓
𝑈𝑟 = ∑𝑓=1 𝜔𝑓𝑟 ([𝐻𝑅𝑜 (𝜔𝑓 )] + [𝐻𝐼𝑜 (𝜔𝑓 )] ) 𝑊𝑓𝐿

in which
𝐻𝑅𝑜 (𝜔𝑓 ) + 𝑖 𝐻𝐼𝑜 (𝜔𝑓 ) = 𝐻 𝑜 (𝑖𝜔𝑓 )
and 𝑁𝑓 is the total number of experimental frequencies.

195
As 𝑄(𝑖𝜔𝑓 ) is not known initially, it is assumed that 𝑊𝑓𝐿 = 1 to obtain first
estimates of the quantities 𝑉𝑟 , 𝑇𝑟 , 𝑆𝑟 and 𝑈𝑟 , and to calculate the coefficients 𝑃𝑟 and
𝑄𝑟 from (5.204). The coefficients 𝑄𝑟 evaluated at iteration (𝐿 − 1) are used to
evaluate 𝑊𝑓𝐿 for the next iteration. The number of iterations necessary would vary
with the nature of the transfer function and the desired accuracy in the values of the
polynomial coefficients.
The book by Vlach [123] presents a series of computer programs which
can be used for the representation of the frequency response functions by series of
orthogonal functions, using Jacobi, Legendre, Chebyshev, Laguerre or Hermite
polynomials.
It can be shown that expressions (5.5), (3.56) and (3.186) are particular
forms of equation (5.187). This observation is the basis of the identification method
used by Dat [151] for systems with nonproportional damping.

196
CHAPTER 6

Identification of Linear Systems Using Multi Point


Harmonic Excitation

This chapter is devoted to the presentation of the identification methods


using excitation with harmonic forces applied simultaneously or successively at the
various points of the tested structure. The material is based on reference [13].
Though the successive application of a force at several points can be considered a
“single-point excitation” method, these procedures are discussed in this section
because they are particular cases of the general procedures described within the
theory of simultaneous multi-point excitation.

6.1. PRINCIPLE OF SIMULTANEOUS MULTI-POINT


EXCITATION METHODS

As shown in Chapter 5, almost all difficulties encountered in the analysis


of a measured frequency response curve are due to the fact that the dynamic
response of a structure is taking place in several modes of vibration simultaneously.
Elimination of one or several “unwanted modes” can be done by the
adequate placement of transducers and exciters, e.g. at nodes, or using anti-
symmetrical or symmetrical forcing distributions. Unfortunately this is seldom
sufficient. It was found useful to isolate individual modes of vibration, because this
facilitates the determination of the modal parameters and mode shapes.
In principle, the study of the response in a pure mode reduces to the analysis
of the response of a single-degree of freedom system using the identification
methods presented in Chapter 2. Therefore, the aim is to determine the necessary
conditions to excite a pure mode {𝛹 (𝑟) } of the associated undamped system.
The equation of motion of a discrete system with linear viscous damping
is (3.26)
[𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} (6.1)
which, expressed in terms of the modal coordinates (3.29), can be written
[ˋ𝑚ˏ] {𝑝̈ } + [𝑐] {𝑝̇ } + [ˋ𝑘ˏ] {𝑝} = {𝐹}. (6.2)
The problem is to find a harmonic excitation able to vibrate the system in
its 𝑟-th principal mode {𝛹 (𝑟) } at some frequency 𝜔, not necessarily the natural
frequency 𝜔𝑟 , i.e. to produce a response of the form
{𝑞} = 𝜅 {𝛹 (𝑟) } 𝑒 𝑖 (𝜔𝑡−𝜑) , (6.3)

197
where 𝜑 is the phase lag between excitation and response, and 𝜅 is some amplitude
factor.
For the sake of simplicity, consider 𝜅 = 1. The problem is to find the
excitation {𝑓} = {𝑓̂} 𝑒 𝑖 (𝜔𝑡+𝜑) able to produce a response of the form
{𝑞} = {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑡 . (6.3,a)
In view of (3.16,a) and (6.3,a)
(𝑟)
{𝑞} = ∑𝑁
𝑟=1{𝛹 } 𝑝𝑟 = {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑡 ,
therefore to isolate the 𝑟-th mode it is necessary to substitute into equation (6.2)
{𝑝} = {𝐼}𝑟 𝑒 𝑖𝜔𝑡 , (6.4)
where {𝐼}𝑟 is the 𝑟-th column of the identity matrix [𝐼] and, at the same time,
{𝐹} = {𝐹̂ } 𝑒 𝑖 (𝜔𝑡+𝜑) . (6.5)
Substituting (6.4) and (6.5) into (6.2), simplifying by 𝑒 𝑖𝜔𝑡 and separating
the real and imaginary parts, we obtain
([`𝑘ˏ] − 𝜔2 [`𝑚ˏ]) {𝐼}𝑟 = {𝐹̂ } 𝑐𝑜𝑠 𝜑 , (6.6)
𝜔 {𝑐}𝑟 = {𝐹̂ } 𝑠𝑖𝑛 𝜑 , (6.7)
where {𝑐}𝑟 is the 𝑟-th column of the modal damping matrix [𝑐].
Using equations (3.6), (3.8), (3.10), (3.20) and (3.28), we can write
(𝜔𝑟2 − 𝜔2 ) [𝛹]𝑇 [𝑀] {𝛹 (𝑟) } = [𝛹]𝑇 {𝑓′}, (6.8)
ω [𝛹]𝑇 [𝐶] {𝛹 (𝑟) } = [𝛹]𝑇 {𝑓′′}, (6.9)
where {𝑓′} and {𝑓′′} are the coincident and quadrature components of the excitation
{𝑓} = {𝑓′ + 𝑖 𝑓′′} 𝑒 𝑖𝜔𝑡 . (6.10)
Simplifying with [𝛹]𝑇 we obtain
{𝑓′} = (𝜔𝑟2 − 𝜔2 ) [𝑀] {𝛹 (𝑟) }, (6.11)
{𝑓′′} = 𝜔 [𝐶] {𝛹 (𝑟) }. (6.12)
It follows that, in order to excite the structure in one of the undamped
principal modes of vibration, the distribution of the excitation forces must have the
form
{𝑓} = {𝑓′ + 𝑖 𝑓′′} 𝑒 𝑖𝜔𝑡 = [(𝜔𝑟2 − 𝜔2 ) [𝑀] + 𝑖𝜔 [𝐶] ] {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑡 . (6.13)
The component {𝑓′}, in-phase with the displacement, is necessary to cancel
the elastic and inertia forces, and the component {𝑓′′}, phase shifted 90° ahead of
the displacements, is needed to counteract the damping.

198
The two components are related by the equation
𝜔
{𝑓′′} = [𝐶][𝑀]−1 {𝑓′} , (6.14)
𝜔𝑟2 −𝜔2

hence the ratio between the ‘active’ forces {𝑓′′} and the ‘reactive’ forces {𝑓′} is a
function of the excitation frequency ω. This makes difficult if not impossible the
practical realization of such a force distribution.
Another way to formulate the problem is to enquire whether it is possible
to excite the pure modes of the associate undamped system by using a distribution
of forces which are all in-phase, of the form
{𝑓} = {𝑓̂} 𝑒 𝑖𝜔𝑡 (6.15)
where {𝑓̂} is a vector with real elements.
There are three distinct occasions on which it is possible to excite real
“classical” modes of vibration:
a) For systems with negligible damping, any principal mode of vibration
(𝑟)
{𝛹 } can theoretically be excited at any frequency 𝜔 by a force distribution of the
form (3.25), which represents a principal mode of excitation

{𝑓̂} = {ℱ̂ (𝑟) } = [𝐾 − 𝜔2 𝑀]{𝛹 (𝑟) } = (𝜔𝑟2 − 𝜔2 ) [𝑀] {𝛹 (𝑟) } (6.16)

even if 𝜔 ≠ 𝜔𝑟 .
b) For systems with proportional damping, a principal mode {𝛹 (𝑟) } can be

excited at any frequency 𝜔 using a forcing of the form ( )

𝜔 2 2𝜔 2
{𝑓̂} = {ℱ̂ (𝑟) } = √(1 − 𝜔2) + (2𝜁𝑟 𝜔 ) 𝜔𝑟2 [𝑀]{𝛹 (𝑟) } (6.17)
𝑟 𝑟

where
𝑐 𝑘
𝜁𝑟 = 2𝑚 𝑟𝜔 , 𝜔𝑟2 = 𝑚𝑟 , (6.18)
𝑟 𝑟 𝑟

hence the forces {ℱ̂ (𝑟) } should be proportional to the inertia forces corresponding
to the modal displacements.
Indeed, if [𝑐] is diagonal, then according to equation (6.7) all modal forces
should be zero except 𝐹𝑟 , because the vector {𝑐}𝑟 contains only the element on the
𝑟-th row.

_____________________________________

( ) In order to simplify the notation, the vector of the appropriate
excitation required to produce a response in a principal mode {𝛹 (𝑟) } will be denoted
{ℱ̂ (𝑟) } in all cases.

199
Equations (6.6) and (6.7) can be written
𝜔 2
𝑘𝑟 (1 − 𝜔2 ) = 𝐹̂𝑟 𝑐𝑜𝑠 𝜑, (6.19)
𝑟
𝜔
2 𝜁𝑟 𝑘𝑟 = 𝐹̂𝑟 𝑠𝑖𝑛 𝜑, (6.20)
𝜔𝑟
hence
2 2 2
𝜔 𝜔
𝐹̂𝑟 = 𝑘𝑟 √(1 − 2 ) + (2𝜁𝑟 𝜔 ) ,
𝜔𝑟
𝐹̂𝑠 = 0. (𝑟 ≠ 𝑠) (6.21)
𝑟

and
2 𝜁𝑟 𝜔𝜔
𝑟
𝑡𝑎𝑛 𝜑 = 2 = 𝑡𝑎𝑛 𝜑𝑟 , (6.22)
1−𝜔2
𝜔𝑟

where 𝜑𝑟 is the characteristic phase lag defined by (3.98) for [𝐻] = [0].
“Equation (6.21) shows that when the damping matrix is diagonal it is
possible to excite a principal mode at any frequency by a set of forces all in phase”
[13].
By definition {𝐹̂ } = [𝛹]𝑇 {𝑓̂} so that
{𝑓̂} = [𝛹]−𝑇 {𝐹̂ }. (6.23)
From equation (3.21) it follows that
[𝛹]−𝑇 = [𝑀] [𝛹] [`𝑚ˏ]−1
which, substituted into (6.23), yields
{𝑓̂} = [𝑀] [𝛹] [`𝑚ˏ]−1 {𝐹̂ }. (6.24)
In the particular case when only 𝐹̂𝑟 ≠ 0
1
{𝑓̂} = [𝑀] {𝛹 (𝑟) } 𝑚 𝐹̂𝑟 . (6.25)
𝑟

Inserting (6.21) into (6.25) we obtain the equation (6.17), in which the
radical is proportional to the magnitude of the complex modal dynamic stiffness.
Therefore, for proportional damping, the forcing required to excite the
vibration
{𝑞} = {𝛹 (𝑟) } 𝑒 𝑖 (𝜔𝑡−𝜑𝑟 ), (6.26)
where 𝜑𝑟 has the value given in equation (6.22), is

2
𝜔2 𝜔 2
{ℱ (𝑟) } = √(1 − 𝜔2 ) + (2𝜁𝑟 𝜔 ) 𝜔2𝑟 [𝑀] {𝛹 (𝑟) }𝑒 𝑖𝜔𝑡 . (6.27)
𝑟 𝑟

In particular, at the phase resonance 𝜔 = 𝜔𝑟 and the characteristic phase


lag 𝜑𝑟 = 90°. The forcing required to excite the resonant vibration

200
𝜋
{𝑞} = {𝛹 (𝑟) } 𝑒 𝑖 (𝜔𝑟 𝑡− 2 ) = −𝑖 {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑟 𝑡 , (6.28)
is
{ℱ (𝑟) } = 2𝜁𝑟 𝜔𝑟2 [𝑀] {𝛹 (𝑟) }𝑒 𝑖𝜔𝑟 𝑡 , (6.29)
hence “the excitation forces must be proportional to the inertia forces
corresponding to the modal displacements”.
Equation (6.27) shows that the ratios of the various forces to be applied at
the various points of the structure do not depend on frequency. They are given by
the ratios of the forces required to excite the principal modes {𝛹 (𝑟) } at the
corresponding phase resonance.
If the mass matrix is diagonal, [𝑀] = [`𝑀ˏ], it follows that
(𝑟)
(𝑟)
𝑀1 𝛹1 𝜔𝑟2 𝑀𝑗 𝛹𝑗 𝜔𝑟2 (𝑟)
𝑀𝑁 𝛹𝑁 𝜔𝑟2
(𝑟) = . . .= (𝑟) =. . . = (𝑟) (6.30)
ℱ̂1 ℱ̂𝑗 ℱ̂𝑁

where 𝑀𝑗 is the mass at point 𝑗.


This conclusion can be reached directly. The necessary and sufficient
condition that all modal forces are zero, except 𝐹𝑟 , is that the vector {ℱ (𝑟) } be
proportional to the vector [𝑀] {𝛹 (𝑟) }, because, in virtue of the orthogonality
condition (3.6), the vector [𝛹]𝑇 {ℱ (𝑟) }, proportional to [𝛹]𝑇 [𝑀] {𝛹 (𝑟) }, contains only
the element of index 𝑟.
c) For systems with nonproportional damping, the 𝑟-th principal mode of
vibration {𝛹 (𝑟) } may be excited by in-phase force distributions only when the
system is driven at the 𝑟-th phase resonance frequency 𝜔𝑟 and the excitation has
the form
{𝑓̂} = {ℱ̂ (𝑟) } = 𝜔𝑟 [𝐶] {𝛹 (𝑟) }. (6.31)
Expression (6.31) is obtained from (3.95) for [𝐻] = [0] because at 𝜔 = 𝜔𝑟
the characteristic phase lag is 𝜑𝑟 = 90°, and the monophase forced mode {𝛷 (𝑟) } is
the same as the principal mode {𝛹 (𝑟) }.
Premultiplying by [𝛹]𝑇 , equation (6.31) becomes

̂ (𝑟) } = 𝜔𝑟 [𝛹]𝑇 [𝐶] {𝛹(𝑟) } = 𝜔𝑟 {𝑐}𝑟 .


[𝛹]𝑇 {ℱ (6.32)

Therefore, for the isolation of a principal mode of vibration, the modal


forces must be proportional to the column of the modal damping matrix of the same
index.
For 𝜔 = 𝜔𝑟 , equations (6.11) and (6.12) become
{𝑓′} = {0}, {𝑓′′} = 𝜔𝑟 [𝐶] {𝛹(𝑟) }, (6.33)

201
which can be used to demonstrate the following theorem: The necessary and
sufficient condition for a vibration to take place in a principal mode {𝛹 (𝑟) } is that
the response be in quadrature with the monophase excitation forces.
At frequencies other than the natural frequencies, the only in-phase
vibrations which may be excited by in-phase force distributions are the monophase
forced modes {𝛷 (𝑟) } discussed in § 3.2.3.

6.2. PHASE RESONANCE METHOD

The phase resonance method is based on equation (6.31) and on the


observations at point (c) above (§ 6.1). For linear systems, with viscous damping,
having a symmetric and positive definite matrix [𝐶], the force distribution given by
equation (6.33) enables the isolation of a pure mode {𝛹 (𝑟) } at the corresponding
phase resonance frequency 𝜔𝑟 .
This applies only to systems with a finite number of degrees of freedom.
In real structures, a finite number of forces cannot cancel the continuously
distributed damping forces. A pure mode cannot be achieved, only approximations
to the correct mode shape.
After the approximate determination of all resonance frequencies 𝜔𝑟
within the frequency range of interest, the most difficult operation is the isolation
of each mode, by the selection of the appropriate excitation.

6.2.1. Apportioning of Excitation Forces

A continuous structure can be excited in a limited number of accessible


points, the number of exciters applied simultaneously being even less.
The ideal condition, when the monophase (or 180° phase shifted) applied
forces produce velocities in-phase with the forces, can be achieved only with some
approximation.
The “complete” phase resonance is obtained only when the column vector
of response velocities, in quadrature with the forces, would be zero. In practice, this
condition is replaced by the minimization of a scalar, function of the velocities in
quadrature with the forces, measured at selected points of the structure.

6.2.1.1. Pure experimental methods

The first method of experimental apportioning of the excitation forces, for


the isolation of a pure mode, was proposed in 1950 by Lewis and Wrisley [70]. In
principle, the method is an iterative procedure to obtain the force distribution (6.33).
The experimental equipment consists of 24 independently controllable
electromagnetic exciters and 24 accelerometers attached at the driving points. The
exciters are driven in-phase or exact antiphase with each other by a source of
variable frequency.

202
First, the total mass of the structure is considered to be divided among the
exciters into several lumped masses, estimating arbitrarily the mass 𝑀𝑗 attached to
point 𝑗. A means is provided for indicating the modified exciter force, which is the
ratio of force to mass, for each exciter. The response velocities at the driving points
are derived integrating the accelerometer signals. The modified exciter force and
velocity for each point are displayed on the vertical and horizontal axes of a
cathode-ray oscilloscope, measuring the ratio of the signals and their phase
difference (by the method of ellipse).
The operation starts out with one exciter. The frequency of excitation is
varied until the local phase resonance is achieved. Then, the other exciters are
activated in turn, at the same frequency, adjusting the level of each force, making
the ratio of modified exciter force to velocity the same for each point, condition
similar to (6.30). A detailed description of the instrumentation and experimental
procedure is given in the original paper [70] and in Chapter13 of Bishplinghoff et
al [15].
Hawkins [51] described an experimental set up for the automatic
appropriation of the excitation forces in order to isolate a pure mode of vibration.
The equipment used for the simultaneous excitation in several points contains
servo-control circuits to maintain the phase resonance condition (zero phase shift
between forces and velocities). One of the feedback circuits controls the driving
frequency of all exciters, the others control only the force amplitude [63].
Bishop and Gladwell [13] suggested a procedure for testing the validity of
this assumption. A pure mode is excited at the phase resonance frequency by the
method of Lewis and Wrisley. Then, keeping constant the ratios of the forces
applied at the various points, the forcing frequency is changed. If the damping is
proportional and if the forcing distribution is correct, then the vibration should
continue to be in one mode, i.e. the points are moving in phase and the
displacements are 90° out of phase with the forces.

6.2.1.2. Semi-experimental methods

In order to avoid the inherent approximations of the iterative method of


Lewis and Wrisley, Traill-Nash [117, 118] developed an alternative procedure for
exciting pure natural modes. Based on experimental data obtained for the tested
structure, the method allows either the calculation of the force distribution which
can excite an undamped normal mode, or the systematic adjustment of excitation
forces by a relaxation technique for the isolation of pure modes. The method can
be applied to structures with nonproportional damping.
The equipment needed to implement this technique is a set of
electromagnetic exciters and pick-ups suitable for measuring the components of the
response in phase with the applied forces.
First, using single-point excitation, the natural frequencies 𝜔𝑟 are
determined using (for example) the method of Kennedy and Pancu [61]. Then, if

203
the system has 𝑁 degrees of freedom, 𝑁 linearly independent in-phase force
distributions
{𝑓̂}1 𝑠𝑖𝑛 𝜔𝑟 𝑡, {𝑓̂}2 𝑠𝑖𝑛 𝜔𝑟 𝑡, {𝑓̂}𝑁 𝑠𝑖𝑛 𝜔𝑟 𝑡, (6.34)
are applied in turn at each natural frequency 𝜔𝑟 and the coincident components
{𝑞𝑅 }1 𝑠𝑖𝑛 𝜔𝑟 𝑡, {𝑞𝑅 }2 𝑠𝑖𝑛 𝜔𝑟 𝑡, {𝑞𝑅 }𝑁 𝑠𝑖𝑛 𝜔𝑟 𝑡, (6.35)
of the corresponding responses are measured.
Based on these experimental data, a linear combination of the force
distributions (6.34) is constructed
{𝑓̂} = ∑𝑁 ̂
𝑠=1 𝛾𝑠 {𝑓 }𝑠 (6.36)
in which the real constants 𝛾𝑠 are determined to produce a null coincident response
{ 𝑞 𝑅 } = ∑𝑁
𝑠=1 𝛾𝑠 {𝑞𝑅 } = 0. (6.37)
𝑠
In view of equation (3.119), any distribution of in-phase forces can be
expressed as a linear combination of the monophase excitation modes {ℱ̂ (𝑟) }

{𝑓} = {𝑓̂} 𝑠𝑖𝑛 𝜔𝑡 = (∑𝑁 ̂ (𝑟) })𝑠𝑖𝑛 𝜔𝑡,


𝑟=1 𝜆𝑟 {ℱ (6.38)

where 𝜆𝑟 are real multipliers given by expression (3.121).


The corresponding response is (3.20)
(𝑟)
{𝑞} = ∑𝑁
𝑟=1 𝜆𝑟 {𝛷 } 𝑠𝑖𝑛 (𝜔𝑡 − 𝜑𝑟 ). (6.39)
The in-phase components of the response may be expressed in the form

{𝑞𝑅 } 𝑠𝑖𝑛 𝜔𝑡 = (∑𝑁 (𝑟)


𝑟=1 𝜆𝑟 𝑐𝑜𝑠 𝜑𝑟 {𝛷 })𝑠𝑖𝑛 𝜔𝑡. (6.40)

The condition (6.37) is satisfied if


𝜆1 = 𝜆2 = . . . = 𝜆𝑟−1 = 𝜆𝑟+1 = . . . = 𝜆𝑁 = 0, and 𝜔 = 𝜔𝑟 . (6.41)

But when 𝜔 = 𝜔𝑟 , 𝜑𝑟 = 90° and {𝛷 (𝑟) } ≡ {𝛹 (𝑟) }, and equation (6.39)


yields
{𝑞} = −𝜆𝑟 {𝛹 (𝑟) } 𝑐𝑜𝑠 𝜔𝑟 𝑡 , (6.42)
hence a natural mode {𝛹 (𝑟) } is isolated at the frequency 𝜔𝑟 .
Note that the distributions (6.34) may contain only one force, applied each
time at another point, hence the practical problem is reduced to the measurement
of the response for single point excitation. For this, it is necessary to know exactly
the number of degrees of freedom of the structure and to use just so many exciters
and pick-ups, which is the main difficulty of the method. Also, unless the forcing
frequency is exactly a natural frequency 𝜔𝑟 , the equations (6.37) have no solution.

204
In the paper [117] Traill-Nash introduces the concept of the “effective
number of degrees of freedom” of a continuous structure relevant to a given natural
frequency 𝜔𝑟 . He indicates the number of exciters and pick-ups necessary for an
efficient measurement, number which is generally different for the isolation of
various natural modes.
In order to avoid the wrong estimation of the number of degrees of freedom
𝑁, Asher [2] proposed another method which combines experiment and
calculations, having essentially the same analytical basis.
Instead of attempting to solve the equations (6.37) to find the correct
forcing distribution, Asher uses it to find the natural frequencies of the system.
According to (3.140), these will be given by the roots of the determinantal equation
𝑑𝑒𝑡 [𝑞𝑅 ] = 0 (6.43)
where
[𝑞𝑅 ] = [ {𝑞𝑅 }1 {𝑞𝑅 }2 . . . {𝑞𝑅 }𝑁 ] (6.44)
is the matrix of the real components of the response, produced by 𝑁 force
distributions, columns of the square matrix
[𝑓] = [ {𝑓}1 {𝑓}2 . . . {𝑓}𝑁 ]. (6.45)
As matrices (6.44) and (6.45) satisfy an equation of the form
[𝑞𝑅 ] = [𝛼𝑅 ] [𝑓], (6.46)
where [𝛼𝑅 ] is the real part of the receptance matrix, equation (6.43) is equivalent
to
𝑑𝑒𝑡 [𝛼𝑅 ] = 0, (6.47)
because the force distributions (6.45) must be linearly independent, hence
𝑑𝑒𝑡 [𝑓] ≠ 0.
The methodology proposed by
Asher is the following: The test starts with
a single exciter and a single displacement
pick-up. The frequency is swept starting at
a low frequency. The value at which the
real part of the single receptance changed
sign gives the approximate value of the
first natural frequency.
A second exciter and pick-up is
then added, so that four receptances can be
measured, and the frequency sweep
repeated over the range encompassing the
first natural frequency. A graph is sketched
(Fig.6.1) showing the variation with frequency of the determinant of the real part
of the receptance matrix. The first zero should give a better approximation to the
first natural frequency, and the second zero should give an estimation of the second
natural frequency.

205
A third exciter and a third pick-up are then added, measuring 3𝑥3 = 9
receptances, redefining the first two natural frequencies and estimating the third.
The process continues in this way throughout the frequency range of interest, and
in this way the correct number of exciters can be systematically determined for a
system whose dynamic characteristics (hence degrees of freedom of interest) were
unknown before the test.
Finally, once the effective number of degrees of freedom and the values of
the natural frequencies are known, one can come back to equations (6.36) and
(6.37) to determine the force distribution able to excite a pure mode.
A variant of Asher’s method was presented by Bishop and Johnson [13],
not sustained by an experimental verification. A generalization has been suggested
by Angélini and Darras [1].
An improved method for the appropriation of the excitation forces has been
reported by Clerc [24, 25]. The method is implemented using an equipment
containing electrodynamic exciters and velocity pick-ups. It consists of the
calculation of a force distribution of the form (6.36)
{𝑓̂} = [𝑓] {𝛾}, (6.48)
where the vector
{𝛾} = { 𝛾1 𝛾2 . . . 𝛾𝑁 }𝑇

is determined from the condition to minimize the column vector of the quadrature
components of velocity {𝑞̇ 𝐼 } at various points, considered as a parasite response.
For this, a relative residual scalar is defined as
𝑇 2
∑𝑁
𝑠=1 𝜔𝑠 (𝑞𝐼𝑠 )
{𝑞̇ 𝐼 } [`𝜔𝑠 ˏ] {𝑞̇ 𝐼 } ̇
𝜀= 𝑇 = (6.49)
{𝑓̂} {𝑓̂} ̂ 𝑇 ̂
{𝑓} {𝑓}

where [`𝜔𝑠 ˏ] is an arbitrary weighting matrix, and the vector {𝛾} is determined for
which 𝜀 has a minimum value. The detailed presentation of this method [24] is
beyond the aim of this book.
A relatively widely used method of semi-automatic appropriation of
excitation forces has been developed by Deck [29]. The equipment contains
velocity pick-ups and an analogue computer is used to calculate a global criterion
quantifying the degree of isolation of a mode, expressed by the positive quantity
∑𝑀
𝑗=1|𝑞̇ 𝐼𝑗 | ∑𝑀 ̂
𝑗=1| 𝑞̇ 𝑗 𝑠𝑖𝑛 𝜃𝑗 |
𝜉= = ∑𝑀 | ̂𝑞̇𝑗 𝑐𝑜𝑠 𝜃𝑗 |
, (6.50)
∑𝑀
𝑗=1|𝑞̇ 𝑅𝑗 | 𝑗=1

where |𝑞̇ 𝑅𝑗 | and |𝑞̇ 𝐼𝑗 | are the magnitudes of the real and imaginary components of
the velocity at the point 𝑗, and 𝜃𝑗 is the phase lag between the velocity and force.
The summations are over a number 𝑀 of pick-ups, larger than the number 𝑁 of
exciters.

206
The quantity 𝜉, which can be considered an average global phase lag, is a
function of the frequency ω and the amplitude 𝑓̂ℓ of the driving forces, having a
minimum value for the optimal distribution of forces.
The evolution of 𝜉 during the adjustment of the forces of three exciters, for
the excitation of a pure mode, is illustrated in Fig.6.2,b.

The procedure starts with an arbitrary force distribution 𝑓̂1′ , 𝑓̂2′ , 𝑓̂3′ .The
frequency is swept around an approximate value of 𝜔𝑟 (determined beforehand) to
find the value 𝜔′ corresponding to the minimum of the function 𝜉(𝜔) – point ①
in Fig.6.2,b. Then, keeping 𝜔′ = 𝑐𝑜𝑛𝑠𝑡., the amplitude of the force 𝑓̂1 is adjusted
to determine the value 𝑓̂1′′ for which 𝜉 is a minimum (point ②). The excitation
frequency is swept again to obtain the value 𝜔′′ at which 𝜉 is a minimum (point
③). The procedure continues, adjusting the force 𝑓̂2 to determine the value 𝑓̂2′′
(point ④) and a new frequency 𝜔′′′ (point ⑤) for which 𝜉 is a minimum. Next, 𝑓̂3
is modified, determining the amplitude 𝑓̂3′′ (point ⑥) and the frequency 𝜔𝘐𝘝 (point
⑦). During the adjustment of each force 𝑓̂𝑗 and the corresponding adjustment of
the frequency 𝜔, the value of 𝜉 decreases, until the optimum excitation (last column
in Fig.6.2,a) is obtained.
The theoretical basis of the method and the demonstration of its
convergence are presented in [29]. A similar method, based on the minimization of
the positive quantity
𝑅 = ∑𝑀 ̂
𝑗=1|𝑞̇ 𝑗 𝑠𝑖𝑛 𝜃𝑗 |

has been proposed by Deck in [30].


A procedure for the iterative adjustment of the excitation, based on
Galerkin’s method, is presented in [139].

207
6.2.1.3. Method of “reinjection”

An interesting solution for the appropriation of excitation forces, which in


fact adapts a structure to a given excitation, has been reported by de Vries [132,
133].
As follows from (6.30), in the case of proportional damping, the forcing
distribution able to excite a
system in one of its principal
modes is independent of the
absolute value of damping,
being the same as for the
undamped structure. These
forces cancel the reactive
(elastic and inertia) internal
forces at any frequency 𝜔,
except at phase resonance
𝜔 = 𝜔𝑟 , where a stationary
response is not possible in
the absence of damping.
Let 𝑓̂𝑗 be the amplitude of the harmonic excitation force and |𝑞̇ 𝑗 | the
absolute value of the response velocity at point 𝑗 (𝑗 = 1, 2, . . . , 𝑁). A family of
|𝑞̇ 𝑗 | 𝑓̂ 𝑗
curves versus are drawn on the same graph (Fig.6.3) for different frequencies
|𝑞̇ 1 | 𝑓̂ 1
𝜔. The curves of any family define a zone of intersection inside which a point can
𝑓̂ 𝑗
be placed whose abscissa defines an approximate value of the ratio , defining a
𝑓̂ 1
first estimation of the distribution of forces able to excite the response in a pure
mode.
If the curves are convergent in a point, the mode is perfectly isolated,
𝑓̂𝑗 |𝑞̇ 𝑗 | 𝛹𝑗(𝑟)
because for any 𝜔 the ratios (𝑓̂ ) produce the same response = .
1 𝑎𝑑𝑗𝑢𝑠𝑡𝑒𝑑 |𝑞̇ 1 | 𝛹1(𝑟)
If not, the existence of the triangular zone in Fig.6.3 proves that the damping is
nonproportional [6].
The presence of non-diagonal elements in the matrix [𝑐] is due to the fact
that the local distribution of damping is different from that of the masses and
stiffnesses.
Using feedback circuits, a signal proportional to the velocity at a point of
the structure is added to the exciter signal, i.e. a component proportional to velocity,
having the nature of a force of viscous damping, is added to the initial force. Using
this method of reinjection, the structure is adapted to a given excitation by means
of a controlled modification of the structure damping.

208
The additional modal forces have the form
{𝐹 ∗ } = [𝑐 ∗ ] {𝑝̇ } (6.51)
therefore equation (6.2) becomes
[ˋ𝑚ˏ] {𝑝̈ } + [𝑐 − 𝑐 ∗ ] {𝑝̇ } + [ˋ𝑘ˏ] {𝑝} = {𝐹}. (6.52)
In order to have proportional damping, the matrix [𝑐 ∗ ]
must be chosen such

that [𝑐 − 𝑐 ] be diagonal. In practice, it suffices to cancel out all non-diagonal
elements of the first row and column of the matrix [𝑐 − 𝑐 ∗ ].
The experimental adaptation of a structure to correspond to a system with
nonproportional damping is carried out by an iterative procedure.
Using the approximate force distribution, determined with the aid of the
curves from Fig.6.2, a frequency very close to the phase resonance is determined,
based on the response of a reference pick-up which measures the velocity 𝑞̇ 1 . The
velocities at the other points are phase shifted with respect to 𝑞̇ 1 with angles 𝜃𝑗1 .
The gain in the feedback (reinjection) circuit of each exciter is varied to determine
the value of the gain for which 𝜃𝑗1 = 0. At the end of this operation, the velocities
at all points are in-phase with the velocity at the reference point 1, but this has a
phase lag with respect to the excitation. The forcing frequency is modified and the
gain of each feedback circuit is adjusted modifying the “damping” introduced
locally, until obtaining the phase resonance.
The modern multi-point excitation techniques are less elaborate than the
classical method [140] and the force appropriation is no more required [1].

6.2.2. Determination of Modal Parameters

6.2.2.1. Method of “work introduced”

This method, presented in 1950 by G. de Vries [125], is based on the


measurement of the work done by the (appropriated) excitation forces on the
displacement in a natural mode of vibration, at phase resonance.
If the forcing
{𝑓} = {ℱ̂ (𝑟) } 𝑒 𝑖𝜔𝑟 𝑡 (6.53)
excites the quadrature response
{𝑞} = −𝑖 {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑟 𝑡 , (6.54)
introducing (6.53) and (6.54) into (6.1) and separating the real and imaginary parts,
we obtain the equilibrium equations of the reactive and active forces
([𝐾] − 𝜔2 [𝑀]) {𝛹 (𝑟) } = {0},
{ℱ̂ (𝑟) } = 𝜔𝑟 [𝐶] {𝛹 (𝑟) }. (6.55)

209
For systems with diagonal modal damping matrix, in view of notations
(6.18), equation (6.55) becomes
{ℱ̂ (𝑟) } = 2 𝜁𝑟 𝜔𝑟2 [𝑀] {𝛹 (𝑟) }. (6.56)
𝑇
Premultiplying through by {𝛹 (𝑟) } and using notation (3.8), we obtain
(𝑟) 𝑇 (𝑟)
{𝛹 ̂ }
} {ℱ
𝑚𝑟 = . (6.57)
2 𝜁𝑟 𝜔2𝑟
Therefore, the modal mass 𝑚𝑟 can be determined from the work introduced
at phase resonance
𝑇 (𝑟) (𝑟)
{𝛹 (𝑟) } {ℱ̂ (𝑟) } = ∑𝑁
𝑗=1 𝛹𝑗 ℱ𝑗
̂

and the value of the damping ratio 𝜁𝑟 determined using a different method (e.g.: in
free vibrations).
The main drawback of the method is the dependence of the modal mass 𝑚𝑟
on the damping ratio 𝜁𝑟 . In actual structures, 𝜁𝑟 does not remain constant in time,
so that small errors on 𝜁𝑟 produce large absolute errors on 𝑚𝑟 [7].
In the case of nonproportional damping, introducing the square matrix
[ℱ̂ ] = [ {ℱ̂ (1) } {ℱ̂ (2) } . . . {ℱ̂ (𝑟) } . . . {ℱ̂ (𝑁) }] ,
equation (6.32) yields the modal damping matrix
1
[𝑐] = [𝛹]𝑇 [ℱ̂ ] [` ˏ]. (6.58)
𝜔 𝑟

6.2.2.2. Method of additional masses

The method of additional masses was developed in 1952 by G. de Vries


[126] as a procedure for the experimental determination of the modal masses. It is
a generalization of the technique described in § 2.6.1.1 for single-degree-of-
freedom systems.
First, a pure mode {𝛹 (𝑟) } is excited at the frequency 𝜔𝑟 using a set of
appropriated forces {ℱ̂ (𝑟) }. Then, known masses ∆𝑀𝑗 are added at various points on
the structure (usually at the driving points).
This destroys the resonance condition. In order to reestablish the phase
resonance it is necessary to vary the frequency (with ∆𝜔𝑟 ) and the magnitude
(eventually the location) of the excitation forces. Knowing {𝛹 (𝑟) } and measuring
∆𝜔𝑟 it is possible to calculate 𝑚𝑟 .
In the mostly used variant of the method, the additional masses ∆𝑀𝑗 are
selected such that the mode shape {𝛹 (𝑟) } remains practically unchanged [73].
The addition of masses is equivalent to the application of inertia forces
proportional to 𝛹𝑗(𝑟) ∆𝑀𝑗 . These forces are added to the initial set of appropriated

210
forces which, in the case of proportional damping, should be proportional to the
products 𝛹𝑗(𝑟) 𝑀𝑗 .
In order not to excite other modal shapes {𝛹 (𝑠) }, the work done per cycle
of vibration by the resultant forces should be zero
(𝑠) (𝑟)
∑𝑁
𝑠=1 𝛹𝑗 𝛹𝑗 (𝑀𝑗 + ∆𝑀𝑗 ) = 0, (𝑠 ≠ 𝑟)
which is equivalent to the orthogonality condition (3.6) when the mass matrix is
diagonal. Hence, the mode shape will remain constant provided that the modal mass
matrix remains diagonal.
The condition is satisfied selecting the additional masses ∆𝑀𝑗 proportional
to the masses 𝑀𝑗 . This cannot be made straightforward, because both the
distribution and the values of these masses are unknown.
The same problem occurs at the excitation of pure modes of vibration,
where, theoretically, the mass distribution should have been known in order to
apply at each point 𝑗 forces proportional to 𝛹𝑗(𝑟) 𝑀𝑗 . Practically it has been proved
that the force appropriation can be done by trial and error, satisfying the phase
criterion.
Because there is a direct analogy between the two problems, the following
conclusion can be drawn:
If, for a given number of excitation points, the optimal force distribution
(which produces a minimum phase lag between various points) has been
determined, the best distribution of additional masses, with an equal number to the
forces, is that in which the masses are proportional to the excitation forces placed
at the same points [126]
(𝑟) (𝑟) (𝑟)
∆𝑀1 𝛹1 ∆𝑀2 𝛹2 ∆𝑀𝑁 𝛹𝑁
(𝑟) = (𝑟) =. . . = (𝑟) . (6.59)
ℱ̂1 ℱ̂2 ℱ̂𝑁
An imperfect excitation gives rise to a sum of modes phase shifted with
each other, while the existence of non-diagonal elements in the matrix of modal
masses determines a distortion of the mode shape by a superposition of in-phase
modes.
By the attachment of additional masses, the equations (6.1) become
[𝑀 + ∆𝑀] {𝑞̈ } + [𝐶] {𝑞̇ } + [𝐾] {𝑞} = {𝑓} . (6.60)
Imposing the invariance of the deformation

{𝑞} = −𝑖 {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑟 𝑡 , (6.61)
where 𝜔𝑟∗ is the new resonance frequency under the action of the readjusted force
distribution

{𝑓} = {ℱ̂ (𝑟)∗ } 𝑒 𝑖𝜔𝑟 𝑡 , (6.62)

211
and introducing (6.61) and (6.62) into (6.60) we obtain
[𝜔𝑟∗ [𝐶] − 𝑖([𝐾] − 𝜔𝑟∗2 [𝑀 + ∆𝑀])] {𝛹 (𝑟) } = {ℱ̂ (𝑟)∗ } (6.63)
or
(𝜔𝑟∗ [𝐶] − 𝑖[𝐾 − 𝜔𝑟∗2 𝑀]) {𝛹 (𝑟) } = {ℱ̂ (𝑟)∗ } − 𝑖𝜔𝑟∗2 [∆𝑀]{𝛹 (𝑟) } . (6.63,a)
Equating the real and imaginary parts we obtain
𝜔𝑟∗ [𝐶] {𝛹 (𝑟) } = {ℱ̂ (𝑟)∗ }, (6.64)
[𝐾 − 𝜔𝑟∗2 𝑀] {𝛹 (𝑟) } = 𝜔𝑟∗2 [∆𝑀]{𝛹 (𝑟) } . (6.65)
The right hand side of equation (6.65) represents the inertia forces
corresponding to the attached masses. They are components of the excitation in
quadrature with the forces {ℱ̂ (𝑟)∗ }. Because for proportional damping, in view of
(6.18), expression (6.64) becomes
2 𝜁𝑟 𝜔𝑟 𝜔𝑟∗ [𝑀] {𝛹 (𝑟) } = {ℱ̂ (𝑟)∗ }, (6.66)
the resultant of two components phase shifted with 90° is proportional to [𝑀] {𝛹 (𝑟) },
which ensures the invariance of the modal vector {𝛹 (𝑟) }.
Premultiplying (6.65) by {𝛹 (𝑟) } 𝑇 and using notations (3.8) and (3.9) we
obtain
𝑘𝑟 − 𝜔𝑟∗2 𝑚𝑟 = 𝜔𝑟∗2 {𝛹 (𝑟) } 𝑇 [∆𝑀] {𝛹 (𝑟) }. (6.67)
Denoting
{𝛹 (𝑟) } 𝑇 [∆𝑀] {𝛹 (𝑟) } = ∆𝑚𝑟 , (6.68)
equation (6.67) becomes
𝑘𝑟 − 𝜔𝑟∗2 𝑚𝑟 = 𝜔𝑟∗2 ∆𝑚𝑟 . (6.69)
Considering 𝑘𝑟 = 𝑐𝑜𝑛𝑠𝑡., implying {𝛹 (𝑟) } 𝑇 [𝐾] {𝛹 (𝑟) } = 𝑐𝑜𝑛𝑠𝑡., hence
{𝛹 (𝑟) } = 𝑐𝑜𝑛𝑠𝑡., in view of (3.10) we obtain
∆𝑚𝑟
𝑚𝑟 = . (6.70)
𝜔 2
( 𝑟∗) −1
𝜔𝑟

If [𝑀] is a diagonal matrix


(𝑟)2
∆𝑚𝑟 = ∑𝑁
𝑗=1 𝛹𝑗 ∆𝑀𝑗 (6.71)
hence the modal masses are
(𝑟)2
∑𝑁
𝑗=1 𝛹𝑗 ∆𝑀𝑗
𝑚𝑟 = 𝜔
2 . (6.72)
( 𝑟∗ ) −1
𝜔𝑟
Denoting
𝜔𝑟∗ = 𝜔𝑟 + ∆𝜔𝑟 , (6.73)

212
for small increments ∆𝜔𝑟 of the resonance frequency, expression (6.70) becomes

𝑚𝑟 ≅ − ∆𝑚
∆𝜔
𝑟 𝜔𝑟
2
= − 2 𝜔∆𝜔𝑟 ∑𝑁
𝑗=1 𝛹𝑗
(𝑟)2
∆𝑀𝑗 . (6.74)
𝑟 𝑟

The modal mass of the 𝑟-th natural mode


𝑚𝑟 = ∑𝑁
(𝑟)2
𝑗=1 𝛹𝑗 𝑀𝑗 (6.75)
(𝑟)
depends on the arbitrary multiplier factor from the expression of 𝛹𝑗 which is
determined only after experimentation, when processing the test data.
For this reason, the concept of equivalent mass is often used, which is
independent of the normalization of the modal vector and can be determined during
testing [103].
The equivalent mass at point 𝑗, for the 𝑟-th mode, is the mass which
displaced with the velocity of point ℓ has the same kinetic energy as the whole
structure when it responds in the respective mode
1
𝑀𝑒𝑞ℓ = (𝑟)2 ∑𝑁
𝑗=1 𝑀𝑗 𝛹(𝑟)2 . (6.76)
𝛹ℓ 𝑗

This yields the following relation between the modal mass and the
equivalent masses at various points
(𝑟)2 (𝑟)2 (𝑟)2
𝑚𝑟 = 𝑀𝑒𝑞1 𝛹1 = 𝑀𝑒𝑞2 𝛹2 = . . . = 𝑀𝑒𝑞ℓ 𝛹ℓ . (6.77)
After the attachment of the additional masses ∆𝑀𝑗 , the variation of the
equivalent mass is
1
∑𝑁
(𝑟)2
∆𝑀𝑒𝑞ℓ = (𝑟)2 𝑗=1 ∆𝑀𝑗 𝛹𝑗 . (6.78)
𝛹ℓ
According to equations (6.72), (6.75) and (6.78)
∆𝑀𝑒𝑞ℓ
𝑀𝑒𝑞ℓ = 2 . (6.79)
𝜔
( 𝑟∗ ) −1
𝜔𝑟

If ∆𝑀𝑒𝑞ℓ is calculated for several values ∆𝑀𝑗


𝜔𝑟 2
and plotted as a function of (( ) − 1), the resulting
𝜔𝑟∗
graph is a straight line (Fig.6.4) passing through the
origin. The inclination to the vertical axis yields 𝑀𝑒𝑞ℓ
and equation (6.77) gives 𝑚𝑟 .
In practice it was noticed that as the masses
∆𝑀𝑗 increase, the expression (6.79) is no more linear,
and it is recommended to limit them to ∆𝑀𝑗 = 2 𝜁𝑟 𝑀𝑗
[126].

213
The application of the method is laborious, requiring many operations and
a reappropriation of the excitation forces after the attachment of additional masses
(for structures with nonproportional damping) [40]. It was replaced by the method
of electrical stiffnesses [130].
Gauzy [44] and Kappus [134] suggested variants of the method of
additional masses in which the modal shape is not conserved after the attachment
of masses.

6.2.2.3. Method of forces in quadrature

The method of forces in quadrature, proposed in 1963 by G. de Vries [131],


is based on the use of reactive forces, in quadrature with the initial excitation forces,
which at resonance have an action similar to that of the inertia forces produced by
the additional masses in the above method.
Introducing in equation (6.1) the condition that an excitation

{𝑓} = {ℱ̂ (𝑟) } (1 + 𝑖 𝜆) 𝑒 𝑖𝜔𝑟 𝑡 , (6.80)
excites a response

{𝑞} = −𝑖 {𝛹 (𝑟) } 𝑒 𝑖𝜔𝑟 𝑡 , (6.81)
we obtain

(𝜔𝑟∗ [𝐶] − 𝑖 [𝐾 − 𝜔𝑟∗2 𝑀]) {𝛹 (𝑟) } = (1 + 𝑖 𝜆){ℱ̂ (𝑟) } . (6.82)

Equating the real and imaginary parts, we obtain

𝜔𝑟∗ [𝐶] {𝛹 (𝑟) } = {ℱ̂ (𝑟) }, (6.83)


[𝐾 − 𝜔𝑟∗2 𝑀] {𝛹 (𝑟) } = −𝜆 {ℱ̂ (𝑟) } . (6.84)
Premultiplying in (6.84) by {𝛹 (𝑟) } 𝑇 and using notations (3.8) and (3.9) we
obtain
𝑘𝑟 , −𝜔𝑟∗2 𝑚𝑟 = −𝜆 {𝛹 (𝑟) } 𝑇 {ℱ̂ (𝑟) }, (6.85)
and, in virtue of (3.10),
𝑚𝑟 (𝜔𝑟2 − 𝜔𝑟∗2 ) = −𝜆 {𝛹 (𝑟) } 𝑇 {ℱ̂ (𝑟) }, (6.86)
which yields a formula for the calculation of modal masses
𝑇 (𝑟)
𝜆 {𝛹 𝑟 } {ℱ
( ) ̂ }
𝑚𝑟 = 𝜔𝑟∗2 −𝜔𝑟2
. (6.87)

In the case of proportional damping, after premultiplying by {𝛹 (𝑟) } 𝑇 ,


equation (6.83) can be written
𝜔𝑟∗ 𝑐𝑟 = {𝛹 (𝑟) } 𝑇 {ℱ̂ (𝑟) }, (6.88)

214
or
2 𝜁𝑟 𝜔𝑟 𝜔𝑟∗ 𝑚𝑟 = {𝛹 (𝑟) } 𝑇 {ℱ̂ (𝑟) }. (6.89)
Dividing (6.86) by (6.89) we obtain
𝜔𝑟∗2 −𝜔𝑟2
=𝜆 (6.90)
2 𝜁𝑟 𝜔𝑟 𝜔𝑟∗
which yields the modal damping ratio
1 𝜔𝑟∗2 −𝜔𝑟2
𝜁𝑟 = . (6.91)
2𝜆 𝜔𝑟∗ 𝜔𝑟
Expression (6.90) can be written
2
𝜔∗
1 ( 𝑟 ) −1 𝜔∗
𝜆=
2𝜁𝑟
𝜔𝑟
𝜔∗𝑟
= 2𝜁1 𝑟 𝑟
𝜔
( 𝜔 𝑟 − 𝜔𝑟∗). (6.92)
𝜔𝑟
𝑟
𝜔𝑟∗
Differentiating with respect to the dimensionless frequency we obtain
𝜔𝑟
𝑑𝜆 1 𝜔𝑟∗2
𝜔∗
= 2𝜁 (1 + ). (6.93)
𝑑( 𝑟) 𝑟 𝜔𝑟2
𝜔𝑟

The value of the derivative (6.93) at 𝜔𝑟 = 𝜔𝑟∗ is

𝑑𝜆 1
[ ] =
𝜔∗
𝑑( 𝑟 ) 𝜔∗ 𝜁𝑟 . (6.94)
𝜔𝑟 𝑟 =1
𝜔𝑟

In the vicinity of the frequency 𝜔𝑟∗ = 𝜔𝑟 , the


relation "𝜔𝑟∗ − 𝜆” has the form
∆𝜆 1
∆𝜔∗𝑟
≅ . (6.95)
𝜁𝑟
𝜔𝑟

It follows that [45]


∆𝜔𝑟∗ ∆𝑇𝑟∗
𝜁𝑟 ≅ 𝜔 = −𝑇 (6.96)
𝑟 ∆𝜆 𝑟 ∆𝜆
where
2𝜋 2𝜋
𝑇𝑟 = , 𝑇𝑟∗ = .
𝜔𝑟 𝜔𝑟∗
The graph of 𝑇𝑟∗ versus λ (Fig.6.5) is a “resonance microcurve” which
crosses the vertical axis at the point of ordinate 𝑇𝑟 . The tangent to the curve at this
point is inclined an angle 𝜃 from a horizontal line. The modal damping ratio 𝜁𝑟 is
given by

215
𝑡𝑎𝑛𝜃
𝜁𝑟 = 𝑇 . (6.97)
𝑟

From (6.89) we obtain 𝑚𝑟 and from (3.10) – the modal stiffness 𝑘𝑟 .


Comparing expressions (6.92) and (6.22) it follows that
𝜆 = −𝑐𝑜𝑡𝑎𝑛 𝜑𝑟 , (6.98)
hence the frequency dependence of the parameter 𝜆 is similar to the variation of the
phase lag around the resonance.
Equation (6.83) can also be written
𝜔𝑟∗ {𝛹 (𝑟) } = [𝐶]−1 {ℱ̂ (𝑟) }. (6.99)
When {ℱ̂ (𝑟) } = 𝑐𝑜𝑛𝑠𝑡., the velocity {𝑞̇̂ } = 𝜔𝑟∗ {𝛹 (𝑟) } at all points is
invariable for any value 𝜆, hence for any 𝜔𝑟∗ .
Reciprocally, if the phase resonance condition is maintained during the
application of forces in quadrature, the invariance of velocity is a criterion for the
quality of mode isolation.

6.2.2.4. Method of polar plots

The complex modal impedance of an isolated mode is


𝑘 𝜔 𝜔
𝑍̅𝑟 = 𝑐𝑟 − 𝑖 ( 𝜔𝑟 − 𝑚𝑟 𝜔) = 𝑚𝑟 𝜔𝑟 [2𝜁𝑟 − 𝑖 ( 𝜔𝑟 − 𝜔 )] (6.100)
𝑟

and the complex modal mobility is


𝜔𝑟 𝜔
̅ 𝑟 = 𝑍𝑟̅ −1 = 1 2𝜁𝑟 +𝑖 ( 𝜔 −𝜔𝑟)
ℳ . (6.101)
𝑚𝑟 𝜔𝑟 4𝜁2 +(𝜔𝑟− 𝜔 )2
𝑟 𝜔 𝜔𝑟

The polar plots of the functions ℳ ̅ 𝑟 and 𝑍𝑟̅ ∗ , where 𝑍̅𝑟∗ is the complex
conjugate of 𝑍𝑟̅ , are illustrated in Fig.6.6.
The graph of the function 𝑍𝑟̅ ∗ is a straight line of equation
ℛℯ(𝑍𝑟̅ ∗ ) = 𝑐𝑟 2𝜁𝑟 𝑚𝑟 𝜔𝑟
which intersects the real axis at the point of frequency 𝜔𝑟 .
𝑘
The point of frequency 𝜔𝑗 has an ordinate (𝜔𝑟 − 𝑚𝑟 𝜔𝑗 ). The line (∆)
𝑗
which connects this point to the origin has a slope
𝜔𝑟 𝜔𝑗
𝑘𝑟
− 𝑚𝑟 𝜔𝑗 −
𝜔𝑗 𝜔𝑗 𝜔𝑟
𝑡𝑎𝑛 𝜑𝑗 = = . (6.102)
𝑐𝑟 2𝜁𝑟

The graph of the function ℳ ̅ 𝑟 is a circle centered on the real axis and
tangent at origin to the imaginary axis. The circle crosses the real axis at the point
of frequency 𝜔𝑟 and the line (∆) at the point of frequency 𝜔𝑗 .

216
Based on the frequencies 𝜔𝑗 of several points along the circle, the
frequency 𝜔𝑟 is determined using the graphical construction from Fig.6.7 [7]. First,
a line (𝐷) is drawn parallel to the imaginary axis. Let 𝑂1 be its crossing point with
the real axis. A frequency axis is drawn, with the origin in 𝑂1 , on which points are
marked at abscissae 𝜔𝑗 . The points of frequencies 𝜔𝑗 marked on the circle are
connected to the origin 𝑂 by straight lines which intersect the line (𝐷) at the points
of ordinates 𝜎𝑗 . The points (𝑗), of coordinates (𝜔𝑗 , 𝜎𝑗 ), are plotted in the reference
system (𝐷 − 𝜔) and a straight line is drawn through the points, which crosses the
𝜔-axis at the point of abscissa 𝜔𝑟 .

Measuring an angle 𝜑𝑗 , the modal damping ratio is calculated from (6.102).


1
The diameter of the modal mobility circle is , hence the modal mass can be
𝑐𝑟
determined as
𝑐
𝑚𝑟 = 2 𝜁 𝑟𝜔 .
𝑟 𝑟

The method provides good results only for linear structures. The
distribution of the parameter 𝜔 along the circle is highly sensitive to non-linearities
(see Ch.7) so it would appear more judicious to operate at constant amplitude, as in
the method of forces in quadrature [128].

6.2.2.5. Method of complex power supplied to the structure

The method of complex power, reported in 1969 by Bonneau [17], has been
developed for the determination of modal parameters of linear structures with
nonproportional viscous damping described by a symmetric matrix.
The complex power supplied to a structure has the expression
1 𝑇 1 1
𝒫 = 2 {𝑓̂} {𝑞̇̃ } = 2 ∑𝑁 ̂̃ 𝑁 ̂
𝑗=1 𝑓𝑗 𝑞̇ 𝑗 = 2 ∑𝑗=1 𝑓𝑗 (𝑞̇ 𝑗𝑅 + 𝑖 𝑞̇ 𝑗𝐼 ). (6.103)

217
For harmonic excitation, using transformation (3.16), we obtain
{𝑞̇̃ } = [𝛹] {𝑝̇̃} = 𝑖𝜔 [𝛹] {𝑝̃}, (6.104)
hence
𝜔 𝑇
𝒫 = 𝑖 2 {𝑓̂} [𝛹] {𝑝̃}. (6.105)
Equation (6.2) can be written under the form

𝑖 [ 𝜔[𝑐] + 𝑖 [`𝑚ˏ] (𝜔2 [𝐼] − [`𝜔𝑟2 ˏ])] {𝑝̃} = [𝛹]𝑇 {𝑓̂} (6.106)
or
𝑖 [𝐺] {𝑝̃} = [𝛹]𝑇 {𝑓̂} (6.107)
where
[𝐺] = 𝜔 [𝑐] + 𝑖 [`𝑚ˏ] (𝜔2 [𝐼] − [`𝜔𝑟2 ˏ]). (6.108)
In order to isolate the 𝑟-th undamped natural mode at the frequency 𝜔𝑟
using in-phase real forces {𝑓̂}, the condition (6.32) should be satisfied.
Substituting the vector of appropriated modal forces (6.32) into equation
(6.107), we obtain
𝑖 {𝑝̃} = [𝐺]−1 [𝛹]𝑇 {𝑓̂} = [𝐺]−1 {𝑐}𝑟 𝜔𝑟 , (6.109)
where {𝑐}𝑟 is the 𝑟-th column of the nondiagonal modal damping matrix [𝑐].
Due to the symmetry of the matrix [𝑐], equation (6.32) yields
𝑇
̂} ]𝑇 = 𝜔𝑟 {𝑐}𝑇𝑟 = 𝜔𝑟 {𝑐}𝑟 ,
{𝑓̂} [𝛹] = [ [𝛹]𝑇 {𝑓 (6.110)
where {𝑐}𝑟 is the 𝑟-th row of the matrix [𝑐].
Substituting (6.109) and (6.110) into (6.105) we obtain
𝜔𝜔𝑟2
𝒫= {𝑐}𝑟 [𝐺]−1 {𝑐}𝑟 . (6.111)
2
Because of interest is the evaluation of the complex power supplied to the
structure
𝒫(𝑖𝜔) = 𝒫𝑅 (𝜔) + 𝑖 𝒫𝐼 (𝜔) (6.112)
for frequencies in the neighborhood of the resonance frequency 𝜔𝑟 , the matrix
[𝐺]−1 is calculated by a series expansion around the frequency 𝜔 = 𝜔𝑟 .
Let denote
[𝐺(𝜔)] = [𝐺(𝜔𝑟 )] + [∆𝐺(𝜀)] (6.113)
where
𝜀 = 𝜔 − 𝜔𝑟 (6.114)
and
[𝐺(𝜔𝑟 )]−1 = [𝐽]. (6.115)

218
We obtain
[𝐺(𝜔𝑟 )] = 𝜔𝑟 [𝑐] + 𝑖 [`𝑚ˏ] (𝜔𝑟2 [𝐼] − [`𝜔𝑟2 ˏ]), (6.116)
[∆𝐺(𝜀)] = 𝜀 [𝑐] + 𝑖 [`𝑚ˏ] (2𝜔𝑟 𝜀 + 𝜀 2 ), (6.117)
hence the series expansion will have the form [17]
[𝐺]−1 = [𝐽] − [𝐽] [∆𝐺] [𝐽] + [𝐽] [∆𝐺] [𝐽] [∆𝐺] [𝐽]− . .. (6.118)
Substituting (6.118) into (6.111) and having in view the following
equations which result from notations (6.112) to (6.117)
[𝐽] {𝑐}𝑟 𝜔𝑟 = {𝐼}𝑟 ,

𝜔𝑟 {𝑐}𝑇𝑟 [𝐽] = {𝐼}𝑇𝑟 ,


{𝐼}𝑇𝑟 {𝐼}𝑟 = 1 ,
[∆𝐺] {𝐼}𝑟 = {∆𝐺}𝑟 = 𝜀 {𝑐}𝑟 + 𝑖 {𝐼}𝑟 𝑚𝑟 (2𝜔𝑟 𝜀 + 𝜀 2 ) ,

{𝐼}𝑇𝑟 [∆𝐺] = {∆𝐺}𝑇𝑟 = 𝜀 {𝑐}𝑇𝑟 + 𝑖 𝑚𝑟 (2𝜔𝑟 𝜀 + 𝜀 2 ) {𝐼}𝑇𝑟 ,


{𝐼}𝑇𝑟 [∆𝐺]{𝐼}𝑟 = ∆𝐺𝑟𝑟 = 𝜀 𝑐𝑟𝑟 + 𝑖 𝑚𝑟 (2𝜔𝑟 𝜀 + 𝜀 2 ) ,
and denoting
𝑐
𝜁𝑟𝑟 = 2 𝑚𝑟𝑟𝜔
𝑟 𝑟
we obtain the following expression for the Taylor series expansion of the complex
power [17] :
(𝜔−𝜔𝑟 )2 𝑖
𝒫(𝑖𝜔) = 𝑘𝑟 [𝜁𝑟𝑟 𝜔𝑟 − 𝑖 (𝜔 − 𝜔𝑟 ) − 2!
(4𝑚𝑟 𝜔𝑟 𝐽𝑟𝑟 − 𝜔 ) +
𝑟
(𝜔−𝜔𝑟 )3
+ ( . . . .) + . . . ] (6.119)
3!

where
𝐽𝑟𝑟 = {𝐼}𝑇𝑟 [𝐽]{𝐼}𝑟
is a complex quantity.
At the phase resonance of the 𝑟-th mode, from (6.112) and (6.119) we
obtain
𝒫𝑅 (𝜔𝑟 ) = 𝜁𝑟𝑟 𝜔𝑟 𝑘𝑟 , (6.120)
𝒫𝐼 (𝜔𝑟 ) = 0 ,
𝑑𝒫
( 𝑑𝜔𝑅) =0,
𝜔=𝜔𝑟
𝑑𝒫
( 𝑑𝜔𝐼) = −𝑘𝑟 , (6.121)
𝜔=𝜔𝑟

219
𝑑2 𝒫
( 𝑑𝜔2𝑅) = −4 𝑚𝑟2 𝜔𝑟3 ℛℯ(𝐽𝑟𝑟 ), (6.122)
𝜔=𝜔𝑟
𝑑2 𝒫
( 𝑑𝜔2𝐼) = 𝑚𝑟 𝜔𝑟 − 4 𝑚𝑟2 𝜔𝑟3 ℐ𝑚(𝐽𝑟𝑟 ). (6.123)
𝜔=𝜔𝑟

During testing, the curves 𝒫𝑅 (𝜔) and 𝒫𝐼 (𝜔) are drawn in the vicinity of
the phase resonance (Fig.6.8) keeping constant the appropriated excitation forces.
The modal parameters are calculated as
follows
𝑑𝒫
𝑘𝑟 = − ( 𝑑𝜔𝐼) = 𝑡𝑎𝑛 𝜃, (6.124)
𝜔=𝜔𝑟
𝑘
𝑚𝑟 = 𝜔𝑟2 , (6.125)
𝑟

𝒫𝑅 (𝜔𝑟 )
𝜁𝑟𝑟 = . (6.126)
𝜔 𝑟 𝑘𝑟

The diagonal elements of the modal


damping matrix are
2
𝑐𝑟𝑟 = 𝜔2 𝒫𝑅 (𝜔𝑟 ). (6.127)
𝑟

The nature of the extremum of the


function 𝒫𝑅 (𝜔) is theoretically dependent on the
sign of the quantity ℛℯ(𝐽𝑟𝑟 ). Practically, it corresponds to a maximum of the local
velocities, hence to a maximum of the active power 𝒫𝑅 .
This procedure offers several advantages which make it preferred to the
methods presented above [7]. The values 𝑚𝑟 and 𝜁𝑟𝑟 are obtained separately from
two pieces of information. It provides a criterion for the quality of appropriation:
the maximum value of 𝒫𝑅 and the zero of 𝒫𝐼 should correspond to the same abscissa
𝜔𝑟 . It is applicable to weakly non-linear systems and when the mode isolation is
not quite perfect.
Other identification methods of structures with nonproportional damping
are presented in the papers by Wittmeyer [139] and [140].

6.3. MULTI-POINT SUCCESSIVE EXCITATION METHODS

This section is devoted to the presentation of identification methods


using multi-point excitation, but considered “single-point excitation methods”
due to the use of only one exciter which is moved successively at various
points on the structure. These methods are based on experimental data
processed after testing by calculations which require the use of computers.
Their practical application was limited to relatively simple models due to the
inversion of matrices with generally ill-conditioned experimentally
determined elements

220
6.3.1. Method of Independent Forces

In 1961, Traill-Nash [118] presented a combined analytical/experimental


method for the determination of the dynamic properties of a linear system with
nonproportional damping. Its presentation requires a reformulation of the equations
developed in § 3.2.3, introducing the matrix of complex receptances (dynamic
influence coefficients).
For harmonic conditions, the excitation-response relation (3.38) can be
written under the form
{𝑞̃} = [𝛼̅] {𝑓̂} = [𝛼𝑅 + 𝑖 𝛼𝐼 ] {𝑓̂} . (6.128)
Let {𝑓̂} = {ℱ̂ } be the forcing which excites the monophase response
{𝑞̃} = {𝑞̂} 𝑒 −𝑖𝜑 , (6.129)
where 𝜑 is the phase lag between excitation and response.
Substituting (6.129) into (6.128) and separating the real and imaginary
parts we obtain
{𝑞̂} 𝑐𝑜𝑠 𝜑 = [𝛼𝑅 ] {ℱ̂ } , (6.130)
−{𝑞̂} 𝑠𝑖𝑛 𝜑 = [𝛼𝐼 ] {ℱ̂ } . (6.131)
Multiplying (6130) by 𝑠𝑖𝑛 𝜑 and (6.131) by 𝑐𝑜𝑠 𝜑 and summing up we
obtain
𝑠𝑖𝑛 𝜑 [𝛼𝑅 ] {ℱ̂ } + 𝑐𝑜𝑠 𝜑 [𝛼𝐼 ] {ℱ̂ } = {0}, (6.132)
or
([𝛼𝑅 ] 𝑡𝑎𝑛𝜑 + [𝛼𝐼 ]) {ℱ̂ } = {0}. (6.133)
Multiplying (6130) by 𝑐𝑜𝑠 𝜑 and (6.131) by (−𝑠𝑖𝑛 𝜑) and summing up we
obtain
{𝑞̂} = ([𝛼𝑅 ] 𝑐𝑜𝑠 𝜑 − [𝛼𝐼 ] 𝑠𝑖𝑛 𝜑) {ℱ̂ }. (6.134)
When 𝜔 = 𝜔𝑟 , 𝑐𝑜𝑠 𝜑𝑟 = 0 , and equations (6.132) and (6.134) take the
form
[𝛼𝑅 ] {ℱ̂ } = {0} , (6.135)
− [𝛼𝐼 ] {ℱ̂ } = {𝑞̂}. (6.136)
Therefore, the natural frequencies 𝜔𝑟 can be determined from (6.135) as
the frequencies for which the determinant of the real part of the complex receptance
matrix vanishes, 𝑑𝑒𝑡 [𝛼𝑅 ] = 0 , which corresponds to Asher’s method.
The method of Traill-Nash is based on the experimental measurement of
the matrix of complex receptances [𝛼̅]. Using only one exciter, placed successively
at 𝑁 points on the structure, the response is measured at the driving points.
Let 𝑓̂ℓ be the force which applied at point ℓ excites the response
{𝑞𝑅 (𝜔)}ℓ + 𝑖 {𝑞𝐼 (𝜔)}ℓ .

221
Next, the following square matrices are constructed, which have the real
and imaginary parts of the measured response as columns
[𝑞𝑅 (𝜔)] = [{𝑞𝑅 (𝜔)}1 {𝑞𝑅 (𝜔)}2 . .. {𝑞𝑅 (𝜔)}ℓ . . . {𝑞𝑅 (𝜔)}𝑁 ], (6.137)

[𝑞𝐼 (𝜔)] = [{𝑞𝐼 (𝜔)}1 {𝑞𝐼 (𝜔)}2 . .. {𝑞𝐼 (𝜔)}ℓ . . . {𝑞𝐼 (𝜔)}𝑁 ], (6.138)
as well as the diagonal matrix
[`𝑓̂ˏ] = 𝑑𝑖𝑎𝑔 [𝑓̂ℓ ]. (6.139)
Equation (3.128) becomes
[𝑞𝑅 (𝜔)] + 𝑖 [𝑞𝐼 (𝜔)] = [𝛼𝑅 + 𝑖 𝛼𝐼 ] [`𝑓̂ˏ], (6.140)
in which the elements of the matrices
−1 1
[𝛼𝑅 (𝜔)] = [𝑞𝑅 (𝜔)] [`𝑓̂ˏ] = [𝑞𝑅 (𝜔)] [` ̂ ˏ], (6.141)
𝑓
−1 1
[𝛼𝐼 (𝜔)] = [𝑞𝐼 (𝜔)] [`𝑓̂ˏ] = [𝑞𝐼 (𝜔)] [` ̂ ˏ], (6.142)
𝑓

are determined experimentally at the excitation frequency 𝜔.


Equation (6.133) represents a problem of latent roots of a matrix pencil in
which the latent roots are 𝑡𝑎𝑛 𝜑. The condition to have nontrivial solutions is
𝑑𝑒𝑡 [ 𝑡𝑎𝑛𝜑 [𝛼𝑅 ] + [𝛼𝐼 ] ] = {0} (6.143)
The 𝑁 roots 𝑡𝑎𝑛 𝜑𝑟 of the equation (6.143) determine the values of the 𝑁
characteristic phase lags 𝜑𝑟 at the frequency 𝜔. Repeating the measurements at
various excitation frequencies, the variation of characteristic phase lags with
frequency is represented graphically as in Fig.6.9,a. for two such values. The
natural frequencies are the abscissae of the points where the curves intersect the
horizontal line of ordinate 𝜑 = 90°.
The procedure continues with the calculation of the modal monophase
excitation vectors {ℱ̂ (𝑟) } from equation (6.133) and the modal monophase response
vectors {𝛷 (𝑟) } from (6.134), at various frequencies. The graphs of the monophase
response vectors versus frequency are shown in Figs.6.9,b and c, and the graphs of
the monophase excitation vectors are depicted in Figs.6.9,d and e.
In particular, the shape of the “classical” principal mode of vibration {𝛹 (𝑟) }
(𝑟)
is determined from the ordinates of the intersection points of the curves 𝛷𝑗 with
the vertical of abscissa 𝜔𝑟 (Fig.6.9,c).
The method has been applied to a system with two degrees of freedom
[119], and found to be more accurate than the method of Kennedy and Pancu [61]
in the case of close natural frequencies.
The identification of the system spatial parameters is based on the elements
of the complex receptance matrix, measured at two frequencies 𝜔 ̅𝑎 and 𝜔̅𝑏 . For this

222
it is necessary to invert the receptance matrix in order to calculate the matrix of
complex dynamic stiffnesses
[𝛽̅ ] = [𝛽𝑅 ] + 𝑖 [𝛽𝐼 ] = [𝛼̅]−1 . (6.144)

In view of equations (3.89), the components of the matrix of dynamic


stiffnesses are
[𝛽𝑅 ] = [𝐾] − 𝜔2 [𝑀], (6.145)
[𝛽𝐼 ] = 𝜔 [𝐶] + [𝐻]. (6.146)
Calculating matrices (6.145) and (6.146) at two frequencies 𝜔
̅𝑎 and 𝜔
̅𝑏 we
obtain
[𝛽𝑅 (𝜔
̅ 𝑎 )]−[𝛽 (𝜔
̅
𝑅 𝑏 )]
[𝑀] = 2 ̅2 , (6.147)
𝜔𝑏 −𝜔𝑎
̅

̅2
𝜔𝑏 [𝛽𝑅 (𝜔 ̅2
̅ 𝑎 )]−𝜔𝑎 [𝛽𝑅 (𝜔
̅ 𝑏 )]
[𝐾] =
̅2 ̅2
, (6.148)
𝜔 𝑏 −𝜔𝑎
[𝛽𝐼 (𝜔
̅ 𝑎 )]−[𝛽 (𝜔
̅
𝐼 𝑏 )]
[𝐶] =
𝜔̅ 𝑎 −𝜔̅𝑏
, (6.149)

223
𝜔
̅ 𝑎 [𝛽 (𝜔
𝐼 𝑏 )]−𝜔𝑏 [𝛽𝐼 (𝜔𝑎 )]
̅ ̅ ̅
[𝐻] =
𝜔𝑎 −𝜔𝑏
̅ ̅
. (6.150)

For pure hysteretic damping [𝐶] = [0] and [𝐻] = [𝛽𝐼 ]. For pure viscous
1
damping [𝐻] = [0] and [𝐶] = 𝜔 [𝛽𝐼 ].
Nissim [83] recommends to calculate first the matrices (6.147) and (6.151)
and then the natural frequencies 𝜔𝑟 and the mode shapes. Because the characteristic
phase lags 𝜑𝑟 are usually small, the computation of 𝜔𝑟 ’s using the method of Traill-
Nash can result in large errors.
The matrix inversion (6.144) can introduce at its turn large errors when the
matrix [𝛼̅] is ill conditioned, so that the method can be applied with good results
only to systems with a small number of degrees of freedom and with relatively close
natural frequencies.
Cottin and Dellinger [26] have published an identification method based on the
theory of complex modes of vibration of systems with viscous damping (§ 3.2.4.1).
The direct complex receptance at the driving point ℓ of a structure with
nonproportional damping is given by expression (3.186)
(𝑟) (𝑟)∗
𝛿 𝛿
𝛼̅ℓℓ = ∑𝑁 ℓℓ
𝑟=1 (𝑖𝜔−𝜎 + 𝑖𝜔−𝜎
ℓℓ
∗ ). (6.151)
𝑟 𝑟

This receptance is measured at 2𝑁 excitation frequencies and the test data


are grouped into two column vectors
{𝛼̅Ⅰ } = { 𝛼̅ℓℓ (𝜔
̅1 ) 𝛼̅ℓℓ (𝜔 ̅ 𝑁 ) }𝑇 ,
̅2 ) . . . 𝛼̅ℓℓ (𝜔
(6.152)
{𝛼̅Ⅱ } = { 𝛼̅ℓℓ (𝜔
̅𝑁+1 ) 𝛼̅ℓℓ (𝜔 ̅2𝑁 ) }𝑇 .
̅𝑁+2 ) . . . 𝛼̅ℓℓ (𝜔
If the same receptances are calculated from expression (6.151), we obtain
2𝑁 equations which can be written in condensed matrix form as
{𝛼̅Ⅰ } [𝑁 ] [𝑄Ⅰ ] {𝛿}
{ }=[ Ⅰ ]{ } (6.153)
{𝛼̅Ⅱ } [𝑁Ⅱ ] [𝑄Ⅱ ] {𝛿 ∗ }
where
(1) (2) (𝑁) 𝑇 (1)∗ (2)∗ (𝑁)∗ 𝑇
{𝛿} = {𝛿ℓℓ 𝛿ℓℓ . . . 𝛿ℓℓ } , {𝛿 ∗ } = {𝛿ℓℓ 𝛿ℓℓ . . . 𝛿ℓℓ } ,
and
1 1 1 1
𝑖𝜔
̅ 1 −𝜎1
… 𝑖𝜔
̅ 1 −𝜎𝑁 ̅ 1 −𝜎∗1
𝑖𝜔
… ̅ 1 −𝜎∗𝑁
𝑖𝜔
[𝑁Ⅰ ] = ⋮ ⋱ ⋮ , [𝑄Ⅰ ] = ⋮ ⋱ ⋮
1 1 1 1
[𝑖𝜔
̅ 𝑁 −𝜎1
… ̅ 𝑁 −𝜎𝑁 ]
𝑖𝜔 ̅ 𝑁 −𝜎∗1
[𝑖𝜔
… ̅ 𝑁 −𝜎∗𝑁 ]
𝑖𝜔

224
1 1 1 1
𝑖𝜔
̅ 𝑁+1 −𝜎1
… 𝑖𝜔
̅ 𝑁+1 −𝜎𝑁 ̅ 𝑁+1 −𝜎∗1
𝑖𝜔
… ̅ 𝑁+1 −𝜎∗𝑁
𝑖𝜔
[𝑁Ⅱ ] = ⋮ ⋱ ⋮ , [𝑄Ⅱ ] = ⋮ ⋱ ⋮ .
1 1 1 1
[ 𝑖𝜔
̅ 2𝑁 −𝜎1
… 𝑖𝜔
̅ 2𝑁 −𝜎𝑁 ] [ ̅ 2𝑁 −𝜎∗1
𝑖𝜔
… ̅ 2𝑁 −𝜎∗𝑁
𝑖𝜔 ]
The set (6.153) contains 2𝑁 equations having 𝑁 unknowns 𝜔𝑟 and 𝑁
(𝑟)
unknowns 𝛿ℓℓ . For the numerical solution of (6.153) it is useful to write it under
the form
{𝛼̅Ⅰ } [𝑁 ] [𝑄Ⅰ ] {𝛿}
{ ∗ } = [ ∗Ⅰ ]{ } (6.154)
{𝛼̅Ⅱ } [𝑄Ⅱ ] [𝑁Ⅱ∗ ] {𝛿 ∗ }
∗ }
where the star superscript denotes the complex conjugate. Using the vector {𝛼̅𝙸𝙸
instead of its conjugate, a better conditioning is obtained for the square matrix in
equation (6.154).
The 𝑁 unknowns 𝜎𝑟 are calculated from the zeros of the function
𝐹(𝜎) = {𝛿} − {𝛿 ∗ }∗ = {0}, (6.155)
where {𝛿 ∗ }∗ is the complex conjugate of {𝛿 ∗ }. Good results are obtained using the
(𝑟)
iterative method of Krawczyk [65] which provides also 𝛿ℓℓ .
(𝑟)
If in equation (3.185) we normalize the coefficients 𝛿ℓℓ taking 𝑢𝑟 = 1, we
(𝑟)
obtain the values 𝑞ℓ .
The modal vectors {𝑞 (𝑟) } can be determined repeating the measurements
for all ℓ points on the structure. Then, from (3.162) we obtain the vectors {𝜉 (𝑟) },
hence the modal matrix [𝜉].
From equation (3.161) we can calculate 𝑣𝑟 . From (3.169) and (3.170) we
obtain the matrices [𝑈] and [𝑉], then the dynamic matrix [𝐷] (3.156). If [𝑀] is
known, from (3.150) we obtain [𝐶] and [𝐾].

6.3.2. Method of Independent Additional Loads

A variant of the method of independent forces was published in 1970 by


Feix [34] and applied by Ratjen and Mölenkamp [99] to a laboratory model.
Consider a proportionally damped linear system with a reduced number of
degrees of freedom, excited by a single harmonic force which might be unknown.
The steady-state harmonic motion of the system is governed by equation
(3.38)
(−𝜔2 [𝑀] + 𝑖𝜔[𝐶] + [𝐾]) {𝑞̃} = {𝑓̂} (6.156)
in which the vector {𝑓̂} has only one element.
If additional masses [∆𝑀]ℓ are attached at the driving points maintaining
the excitation force constant, they produce an additional response {∆𝑞̃}. Equation
(6.156) becomes

225
(−𝜔2 [𝑀 + ∆𝑀ℓ ] + 𝑖𝜔[𝐶] + [𝐾]) {𝑞̃ + ∆𝑞̃ℓ } = {𝑓̂}. (6.157)
The left hand term due to the additional masses can be shifted to the right
hand side as an additional excitation vector
{∆𝑓̂}ℓ = 𝜔2 [∆𝑀]ℓ {𝑞̃ + ∆𝑞̃ℓ } (6.158)
which results in
(−𝜔2 [𝑀] + 𝑖𝜔[𝐶] + [𝐾]){𝑞̃ + ∆𝑞̃ℓ } = {𝑓̂ + ∆𝑓̂ℓ } (6.159)
Subtracting equations (6.151) and (6.154) we obtain
(−𝜔2 [𝑀] + 𝑖𝜔[𝐶] + [𝐾]){∆𝑞̃}ℓ = {∆𝑓̂} (6.160)

For ℓ = 1, 2, . . , 𝑁 distributions of additional masses [∆𝑀]ℓ chosen to


obtain 𝑁 linearly independent force distributions {∆𝑓̂}ℓ , we can construct the
square matrices
[∆𝑞̃] = [{∆𝑞̃}1 {∆𝑞̃}2 . . . {∆𝑞̃}ℓ . . . . {∆𝑞̃}𝑁 ], (6.161)

[∆𝑓̂] = [{∆𝑓̂}1 {∆𝑓̂}2 . . . {∆𝑓̂}ℓ . . . . {∆𝑓̂}𝑁 ], (6.162)

hence the 𝑁 equations (6.160) can be written


(−𝜔2 [𝑀] + 𝑖𝜔[𝐶] + [𝐾])[∆𝑞̃] = [∆𝑓̂]. (6.163)
Using the matrix of complex dynamic stiffnesses [𝛽̅ ] , equation (6.163)
becomes
[𝛽̅ ] [∆𝑞̃] = [∆𝑓̂]. (6.164)
wherefrom we obtain
[𝛽̅ ] = [∆𝑓̂] [∆𝑞̃]−1. (6.165)
Evaluating the matrix [𝛽̅ ] at two frequencies 𝜔 ̅𝑎 and 𝜔 ̅𝑏 , the system
matrices can be calculated from expressions (6.147) to (6.150). The two frequencies
are selected in the vicinity of natural frequencies and not higher than 𝜔𝑁 . If the
excitation forces are known, it is possible to use only 𝑁 − 1 distributions of
additional masses.
Like the method of independent forces, the method of independent
additional loads is adequate for the laboratory study of the dynamic response of
simple small scale models and lumped parameter models with a reduced number of
degrees of freedom.
An improvement of the above method has been reported by Feix [35].

226
CHAPTER 7

Parameter Identification of Simple Nonlinear


Systems
It is known that experimentally plotted vector diagrams for aircraft
structures may become distorted if resonance tests at relatively large response
amplitudes are carried out. De Vries [131] showed that the frequency shift along
the polar diagrams of mobility can be explained by the presence of nonlinear
stiffness characteristics, while the distortion of the circular shape is due to nonlinear
damping.
Generally, it is assumed that small nonlinearities do not affect the basic
principles of modal analysis and the linear summation of natural modes, while the
usual criteria of resonance location (more exactly – of the principal “resonance”)
cannot be used.
For systems with well-spaced natural frequencies and weak nonlinearities,
it can be supposed, as a first approximation, that in the vicinity of “resonance” the
dynamic response can be described by the equation of motion of a single-degree-
of-freedom system with (equivalent) hysteretic damping and cubic stiffness
characteristic, of the form
𝑓𝑒 (𝑥) = 𝑘 (𝑥 + 𝜇 𝑥 3 ),
where 𝜇 is a coefficient of nonlinearity.
Based on the analysis of the frequency response of this type of nonlinear
oscillator, we developed several methods for the parameter identification of weakly
nonlinear systems applied to the determination of the dynamic properties of some
materials used in vibration isolation applications [92, 94, 97].

7.1. HARMONIC EXCITATION METHODS

7.1.1. Polar Plot of the Frequency Response

The Duffing-type equation of motion of an oscillator with hysteretic


damping and cubic stiffness, excited by a harmonic force 𝑓(𝑡) = 𝑓̂ 𝑐𝑜𝑠 𝜔𝑡, may be
written as

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘 (𝑥 + 𝜇 𝑥 3 ) = 𝑓̂ 𝑐𝑜𝑠 𝜔𝑡, (7.1)
where 𝑘 is the slope of the elastic characteristic at the origin and 𝜇 is a coefficient
of nonlinearity (positive for hardening systems).
A first harmonic approximation of the response is chosen of the form

227
𝑥(𝑡) = 𝑥̂1 𝑐𝑜𝑠 (𝜔𝑡 + 𝜑). (7.2)
Using the method of harmonic linearization, the higher harmonic terms are
3
neglected and it is considered that 𝑥 3 ≅ 𝑥̂12 𝑥.
4

Substitution of (7.2) into (7.1) yields


3 1
𝛺2 = 1 + 4 𝛾 𝑎2 ∓ √𝑎2 − 𝑔2 , (7.3)
𝑔 𝑔
𝜑 = 𝑡𝑎𝑛−1 3 = 𝑡𝑎𝑛−1 , (7.4)
𝛺2 −1− 𝛾𝑎2 1
∓√ 2 −𝑔2
4
𝑎
where
ℎ 𝜔 𝑘 𝑓̂2 𝑘
𝑔= , 𝛺= , 𝜔𝑛 = √ , 𝛾=𝜇 , 𝑎 = 𝑥̂1 ̂ . (7.5)
𝑘 𝜔𝑛 𝑚 𝑘2 𝑓

It should be noticed that the function 𝑎 defined by (7.3) is not a genuine


dimensionless receptance, because 𝑥̂1 is not the response amplitude (which is
periodic, but not sinusoidal). Due to the first harmonic approximation, 𝑎 becomes
a “describing function” of the response, which is a function of both ω and 𝑓̂,
because there is not cause-effect proportionality, as for linear systems. The transfer
locus of the nonlinear system is no more a single curve with frequency graduations,
but a family of curves having as parameter the force amplitude 𝑓̂ .
Consider the dimensionless complex function
𝑎̃ = 𝑎𝑅 + 𝑖 𝑎𝐼 = 𝑎 𝑐𝑜𝑠𝜑 + 𝑖 𝑎 𝑠𝑖𝑛𝜑. (7.6)
The vector components are
3
𝑎𝑅 = (1 + 4 𝛾 𝑎2 − 𝛺2 ) 𝑎2 , (7.7)

𝑎𝐼 = −𝑔 𝑎2 , (7.8)
where 𝑎 is defined by (7.3).
Eliminating 𝑎 and 𝛺2 among (7.3), (7.7) and (7.8) we obtain the equation
of a circle
1 2 1
𝑎𝑅2 + (𝑎𝐼 + ) = , (7.9)
2𝑔 4𝑔2

which is the locus of the tip of the vector 𝑎̃ in the Argand plane. Equation (7.9) is
identical to equation (2.37) for linear systems (𝜇 = 0); hence, for the same values
𝑓̂⁄𝑘 and 𝑔 the polar diagrams will have the same shape.
For comparison, Fig.7.1 shows the polar plots of 𝑎̃, for 𝑔 = 0.04, 𝛾 = 0
and 𝛾 = 10−4, graduated with the values of the dimensionless frequency 𝛺. For the
nonlinear system (Fig.7.1,b) a frequency shift along the circle can be noticed with
respect to the linear system (Fig.7.1,a), anticlockwise for a hardening system. The

228
ratio ∆𝑠⁄∆𝜔 has not a maximum at 𝜑 = −90°, hence the Kennedy and Pancu
criterion cannot be applied.

Figure 7.2 shows the significant points of the two diagrams with frequency
graduations. For the linear system (Fig.7.2,a) the point 𝑀 defines the resonance
frequency. For nonlinear systems (Fig.7.2,) the frequency of the crossing point 𝑀
with the negative imaginary semiaxis is
3 𝛾
𝜔𝑟 = 𝜔𝑛 √1 + 4 𝑔2 ≠ 𝜔𝑛 . (7.10)

The frequencies of the points where the diameter 𝐵𝐶, perpendicular to 𝑂𝑀,
intersects the circle are

′ 3 𝛾
𝜔1,2 = 𝜔𝑛 √1 ∓ 𝑔 + 8 𝑔2 ≠ 𝜔1,2. (7.11)

229
For systems with a hardening characteristic (Fig.7.2,b), 𝜇 > 0, hence 𝛾 >
0 , therefore 𝜔𝑟 > 𝜔𝑛 , 𝜔1′ > 𝜔1 , 𝜔2′ > 𝜔2 and the frequencies are “shifted”
anticlockwise along the circle.
For systems with softening characteristic, 𝜇 < 0, hence 𝛾 < 0 , therefore
𝜔𝑟 < 𝜔𝑛 , 𝜔1′ < 𝜔1 , 𝜔2′ < 𝜔2 and the frequency shift is clockwise.
The frequency shift on the polar plots of the dimensionless complex
function 𝑎̃ is more evident if several polar diagrams are drawn on the same graph,
for different values of the amplitude of the excitation force.

In figure 7.3 the constant-frequency lines (isochrones) are drawn by dashed


lines. For linear systems (Fig.7.3,a) these are straight lines radiating from the
origin. For hardening systems (Fig.7.3,b) they are bent towards low frequencies,
and for softening systems (Fig.7.3,c) – towards higher frequencies.
The distortion of the polar plots provides an indication of the nonlinearity
of the damping characteristic. For linear damping, the polar plots are circles. The
Coulomb damping tends to elongate the polar plots in the direction of the negative
imaginary semiaxis, while the quadratic damping (proportional to the velocity
squared) tends to elongate the polar plots in the direction of the real axis,
“flattening” the otherwise circular diagrams.

7.1.2. Jump Phenomena

For nonlinear systems it is convenient to plot response displacement


components and not receptances. Each point represented in the Argand plane is
defined by two parameters: the excitation frequency and the amplitude of the input
force. Consequently, two sets of response loci can be drawn: Nyquist plots –
connecting points of constant excitation level and isochrones – connecting constant
frequency points.
If the polar plots of the displacement
𝑥̃1 = 𝑥1𝑅 + 𝑖 𝑥1𝐼 = 𝑥̂1 𝑐𝑜𝑠𝜑 + 𝑖 𝑥̂1 𝑠𝑖𝑛𝜑

230
are drawn for different values 𝑓̂⁄𝑘 (Fig.7.4), as 𝑓̂ grows the isochrones are so much
bent that they become tangent to the Nyquist plots.

The locus of the tangency points of Nyquist plots with the isochrones (solid
thick line) is a “stability boundary”, which separates the range of stable vibrations
from the range of unstable vibrations, defined by the equations
3 9
𝛺2 = 1 + 2 𝜇 𝑥̂12 ± √16 𝜇2 𝑥̂14 − 𝑔2 , (7.12)

1 3 9
𝜑 = 𝑡𝑎𝑛−1
𝑔
(4 𝜇 𝑥̂12 ± √16 𝜇2 𝑥̂14 − 𝑔2 ). (7.13)

The stability boundary 𝑋𝐿𝑌, plotted for a given value 𝜇, is a hyperbola


symmetrical with respect to the line 𝜑 = −135° and tangent to the circle with the
4𝑔
center at origin and radius 𝑥̂1𝐿 = √3𝜇 which represents the lowest displacement
amplitude at which jump phenomena can occur. It is tangent at point 𝐾 to the polar
𝑓̂ 32 𝑔3
plot of parameter =√ and to the isochrone of parameter 𝛺𝐾 = √1 + √3𝑔
𝑘 9√3𝜇
and corresponds to a phase angle 𝜃𝐾 = −120°.

231
32 𝑔3
For forces 𝑓̂ > 𝑘√ and for frequencies 𝜔 > 𝜔𝑛 √1 + √3𝑔 the
9√3𝜇
stability boundary crosses the response circles and the isochrones, so that jump
phenomena may occur. The points located on the arcs of curve between these
crossing points define unstable vibrations.
Based on the diagrams from Fig.7.4 it is possible to explain the two types
of discontinuities which might occur in the variation of the amplitude and phase of
the harmonic vibration when measurements are carried out with frequency sweep
(at 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡.) or with variable force amplitude (for 𝜔 = 𝑐𝑜𝑛𝑠𝑡.).
If 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡. and the frequency ω is gradually increased from rest, the
tip of the response vector follows the portion 𝑁𝐹 of the respective Nyquist plot
until the stability boundary is reached at point 𝐶, where the circle 𝑓̂ = 𝑐𝑜𝑛𝑠𝑡.
is tangent to the isochrone 𝜔′′ = 𝑐𝑜𝑛𝑠𝑡. Then, there is a jump in amplitude and
phase from 𝐶 to 𝐷, along the isochrone 𝜔′′ = 𝑐𝑜𝑛𝑠𝑡., followed by a continuous
change in amplitude and phase along the arc 𝐷𝑂 of the polar plot.
When the frequency is decreased, the tip of the response vector 𝑥̃1 moves
anticlockwise along 𝑂𝐷𝐸 until the point of tangency with the isochrone 𝜔′ =
𝑐𝑜𝑛𝑠𝑡., and jumps along the curve 𝜔′ = 𝑐𝑜𝑛𝑠𝑡. from 𝐸 to 𝐹, where the isochrone
crosses the circle, then moves along the circle until point 𝑁 where 𝜔 = 0.

The arcs 𝑁𝐹 and 𝐷𝑂 of the polar plot define stable regimes of vibration,
the arcs 𝐹𝐶 and 𝐸𝐷 define conditionally stable regimes of vibration, while the arc
𝐶𝐸 defines unstable regimes of vibration. Generally, jumps do not occur suddenly.
The response amplitude and phase vary with finite speed (function of system
damping and rate of change of 𝛺) during a transitory regime until a stable steady-
state regime is reached.

232
Figure 7.6 illustrates the jump phenomenon which occurs when the
amplitude of the exciting force is varied while the forcing frequency is constant.
When 𝑓̂ is gradually increased from zero, the tip of the displacement vector moves
along the isochrone 𝜔 = 𝑐𝑜𝑛𝑠𝑡. from 𝑂 to 𝑉 until the stability boundary is reached
at 𝑆, the point of tangency with the circle 𝑓̂ ′′ = 𝑐𝑜𝑛𝑠𝑡. At a subsequent increase of
the force, the amplitude and phase of the response have jumps to the values
corresponding to point 𝑇, where the isochrone intersects the circle 𝑓̂ ′′ = 𝑐𝑜𝑛𝑠𝑡.,
then vary continuously along 𝜔 = 𝑐𝑜𝑛𝑠𝑡.
On decreasing the force amplitude, the tip of the response vector 𝑥̃1 moves
along the curve 𝜔 = 𝑐𝑜𝑛𝑠𝑡. in opposite direction, through 𝑇 until 𝑈 on the stability
boundary, the point of tangency with the circle 𝑓̂ ′ = 𝑐𝑜𝑛𝑠𝑡. Then it jumps to the
point 𝑉, following the same response curve 𝑓̂ ′ = 𝑐𝑜𝑛𝑠𝑡., and then moves again
along the isochrone from 𝑉 to the origin 𝑂.
An understanding of the characteristics of jump phenomena allows the
completion of measured polar diagrams with the conditionally unstable parts and a
correct interpretation of some apparent anomalies which arise during the automatic
plotting of these diagrams (on an 𝑋-𝑌 plotter with frequency sweep).

7.1.3. Parameter Estimation

One method for determining the system dynamic parameters used by the
present author [92] requires at least two response circles (Fig.7.7) plotted for
different amplitudes 𝑓̂ ′ and 𝑓̂ ′′ of the harmonic excitation force.
The procedure starts by plotting several polar
plots of the displacement, for different values 𝑓̂ of the
amplitude of the excitation force, on which are
marked the points of maximum displacement
amplitude. For systems with proportional damping
and a perfect isolation of the examined mode of
vibration, these points correspond to a phase lag of
−90°. If the diagrams can be approximated by circles
and if the following relations can be established
𝑓̂′ 𝑓̂′′ 𝑓̂′′′
̅̅̅̅̅̅ = ̅̅̅̅̅̅̅ = ̅̅̅̅̅̅̅̅ =. . .
𝑂𝑀′ 𝑂𝑀 ′′ 𝑂𝑀 ′′′

then the model of hysteretic damping is valid.


Let 𝜔𝑟′ and 𝜔𝑟′′ be the frequencies of the
points 𝑀′ and 𝑀′′ of maximum response amplitude
on two diagrams plotted for excitation forces 𝑓̂ ′ and 𝑓̂′′ > 𝑓̂ ′, respectively
(Fig.7.7).
Based on equation (7.11), it follows that the diameter 𝐵′ 𝐶 ′ , perpendicular
to 𝑂 𝑀 , intersects the circle 𝑓̂ ′ = 𝑐𝑜𝑛𝑠𝑡. at the points of frequencies
′ ′

233
3 2
𝜔1′ = 𝜔𝑛 √1 − 𝑔 + 4 𝜇 ̅̅̅̅̅
𝑂𝐵′ ,

3 2
𝜔2′ = 𝜔𝑛 √1 + 𝑔 + 𝜇 ̅̅̅̅̅
𝑂𝐶 ′ .
4

Because ̅̅̅̅̅
𝑂𝐵′ = ̅̅̅̅̅
𝑂𝐶 ′
𝜔2′2 − 𝜔1′2 = 2 𝑔 𝜔𝑛2
wherefrom we obtain the equivalent hysteretic damping factor
𝜔′2
2 −𝜔1
′2
𝑔= 2 . (7.14)
2 𝜔𝑛
Denoting
̅̅̅̅̅̅
𝑂𝑀′ = 𝑥𝑟′ and ̅̅̅̅̅̅
𝑂𝑀′′ = 𝑥𝑟′′
equation (7.10) yields
3
𝜔𝑟′ = 𝜔𝑛 √1 + 4 𝜇𝑥𝑟′2 ,

3
𝜔𝑟′′ = 𝜔𝑛 √1 + 4 𝜇𝑥𝑟′′2 ,

hence
2
𝑥𝑟 ′
𝜔′2 ′′2
𝑟 −( ′′) 𝜔𝑟
𝑥𝑟
𝜔𝑛2 = 2 (7.15)

1−(𝑥′′𝑟)
𝑥𝑟
and the coefficient of nonlinearity
𝜔𝑟 ′ 2
1−( ′′ )
4 𝜔𝑟
𝜇= ′ 2 . (7.16)
3 𝜔𝑟
𝑥𝑟′′2 ( ′′ ) −𝑥𝑟′2
𝜔𝑟
Substituting (7.15) into (7.14) we obtain the damping factor 𝑔.
The slope of the elastic characteristic at the origin is
′ ′′
1 𝑓̂ 𝑓̂
𝑘=
𝑔 𝑥′𝑟
= 𝑔1 . (7.17)
𝑥′′𝑟
The other parameters can be determined as
𝑘
𝑚=
𝜔2𝑛
, ℎ = 𝑔𝑘.

The practical application of this method to the measurement of the dynamic


characteristics of polyurethane foams is presented in [92].

234
When jump phenomena occur during testing, the frequency marks on the
polar plots are incomplete, sometimes the frequency 𝜔𝑟 cannot be located and a
decrease of the amplitude of the excitation force is required.

7.2. HARMONIC EXCITATION WITH FORCES IN QUADRATURE

7.2.1. Polar Diagrams of the Frequency Response

The equation of motion of an oscillator with linear hysteretic damping and


cubic stiffness, excited by a harmonic force with a basic component 𝑓̂ 𝑐𝑜𝑠 𝜔𝑡 and
𝜋
a component in quadrature 𝜆𝑓̂𝑐𝑜𝑠 (𝜔𝑡 + 2 ) , may be written as

𝑚𝑥̈ + 𝜔 𝑥̇ + 𝑘 (𝑥 + 𝜇 𝑥 3 ) = 𝑓̂ (𝑐𝑜𝑠 𝜔𝑡 − 𝜆 𝑠𝑖𝑛 𝜔𝑡), (7.18)
where 𝜆 is the ratio of the amplitudes of the force components. The other notations
are the same as used in Chapter 2 and § 7.1.
Using the method of harmonic linearization, an approximate trial solution
is chosen of the form (7.2). Neglecting the higher harmonic terms and using
notations (7.5) we obtain
3 1+𝜆2
𝛺2 = 1 + 4 𝛾 𝑎2 ∓ √ 𝑎2
− 𝑔2 , (7.19)
𝑔
𝜆−
1+3 𝛾𝑎2−𝛺2
4
𝜑= 𝑡𝑎𝑛−1 . (7.20)
1+𝜆 3 𝑔 2
1+ 𝛾𝑎2−𝛺
4
The complex function (7.6) has components
𝑎2 3
𝑎𝑅 = (1 + 𝛾 𝑎2 − 𝛺2 + 𝜆𝑔) , (7.21)
1+𝜆2 4
𝑎2 3
𝑎𝐼 = − 1+𝜆2 [ 𝑔 − 𝜆 (1 + 4 𝛾 𝑎2 − 𝛺2 )] , (7.22)

where 𝑎 = √𝑎𝑅2 + 𝑎𝐼2 is given by equation (7.19).


Eliminating 𝑎 and 𝛺 between equations (7.21) and (7.22), we obtain circles
of equation
𝜆 2 1 2 1+𝜆2
(𝑎𝑅 − 2𝑔) + (𝑎𝐼 + 2𝑔) = (7.23)
4𝑔2

representing the loci of the tip of the vector 𝑎̃ in the Argand plane. Equation (7.23)
is identical to that obtained for the linear oscillator (2.112) so that, for the same
values 𝑓̂⁄𝑘, 𝑔 and 𝜆, the polar diagrams will be circles of the same diameter.
Two response circles (7.23), calculated for 𝑔 = 0.04, 𝛾 = 10−4, 𝜆 = 0 and
𝜆 = +1, respectively, are shown in Fig.7.8,a. For comparison, the same circles

235
calculated for 𝑔 = 0.04 and 𝛾 = 0 (for the corresponding linear system) are plotted
in Fig.7.8,b.

In the case of the nonlinear system (Fig.7.8,a) a frequency shift along the
circle can be noticed (anticlockwise for a hardening system) with respect to the
linear system (Fig.7.8,b).
For equal increments ∆𝛺 = 0.01 of the excitation frequency, the arc of
maximum length does not occur at “resonance”, i.e. at the frequency of maximum
response amplitude, so that the method of Kennedy and Pancu cannot be applied.
On the other hand, the circle 𝜆 = +1 crosses the circle 𝜆 = 0 just at the point of
maximum displacement amplitude, fact which suggests application of the “method
of three circles” (see § 2.3.1.1) to nonlinear systems too.
However, as can be seen from Fig.7.8,a, its use is limited by the jump
phenomena which occur with increasing λ, hence of the amplitude 𝑓̂√1 + 𝜆2 of the
total excitation force. Thus, as will be shown in detail in the next section, the points
on the arc 𝐶𝐸 of the circle 𝜆 = +1 correspond to unstable vibration regimes, and
cannot be plotted experimentally. For a continuous frequency sweep, when 𝛺 is
increased, the tip of the vector 𝑎̃ travels along the circle clockwise and at 𝛺 =
1.04797 jumps from 𝐶 to 𝐷; when 𝛺 is decreased, the vector rotates anticlockwise,
and a jump occurs at 𝛺 = 1.04228 from 𝐸 to 𝐹. It comes out that the frequency of
the crossing point of the two circles cannot be determined (on the circle 𝜆 = +1)
which requires the lowering of the excitation level or of the parameter λ.
Figure 7.9 shows the polar diagrams of the displacement (7.23) plotted for
𝑔 = 0.04, 𝛾 = 10−4 and several values λ. All circles are passing through the origin
1
and the point 𝑀 (0, − 𝑔) having the imaginary axis as a radical axis.
For several values of the dimensionless excitation frequency 𝛺, the points
of constant frequency on different circles have been joined by dashed lines. The

236
locus of these points is obtained eliminating 𝜆 between equations (7.21) and (7.22).
This gives
1 3 1
𝑎𝐼 = 𝑔 (1 + 4 𝛾 𝑎2 − 𝛺2 ) 𝑎𝑅 − 𝑔 . (7.24)

For 𝛾 = 0, i.e. for linear systems, equation (7.24) describes straight lines
1
radiating from the point 𝑀 (0, − 𝑔). For 𝛾 ≠ 0, equation (7.24) describes curves
passing also through the point 𝑀 and more distorted as |𝜆|, hence 𝑓̂√1 + 𝜆2 ,
increases.

From Fig.7.9 it can be seen that, as |𝜆| increases, the isochrones are so
much bent that become tangent to the response circles (7.23).
The locus of these tangency points (thick solid line), defined by the
equations
3 9
𝛺2 = 1 + 2 𝜇𝑎2 ± √16 𝛾 2 𝑎4 − 𝑔2 , (7.25)

2
−𝑔−34𝜆𝛾𝑎2 [1±√1−( 4𝑔2) ]
3𝛾𝑎
𝜑 = 𝑡𝑎𝑛−1 , (7.26)
2
1±√1−( 4𝑔2 ) −𝜆 4𝑔2
3𝛾𝑎 3𝛾𝑎

is represented by two “limit curves” (curve 𝑋𝐾𝑌 for 𝜆 < 0 and curve 𝑋′𝐾′𝑌′ for
𝜆 > 0) which separate the domains of stable and unstable vibrations.

237
The limit curves are tangent at 𝐾 and 𝐾′ to the polar diagrams of parameters
32𝑔3
𝜆𝐾 = ±√9 − 1 and to the isochrone 𝛺𝐾 = √1 + √3𝑔. Thus, for either |𝜆| > 𝜆𝐾
√3𝛾
or 𝛺 > 𝛺𝐾 the limit curves intersect the response circles and the isochrones. The
points located on the arcs of curves bounded by these crossings define unstable
regimes of harmonic vibration.

7.2.2. Jump Phenomena

Figure 7.10 illustrates the discontinuities which occur in the variation of


the amplitude and phase of the harmonic response during a frequency sweep (at
𝜆 = 𝑐𝑜𝑛𝑠𝑡. ).
On a diagram 𝜆 ≠ 0, if the frequency 𝛺 varies continuously from zero, the
tip 𝑃 of the response vector 𝑎̃ moves clockwise
from 𝑂 to 𝐶, where the circle is tangent to the
curve 𝛺′′ = 𝑐𝑜𝑛𝑠𝑡. Afterword it jumps to
point 𝐷 where the curve 𝛺′′ = 𝑐𝑜𝑛𝑠𝑡.
intersects the circle, moving then along the arc
𝐷𝑂.
When 𝛺 decreases continuously, point 𝑃
moves anticlockwise along 𝑂𝐷 to 𝐸, the point
of tangency with the curve 𝛺′ = 𝑐𝑜𝑛𝑠𝑡.,
wherefrom it jumps to point 𝐹, where the
curve 𝛺′ = 𝑐𝑜𝑛𝑠𝑡. crosses the circle, moving
then through 𝑀 to the origin. The arc 𝐶𝐸
defines unstable vibrations, while arcs 𝐹𝐶 and
𝐸𝐷 define conditionally stable vibrations. The
points on the arcs 𝑁𝐹 and 𝐷𝑂 define stable
vibrations.
In the paper [97] are described the jump phenomena which occur when 𝜆
varies continuously and 𝛺 = 𝑐𝑜𝑛𝑠𝑡., i.e. when the tip of the response vector moves
along a constant frequency curve.

7.2.3. Parameter Identification

7.2.3.1. Method of the three circles

Figure 7.11 shows two circles (7.23) plotted for two values 𝜆 ≠ 0 (solid
lines) and the circle corresponding to 𝜆 = 0 (dashed line). The three circles cross
each other at point 𝑀(0, −𝑥𝑟 ) on the negative imaginary semiaxis. On the circle
𝜆 = 0 point 𝑀 has the frequency

238
3 𝛾
𝜔𝑟 = 𝜔𝑛 √1 + 4 𝑔2 , (7.27)

on the circle 𝜆 = 𝜆1 it has the frequency


3 𝛾
𝜔′ = 𝜔𝑛 √1 + 4 𝑔2 + 𝜆1 𝑔

and on the circle 𝜆 = 𝜆2 it has the


frequency
3 𝛾
𝜔′′ = 𝜔𝑛 √1 + 4 𝑔2 + 𝜆2 𝑔 .

The equivalent hysteretic damping factor is given by


(𝜆1 +𝜆2 ) (𝜔′2 −𝜔′′2 )
𝑔= 2 2 2 , (7.28)
(1+𝜆1 )𝜔2𝑟2 −(1+𝜆2 )𝜔𝑟1

where 𝜔𝑟1 is the frequency of point 𝑀1 on the circle 𝜆 = 𝜆1 , and 𝜔𝑟2 is the
frequency of point 𝑀2 on the circle 𝜆 = 𝜆2 . Points 𝑀1 and 𝑀2 are located where
the line perpendicular to 𝑂𝑀 intersects the circles 𝜆1 and 𝜆2 .
Once 𝑔 and 𝜔𝑟 are determined, from equation (7.27) we obtain
4 𝜔2
𝛾 = 3 𝑔2 (𝜔𝑟2 − 1), (7.29)
𝑛
where
2 2 2
(1+𝜆1 )𝜔2
𝑟2 −(1+𝜆2 )𝜔𝑟1
𝜔𝑛2 = 2 2 . (7.30)
𝜆1 −𝜆2
Evaluating the slope of the elastic characteristic at the origin
1 𝑓̂
𝑘 = 𝑔𝑥 , (7.31)
𝑟

where 𝑥𝑟 = 𝑂𝑀, the coefficient of nonlinearity


2
𝑘
𝜇=𝛾 2 . (7.32)
𝑓̂

7.2.3.2. Method of the two circles

In Fig.7.12 two circles of equation (7.24) are shown, one for 𝜆 > 0 and the
other for 𝜆 = 0, crossing each other at the origin and at the point 𝑀(0, −𝑥𝑟 ). The
perpendicular 𝑀𝑀1 to the diameter 𝑂𝑀 of the circle 𝜆 = 0 crosses the circle 𝜆 > 0
at the point 𝑀1 . The damping factor has the expression

239
𝜆 (𝜔′2 −𝜔2𝑟 )
𝑔= 2 2
, (7.33)
(1+𝜆 )𝜔2
𝑟 −𝜔𝑟1

where 𝜔𝑟 is the frequency of the point 𝑀 on the circle 𝜆 = 0, 𝜔′ is the frequency


of the point 𝑀 on the circle 𝜆 > 0, and 𝜔𝑟1 is the frequency of the point 𝑀1 on the
circle 𝜆 > 0.
The constant 𝑘 is given by (7.31) and the
coefficient of nonlinearity – by equations (7.32)
and (7.29) where
2 2
(1+𝜆 ) 𝜔2
𝑟 −𝜔𝑟1
𝜔𝑛2 = 2 . (7.33)
𝜆
This method is simpler than the method of
the three circles and can lead to an inaccurate
location of the point 𝑀 on the diagram 𝜆 = 0.
This happens when, in order to avoid unstable
vibrations, the circle 𝜆 ≠ 0 is plotted for a
relatively small value of 𝜆, thus crossing the
circle 𝜆 = 0 at a small angle.
An application of this method to the determination of the damping
properties of a polyurethane foam vibration isolator is described in the paper [97].
It was shown that, neglecting the material nonlinearity, using equation (2.113)
established for linear systems instead of equation (7.33), an error of 5.5% would
have been obtained.
Pre-loading of a nonlinear system requires introduction of an additional
quadratic term in the expression of the cubic elastic restoring force. Further research
work will consider this type of stiffness function.
Other methods for parameter identification of nonlinear systems have been
proposed by Lorenz [72], Ibáñez [55] and Panik [84].

240
References

1 ANGÉLINI, J., DARRAS, B., Determination des modes propres de l’avion RF8 à
partir d’un essai de vibration au sol avec excitation non appropriée, O.N.E.R.A.
Technical Note 1/1984 RY, 1973.
2 ASHER, G.W., A Method of Normal Mode Excitation Utilizing Admittance
Measurements, Proc. National Specialists’ Meeting on Dynamics and
Aeroelasticity, Inst. of Aeronaut. Sci., Fort Worth, Texas, pp 69-76, 1958.
3 ǺSTRȌM, K., EYKHOFF, P., System Identification: A Survey, I.F.A.C.-
Automatica, vol.7, pp 123-162, 1971.
4 BA HLI, F., A General Method for Time Domain Network Synthesis, Trans. I.R.E.
on Circuit Theory, vol.1, pp 21-28, 1954.
5 BASILE, R., Recherche des charactéristiques dynamiques des systèmes continus,
Colloque International de Mécanique, Poitiers, vol.4, Publications Scientifiques et
Techniques du Ministère de l’Air, No. 261, 1950.
6 BÉATRIX, CH., Les procédés expérimentaux de l’essai global de vibration d’une
structure , La Recherche Aérospatiale, vol.109, pp 57-64, 1965.
7 BÉATRIX, CH., Experimental Determination of the Vibratory Characteristics of
Structures, O.N.E.R.A. Technical Note No.212E, 1974.
8 BEKEY, G.A., System Identification – an Introduction and a Survey, Simulation,
vol.15, nr.4, pp 151-166, Oct 1970.
9 BERMAN, A., FLANNELLY, W.G., Theory of Incomplete Models of Dynamic
Structures, A.I.A.A. Journal, vol.9, nr.8, pp 1481-1487, 1971.
10 BERMAN, A., Determining Structural Parameters from Dynamic Testing, Shock
and Vibration Digest, vol.7, nr.1, pp 10-17, 1975.
11 BERT, C.W., Material Damping: An Introductory Review of Mathematical Models,
Measures and Experimental Techniques, J. Sound Vib., vol.29, nr.2, pp 129-153,
1973.
12 BISHOP, R.E.D., The Treatment of Damping Forces in Vibration Theory, J. Royal
Aeron. Soc., vol.59, p.738, Nov 1955.
13 BISHOP, R.E.D., GLADWELL, G.M.L., An Investigation into the Theory of
Resonance Testing, Phil. Trans. Royal Soc. London, vol.255, A.1055, pp 241-273,
1963.
14 BISHOP, R.E.D., GLADWELL, G.M.L., MICHAELSON, S., The Matrix Analysis
of Vibration, University Press, Cambridge, 1965.

241
15 BISPLINGHOFF, R.L., ASHLEY, H., HALFMAN, R.L., Aeroelasticity, Addison-
Wesley, Cambridge, Mass., 1955.
16 BLEAKNEY, W.M., HAMM, J.D., Vector Methods for Flutter Analysis, J.
Aeronautical Sciences, vol.9, nr.12, pp 439-451, 1942.
17 BONNEAU, E., Determination of the Vibratory Characteristics of a Structure from
the Expression of the Complex Power Supplied, La Recherche Aérospatiale,
vol.130, pp 45-51, May-June 1969.
18 BROADBENT, E.G., HARTLEY, E.V., Vectorial Analysis of Flight Flutter Test
Results, R.A.E. Tech. Note: Struct. No.233, 1958.
19 BUZDUGAN, Gh., MIHĂILESCU, E., RADEŞ, M., Vibration Measurement,
Editura Academiei R.S. Romania, Bucureşti, 1979 (in Romanian).
20 BUZDUGAN, Gh., FETCU, L., RADEŞ, M., Vibration of Mechanical Systems,
Editura Academiei R.S. Romania, Bucureşti, 1975 (in Romanian).
21 CAUGHEY, T.K., Classical Normal Modes in Damped Linear Dynamic Systems, J.
Appl. Mechanics, Series E, vol.27, nr.2, Trans. ASME, vol.82, pp 269-271, June
1960.
22 CAUGHEY, T.K., O’KELLY, M.E.J., Classical Normal Modes in Damped Linear
Dynamic Systems, J. Appl. Mech., Series E, vol.32, nr.1, Trans. ASME, vol.87, pp
583-588, Sept 1965.
23 CHLADNI, E.F.F., Entdeckungen über die Theorie des Klanges, Weidmanns Erben
und Reich, Leipzig, 1787.
24 CLERC, D., Une méthode d’appropriation des forces d’excitation aux modes propres
non amortis d’une structure, La Recherche Aéronautique, No.85, pp 59-66, Nov-
Dec 1961.
25 CLERC, D., Sur l’appropriation des forces d’excitation lors des essais de vibration
en regime harmonique, La Recherche Aéronautique, No.87, pp 55-58, 1962.
26 COTTIN, N., DELLINGER, E., Bestimmung der dynamischen Kenngröβen linearer
elastomechanischer Systeme aus Impulsantworten, Zeitschrift für
Flugwissenschaften, vol.27, nr.8, pp 259-266, Aug 1974.
27 CUNNINGHAM, H.J., In Defense of Modern Damping Theory in Flutter Analysis,
J. Sound Vib., vol.14, pp 142-144, 1971.
28 DAVIS, J.C., Modal Modeling Techniques for Vehicle Shake Analysis, S.A.E. Paper
No 720045, Automotive Engineering Congress, Detroit, Mich., Jan.10-14, 1972.
29 DECK, A., Méthode automatique d’appropriation des forces d’excitation dans l’essai
au sol d’une structure d’avion, O.N.E.R.A. Tech. Paper No.870, 1970, presented at
Congress Euromech 22, Newcastle-upon-Tyne, Sept 15-16, 1970.
30 DECK, A., Contribution à l’étude d’une méthode semiautomatique d’appropriation
des forces d’excitation dans l’essai au sol d’une structure d’avion, Note Technique
O.N.E.R.A., No.129, 1968.

242
31 DUDNIKOV, E.E., Determination of Transfer Function Coefficients of a Linear
System from the Initial Part of the Experimental Vector Diagram, Avtomatika i
telemechanika, vol.20, nr.5, pp 576-582, 1959.
32 EYKHOFF, P., System Identification, John Wiley & Sons, London, 1974.
33 EWINS, D.J., Some Whys and Wherefores of Impedance Testing, Dynamic Testing
Symposium of Soc. Env. Engrs., Imperial College, London, Jan 5-6, 1971.
34 FEIX, M., Détermination expérimentale des parametres d’une structure élastique
linéaire à faible nombre de degrés de liberté, La Recherche Aérospatiale, No.4, pp
221-222, 1970.
35 FEIX, M., Une méthode iterative et autoguidée pour la determination des
charactéristiques dynamiques d’une structure élasto-mécanique, La Recherche
Aérospatiale, nr.1, pp 21-33, Jan-Feb 1975.
36 FLANNELLY, W.G., BERMAN, A., GIANSANTE, N., Research on Structural
Dynamic Testing by Impedance Methods, U.S.Army A.M.R.D.L. Tech. Rept. 72-63,
Nov.1972.
37 FOSS, K.A., Coordinates which Uncouple the Equations of Motion of Damped
Linear Systems, Tech. Report 25-30, Massachussetts Institute of Technology,
Aeroelastic and Structures Research Laboratory, March 1956.
38 FOSTER, R.M., A Reactance Theorem, Bell System Tech. Journal, vol.3, nr.2, pp
259-267, April 1924.
39 FÖRSCHING, H., Die Schwingungsanalysis elastomechanischer Systeme mittels
vektorieller Ortskurven, VDI-Zeitschrift, vol.105, nr.27, pp 1269-1278, Sept 1963.
40 FÖRSCHING, H., Kritischer Vergleich der Methoden zur experimentellen
Bestimmung der generalisierten Massen von Eigenschwingungsformen
elastomechanischer Systeme, VDI-Berichte Nr.88, pp 65-72, 1965.
41 FRAZER, R.A., DUNCAN, W.J., COLLAR, A.R., Elementary Matrices, Macmillan
Comp., New York, 1946.
42 GAUKROGER, D.R., SKINGLE, C.W., HERON, K.H., Numerical Analysis of
Vector Response Loci, J. Sound Vibration, vol.29, nr.3, pp 341-353, 1973.
43 GAUKROGER, D.R., HERON, K.H., SKINGLE, C.W., The Processing of
Response Data to Obtain Modal Frequencies and Damping Ratios, J. Sound
Vibration, vol.35, nr.4, pp 559-571, 1974.
44 GAUZY, H., Le glissement des fréquences propres d’un système surchargé, La
Recherche Aéronautique, No.59, pp 39-45, Jul-Aug 1957.
45 GAUZY, H., Détermination des masses généralisées à partir de la mesure de la phase,
La Recherche Aérospatiale, No.101, pp 57-63, 1964.
46 GAUZY, H., Vibration Testing by Harmonic Excitation, in Manual on Aeroelasticity
(W.P. Jones, ed.), Part 4, Ch.4A, pp 1-21, 1960.
47 GLADWELL, G.M.L., A Refined Estimate for the Damping Coefficient,
Aeronautical Journal, vol.66, pp 124-125, 1962.

243
48 GOLUB, G.H., REINSCH, C., Singular Value Decomposition and Least Squares
Solutions, in Handbook of Automatic Computation, pp 134-151, Springer Verlag,
1971.
49 GRINSTED, B., Nodal Pattern Analysis, Proc. Inst. Mech. Engrs. (A), vol.166, nr.3,
pp 309-326, 1952.
50 HART, G.C., COLLINS, J.D., Study of Modeling of Substructure Damping
Matrices, S.A.E. Transactions, vol.81, nr.4, pp 2408-2416, 1972.
51 HAWKINS, F.J., An Automatic Shake Testing Technique for Exciting the Normal
Modes of Vibration of Complex Structures, A.I.A.A. Symp. “Structural Dynamics
and Aeroelasticity”, Boston, Mass., pp 452-464, Sept 1965.
52 HERTWIG, A., FRÜH, G., LORENZ, H., Die Ermittlung der für das Bauwesen
wichtigsten Eigenschaften des Bodens durch erzwungene Schwingungen, Veröff.
DeGeBo, T. H. Berlin, Heft 1, Springer Verlag, 1933.
53 HEUKELOM, W., Über die mechanischen Elemente die die Steifigkeit von
Strassenkonstruktionen beherrschen, Report M 2302, Koninklijke Shell Lab.,
Amsterdam, 1957.
54 HILLYER, D.F., Identification of Dynamic Systems from Experimental Data,
UCRL-72818, Paper 7101 presented at S.E.S.A. Spring Meeting, Salt Lake City,
May 18-21, 1971.
55 IBÁÑEZ, P., Identification of Dynamic Structural Models from Experimental Data,
UCLA-ENG 7225, March 1972.
56 * * * Identification in Automatic Control Systems, Proc. First I.F.A.C. Symposium,
Prague, 1967.
57 * * * Identification and Process Parameter Estimation, Proc. Second I.F.A.C.
Symposium, Prague, 1970.
58 * * * Identification and System Parameter Estimation, Proc. Third I.F.A.C.
Symposium, The Hague/Delft, 1973.
59 ISERMANN, R., Experimentelle Analyse der Dynamik von Regelsystemen,
Identification I, Bibliographisches Institut, Mannheim, 1971.
60 ISERMANN, R., Prozessidentification und Parameterschätzung, Springer Verlag,
Berlin, 1974.
61 KENNEDY, C.C., PANCU, C.D.P., Use of Vectors in Vibration Measurement and
Analysis, Journal Aeronautical Sciences, vol.14, nr.1, pp 603-625, 1947.
62 KIMBALL, A.L., LOVELL, D.E., Internal Friction in Solids, Physical Review,
Series 2, vol.30, pp 948-959, Dec 1927.
63 KLOSTERMAN, A.L., On the Experimental Determination and Use of Modal
Representations of Dynamic Characteristics, Ph.D. Thesis Univ. Cincinnati, 1971.
64 KLOSTERMAN, A.L., LEMON, J.R., Building Block Approach to Structural
Dynamics, A.S.M.E. Publ. VIBR-30, 1969.
65 KRAWCZYK, R., Über Iterationsverfahren bei nichtlinearen Gleichungssystemen,
Z.A.M.M., vol.49, pp 341-349, 1969.

244
66 LANCASTER, R., Lambda-Matrices and Vibrating Systems, Pergamon Press,
Oxford, 1966.
67 LANG, G.F., Application Note 9, Nicolet Sci. Corp., 1975.
68 LAZAN, B.J., Damping of Materials and Members in Structural Mechanics,
Pergamon Press, New York, 1968.
69 LEVY, E.C., Complex-Curve Fitting, I.R.E. Trans. on Automatic Control, vol. AC-
4, pp 37-43, May 1959.
70 LEWIS, R.C., WRISLEY, D.L., A System for the Excitation of Pure Natural Modes
of Complex Structures, J. Aeronautical Sciences, vol.17, nr.11, pp 705-722, 1950.
71 LORENZ, H., Zeichnerische Auswertung von Resonanzkurven, Ingenieur-Archiv,
vol.5, pp 376-394, 1934.
72 LORENZ, H., Grundbau Dynamik, Springer Verlag, Berlin, 1960.
73 MAZET, R., Some Aspects of Ground and Flight Vibration Tests, A.G.A.R.D. Report
40T, 1956.
74 MEAD, D.J., The Internal Damping due to Structural Joints and Techniques for
General Damping Measurement, A.R.C. Report 19870, Jan 31,1958.
75 MIHĂILESCU, E., RADEŞ, M., Dynamic Structural Measurement Using the
Method of Forces in Quadrature, Symposium “Experimental Techniques in Applied
Mechanics”, Bucureşti, Romania, pp 21-34, Nov 1-3, 1972.
76 MORSE, I.E., SHAPTON, W.R., WOOD, D.M., Dynamic Machine Tool Testing,
Interim Engineering Progress Report, AFML-Contract F 33615-67-C-1727, 1969.
77 NATKE, H.G., Ein Verfahren zur rechnerischen Ermittlung der
Eigenschwingungsgröβen aus der Ergebnissen eines Schwingungsversuches in
einer Erregerkonfiguration, Dissertation, T. H. München, 1968.
78 NATKE, H.G., Die Berechnung der Eigenschwingungsgröβen eines gedämpften
Systems aus der Ergebnissen eines Schwingungsversuchs in einer
Erregerkonfiguration, Habilitationsschrift, T. U. Berlin, 1971. Gekürzte Fassung:
Jahrbuch 1971 der D.G.L.R., pp 98-120.
79 NATKE, H.G., STRUTZ, K.-D., Determination of Damping from the Vibration
Response of Elastomechanical Systems with Interactive Degrees of Freedom, in
“Dämpfungsverhalten von Werkstoffen und Bauteilen – Viscoelastische Systeme”
Kolloquium T. U. Berlin, pp 305-348, Oct 13-14, 1975 (in German).
80 NATKE, H.G., System Identification, in Handbuch zum Lehrgang “Berechnung von
Maschinenschwingungen”, V.D.I. Fachbereich Schwingungstechnik, Düsseldorf,
Apr 14-18, 1975 (in German).
81 NATKE, H.G., Survey of European Ground and Flight Vibration Test Methods,
S.A.E. Paper No.760878, 1976,
82 NICOLAI, D., Prozeβidentifikation, V.D.I.-Zeitschrift, vol.115, nr.3, pp 195-202,
1973.
83 NISSIM, E., On a Simplified Technique for the Determination of the Dynamic
Properties of a Linear System with Damping, J. Royal Aeronautical Society, vol.71,
pp 126-128, Feb 1967.

245
84 PANIK, F., Ein Verfahren zur experimentellen Ermittlung von Ersatzmodellen
kontinuierlichwe, viscoelastischer mechanischer Systeme, Dissertation, T.U.
Berlin, 1972.
85 PENDERED, J.W., BISHOP, R.E.D., A Critical Introduction to Some Industrial
Resonance Testing Techniques, J. Mechanical Engineering Science, vol.5, nr.4, pp
345-367, 1963.
86 PENDERED, J.W., Theoretical Investigation into the Effects of Close Natural
Frequencies in Resonance Testing, J. Mechanical Engineering Science, vol.7, nr.4,
pp 372-379, 1965.
87 PENESCU, C., IONESCU, G., TERTIŞCO, M., CEANGĂ, E., Identificarea
experimentală a proceselor automatizate, Editura Tehnică, Bucureşti, 1971.
88 PENROSE, R., On Best Approximate Solution of Linear Matrix Equations, Proc.
Cambridge Phil. Soc., vol.52, pp 17-19, 1956.
89 RADEŞ, M., Resonance Methods for the Dynamic Analysis of Deformable
Structures, Studii Cercetări Mecanică Aplicată, vol.32, nr.3, pp 607-633, 1973 (in
Romanian).
90 RADEŞ, M., Methods for Graphical Analysis of the Dynamic Response of
Mechanical Systems, St. Cerc. Mec. Apl., vol.33, nr.1, pp 75-103, 1974 (in
Romanian).
91 RADEŞ, M., A Technique for the Dynamic Analysis of Anti-Vibration Mountings,
Colloq. Euromech 44 “Dynamics of Machine Foundations”, Bucureşti, Romania,
pp 363-378, Oct 29-31, 1973.
92 RADEŞ, M., A Technique for Measuring the Dynamic Properties of Polyurethane
Foam Layers, Second Int. Symp. RILEM “New Developments in Non-Destructive
Testing of Non-Metallic Materials”, Constanţa, Romania, vol.1, pp 133-140, Sept
4-7, 1974.
93 RADEŞ, M., Methods for the Analysis of Structural Frequency-Response
Measurement Data, Shock and Vibration Digest, vol.8, nr.2, pp 73-88, 1976.
94 RADEŞ, M., Effect of Non-Linearity on the Polar Diagrams of the Frequency
Response of Structures, St. Cerc. Mec. Apl., vol.35, nr.1, pp 73-91, 1976 (in
Romanian).
95 RADEŞ, M., Mass Compensation at Mechanical Impedance Measurement, St. Cerc.
Mec. Apl., vol.35, nr.4, pp 547-568, 1976 (in Romanian).
96 RADEŞ, M., Measurement of Damping in Machine Tools, Lucrările a 2-a Conf. Nat.
Maşini Unelte, Bucureşti, Romania, pp 23-32, Dec 27-28, 1976 (in Romanian).
97 RADEŞ, M., Dynamic Testing of Non-Linear Materials Using Harmonic Excitation
with Forces in Quadrature, Rev. Roum. Sci. Tech.-Mécanique Appliquée, vol.22,
nr.4, pp 593-606, 1977.
98 RANEY, J.P., HOWLETT, J.T., Identification of Structural Systems by Use of Near-
Resonance Testing, NASA Rept. TN D-5069, 1969.

246
99 RATJEN, H., MOHLENKAMP, J., Untersuchung des Verfahrens von M. Feix zur
rechnerischen Ermittlung der Eigenschwingungsgröβen aus den Ergebnissen eines
Standschwingungsversuches, Diplomarbeit, T. U. Hannover, 1971.
100 RAYLEIGH, Lord, The Theory of Sound, Macmillan, London, 1894, Dover, New
York, 1950.
101 RICHARDSON, M., POTTER, R., Identification of the Modal Properties of an
Elastic Structure from Measured Transfer Function Data, I.S.A.- A.S.I. Reprint
74250, pp 239-246, May 1974.
102 RODDEN, W.P., WHITTIER, J.S., Damping of Shaker-Excited Beams Calculated
Solely from Displacement-Amplitude Measurements, J. Acoust. Soc. Amer.,
vol.34, pp 469-471, 1962.
103 ROST, J., Détermination de la masse généralisée à l’aide du procédé “par
l’impédance”, La Recherche Aéronautique, No.28, pp 49-52, 1952.
104 RUZICKA, J.E., DERBY, T.F., Influence of Damping in Vibration Isolation,
Shock and Vibration Monograph, No.7, 1971.
105 SAGE, A.P., MELSA, J.L., System Identification, Academic Press, New York,
1971.
106 SALTER, J.P., Steady-State Vibration, Mason, London, 1969.
107 SANATHANAN, C.K., KOERNER, J., Transfer Function Synthesis as a Ratio of
Two Complex Polynomials, I.E.E.E. Transactions on Automatic Control, AC-8, pp
56-58, Jan 1963.
108 von SANDEN, H., Graphical Representation of Forced Vibrations Due to C.
Runge, Ingenieur-Archiv, vol.1, p.645, 1930.
109 SCANLAN, R.H., ROSENBAUM, R., Introduction to the Study of Aircraft
Vibration and Flutter, Macmillan, New York, 1951.
110 SIMOYU, M.P., Determination of the Transfer Function Coefficients of Linearized
Units and Control Systems, Avtomatika i Telemechanika, vol.18, nr.6, pp 514-528,
1957 (in Russian).
111 SNOWDON, J.C., Representation of the Mechanical Damping Possessed by
Rubberlike Materials and Structures, J. Acoust. Soc. Amer., vol.35, nr.6, pp 821-
829, 1963.
112 STAFFIN, H.K., STAFFIN, R., Approximating Transfer Functions from Frequency
Response Data, Instruments and Control Systems, vol.38, nr.2, pp 137-144, Feb
1965.
113 STAHLE, C.V. FORLIFER, W.R., Ground Vibration Testing of Complex
Structures, Proc. A.I.A.-A.F.O.S.R. Flight Flutter Testing Symposium, May 1958.
114 STREJC, V., Auswertung der dynamischen Eigenschaften von Regelstrecken bei
gemessenen Ein- und Ausgangsignalen allgemeiner Art, Zeitschrift für Messen,
Steuren, Regeln, vol.3, nr.1, pp 7-11, 1960.
115 * * * System Identification of Vibrating Structures (W. D. Pilkey & R. Cohen, eds.),
A.S.M.E., New York, 1972.

247
116 THOREN, A.R., Derivation of Mass and Stiffness Matrices from Dynamic Test
Data, A.I.A.A. Paper 72-346, 1972.
117 TRAILL-NASH, R.W., On the Excitation of Pure Natural Modes in Aircraft
Resonance Testing, J. Aeronautical Sciences, vol.25, pp 775-778, 1958.
118 TRAILL-NASH, R.W., Some Theoretical Aspects of Resonance Testing and
Proposals for a Technique Combining Experiment and Computation,
Commonwealth of Australia, Aeronautical Research Laboratories, Report SM280,
April 1961.
119 TRAILL-NASH, R.W., LONG, G., BAILEY, C.M., Experimental Determination
of the Complete Dynamical Properties of a Two-Degree-of-Freedom Model
Having Nearly Coincident Natural Frequencies, J. Mechanical Engineering
Science, vol.9, nr.5, pp 402-413, 1967.
120 de VEUBEKE, B.M. FRAEJIS, Déphasages caractéristiques et vibrations forcées
d’un système amorti, Bulletin de la Classe des Sciences, Acad. Roy. de Belgique,
5e série, vol.34, pp 626-641, 1948.
121 de VEUBEKE, B.M. FRAEJIS, A Variational Approach to Pure Mode Excitation
Based on Characteristic Phase Lag Theory, A.G.A.R.D. Report No.39, April 1956.
122 de VEUBEKE, B.M. FRAEJIS, Influence of Internal Damping on Aircraft
Resonance, in Manual on Aeroelasticity (W.P.Jones, ed.), Part I, Ch.3, 1960.
123 VLACH, J., Computerized Approximation and Synthesis of Linear Networks, John
Wiley & Sons, New York, 1969.
124 de VRIES, G., Beitrag zur Bestimmung der Schwingungseigenshaften von
Flugzeugen im Standversuch unter besonderer Berücksichtigung eines neuen
Verfahrens zur Phasenmessung, Z.W.B. Forschungs Bericht 1882, 1942.
125 de VRIES, G., Dètermination experimentale des masses généralisées, La Recherche
Aéronautique, No.18, pp 11-22, 1950.
126 de VRIES, G., Sondage des systems vibrants par masses additionnelles, La
Recherche Aéronautique, No.30, pp 47-49, 1952.
127 de VRIES, G., L’analyse harmonique appliquée aux essais de vibration des
structures non-linéaires, La Recherche Aéronautique, No.44, pp 55-59, 1955.
128 de VRIES, G., Emploi de la méthode vectorielle d’analyse dans les essais de
vibration, La Recherche Aéronautique, No.74, pp 41-47, 1960.
129 de VRIES, G., Adaptation par correlation d’une structure vibrante au moule
linéaire, La Recherche Aéronautique, No.90, pp 59-67, Sept-Oct 1962.
130 de VRIES, G., Les raideurs électriques et leur employ dans les essais de vibration,
La Recherche Aéronautique, No.92, pp 49-55, 1963.
131 de VRIES, G., Remarques sur l’analyse des courbes d’admittance d’une structure
mécanique, La Recherche Aérospatiale, No.95, pp 49-55, 1963.
132 de VRIES, G., Le problème de l’appropriation des forces d’excitation dans l’essai
de vibration, La Recherche Aérospatiale, No.102, pp 43-49, 1964.

248
133 de VRIES, G., BEATRIX, Ch., General Measuring Process of Vibration
Characteristics of Lightly Damped Linear Structures, in Progress in Aeronautical
Sciences, vol.9, Pergamon Press, Oxford, New York, 1968.
134 WAGNER, K.W., Lehre von den Schwingungen und Wellen, Dietrich Verlag,
Wiesbaden, 1957.
135 WHITE, R.G., Measurement of Structural Frequency Response by Transient
Excitation, Inst. Sound Vib. Res., Southampton, Tech. Rep. No.12, 1969.
136 WOLFE, M.O., KIRKBY, W.T., Flight Flutter Tests, in Manual on Aeroelasticity
(W.P.Jones, ed.), Part 4, Chapter 10, 1960.
137 WOODCOCK, D.L., On the Interpretation of the Vector Plots of Forced Vibrations
of a Linear System with Viscous Damping, The Aeronautical Quarterly, vol.14,
nr.1, pp 45-62, Feb 1963.
138 YOUNG, J.P., ON, F.J., Mathematical Modelling Via Direct Use of Vibration Data,
S.A.E. Paper 69615, Nat. Aeron. Space Engr. Manufact. Meeting, Los Angeles, Oct
6-10 1969.
139 WITTMEYER, H., An Iterative, Experimental/Analytical Method for Determining
the Dynamic Parameters of a Highly Damped Elastic Body, Zeitschrift für
Flugwissenschaften, vol.19, nr.6, pp 219-224, June 1971 (in German).
140 WITTMEYER, H., Standschwingungsversuch einer Struktur mit
Dämpfungskopplung und Frequenznachbarschaft, Zeitschrift für
Flugwissenschaften, vol.24, nr.3, pp 139-151, 1976.
141 ZIMMERMANN, H., Interpretation der Ergebnisse von
Flugschwingungsversuchen anhand einer verallgemeinerten Flatterbeziehung,
Aeroelastik Kolloquium der T.U. Berlin, July 1971.
142 DAT, R., MEURZEC, J.-L., Exploitation par lissage mathématique d’admittance
d’un système linéaire, La Recherche Aérospatiale, No.4, pp 209-215, Jul-Aug
1972.
143 OVERTON, J.A., MILLS, B., The Use of Mobility Methods for the Dynamic
Analysis of Mechanical Systems, J. Soc. Environ. Engr., nr.46, pp 7-12, 1970.
144 GAUZY, H., Measurement of Inertia and Structural Damping, Manual on
Aeroelasticity, Part 4, Ch.3, O.N.E.R.A., 1960.
145 BOULLION, T., ODELL, P., Generalized Inverse Matrices, Wiley (Interscience),
New York, 1971.
146 RAO, C.R., MITRA, S.K., Generalized Inverse of Matrices and Its Applications,
Wiley, New York, 1971.
147 ALBERT, A., Regression and the Moore-Penrose Pseudoinverse, Academic Press,
New York, London, 1972.
148 IONESCU, V., LUPAȘ, L., Tehnici de calcul ȋn teoria sistemelor, I, Sisteme liniare,
Editura tehnică, Bucureşti, 1974 (p.54).
149 RUSU, ST., ROZSA, F., Metoda descompunerilor successive la identificarea
structural a Maşinilor unelte, Lucrările celei de-a doua Conferinţe Naţionale de
Maşini Unelte, Bucureşti, 27-28 dec. 1976, pp 13-22.

249
150 RADEȘ, M., Determinarea caracteristicilor dinamice ale materialelor şi structurilor,
Lucrările Conferinţei “Vibraţii ȋn construcţia de maşini”, Timişoara, 1975, pp 347-
352.
151 DAT, R., Détermination des modes propres d’une structure par essai de vibration
avec excitation non appropriée, La Recherche Aérospatiale, nr.2, pp 99-108, 1973.
152 DAT, R., Détermination des caractéristiques dynamiques d’une structure à partir
d’un essai de vibration avec excitation ponctuelle, La Recherche Aérospatiale, nr.5,
pp 301-306, 1973.

250
Redactor : LIA CREŢOIU
Tehnoredactor : TUDOR IONEL
Coperta de : OLARU ION
_________________________________________
Bun de tipar 21 XII 1978. Tiraj 2.720
Format 16/70 X 100. Coli de tipar 15,50.
C. Z. pentru biblioteci mari 534.1:62.
C. Z. pentru biblioteci mici 53:62.
_________________________________________
C. 1852 – I. P. Informaţia,
Str. Brezoianu nr. 23-25,
Bucureşti

251

Potrebbero piacerti anche