You are on page 1of 293

Preface

This textbook represents the Mechanical Vibrations lecture course given to


students in the fourth year at the Department of Engineering Sciences (now
F.I.L.S.), English Stream, University Politehnica of Bucharest, since 1993.
It grew in time from a course taught in Romanian since 1972 to students in
the Production Engineering Department, followed by a special course given
between 1985 and 1990 to postgraduate students at the Strength of Materials Chair.
Mechanical Vibrations, as a stand alone subject, was first introduced in the
curricula of mechanical engineering departments in 1974. To sustain it, we
published with Professor Gh. Buzdugan the book Vibration of Mechanical Systems
in 1975, at Editura Academiei, followed by two editions of Mechanical Vibrations,
in 1979 and 1982, at Editura didactică şi pedagogică. In 1984 we published
Vibration Measurement at Martinus Nijhoff Publ., Dordrecht, which was the
English updated version of a book published in 1979 at Editura Academiei.
As seen from the Table of Contents, this book is application oriented and
limited to what can be taught in an one-semester (28 hours) lecture course. It also
contains material to support the tutorial that includes the use of finite element
computer programs and basic laboratory experiments. The course syllabus changed
in time due to the growing use of computers. We wrote simple finite element
programs to assist students in solving problems as homework. The course aims to:
(a) increase the knowledge of vibration phenomena; (b) further the understanding
of the dynamic behaviour of structures and systems; and (c) provide the necessary
physical basis for analytical and computational approaches to the development of
engineering solutions to vibration problems.
As a course taught for non-native speakers, it has been considered useful to
reproduce as language patterns some sentences from English texts.
Computational methods for large eigenvalue problems, model reduction,
estimation of system parameters based on the analysis of frequency response data,
transient responses, modal testing and vibration testing are treated in the second
volume. No reference is made to the dynamics of rotor-bearing systems and the
vibration of discs, impellers and blades which are studied in the Dynamics of
Machinery lecture course.

April 2006 Mircea Radeş


Prefaţă

Lucrarea reprezintă cursul de Vibraţii mecanice predat studenţilor anului


IV al Facultăţii de Inginerie în Limbi Străine, Filiera Engleză, la Universitatea
Politehnica Bucureşti, începând cu anul 1993. Conţinutul cursului s-a lărgit în timp,
pornind de la un curs predat din 1972 studenţilor de la facultatea T. C. M. (în
prezent I.M.S.T.), urmat de un curs postuniversitar organizat între 1985 şi 1990 în
cadrul Catedrei de Rezistenţa materialelor.
Vibraţiile mecanice au fost introduse în planul de învăţământ al facultăţilor
cu profil mecanic ca un curs de sine stătător în 1974. Pentru a susţine cursul, am
publicat, sub conducerea profesorului Gh. Buzdugan, monografia Vibraţiile
sistemelor mecanice la Editura Academiei în 1975, urmată de două ediţii ale
manualului Vibraţii mecanice la Editura didactică şi pedagogică în 1979 şi 1982. În
1984 am publicat Vibration Measurement la Martinus Nijhoff Publ., Dordrecht,
reprezentând versiunea revizuită în limba engleză a monografiei ce a apărut în
1979 la Editura Academiei.
După cum reiese din Tabla de materii, cursul este orientat spre aplicaţii
inginereşti, fiind limitat la ceea ce se poate preda în 28 ore. Materialul prezentat
conţine exerciţii rezolvate care susţin seminarul, în cadrul căruia se utilizează
programe cu elemente finite elaborate de autor şi se prezintă lucrări demonstrative
de laborator, fiind utile şi la rezolvarea temelor de casă. Cursul are un loc bine
definit în planul de învăţământ, urmărind a) descrierea fenomenelor vibratorii
întâlnite în practica inginerească; b) modelarea sistemelor vibratoare şi analiza
acestora cu metoda elementelor finite; şi c) înarmarea studenţilor cu baza fizică
necesară în modelarea analitică şi numerică a structurilor în vibraţie şi a maşinilor,
pentru elaborarea soluţiilor inginereşti ale problemelor de vibraţii.
Fiind un curs predat unor studenţi a căror limbă maternă nu este limba
engleză, au fost reproduse expresii şi fraze din cărţi scrise de vorbitori nativi ai
acestei limbi.
În volumul al doilea se vor prezenta metode de calcul pentru probleme de
valori proprii de ordin mare, reducerea ordinului modelelor, răspunsul tranzitoriu,
estimarea pametrilor sistemelor vibratoare pe baza analizei funcţiilor răspunsului în
frecvenţă, analiza modală experimentală şi încercările la vibraţii. Nu se tratează
dinamica sistemelor rotor-lagăre şi vibraţiile discurilor şi paletelor, acestea fiind
studiate în cadrul cursului de Dinamica maşinilor.

Aprilie 2006 Mircea Radeş


Contents

Preface 1
Prefaţă 2
Contents 3
1. Modelling Vibrating Systems 5
1.1 Vibrations vs. Oscillations 5
1.2 Discrete vs. Continuous Systems 6
1.3 Simple Vibrating Systems 7
1.4 Vibratory Motions 8
1.5 Damping 10

2. Simple Linear Systems 11


2.1 Undamped Free Vibrations 11
2.2 Undamped Forced Vibrations 22
2.3 Damped Free Vibrations 35
2.4 Damped Forced Vibrations 42
Exercices 73

3. Simple Non-Linear Systems 79


3.1 Non-Linear Harmonic Response 79
3.2 Cubic Stiffness 81
3.3 Combined Coulomb and Structural Damping 92
3.4 Quadratic Damping 97
3.5 Effect of Pre-Loading 103
4 MECHANICAL VIBRATIONS

4. Two-Degree-of-Freedom Systems 105


4.1 Coupled Translation 106
4.2 Torsional Systems 119
4.3 Flexural Systems 130
4.4 Coupled Translation and Rotation 145
4.5 Coupled Pendulums 151
4.6 Damped Systems 156
Exercices 179

5. Several Degrees of Freedom 183


5.1 Lumped Mass Systems 184
5.2 Plane Trusses 210
5.3 Plane Frames 220
5.4 Grillages 234
5.5 Frequency Response Functions 241
Exercices 247

6. Continuous Systems 259


6.1 Lateral Vibrations of Thin Beams 259
6.2 Longitudinal Vibration of Rods 275
6.3 Torsional Vibration of Rods 278
6.4 Timoshenko Beams 280

References 281
Index 289
1.
MODELLING VIBRATING SYSTEMS

Vibrations are dynamic phenomena encountered in everyday life, from the


heart beating and walking, trees shaking in gusty winds or boats floating on rough
waters, vibration of musical instruments and loudspeaker cones, to bouncing of
cars on corrugated roads, swaying of buildings due to wind or earthquakes,
vibrations of conveyers and road drills.
It is customary to term ‘vibrations’ only the undesired repetitive motions,
giving rise to noise or potentially damaging stress levels. The effect of vibrations
on humans, buildings and machines are of main concern. Modelling vibration
phenomena implies describing the structure and parameters of the vibrating body,
the excitation function and the response levels.
This introductory chapter focuses on definitions and classifications, to give
an overview of the main notions used in vibration analysis.

1.1 Vibrations vs. Oscillations

The Oxford Dictionary gives “vibration, n. Vibrating, oscillation; (phys)


rapid motion to and fro, esp. of the parts of a fluid or an elastic solid whose
equilibrium is disturbed”. It comes out that all matter, gaseous, liquid or solid is
capable of executing vibrations and, in fact, so are the elementary particles of
which the matter is composed.
Generally, oscillations are variations of a state parameter about the value
corresponding to a stable equilibrium position (or trajectory). Vibrations are
oscillations due to an elastic restoring force. To save confusion, a flexible beam or
string vibrates while a pendulum oscillates.
For practical engineering purposes it is usual to allocate the term
‘vibration’ predominantly to unwanted periodic motions. In music, the opposite is
the case, since all musical instruments use periodic vibrations to make sound. We
might say that vibration in engineering is more akin to noise in acoustics: an
6 MECHANICAL VIBRATIONS

annoying, but to a degree, inescapable by-product of the machine, either in terms of


external sound or damage within itself. Apart from harmful vibrations, there are
installations whose operation is based on vibratory motions, namely: concrete
tampers, pile driving vibrators, soil compaction machines, vibrating screens,
fatigue testing machines, etc.
All bodies possessing mass and elasticity are capable of vibration. A
vibrating system has both kinetic energy, stored in the mass by virtue of its
velocity, and potential energy, stored in the elastic element as strain energy. A
major feature of vibrations is the cyclic transformation of potential energy into
kinetic and back again. In a conservative system, when there is no dissipation of
energy, the total energy is constant. At the point of maximum displacement
amplitude, the instantaneous velocity is zero, the system has only potential energy.
At the static equilibrium position, the strain energy is zero and the system has only
kinetic energy. The maximum kinetic energy must equal the maximum potential
energy. Equating the two energies it is possible to obtain the natural frequency of
vibration. This is the basis of Rayleigh’s method.
Vibrating systems are subject to damping because energy is removed by
dissipation or radiation. Damping is responsible for the decay of free vibrations, for
the phase shift between excitation and response, and provides an explanation for
the fact that the forced response of a vibratory system does not grow without limit.

1.2 Discrete vs. Continuous Systems

The number of independent coordinates needed to specify completely the


configuration of a vibrating system at any instant gives the number of degrees of
freedom of the system.
It follows that, in order to describe the motion of every particle of a
system, the number of degrees of freedom has to be infinite. However, for practical
purposes, it is useful to use systems of approximate dynamical similarity to the
actual system, which have a small number of degrees of freedom.
The criteria used to determine how many degrees of freedom to ascribe to
any system under analysis are practical in nature. For instance, some of the
possible system motions may be so small that they are not of practical interest.
Some or most of the motions of particles in the system may be practically similar,
allowing such particles to be lumped into a single rigid body. The frequency range
of the excitation forces may be so narrow that only one, or at most a few, of the
natural frequencies of the system can give rise to resonances. Groups of particles
experiencing similar motions may be considered single bodies, thereby reducing
the number of degrees of freedom necessary to consider. All these practical
considerations lead to the concept of lumped masses which are rigid bodies
1. MODELLING VIBRATING SYSTEMS 7

connected by massless flexible members. The motions predicted by using such


approximate lumped-parameter or discrete systems are often close enough to the
actual vibrations to satisfy all practical demands and to provide useful design data
and allowable vibration limits.
In some systems, a second approximation can be made, by taking into
account the mass of the elastic members. This is necessary only when the flexible
members have distributed masses which are comparable in magnitude with the
masses of system components modelled as rigid bodies.
Finally, there are many systems of practical interest which have such
simple shapes that they can be considered as systems possessing an infinite number
of degrees of freedom. Such distributed-parameter or continuous systems may be
modelled as strings, beams, plates, membranes, shells and combinations of these.
In most engineering applications, geometrically complex structures are
replaced by discretized mathematical models. A successful discretization approach
is the finite element method. The infinite degree of freedom system is replaced by a
finite system exhibiting the same behaviour. The actual structure is divided
(hypothetically) into well-defined sub-domains (finite elements) which are so small
that the shape of the displacement field can be approximated without too much
error, leaving only the amplitude to be found. All individual elements are then
assembled together in such a way that their displacements are mating each other at
the element nodes or at certain points at their interfaces, the internal stresses are in
equilibrium with the applied loads reduced at nodes, and the prescribed boundary
conditions are satisfied. Modelling errors include inappropriate element types,
incorrect shape functions, improper supports and poor mesh.

1.3 Simple Vibrating Systems

A surprisingly large number of practical vibration problems which arise in


the machines and structures designed by engineers can be treated with a
sufficiently high degree of accuracy by imagining the actual system to consist of a
single rigid body, whose motion can be described by a single coordinate.
In reality, the simplest imaginable system consists of the body whose
motion is of interest and the fixed surrounding medium, relative to which the
motion is measured. The problem of treating such a simplified system is fourfold.
The first part consists in deciding what part of the system is the rigid body and
what part are the flexible members. The second part consists in calculating the
values of the dynamic parameters of the rigid body and flexible parts. The third
part consists in writing the equations of motion of the equivalent system, Finally,
the fourth part consists in solving the equations for the prescribed conditions of
8 MECHANICAL VIBRATIONS

free or forced vibrations. Alternatively, methods using the kinetic and potential
energies may be used in the place of the last two stages.
The first two parts require judgement and experience which come with
practice, that is, with the repeated process of assuming equivalent systems,
predicting their motions and checking the predictions against actual measurements
on the real systems. Model verification and validation may require updating of
system parameters or even of the model structure. The adequacy of the solution
depends largerly on the skill with which the basic simplifying assumptions are
made. A basic choice is between linear and non-linear models. Damping estimation
is another source of error, because damping cannot be calculated like the mass and
stiffness properties. The last two steps consist in applying procedures worked out
by mathematicians. The real engineering work lies in the first two stages, while the
last two stages may be considered as mere applications of recipies.
One degree of freedom systems are considered in Chapters 2 and 3.
Discrete systems are treated in Chapters 4 and 5. Chapter 6 is devoted to straight
beams and bars.

1.4 Vibratory Motions

According to the cause producing or sustaining the vibratory motion, one


can distinguish: free vibrations, produced by an impact or an initial displacement;
forced vibrations, produced by external forces or kinematic excitation; parametric
vibrations, due to the change, produced by an external cause, of a system
parameter; self-excited vibrations, produced by a mechanism inherent in the
system, by conversion of an energy obtained from a uniform energy source
associated with the system oscillatory excitation.
If the system is distorted from the equilibrium configuration and then
released, it will vibrate with free vibrations. If any part of the system is struck by a
blow, the system will vibrate freely. Musical instruments like drums are struck and
strings are plucked. Free vibrations exist when the forces acting on the system arise
solely from motion of the system itself. The frequencies of the free vibrations are
fixed functions of the mass, stiffness, and damping properties of the system itself.
They are called natural frequencies. For any particular system they have definite
constant values. When all particles of a body vibrate in a synchronous harmonic
motion, the deflected shape is a natural mode shape.
Vibrations which take place under the excitation of external forces are
forced vibrations. External forces in any system are forces which have their
reactions acting on bodies which are not parts of the system isolated for study. The
forcing function can be harmonic, complex periodic, impulse, transient, or random.
1. MODELLING VIBRATING SYSTEMS 9

When a system is excited by a periodic external force which has one


frequency equal to or nearly equal to a natural frequency of the system, the ensuing
vibratory motion becomes relatively large even for small amplitudes of the
disturbing force. The system then is in a state of resonance. An example is the
swing pushed at the right intervals. Other examples include vibrations of geared
systems at the tooth-meshing frequency, torsional vibrations of multi-cylinder
engine shafts at the firing frequency, vibrations of rolling element bearings at the
ball passing frequencies, etc.
There is an effect arising from the damping which causes the resonance
frequency to differ slightly from the natural frequency by an amount which
increases with the damping. Fortunately the distinction in practice is very small and
can be neglected in most engineering structures, unless very high damping is
provided on purpose.
Resonance relates to the condition where either a maximum motion is
produced by a force of constant magnitude, or a minimum force is required to
maintain a prescribed motion level. A resonance is defined by a frequency, a
response level and a bandwidth of the frequency response curve. Avoidance of
large resonant vibration levels can be accomplished by: a) changing the excitation
frequency; b) making stiffness and/or mass modifications to change the natural
frequencies; c) increasing or adding damping; and d) adding a dynamic vibration
absorber.
When the driving frequency is an integer multiple of the natural frequency
of the associated linear system, non-linear single-degree-of-freedom systems
described by Mathieu equations exhibit parametric instabilities, referred to as
parametric resonances.
The principal parametric resonance occurs when the excitation frequency
is twice the natural frequency. Parametric resonances of fractional order also exist.
Multi-degree-of-freedom systems can experience parametric resonance if the
driving frequency and two or more natural frequencies satisfy a linear relation with
integer coefficients.
Parametric resonance is a state of vibration in which energy flows into the
system from an external source at resonance, increasing the amplitude of the
system’s response. This energy is dependent upon both the natural frequency of the
system and the frequency of the parameter variation.
During resonant vibrations and self-excited vibrations, the system vibrates
at its own natural frequency. But while the former are forced vibrations, whose
frequency is equal to a whole-number ratio multiple of the external driving
frequency, the latter is independent of the frequency of any external stimulus.
In a self-excited vibration, the alternating force that sustains the motion is
created or controlled by the motion itself. When the motion stops, the alternating
force disappears. Well-known examples include the vibrations of a violin string
10 MECHANICAL VIBRATIONS

being excited by a bow, the ‘chatter’ of cutting tools, of a chalk on the blackboard,
of a door that screeches when opened or of a water glass whose rim is rubbed with
a wet finger. One can add vortex induced vibrations of industrial smokestacks,
galloping and flutter of electric transmission lines, the oil-whirl of rotors in
hydrodynamic bearings, vibrations of poppet valves, the wheel shimmy, etc.
Parametric vibrations occur in systems with variable stiffness like rotating
shafts with non-circular cross-section, pendulums of variable length, geared
torsional systems, etc.

1.5 Damping
Damping represents the dissipation of energy from a system, generally as a
result of energy of motion converted into thermal energy. The loss of energy by
radiation, sometimes referred to as geometric damping, is not considered herein.
Four of the most common damping mechanisms are: a) Coulomb (sliding
friction), in which the force magnitude is independent of velocity, b) viscous,
where the force is proportional to velocity, c) velocity-nth power, when the force is
proportional to the nth power of velocity across the damper, and d) structural
(hysteretic, internal, material), in which the force is proportional to the magnitude
of displacement from some quiescent position. Hereditary damping and clearance
damping are other possible damping mechanisms.
From a microscopic point of view, most damping mechanisms involve
frictional forces that oppose the motion (velocity) of some part of a physical
system, resulting in heat loss. For example, the Coulomb friction force is caused by
two surfaces sliding with respect to one another, and this sliding force is
independent of velocity, once the initial static friction (stiction) is overcome.
Hysteretic damping may be viewed as a sliding friction mechanism
between molecular layers in a material, or between components of a riveted or
bolted structure, in which the friction force is proportional to the displacement
from the undisturbed position but in phase with the velocity.
Viscous damping occurs when molecules of a viscous fluid rub together,
causing a resistive force that is proportional to, and opposing the velocity of an
object moving through the fluid. Actual oil dampers and shock absorbers provide
friction forces proportional to some non-integer power of the relative velocity.
The influence of structural and non-linear damping mechanisms on the
response of mass-excited single-degree-of-freedom systems is treated in Chapter 3.
In the study of discrete vibrating systems only viscous and structural damping is
considered.
2.
SIMPLE LINEAR SYSTEMS

Any vibrating system has mass and elasticity. The simplest vibrating
system consists of a mass attached to a linear spring. When its motion can be
described by a single coordinate it has a single degree of freedom. Using this
simple model, it is possible to introduce basic concepts such as natural frequency,
resonance, beats and antiresonance. During vibration, energy is dissipated by
damping. This limits the motion at resonance, decreases the amplitude of free
vibration, and introduces phase shifts between excitation and response.
Measurement of damping is an important issue because it cannot be calculated like
the mass and stiffness properties.

2.1 Undamped Free Vibrations


The free vibration of a mass-spring system, that takes place in the absence
of any external excitation, is a harmonic motion whose frequency depends solely
upon the system parameters, the mass and the stiffness, being independent of the
motion initial conditions. It is referred to as a natural frequency because it is an
intrinsic (natural) system property. Calculation of natural frequencies is based on
values of the stiffness of spring elements and of inertia of mass elements.

2.1.1 The Mass-Spring System

The system shown in Fig. 2.1 consists of a linear spring of stiffness k and a
weight W having a mass m = W g , where g is the acceleration of gravity. The
weight is restricted to move in the vertical direction without rotation. The stiffness
k is defined as the change in force per unit change in length of the spring.
Figure 2.1, a shows the unstretched spring. When the mass m is suspended
from the spring (Fig. 2.1, b), its lower end moves downwards and stops in the static
equilibrium position, determined by the spring static deflection δ st . In this
position, the gravitational force W = mg acting on the mass downwards is
12 MECHANICAL VIBRATIONS

balanced by the spring force k δ st acting upwards (Fig. 2.1, c), so that the static
deflection is
mg
δ st = . (2.1)
k
If the mass is disturbed from the rest position, the system free vibrations
will take place. In order to write the equation of motion, the origin of vibration
displacements is chosen at the static equilibrium position, so that only forces due to
displacement from this position need be considered.

Fig. 2.1

Letting all vector quantities in the downward direction be positive, in


position x the elastic force acting on the mass is − k x (Fig. 2.1, d). Its motion is
described by Newton’s second law
m x&& = −k x ,
which can be written
m x&& + k x = 0 , (2.2)
where a dot above a letter denotes differentiation with respect to time.
Equation (2.2) is a homogeneous second order differential equation. Its
general solution has the form
x = C1 sin ω n t + C 2 cos ω n t , (2.3)
where
ωn = k m [rad/sec] (2.4)

is the undamped natural circular frequency of the system.


The undamped natural frequency is
1 k
fn = . [Hz] (2.5)
2π m
2. SIMPLE LINEAR SYSTEMS 13

The arbitrary constants C1 and C 2 are evaluated from the initial


conditions of the motion. In the most general case, the system may be started from
position x0 with velocity v 0 so that the general solution becomes
v0
x= sin ωn t + x0 cos ωn t . (2.6)
ωn
Another form of the general solution is
x = A sin (ωn t + φ ) (2.7)
where the two arbitrary constants are given by
ωn x0
A= x 20 + (v0 ωn ) 2 , φ = tan −1 . (2.8)
v0
Equation (2.7) indicates that the free vibration of the spring-mass system is
harmonic and occurs at a natural frequency f n . The quantity A represents the
displacement amplitude from the static equilibrium position and φ is the phase
angle. The circular frequency ω n defines the rate of vibration in terms of radians
per unit time, 2π rad being equal to one complete cycle of vibration.
The frequency of vibration is the number of complete cycles of motion in a
unit of time, and is the reciprocal of the period
T = 1 f n = 2π ω n . [sec] (2.9)

The period of vibration is the time required for the motion to begin repeating itself.
The undamped natural frequency may be expressed as a function of the
static deflection using equation (2.1)
1 g
fn = , [Hz] (2.10)
2π δ st
where g = 9.81 m s 2 .

2.1.2 Stiffness of Elastic Elements

Although it is convenient to model a single-degree-of-freedom system as a


mass attached to a single helical spring, in many actual systems the spring can take
different forms and can also represent an assemblage of several elastic elements.
In Fig. 2.2 the stiffnesses of several elastic elements are calculated as the
applied force divided by the displacement of its point of application.
14 MECHANICAL VIBRATIONS

Fig. 2.2

In Fig. 2.3 two general types of spring combinations are shown.

Fig. 2.3

For the series arrangement (Fig. 2.3, a) there is a condition of equal force
in each spring. Two linear springs, having stiffnesses k1 and k 2 , will deflect
statically when loaded by a weight W by an amount
W W ⎛1 1 ⎞
δ st = + =W ⎜⎜ + ⎟⎟ .
k1 k 2 ⎝ k1 k 2 ⎠
2. SIMPLE LINEAR SYSTEMS 15

The equivalent spring constant, representing the combined effect of k 1 and


k 2 , is
W 1
kS = = . (2.11)
δ st 1
+
1
k1 k 2
For a system with n springs connected in series, the equivalent stiffness k S
is given by
1 1 1 1
= + + ... + . (2.12)
kS k 1 k 2 kn

The parallel spring arrangement (Fig. 2.3, b) must satisfy the condition of
equal displacement in each spring and the sum of forces in each spring must equal
the weight W :
W = k 1 δ st + k 2 δ st .

Thus, for parallel springs, the equivalent stiffness is


W
kP = = k1 + k 2 . (2.13)
δ st
In general, a system with n parallel springs has an equivalent stiffness
given by
k P = k 1 + k 2 + ... + k n . (2.14)

These rules for compounding spring stiffnesses are exactly the same as those for
finding the total capacitance of series or parallel circuits in electrical engineering.

2.1.3 Torsional System

Consider the torsional system of Fig. 2.4 consisting of a disc of mass


moment of inertia J, kg m 2 , suspended from a bar or wire of torsional stiffness K,
N m rad . The system is restricted to undergo angular vibrations around the
vertical axis.
If the instantaneous angular position of the disc is given by the angle θ ,
the torque acting on the disc is − Kθ so that Newton’s second law for angular
motion is

J θ&& = − K θ ,
which can be written
16 MECHANICAL VIBRATIONS

J θ&& + K θ = 0 , (2.15)

where a dot above a letter denotes differentiation with respect to time.

Fig. 2.4

Equation (2.15) has been established by Ch. O. Coulomb in 1784. It has


the general solution of the form

θ (t ) = C1 sin ωn t + C2 cos ωn t ,
where
ωn = K J [rad/sec] (2.16)

is the undamped natural circular frequency of the torsional system.

The undamped natural frequency is


1 K
fn = . [Hz] (2.17)
2π J
From Mechanics of Materials it is known that a uniform shaft of diameter
d and length l , from a material with shear modulus of elasticity G, acted upon by a
Mt l πd4
torque M t will twist an angle θ = , where I p = is the polar second
GIp 32
moment of area of the shaft cross section. The torsional stiffness is then
M GIp
K= t = .
θ l
In fact, there is complete analogy between systems in axial and torsional
vibration, with the counterparts of springs and masses being torsional springs and
rigid discs possessing polar mass moments of inertia.
2. SIMPLE LINEAR SYSTEMS 17

2.1.4 The Energy Method

Assuming that the vibrational motion is harmonic, the frequency can be


calculated from an energy consideration. When there is no dissipation of energy,
the system is called conservative. At any instant, the energy of a conservative
system is the constant sum of potential and kinetic energies
U + T = const . (2.18)
The maximum potential energy, which occurs in an extreme position,
where the mass stands still for a moment, must equal the maximum kinetic energy,
which occurs when the mass passes through the static equilibrium position with
maximum velocity.
The spring force is k x , and the work done on an infinitesimal
displacement dx is k x d x . The potential energy in the spring, when stretched over
x
1
∫ k x dx = 2 k x
2
a distance x , is U = . Assuming the vibratory motion of the form
0
1
x = A sin ωn t , the maximum potential energy is U max = k A2 .
2
1
The kinetic energy at any instant is T = m v 2 . The velocity is
2
1
v = A ωn cos ωn t , so that the maximum kinetic energy is Tmax = m ωn2 A2 .
2
1 1
Equating U max = Tmax , we obtain k A2 = m ωn2 A2 wherefrom the
2 2
natural frequency ωn = k m is obtained, independent of the amplitude A .

Example 2.1
Determine the natural frequency of the fluid oscillations in a U tube
(Fig. 2.5).
Solution. Let the total length of the fluid column be l , the tube cross
section be A and the fluid mass density be ρ .
Assuming all fluid particles to have the same speed at any instant, the
1
kinetic energy can be written T = ρ A l x& 2 . If the fluid oscillates back and forth,
2
the work done is the same as if the fluid column of length x has been transferred
from the left side to the right side of the tube, leaving the remaining fluid
undisturbed.
18 MECHANICAL VIBRATIONS

The instantaneous potential energy is U = g ρ A x 2 . Substituting the two


energies in the condition that the rate of change of total energy must be zero
d
(T + U ) = 0
dt
and dividing out x& , we obtain the differential equation of motion of the fluid
2g
&x& + x =0.
l

Fig. 2.5

Therefore the natural frequency


ωn = 2g l
is independent of the kind of fluid used, of the tube shape and its cross-sectional
area.

2.1.5 Rayleigh’s Method

An application of the energy method to systems with distributed mass


and/or elasticity is Rayleigh’s method. It is used to reduce a distributed system into
an equivalent spring-mass system and to determine its fundamental natural
frequency.
The kinetic and potential energies are calculated assuming any reasonable
deflection curve that satisfies the geometric boundary conditions. If the true
deflection curve of the vibrating system is assumed, the fundamental frequency
found by Rayleigh’s method will be the correct frequency. For any other curve, the
frequency determined by this method will be higher than the correct frequency.
This is explained by the fact that any deviation from the true curve requires
additional constraints, a condition that implies greater stiffness and higher
2. SIMPLE LINEAR SYSTEMS 19

frequency. In the following, Rayleigh’s method is applied to beam flexural


vibrations. A prismatic beam has a bending rigidity E I (where E is Young’s
modulus and I is the second moment of area of the cross section) and a mass per
unit length ρ A (where ρ is the mass density and A is the area of the cross
section). The lateral deflection is assumed harmonic, with frequency ω 1 ,
synchronous in all points along the beam
y (x ,t ) = v (x ) cos ω 1 t .

The instantaneous potential energy is


2
M 2 dx 1 ⎛ ∂2 y ⎞
U= ∫ 2 EI
=
2 ∫ EI ⎜ 2 ⎟ dx
⎜∂x ⎟
⎝ ⎠
where the linearized differential equation (5.65) of the beam elastic line
( )
M = E I ∂ 2 y ∂ x 2 has been used.
Its maximum value is
2
1 ⎛ ∂2v ⎞
U max =
2 ∫ E I ⎜ 2 ⎟ dx .
⎜∂x ⎟
⎝ ⎠
The instantaneous kinetic energy is
2
1 ⎛∂ y⎞ 1
∫ ⎟⎟ dm = ω 12 ρ A ∫ y dx ,
2
T= ⎜⎜
2 ⎝ ∂t ⎠ 2
with the maximum value
1
Tmax = ω 12 ∫ ρ A v dx .
2
2
Equating the maximum potential energy to the maximum kinetic energy,
we obtain the expression of the fundamental natural frequency

∫ E I (∂ v ∂ x ) d x .
2 2 2
ω12 = (2.19)
∫ ρ A v dx
2

Example 2.2
Determine the fundamental natural frequency of the uniform cantilever
beam shown in Fig. 2.6.
Solution. Consider the deflection curve of the form
20 MECHANICAL VIBRATIONS

⎛ πx⎞
v = v0 ⎜1 − cos ⎟.
⎝ 2l ⎠
It can be seen that this function satisfies the boundary conditions x = 0 ,
v = 0 , dv dx = 0 , and x = l , d 2 v d x 2 = 0 , but not the condition x = l ,
d 3 v d x 3 = 0 (zero shear force), so that it is an approximate admissible function.

Fig. 2.6

π4 E I 2
The maximum potential energy is U max = v0 . The maximum
64 l 3
⎛3 2⎞ ρA 2 2
kinetic energy is Tmax = ρ A ω 21 v02 l ⎜ − ⎟ , or Tmax = ω 1 v0 l ⋅ 0.23 .
⎝4 π ⎠ 2
Equating the two energies, the fundamental frequency of vibration (in
rad/sec) is obtained as
3.6638 E I
ω1= .
l2 ρA
3.515 EI
The true solution (6.16) is ω 1 = , so that the value based on
l2 ρA
Rayleigh’s solution is 4 % higher.
If the assumed function is the static deflection curve of the massless
cantilever beam with a concentrated load at the end

1 ⎡ ⎛ x⎞ ⎛ x⎞ ⎤
2 3
v = v0 ⎢3 ⎜ ⎟ − ⎜ ⎟ ⎥ ,
2 ⎣⎢ ⎝ l ⎠ ⎝ l ⎠ ⎦⎥
3 EI 2 1
the maximum potential energy is U max = 3
v0 = k v02 and the maximum
2 l 2
1 ⎛ 33ρ Al ⎞ 2 2 1
kinetic energy is Tmax = ⎜
2 ⎝ 140 ⎠
(
2
⎟ ω 1 v0 = mred ω 1 v0 .
2
)
Equating the two energies, the fundamental frequency given by Rayleigh’s
formula is
2. SIMPLE LINEAR SYSTEMS 21

3E I l3 k 3.5675 EI
ω1 = = = ,
(33 140)ρ Al mred l2 ρA
which is only 1.47 % higher than the true solution (6.16).
The above equation indicates that, for the assumed deflection curve, the
beam with uniformly distributed mass has the same natural frequency as a massless
beam with a concentrated mass (33 140 )ρ A l attached at the end. This is called a
reduced mass.

Example 2.3
Determine the fundamental natural frequency of the free-free uniform
beam shown in Fig. 2.7.

Fig. 2.7

Solution. The assumed deflected shape can be taken of the form


πx
v = v 0 sin − a.
l
The constant a has to be determined from the conservation of momentum
for the free-free beam
l l l

∫ (velocity)⋅ d(mass ) = ∫ (ω v)(ρ A dx) = ω ρ A ∫ v dx = 0 ,


0 0 0

which yields a = 2 v 0 π .
Using the deflected shape of the form
⎛ πx 2⎞
v = v 0 ⎜ sin − ⎟,
⎝ l π⎠
equation (2.19) yields the fundamental natural frequency
22.6 EI
ω1 = .
l2 ρA
22 MECHANICAL VIBRATIONS

22.4 EI
The true solution (6.21) is ω 1 = so that the discrepancy is only
l2 ρA
0.9 %.

2.2 Undamped Forced Vibrations

Undamped forced vibrations are produced by variable forces or imposed


displacements. If the mass is subjected to a harmonic force of constant amplitude
and variable frequency, when the driving frequency approaches the system natural
frequency, the response tends to increase indefinitely. This condition is called
resonance and is characterised by violent vibrations. For undamped systems,
resonance frequencies are equal to the system natural frequencies and in most cases
operation at resonance has to be avoided. For damped systems, the response at
resonance has finite magnitude.
A swing pushed at the right intervals exhibits resonant oscillations.
Operation of soil compactors, concrete tampers, vibration conveyers, road drills
and vibrating screens is often based on resonant vibrations. However, the main
concern with resonance relates to its adverse effects. While operating at resonance,
excessive motion and stress amplitudes are generated, causing structural fatigue
and failure, harmful effects or discomfort to humans, and a decrease in product
accuracy. The nuisance of a noisy component vibrating at resonance can be an
obstacle to the sale of a car or a household appliance.
When the harmonic force is applied to the spring, the driving point
displacement decreases to zero at the system natural frequency. This condition is
called antiresonance. Generally, it is a local property, dependent upon the driving
location. It helps obtaining points with very low vibration amplitudes.

2.2.1 Mass Excitation with Arbitrary Force

Consider a force F (t ) with an arbitrary general time variation (Fig. 2.8).

During the short time interval dτ , the force F (τ ) can be considered


constant. The cross-hatched area represents an infinitesimal impulse F (τ ) dτ
which produces a velocity variation
F (τ ) dτ
d x& = .
m
The response of mass m due to the differential impulse, over the entire
response history for t > τ , is
2. SIMPLE LINEAR SYSTEMS 23

F (τ ) dτ 1
dx= sin ω n (t − τ ) , (2.20)
m ωn
which can be deduced from (2.6) considering that at t = τ , x0 = 0 and v0 = dx& .
The entire loading history may be imagined to consist of a succession of
such infinitesimal impulses, each producing its own differential response of the
form (2.20).

Fig. 2.8

For a linear system, the total response can be obtained by summing all the
differential responses developed during the loading history, that is, by integrating
equation (2.20) as follows
t
1
x (t ) = ∫ F (τ ) sin ωn (t − τ ) dτ . (2.21)
mωn
0

Equation (2.21) is generally known as the Duhamel integral for an


undamped system.

2.2.2 Mass Excitation with Harmonic Force

The mass-spring system from Fig. 2.9, a is excited by a harmonic force


f (t ) = F0 cosω t of constant amplitude F0 and driving frequency ω , applied to
the mass.
Based on the free body diagram of Fig. 2.9, b, its motion is described by
Newton’s second law
m x&& = −k x + F0 cos ω t ,
which can be written
m x&& + k x = F0 cosω t . (2.22)
24 MECHANICAL VIBRATIONS

The general solution of the linear non-homogeneous equation (2.22) is the


sum of the homogeneous solution (2.3) of the equation with zero right-hand side
and a particular solution. The particular solution can be found by assuming that it
has the same form as the forcing function

x P (t ) = X cos ω t , (2.23)

where X is the amplitude of the forced response in steady-state conditions.

Fig. 2.9

On substitution of the particular solution (2.23), equation (2.22) becomes

− mω 2 X cosω t + k X cosω t = F0 cosω t

which can be divided throughout by cosω t yielding

( k − mω ) X = F
2
0

F0 F0 k X st
or X = = = . (2.24)
k − mω 2
1 − mω k 2
1 − ( ω ωn )2
In (2.24)
F0
X st = (2.25)
k
is the static deflection of the spring under the (constant) load F0 and ω n = k m
is the undamped natural circular frequency (2.4).
Provided that ω ≠ ω n , the general solution of equation (2.22) is
2. SIMPLE LINEAR SYSTEMS 25

X st
x (t ) = C1 sin ωn t + C2 cos ωn t + cos ω t . (2.26)
1 − (ω ωn ) 2
Being the sum of two harmonic waves of different frequencies, the solution
(2.26) is not a harmonic motion.
Let the initial displacement and velocity be given by the constants x0 and
v 0 . Equation (2.26) yields
X st
x (0 ) = C2 + = x0 , x& (0) = C1ωn = v0 ,
1 − (ω ωn ) 2
so that the total response is

v0 ⎡ X st ⎤ X st
x (t ) = sinωn t + ⎢ x0 − ⎥ cosωnt + cos ω t . (2.27)
ωn ⎣⎢ 1 − (ω ωn ) ⎥⎦
2
1 − (ω ωn ) 2

For zero initial conditions, x0 = v 0 = 0 , the response (2.27) becomes

X st
x (t ) = (cos ω t − cos ωnt ) . (2.28)
1 − (ω ωn ) 2

2.2.3 Beats

The difference of cosines in equation (2.28) can be expressed as a product


2 X st
x (t ) = sin ωm t sin Δω t , (2.29)
1 − (ω ωn ) 2
where
ωn + ω ωn − ω
ωm = and Δω = .
2 2
In the case when Δω becomes very small, since ω m is relatively large, the
product in equation (2.29) represents an amplitude modulated oscillation. The
harmonic motion with higher frequency ω m is amplitude modulated by the
harmonic motion with lower frequency Δω (Fig. 2.10). The resulting motion,
which is a rapid oscillation with slowly varying amplitude, is known as beats.

The terminology is derived from acoustics. For instance when two strings
for the same note on a pianoforte are slightly out of tune, a listener hears the sound
waxing and waning (beating). The beats disappear when the strings are in unison,
and there is then only one frequency audible.
26 MECHANICAL VIBRATIONS

Fig. 2.10
Beats can be heard in an airplane when the two engines have slightly
different speeds. It occurs also in electric power stations when a generator is
started. Just before the generator is connected to the line, the electric frequency of
the generator is slightly different from the line frequency. Thus the hum of the
generator and the hum of other generators or transformers are of different pitch,
and beats can be heard.

2.2.4 Frequency Response Curves

It is of interest to examine more closely the frequency dependence of the


steady-state response amplitude
1
X= X st . (2.30)
1 − (ω ω n ) 2

The absolute value of the coefficient of X st in the right hand side of Eq.
(2.30) is referred to as the dynamic magnification factor.
Figure 2.11, a is a plot of the amplitude X as a function of the driving
frequency ω . For ω ωn < 1 the ordinates are positive, the force and motion are in
phase, while for ω ωn > 1 the ordinates are negative, the force and motion are
180 0 out of phase (Fig. 2.11, b). Whereas for ω ωn < 1 the mass is below the
static equilibrium position when the force pushes downward, for ω ωn > 1 the
mass is above the equilibrium position while the force is pushing downward.
2. SIMPLE LINEAR SYSTEMS 27

Fig. 2.11

Usually this phase relation is considered of slight interest, therefore the


resonance curve is plotted as in Fig. 2.11, c with the modulus of amplitude in the
ordinate axis. This is often referred to as a frequency response curve.

2.2.5 Resonance

At ω ωn = 1 , when the forcing frequency coincides with the system


natural frequency, the amplitude becomes infinitely large (because the system is
undamped). This phenomenon is known as “resonance”, and the natural frequency
is sometimes also called the “resonance frequency”.
At ω = ω n the spring force and the inertia force balance each other and
the exciting force increases the amplitude of motion of the undamped system
without bound. Damped systems have finite amplitudes at resonance and the phase
angle between force and displacement is 90 0 (Fig. 2.28).
Consider the case when, starting from rest, the mass-spring system is
subjected to a force of instantaneous magnitude F0 cos ωn t , where ω n is the
28 MECHANICAL VIBRATIONS

natural frequency. As ω becomes exactly equal to ω n , the solution (2.27) is no


longer valid. Substitution of F (τ ) = F0 cos ωn τ into equation (2.21) yields
t
F
x (t ) = 0 ∫ cos ω τ sin ω (t − τ )dτ ,
n n
mωn
0

F ⎡ ⎤
t t
x (t ) = 0 ⎢ sin ωn t ∫
2
cos ωn τ dτ − cos ωn t ∫ cos ωn τ sin ωnτ dτ ⎥ ,
mωn ⎢ ⎥
⎣ 0 0 ⎦
F0 t
x P (t ) = sin ωn t . (2.31)
mωn 2
Thus, when excited at resonance, the amplitude of an undamped system
increases linearly with time. Because the excitation is a cosine function and the
response is a sine function, there is a 90 0 phase angle between them. The same
result can be obtained using the limit theorems from calculus.

Fig. 2.12

The total solution for non-zero initial conditions is now of the form

v0 F0
x (t ) = sin ωn t + x 0 cos ωn t + t sin ωn t . (2.32)
ωn 2 m ωn

A plot of x (t ) versus time is given in Figure 2.12 for zero initial


conditions. It can be seen that x (t ) grows without bound, but it takes a time for the
displacement amplitude to build-up.
2. SIMPLE LINEAR SYSTEMS 29

2.2.6 Acceleration through Resonance

For most practical vibrating systems, the steady amplitude is achieved


quickly and the rate at which it is approached is of little interest.
However, when a vibrating system is driven through the resonance, i.e.
when the forcing frequency is swept with some speed ε = dω dt , there is no time
to reach a steady-state condition and the resonance amplitude is finite even for
undamped systems. Thus the response to a force of variable frequency may be of
major interest when running through a resonance.
The response exhibits a resonance-like peak, sometimes followed by a
beating-like response. If the sweep is upwards in frequency (Fig. 2.13), the peak
frequency is higher that that obtained for steady-state conditions, the peak
amplitude is lower and the width of the resonance curve is larger. If the sweep is
downwards in frequency, the peak frequency is lower than the steady-state
⎛1 π⎞
resonance frequency. In Fig. 2.13, f (t ) = F0 sin ⎜ ε t 2 + ⎟ and ε = const .
⎝2 2⎠

Fig. 2.13

The effect of sweep rate is dependent on the system damping, because the
lighter the damping, the longer the time to reach the steady-state level of vibration.
Figure 2.13 is plotted for zero damping.

2.2.7 Resonance for Constant Displacement Amplitude

Resonance relates to the condition where either a maximum motion is


produced by a force of constant magnitude, or a minimum force is required to
maintain a given motion amplitude.
30 MECHANICAL VIBRATIONS

When the force amplitude F is variable and the displacement amplitude


X 0 is kept constant, equation (2.24) can be written

[
F = k X 0 1 − (ω ωn ) 2 . ] (2.33)
Figure 2.14 is a plot of the force modulus as a function of the driving
frequency for X 0 = const. For an undamped system, the force at resonance is zero,
because the spring force is balanced by the inertia force.

Fig. 2.14

Resonance is a condition whereby a minimum of excitation is required to


produce a maximum of dynamic response.

2.2.8 Excitation with Unbalanced Rotating Masses

For many systems, vibrations are produced by driving forces from


unbalanced rotating masses. In contrast to the constant-force-amplitude case
previously discussed, the rotating-mass-type force has an amplitude proportional to
the square of the frequency of vibration. The vibratory force is thus m1eω 2 cosω t ,
where m1 is the eccentric mass located at an eccentricity e (Fig. 2.15, a).
The amplitude of the forced vibrations produced by such a force can be
obtained by a substitution of m1eω 2 for F0 in equation (2.24). Then

X=
m 1 eω 2
=
m1eω 2 k
=e
(ω ωn ) 2 . (2.34)
k − m ω2 1 − (ω ωn ) 2 1 − (ω ωn ) 2
2. SIMPLE LINEAR SYSTEMS 31

It should be pointed out that m is the total vibrating mass and includes the
mass m1 .

a b
Fig. 2.15

Figure 2.15, b is a plot of the absolute value of X from equation (2.34) as


a function of the circular frequency ω , for e = const . The curve starts from zero,
goes to infinity at resonance and decreases to e for high frequencies.

2.2.9 Antiresonance

Consider the mass-spring ungrounded system from Fig. 2.16, subjected to


a harmonic force applied to the base. The equations of motion can be written

− m x&&2 = k (x 2 − x1 ) = F0 cosω t .

The magnitude of the driving point displacement is given by

F0 k − m ω 2 F0 1 − (ω ωn ) 2
X1 = = .
k mω 2 k (ω ωn ) 2
For constant force amplitude F0 = const . , its modulus has a minimum zero
value at the natural frequency.
This is a condition of antiresonance. Generally, it takes place at a
frequency at which a maximum of force magnitude produces a minimum of
motion.
Unlike the resonance, which is a global property of a vibrating system,
independent of the driving point, antiresonance is a local property, dependent on
the driving location.
32 MECHANICAL VIBRATIONS

Fig. 2.16

In the absence of damping, the antiresonance frequency of the base excited


sprung-mass system is the same as the resonance frequency of the grounded mass-
excited system. If a second mass is attached at the driving point, the resulting mass-
spring-mass system exhibits both a resonance and an antiresonance in the driving-
point response.

2.2.10 Transmissibility

If the mass-spring system is excited by a prescribed motion


x 1 = X 1 cosω t applied to the spring end not connected to the mass, then the
motion transmitted to the mass x 2 = X 2 cosω t is defined by the amplitude ratio

X2 1
= . (2.35)
X 1 1 − (ω ωn ) 2

The ratio TR = X 2 X 1 is called transmissibility and is plotted in Fig. 2.17


as a function of the frequency ratio ω ωn .

For ω ωn > 2 , the transmissibility is less than unity and the sprung mass
is said to be isolated from the base motion. Vibration isolation is possible only
above resonance, for frequencies ω > 2ω n . The spring between the mass and the
vibrating base can be designed to ensure a given degree of isolation, by imposing
the value of TR . This shows how much the motion of the isolated mass is reduced
with respect to the case when it had been directly mounted on the vibrating base.
2. SIMPLE LINEAR SYSTEMS 33

Fig. 2.17

2.2.11 Critical Speed of Rotating Shafts

Consider the rotor shown in Fig. 2.18, consisting of a single rigid disc
symmetrically located on a uniform massless shaft supported by two rigid bearings.
The disc centre of mass G is at a radial distance e from its geometric centre C. The
centre line of the bearings intersects the plane of the disc at point O.
As the shaft starts to rotate about the bearing axis, the disc rotates in its
own plane about its geometric centre C. A centrifugal force m rG ω 2 is thus
applied to the disc, where ω is the speed of rotation, m is the mass of the disc and
rG = OG . This force causes the shaft to deflect in its bearings and the shaft is said
to be in a state of unbalance. The shaft reacts with a restoring force k rC acting in
C, where k is the stiffness of the shaft at the disc and rC = OC .
Neglecting the effect of gravity and damping, the disc is under the action
of only these two forces. In order to be in equilibrium, these forces must be
collinear, equal in magnitude, and opposite in direction

k rC = mω 2 (rC + e ) .

Solving for rC , we obtain

m ω 2e e (ω ω n ) 2
rC = = . (2.36)
k − mω 2 1 − (ω ω n ) 2
where ω n = k m is the natural circular frequency of the rotor lateral vibration at
zero speed.
34 MECHANICAL VIBRATIONS

This expression represents the radius of the orbit along which the point C
moves about the bearing axis with an angular velocity ω . Because at the same
time the disc rotates in its own plane about C with the same angular velocity, the
shaft whirling is called synchronous precession.

Fig. 2.18

The radius of the circular orbit of point G is


e
rG = rC + e = . (2.37)
1 − (ω ωn ) 2

A plot of rC (solid line) and rG (broken line) as a function of ω is given


in Fig. 2.19. At a speed ω 1 < ωn the system rotates with the heavy side G1 outside
C1 , whereas for ω 2 > ω n the light side, or the side opposite G 2 , is outside C 2 .
For very high speeds, ω >> ω n , the radius rC becomes equal to the eccentricity
and the points O and G coincide; the disc rotates about its centre of gravity.
When ω = ω n , the radii rC and rG grow without bound, a state defined as
a critical speed. Equations (2.36) and (2.37) indicate that the critical speed of the
shaft is equal to the natural frequency of the lateral vibration of the rotor.
The sudden change of the relative position of points O, C and G at the
critical speed is due to the neglection of damping. In damped systems, the segment
CG rotates continuously with respect to OC, when the shaft speed varies, so that
the “high point” does not coincide with the “heavy point”. At the critical speed, the
angle between the two segments is 90 0 (see Sec. 2.4.11)
2. SIMPLE LINEAR SYSTEMS 35

Fig. 2.19

Although there is an obvious analogy between the analytical results (2.36)


and (2.37) on one hand, and the steady-state response of a linear mass-spring
system (2.30) and (2.34) on the other hand, the forced motion of the shaft is not a
genuine vibration. The shaft does not experience any alternating stresses while
executing this motion. It just bows out in a simple bend. The bend is greatest when
the angular speed is equal to the circular frequency of bending vibration that the
shaft would have if it did not rotate and were simply executing free undamped
flexural vibrations.

2.3 Damped Free Vibrations

During vibration, energy is dissipated by friction or other resistances. The


motion amplitude in free vibration diminishes with time, while the steady
amplitude can be maintained only by external forcing. The dissipation of energy is
generally termed damping. It is produced by internal friction in materials, by
friction between structural components, by fluid-structure interactions, by radiation
or by movement in electric or magnetic fields.
The simplest damping mechanism is due to movement in a viscous
medium, and the viscous damping force is directly proportional to velocity. It is
convenient to replace all damping forces by a single equivalent viscous damping
force based on the same value of energy dissipated during a cycle of vibration.
Structural or hysteretic damping is described by a damping force in phase with
velocity but proportional to the displacement. Experience has indicated that in
aircraft structures the damping loss is better represented by the hysteretic damping.
More complicated mechanisms, such as hereditary damping, can be used to better
describe the behaviour of actual systems.
36 MECHANICAL VIBRATIONS

2.3.1 Viscous Damping

The system shown in Fig. 2.20, a consists of a linear spring of stiffness k, a


mass m and a viscous damper or dashpot. The force in the dashpot is directly
proportional to velocity and of opposite sign. The proportionality coefficient is
referred to as the viscous damping coefficient, c, having units of N (m sec ) .

Fig. 2.20

For free vibrations, the differential equation of motion can be obtained by


use of Newton’s second law and the free body diagram from Fig. 2.20, b
m x&& = −c x& − k x ,
which can be written
m x&& + c x& + k x = 0 . (2.38)

Assuming solutions of the form x = e s t , we obtain the characteristic


equation
c k
s2 + s + = 0 , (2.39)
m m
which has two roots
2
c ⎛ c ⎞ k
s 1, 2 = − ± ⎜ ⎟ − . (2.40)
2m ⎝ 2m ⎠ m
The general solution for the damped free vibrations is

x ( t ) = C1 e s1 t + C2 e s 2 t , (2.41)
where the integration constants are determined from the initial conditions.
As a reference quantity, we define critical damping as corresponding to the
value of c for which the radical in (2.40) is zero
2. SIMPLE LINEAR SYSTEMS 37

cc k
= = ωn ,
2m m

or cc = 2 k m = 2mωn . (2.42)

The actual damping of the system can be specified by a dimensionless quantity,


which is the ratio of the actual system damping to the critical system damping
c
ζ= (2.43)
cc
referred to as the damping ratio (or percent of critical damping)
Using this notation, equation (2.40) becomes

s 1, 2 = ⎛⎜ − ζ ± ζ 2 − 1 ⎞⎟ ω n . (2.44)
⎝ ⎠
Three possible cases must be considered for the above equations,
depending on whether the roots (2.44) are real, complex, or equal.

Case I: Underdamped system, ζ < 1


For ζ < 1 , equation (2.44) can be written

s 1, 2 = ⎛⎜ − ζ ± i 1 − ζ 2 ⎞⎟ ω n . (2.45)
⎝ ⎠
Substitution of (2.45) into (2.41) and conversion to trigonometric form
with the aid of Euler’s formula eiβ = cosβ + i sinβ , yields

⎛ ⎞
x (t ) = e − ζ ω n t ⎜ C1 e
i 1− ζ 2 ω n t - i 1− ζ 2 ω n t
+ C2 e ⎟,
⎝ ⎠
or
x (t ) = A e − ζ ω n t sin ⎛⎜ 1 − ζ 2 ωn t + φ ⎞⎟ . (2.46)
⎝ ⎠
Equation (2.46) indicates that the motion is oscillatory with diminishing
amplitude. The decay in amplitude with time is proportional to e − ζ ωn t , as shown
by the dashed curves in Fig. 2.21.
The frequency of the damped oscillation

ω d = 1− ζ 2 ω n (2.47)

is less than the undamped natural frequency ω n and is called the damped natural
frequency. As ζ → 1 , ω d approaches zero and the motion is no more oscillatory.
38 MECHANICAL VIBRATIONS

Equation (2.44) can be written


s 1, 2 = −σ ± iω d (2.48)

where
σ = ζ ωn (2.49)
is the rate of decay of amplitude (slope of tangent to the exponential curve at
t = 0 ).

Fig. 2.21

The following equations are useful

σ σ
ζ= , ωn = = ωd2 + σ 2 . (2.50)
ωd2 +σ 2 ζ

Fig. 2.22
2. SIMPLE LINEAR SYSTEMS 39

Case II: Overdamped system, ζ > 1


For ζ > 1 , substitution of (2.44) into (2.41) yields
⎛ −ζ + ζ 2 −1 ⎞⎟ ω n t ⎛ −ζ − ζ 2 −1 ⎞⎟ ω n t
⎜ ⎜
x (t ) = C1 e ⎝ ⎠ +C 2e
⎝ ⎠ .
The motion is no longer oscillatory (Fig. 2.22) and is referred to as
aperiodic.

Fig. 2.23

Case III: Critically damped system, ζ = 1


Critical damping represents the transition between the oscillatory and
nonoscillatory motions. In this case, the general solution is

x ( t ) = ( C1 + C2 t ) e −ω n t .
The motion is similar to that with damping greater than critical (Fig. 2.23)
but returns to rest in the shortest time without oscillation. This is used in electrical
instruments whose moving parts are critically damped to return quick on the
measured value.

2.3.2 Logarithmic Decrement

A way to determine the amount of damping in a vibrating system is to


measure the rate of decay of oscillations. This is conveniently expressed by the
logarithmic decrement, which is defined as the natural logarithm of the ratio of any
two successive amplitudes. For viscous damping, this ratio is a constant.
Consider the recorded curve of a damped vibration (Fig. 2.24), expressed
by equation (2.46).
40 MECHANICAL VIBRATIONS

Fig. 2.24

The decaying sinusoid is tangent to the exponential envelope at points that


are slightly to the right of the points of maximum amplitude, where the sine
function is equal to 1. However, this difference is negligible, so the ratio of two
successive amplitudes can be replaced by the ratio of exponential ordinates at a
period distance
x1 A e − ζ ωn t
= − ζ ω (t + T )
= e ζ ωn Td
x2 A e n d

where the period of the damped vibration is


2π 2π
Td = = .
ωn 1− ζ 2 ωd

The logarithmic decrement is


x1 2πζ
δ = ln = ζ ωn Td = . (2.51)
x2 1− ζ2
For ζ << 1 , δ ≅ 2π ζ .
Sometimes the decay after one cycle of vibration is too small so that it is
possible to distinguish a smaller amplitude only after n cycles. The ratio
x0 x 0 x1 x 2 xn −1
xn
=
x1 x 2 x3
⋅⋅
xn
= eδ ( ) n
= e nδ

so that the logarithmic decrement is given by


1 x0
δ = ln . (2.51, a)
n xn
2. SIMPLE LINEAR SYSTEMS 41

If the peak amplitude of vibration is plotted on a logarithmic scale against


the cycle number on an arithmetic scale, the points will fall on a straight line if the
damping is of viscous type as assumed in equation (2.38).
In practice, the envelopes of peaks and troughs are first drawn (Fig. 2.25).
The height between the envelope lines is then measured at each peak and trough.
The height is plotted on a logarithmic scale against the number of semicycles and
the best line is drawn through the points. The slope of this line is then used to
determine the damping ratio.

Fig. 2.25

For ζ << 1 , equation (2.51, a) yields


42 MECHANICAL VIBRATIONS

ln x n = ln x 0 − 2 ζ π ⋅ n
so that the damping ratio ζ is equal to the slope of the line divided by 2π (or by
π , if measurements are made at heights and troughs as in Fig. 2.25).

2.3.3 Loss Factor

A convenient measure of damping is obtained by the loss factor which is


defined as the ratio of energy lost per cycle (or energy that must be supplied to the
system to maintain steady-state conditions) ΔU to the peak potential energy U
stored in the system during that cycle
ΔU
η= . (2.52)
U
In general, the loss factor depends on both the amplitude and frequency of
the vibration, and can be applied also to nonlinear systems and to systems with
frequency-dependent parameters.
If X 1 and X 2 are two consecutive amplitudes of a damped free vibration,
1
the energy stored in the spring at maximum displacement is U 1 = k X 12 ,
2
1
U 2 = k X 22 . The loss of energy divided by the original energy is
2
2
U1 − U 2 U ⎛X ⎞
= 1 − 2 = 1 − ⎜⎜ 2 ⎟⎟ = 1 − e − 2δ ≅ 2δ
U1 U1 ⎝ X1 ⎠
where δ is the logarithmic decrement. Hence, for small damping, the loss factor is
approximately twice the logarithmic decrement
η ≅ 2δ . (2.52, a)

2.4 Damped Forced Vibrations

During damped forced vibrations, the response lags the excitation due to
the energy dissipation by damping. The response at phase resonance has finite
magnitude and is 90 0 phase shifted with respect to the forcing. The amplitude of
motion at resonance is related to damping and the width of the resonance curve is
directly proportional to the system damping. In the case of harmonic vibrations, the
displacement-force diagram is a closed hysteretic loop which for viscous damping
is an ellipse whose area is a measure of the energy dissipated by damping.
2. SIMPLE LINEAR SYSTEMS 43

2.4.1 Steady State Vibrations with Viscous Damping

Consider the spring-mass-dashpot grounded system subjected to a


harmonic force F0 cos ω t applied to the mass (Fig. 2.26, a).

Fig. 2.26

Based on the free body diagram from Fig. 2.26, b the differential equation
of motion can be written as
m x&& + c x& + k x = F0 cos ω t . (2.53)
The complete solution of equation (2.53) consists of the sum of the
solution (2.46) of the homogeneous equation (2.38) and a particular solution which
corresponds to the type of excitation in the right hand side.
Due to the damping, the homogeneous solution soon dies out, leaving only
the particular solution which is a harmonic motion having the same frequency as
the exciting force and a phase lag due to damping
x (t ) = X cos (ω t − ϕ ) . (2.54)
The displacement amplitude X and the phase shift ϕ between displacement
and force are found by substituting the above solution into the equation (2.53).
Shifting all the terms to the right hand side, we obtain

mω 2 X cos (ω t − ϕ ) + cω X sin (ω t − ϕ ) − k X cos (ω t − ϕ ) + F0 cos ω t = 0 .

The terms in the above equations are projections of force vectors on a


(horizontal) line at an angle ω t with respect to the force vector (Fig. 2.27).

The force vector F0 is ϕ degrees ahead of the displacement vector X .


The spring force k X is opposite to the displacement, while the inertia force
44 MECHANICAL VIBRATIONS

mω 2 X is in phase with the displacement. The damping force cω X is 90 0 ahead


the spring force. The vectors remain fixed with respect to each other and rotate
together with angular velocity ω . The rotating vector (phasor) diagram in Fig. 2.27
is drawn for a forcing frequency lower than the resonance frequency.

Fig. 2.27

Summation of vector projections in the displacement direction and in the


normal direction provides the equilibrium equations

k X − mω 2 X = F0 cos ϕ , cω X = F0 sin ϕ . (2.55)

A component of the driving force balances the damping force while the
other component is necessary to balance the reactive force, i.e. the difference
between the elastic force and the inertia force.
Solving for X and ϕ yields the amplitude of the forced vibration

F0 X st
X= = , (2.56)
( k − mω )2 2
+ ( cω )2
[ 1 − (ω ω ) ]
n
2 2
+ ( 2 ζ ω ωn ) 2

and the phase shift


cω 2ζ ω ωn
tan ϕ = = , (2.57)
k − mω 2
1 − (ω ωn ) 2

where ω n = k m and ζ = c 2 k m .
These expressions are plotted in Fig. 2.28 for several values of the
damping ratio ζ .
2. SIMPLE LINEAR SYSTEMS 45

The amplitude-frequency diagrams are called resonance curves or


frequency response curves. Such a curve starts at a value X st , increases to a
maximum at the resonance frequency, decreases through the damped natural
frequency (2.47) and the undamped natural frequency (2.4), and continues to
decrease, approaching zero asymptotically, as the frequency increases.

Fig. 2.28

The phase angle between the force and the displacement varies from zero,
at zero frequency, through 90 0 at the undamped natural frequency, to approach
180 0 asymptotically as the frequency increases. When the damping is small, the
rate of change of phase shift in passing through a natural frequency is very sharp.
For subcritical damping, the frequency response diagram (Fig. 2.28, a)
exhibits a resonance peak which is said to occur at the resonance frequency. For
ζ > 0.707 , the resonance peak is completely smoothed. Overcritically damped
systems do not exhibit resonances.
It is important to note that “amplitude resonance” is defined at the peak
X st
frequency ωr = ωn 1− 2ζ 2 where the peak value X max = of the
2 ζ 1 − ζ2
steady-state response occurs.
46 MECHANICAL VIBRATIONS

The “phase resonance” is defined to occur at the undamped natural


frequency ω = ω n (when the phase shift is 90 0 ) when the displacement amplitude
X
is X res = st . For small values of damping the two resonances coincide.

Fig. 2.29

The force vector diagram at phase resonance is shown in Fig. 2.29. The
spring force balances the inertia force of the mass, and the excitation force
overcomes the damping force only. There is a continuous interchange of potential
and kinetic energy between the spring and the mass. The only external force that
has to be applied to maintain the system vibrating is that needed to supply the
energy dissipated by damping.
At resonance, the reactive energy (in spring and mass) is zero and the
active energy (actually dissipated) is maximum. That is why a minimum of force is
required to maintain a given displacement amplitude. On a plot of the dynamic
stiffness (force required to produce unit displacement at the driving point) versus
frequency, the resonance appears as a trough (as in Fig. 2.14).

2.4.2 Displacement-Force Diagram

Consider for convenience the steady-state displacement


x (t ) = X cos ω t (2.58)
lagging by an angle ϕ the driving force applied to the mass
f (t ) = F0 cos (ω t + ϕ ) . (2.59)
2. SIMPLE LINEAR SYSTEMS 47

The two equations above are the parametric equations for an ellipse.
Eliminating the time between equations (2.58) and (2.59) yields
2
x2 f x f
2
+ −2 cos ϕ = sin 2ϕ . (2.60)
X F02 X F0

The displacement-force response forms an elliptical hysteresis loop as


shown in Fig. 2.30, which is traversed in the anticlockwise direction.

Fig. 2.30

The area inside this loop is the energy dissipated during a cycle of motion.
It is equal to the work done by the force (2.59) acting on the displacement (2.58)
2π ω 2π
dx
Wd = ∫ f dx = ∫ f
dt
d t = − X F0 ∫ cos (ω t + ϕ ) sin ω t d (ω t ) ,
0 0

Wd = π F0 X sin ϕ .
Using the second equation (2.55), the above expression becomes

Wd = π c ω X 2 . (2.61)

In order to produce work, the damping force f d = −c x& = ω X sin ω t must


be 90 0 phase shifted with respect to the displacement x (t ) = X cos ω t .
If the displacement and force are measured with appropriate transducers
and the signals are fed to an oscilloscope (displacement as ordinate and force as
abscissa) the resulting image is a Lissajous’ figure. At low frequencies the figure is
a straight line, its slope depending on the ratio of amplitudes of the two signals
(Fig. 2.31, a). As the frequency increases the straight line opens into an ellipse
(Fig. 2.31, b) whose major axis increases with frequency. At the undamped natural
frequency (Fig. 2.31, c) the ellipse major axis is very large and vertical. As the
48 MECHANICAL VIBRATIONS

frequency continues to increase, the major axis continues rotating but decreases in
magnitude (Fig. 2.31, d). The width of the ellipse decreases until at frequencies
well above resonance the ellipse is again reduced to a line which lies almost
parallel to the horizontal axis (Fig. 2.31, e).

Fig. 2.31

At the phase resonance, ω = ω n , ϕ = 90 0 , X res = F0 2 ζ k , the ellipse


major axis is vertical and the energy dissipated by damping is
2
Wd = π F0 X res = π c ω n X res . (2.62)
The energy dissipated per cycle by viscous damping is directly
proportional to the excitation frequency (eq. 2.61).

2.4.3 Structural Damping

Experiments with aircraft structures and various materials indicate that the
energy dissipated per cycle of vibration is independent of frequency and
proportional to the square of displacement amplitude. Damping values for
engineering structures are relatively low even at high resonant frequencies. Also, if
all damping were viscous, then small, high frequency bells would react to a strike
with a dull thud, instead of a clear tinkle.
This means that the viscous damping, adopted first for its mathematical
tractability, should be replaced by a model in which the energy dissipated by
damping is independent of frequency. This type of damping is called hysteretic or
structural damping.
2. SIMPLE LINEAR SYSTEMS 49

The use of the term “hysteretic” damping is somewhat confusing, since all
damping mechanisms involve a hysteresis curve of some sort. Thus the word
“structural” is preferred herein to describe this particular mechanism. It implies a
resisting force which is in phase with velocity but, unlike the viscous damping, has
a magnitude which is not proportional to the velocity but to the displacement. The
damping coefficient is inversely proportional to frequency so that the damping
force is − h x& ω (rather than − cx& ). Equation (2.53) becomes

h
m x&& + x& + k x = F0 cos ω t , (2.63)
ω
where h is the coefficient of structural damping. The inclusion of ω in the
coefficient of x& implies that only solutions with this frequency may be thought.
Alternatively, this equation of motion may be written in terms of the complex

stiffness k ∗ = k + i h , since it is decided that a harmonic solution is required

m x&& + (k + i h ) x = F0 eiω t . (2.64)

Since c is replaced by h ω , the dissipation of energy per cycle is

Wd = π h X 2 , (2.65)
which is independent of frequency.
Equations (2.56) and (2.57) become
F0 X st
X= = , (2.66)
( k − mω ) 2 2
+h 2
[1 − (ω ω ) ]
n
2 2
+g 2

h g
tan ϕ = = , (2.67)
k − mω 2
1 − (ω ωn ) 2

where g = h k is the structural damping factor.

2.4.4 The Half-Power Points Method

The resonance curve of the spring-mass-dashpot system can be used to


determine the damping ratio (Fig. 2.32).
X st
When ω = ω n the resonant amplitude is X res = . For small values of

damping, the peak M coincides with the point of phase resonance. The points B and
50 MECHANICAL VIBRATIONS

C, of ordinate ( )
2 2 X res are referred to as the half-power points. The amplitude
squared is (1 2) X res
2
, so that the power dissipated by damping at the corresponding
frequencies ω 1 and ω 2 is half the power dissipated at resonance.

Fig. 2.32

Substituting into equation (2.56) we obtain


2
1⎛ 1 ⎞ 1
⎜ ⎟ = ,
2 ⎜⎝ 2 ζ ⎟⎠
(1 − (ω ω ) )
n
2 2
+ (2 ζ ω ωn ) 2

which yields the equation

(ω ( ) (
ωn ) 4 − 2 1 − 2 ζ 2 (ω ωn ) 2 + 1 − 8 ζ 2 = 0 .)
Solving, we obtain the frequencies of half-power points

(ω ( )
ωn ) 12, 2 = 1 − 2 ζ 2 ± 2 ζ 1 + ζ 2

which for ζ << 1 can be approximated by

(ω ωn ) 12, 2 ≅ 1 ± 2 ζ .

Denoting ω 12 ≅ ωn2 ( 1 − 2 ζ ) and ω 22 ≅ ωn2 ( 1 + 2 ζ ) , we obtain


2. SIMPLE LINEAR SYSTEMS 51

ω 22 − ω 12
2ζ ≅ (2.68)
ω 22 + ω 12
or
ω 22 − ω 12 ω 2 − ω1 ω 2 + ω1 ω 2 − ω1
2ζ ≅ = ≅
2 ωn2 ωn 2 ωn ωn
so that the damping ratio is given by
Δω
ζ≅ , (2.69)
2 ωn
where Δω = ω 2 − ω1 is the bandwidth of the resonance curve.
From the shape of the resonance curve it is difficult to establish if the
damping is really of the viscous type. If only a single degree of freedom is
considered, and the motion is to be harmonic, it is most convenient to use the
concept of “equivalent viscous damping”, in which the viscous damping coefficient
has such a value that the energy dissipated in a harmonic displacement cycle of a
certain amplitude and frequency is the same as that of the actual damping
mechanism in the same displacement cycle. In equation (2.43) the coefficient c is
then the coefficient of equivalent viscous damping.

2.4.5 The Added Mass Method

In the neighbourhood of an isolated resonance, the behaviour of a general


vibrating system resembles the response of a single-degree-of-freedom system. The
equivalent mass and equivalent stiffness of the substitute system can be determined
by the additional mass method.

Fig. 2.33

Two frequency response curves are experimentally drawn, one for the
actual system, and the other for the system with a known additional mass ma
52 MECHANICAL VIBRATIONS

(Fig. 2.33). The natural frequencies ω n1 and ω n 2 are determined at peak


displacement amplitudes.
From the corresponding equations (2.4)

k = m ω n21 , (2.70)
k = (m + m a )ω n2 2 , (2.71)
it is possible to obtain the equivalent mass
ma
m= , (2.72)
(ω 2
n1 )
ω 2n 2 − 1

then, from equation (2.70), the equivalent stiffness k.

Note that the resonance response of the system with added mass is larger
because for the actual system
F0 1 F F m
X res 1 = = 0 = 0
k 2 ζ1 cω n1 c k
and for the system with the added mass
F0 1 F F m + ma
X res 2 = = 0 = 0 .
k 2 ζ 2 cω n 2 c k
If the system operating frequency is near ω n 1 , then the forced response of
the vibrating system may be decreased by adding a mass ma .

2.4.6 Solution by Complex Algebra

For harmonic excitation, the force acting on the mass of the system of Fig.
2.26 can be written
f (t ) = F0 e iω t , (2.73)
so that the steady-state solution (2.54) becomes

x (t ) = X eiω t , (2.74)
where

X = X eiθ = X R + i X I (2.75)
is the complex displacement amplitude.
In equation (2.75), X is the modulus, θ is the phase angle, X R is the real
2. SIMPLE LINEAR SYSTEMS 53

(in-phase) component, and X I is the imaginary (in-quadrature) component

X R = X cos θ , X I = X sin θ , (2.76)


X = X R2 + X I2 , tan θ = X I X R . (2.77)

For structural damping, the equation of motion (2.63) becomes


h
m x&& + x& + k x = F0 eiω t . (2.78)
ω
Substitution of (2.74) in (2.78) yields

(− m ω 2
)
+ i h + k X = F0 .

The complex amplitude X is determined as


F0 X st
X= = , (2.79)
k − mω + i h 2
1 − (ω ωn ) 2 + i g
where ω n = k m , g = h k , so that

1 − (ω ωn ) 2 −g
XR = X st , X I = X st (2.80)
[ 1 − (ω ω ) ]
n
2 2
+ g2 [ 1 − (ω ω ) ] n
2 2
+ g2

1 −g
X= X st , tan θ = . (2.81)
[ 1 − (ω ω ) ]
n
2 2
+g 2 1 − (ω ωn ) 2

Eliminating ω between the expressions of X R and X I yields


2 2
⎛ 1 ⎞ ⎛ 1 ⎞
⎜⎜ X I + X st ⎟⎟ + X R2 = ⎜⎜ X st ⎟⎟ . (2.82)
⎝ 2g ⎠ ⎝ 2g ⎠
This circle is the locus of the end of the response vector in the complex plane.

2.4.7 Frequency Response Functions

Depending on whether the response is a displacement, velocity or


acceleration, there are several frequency response functions (FRFs) defined as the
complex ratios response/excitation or excitation/response. The following
definitions are almost generally accepted and even standardized :
displacement / force = receptance,
velocity / force = mobility,
acceleration / force = accelerance (or inertance),
54 MECHANICAL VIBRATIONS

force / displacement = dynamic stiffness,


force / velocity = mechanical impedance,
force / acceleration = apparent mass.

Fig. 2.34

Because of the harmonic nature of all of the parameters involved, these


functions contain basically the same information about the vibrating system, and
simple relationships can be established between them.
2. SIMPLE LINEAR SYSTEMS 55

Basically, three different types of plots are used:


a) Bodé diagrams of the FRF modulus versus frequency plus the FRF
phase angle versus frequency;
b) diagrams of the FRF real part versus frequency plus the FRF imaginary
part versus frequency;
c) Nyquist diagrams of the FRF real part versus the FRF imaginary part.
For a system with structural damping, plots of the receptance α = X F0
are presented in Fig. 2.34 for a fixed value of the structural damping factor.
Resonance occurs at point M, while the half-power points are denoted B and C.
The Nyquist plot (Fig. 2.34, e), which is the locus of the tip of vector α in
the complex plane, is a circle. It contains on a single plot the amplitude and phase
angle information. In the vicinity of resonance the frequency scale is most
expanded so that half of the circle represents the response between the half-power
points, irrespective of the level of damping. The effect of lowering the damping is
that of increasing the diameter of the circle and of expanding the frequency scale.
Resonance is indicated by maxima in α (Fig. 2.34, a) and in α I (Fig.
2.34, d), and by points of inflection (maximum slope or maximum derivative with
respect to ω 2 ) in θ (Fig. 2.34, b) and α R (Fig. 2.34, c). The peak in α I is
sharper than that in α . At resonance, θ = −90 0 and α R = 0 . In a Nyquist plot
(Fig. 2.34, e), resonance occurs at the crossing of the circle with the imaginary
axis, where the rate of change of the arc length with respect to frequency attains a
maximum. This is based on the observation that the derivative
ds 1 dθ 1
=− = = k α 2 (2.83)
(
d ω 2 ωn2 ) h
(
d ω 2 ωn2 ) k⎡
⎢⎣
( 1 − ω 2 ωn2)2
+ g2⎤
⎥⎦

has a maximum value at resonance. If the system is excited by a harmonic force


and the receptance is plotted point by point, at equal frequency increments Δω ,
then the arc length Δ s between two successive points is a maximum at resonance.
This property is the basis of a method for locating natural frequencies developed by
Kennedy and Pancu.
The structural damping factor can be calculated from
ω 22 − ω 12
g= (2.84)
ω 22 + ω 12
where ω1 and ω 2 are the frequencies of the peaks of α R (ω ) or of the ends of the
diameter BC, perpendicular to OM, in the Nyquist plot.
56 MECHANICAL VIBRATIONS

The stiffness can be calculated from the value α res at resonance

1 1 1 F0
k= = . (2.85)
g α res g X res

Because in the complex plane the velocity is 900 phase shifted ahead the
displacement and the acceleration is 900 phase shifted ahead the velocity, the
Nyquist plots of mobility and accelerance are rotated 90 0 and 180 0 , respectively,
anticlockwise with respect to the polar plot of receptance.
The Nyquist plot of the mobility M = i ω X F0 = M R + i M I , which is not a
circle, is shown in Fig. 2.35, a and is described by the following equation

(M 2
R + M I2 ) 2

M R2
2
g km
M M
+ R I =0.
gkm

a b
Fig. 2.35

The Nyquist plot of the accelerance η = − ω 2 X F0 = η R + iη I is shown in


Fig. 2.35, b. It is a circle of equation
2 2
⎛ 1 ⎞ ⎛ 1 ⎞ 1+ g2
η
⎜ R − ⎟ + ⎜η − ⎟ = .
2m ⎠ ⎜⎝ 2 g m ⎟⎠
I
⎝ 4 g 2m2
The half-power points and the point of maximum response amplitude are
shown in both figures.
2. SIMPLE LINEAR SYSTEMS 57

2.4.8 Receptance Polar Plot for Viscous Damping

The receptance can be expressed as the ratio of the complex displacement


amplitude X to the force amplitude F0 . Using, instead of α , the general notation
for frequency response functions H (iω ) , we obtain for viscous damping

X 1 1m
H (iω ) = = 2
= 2 2
. (2.86)
F0 k − mω + iω c ωn − ω + i 2ζ ω ωn

Its Nyquist plot is not a circle, which is a drawback for the identification of
system parameters. However, it will be shown that it can be decomposed into two
circles.
Equation (2.86) can be written under the form
1m
H (iω ) = . (2.87)
(i ω − s1 )(i ω − s2 )
where s1, 2 = −σ ± iω d (2.48) are the roots of the characteristic equation (2.39).

Equation (2.87) can be expressed in terms of partial fractions


1m C1 C2
= + . (2.88)
(i ω − s1 )(i ω − s2 ) i ω − s1 i ω − s2

Multiplying both sides of equation (2.88) by (i ω − s1 ) and evaluating the


result at iω = s1 we obtain

1m i ω − s1
= C1 + C2 .
(i ω − s2 ) i ω = s1
i ω − s2 i ω = s1

Thus
1m 1m 1m
C1 = = =
s1 − s 2 (− σ + iω d ) − (− σ − iω d ) i 2 ω d
and similarly
1m
C2 = − ,
i 2ωd
1
so that equation (2.88) can be written, by factoring a constant of out of C1 and
2i
C 2 , under the standard form
58 MECHANICAL VIBRATIONS

R R∗
H (iω ) = − . (2.89)
2 i (i ω − s1 ) 2 i (i ω − s2 )
where the star denotes the complex conjugate.
In this case, the residues are purely real
1m
R = R∗ = . (2.90)
ωd
For multi-degree-of-freedom systems they are complex conjugates.
Equation (2.89) can be written
U U∗
H (iω ) = + . (2.91)
σ + i ( ω − ωd ) σ + i ( ω + ωd )
where
1
U = −U ∗ = − i . (2.92)
2 mωd
It is useful to analyse the Nyquist plot obtained from the analytical
expression (2.91).

a b
Fig. 2.36

In order to draw the plot of the first term in the summation


U
, (2.93)
σ + i ( ω − ωd )
consider the plot of
2. SIMPLE LINEAR SYSTEMS 59

1
. (2.94)
σ + i ( ω − ωd )
In the complex plane, expression (2.94) represents a circle (Fig. 2.36, a)
with centre (1 2σ , 0 ) and diameter 1 σ . At the point M of maximum amplitude, i.e.
at the crossing point of the circle with the real axis, the frequency is ω d , the
damped natural frequency. The decay rate σ equals the frequency spacing
measured from M to points B and C whose response vectors make angles of ± 45 0
with the response vector of M. For negative frequencies the circle is drawn with
dotted line.
Next consider the effect of the imaginary number U in the numerator of
expression (2.93). Multiplication of the former plot by this imaginary number
results in a clockwise rotation of the diagram by 90 0 and expansion or contraction
by an amount 1 2mω d (Fig. 2.36, b). The obtained circle is called the circle with
predominantly positive frequencies. The centre of this circle is at (0, − U 2σ ) and
U 1 1
its diameter equals = . The portion drawn with broken line
σ 2k ζ 1 − ζ 2
corresponds to negative frequencies.

a b
Fig. 2.37

Consider now the second term in the summation


U∗
. (2.95)
σ + i ( ω + ωd )
The plot of
1
. (2.96)
σ + i ( ω + ωd )
is shown in Fig. 2.37, a.
60 MECHANICAL VIBRATIONS

The expression (2.95) represents also a circle (Fig. 2.37, b), called the
circle with predominantly negative frequencies. This circle has the same diameter
as that shown in Fig. 2.36, b but is rotated 90 0 anticlockwise from the real axis.
The arc of circle corresponding to positive frequencies represents only a very small
part of the circle. The remaining part, corresponding to negative frequencies, is
drawn with broken line.
Combining the diagrams from Fig. 2.36, b and Fig. 2.37, b, the Nyquist
plot of Fig. 2.38, a is obtained, which is no more a circle. Several such diagrams
are presented in Fig. 2.38, b for different values of ω ωn and ζ .

a b
Fig. 2.38

The value of the FRF (2.89) at the damped natural frequency is

R 2i R∗ 2 i
H (iω d ) = − ,
σ σ + i 2 ωd
1 ⎡ 1 1 ⎤
H (iωd ) = ⎢ − ⎥. (2.97)
2i ⎣ mωd σ mωd (σ + i 2 ωd ) ⎦
This can be approximated as
R 1
H (iω d ) ≅ = . (2.98)
i 2 σ i 2 σ mω d

since the second term on the right of equation (2.97) approaches zero as ω d gets
large.
2. SIMPLE LINEAR SYSTEMS 61

Therefore, many single-degree-of-freedom models can be represented


simply as
R R
H (iω ) ≅ = . (2.99)
2 i (ω − s1 ) 2 i [ω − (σ − iωd )]

2.4.9 Transmissibility in Damped Systems

If the mass-spring-dashpot system is base-excited by a prescribed motion


x1 = X 1e i ω t , the motion transmitted to the mass is x 2 = X 2 e i ω t , where
X 2 = X 2 e − i ϕ is a complex amplitude. The differential equation of motion is

m &x&2 = −k (x 2 − x1 ) − c (x& 2 − x&1 )


which may be rearranged as
m &x&2 + c x& 2 + k x 2 = c x&1 + k x1 . (2.100)
The amplitude ratio is
X2 k + iω c
= . (2.101)
X 1 k − mω 2 + i ω c

Fig. 2.39
62 MECHANICAL VIBRATIONS

The motion transmissibility is

X2 1 + (2ζ ω ωn ) 2
TR = = , (2.102)
X1
[ 1 − (ω ω ) ]
n
2 2
+ (2ζ ω ωn ) 2

The phase shift is given by

2 ζ (ω ωn ) 3
tan ϕ = . (2.103)
1 − (ω ωn ) 2 + (2ζ ω ωn ) 2
Equations (2.102) and (2.103) are graphically represented in Fig. 2.39 for
various values of the damping ratio ζ .
For ω ωn > 2 , TR is less than unity, as in Fig. 2.17, but as the damping
increases, the transmissibility grows, which means a deterioration of the isolation
properties. Reducing the damping is not a good solution because, in order to
operate in the region ω ωn > 2 , the system has to pass through the resonance,
where the amplitude is reduced by damping. In some cases, there are provisions for
some light damping and the large amplitudes are limited by stops or by
acceleration through resonance.
A similar problem can be formulated for a mass-excited grounded system
(Fig. 2.26). If the driving force to be isolated is F0 ei ω t (2.73), and the steady-state
complex displacement amplitude is X (2.74), then the transmitted force through
the spring and damper is also harmonic with an amplitude

FT = (k X )2 + (cω X )2 (2.104)

so that the force transmissibility


TR = FT F0 (2.105)
is given by equation (2.102).
Note that the force transmitted through the spring and damper is phase
shifted with respect to the elastic force and the damping force.

2.4.10 Theory of Seismic Instruments

There are two basically different instruments for vibration measurement: a)


fixed reference instruments or quasistatic devices, in which the vibratory motion is
measured relative to some fixed reference point, and b) seismic instruments, in
which the vibratory motion is measured relative to the mass of a mass-spring-
dashpot system attached to the vibrating structure.
2. SIMPLE LINEAR SYSTEMS 63

The seismic instrument (Fig. 2.40) consists of the casing S, rigidly attached
to the vibrating system, the mass-spring-dashpot m-k-c system, and the transducer
T, that measures the relative motion between the seismic mass and the casing.
It is assumed that the vibrating system, hence the instrument base,
experiences a harmonic motion
x 1 (t ) = X 1 cos ω t . (2.106)

Neglecting transient terms, the relative displacement between the mass m


and the casing S can be defined by
x r (t ) = X r cos (ω t − ϕ ) . (2.107)
The absolute displacement of the mass m, relative to a fixed reference
point, is
x 2 = x1 + x r
and the absolute acceleration is
&x&2 = &x&1 + &x&r .
The equation of motion of the mass m can be written
m (&x&1 + &x&r ) + c x&r + k xr = 0
or
m &x&r + c x&r + k xr = −m &x&1 = m X1 ω 2cos ω t . (2.108)

Fig. 2.40

Equation (2.108) has a steady-state solution for which

Xr
=
(ω ωn ) 2
, (2.109)
X1
[1 − (ω ω ) ]n
2 2
+ (2ζ ω ωn ) 2
64 MECHANICAL VIBRATIONS

2ζ ω ω n
tan ϕ = . (2.110)
1 − (ω ωn ) 2

Fig. 2.41

Fig. 2.42
2. SIMPLE LINEAR SYSTEMS 65

Figure 2.41 shows the variation of the amplitude ratio (2.109) plotted
against ω ωn for two values of the damping ratio. Figure 2.42 shows the variation
of the phase shift ϕ plotted against ω ωn .
Depending on the frequency range utilized, the instrument indicates
displacement, velocity or acceleration.
Vibrometer. Within the range III, when ω >> ω n , it can be seen that
X r ≅ X 1 , so that the relative motion X r between the mass and casing, sensed by
the transducer, is essentially the same as the displacement X 1 of the structure
being measured. Figure 2.42 shows that, within this frequency range, the phase
shift is ϕ = π for light damping (ζ → 0 ) , so that the casing and the mass m are
vibrating 180 0 out of phase. Relative to an inertial frame (fixed reference point)
the mass m remains nearly stationary (becomes a fixed point in space) and the
casing motion is measured with respect to it.
When T is a displacement transducer, the instrument is a seismic absolute
displacement pickup (vibrometer). When T is a velocity transducer, the instrument
becomes a velocity pickup.
Seismic displacement-measuring instruments should have very low natural
frequencies (1 to 5 Hz) which are obtained with low values of k, hence with a soft
suspension of the seismic mass, respectively, with relatively large masses m.

Accelerometer. Within the range I, for ω << ω n , equation (2.109)


becomes
X r X 1 ≅ (ω ωn ) 2 ,
or
Xr ≅
1
ω n2
(X ω ) ,
1
2
(2.111)

where X 1ω 2 is the acceleration of the structure which is being measured.


In this case, the instrument measures a quantity directly proportional to the
absolute acceleration of the structure and is called an accelerometer. It has a high
natural frequency, obtained with a small seismic mass and a hard spring.
In the frequency range II, at ω ≅ ω n , the mass exhibits large amplitude
vibrations, property used in the design of reed-type frequency-indicating
vibrometers and accelerometers used for measuring rolling bearing defects.

Amplitude distortions. To reproduce a complex signal without distortion,


all harmonic components must be amplified equally along the frequency axis. This
can be accomplished if the amplitude ratio X r X 1 is almost constant.
66 MECHANICAL VIBRATIONS

Consequently, frequency response ranges are given for each pickup so that the
distortions might remain within prescribed limits.
Figure 2.43 shows an augmented part of Fig. 2.41, with curves plotted for
four damping ratios. The instrument with ζ = 0.7 has a horizontal response curve
down to ω ωn = 4 . Amplitude distortions set a lower frequency limit to the
displacement-measuring instrument.

Fig. 2.43

Phase distortions. To reproduce a complex signal without a change in its


shape, the phase of its harmonic components must be shifted equally along the time
axis. This can be accomplished if the phase angle ϕ increases linearly with
frequency.
For the vibrometer, the ratio ω ωn is relatively large, and ϕ is
approximately 180 0 for all harmonics, thus no phase distortion occurs. For an
accelerometer, when the damping ratio is about 0.7 , a nearly linear relationship
exists between phase angle and frequency, ϕ ≅ (π 2)(ω ωn ) . This value of
damping is also used to minimize the transient response of the instrument.

2.4.11 Rotating Shaft with External Damping

Consider the rotor from Fig. 2.18 under the action of a friction force
generated by the motion relative to its stationary environment. In synchronous
precession, points C and G rotate around the bearing axis O with an angular
velocity ω , the same as the shaft speed of rotation about C. The damping force f d
can be assumed to be proportional to the tangential velocity rC ω , so that
f d = − c rC ω , where c is the coefficient of external viscous damping.
2. SIMPLE LINEAR SYSTEMS 67

The free body diagram of the disc is shown in Fig. 2.44. The elastic
restoring force due to the shaft bending k rC acts along OC. The centrifugal force
mω 2 rG , due to the offset CG = e , acts along OG. The viscous damping force is
perpendicular to OC. Points O, C and G are no more collinear, and the line CG
leads the line OC by an angle ϕ .
Dynamic equilibrium of the three forces implies

mω 2 ( rC + e cosϕ ) = k rC , mω 2 e sinϕ = c rC ω . (2.112)

Fig. 2.44

The circular orbit of point C has a radius

rC =
mω 2 e cosϕ
=e
(ω ωn ) 2
. (2.113)
k − mω 2
[1 − (ω ω ) ]
n
2 2
+ (2ζ ω ωn ) 2

At the critical speed, when ω = ω n , the orbit radius is rC = e 2ζ .

Fig. 2.45
68 MECHANICAL VIBRATIONS

The angle between the line CG and the line OC is given by


cω 2 ζ ω ωn
tan ϕ = = . (2.114)
k − mω 2
1 − (ω ωn ) 2
This angle ϕ grows with the speed ω . When ω < ω n (Fig. 2.45, a), the
disc rotates with G outside C. With increasing speed, the segment OC increases,
and CG rotates with respect to OC. When ω = ω n , the line CG leads OC by 90 0 .
When ω > ω n , the disc rotates with G inside C, and OC decreases. At very high
speeds, point G coincides with point O, the radius rC approaches e, and the shaft
rotates about its centre of mass.

2.4.12 Hereditary Damping

Viscous damping has been considered in the simplest model, consisting of


a dashpot in parallel with a spring (Fig. 2.26). It is said that the Kelvin-Voigt model
has directly-coupled damping.

Fig. 2.46

Other simple models incorporate elastically-coupled viscous damping


mechanisms. In the three-parameter Maxwell model, the dashpot is introduced in
series with another spring (Fig. 2.46, a). The system has two degrees of freedom.
The equations of motion can be written
m &x& + k x + c (x& − x&1 ) = f , c (x& − x&1 ) = k1 x1 . (2.115)

For an excitation f = F0 ei ω t , assume solutions of the form

x = X e i ω t , x1 = X 1 ei ω t , (2.116)

where X and X 1 are complex amplitudes.


2. SIMPLE LINEAR SYSTEMS 69

Equations (2.115) become


( k − mω 2
)
+ i ω c X − i ω c X 1 = F0 ,
(2.117)
− i ω c X + ( k1 + i ω c ) X 1 = 0 ,
and can be written
(1 − β 2
)
+ i 2 ζ β X − i 2 ζ β X 1 = F0 k ,
(2.118)
− i 2 ζ β X + ( N + i 2 ζ β ) X 1 = 0,
where
ω n = k m , β = ω ωn , ζ = c 2 m ωn , N = k1 k . (2.119)

Fig. 2.47

The complex displacement amplitude of the mass m is


70 MECHANICAL VIBRATIONS

X 1+ i 2ζβ N
=
F0 k (
1 − β + i (2 ζ β N ) N + 1 − β 2
2
.
) (2.120)

The displacement amplification factor

X 1 + (2 ζ β N ) 2
= . (2.121)
F0 k
(1 − β )
2 2
(
+ (2 ζ β N )2 N + 1 − β 2 ) 2

is graphically presented in Fig. 2.47 for a stiffness ratio N = 5 and different values
of the damping ratio. The corresponding phase angle is presented in Fig. 2.48.

Fig. 2.48

The expression (2.121) can be squared and written under the form
2
⎛ X ⎞ 2
⎜ ⎟ = ψ 2 = C1 + ζ C2 . (2.122)
⎜ F0 k ⎟ C3 + ζ 2 C4
⎝ ⎠

Equation (2.122) can also be written as

( )
C3 ψ 2 − C1 + C4 ψ 2 − C2 ζ 2 = 0 . (2.123)
All curves represented by equation (2.123) are passing through the
crossing point of the curves of equations

C 3 ψ 2 − C1 = 0 , C4 ψ 2 − C2 = 0 .
These can be expressed as
2. SIMPLE LINEAR SYSTEMS 71

ψ= C1 C3 , ψ= C2 C4 ,

or
X 1
= , (2.124)
F0 k 1− β 2
and
X 1
= . (2.125)
F0 k N +1− β 2
Equation (2.124) represents the curve (2.123) of parameter ζ = 0 . Equation
(2.125) represents the curve (2.123) of parameter ζ = ∞ . The two curves intersect
each other at a point of frequency ratio β = (N + 2) 2 and of ordinate
X (F0 k ) = 2 N . All frequency response curves are passing through this point.

Figure 2.47 indicates that small variations in damping may result in


pronounced variations in resonance frequency. This is totally different from
systems with directly coupled viscous damping (Fig. 2.28), where the variation of
the resonance frequency with damping is negligible.
The resonance frequency grows from β = 1 for ζ = 0 , to β = 1 + N for
ζ = ∞ . When the damping increases, the peak of the response first decreases, then
increases, thereby indicating that an optimum degree of damping exists
ζ opt = N 2 (N + 2 ) ,

for which the resonant response is a minimum, equal to the ordinate of the crossing
point of all curves drawn for various damping ratios.
For N > 2 , over the damping range of 0.7 < ζ < ζ opt , no resonance appears
in the frequency response curve.
This behaviour can be explained considering the dynamic response of the
three parameter spring-dashpot model from Fig. 2.46, b. Its behaviour is described
by two equations
f1 = k x + c ( x& − x&1 ) , (2.126)

c ( x& − x&1 ) = k1 x1 . (2.127)

Since we are not interested in the “hidden” coordinate x1 , containing the


internal degree of freedom, we solve equation (2.127) for x1 and substitute the
result in equation (2.126) to obtain
72 MECHANICAL VIBRATIONS

t
f 1 = k x + G ( t − τ ) x& (τ ) dτ ,
∫ (2.128)
0
where
k1
− t
G (t ) = k1 e c (2.129)
with the underlying assumption that for t = 0 the model is unstrained.
In equation (2.128) the damping term depends on the past history of the
velocity. For this reason it is called “hereditary damping”.
When the force f1 is given as a function of time, the solution of equations
(2.126) and (2.127) is
2 t τ
f (t ) 1 ⎛⎜ k 1 ⎞⎟

∫e
τ1
x (t ) = 1 + f 1 (t − τ ) dτ , (2.130)
k + k 1 c ⎜⎝ k + k 1 ⎟⎠
0
where
⎛1 1 ⎞⎟
τ 1 = c⎜ +
⎜k k 1 ⎟⎠

is called the “time constant” of the model.
The first term in the right of equation (2.130) describes the instantaneous
response, actually noticed for many damped systems.
Next consider the forced response to harmonic excitation. Substituting the
complex solutions (2.116) into equations (2.126) and (2.127), then eliminating the
internal coordinate, we obtain
f1 = k x , (2.131)

where the complex stiffness k is


ω c k1
k = k +i . (2.132)
k1 + iω c

When equation (2.132) is split into its real and imaginary parts, it may be
written in the form obtained for the model with directly-coupled viscous damping
k = ke + iω ce , (2.133)
where the equivalent stiffness and equivalent coefficient of viscous damping are

ω 2c 2 k12
k e = k + k1 , ce = c . (2.134)
k12 + ω 2 c 2 k12 + ω 2 c 2
2. SIMPLE LINEAR SYSTEMS 73

The model with hereditary damping is thus reduced to a Kelvin-Voigt


model with frequency-dependent parameters. The equivalent spring stiffness k e
increases with frequency from k to the asymptotic value k + k1 , whilst the
equivalent coefficient of viscous damping ce decreases from c to zero.
The energy dissipated per cycle is
ω c k12
Wd = π X 2ω ce = π X 2 (2.135)
k12 + ω 2 c 2

which is zero for ω = 0 and ω = ∞ , having a maximum value at ω 0 = k1 c where


ce = c 2 .

Fig. 2.49

This is also reflected in the hysteresis curves (Fig. 2.49). At zero frequency
we have a straight line corresponding to a pure spring of stiffness k. For ω = ω 0
there is an ellipse of maximum area. When the frequency tends to infinity, we have
a straight line of smaller slope, corresponding to a pure spring of stiffness k + k1 .

Exercises
2.E1 An unknown mass m is hung on a spring of unknown stiffness k.
When a mass m 1 = 0.5 kg is added to m, the system natural frequency is lowered
from 50 Hz to 49 Hz . a) Determine the values of m and k. When second spring of
stiffness k ′ is added in parallel with the first spring, the natural frequency is
increased to 50 Hz . b) Determine the value of k ′ .

Answer: m = 12.126 kg , k = 1.19 ⋅ 10 6 N m , k ′ = 56 N m .


74 MECHANICAL VIBRATIONS

2.E2 A mass m = 0.5 kg vibrating in a viscous medium has a period


T = 0.15 sec and an initial amplitude a 0 = 10 mm . a) Determine the stiffness k and
the viscous damping coefficient c if the amplitude after 12 cycles is a12 = 0.2 mm .
b) Determine the amplitude a12′ when a mass m1 = 0.3 kg is added to m.
′ = 0.45 mm .
Answer: k = 877 N m , c = 2.173 N s m , a12

2.E3 A spring-mass system with a natural frequency f = 5 Hz vibrates in


a viscous medium having a damping coefficient c = 0.002 N s mm . The
logarithmic decrement is δ = 0.31 . a) Determine the mass m and the spring
stiffness k. b) Find the new value of the logarithmic decrement if a mass m 1 = 1 kg
is added to the first mass.
Answer: m = 0.645 kg , k = 636.4 N m , δ ′ = 0.194 .

2.E4 A spring-mass system with viscous damping is displaced a distance


a 0 = 20 mm and released. After 8 cycles of vibration the amplitude decays to
a8′ = 4 mm . A mass m 1 = 2 kg is attached to the initial mass producing a static
displacement of 4 mm . If the new system is displaced a distance a 0 = 20 mm and
released, after 8 cycles of vibration the amplitude decays to a8′′ = 5 mm . Determine
the mass m, the stiffness k and the damping coefficient c.
Answer: m = 5.77 kg , k = 4905 N m , c = 10.76 Ns m .

2.E5 A vibrating system of mass m = 5 kg and spring stiffness


k = 1 N mm is acted upon by a harmonic force of amplitude F0 = 10 N and
frequency 2 Hz . Determine a) the displacement amplitude X; b) the displacement
amplitude X ′ when a mass m 1 = 2 kg is added to m; c) for the system with the
added mass, how should k be modified so as the displacement amplitude to become
X again.
Answer: X = 47.34 mm , X ′ = 95.85 mm , add k ′ = 315.5 N m in parallel,
or k ′ = 8546.5 N m in series.

2.E6 A weight m g = 20 N attached to a light spring elongates it 1 mm .


Determine a) the amplitude F0 of the force that, acting with a frequency of 20 Hz ,
produces vibrations with an amplitude X = 3.5 mm . b) Find another excitation
2. SIMPLE LINEAR SYSTEMS 75

frequency at which the force of magnitude F0 produces the same displacement


X = 3.5 mm .
Answer: F0 = 42.68 N , ω ′ = 61.87 rad sec .

2.E7 A small motor of mass m = 2 kg is found to be transmitting a force


of 14 N to its supporting springs when running at a speed of 30 rad sec and has
an amplitude of vibration of 3.5 mm . Determine a) the amplitude of the
unbalanced force F0 ; b) Find another value of the running speed at which the
amplitude of vibration has the same magnitude.
Answer: F0 = 7.7 N , ω ′ = 55.67 rad sec .

2.E8 A fragile instrument of mass m = 1 kg is used on a table that is


vibrating because of its proximity to machines running in the area. The table has a
harmonic motion of amplitude X 1 = 0.2 mm and frequency f = 15 Hz . To reduce
the vibration of the instrument, it is isolated from the table by means of rubber pads
of stiffness k = 7,000 N m . Determine a) the amplitude X of the displacement of
the instrument: b) the value X ′ of the amplitude, if a pad of stiffness
k ′ = 2,000 N m is added in parallel with the first pad.
Answer: X = 0.743 mm , X ′ = 15.148 mm .

2.E9 A mass m = 4 kg is supported between two springs attached to fixed


points is excited by a harmonic force of amplitude F0 = 10 N . The upper spring
has a stiffness k = 5,000 N m and the lower spring has a stiffness 2k. Determine a)
the frequencies at which the displacement amplitude is X = 20 mm , and b) the
amplitude F0′ of the force transmitted to the lower support through the spring 2k.

Answer: ω 1 = 60.2 rad sec , ω 2 = 62.25 rad sec , F0′ = 200 N .

2.E10 An undamped vibrating system has an equivalent mass m and an


equivalent stiffness k. A harmonic force of amplitude F0 = 1 N and frequency
ω = 14 rad sec produces steady-state vibrations of amplitude X. When a mass
m1 = 2 kg is added to m, the force of amplitude F0 = 1 N must have a frequency of
either ω = 10 rad sec or ω = 12 rad sec in order to produce the same displacement
amplitude X. Determine the values of m and k.
76 MECHANICAL VIBRATIONS

Answer: For ω < ω1 , m = 2.083 kg , k = 498.1 N m ; for ω > ω1 ,


m = 5.538 kg , k = 919.6 N m .

2.E11 A machine weighing m g = 12,000 N is mounted at the middle of a


beam of length l = 2 m , cross section second moment of area I = 20 mm 4 and
Young’s modulus E = 2.1 ⋅ 10 5 N mm 2 , simply supported at the ends. If the
machine has a rotating unbalanced weight m0 g = 2,000 N with eccentricity
e = 0.1 mm and speed n = 1500 rpm , determine a) the amplitude X of the forced
vibrations of the machine, and b) the magnitude of the dynamic force transmitted to
only one of the supports.
Answer: X = 50.22 μm , Fdin = 632.77 N .

2.E12 A motor of mass m = 10 kg , supported on springs of total stiffness


k = 2,000 N m , generates an unbalanced force of magnitude F0 = 20 N at a
frequency of 2 Hz . Determine the stiffness k ′ of a spring and how should it be
mounted (in series or in parallel) so as to lower to a half the magnitude of the
transmitted force FT 0 . Find the value of another frequency of the excitation force
having the same effect.
Answer: k ′ = 731.6 N m in parallel, ω ′ = 10.747 rad sec .

2.E13 A vibrating system consists of a mass m = 0.2 kg and a spring of


stiffness k = 0.2 N mm . It is acted upon by a harmonic force of amplitude
F0 = 5 N and frequency ω = 35 rad sec . Determine a) the amplitude FT 0 of the
force transmitted to the support, and b) the amplitude X of the displacement of the
mass m.
Answer: FT 0 = 22.22 N , X = 0.11 m .

2.E14 A spring-mass system with viscous damping is acted upon by a


harmonic force of amplitude F0 = 10 N . Varying the frequency of the excitation
force, a peak amplitude X 1 = 60 mm is measured at the frequency ω 1 = 14 rad sec .
When a mass m 1 = 2 kg is added to the system, the peak amplitude becomes
X 2 = 80 mm . Determine the mass m, the stiffness k and the damping coefficient c.

Answer: m = 2.57 kg , k = 503.72 N m , c = 11.9 N s m .


2. SIMPLE LINEAR SYSTEMS 77

2.E15 A mass m = 5 kg is attached to a spring of stiffness k = 1 N mm


and is acted upon by a harmonic force of amplitude F0 = 10 N and frequency
f = 2 Hz . Determine: a) the amplitude X 1 of the forced vibrations of the mass m;
b) the amplitude X 2 of the forced vibrations when a mass m 1 = 2 kg is added to
m; c) the coefficient of viscous damping c of a dashpot connected in parallel with
the spring which reduces the amplitude of vibrations to the initial value.
Answer: X 1 = 47.3 mm , X 2 = 92.5 mm , c = 10.3 N s m .

2.E16 Using Rayleigh’s method, estimate the fundamental frequency of


the vibrating beams shown in Fig. 2.50 using the given deflection function.

⎛ π x⎞
v (x ) = v0 ⎜1 − cos ⎟.
a ⎝ 2l ⎠

⎡ x ⎛ x⎞ ⎤
3
v (x ) = v0 ⎢3 − 4 ⎜ ⎟ ⎥ ,
b ⎢⎣ l ⎝ l ⎠ ⎥⎦

0≤ x≤l 2

πx
c v (x ) = v0 sin .
l

⎛ 2π x ⎞
d v (x ) = v0 ⎜1 − cos ⎟.
⎝ l ⎠

Fig. 2.50

EI EI
Answer: a) ω 1 = 1.612 ; b) ω 1 = 6.928 ;
ml 3
ml (1 + 0.4857 ρ Al m )
3

EI 22.79 EI
c) ω 1 = 4.935 ; d) ω 1 = .
ml 3
l2 ρA
78 MECHANICAL VIBRATIONS

2.E17 Using Rayleigh’s method, estimate the fundamental frequency for


the axial vibrations of the beam shown in Fig. 2.51 using a linear deflection
function for the longitudinal displacement. Consider m = 2 ρ Al .

Fig. 2.51
0.655 E
Answer: ω 1 = .
l ρ
3.
SIMPLE NON-LINEAR SYSTEMS

Methods for parameter estimation are presented, based on the analysis of


frequency response curves. Only simple non-linear systems are considered, to show
the distortion of frequency response curves due to the system slightly non-linear
properties. This is not a full account of the non-linear dynamic behaviour of
vibrating structures. A comprehensive treatment requires a separate book.

3.1 Non-Linear Harmonic Response

If a non-linear system is excited by a harmonic force, the steady-state


response is not harmonic, as for linear systems, but it is periodic, so that it can be
expressed as a sum of harmonic components. For systems with local weak non-
linearities, a convenient analysis tool is the Harmonic Balance Method in which
the basic assumption is that the response is dominated by the fundamental
harmonic component. The steady-state response is considered to be a single
harmonic at the excitation frequency, neglecting sub-harmonics or supra-
harmonics. The forced response is studied only in the neighbourhood of the so-
called principal resonance. Similar results are obtained using the Equivalent
Linearization Method and the concept of Describing Functions.
The non-linear restoring force function is approximated by equivalent
spring and damper forces. For linear systems, the restoring force function (minus
sign omitted) may be taken of the form
f R ,lin (x , x& , t ) = c x& ( t ) + k x ( t ) ,

containing the contribution of a viscous damper and a linear spring.


For slightly non-linear systems, the restoring force can be expressed as
f R (x , x& , t ) ≅ ceq x& ( t ) + keq x ( t ) ,
80 MECHANICAL VIBRATIONS

where ceq is the equivalent viscous damping coefficient, and k eq is an equivalent


stiffness. An equivalent structural damping model may be used as well.

Cubic Stiffness
For systems without pre-loading or clearance, the elastic force may be
represented by a cubic stiffness law

(
fe = k x + μ x3 , ) (3.1)
where k is the slope of the stiffness function at the origin and μ is a coefficient of
non-linearity, positive for hardening and negative for softening springs.
For a harmonic displacement
x ( t ) = a cos ω t , (3.2)
( )
the elastic force (3.1) becomes f e = k a cos ω t + μ a cos ω t .
2 3

3 1
Substituting cos3ω t = cos ω t + cos 3ω t , and neglecting the higher
4 4
harmonic term in cos 3ω t , yields
⎛ 3 ⎞
f e ≅ k a ⎜ cos ω t + μ a 2 cosω t + ....⎟ = keq x ,
⎝ 4 ⎠
where the equivalent stiffness is
⎛ 3 ⎞
keq = k ⎜1 + μ a 2 ⎟ . (3.3)
⎝ 4 ⎠
The dynamic response of the non-linear system is obtained substituting this
amplitude-dependent stiffness into the equations obtained for the linear system.

Non-Linear Damping
Non-linear damping may be studied using the equivalent viscous damping
concept. This involves the approximation of a non-linear damping force by an
equivalent linear viscous damping force. The criterion for equivalence is that the
energy Wd dissipated per cycle of vibration by the non-linear damping element be
equal to the energy π ω ceq a 2 dissipated by an equivalent viscous damper
experiencing the same harmonic relative displacement.
Similar analysis can be carried out based on a concept of equivalent
structural damping when Wd = π heq a 2 .

Steady-state solutions obtained with the assumption of equivalent viscous


damping have been shown to be identical to those obtained by use of the averaging
3. SIMPLE NON-LINEAR SYSTEMS 81

method of Ritz. The same value is obtained using the coefficient of the first term of
a Fourier series expansion of the non-linear damping force time history.
A generalized non-linear damping force can be described mathematically
as being proportional to the nth power of the relative velocity across the damper

sgn ( x& ) ,
n
f d = cn x& (3.4)

where cn is defined as the velocity-nth power damping coefficient.

The exponent n and the damping coefficient cn are determined according


to the nature of the damping element. A value of n = 0 represents a Coulomb
damper where the damping coefficient c0 equals the dry-friction force R .
Similarly, values of n = 1 and n = 2 represent viscous and quadratic damping,
where cn becomes the damping coefficients c and c2 , respectively.
Other values of the exponent n may be selected to represent other non-
linear damping characteristics. For example, the damping developed in a car shock-
absorber may be described by a value of the exponent n between 2.0 and 3.0 ,
depending on the particular system configuration.

3.2 Cubic Stiffness

A single-degree-of-freedom system with a non-linear spring and a linear


structural damping element is shown in Fig. 3.1. A cubic law with positive
coefficient will be taken as a first approximation for describing a hardening spring
characteristic.

Fig. 3.1

Structural damping is a linear damping phenomenon for which the


damping force is in phase with the relative velocity but is proportional to the
relative displacement across the damper.
82 MECHANICAL VIBRATIONS

3.2.1 Harmonic Response

If the mass is acted upon by a harmonic force of constant amplitude F0


and frequency ω , the Duffing-type equation of motion of the vibrating mass may
be written as
h
( )
m &x& + x& + k x + μ x 3 = F0 e iω t ,
ω
(3.5)
where h = g k , and g can be an equivalent structural damping factor.
A first harmonic approximation of the response is chosen of the form

x = a~ e iω t = ( a R + i a I ) e iω t = a e i (ω t +θ ) . (3.6)
Using the method of harmonic linearization, the higher harmonic terms are
neglected, so that it is considered that
3
x3 ≅ a 2 x . (3.7)
4
Substitution of (3.6) and (3.7) into (3.5) yields the real and imaginary
components of displacement
2
⎛ 3 ⎞ k ⎛ g k a2 ⎞
a R = ⎜1 + μ a 2 − η 2 ⎟ a 2 = m a − ⎜⎜
2 ⎟ ,
⎟ (3.8)
⎝ 4 ⎠ F0 ⎝ F 0 ⎠
k 2
aI = − g a , (3.9)
F0
where
ω k
η= , ωn = . (3.10)
ωn m
The displacement magnitude

a = a R2 + a I2 (3.11)
is implicitly given by
3 F02
η 2 = 1 + μ a2 ± 2 2
− g2 . (3.12)
4 k a

The phase angle is calculated from


g g
tan θ = = . (3.13)
3 F02
η −1 − μ a2
2
± − g2
4
k 2 a2
Elimination of a and η 2 between equations (3.8) and (3.9) yields the
locus of the end of the vector a~ in the Argand plane which is a circle of equation
3. SIMPLE NON-LINEAR SYSTEMS 83

2 2
⎛ 1 F0 ⎞ ⎛ 1 F0 ⎞
a R2 + ⎜⎜ a I + ⎟⎟ = ⎜⎜ ⎟⎟ . (3.14)
⎝ 2g k ⎠ ⎝2g k ⎠
It is identical to equation (2.82) derived for linear systems.
Elimination of a between equations (3.9) and (3.11) yields the frequency
dependence of the quadrature component a I of response

η 2 = 1+ μ
3 F0
(− aI ) ± g F0 1
−1 . (3.15)
4 gk g k (− aI )

Similarly, the frequency dependence of the in-phase component a R of the


response may be obtained under the form
2
⎛ F ⎞ F
2 3μ ⎜⎜ 0 ⎟⎟ aR2 + 0 aR
⎛ F ⎞ ⎝ 2g k ⎠ k
η 2 = 1 + 3μ ⎜⎜ 0 ⎟⎟ − . (3.16)
⎝ 2g k ⎠ 2⎡ 2 ⎤
⎛ F ⎞ ⎛ 2g k ⎞ 2 ⎥
2 ⎜⎜ 0 ⎟⎟ ⎢ 1 ± 1 − ⎜⎜ ⎟⎟ aR
⎝ 2g k ⎠ ⎢⎢ ⎝ F0 ⎠ ⎥
⎣ ⎦⎥

3.2.2 Frequency Response Characteristics

Based on equation (3.12), the magnitude-frequency curves are illustrated in


Fig. 3.2 for a fixed value of damping g = const. and several values F0 of the
amplitude of harmonic force.
The response curves are symmetrically disposed with respect to the
“skeleton curve” of equation
⎛ 3 ⎞
ω 2 = ωn2 ⎜ 1 + μ a 2 ⎟ (3.17)
⎝ 4 ⎠
which passes through the points of maximum amplitude. For linear systems it is a
vertical line of abscissa ω = ωn . For systems with hardening stiffness (μ > 0) , the
skeleton curve is bent towards higher frequencies. For systems with softening
stiffness (μ < 0) , it is bent towards lower frequencies.
The locus of the points of vertical tangency of the magnitude-frequency
curves is given by equation
3 9 2 4
η 2 = 1+ μ a2 ± μ a − g2 (3.18)
4 16
and defines the “stability boundary” XLKY (Fig. 3.2).
84 MECHANICAL VIBRATIONS

Points within the region whose limits are marked by this curve define unstable
regimes of vibration. On the response curves they are drawn with broken lines.

Fig. 3.2 Fig. 3.3

The same information is contained in Fig. 3.3 where force-displacement


curves are plotted at several frequencies for g = const. They are called constant-
frequency lines or isochrones. The locus of the points of horizontal tangency of
these curves defines the stability boundary. Point K defines the maximum force
amplitude for which vibrations are stable irrespective of the displacement
magnitude. Point L defines the maximum displacement amplitude for which
vibrations are stable irrespective of the force level.

Fig. 3.4

The phase-frequency curves of Fig. 3.4 are based on equation (3.13).


3. SIMPLE NON-LINEAR SYSTEMS 85

Again, a stability boundary XLKY can be defined, of equation


⎛ g 3 ⎞
η 2 = 1+ ⎜⎜ tan θ + ⎟ (3.19)
⎝2 tan θ ⎟⎠
which is the locus of the points of vertical tangency of the phase-frequency curves.

The frequency response curves of the coincident (real) component are


represented in Fig. 3.5, based on equation (3.16). In this case, the equation of the
stability boundary XKY is
g 2 1 9μ 2
η2 = 1+ + aR (3.20)
3μ aR2 4
and is valid for aR < 0 .

Fig. 3.5 Fig. 3.6

The frequency response curves of the quadrature (imaginary) component


are illustrated in Fig. 3.6, based on equation (3.15). The stability boundary XKY is
given by the equation
3 g2
η 2 = 1 + μ aI2 + (3.21)
4 μ aI2
and holds for aR < 0 .
The best way to represent the frequency response data is to plot the vector
components of the displacement a R and a I , equations (3.8) and (3.9), on an
Argand diagram, as in Fig. 3.7. The result is a family of circles of equation (3.14).
86 MECHANICAL VIBRATIONS

Polar plots have the advantage of combining the information about


magnitude, phase and forcing frequency on a single diagram. At the same time, the
region of interest in the neighbourhood of the principal resonance is enlarged and
both types of “jump” phenomena can be easier explained.
For non-linear systems it is important to plot response displacement
components and not receptances (displacement/force) or other FRFs. Each point
represented in the complex plane is defined by two parameters – the excitation
frequency, ω , and the amplitude F0 of the input force. Consequently, two sets of
response loci have to be drawn, i.e. isochrones - connecting constant-frequency
points, and Nyquist plots - connecting points of constant excitation level. Since
both the displacement amplitude and the phase angle are sensitive to the force
amplitude, the distortion of these curves may be employed to provide an indication
of the non-linear behaviour. The phase is more sensitive to non-linearities than the
amplitude.

Fig. 3.7

In Fig. 3.7, isochrones are drawn by broken lines. Their equation is


obtained by eliminating F0 between equations (3.8) and (3.9). This yields

⎡ 3
( ) ⎤ a
a R = ⎢η 2 − 1 − μ a R2 + a I2 ⎥ I .
⎣ 4 ⎦ g
(3.22)

For μ = 0 , i.e. for linear systems, equation (3.22) describes straight lines
diverging from the origin of coordinates. For μ ≠ 0 , equation (3.22) describes
curves passing through the origin, more distorted as F0 increases. As F0 grows,
the isochrones are so much bent that they become tangent to the response curves.
3. SIMPLE NON-LINEAR SYSTEMS 87

The locus of the tangency points of Nyquist plots with the isochrones
defines the stability boundary XLKY. This is a hyperbola of equation
2g
aR aI = (3.23)

(defined only for aR < 0 , aI < 0 ), which is symmetrical with respect to the
bisector aR = aI of the coordinate axes.

The stability boundary XLKY intersects the bisector aR = aI at the point L,


of frequency ω L = ω n 1 + 2 g , at a distance a L = 4 g 3μ from the origin. For
lower frequencies or smaller displacement amplitudes jump phenomena cannot
occur.
Point K, where the stability boundary is tangent to the polar plot of
parameter F0 = k 32 g 3 (9 3μ ) and to the isochrone of parameter
ω K = ωn 1 + 3 g , indicates the lowest force level and forcing frequency at
which unstable vibrations can occur. It corresponds to a phase angle θ K = −120 0 .
Points K and L correspond to those marked on the diagrams of Figs. 3.5 - 3.7.

Fig. 3.8

The effect of non-linear stiffnesses is a shift of frequencies along the


circular Nyquist plots, clockwise - for softening springs, and anti-clockwise - for
hardening springs. The main resonance frequency is no more at maximum response
amplitude, as shown in Fig. 3.8. For equal frequency increments, the maximum
spacing between successive points is no more at the principal resonance, so that the
Kennedy-Pancu criterion cannot be used for resonance location.
88 MECHANICAL VIBRATIONS

Introducing a new dimensionless coefficient of non-linearity


γ = μ F02 g 3k 2 , it means that for either γ > γ K = 32 9 3 or ω > ω K the stability
boundary crosses the Nyquist plots and the isochrones. We may consider systems
with weak non-linearities those for which γ < γ K , and systems with strong non-
linearities those for which γ > γ K . According to this classification, weak non-
linearities will produce only a frequency shift along the response curves, while
strong non-linearities will produce jump phenomena.

Fig. 3.9

Generally, for single-degree-of-freedom systems with cubic stiffness and


structural damping, Nyquist plots are circles as in the case of linear systems, and
only the “vertex” shape of isochrones denotes the non-linear behaviour. They are
bent anti-clockwise for hardening stiffness, and clockwise, in the case of softening
3. SIMPLE NON-LINEAR SYSTEMS 89

stiffness (Fig. 3.9). Usually all isochrones are curved in the same way and there is
not a straight isochrone at the principal resonance.

3.2.3 Jump Phenomena

It is now possible to consider the jump phenomena in some detail. One


kind of jump phenomenon occurs when the forcing frequency is changed while
keeping constant the amplitude F0 of the excitation force (Fig. 3.10, a).
As the frequency is gradually increased from rest, the displacement
amplitude increases, the end of the response vector follows the portion BF of the
respective polar plot until the stability boundary is reached at point C. Then, there
is a “jump” in amplitude and phase from C to D, along the isochrone ωu = const . ,
followed by a continuous change in amplitude along the arc DO of the polar plot.
When the frequency is decreased, the end of the response vector moves
along ODE and FB on the Nyquist plot, and jumps from E to F along the isochrone
ωl = const . The arcs BF 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. This means that, in the
presence of strong non-linearities, large portions of the response curve cannot be
experimentally obtained.

a b
Fig. 3.10

Another jump phenomenon may be noticed when the amplitude of the


exciting force is varied while the forcing frequency is constant (Fig. 3.10, b).
When the force amplitude F0 is gradually increased, the response
amplitude increases. The end of the displacement vector moves along the
respective isochrone ( arc OVS ) until the stability boundary is reached at point S.
90 MECHANICAL VIBRATIONS

It jumps to the point T, following the response curve F0′′ = const . , and then moves
again along the isochrone ( arc TZ ).
On decreasing the force amplitude, the end of the response vector moves
along the portion ZTU of the isochrone until the stability boundary is reached at
point U, when it jumps to the point V, following the response curve F0′ = const.
and then moves again along the isochrone from V to O.

a b
Fig. 3.11

The same jump phenomena are represented in Fig. 3.11, a on a


displacement-frequency diagram, and in Fig. 3.11, b on a force-displacement
diagram, using the same notations for the points.

3.2.4 Parameter Estimation

One method for determining the system dynamic parameters requires at


least two response circles (Fig. 3.12), plotted for different amplitudes F0′ and
F0′′ = f ⋅ F0′ ( f > 1) of the harmonic excitation force.
Let denote by
F0′ F0′′
OM ' = = a1 , OM ′′ = = a2 ,
gk gk
the maximum displacement amplitudes on the two response circles.
The above equations yield
F0′′ a 2
f = = . (3.24)
F0′ a 1
3. SIMPLE NON-LINEAR SYSTEMS 91

For linear systems, points M ' and M " have the same frequency
ω n = k m . At systems with non-linear stiffness characteristic, the frequencies
ω ′r and ω ′r′ of these points are given by

3 3
ω ′r = ωn 1 + μ a 21 , ω ′r′ = ωn 1 + μ a 22 . (3.25)
4 4

Fig. 3.12

Using equations (3.24), from (3.25) we obtain the natural frequency of the
corresponding linear system
f 2ω ′r2 − ω ′r′2
ω 2n = (3.26)
f 2 −1
and the coefficient of non-linearity

4 ω ′r′2 − ω ′r2
μ= . (3.27)
3 a 21 f 2ω ′r2 − ω ′r′2

An arc of circle of radius OM ' = a 1 and centre at the origin O crosses the
second circle at the points P and Q, of frequencies ω P and ω Q , respectively,
given by
⎛ 3 ⎞
ω 2P ,Q = ω 2n ⎜ 1 + μ a 21 m g f 2 −1⎟ . (3.28)
⎝ 4 ⎠
The structural damping factor is calculated from

1 ω 2Q − ω 2P ω 2n
g= . (3.29)
f 2 − 1 ω Q + ω P ω ′r
2 2 2
92 MECHANICAL VIBRATIONS

The stiffness and mass parameters are then given by

k = F0′ g a 1 , m = k ω 2n . (3.30)

It is recommended to draw several Nyquist plots on the same diagram. The


adopted model is valid only if the same values are obtained for the dynamic
parameters regardless of the pair of polar plots used.

3.3 Combined Coulomb and Structural Damping

Coulomb damping results from the relative motion of two bodies sliding
one upon the other in the presence of a normal force N holding them in contact.
The friction force R is proportional to N, the proportionality constant being the
coefficient of friction. Usually, the difference between the static and dynamic
values of the coefficient of friction is neglected, and the normal force N is assumed
constant and independent of frequency and (relative displacement) amplitude.

Fig. 3.13

The Coulomb damping force can be written


x&
fd = R = R sgn (x& ) (3.31)
x&

so that it is + R or − R depending on whether the relative velocity is positive or


negative.
The energy dissipated per cycle of vibration by a Coulomb damper in
which a force f d acts through a relative displacement

x ( t ) = a sin ω t (3.32)
is given by
T π 2
Wd = ∫ f d dx = ∫ f d x& dx = 4 a ∫ f d cosω t d (ω t ) = 4 R a (3.33)
0 0
3. SIMPLE NON-LINEAR SYSTEMS 93

and is independent of frequency.


There is practical evidence that the dynamic behaviour of bolted or riveted
structures and systems incorporating hydraulic actuators can approximately be
described by a combined structural and dry-friction model.
The energy dissipated per cycle by a hysteretic damper (2.65) is

Wd = π h a 2 . (3.34)
In order to approximate the non-linear Coulomb damping by a linear one,
the concept of “equivalent structural damping” may be used. The equivalent
structural damping coefficient heq has such a value that the energy dissipated in a
harmonic displacement cycle of a given amplitude and frequency is the same as the
energy loss of the Coulomb damper, in the same displacement cycle. Whence,
equating (3.33) and (3.34) yields

π heq a 2 = 4 R a ,
wherefrom
4R
heq = (3.35)
πa
which is amplitude dependent.

3.3.1 Harmonic Response

The equation of motion for a single-degree-of-freedom system with


combined structural and Coulomb damping (Fig. 3.13) and excited by a harmonic
force may be written

x& + R sgn (x& ) + k x = f (t ) .


h
m &x& + (3.36)
ω
A first harmonic approximation of the response is chosen of the form
(3.32)
x (t ) = a sin ω t
and the force is conveniently expressed as
f (t ) = F0 sin (ω t − θ ) . (3.37)
Upon substitution of (3.37) and use of the concept of equivalent structural
damping, equation (3.36) becomes
h + heq
m &x& + x& + k x = F0 sin (ω t − θ ) . (3.38)
ω
94 MECHANICAL VIBRATIONS

Substituting the solution (3.32) into equation (3.38) and equating the
coefficients of the terms in cos ω t and sin ω t of both sides, the following
equations are obtained
sin θ = − a g − r ,
(
cos θ = a 1 − η 2 , ) (3.39)

where
4R ak ω k h
r= , a= , η= , ωn = , g= . (3.40)
π F0 F0 ωn m k
Equations (3.39) give the magnitude of the dimensionless displacement

a=
−gr+ (1 − r )( 1 − η )
2 2 2
+ g2
(1 −η ) + g
(3.41)
2 2 2

and the phase angle


r
g+
θ = tan -1 2 a . (3.42)
η −1
A solution for a is only possible when r < 1 (F0 > 4 R π ) .
Equations (3.41) and (3.42) can be directly obtained from (2.66) and (2.67)
if the structural damping factor g is replaced by g + g eq = g + r a .

Considering a dimensionless complex displacement amplitude

a~ k
= a R + i a I = a cosθ + i a sinθ (3.43)
F0
we obtain

( )
aR = a 2 1 − η 2 ,
⎛ r⎞
aI = a 2 ⎜ g + ⎟ ,
⎝ a⎠
(3.44)

where the magnitude a = a R2 + a I2 is given by equation (3.41).

3.3.2 Nyquist Plots and Isochrones

Based on equations (3.41) and (3.42), the polar diagram of the frequency
response for the system of Fig. 3.13 can be drawn in the Argand plane. This
Nyquist plot has the pear-shape illustrated in Fig. 3.14. It shows that the
3. SIMPLE NON-LINEAR SYSTEMS 95

elongation, parallel to the imaginary axis, of the otherwise circular diagrams of


lightly damped linear systems, is the effect of Coulomb damping.
The principal resonance frequency ωn may be determined as the parameter
of point M, where the diagram crosses the imaginary negative semi-axis. The
displacement amplitude at resonance is

aM = OM =
1
(1 − r ) . (3.45)
g

Fig. 3.14

The lines OB and OC, drawn from the origin at angles of 450 each side of
the imaginary negative semi-axis, cross the Nyquist plot at points B and C, of
frequencies
g
ω 1, 2 = ωn 1m . (3.46)
1− 2 r

The corresponding dimensionless vector lengths are

1 ⎛ 2 ⎞
OB = OC = ⎜ ⎟
g ⎜ 2 − r⎟
⎝ ⎠
so that
BC = 2 OB =
1
g
(
1− 2 r ) (3.47)

which may be used for determining the Coulomb friction force.


It is recommended to plot several response diagrams for different values of
the excitation force amplitude F0 , using as coordinates the coincident and
96 MECHANICAL VIBRATIONS

quadrature components of the actual displacement (Fig. 3.15). If isochrones are


drawn on the same diagram, they exhibit a “fir tree” pattern. The only straight line
is the isochrone corresponding to phase resonance. Isochrones lying in the real
positive half-plane are bent anticlockwise to the right, while those lying in the real
negative half-plane are bent clockwise to the left. The pattern of isochrones can be
used to recognize the presence of Coulomb damping in a system. Their distortion
from straightness is a less ambiguous indication of non-linear behaviour than the
deviation from circularity of Nyquist plots.

Fig. 3.15

3.3.3 Parameter Estimation

The method of two polar plots can be applied to systems with non-linear
damping too. Consider two Nyquist plots drawn for two distinct values F0′ and
F0′′ = f ⋅ F0′ ( f > 1) of the harmonic excitation force (Fig. 3.16).
The resonance points M ' and M " , of frequency ωn , are located at the
crossing points of the plots with the only isochrone which is a straight line.
Denoting α = OM " OM ' , the following equation can be established for
the calculation of the Coulomb friction force

π F0′ α − f
R= . (3.48)
4 α −1

If an arc of circle of radius a1 = OM ' and centre in the origin O is drawn,


it crosses the second polar plot at points P and Q, of frequencies given by
3. SIMPLE NON-LINEAR SYSTEMS 97

⎛ ⎞
⎜ ⎟
2
− ⎛ f + 1 ⎞⎟
= ωn2 ⎜ 1 m ⎟ = ω2
g f 1
ω 2P ,Q ⎜1 m g (α − 1) . (3.49)
⎜ 4R ⎟ n ⎜ f − 1 ⎟⎠
⎜ − ⎝
π F0′ ⎟⎠
1

The structural damping factor g can be calculated from

ω 2Q − ω 2P 1 f −1
g= (3.50)
ω 2Q + ω 2P α −1 f +1
where

ω 2P + ω 2Q = 2 ω 2n . (3.51)

Fig. 3.16

The stiffness is further given by


F0′ ⎛ 4R ⎞
k= ⎜⎜1 − ⎟⎟ . (3.52)
g a1 ⎝ π F0′ ⎠

and the mass by m = k ω n2 .

3.4 Quadratic Damping


Quadratic damping is a particular case of velocity-nth power damping
which can be associated with the turbulent flow of a fluid through an orifice. It
practically occurs at relatively high flow velocities, but can be used as a first
98 MECHANICAL VIBRATIONS

approximation in describing a certain class of non-linear damping phenomena


arising in systems including oil dampers and shock absorbers.
The damping force f d is proportional to the square of the relative velocity
across the damper (minus sign omitted)

f d = c2 x& 2 sgn ( x& ) , (3.53)

where c2 is defined as the quadratic damping coefficient, being a function of


damper geometry and fluid properties.
The energy dissipated per cycle of vibration by a quadratic damper in
which a force f d acts through a relative displacement x (t ) = a sin ω t is given by
T π 2
f d cosω t d(ω t ) = c2 ω 2 a 3 , (3.54)
8
Wd = ∫ f d dx = ∫ f d x& dx = 4 a ∫ 3
0 0

being dependent upon the frequency and amplitude of vibration.

3.4.1 Harmonic Response

The equation of motion for a single-degree-of-freedom system with


directly coupled quadratic damping (Fig. 3.17) and excited by a harmonic force
acting upon the mass may be written

m &x& + c2 x& 2 sgn (x& ) + k x = f (t ) . (3.55)

A first harmonic approximation of the response is chosen of the form


x (t ) = a sin ω t and the force is conveniently expressed as f (t ) = F0 sin (ω t − θ ) .

Fig. 3.17

Upon substitution in (3.55), the following equations are obtained


3. SIMPLE NON-LINEAR SYSTEMS 99

sin θ = − a 2
8

α2 η 2 , (
cos θ = a 1 − η 2 , ) (3.56)

in which
c2 F0 ak ω k
α2 = , a= , η= , ωn = , (3.57)
km F0 ωn m

and where α 2 is defined as the quadratic damping parameter.


Equations (3.56) give the dimensionless frequency η and the phase shift
θ as functions of the dimensionless displacement amplitude a :

2 2
1 ⎛ 8α 2 ⎞ ⎛ 8α 2 ⎞
1± − a2 ⎜⎜ ⎟⎟ + ⎜⎜ ⎟⎟
⎝ 3π ⎠ ⎝ 3π
2
a ⎠
η =
2
2
, (3.58)
⎛ 8α ⎞
1 + a 2 ⎜⎜ 2 ⎟⎟
⎝ 3π ⎠

8α 2 2
a η

θ = tan -1 . (3.59)
η 2 −1

Considering a complex displacement amplitude

a~ = aR + i aI = a cosθ + i a sinθ
we obtain

aR = a 2
k
F0
(
1 −η 2 , ) aI = − a 3
k 8 c2 2
F0 3 π m
η , (3.60)

where the magnitude a = a R2 + a I2 is given by

F0
2
a= k . (3.61)

(1 − η ) (1 − η )
2 2
2 2 2 4 ⎛ 16 c2 ⎞ ⎛ F0 ⎞ 4
+ + ⎜⎜ ⎟⎟ ⎜ ⎟ η
⎝ 3π m ⎠ ⎝ k ⎠
100 MECHANICAL VIBRATIONS

3.4.2 Nyquist Plots and Isochrones

Based on equations (3.60) and (3.61), the polar diagram of the frequency
response for the system of Fig. 3.17 can be plotted in the Argand plane as the
geometric locus of the affix of the vector a~ .
Figure 3.18 shows three such diagrams. The quadratic damping tends to
elongate the polar plots in the direction of the real axis, “flattening” the otherwise
circular diagrams of systems with slight viscous damping. The isochrones exhibit a
“fireworks” pattern. The only straight line is at phase resonance. Isochrones
defined by under-resonance frequencies are bent up-frequency, while those defined
by over-resonance frequencies are curved down-frequency.

Fig. 3.18

Analysing in more detail the polar plot in Fig. 3.19, it can be seen that the
phase resonance frequency ωn is determined at the point M, where the diagram
crosses the imaginary negative semi-axis.

Fig. 3.19
The displacement amplitude at phase resonance is
3. SIMPLE NON-LINEAR SYSTEMS 101

F0 3π 3π m
aM = OM = = F0 . (3.62)
k 8α 2 8 k c2

The lines OB and OC, drawn from the origin at angles of 450 on each side
of the imaginary negative semi-axis, cross the Nyquist plot at points B and C, of
frequencies

η B ,C = 1+ γ 2 m γ , (3.63)
where
2
γ2 = α2 (3.64)

which may be used for the evaluation of damping.

3.4.3 Parameter Estimation

The method of two polar plots can be used for the estimation of system
parameters. Figure 3.20 illustrates two Nyquist plots drawn for two distinct values
F0′ and F0′′ = f ⋅ F0′ ( f > 1) of the excitation force amplitude. The points M ' and
M " , of frequency η = 1 , are the crossing points of the plots with the only
isochrone which is a straight line.

Fig. 3.20

It is useful to define a new dimensionless damping parameter


8α 2 8 c2 F0
α= = (3.65)
3π 3π k m
102 MECHANICAL VIBRATIONS

that is a function of the amplitude F0 of the excitation force.


For the two polar plots of Fig. 3.20, one can write
8 c2 8 c2
α′ = F0′ , α ′′ = F0′′ (3.66)
3π k m 3π k m
so that the ratio of the excitation force amplitudes can be denoted

F ′′ α ′′ ⎛ OM " ⎞
2
f = 0 = =⎜ ⎟ . (3.67)
F0′ α ′ ⎝ OM ' ⎠

Fig. 3.21
3. SIMPLE NON-LINEAR SYSTEMS 103

If an arc of circle of radius a 1 = OM ' and centre in the origin O is drawn,


it will cross the second polar plot at the points P and Q, of dimensionless
frequencies η P and η Q , given by

η 2P ,Q =
1m ( )
α ′ f 2 − 1 + f 2α ′2
. (3.68)
1+α′
Equations (3.68) yield the damping parameter
2
α′ = −1 . (3.69)
η 2P + η 2Q

The quadratic damping factor α 2′ is given by


α 2′ = α′ . (3.70)
8
The stiffness is further obtained from
F0′ 1
k= . (3.71)
a1 α′
and the quadratic damping coefficient from
α 2′ k 2
c2 = . (3.72)
ωn2 F0′
Figure 3.21 summarizes some of the effects of Coulomb and quadratic
damping on the Nyquist plots and the isochrones of single-degree-of-freedom
systems.

3.5 Effect of Pre-Loading


The stiffness function (3.1) is anti-symmetric with respect to the origin. It
can be easily shown that, replacing x by x + x0 , where x0 is the deflection
produced by pre-load, the stiffness function is represented by three terms

(
f e = k0 x + μ 0 x 2 + μ x 3 . ) (3.73)

The square term due to pre-load has a softening effect irrespective if


whether μ 0 is positive or negative. A positive cubic term can induce a hardening
effect at larger displacement amplitudes. In this case, isochrones plotted in the
Argand plane are represented by double bend curves.
104 MECHANICAL VIBRATIONS

The analysis is complicated by the fact that the first approximation solution
must contain also a constant term, because the system vibrates about a new
equilibrium position which is displaced from the stiffness equation origin. It is also
very difficult in this case to evaluate the constants k, μ 0 and μ from experimental
results.
Double-bend behaviour of isochrones as an indication of pre-load has been
noticed at machine-tools with slackened guides and at systems with loose joints.
Similar patterns of isochrones exhibit polyurethane foam isolator pads, but
this is determined by the force-deflection characteristic, which is softening at low
force levels, then hardening, at larger displacements.
4.
TWO-DEGREE-OF-FREEDOM SYSTEMS

The degrees of freedom of a vibrating system are equal to the number of


independent coordinates required to describe its motion completely.
In this chapter, two-degree-of-freedom systems are analyzed as the
simplest case of discrete systems, and as an introduction to systems with a larger
number of degrees of freedom. As a preparation for multi-degree-of-freedom
systems, the matrix notation is introduced as a compact way of expressing the
equations of motion and the dynamic response.
In Chapter 2 it is shown that the free vibration of an undamped single-
degree-of-freedom system is a harmonic motion at the system natural frequency,
referred to as a natural vibration. In contrast, the free vibration of an undamped
multi-degree-of-freedom system is periodic, and consists of several simultaneous
vibrations at the various natural frequencies. It can be expressed as a sum of
harmonic components, each implying a certain natural frequency and displacement
configuration, called natural modes of vibration. The motion in a natural mode of
vibration is synchronous and harmonic at all system coordinates.
The dynamic response of a discrete system can be described by
simultaneous ordinary differential equations. For a proper choice of coordinates,
known as principal or modal coordinates, the equations can be decoupled and
solved independently. The modal coordinates represent linear combinations of the
actual displacements. Conversely, the motion can be regarded as a superposition of
vibrations in the natural modes of vibration defined by the modal coordinates.
A two-degree-of-freedom system has two natural frequencies and, in
forced vibrations, for small damping it may have two resonances. The free
response to initial excitation and the forced response to external excitation can be
expressed in terms of the natural modes of vibration whose shapes are defined by
mutually orthogonal vectors with respect to the mass and stiffness matrices. The
forced response to harmonic excitation can also be calculated by simply using
Cramer’s rule. Resonance occurs when the forcing frequency equals any of the two
natural frequencies of the system. Antiresonance can also occur at a frequency
equal to the natural frequency of a mass-spring subsystem.
106 MECHANICAL VIBRATIONS

The material contained in this chapter refers primarily to the computation


of natural frequencies and mode shapes. The forced vibrations are studied first for
undamped systems using both the modal analysis and the direct spectral analysis
techniques. Then damped forced vibrations are considered.

4.1 Coupled Translation


In the following, two-degree-of-freedom mass-spring systems are
considered, in which the lumped masses have unidirectional translational motions.

4.1.1 Equations of Motion

Consider the system of Fig. 4.1, a which consists of two masses m1 and
m 2 , attached to fixed points by springs k1 and k 3 , and tied together by a
“coupling spring” k 2 .

Fig. 4.1

Assuming that masses are guided so as to move only in the horizontal


direction, the configuration is entirely determined by their instantaneous
displacements x1 and x 2 from the equilibrium positions. The system has two
degrees of freedom.
Using the free-body diagrams of Fig. 4.1, b and d’Alembert’s principle
(dynamic equilibrium of impressed and inertia forces), the equations of motion can
be written
( )
m 1 &x&1 + k 1 x 1 + k 2 x 1 − x 2 = 0 ,
m 2 &x& 2 + k 3 x 2 − k 2 (x 1 − x 2 ) = 0 .
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 107

On rearranging, these equations become


( )
m 1 &x&1 + k 1 + k 2 x 1 − k 2 x 2 = 0,
m 2 &x& 2 − k 2 x 1 + ( k 3 + k 2 ) x 2 = 0.
(4.1)

This is a set of “coupled” linear differential equations of second order, with


constant coefficients. The coupling between the two coordinates is due to the
stiffness k 2 . If k 2 = 0 , equations (4.1) become independent, and the system of Fig.
4.1 degenerates into two one-degree-of-freedom systems.
Equations (4.1) can be written
⎡m 1 0 ⎤ ⎧ &x&1 ⎫ ⎡k1 + k 2 − k 2 ⎤ ⎧ x 1 ⎫ ⎧0⎫
⎢ 0 m ⎥ ⎨&x& ⎬ + ⎢ ⎨ ⎬ = ⎨ ⎬, (4.2)
⎣ 2 ⎦ ⎩ 2 ⎭ ⎣ − k2 k3 + k 2 ⎥⎦ ⎩ x 2 ⎭ ⎩0⎭

or in compact form
[ m ]{ &x& } + [ k ]{ x } = { 0 } , (4.3)

where [ m ] is the mass matrix, [ k ] is the stiffness matrix and { x } is the column
vector of displacements. Note that square matrices are denoted by brackets while
column vectors are denoted by braces. The mass and stiffness matrices are always
symmetrical so that they are equal to their transposes

[ m ] = [ m ]T , [ k ] = [ k ]T . (4.4)
The mass matrix is diagonal. The coupling is produced by the off-diagonal
elements of the stiffness matrix.

4.1.2 Free Vibration. Natural Modes

Let examine the conditions under which the two masses have synchronous
harmonic motions, i.e. when the system behaves like a single degree of freedom
system in natural vibration.
We assume solutions of the form
x 1 ( t ) = a 1 cos (ω t − ϕ ),
(4.5)
x 2 ( t ) = a 2 cos (ω t − ϕ ),

and examine the conditions under which such motion is possible. The motion we
are seeking is one in which the ratio between the two instantaneous displacements
remains constant throughout the motion
x1 (t ) x 2 (t ) = a1 a 2 = const . (4.6)
The shape of the system configuration does not change during the motion, the
deflected shape resembles itself at any time.
108 MECHANICAL VIBRATIONS

Substituting solutions (4.5) into the differential equations (4.1), the


resulting algebraic equations are

(k 1 + k2 )
− m 1 ω 2 a 1 − k2 a2 = 0 ,
(4.7)
( )
− k2 a 1 + k3 + k2 − m 2 ω 2 a2 = 0.
The simultaneous homogeneous equations (4.7) admit non-trivial solutions
if the determinant of the coefficients a1 and a2 is zero
k 1 + k2 − m 1 ω 2 − k2
= 0. (4.8)
− k2 k3 + k 2 − m 2 ω 2

This can be written


⎛ k1 + k 2 k3 + k 2 ⎞ 2 k1 k3 + (k1 + k3 ) k 2
ω4 − ⎜ + ⎟ω + =0 (4.9)
⎜ m1 m ⎟ m 1 m2
⎝ 2 ⎠

which represents a quadratic equation in ω 2 called the characteristic equation, or


frequency equation. Its roots ω 12 and ω 22 are real and positive. The quantities ω 1
and ω 2 are called natural frequencies or eigenfrequencies because they depend
solely upon the system mass and stiffness parameters.
Because the system (4.7) is homogeneous, the amplitudes a 1 and a 2
cannot be determined, only their ratio μ = a 2 a 1 .

For ω 1 , the amplitude ratio is


⎛ a2 ⎞ k + k − m2 ω 12
μ1 = ⎜ ⎟ = 3 2 (4.10)
⎜ a1 ⎟ k2
⎝ ⎠1

and for ω 2 , the amplitude ratio is


⎛ a2 ⎞ k3 + k 2 − m2 ω 22
μ2 = ⎜ ⎟ = . (4.11)
⎜ a1 ⎟ k2
⎝ ⎠2
The ratios (4.10) and (4.11) determine the shape of the system during
synchronous motion with frequencies ω 1 and ω 2 , respectively. If one element in
each ratio is assigned a certain arbitrary value, then the value of the other element
results from the above expressions. This process is called normalisation.
The natural frequencies and corresponding amplitude ratios determine the
conditions to have synchronous harmonic motions, i.e. the natural modes of
vibration. A mode of vibration is defined by two parameters: the natural frequency
and the mode shape. The mode shapes can be represented by column vectors
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 109

⎧1⎫
{ a }1 = ⎨
⎧a 1 ⎫
( )
⎬ = a 1 1 ⎨ ⎬ = C1 { u }1 ,
⎩a2 ⎭1 ⎩u 1 ⎭
(4.12)
⎧1⎫
{ a }2
⎧a ⎫
( )
= ⎨ 1 ⎬ = a 1 ⎨ ⎬ = C2 { u }2 ,
2 u
⎩a2 ⎭ 2 ⎩ 2⎭
referred to as modal vectors.
In (4.12) the modal vectors are normalised with the first element equal to
unity. The normalised vectors are said to represent the shape of a normal mode.
The two possible synchronous motions are given by
{ x (t ) }1 = C1 { u }1 cos (ω 1 t − ϕ 1 ),
{ x (t ) }2 = C2 { u }2 cos (ω 2 t − ϕ 2 )
(4.13)

and the general solution of free vibrations is


{ x (t ) } = { x (t ) }1 + { x (t ) }2 = C1 {u }1 cos (ω 1 t − ϕ 1 ) + C2 { u }2 cos (ω2 t − ϕ 2 ).(4.14)
In (4.14) the four integration constants C1 , C 2 , ϕ 1 , ϕ 2 are determined
from the four initial conditions, the displacements and velocities at t = 0 .
For arbitrary initial conditions, the free vibration of the two-degree-of-
freedom system is a periodic motion obtained as the superposition of the two
natural modes of vibration, i.e. two harmonic motions with frequencies equal to the
natural frequencies of the system. It can be shown that for zero initial velocities
and initial displacements resembling a mode shape, the free motion is synchronous
and purely harmonic, and takes place at the natural frequency of the respective
mode. A system can vibrate in a pure mode of vibration if the initial deflected
shape is similar to the mode shape.

Example 4.1
Consider the system of Fig. 4.2 and obtain the natural modes of vibration.

Fig. 4.2

Solution. The equations of motion can be written


110 MECHANICAL VIBRATIONS

⎡2m 0 ⎤ ⎧ &x&1 ⎫ ⎡ 3k − k ⎤ ⎧ x1 ⎫ ⎧0⎫


⎢ 0 m ⎥ ⎨ &x& ⎬ + ⎢− k ⎨ ⎬=⎨ ⎬.
k ⎥⎦ ⎩ x 2 ⎭ ⎩0⎭
⎣ ⎦ ⎩ 2⎭ ⎣
Assuming solutions of the form
{ x } = { u } cos (ω t − ϕ )
we obtain

(3 k − 2ω m) u − k u
2
1 2 = 0,
− k u + ( k − ω m) u
1
2
2 = 0,

or, dividing by k,
( 3 − 2α ) u 1 − u 2 = 0,
− u 1 + ( 1 − α ) u 2 = 0,

where

α = mω 2 k .
The condition to have nontrivial solutions is
3 − 2α −1
= 0, 2α 2 − 5α + 2 = 0
−1 1−α
with solutions
α1 = 1 2 , α2 = 2 .
The natural frequencies are
1 k k
ω1= , ω2 = 2 .
2 m m
Assigning the first element to be unity, the first modal vector is given by

⎡ 3 − 2 α1 − 1 ⎤ ⎧ u1 ⎫ ⎡ 2 − 1⎤ ⎧1⎫
⎢ −1 ⎨ ⎬ =⎢ 1 ⎥⎨ ⎬=0
1 − α 1 ⎥⎦ ⎩u 2 ⎭1 ⎢− 1 ⎥ 2
⎣ ⎣ 2 ⎦⎩ ⎭
and the second modal vector is calculated from
⎡ 3 − 2α 2 − 1 ⎤ ⎧ u1 ⎫ ⎡ − 1 − 1⎤ ⎧ 1 ⎫
⎢ −1 ⎥ ⎨ ⎬ =⎢ ⎨ ⎬=0.
⎣ 1 − α 2 ⎦ ⎩u 2 ⎭ 2 ⎣ − 1 − 1 ⎥⎦ ⎩− 1⎭

The mode shapes are graphically presented in Fig. 4.3. In the first mode,
the two masses are moving in the same direction, either both to the right, or both to
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 111

the left, the displacement of the mass m being always twice the displacement of the
mass 2 m . In the second mode, the two masses move in opposite directions
through the same distance. The midpoint of the spring k does not move, hence it is
a nodal point. If this point were clamped, no change in the motion would take
place. The second spring is split into two springs of stiffness 2k . The mass 2m is
thus connected to ground by two springs of stiffness 2k (in parallel) with an
equivalent stiffness 4k , while the mass m is connected to one spring of stiffness
2k , both subsystems having the same natural frequency equal to ω 2 .

Fig. 4.3
1 k
Summarising, the first normal mode has a natural frequency ω 1 =
2 m
1⎫
and a mode shape { u }1 = ⎧⎨ ⎬ , while the second normal mode has a natural
⎩2 ⎭
k ⎧1⎫
frequency ω 2 = 2 and a mode shape { u } 2 = ⎨ ⎬ . If the system is given
m ⎩− 1⎭
an initial disturbance of x 1 = 1 and x2 = 2 and then released, the ensuing motion
will be purely harmonic with the frequency ω 1 = 0.207 k m . If the initial
displacement is x 1 = 1 and x 2 = −1 , the motion will be harmonic with the
frequency ω2 = 1.414 k m .

4.1.3 Orthogonality of Natural Modes

Substituting solutions (4.13) in equation (4.3) we obtain

[ k ] { u }1 = ω 21 [ m ]{ u }1 , (4.15, a)
112 MECHANICAL VIBRATIONS

[ k ] { u }2 = ω22 [ m ]{ u }2 . (4.15, b)

Multiplying (4.15, a) on the left by { u }T2 we get

{ u }T2 [ k ] { u }1 = ω 12 { u }T2 [ m ]{ u }1 . (4.16)

Taking the transpose of (4.15, b) and multiplying on the right by { u }1 we


obtain
{ u }T2 [ k ] { u }1 = ω22 { u }T2 [ m ]{ u }1 . (4.17)

Subtracting (4.16) and (4.17) from each other, for ω 1 ≠ ω 2 , we get

{ u }T2 [ m ]{ u }1 = 0 , (4.18)

and taking the transpose


{ u }T1 [ m ]{ u }2 = 0 . (4.19)

Substituting (4.18) in (4.16) we obtain

{ u }T2 [ k ] { u }1 = 0 (4.20)
and taking the transpose
{u }1T [ k ] {u }2 = 0 . (4.21)

Equations (4.18)-(4.21) show that the modal vectors are orthogonal with
respect to the mass matrix and the stiffness matrix. Note the difference from
normal orthogonality of two vectors { a } and { b } , which is written { a }T { b } = 0.

4.1.4 Modal Coordinates

Let denote in (4.13)


( )
q1 (t ) = C1 cos ω 1 t − ϕ 1 ,
q2 (t ) = C2 cos (ω 2 t − ϕ 2 )
(4.22)

so that (4.14) becomes


{ x (t ) } = {u }1 q 1 + {u }2 q 2 . (4.23)

Equation (4.23) can be written


⎧ q1 ⎫
{ x (t ) } = [ { u }1 { u }2 ] ⎨ ⎬ = [ u ]{q } (4.23, a)
⎩q 2 ⎭
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 113

where
[ u ] = [ { u }1 { u }2 ] (4.24)

is called the modal matrix.

Substituting (4.23) in (4.3), multiplying on the left by { u }Tr (r = 1, 2 ) and


considering the orthogonality we obtain
M r q&&r + K r qr = 0 , (4.25)
where
M r = { u }Tr [ m ] { u }r , K r = { u }Tr [ k ] { u }r (4.26)

are modal masses and modal stiffnesses, respectively.


The values of modal masses and stiffnesses depend on the normalization of
modal vectors. When the values of the modal masses are imposed the vectors are
said to be mass-normalized. For unit modal masses, the modal stiffnesses are equal
to the square of the respective natural frequency.
From equations (4.15) we obtain the natural frequency in terms of the
modal vector
2 { u }Tr [ k ] { u }r
ωr = . r = 1, 2 (4.27)
{ u }Tr [ m ] { u }r
Rayleigh’s quotient is defined as

R ( {u } ) =
{ u }T [ k ] { u } . (4.28)
{ u }T [ m ] { u }
If the vector {u } coincides with one of the system modal vectors, then the
quotient reduces to the associated natural frequency squared. Rayleigh’s quotient
has a stationary value in the neighbourhood of a modal vector. If {u } is a trial
vector with differs slightly from the first modal vector, then R ( {u } ) is very close
to the fundamental natural frequency squared, and always higher.
The linear transformation (4.23) uncouples the equations of motion. The
coordinates (4.22) for which the equations of motion are independent are called
principal coordinates or modal coordinates.

Substituting (4.23, a) in (4.3) and multiplying on the left by [ u ] T yields

[ u ] T [ m ] [ u ] { q&& } + [ u ] T [ k ] [ u ]{ q } = { 0 } , (4.29)
or
[ M ] { q&& } + [ K ] { q } = { 0 }, (4.29, a)
114 MECHANICAL VIBRATIONS

where the diagonal matrices

[ M ] = [ u ]T [ m ] [ u ] and [ K ] = [ u ] T [ k ] [ u ] (4.30)
are the modal mass matrix and the modal stiffness matrix, respectively.
The coordinate transformation (4.23, a) simultaneously diagonalizes the
mass matrix and the stiffness matrix. After solving separately the decoupled
equations (4.29, a), the modal coordinates may be substituted back into (4.23, a) to
obtain the physical coordinates in the configuration space. This technique is called
modal analysis. The modal analysis uses a linear coordinate transformation based
on the modal matrix to uncouple the equations of motion of a vibrating system.

4.1.5 Response to Harmonic Excitation

Consider the forced vibrations of the system of Fig. 4.4 under the action of
forces f1 (t ) and f 2 (t ) applied on mass m1 , respectively m 2 .

Fig. 4.4

The equations of motion are


m 1 &x&1 + (k1 + k 2 ) x 1 − k 2 x 2 = f 1 ,
(4.31)
m 2 &x&2 − k 2 x 1 + (k3 + k 2 ) x 2 = f 2 ,

or, in compact matrix form,


[ m ] { &x& } + [ k ] { x } = { f } , (4.31, a)
where { f } is the column vector of impressed forces.
4.1.5.1 Solution by Modal Analysis
Substituting (4.23, a) in equation (4.31, a), and multiplying on the left by
[u ] T
we obtain
[ M ] { q&& } + [ K ] { q } = { F } , (4.32)
where
{ F } = [ u ]T { f } (4.33)
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 115

is the column vector of modal forces.


For harmonic excitation
{ f } = { f̂ } cos ω t (4.34)
the steady-state response is
{ x } = { x̂ } cos ω t , { q } = { q̂ } cos ω t , (4.35)
where the a hat above a letter denotes amplitude.
Substituting (4.34) and (4.35) in (4.32) we obtain

[− ω 2
[ M ]+ [ K ] ] { q̂ } = [ u ] T { f̂ }= {F̂ }, (4.36)

so that the amplitudes of modal coordinates are

F̂1 { u }T1 { f̂ }
q̂ 1 = = , (4.37)
K1 − ω 2 M 1 { u }T1 [ k ] { u }1 − ω 2 { u }T1 [ m ] { u }1

F̂ 2 { u }T2 { f̂ }
q̂ 2 = = . (4.38)
K2 − ω 2M 2 { u }T2 [ k ] { u }2 − ω 2 { u }T2 [ m ] { u }2
Equation (4.23) for amplitudes becomes

{ x̂ } = { u }1 q̂1 + { u }2 q̂2 . (4.39)

Substituting (4.37) and (4.38) in (4.39) we obtain the vector of amplitudes


in physical coordinates

{ u }1T { f̂ }{ u }1 { u }T2 { f̂ }{ u }2
{ x̂ } = + .
{ u }1T [ k ] { u }1 − ω 2 { u }1T [ m ] { u }1 { u }T2 [ k ] { u }2 − ω 2 { u }T2 [ m ] { u }2
(4.40)
The elements of { x̂ } are of the form

{ u }T1 {f̂ }u { u }T2 { f̂ }u


x̂ 1 = 2 11 + 2 12 , (4.41)
K1 − ω M 1 K2 − ω M 2

{u }T1 { f̂ } {u }T2 { f̂ }
x̂ 2 = u 21 + u 22 . (4.42)
K1 − ω 2 M 1 K2 − ω 2M 2
116 MECHANICAL VIBRATIONS

Example 4.2
Consider the system of Fig. 4.2 acted upon by harmonic forces
f1 = f̂1 cos ω t and f 2 = f̂ 2 cos ω t and obtain the steady-state response
amplitudes. Plot the frequency response curves of receptance when f 2 = 0 .
Solution. The equations of motion can be written
⎡2m 0 ⎤ ⎧ &x&1 ⎫ ⎡ 3k − k ⎤ ⎧ x1 ⎫ ⎧ f1 ⎫
⎢ 0 m⎥ ⎨ &x& ⎬ + ⎢− k ⎨ ⎬=⎨ ⎬.
⎣ ⎦ ⎩ 2⎭ ⎣ k ⎥⎦ ⎩ x 2 ⎭ ⎩ f 2 ⎭

Using the coordinate transformation


⎧1⎫ ⎧ 1⎫ ⎡ 1 1 ⎤ ⎧ q1 ⎫
{ x } = [ u ] {q } = ⎨ ⎬ q1 + ⎨ ⎬ q2 = ⎢ ⎥⎨ ⎬
⎩2⎭ ⎩− 1 ⎭ ⎣ 2 − 1 ⎦ ⎩q 2 ⎭
and multiplying on the left with the transpose of the modal matrix, we obtain
⎡1 2 ⎤ ⎡2m 0 ⎤ ⎡1 1 ⎤ ⎧ q&&1 ⎫ ⎡1 2 ⎤ ⎡ 3k − k ⎤ ⎡1 1 ⎤ ⎧ q1 ⎫ ⎡1 2 ⎤ ⎧ f1 ⎫
⎢1 − 1⎥ ⎢ 0 m ⎥ ⎢2 − 1⎥ ⎨q&& ⎬ + ⎢1 − 1⎥ ⎢− k ⎨ ⎬= ⎨ ⎬
⎣ ⎦⎣ ⎦⎣ ⎦⎩ 2⎭ ⎣ ⎦⎣ k ⎥⎦ ⎢⎣2 − 1⎥⎦ ⎩q2 ⎭ ⎢⎣1 − 1⎥⎦ ⎩ f 2 ⎭
or
⎡6m 0 ⎤ ⎧ q&&1 ⎫ ⎡3k 0 ⎤ ⎧ q1 ⎫ ⎧ f1 + 2 f 2 ⎫ ⎧ F1 ⎫
⎢ 0 3m⎥ ⎨q&& ⎬ + ⎢ 0 ⎨ ⎬=⎨ ⎬=⎨ ⎬.
6k ⎥⎦ ⎩q 2 ⎭ ⎩ f1 − f 2 ⎭ ⎩ F2 ⎭
⎣ ⎦ ⎩ 2⎭ ⎣

For harmonic excitation { f } = { f̂ } cos ω t , the steady-state response is


{ q } = { q̂ } cos ω t . The amplitudes of modal coordinates are
F̂1 f̂1 + 2 f̂ 2 F̂2 f̂1 − f̂ 2
q̂1 = = , q̂2 = = .
3k − ω 6m2
(
3k 1 − ω 2
ω 12 ) 6k − ω 2 3m 6k 1 − ω( 2
ω 22 )
The amplitudes in the configuration space are given by

{ }
⎧1⎫
b 1 2 c f̂ ⎨ ⎬ b 1 − 1 c f̂ { } ⎧⎨−11⎫⎬
{ x̂ } = ⎩2⎭+ ⎩ ⎭.
(
3k 1 − ω 2 ω 12 6k 1 − ω 2 ) ( ω 2
2 )
When f 2 = 0 and f1 = f , the vector of amplitudes can be written

⎧ f̂ ⎫ ⎧1⎫ ⎧ f̂ ⎫ ⎧ 1 ⎫
b1 2 c ⎨ ⎬ ⎨ ⎬ b1 − 1c ⎨ ⎬ ⎨ ⎬
{ x̂ } = ⎩ 0 ⎭ ⎩2 ⎭ + ⎩ 0 ⎭ ⎩− 1⎭
(
3k 1 − ω 2 ω 12 )
6k 1 − ω 2 ω 22 ( )
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 117

and the receptances are


x̂ 1 1 1 x̂2 2 −1
α11 = = + , α 21 = = + ,
f̂ ⎛ ω ⎞ 2 ⎛ ω ⎞ 2
f̂ ⎛ ω ⎞ 2 ⎛ ω2 ⎞
3k ⎜1 − 2 ⎟ 6 k ⎜1 − 2 ⎟ 3k ⎜1 − 2 ⎟ 6 k ⎜1 − 2 ⎟
⎜ ω1 ⎟ ⎜ ω2 ⎟ ⎜ ω ⎟ ⎜ ω ⎟
⎝ ⎠ ⎝ ⎠ ⎝ 1 ⎠ ⎝ 2 ⎠

or

x̂ 1 1 1 x̂2 1 1
= + , = − .
f̂ (
6m ω − ω2
1
2
) (
3m ω − ω
2
2
2
) f̂ (
3m ω 1 − ω
2 2
)
3m ω2 − ω
2 2
( )
The frequency response curves are given in Fig. 4.5.

a b
Fig. 4.5

1 k k
Resonances occur at ω 1 = and ω 2 = 2 , when the
2 m m
excitation frequency equals any of the natural frequencies of the system. When
k
ω = ωa = , the first mass stands still in space, x̂ 1 = 0 , while the second mass
m
moves, x̂ 2 ≠ 0 , condition defined as an antiresonance. The antiresonance
frequency is equal to the natural frequency of the subsystem consisting of the
spring k and the mass m. This subsystem is called a dynamic vibration absorber.
The energy introduced per cycle in the system by the impressed force goes into this
part of the vibrating system, keeping the mass 2m still in space, condition
desirable in many practical applications.
118 MECHANICAL VIBRATIONS

4.1.5.2 Solution by Spectral Analysis

Substituting (4.34) and (4.35) in equation (4.31, a), we obtain

([ k ] − ω 2
[ m ] ){ x̂ } = { f̂ }, (4.43)
or

{ x̂ } = ( [ k ] − ω 2 [ m ] ) { f̂ }.
−1
(4.44)
Equation (4.43) represents a linear set of algebraic equations that can be
solved using Cramer’s rule. The inversion in equation (4.44) is never performed.

Example 4.3
Consider the system of Fig. 4.2 acted upon by a harmonic driving force
f1 = f̂1 cos ω t and obtain the steady-state response amplitudes by direct spectral
analysis.
Solution. The equations of motion (4.31) are
2 m &x&1 + 3 k x 1 − k x 2 = f̂ 1 cosω t ,
m &x& 2 − k x 1 + k x 2 = 0.

Substituting solution (4.35) into above equations, we obtain a set of two


algebraic equations

(3k − ω 2
)
2 m x̂ 1 − k x̂ 2 = f̂ 1 ,
( )
− k x̂ 1 + k − ω 2 m x̂ 2 = 0.

Using Cramer’s rule, the amplitudes x̂ 1 and x̂ 2 are

1 −k

x̂ 1 =
0 k − ω 2m

=
( k − ω m) f̂2
,
3k − ω2 2 m −k ( 3 k − ω 2 m)( k − ω m)− k
2 2 2

−k k − ω 2m

3k −ω2 2m 1

−k 0 k f̂
x̂ 2 = = ,
2
3k − ω 2 m −k (3 k − ω 2
)( 2
2m k −ω m − k ) 2

−k k − ω 2m
or, in a form close to that obtained by modal analysis
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 119

⎛ k 2⎞ k
⎜ − ω ⎟ f̂ f̂
x̂ 1 = ⎝ m ⎠ , x̂ 2 = m .
⎛ k 2 ⎞ ⎛ 2k 2⎞ ⎛ k 2 ⎞ ⎛ 2k 2⎞
2 m ⎜⎜ − ω ⎟⎟ ⎜ −ω ⎟ 2 m ⎜⎜ − ω ⎟⎟ ⎜ −ω ⎟
⎝ 2m ⎠⎝ m ⎠ ⎝ 2m ⎠⎝ m ⎠
The denominator can be recognized as the characteristic determinant,
which makes the amplitudes to grow indefinitely when the forcing frequency
equals either of the natural frequencies. The system has two resonances.

4.2 Torsional Systems


In the following, two-degree-of-freedom disc-shaft systems are considered,
in which the rigid discs have angular vibrations with respect to the shaft axis.

4.2.1 Equations of Motion

Consider the system of Fig. 4.6, a which consists of two rigid discs of
polar mass moment of inertia J 1 and J 2 , kg m 2 , attached to massless shafts of
torsional stiffness K1 , K 2 and K 3 , N m rad .

Fig. 4.6

The instantaneous angular position of discs with respect to the equilibrium


position is denoted by θ 1 and θ 2 . Using the free-body diagrams of Fig. 4.6, b and
120 MECHANICAL VIBRATIONS

d’Alembert’s principle (dynamic equilibrium of impressed and inertia torques), the


equations of motion can be written
( )
J1 θ&&1 + K1 θ 1 + K 2 θ 1 − θ 2 = 0 ,
( )
J 2 θ&&2 + K 3 θ 2 − K 2 θ 1 − θ 2 = 0 ,
or
J1 θ&&1 + (K1 + K 2 ) θ 1 − K 2 θ 2 = 0,
(4.45)
J 2 θ&&2 − K 2 θ 1 + (K 3 + K 2 ) θ 2 = 0.

This is a set of coupled differential equations resembling equations (4.1).


There is a complete analogy between systems in translational and angular
vibration, the counterparts of springs, masses and forces being torsional springs,
discs possessing mass moments of inertia and torques. All results established in
Section 4.1 apply to torsional systems. In the following, only systems admitting
rigid-body motions are considered.

4.2.2 Two-Disc Free-Free Systems

The shafting of a motor-driven fan or pump may rotate in its bearings as a


rigid body. Many engineering systems can be modelled by a two-disc torsional
system not tied rigidly to the ground (Fig. 4.7). The two discs, modelling the rotors
of the driving and driven machines, are joined by a torsional spring representing
the two shafts and the coupling.
Let the discs have polar mass moments of inertia J 1 and J 2 , and the
massless shaft have a torsional stiffness K = G I p l .

The equations of motion


J 1 θ&&1 + K θ 1 − K θ 2 = 0,
(4.46)
J 2 θ&&2 − K θ 1 + K θ 2 = 0,
can be written in matrix form

⎡ J1 0 ⎤ ⎧⎪θ&&1 ⎫⎪ ⎡ K − K ⎤ ⎧θ 1 ⎫ ⎧0⎫
⎢0 ⎨ ⎬+ ⎨ ⎬ = ⎨ ⎬, (4.46, a)
⎣ J 2 ⎥⎦ ⎪⎩θ&&2 ⎪⎭ ⎢⎣− K K ⎥⎦ ⎩θ 2 ⎭ ⎩0⎭

and in compact form


[ J ]{θ&& }+ [ K ]{ θ } = { 0 } .
The stiffness matrix [ K ] is positive semidefinite. Because the system is
ungrounded, the stiffness matrix is singular. The system can rotate freely, having a
rigid-body motion in which the potential energy is zero.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 121

Apparently, this is a two-degree-of-freedom system. However, summing


up equations (4.46) we obtain
J1 θ&&1 + J 2 θ&&2 = 0 , (4.47)

so that the two coordinates θ 1 and θ 2 are not independent. Integrating, we may
obtain a constraint equation that can be used to eliminate one coordinate from the
problem formulation.

Fig. 4.7

Dividing the first equation (4.46) by J1 , the second equation by J 2 and


subtracting the results from each other we obtain
⎛K K ⎞
θ&&1 − θ&&2 + ⎜⎜ + ( )
⎟⎟ θ 1 − θ 2 = 0 . (4.48)
⎝ J1 J 2 ⎠
Denoting the twist angle θ 1 − θ 2 = θ , equation (4.48) becomes

J1 J 2 &&
θ + Kθ = 0 (4.48, a)
J1 + J 2
which is the equation of motion of a single-degree-of-freedom system.

4.2.2.1 Normal Modes

Assuming solutions of the form


θ 1 (t ) = a 1 cos (ω t − ϕ ) ,
(4.49)
θ 2 (t ) = a 2 cos (ω t − ϕ ) ,
122 MECHANICAL VIBRATIONS

we obtain the set of algebraic equations

(K − J ω ) a − K a = 0,
1
2
1 2
(4.50)
− K a + (K − J ω ) a = 0.
1 2
2
2

Dividing by K and denoting

ω 2 J1 K = α (4.51)
equations (4.50) become
(1 − α ) a1 − a2 = 0,
⎛ J ⎞ (4.52)
− a1 + ⎜⎜1 − 2 α ⎟⎟ a2 = 0.
⎝ J1 ⎠

The simultaneous homogeneous equations (4.52) admit non-trivial


solutions if the determinant of the coefficients a 1 and a 2 is zero

1−α −1
J =0 (4.53)
−1 1− 2 α
J1

J2 2 ⎛ J2 ⎞
or α − ⎜⎜1 + ⎟⎟ α = 0 . (4.53, a)
J1 ⎝ J1 ⎠
The solutions are
α1 = 0 , α 2 = 1 + J1 J 2 . (4.54)

The first natural frequency is given by ω 21 = 0 and the second by

⎛1 1 ⎞
ω 22 = ⎜⎜ + ⎟⎟ K . (4.55)
⎝ J1 J 2 ⎠

The root ω 21 = 0 indicates that rigid body displacement is possible. This


may be due to a static angular displacement or to a uniform rotating speed. It is not
a genuine vibration. The solution of equations (4.46) is of the form
θ 1 (t ) = C1 + C2 t + C3 sin ω 2 t + C4 cos ω 2 t ,
( )
θ 2 (t ) = C1 + C2 t + 1 − J1ω 22 K (C3 sin ω 2 t + C4 cos ω 2 t )

where C1 ,..,C4 are integration constants.


The mode shapes are determined by the ratio μ = a2 a1 = 1 − α .
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 123

For the first mode


μ 1 = (a2 a1 ) 1 = 1 − α1 = 1 , (4.56, a)

both discs have the same angular displacement defining the rigid-body rotation in
which the shaft is not twisted.
For the second mode
μ 2 = (a2 a1 ) 2 = 1 − α 2 = − J1 J 2 , (4.56, b)

the discs vibrate in opposition. The shaft has a nodal point which is closer to the
larger disc.
The mode shapes are graphically presented in Fig. 4.8.

Fig. 4.8

4.2.2.2 Response to Harmonic Excitation

Consider the system of Fig. 4.6 where on the second disc acts a harmonic
torque M (t ) = M 0 cos ω t (not shown). The equations of motion are

J1 θ&&1 + K θ 1 − K θ 2 = 0 ,
(4.57)
J 2 θ&&2 − K θ 1 + K θ 2 = M 0 cos ω t .

The steady-state vibration of this system is of the form


θ 1 (t ) = Θ1 cos ω t , θ 2 (t ) = Θ 2 cos ω t . (4.58)
124 MECHANICAL VIBRATIONS

The resulting algebraic equations are

(K − J ω ) Θ − K Θ = 0,
1
2
1 2
(4.59)
− K Θ + (K − J ω ) Θ = M
1 2
2
2 0.

a b
Fig. 4.9

Dividing by K and denoting ω 2 J1 K = α we obtain

(1 − α ) Θ 1 −Θ 2 = 0,
⎛ J ⎞ M (4.60)
− Θ 1 + ⎜⎜1 − 2 α ⎟⎟ Θ 2 = 0 .
⎝ J 1 ⎠ K

Solving for Θ 1 and Θ 2 using Cramer’s rule

0 −1
M0
J
1 1− 2 α K
J1 1 M0
Θ1 = = , (4.61)
1−α −1 J2 ⎛ J1 + J 2 ⎞ K
⎜ α− ⎟ α
J
−1 1− 2 α J1 ⎜⎝ J 2 ⎟⎠
J1

1−α 0 M0
−1 1 K 1−α M0
Θ2 = = . (4.62)
1−α −1 J2 ⎛ J1 + J 2 ⎞ K
⎜ α− ⎟ α
J
−1 1− 2 α J1 ⎜⎝ J 2 ⎟⎠
J1
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 125

The amplitudes Θ 1 and Θ 2 are graphically presented in Fig. 4.9 as a


function of the disturbing frequency. Both angular amplitudes become infinite
when the denominator is zero, i.e. when the frequency of the driving torque equals
one of the natural frequencies. There is a sort of “resonance at zero frequency”
corresponding to the rigid-body mode and a genuine resonance at
⎛ 1 1 ⎞
ω2 = ⎜⎜ + ⎟⎟ K .
⎝ J1 J 2 ⎠
The amplitude of the second disc is zero when α = 1 . This antiresonance
occurs when ω = K J1 , the natural frequency of the subsystem consisting of the
shaft and the first disc, which acts as a dynamic absorber and keeps the second disc
still in space.

4.2.2.3 Dynamic Stresses

The amplitude of the angle of twist is


1 M0
ΔΘ = Θ 2 − Θ 1 = . (4.63)
J 2 ⎛ J1 + J 2 ⎞ K
⎜ − α ⎟⎟
J1 ⎜⎝ J 2 ⎠
Denoting the torque in the shaft by M t = M t 0 cos ω t , its amplitude is

M0
M t 0 = K ΔΘ = . (4.64)
J 2 ⎛ J1 + J 2 ⎞
⎜ − α ⎟⎟
J1 ⎜⎝ J 2 ⎠
The dynamic shear stresses due to the torsion are τ = τ 0 cos ω t , and the
amplitude is
τ 0 = M t 0 Wp (4.65)

where W p is the polar modulus of the cross section.

If the shaft transmits a power N at constant angular speed ω N , then the


“static” shear stress is
τ N = M t st W p = N (ω N W p ) . (4.66)

A fatigue calculation can be carried out based on values of τ 0 and τ N ,


considering a time history as in Fig. 4.10.
126 MECHANICAL VIBRATIONS

Fig. 4.10

Example 4.4
The torsional system of Fig. 4.6 is acted upon by a harmonic torque of
amplitude M 0 = 10 4 N m and frequency ω = 314 rad sec (not shown). Both discs
have the polar mass moment of inertia J = 57 kg m 2 . The shaft has a length
l = 0.4 m , diameter d = 0.14 m and shear modulus G = 81 GPa . Determine the
amplitude of the dynamic shear stresses in the shaft.

Solution. The shaft cross-section has I p = π d 4 32 = 0.377 ⋅ 108 mm 4 and


W p = π d 3 16 = 0.538 ⋅ 106 mm3 . The torsional stiffness is K = G I p l = 7.63 ⋅ 109
N mm rad . The ratio α = J ω 2 K = 0.735 . The twist angle is
ΔΘ = M 0 K (2 − α ) = 0.001036 rad . The amplitude of the dynamic torque is
M t 0 = M 0 (2 − α ) = 7910 N m . The amplitude of dynamic shear stresses is
τ 0 = M t 0 W p = 14.7 N mm 2

4.2.3 Geared Systems

Consider the geared torsional system of Fig. 4.11, a with a gear of pitch
radius r1 on shaft 1 and a gear of pitch radius r 2 on shaft 2. Assume the gears are
rigid, of negligible inertia and their teeth remain in contact. The gear ratio is
r1 n2 θ2
i= =− =− , (4.67)
r2 n1 θ1

where n 1 and n 2 are the rotational speeds of the two shafts, θ 1 and θ 2 are the
corresponding angular displacements.
The geared system is conveniently reduced to an equivalent non-geared
system (Fig. 4.11, b) in which the gears are omitted. In the reduction process, the
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 127

stiffness of the equivalent shaft is determined from the condition of equal potential
energies
(K θ ) 2
actual (
= Kθ2 ) eq ,
wherefrom
2 2
⎛θ ⎞ ⎛ ⎞
K eq = ⎜ actual ⎟ K actual = ⎜ nactual ⎟ K actual = i 2 K actual . (4.68)
⎜ θ eq ⎟ ⎜ neq ⎟
⎝ ⎠ ⎝ ⎠

Fig. 4.11

The polar mass moment of inertia of the equivalent disc is determined from
the condition of equal kinetic energies

( J θ& )2
actual (
= J θ& 2 ) eq ,
wherefrom
2 2
⎛ θ& ⎞ ⎛ ⎞
J eq = ⎜ actual ⎟ J actual = ⎜ nactual ⎟ J actual = i 2 J actual . (4.69)
⎜ θ&eq ⎟ ⎜ neq ⎟
⎝ ⎠ ⎝ ⎠
Choosing shaft 1 as reference, the equivalent parameters of shaft 2 are
(Fig. 4.11, b)
K 2 eq = i 2 K 2 , J 2 eq = i 2 J 2 . (4.70)

The following rule applies for the equivalent systems when gears have
negligible inertia : Remove all gears and multiply all stiffnesses and all inertias by
i 2 , where -i is the speed ratio of the geared shaft to the reference shaft.
128 MECHANICAL VIBRATIONS

After determining the mode shapes of the equivalent system, the mode
shapes and the torques of the actual system are recovered from the compatibility
equations
θ actual θ eq = −i , M eq M actual = −i . (4.71)

4.2.4 Geared-Branched Systems

Consider the branched system with gears of negligible inertia and massless
shafts shown in Fig. 4.12, a. It can be converted to the model with an one-to-one
gear shown in Fig. 4.12, b, by multiplying all the stiffnesses and inertias of the
branch 2-3 by the square of the speed ratio i. Note that torques and angular
displacements in the reduced branch are different from the actual values, according
to equations (4.71).

Fig. 4.12

The equations of motion can be written using the finite element approach.
A uniform shaft is considered a two-noded finite element of torsional stiffness K.
The points of attachment of the shaft to other parts of the vibrating system are
called nodes (not to be confused with the stationary points of mode shapes) and are
denoted 1 and 2 in Fig. 4.13.
The torques M 1 and M 2 may be related to the rotation angles θ 1 and θ 2
using the equilibrium and the torque/rotation equations
M1 = − M 2 = K θ 1 when θ 2 = 0,
(4.72)
M 1 = − M 2 = − K θ 2 when θ 1 = 0.

Equations (4.72) may be written in matrix form as


4. TWO-DEGREE-OF-FREEDOM SYSTEMS 129

⎧ M1 ⎫ ⎡ K − K⎤ ⎧ θ1 ⎫
⎨ ⎬=⎢ ⎨ ⎬ (4.73)
⎩ M 2 ⎭ ⎣− K K ⎥⎦ ⎩ θ 2 ⎭

or in shorthand form { M }= [ k ] { θ }, where [ k ] is referred to as the element


e e

stiffness matrix.

Fig. 4.13

Using equation (4.73), the torque-rotation equation for each shaft in Fig.
4.12, b can be written

⎧ M 1 ⎫ ⎡ K1 − K1 ⎤ ⎧ θ 1 ⎫ ⎧ M 2 ⎫ ⎡ K 2 − K2 ⎤ ⎧ θ 2 ⎫
⎨ ⎬=⎢ ⎨ ⎬, ⎨ ⎬= ⎨ ⎬,
⎩ M 3 ⎭ ⎣− K1 K1 ⎥⎦ ⎩ θ 3 ⎭ ⎩ M 3 ⎭ ⎢⎣− K 2 K 2 ⎥⎦ ⎩ θ 3 ⎭

⎧ M 3 ⎫ ⎡ K3 − K3 ⎤ ⎧ θ 3 ⎫
⎨ ⎬=⎢ ⎨ ⎬.
⎩ M 4 ⎭ ⎣− K 3 K 3 ⎥⎦ ⎩ θ 4 ⎭

Each of these equations can be expanded as follows

⎧ M1 ⎫ ⎡ K1 0 − K1 0⎤ ⎧ θ 1 ⎫ ⎧ M1 ⎫ ⎡0 0 0 0⎤ ⎧ θ 1 ⎫
⎪M
⎪ 2
⎪ ⎢ 0
⎪ ⎢ 0 0 0⎥⎥ ⎪⎪ θ 2 ⎪⎪ ⎪M
⎪ 2
⎪ ⎢0 K
⎪ ⎢ 2 − K2 0⎥⎥ ⎪⎪ θ 2 ⎪⎪
⎨ ⎬= ⎨ ⎬, ⎨ ⎬= ⎨ ⎬,
⎪ M3 ⎪ ⎢− K1 0 K1 0⎥ ⎪ θ 3 ⎪ ⎪ M3 ⎢
⎪ 0 − K2 K2 0⎥ ⎪ θ 3 ⎪
⎪⎭ ⎢⎣ 0 ⎥ ⎪⎭ ⎢⎣0 ⎥
⎪⎩ M 4 0 0 0⎦ ⎪⎩ θ 4 ⎪⎭ ⎪⎩ M 4 0 0 0⎦ ⎪⎩ θ 4 ⎪⎭

⎧ M1 ⎫ ⎡0 0 0 0 ⎤ ⎧θ1 ⎫
⎪M ⎪ ⎢0 0 0 0 ⎥⎥ ⎪⎪ θ 2 ⎪⎪
⎪ 2 ⎪ ⎢
⎨ ⎬=⎢ ⎨ ⎬.
⎪ M3 ⎪ 0 0 K3 − K3 ⎥ ⎪ θ3 ⎪
⎪⎩ M 4 ⎪⎭ ⎢⎣0 0 − K3

K 3 ⎦ ⎪⎩ θ 4 ⎪⎭

The torques in the overall system are obtained by adding all the torques at
each node. This can be obtained by adding the expanded stiffness matrices to give
130 MECHANICAL VIBRATIONS

⎧ M1 ⎫ ⎡ K1 0 − K1 0 ⎤ ⎧θ1 ⎫
⎪M ⎪ ⎢ 0 K2 − K2 0 ⎥⎥ ⎪⎪ θ 2 ⎪⎪
⎪ 2 ⎪ ⎢
⎨ ⎬=⎢ ⎨ ⎬. (4.74)
⎪ M3 ⎪ − K1 − K 2 K1 + K 2 + K 3 − K3 ⎥ ⎪ θ3 ⎪
⎪⎩ M 4 ⎪⎭ ⎢⎣ 0 0 − K3

K 3 ⎦ ⎪⎩ θ 4 ⎪⎭

Using the boundary condition θ 4 = 0 , the reduced system stiffness matrix


is obtained, so the equations of motion can be written

⎡ J1 0 0⎤ ⎧ θ&&1 ⎫ ⎡ K1 0 − K1 ⎤ ⎧ θ1 ⎫ ⎧0⎫
⎢0 ⎪ ⎪ ⎢ ⎥ ⎪θ ⎪ ⎪ ⎪
⎢ J2 0⎥⎥ ⎨ θ&&2 ⎬+⎢ 0 K2 − K2 ⎥⎨ 2 ⎬ = ⎨ 0 ⎬. (4.75)
⎢⎣ 0 0 0⎥⎦ ⎪⎩ θ&&3 ⎪ ⎢− K − K K1 + K 2 + K 3 ⎥⎦ ⎪⎩ θ 3 ⎪ ⎪0⎪
⎭ ⎣ 1 2 ⎭ ⎩ ⎭

Eliminating the coordinate θ 3 using the third equation


− K1 θ 1 − K 2 θ 2 + (K1 + K 2 + K 3 )θ 3 = 0

we obtain the two equations of motion


K1 (K 2 + K 3 ) K1 K 2
J1θ&&1 + θ1 − θ 2 = 0,
K1 + K 2 + K 3 K1 + K 2 + K 3
(4.76)
K1 K 2 K (K + K 3 )
J 2 θ&&2 − θ1 + 2 1 θ 2 = 0.
K1 + K 2 + K 3 K1 + K 2 + K 3
After solving the eigenvalue problem, in order to draw the actual mode
shapes, the angular amplitudes Θ 2 determined for the equivalent system must be
transformed back to actual values using equation (4.71). A more straightforward
approach is presented in Section 5.1.3.

4.3 Flexural Systems


Herein, massless beams and planar frames are considered, with attached
rigid masses. For such systems it is easier to calculate flexibilities instead of
stiffnesses, the former being measurable quantities.

4.3.1 Flexibility Coefficients

Let an elastic beam be subjected to forces f1 and f 2 at points 1 and 2


(Fig. 4.14, a). The deflections in the directions of f1 and f 2 are y 1 and y2 ,
respectively.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 131

Consider the unloaded beam (Fig. 4.14, b) and apply a single unit force at
point 1 in the direction of f1 . Let the deflections at 1 and 2, in the directions y 1
and y2 be δ 11 and δ 21 . Similarly, let a unit force be applied at point 2 in the
direction of f 2 and denote the deflections δ 12 and δ 22 (Fig. 4.14, c).

The relationship between the forces f1 and f 2 and the total deflections y 1
and y2 can be expressed by the equations
y 1 = δ 11 f1 + δ 12 f 2 ,
(4.77)
y2 = δ 21 f1 + δ 22 f 2 .

The coefficients δ ij are called flexibility (influence) coefficients.

By definition, δ ij is the deflection of coordinate i due to unit load applied


to coordinate j.

Fig. 4.14

In matrix notation, equations (4.77) become

⎧ y 1 ⎫ ⎡δ 11 δ 12 ⎤ ⎧ f1 ⎫
⎨ ⎬ = ⎢δ ⎥⎨ ⎬ (4.77, a)
⎩ y2 ⎭ ⎣ 21 δ 22 ⎦ ⎩ f 2 ⎭
or
{ y } = [ δ ]{ f }, (4.78)
where [ δ ] is known as the flexibility matrix.
The flexibility matrix is symmetrical [ δ ] T = [ δ ] , according to Maxwell’s
reciprocal theorem: The deflection at one point in a structure due to a unit load
applied at another point equals the deflection at the second point when a unit load
is applied at the first (deflections are measured in the same direction as the load).
132 MECHANICAL VIBRATIONS

Because in an equation of the type { f } = [ k ] { y } the matrix [ k ] is


known as the stiffness matrix, it comes out that [ k ] = [ δ ] −1 or alternatively

[ δ ] = [ k ] −1 . (4.79)

When rigid-body motions are possible, [ δ ] = [ k ] −1 does not exist. There


are no “ground” reactions to counterbalance the unit forces that must be applied to
the structure to determine [δ ] and equilibrium is not possible.

4.3.2 Equations of Motion

The massless beam of Fig. 4.15, a has constant bending rigidity EI and
carries masses m 1 and m 2 at points 1 and 2. If only the two lateral displacements
of the two masses are of interest, the motion is completely defined by deflections
y 1 and y2 , hence the system has two degrees of freedom.

Fig. 4.15

In free vibrations (Fig. 4.15, b), applying d’Alembert’s principle, the only
external forces acting on masses are the inertia forces
f1 = −m 1 &y&1 , f 2 = − m 2 &y& 2 . (4.80)

Substituting these forces in equations (4.77) yields the differential


equations of motion
δ 11 m 1 &y&1 + δ 12 m 2 &y&2 + y 1 = 0 ,
(4.81)
δ 21 m 1 &y&1 + δ 22 m 2 &y&2 + y2 = 0.

In matrix form
[ b ] { &y& } + { y } = { 0 }. (4.82)
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 133

The matrix [ b ] can be written

b11 b12 ⎤ ⎡δ 11 m1 δ 12 m2 ⎤
[ b ] = ⎡⎢ ⎥=⎢ ⎥ = [δ ] [ m ], (4.83)
⎣b21 b22 ⎦ ⎣δ 21 m1 δ 22 m2 ⎦
where [ m ] is the diagonal mass matrix.
In equations (4.81) the coupling is due to the off-diagonal elements of the
flexibility matrix.

4.3.3 Normal Modes

Assuming solutions of the form


{ y } = { a } cos (ω t − ϕ ) , (4.84)
equation (4.81) becomes

− ω 2 [ b ]{ a } + { a } = { 0 }
or
1
[ b ]{a } = {a }. (4.85)
ω2
This is the standard eigenvalue problem, in which λ = 1 ω 2 are the
eigenvalues and { a } are the eigenvectors. The eigenvalues are the inverses of the
natural frequencies squared, and the eigenvectors are the modal vectors that define
the shape of the natural modes of vibration.
Equations (4.85) may be written under the form

(b 2
11ω − 1 ) a + b ω a = 0,
1 12
2
2
(4.85, a)
b21ω 2 a1 + ( b ω − 1) a = 0.
22
2
2

The condition to have non-trivial solutions gives the frequency equation

(b11ω
2
)( )
− 1 b22 ω 2 − 1 − b12 b21ω 4 = 0
or
( b11 b22 − b12 b21 )ω 4 − (b11 + b22 )ω 2 + 1 = 0 . (4.86)

Equation (4.86) has two real positive roots ω 12 and ω22 , the natural
frequencies squared.
The mode shapes are defined by the ratio μ = a2 a1 , so that
134 MECHANICAL VIBRATIONS

⎛ a2 ⎞ 1 − b11ω 12 ⎛ a2 ⎞ 2
μ1 = ⎜ ⎟ = , μ2 =⎜ ⎟ = 1 − b11ω2 . (4.87)
⎜ a1 ⎟ b ω 2 ⎜ a1 ⎟ b12 ω22
⎝ ⎠1 12 1 ⎝ ⎠2

Example 4.5
Calculate the natural modes of vibration for the beam of Fig. 4.15, a where
l1 = l 2 = l 2 , m 1 = m2 = m and E I = const .

Solution. The flexibility coefficients are

δ11 = l3 24 EI , δ12 = δ 21 = 5l 3 48EI , δ 22 = l 3 3EI .


The flexibility matrix is
l3 ⎡2 5 ⎤
[δ ] = ⎢5 16⎥ .
48 EI ⎣ ⎦
Equation (4.85) can be written
⎡2 5 ⎤ ⎧ a1 ⎫ ⎧ a1 ⎫
⎢5 16⎥ ⎨a ⎬ = λ ⎨ ⎬
⎣ ⎦ ⎩ 2⎭ ⎩a 2 ⎭
where
λ = 48 E I ml 3ω 2 .

Fig. 4.16
The frequency equation is
2−λ 5
=0 or λ2 − 18 λ + 7 = 0
5 16 − λ
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 135

with solutions λ1 = 17.602 and λ2 = 0.3976 .


The natural frequencies are

ω 1 = 1.6503 E I ml 3 , ω 2 = 10.986 E I ml3 .

The mode shapes are defined by the ratio μ = a2 a1 = λ − 2 5

⎛ a2 ⎞ λ 1 − 2 ⎛ a2 ⎞ λ −2
μ1 = ⎜ ⎟ = = 3.12 , μ2 =⎜ ⎟ = 2 = − 0.32 .
⎜ a1 ⎟ 5 ⎜ a1 ⎟ 5
⎝ ⎠1 ⎝ ⎠2

They are graphically presented in Fig. 4.16.

4.3.4 Free Vibrations

Let denote by

[ u ] = [{ u }1 { u } 2 ] = ⎡⎢
u11 u12 ⎤
⎥ (4.88)
⎣u 21 u22 ⎦
the matrix of normalized mode shapes (modal matrix).
The actual deflections, denoted in the following by { x } instead of { y },
may be expressed in terms of the modal coordinates as in (4.23)
2
{ x } = [ u ]{ q } = {u }1 q 1 + {u } 2 q 2 = ∑ Ci cos (ω i t − ϕi ) {u }i . (4.89)
i =1

The constants Ci and ϕi may be evaluated from the initial conditions of


the motion
2
{ x (0) } = ∑ Ci cos ϕi {u }i , (4.90)
i =1

2
{ x& (0) } = ∑ ω i Ci sin ϕi { u } i . (4.91)
i =1

Premultiplying equations (4.90) and (4.91) by { u }Tj [ m ] , using the


orthogonality relationships (4.18) - (4.19) and the definitions (4.25) and (4.28) of
the modal masses

{ u }Tj [ m ] { u }i = 0 , i≠ j M i = { u }Ti [ m ] { u } i , i = 1, 2 (4.92)


136 MECHANICAL VIBRATIONS

we obtain

{ u }Tj [ m ] { x (0) } = C j M j cos ϕ j , j = 1, 2 (4.93)

{ u }Tj [ m ]{ x& (0) } = ω j C j M j sin ϕ j . j = 1, 2 (4.94)

Combining equations (4.93) and (4.94) and renaming the index yields

{ u }Ti [ m ] { x& (0) }


tan ϕi = , i = 1, 2 (4.95)
ωi { u }Ti [ m ] { x (0) }

and
{ u }Ti [ m ] { x(0) } { u }Ti [ m ] { x& (0) }
Ci = = . i = 1, 2 (4.96)
M i cos ϕi ωi M i sin ϕi

Example 4.6
For the system of Fig. 4.17: a) determine the natural modes of vibration;
b) derive the equations of the free vibrations when the mass has an initial vertical
velocity v and calculate the trajectory of the mass.

Fig. 4.17

Solution. a) Let y and z be the vertical and horizontal components of the


instantaneous displacement of mass m.
The flexibility coefficients are

δ yy = 4l 3 3EI , δ yz = δ zy = l 3 2 EI , δ zz = l 3 3EI .
The equations of motion can be written
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 137

4l 3 l3
m &y& + m &z& + y = 0 ,
3E I 2E I
l3 l3
m &y& + m &z& + z = 0.
2E I 3E I
Looking for solutions of the form
y ( t ) = u 1 cos (ω t − ϕ ), z ( t ) = u2 cos (ω t − ϕ )

we obtain
( 8 − β ) u 1 + 3 u 2 = 0,
3 u 1 + ( 2 − β ) u 2 = 0,
where
β = 6 E I m l 3ω 2 .
The frequency equation is
8−β 3
=0, β 2 − 10 β + 7 = 0 ,
3 2−β
with solutions
β 1 = 9.2426 , β 2 = 0.7574 .
The natural frequencies are

ω 1 = 0.8057 E I ml3 , ω 2 = 2.8146 E I ml3 .


The mode shapes are given by
⎛ u2 ⎞ u21 β1 − 8 ⎛u ⎞ u β 2 −8
μ 1 = ⎜⎜ ⎟⎟ = = = 0.4142 , μ 2 = ⎜⎜ 2 ⎟⎟ = 22 = = −2.4142 .
⎝ u1 ⎠1 u11 3 ⎝ u1 ⎠ 2 u12 3

The modal vectors, normalized with unit first element, are

{ u }1 = ⎧⎨ ⎫
{ u }2 = ⎧⎨ ⎫
1 1
⎬, ⎬.
⎩0.4142⎭ ⎩− 2.4142⎭
It can be seen that
u21
= tan γ 1 = 0.4142 , γ 1 = 22.50 ,
u11

u22
= tan γ 2 = −2.4142 , γ 2 = 112.50 = γ 1 + 900 .
u12
138 MECHANICAL VIBRATIONS

The mass m has unidirectional motion in the natural modes of vibration.


The modal vectors are orthogonal { u }T1 { u }2 = 0 . The motion in the first mode is
along direction 1 at 22.50 measured from the vertical direction. The motion in the
second mode is along direction 2 at 122.50 , hence normal to direction 1. This
happens because the motion in modal coordinates is along the directions of
principal flexibilities.
The flexibility matrix
l 3 ⎡8 3⎤
[δ ] = .
6 E I ⎢⎣3 2⎥⎦
represents a deflection-force relation between the components y and z of the
deflection and the components f y and f z of a force applied to the point where the
mass is located
⎧ y ⎫ ⎡δ yy δ yz ⎤ ⎧ f y ⎫
⎨ ⎬=⎢ .
⎩ z ⎭ ⎣δ zy δ zz ⎥⎦ ⎨⎩ f z ⎬⎭

Consider a reference frame y ∗Oz ∗ rotated through an angle γ with respect


to the frame yOz . The transformation of displacements can be written

⎧⎪ y ∗ ⎫⎪ ⎡ cosγ sinγ ⎤ ⎧ y ⎫
⎨ ∗ ⎬=⎢ ⎨ ⎬
⎪⎩ z ⎪⎭ ⎣− sinγ cosγ ⎥⎦ ⎩ z ⎭

and the transformation of forces is defined by


⎧⎪ f y∗ ⎫⎪ ⎡ cosγ sinγ ⎤ ⎧ f y ⎫
⎨ ∗ ⎬=⎢ ⎨ ⎬.
⎪⎩ f z ⎪⎭ ⎣− sinγ cosγ ⎥⎦ ⎩ f z ⎭

The new deflection-force relation is

[ ] ⎧⎪⎨⎪ ff
⎧⎪ y ∗ ⎫⎪ ∗ ⎫⎪

⎨ ∗ ⎬= δ
y
∗ ⎬
⎪⎩ z ⎪⎭ ⎩ z ⎪⎭
where

[δ ] = ⎡⎢−cs
∗ s⎤

c⎦
[ δ ] ⎡⎢
c − s⎤

⎣ ⎣s c ⎦
with c = cosγ and s = sinγ .
The flexibility matrix in the rotated reference frame is
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 139

[δ ] (

⎡ δ yy c 2 + δ zz s 2 + 2δ yz cs (δ zz − δ yy )sc + δ yz ( c 2 − s 2 )⎤⎥ .
=⎢
) 2
⎢⎣ δ zz − δ yy sc + δ yz c − s
2
( ) δ yy s 2 + δ zz c 2 − 2δ yz cs ⎥⎦

It can be seen that there are two angles γ ∗ for which the off-diagonal
elements vanish, given by
2 δ yz 2⋅3
tan 2 γ ∗ = = = 1 , γ 1∗ = 22.50 , γ 2∗ = 112.50 .
δ yy − δ zz 8−2
The two solutions γ 1∗ and γ 2∗ define the principal directions of flexibility.
Substituting these angles into the expression of diagonal elements we obtain the
principal flexibilities
2
δ yy + δ zz ⎛ δ yy − δ zz ⎞ 2 5 ± 3 2 l3
δ1, 2 = ± ⎜⎜ ⎟⎟ + δ yz = ,
2 ⎝ 2 ⎠ 6 EI
l3 l3
δ1 = 1.5404 , δ 2 = 0.1262 .
EI EI
Their meaning is straightforward. A force applied along 1 (or 2) produces a
deflection only along 1 (or 2). The principal directions of flexibility coincide with
the directions of vibration in the natural (principal) modes of vibration.
The natural frequencies are given by
1 1 EI EI
ω1 = = = 0.805 ,
m δ1 1.5404 ml 3
m l3

1 1 EI EI
ω2 = = = 2.815 .
mδ2 0.1262 ml 3
m l3

b) To determine the free response to a vertical impulse of velocity v , let


calculate first the modal masses
2
(
M 1 = m u11 2
+ u21 ) (
= m 1 + 0.4142 2 = 1.1716 m , )
M2 = m (u 2
12
2
+ u22) = m (1 + 2.4142 ) = 6.8284 m .
2

The initial conditions are

{ x (0) } = ⎧⎨
0⎫
{ x& (0) } = ⎧⎨
v⎫
⎬, ⎬.
⎩0⎭ ⎩0⎭
From (4.93) and(4.96) we obtain
140 MECHANICAL VIBRATIONS

cos ϕ 1 = 0 , cos ϕ 2 = 0 ,
mv mv
C1 sin ϕ 1 = , C2 sin ϕ 2 = .
ω 1 M1 ω2 M 2

a b
Fig. 4.18

Fig. 4.19

The vertical component of the instantaneous displacement is

( ) (
y (t ) = C1 cos ω 1 t − ϕ1 u11 + C2 cos ω 2 t − ϕ 2 u12 , )
mv mv
y (t ) = sin ω 1t + sin ω 2t ,
ω 1M 1 ω 2M 2

(
y (t ) = C 1.0593 sin ω 1 t + 0.0520 sin ω 2 t , )
where
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 141

ml 3
C=v , ω 1 = 0.805 v C , ω 2 = 2.815 v C .
EI

The horizontal component of the instantaneous displacement is

( ) ( )
z (t ) = C1 cos ω 1 t − ϕ 1 u 21 + C2 cos ω 2 t − ϕ 2 u22 ,

mv mv
z (t ) = u 21 sin ω 1 t + u 22 sin ω 2 t ,
ω1 M1 ω2 M 2

(
z (t ) = C 0.4387 sin ω 1 t − 0.1256 sin ω 2 t . )
The trajectory of the mass m is plotted in Fig. 4.18, a for a time duration
equal to 2π ω 1 and in Fig. 4.18, b for 4π ω 1 . The two components y and z are
plotted in Fig. 4.19 as a function of time. It is seen that the horizontal component is
almost harmonic, the second component having a relatively small amplitude.

4.3.5 Response to Harmonic Excitation

Consider the steady-state vibrations of the beam of Fig. 4.20, a under the
action of the force f (t ) = F0 cos ω t acting on mass m2 .

Fig. 4.20

The equations of motion are


142 MECHANICAL VIBRATIONS

( )
y 1 = −δ 11 m 1 &y&1 + f − m 2 &y&2 δ 12 ,
y2 = −δ 21 m 1 &y&1 + ( f − m 2 &y&2 ) δ 22
or
δ 11 m 1 &y&1 + δ 12 m 2 &y&2 + y 1 = δ 12 f (t ) ,
(4.97)
δ 21 m 1 &y&1 + δ 22 m 2 &y&2 + y2 = δ 22 f (t ) .

Substituting the steady-state solutions


y1 (t ) = Y1 cos ω t , y2 (t ) = Y 2 cos ω t ,

into equations (4.97) yields

(1 − ω δ 2
)Y −ω δ
11 m 1 1
2
12 m 2 Y2 = δ 12 F0 ,
(4.98)
− ω 2δ 21 m Y + (1 − ω δ
1 1
2
)
22 m 2 Y2 = δ 22 F0 .

The solutions are


δ 12
Y1 = F0 ,
( )
1 − δ 11 m 1 + δ 22 m 2 ω 2 + m 1 m 2 δ 11δ 22 − δ12
2
(
ω4 )
Y2 =
(
δ 22 − m 1ω 2 δ 11δ 22 − δ12
2
) F0 .
(4.99)

( )
1 − δ 11 m 1 + δ 22 m 2 ω 2 + m 1 m 2 δ 11δ 22 − δ12
2
(
ω4 )
The denominator can be recognized as the characteristic polynomial (4.86).

a b
Fig. 4.21

The absolute values of the amplitudes Y1 and Y2 are graphically presented


in Fig. 4.21. When the forcing frequency equals either of the natural frequencies,
the amplitudes grow indefinitely. The system has two resonances marked by peaks
in the frequency response curves.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 143

Antiresonance occurs for Y2 = 0 when

δ 22
ω 2 = ωa2 = . (4.100)
( 2
m 1 δ 11δ 22 − δ 12 )
If the amplitudes of the forced response are known, the amplitudes of the
inertia forces can be calculated so that the amplitude of the dynamic forces acting
on the beam (Fig. 4.20, b) are

Φ 1 = m1ω 2 Y1 , Φ 2 = m2 ω 2 Y2 + F0 . (4.101)

The diagram of dynamic bending moments (Fig. 4.20, c) may be then


constructed, and the dynamic stresses produced by the harmonic force may be
calculated.

Example 4.7
The massless beam of Fig. 4.22, a has diameter d = 40 mm , l = 1 m ,
E = 210 GPa and m = 50 kg . a) Calculate the natural frequencies; b) Determine
the amplitudes of forced vibrations produced by a harmonic force of amplitude
F0 = 20 N and frequency 0.179 Hz ; c) Draw the diagram of static bending
moments and determine the maximum static stress ; d) Draw the diagram of the
dynamic bending moments and calculate the amplitude of the maximum dynamic
stress.
Solution. The flexibility coefficients are

δ 11 = l 3 6 E I , δ 12 = δ 21 = − l3 8 E I , δ 22 = 5 l 3 24 E I .
Denoting
24 E I
λ= ,
ω 2m l3
the frequency equation is
λ −8 3
=0, λ2 − 13λ + 22 = 0 ,
6 λ −5
with solutions
λ 1 = 11 , λ2 = 2.
The natural frequencies are

ω 1 = 1.477 E I ml 3 = 1.073 rad sec ,


144 MECHANICAL VIBRATIONS

ω 2 = 3.464 E I ml 3 = 2.516 rad sec .

Fig. 4.22

For the given numerical data, the forcing frequency corresponds to λ = 10 .


From (4.99) we obtain the vibration amplitudes
3λ F0
Y1 = − = 0.118 mm ,
(λ − 11) (λ − 2)
mω 2
22 − 5 λ F0
Y2 = − = −1.105 mm .
(λ − 11) (λ − 2) mω 2
For the static loading shown in Fig. 4.22, b, the diagram of static bending
moments is presented in Fig. 4.22, c. The maximum bending moment is 368 Nm
and the maximum static stress is σ st = 58.5 N mm 2 .
The amplitudes (4.101) of dynamic forces are
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 145


Φ 1 = 2m ω 2 Y1 = − F0 = 150 N ,
(λ − 11) (λ − 2)
⎡ 22 − 5 λ ⎤
Φ 2 = m ω 2 Y2 + F0 = ⎢1 − F0 = −50 N .
⎣ (λ − 11) (λ − 2) ⎥⎦
For the dynamic loading shown in Fig. 4.22, d, the diagram of dynamic
bending moments is given in Fig. 4.22, e. The maximum bending moment is
87.5 Nm and the maximum dynamic stress is σ d = 14 N mm 2 .

4.4 Coupled Translation and Rotation


When the forces and reactions acting on an elastically supported single
mass do not coincide with the centre of gravity of the mass, the translational and
rotational modes of vibration are coupled. Applications are found at the rigid-body
bouncing and pitching of cars on suspensions and car engines on flexible
mountings, vibrations of rigid rotors in two bearings, vibrations of long masses at
the tip of short cables in Stockbridge dynamic absorbers, vibrations of beams with
end discs.

4.4.1 Equations of Motion

Consider a rigid bar of mass m and mass moment of inertia about the
centre of gravity J supported at its ends on springs of stiffnesses k 1 and k 2 (Fig.
4.23, a). The bar is constrained so that any point can only translate in the vertical
direction. Let the bar motion be defined by two coordinates: x − the linear
displacement of the centre of mass G, and θ − the angle of rotation about G.
Using the free-body diagram of Fig. 4.23, b, the equations of motion can
be written
m &x& + k 1 (x − l 1θ ) + k 2 ( x + l 2 θ ) = 0 ,
Jθ&& − k 1 ( x − l 1θ ) l 1 + k 2 ( x + l 2 θ ) l 2 = 0
or
( ) (
m &x& + k 1 + k 2 x + k 2 l 2 − k 1 l 1 θ = 0, )
J θ&& + ( k 2 l 2 − k 1 l 1 ) x + ( k 1 l 12 + k 2 l 22 )θ = 0.
(4.102)

In matrix form
⎡m 0 ⎤ ⎧ &x& ⎫ ⎡ k1 + k 2 k 2 l 2 − k1 l 1 ⎤ ⎧ x ⎫ ⎧ 0 ⎫
⎢ 0 J ⎥ ⎨ θ&& ⎬ + ⎢k l − k l k l 2 + k l 2 ⎥ ⎨ θ ⎬ = ⎨ 0 ⎬ . (4.102, a)
⎣ ⎦⎩ ⎭ ⎣ 2 2 1 1 1 1 2 2⎦ ⎩ ⎭ ⎩ ⎭
146 MECHANICAL VIBRATIONS

Note in the above equations that if k1l 1 = k 2l 2 the system is uncoupled,


and has two independent natural frequencies – one for translation and one for
rotation. These are defined by

ωx = (k1 + k2 ) m and ω θ = (k l2
1 1 + k 2 l 22 )J. (4.103)

With zero coupling, a force applied to the centre of gravity produces only
up-and-down bouncing x, whereas a torque applied to the bar produces only
pitching motion θ . The coupling is given by the off-diagonal elements of the
stiffness matrix, hence it is referred to as static coupling.

Fig. 4.23

Let the bar motion be defined by other two coordinates, the linear
displacements of the bar ends (points of spring connection) x1 and x2 . The
transformation of coordinates is defined by

⎧x⎫ 1 ⎡ l 2 l 1 ⎤ ⎧ x1 ⎫
⎨ ⎬= ⎢ ⎥ ⎨ ⎬. (4.104)
⎩ θ ⎭ l 1 + l 2 ⎣ − 1 1 ⎦ ⎩ x2 ⎭

Substituting (4.104) into (4.102, a) and multiplying to the left by the


transpose of the transformation matrix from (4.104) yields the equations of motion

⎡ ml 22 + J ml 1l 2 − J ⎤ ⎧&x&1 ⎫ ⎡k 1 (l 1 + l 2 ) 0 ⎤ ⎧ x 1 ⎫ ⎧0⎫
⎢ ⎥ ⎨ ⎬+⎢ ⎨ ⎬ = ⎨ ⎬ . (4.105)
⎣⎢ml 1l 2 − J
2
ml1 + J ⎦⎥ ⎩ &x&2 ⎭ ⎣ 0 k 2 (l 1 + l 2 )⎥⎦ ⎩ x2 ⎭ ⎩0⎭

The coupling is given by the off-diagonal elements of the mass matrix,


hence it is referred to as dynamic coupling. When J = m l 1l 2 the coupled
translation and rotation can be expressed as a sum of two pitching vibrations, one
about the the right end and the other about the left end.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 147

4.4.2 Normal Modes

Denoting

( )
a = (k1 + k 2 ) m , b = (k 2l 2 − k1l 1 ) m , c = k1l 12 + k 2 l 22 J ,
equations (4.97) become
&x& + a x + bθ = 0 ,
(4.106)
r 2θ&& + b x + c r 2θ = 0 ,

where r = J m is the radius of gyration.

Assuming solutions of the form


x (t ) = A x cos (ω t − ϕ ) , θ (t ) = Aθ cos (ω t − ϕ ) , (4.107)

we obtain the set of algebraic equations

( a − ω ) A + b Aθ = 0 ,
2
x
(4.108)
b A + ( c − ω ) r Aθ = 0 .
x
2 2

Equations (4.108) admit non-trivial solutions if the determinant of the


coefficients A x and Aθ is zero

a −ω2 b
=0
b (c − ω ) r 2 2
(4.109)

or

(
ω 4 − (a + c )ω 2 + a c − b 2 r 2 = 0 . ) (4.109, a)
The solutions of the frequency equation (4.109, a) are given by

ω 12, 2 = (a + c ) 2 ± (a − c )2 4 + b2 r 2 . (4.110)

The first natural frequency ω 1 is always less than ω x or ω θ , whichever is


smaller, and the second natural frequency ω 2 is always greater than ω x and ω θ .

The mode shapes are obtained substituting the natural frequencies in turn
in the amplitude ratio

( )
μ 1 = ( A x Aθ )1 = − b a − ω 12 ,
(4.111)
μ 2 = ( Ax Aθ ) = − b ( a − ω ) .
2
2
2
148 MECHANICAL VIBRATIONS

Example 4.8
For the system of Fig. 4.23 determine the natural modes of vibration if
l 1 = 3l 4 , l 2 = l 4 , k1 = k 2 = k , J = m l 2 8 .
Solution. Equations (4.108) have the form
⎛ 2k ⎞ kl
⎜ −ω2 ⎟ Ax + Aθ = 0 ,
⎝ m ⎠ 2m
kl ⎛ 5k ⎞ l2
A x + ⎜⎜ − ω 2 ⎟⎟ Aθ = 0 .
2m ⎝ m ⎠ 8
Denoting
ω 2m
α= ,
k
the equations become
( 2 − α ) A x + l Aθ = 0,
2
l l2
Ax + ( 5 −α ) Aθ = 0 .
2 8
The frequency equation is

α 2 − 7α + 8 = 0
with solutions
α 1 = 1.438 , α 2 = 5.561 .
The natural frequencies are
ω 1 = 1.199 k m , ω 2 = 2.358 k m .

Fig. 4.24
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 149

The mode shapes (shown in Fig. 4.24) are defined by

⎛ Aθ l 4 ⎞ 2 − α 1 ⎛ Aθ l 4 ⎞ 2 −α 2
⎜ ⎟ = = 0.28 , ⎜ ⎟ = = −1.78 .
⎜ Ax ⎟ 2 ⎜ Ax ⎟ 2
⎝ ⎠1 ⎝ ⎠2
Negative inverses of the above amplitude ratios define the location of the
nodal point with respect to the centre of mass, taking a positive value to the right
⎛ Ax ⎞ ⎛ Ax ⎞
d 1 = −⎜ ⎟ = − 1 = −3.56 , d 2 = −⎜ ⎟ = 1 = 0.56 .
⎜ Aθ l 4 ⎟ 0.28 ⎜ ⎟
⎝ ⎠1 ⎝ Aθ l 4 ⎠ 2 1.78
The frequencies of the uncoupled pure translation and pure rotation are

ω x = 1.414 k m , ω θ = 2.236 k m .
The following inequalities hold in this case
ω1 < ω x < ωθ < ω 2 .

Example 4.9
Calculate the natural modes of vibration for the system of Fig. 4.25, a
which consists of a long rigid body of mass m and mass moment of inertia
J = m l 2 8 attached at the end of a cantilever beam of bending rigidity EI.

a b
Fig. 4.25

Solution. Let the rigid body motion be defined by the linear displacement x
of the centre of mass G, and by the angle of rotation θ about G.
150 MECHANICAL VIBRATIONS

The relationship between the inertia force F = mω 2 x , the inertia torque


M = Jω 2θ and the linear and angular displacements x and θ can be expressed by
the equations
x = δ 11 F + δ 12 M ,
θ = δ 21 F + δ 22 M .

Two bending moment diagrams are constructed (Fig. 4.25, b), mx − for
loading with a unit force at point G in the direction of F , and mθ − for loading
with a unit torque at point G in the direction of M . They are used to calculate the
flexibility coefficients using Mohr-Maxwell’s method.
Let the resulting deflections in the directions x and θ be δ 11 and δ 21 ,
respectively δ 12 and δ 22 . The flexibility influence coefficients are

δ11 = 13 l 3 12 E I , δ12 = δ 21 = l 2 E I , δ 22 = l E I .
Substitution in the equations of equilibrium yields
3
β x = 13 x + lθ ,
2
12 3
βθ = x+ θ ,
l 2
where
β = 12 E I m l 3 ω 2 .

Fig. 4.26

The frequency equation is

2β 2 − 29β + 3 = 0
with solutions β 1 = 14.396 and β 2 = 0.1042 .
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 151

The natural frequencies are

ω 1 = 0.913 E I ml 3 , ω 2 = 10.73 E I ml 3 .

The mode shapes (Fig. 4.26) are given by


⎛ Aθ l ⎞ β 1 − 13 ⎛ Aθ l ⎞ β − 13
⎜⎜ ⎟⎟ = = 0.930 , ⎜⎜ ⎟⎟ = 2 = −8.597 .
⎝ Ax ⎠1 32 ⎝ Ax ⎠ 2 32

4.5 Coupled Pendulums


An interesting phenomenon arises in the free vibrations of coupled
pendulums, where a continuous transfer of motion occurs from one pendulum to
the other due to the weak coupling by an elastic element.

4.5.1 Equations of Motion

Consider two simple pendulums (Fig. 4.27, a), each of length l and mass
m, swinging in the vertical plane and coupled together by a light spring of stiffness
k attached at a distance d from the supporting points and unstrained when the
pendulums are in the vertical position.

Fig. 4.27

Using the angular coordinates θ 1 and θ 2 and assuming small amplitudes,


the equations for the kinetic and potential energies are

T=
2
(
1 2 &2 &2
ml θ 1 + θ 2 , ) (4.112)
152 MECHANICAL VIBRATIONS

( )
U = m gl 1 − cosθ 1 + m gl ( 1 − cosθ 2 ) +
1 2
2
(
k d θ2 − θ 1 )2 ,
or
U=
1
2
( 1
)
m g l θ 12 + θ 22 + k d 2 θ 2 − θ 1
2
( )2 . (4.113)

Using Lagrange’s equations

d ∂T ∂T ∂U
− + = 0, r = 1, 2 (4.114)
d t ∂q& r ∂qr ∂qr
we obtain the equations of motion

( )
ml 2θ&&1 + m glθ 1 + k d 2 θ 1 − θ 2 = 0 ,
(4.115)
ml 2θ&&2 + m glθ 2 − k d 2 (θ 1 − θ 2 ) = 0 .

In matrix form, the equations of motion are written


⎡m l 2 0 ⎤ ⎧ θ&&1 ⎫ ⎡ m g l + kd 2 − kd 2 ⎤ ⎧θ1 ⎫ ⎧ 0 ⎫
⎢ ⎥ ⎨ ⎬+⎢ ⎥ ⎨ ⎬ = ⎨ ⎬ . (4.115, a)
⎢⎣ 0 m l 2 ⎥⎦ ⎩ θ&&2 ⎭ ⎢⎣ − k d 2 m g l + k d 2 ⎥⎦ ⎩ θ 2 ⎭ ⎩ 0 ⎭

The coupling is due to the spring k.

4.5.2 Normal Modes

Assuming solutions of the form


θ 1 = a 1 cos (ω t − ϕ ) , θ 2 = a2 cos (ω t − ϕ ) ,
and substituting into the differential equations (4.110), we obtain

(mgl − ω m l + k d ) a
2 2 2
1 − k d 2 a2 = 0 ,
− k d a + (mgl − ω m l
2
1
2 2
)
+ k d 2 a2 = 0.

The frequency equation is

(mgl − ω m l 2 2
+ kd2 ) − (kd )
2 2 2
=0
or
⎛g k d 2 ⎞⎟ 2 ⎛⎜ g 2 k d 2 g ⎞⎟
ω 4 − 2 ⎜⎜ + ω + + 2 =0.
⎝l ml 2 ⎟⎠ ⎜ l2
⎝ ml 3 ⎟⎠

The natural frequencies are


4. TWO-DEGREE-OF-FREEDOM SYSTEMS 153

g g kd2
ω1 = , ω2 = +2 2 . (4.116)
l l ml

The mode shapes are defined by the amplitude ratios


⎛ a2 ⎞ ⎛ a2 ⎞
μ1 = ⎜ ⎟ = +1 , μ2 =⎜ ⎟ = −1 . (4.117)
⎜ a1 ⎟ ⎜ a1 ⎟
⎝ ⎠1 ⎝ ⎠2
In the first mode (Fig. 4.27, b), the pendulums are swinging in phase with
equal amplitudes. The coupling spring is not strained and the pendulums move as if
they were uncoupled. The system natural frequency is equal to that of a single
pendulum g l . In the second mode (Fig. 4.27, c), the two pendulums swing
against each other with equal amplitudes. Due to the stiffening effect of the
coupling spring, the natural frequency is higher than in the first mode.

4.5.3 Free Vibrations

The general solution of free vibrations (4.14) is


θ 1 (t ) = C1 cos (ω 1 t − ϕ1 ) u11 + C2 cos (ω 2 t − ϕ 2 ) u12 ,

θ 2 (t ) = C1 cos (ω 1 t − ϕ1 ) u21 + C2 cos (ω 2 t − ϕ 2 ) u22 ,


or
θ 1 (t ) = a 1 cos (ω 1 t − ϕ 1 ) + a 2 cos (ω 2 t − ϕ 2 ) ,

θ 2 (t ) = μ 1 a 1 cos (ω 1 t − ϕ 1 ) + μ 2 a 2 cos (ω 2 t − ϕ 2 ) .
Differentiating with respect to time yields
θ&1 (t ) = −ω 1 a 1 sin (ω 1 t − ϕ 1 ) − ω 2 a2 sin (ω 2 t − ϕ 2 ) ,

θ& 2 (t ) = − μ 1 ω 1 a 1 sin (ω 1 t − ϕ 1 ) − μ 2 ω 2 a 2 sin (ω 2 t − ϕ 2 ) .


The integration constants are determined from the initial conditions
θ 1 (0 ) = θ 0 , θ 2 (0 ) = 0 , θ&1 (0 ) = 0 , θ& 2 (0 ) = 0 . (4.118)

Because μ1 = 1 and μ 2 = −1 , we obtain

a 1 cos ϕ 1 + a 2 cos ϕ 2 = θ 0 , a 1 cos ϕ 1 − a 2 cos ϕ 2 = 0 ,

ω 1 a 1 sinϕ 1 + ω 2 a 2 sinϕ 2 = 0 , ω 1 a 1 sinϕ 1 − ω 2 a 2 sinϕ 2 = 0 ,

so that sin ϕ 1 = sin ϕ 2 = 0 , ϕ 1 = ϕ 2 = 0 ,


154 MECHANICAL VIBRATIONS

2 a 1 cosϕ 1 = 2 a 2 cosϕ 2 = θ 0 , a 1 = a2 = θ 0 2 .

The instantaneous values of the angular displacements are


θ0
θ 1 (t ) =
2
( cos ω 1 t + cos ω 2 t ) = θ0 cos ω m t ⋅ cos Δω t ,
(4.119)
θ0
θ 2 (t ) =
2
( cos ω 1 t − cos ω 2 t ) = θ0 sin ω m t ⋅ sin Δω t ,
(
where ωm = ω 1 + ω 2 ) 2 and Δω = ( ω 2 − ω 1 ) 2 .
When Δω is small with respect to ωm , the products in the above
equations represent amplitude modulated oscillations known as beats. This
condition is equivalent to
g kd2 g g g ml
ω22 − ω 12 << ω 12 , + 2 2 − << or k << ,
l ml l l 2d2
so that beats can occur only for small stiffness values k of the coupling spring.

Fig. 4.28

Equations (4.119) show that θ 1 and θ 2 are given by sine and cos
functions which are 90 0 phase shifted. When θ 1 = θ 0 , θ 2 = 0 and vice-versa
(Fig.4.28).
The phase difference between the two motions is
ϕ = (ω 2 t − ϕ 2 ) − (ω 1 t − ϕ 1 ) = (ω 2 − ω 1 ) t − (ϕ 2 − ϕ 1 ) = 2Δω t − Δϕ
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 155

so it varies slowly with time.


At t = 0 , θ1 = θ 0 , θ 2 = 0 .

ωm π ω
At Δω t = π 2 , θ1 = 0 , θ 2 = θ 0 sin = θ 0 if m = n is an integer.
Δω 2 Δω
ωm ω
At Δω t = π , θ 2 = 0 , θ 1 = θ 0 cos π = −θ 0 if m = n is an integer.
Δω Δω
The initial conditions (4.118) imply that the left pendulum is pulled out
while keeping the other in place, then released (Fig. 4.29, a). For the first few
cycles the left pendulum will swing, the right pendulum stands almost still. Then
the right pendulum starts oscillating with increasing amplitudes, while the
amplitude of oscillation of the left pendulum decreases. After a sufficient time
interval, the left pendulum stands still, while the right pendulum swings with the
full amplitude (Fig. 4.29, d). Then the phenomenon repeats itself. There is a
continuous transfer of energy from one pendulum to the other until the inherent
damping (neglected in this analysis) brings the system to rest.

Fig. 4.29

In terms of the component mode shapes, the motion can be regarded as the
sum of two harmonic motions at the natural frequencies ω 1 and ω 2 . Let initially
the motion in the second mode be such that the pendulums are departed from each
other (Fig. 4.29, b). The phase difference between modes grows with the time
elapsed. The motion in the second mode is faster than in the first mode, until it is
1800 degrees in advance (Fig. 4.29, c). The motion in the second mode is such that
the pendulums are closer to each other. Summing up the motions, it turns out that
at a certain time the left pendulum stands still, while the right one swings with
maximum amplitude. Then the amplitude wanders to the left pendulum and the
whole sequence repeats itself.
156 MECHANICAL VIBRATIONS

4.6 Damped Systems


In the preceding sections only undamped systems have been considered.
Herein the effect of viscous damping is taken into account. Arbitrary damping
introduces coupling of coordinates so that only the special case of proportional
damping is considered. Low damping values limit the vibration amplitude at
resonance. Larger values may make modes to be overdamped so that the
corresponding motion is aperiodic.

4.6.1 Proportional Viscous Damping


Consider the system of Fig. 4.30, a. Energy dissipation is conveniently
modelled by viscous damping represented by the dashpots c1 and c2 . Viscous
damping forces are proportional and opposite to the relative velocity of the dashpot
ends (Fig. 4.30, b). The equations of motion may be written by summing the forces
acting on each mass:
( ) ( )
m 1 &x&1 + k 1 x 1 + c 1 x& 1+ k 2 x 1 − x 2 + c2 x& 1 − x& 2 = 0 ,
m 2 &x& 2 − k 2 ( x 1 − x 2 ) − c2 ( x& 1 − x& 2 ) = 0

or
m 1 &x&1 + ( c 1 + c2 ) x& 1− c2 x& 2 + ( k1 + k 2 ) x 1 − k 2 x 2 = 0 ,
(4.120)
m 2 &x& 2 − c2 x& 1 + c2 x& 2 − k 2 x 1 + k 2 x 2 = 0 .

Fig. 4.30

The coupling between the two coordinates is due to both the stiffness k 2
and the damping coefficient c2 .
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 157

In matrix form, these equations can be written


⎡m1 0 ⎤ ⎧ &x&1 ⎫ ⎡c1 + c2 − c2 ⎤ ⎧ x& 1 ⎫ ⎡k1 + k 2 − k 2 ⎤ ⎧ x 1 ⎫ ⎧0⎫
⎢ 0 m ⎥ ⎨ &x& ⎬ + ⎢ − c ⎨ ⎬+ ⎨ ⎬ = ⎨ ⎬, (4.121)
⎣ 2⎦ ⎩ 2⎭ ⎣ 2 c2 ⎥⎦ ⎩ x&2 ⎭ ⎢⎣ − k 2 k 2 ⎥⎦ ⎩ x2 ⎭ ⎩0⎭

or in compact form
[ m ]{ &x& } + [ c ]{ x& } + [ k ]{ x } = { 0 }, (4.121, a)

where [ c ] is the damping matrix, and { x& } is the column vector of velocities.
Note that the damping matrix is symmetrical so that

[ c ] = [ c ]T .
The mass matrix is diagonal. The coupling is produced by the off-diagonal
elements of the stiffness matrix and the damping matrix.
Modal analysis may be used to solve equations (4.121) if the linear
transformation based on the modal matrix diagonalizes the damping matrix
simultaneously with the mass and the stiffness matrix. This is simply achieved if
the damping matrix can be expressed as a linear combination of the mass and
stiffness matrices, that is, if
[ c ]= α [ m ] + β [ k ] , (4.122)
where α and β are constants. This form of damping is called proportional
damping or Rayleigh damping. There are also other conditions when the modal
damping matrix becomes diagonal, but they are only special cases which occur
seldom.

4.6.2 Damped Free Vibrations


If the undamped problem (4.3) is first solved, one obtains the undamped
modes of vibration. The modal matrix (4.24) is constructed having the normal
modes as columns. Using the transformation of coordinates
{ x } = [ u ] {q } (4.123)

and multiplying on the left by [ u ] T we obtain


[ M ] { q&& } + [ C ] { q& } + [ K ] { q } = { 0 } , (4.124)
where

[ M ] = [ u ]T [ m ] [ u ] , [ C ] = [ u ]T [ c ] [ u ], [ K ] = [ u ]T [ k ] [ u ]. (4.125)

The modal matrices [ M ] and [ K ] are diagonal while the matrix [ C ]


is diagonal only for proportional damping when
158 MECHANICAL VIBRATIONS

[ C ]= α [ M ] + β [ K ] . (4.126)
In this case, the following orthogonality relations can be established

{ u }Ts [ c ] { u }r = 0 , r ≠ s , r , s = 1, 2 (4.127)

For proportional damping, the decoupled modal equations are


M r q&&r + Cr q& r + K r qr = 0 , r = 1, 2 (4.128)

where M r and K r are defined by (4.26) and (4.28), and

Cr = { u }Tr [ c ] { u }r , r = 1, 2 (4.129)

are modal damping coefficients.


For unit modal masses, the modal stiffnesses are equal to the square of the
respective natural frequency and the modal damping coefficients can be expressed
as 2 ζ r ωr where ζ r is the rth modal damping ratio and ωr is the rth natural
frequency.
Equations (4.128) become

q&&r + 2ζ r ωr q& r + ω 2r qr = 0 , r = 1, 2 (4.130)

which, for 0 < ζ r < 1 have solutions of the form (2.46)

qr (t ) = Ar e − ζ r ω r t sin ⎛⎜ 1 − ζ r2 ωr t + φr ⎞⎟ . r = 1, 2 (4.131)
⎝ ⎠
The solutions (4.131) can also be obtained directly. Looking for solutions
of the form xr = ar e s t , equations (4.121) become

m 1 s 2 a 1 + (c1 + c2 ) s a 1 − c2 s a2 + (k1 + k 2 ) a 1 − k 2 a2 = 0 ,
m2 s 2 a2 − c2 s a 1 + c2 s a2 − k 2 a 1 + k 2 a2 = 0 ,

or

[m s 1
2
]
+ (c1 + c2 ) s + (k1 + k 2 ) a1 − ( c2 s + k 2 ) a2 = 0,
(4.132)
(
− ( c2 s + k 2 ) a1 + m 2 s 2 + c2 s + k 2 a2 = 0. )
The condition to have non-trivial solutions leads to the characteristic
equation

[m s
1
2
+ (c1 + c2 ) s + (k1 + k 2 ) ] (m s
2
2
)
+ c2 s + k 2 − ( c2 s + k 2 ) 2 = 0 . (4.133)
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 159

For underdamped systems, equation (4.133) has two pairs of complex


conjugate roots
s 1, 2 = −σ 1 ± i ω d 1 , s 3, 4 = −σ 2 ± iω d 2 , (4.134)

where ω d 1 and ω d 2 are the damped natural frequencies and σ 1 and σ 2 are the
damping factors (exponential decay rates).
The above parameters can be related to the absolute values of the
eigenvalues (equal to the undamped natural frequencies for systems with
proportional damping) and the damping ratios as follows

σ r = ζ r ωr , ω d r = ωr 1 − ζ 2r r = 1, 2 (4.135, a)

σr σr
ζr = , ωr = = ω 2d r + σ r2 . r = 1, 2 (4.135, b)
ω 2d r + σ r2 ζr

With these notations, expression (4.131) becomes

(
qr ( t ) = Ar e − σ r t sin ω d r t + φr . ) r = 1, 2 (4.131, a)

For relatively high damping values, equation (4.133) may have two real
roots and two complex conjugate roots, when one mode of vibration is
overdamped, or two pairs of real roots, when both modes are overdamped and the
system has no vibratory motion. These cases are not considered herein.
Substituting solutions (4.134) into (4.132) we obtain the amplitude ratios
(a2 a1 )r which, for proportional damping, define real modes of vibration. The
above analysis is valid for distinct eigenfrequencies. The case of equal eigenvalues
is treated elsewhere.

Example 4.10
Calculate the modes of vibration for the system of Fig. 4.30 taking for
simplification m 1 = 2 , k 1 = 2 , c 1 = 1 , m 2 = 1 , k 2 = 1 , c 2 = 0.5 , in the
appropriate units.
Solution. The equations of free vibrations (4.121) are
⎡2 0⎤ ⎧ &x&1 ⎫ ⎡ 1.5 − 0.5⎤ ⎧ x& 1 ⎫ ⎡ 3 − 1⎤ ⎧ x 1 ⎫ ⎧ 0 ⎫
⎢0 1⎥ ⎨ &x& ⎬ + ⎢− 0.5 0.5 ⎥ ⎨ x& ⎬ + ⎢− 1 1 ⎥ ⎨ x ⎬ = ⎨ 0 ⎬ .
⎣ ⎦⎩ 2⎭ ⎣ ⎦⎩ 2⎭ ⎣ ⎦⎩ 2⎭ ⎩ ⎭
It is seen that
[ c ]= 0.5 [ k ] .
Equations (4.132) become
160 MECHANICAL VIBRATIONS

(2 s 2
)
+ 1.5 s + 3 a 1 − ( 0.5 s + 1 ) a 2 = 0 ,
(
− ( 0.5 s + 1) a 1 + s 2 + 0.5 s + 1 a 2 = 0. )
The characteristic equation is

(2 s 2
+ 1.5 s + 3 )( s 2
)
+ 0.5 s + 1 − ( 0.5 s + 1 )2 = 0 ,
or
2 s 4 + 2.5 s 3 + 5.5 s 2 + 2 s + 2 = 0 ,
with roots
1 31 1 7
s 1, 2 = − ± i , s 3,4 = − ± i .
8 8 2 2
The imaginary parts are the damped natural frequencies
ωd 1 = 31 8 = 0.6960 , ωd 2 = 7 2 = 1.3229 .
The real parts are the damping factors (decay rates)
σ 1 = 0.125 , σ 2 = 0 .5 .
The undamped natural frequencies are equal to the absolute value of the
eigenvalue

ω 1 = ωd21 + σ 12 = ( 31 8) 2
+ (1 8) 2 = 1 2,

ω 2 = ωd2 2 + σ 22 = ( 7 2) 2
+ (1 2) 2 = 2 .

The modal damping ratios are


σ1 18 σ2 12
ζ1 = = = 0.176 , ζ 2 = = = 0.353 .
ω1 1 2 ω2 2

The amplitude ratio


a2 2 s 2 + 1.5 s + 3 0.5 s + 1
= = 2
a1 0.5 s + 1 s + 0.5 s + 1
yields
⎛ a2 ⎞ ⎛ a2 ⎞
⎜ ⎟ =2, ⎜ ⎟ = −1 .
⎜ a1 ⎟ ⎜ a1 ⎟
⎝ ⎠1 ⎝ ⎠2
Solving the undamped problem in Example 4.1, the same undamped
natural frequencies and mode shapes have been obtained.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 161

Example 4.11
Calculate the modal parameters for the system of Fig. 4.30 using the same
values for masses and stiffnesses but lower damping [ c ]= 0.1 [ k ] .
Solution. Equations (4.132) become

(2 s 2
)
+ 0.3 s + 3 a 1 − (0.1 s + 1 ) a 2 = 0 ,
( )
− ( 0.1 s + 1) a 1 + s 2 + 0.1 s + 1 a 2 = 0.

The characteristic equation


2 s 4 + 0.5 s 3 + 5.02 s 2 + 0.4 s + 2 = 0 ,
has the roots
1 799 1 199
s 1, 2 = − ±i , s 3, 4 = − ± i
40 40 10 10
or
s 1, 2 = −0.025 ± i 0.706665 , s 3, 4 = −0.1 ± i 1.410674 .

The damped natural frequencies are


ωd 1 = 0.70666 , ωd 2 = 1.4107 .
The damping factors have the values
σ 1 = 0.025 , σ 2 = 0 .1 .
The modal damping ratios are
ζ 1 = 0.0353 , ζ 2 = 0.0707 .
The undamped natural frequencies and the mode shapes are the same as in
Example 4.10.

4.6.3 Response to Harmonic Excitation


Consider the vibrations of the two-degree-of-freedom system from Fig.
4.31 under the action of forces f1 (t ) and f 2 (t ) .

Fig. 4.31
162 MECHANICAL VIBRATIONS

The equations of motion are


m 1 &x&1 + (c1 + c2 ) x& 1 − c2 x&2 + (k1 + k 2 ) x 1 − k 2 x2 = f1 (t ) ,
(4.136)
m 2 &x&2 − c2 x& 1 + c2 x&2 − k 2 x 1 + k 2 x2 = f 2 (t ) ,

or in compact matrix form


[ m ] { &x& } + [ c ] { x& } + [ k ] { x } = { f } . (4.136, a)

4.6.3.1 Transfer Functions


Taking the Laplace transform of (4.136, a) and assuming that all initial
conditions are zero, yields

[[ m ] s 2
+ [c ] s + [k ] ] { X (s ) } = { F (s ) } (4.137)
or
[ B (s ) ] { X (s ) } = { F (s ) } (4.138)

where [ B (s ) ] is referred to as the system matrix.

Multiplying on the left by [ B(s ) ] −1 = [ H (s ) ] yields


[ H (s ) ] { F (s ) } = { X (s ) } (4.139)

where [ H (s ) ] is referred to as the transfer function matrix. It is the inverse of the


system matrix.
For the system of Fig. 4.31 this matrix has the form

⎡ m 2 s 2 + c2 s + k 2 c2 s + k 2 ⎤
⎢ ⎥
⎢⎣ c2 s + k 2 m 1 s + (c1 + c2 ) s + k1 + k 2 ⎥⎦
2
[H ( s )] = (4.140)
[m s
1
2
+ (c1 + c2 ) s + k1 + k 2 ] [m 2s
2
]
+ c 2 s + k 2 − ( c2 s + k 2 ) 2

The denominator of (4.140) is det [B (s )] which is the system characteristic


polynomial. This can be expressed as a product

( )( )(
det [ B (s ) ] = A s − s 1 s − s 2 s − s 3 s − s 4 )( )
where A is a constant and s 1 ,.., s 4 are the roots of the characteristic equation
(4.133). Since the coefficients of the characteristic equation are real, for low
damping levels the roots appear as complex conjugate pairs. They are called the
poles of the transfer function.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 163

We can also express (4.139) as

⎡ h11 (s ) h12 (s ) ⎤ ⎧ F1 (s ) ⎫ ⎧ X 1 (s ) ⎫
⎢ h (s ) h (s ) ⎥ ⎨ F (s ) ⎬ = ⎨ ⎬. (4.141)
⎣ 21 22 ⎦ ⎩ 2 ⎭ ⎩ X 2 (s ) ⎭
where h l j (s ) is obtained by exciting the system at point j and measuring the
response at point l . For instance h11 (s ) is a driving point transfer function
measured by exciting the system with F1 (s ) and measuring the response X 1 (s )

X 1 (s ) m2 s 2 + c2 s + k 2
h11 (s ) = = . (4.142)
F1 (s )
( )( )( )(
A s − s1 s − s 2 s − s 3 s − s 4 )
4.6.3.2 Frequency Response Functions
A frequency response function (FRF) is the transfer function evaluated
along the iω (frequency) axis. Substituting s = iω into (4.141) we obtain

⎡ h11 (iω ) h 12 (iω ) ⎤ ⎧ F1 (ω ) ⎫ ⎧ X 1 (ω ) ⎫


⎢ h (iω ) h (iω ) ⎥ ⎨ F (ω ) ⎬ = ⎨ ⎬
⎣ 21 22 ⎦⎩ 2 ⎭ ⎩ X 2 (ω ) ⎭
or
[ H (iω ) ] { F (ω ) } = { X (ω ) } (4.143)
where [ H (iω ) ] is the frequency response function matrix and X j (ω ) , F j (ω ) are
Fourier transforms of the response and excitation, respectively.
The response of the system can be defined in the frequency domain by the
multiplication and addition of measured frequency response functions and forcing
functions
X 1 (ω ) = h 11 (iω ) F1 (ω ) + h12 (iω ) F 2 (ω ),
(4.143, a)
X 2 (ω ) = h 21 (iω ) F1 (ω ) + h 22 (iω ) F 2 (ω ).

The time domain forced response of the system, x 1 (t ) and x 2 (t ) , can then
be computed by taking the inverse Fourier transform of X 1 (ω ) and X 2 (ω ) .

Example 4.12
Calculate the matrix of frequency response functions for the two-degree-
of-freedom system of Example 4.11 and plot the FRF curves of h11 (iω ) .
164 MECHANICAL VIBRATIONS

Solution. For s = iω , the matrix of FRFs (4.140) becomes

⎡1 − ω 2 + i 0.1ω 1 + i 0.1ω ⎤
⎢ ⎥
⎢⎣ 1 + i 0.1ω 3 − 2ω 2 + i 0.3ω ⎥⎦
[H (iω )] = .
(
2 ω 4 − 5.02 ω 2 + 2 + i − 0.5 ω 3 + 0.4 ω )

Fig. 4.32

Fig. 4.33
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 165

Fig. 4.34

The plots of the FRF magnitude and phase angle are given in Fig. 4.32.
The magnitude plot has two resonance peaks. Because of the larger value of
damping in the second mode, the respective resonance peak is lower. At
resonances, the phase angle is about − 900 . The plots of the real part and imaginary
part are presented in Fig. 4.33. The Nyquist plot is illustrated in Fig. 4.34 with
points marked at equal frequency increments.

4.6.3.3 Solution by Modal Analysis


Using the transformation of coordinates (4.123) { x } = [ u ]{ q } and
multiplying on the left by [ u ] T equations (4.136, a) become

[ M ] { q&& } + [ C ] { q& } + [ K ] { q } = [ u ] T { f } = { F }. (4.144)

For proportional damping, the decoupled modal equations are

M r q&&r + Cr q&r + K r qr = Fr , r = 1, 2 (4.145)


or
q&&r + 2ζ r ωr q& r + ω 2r qr = Fr M r . r = 1, 2 (4.146)

For harmonic excitation and steady-state response, denote

{ f } = { f̂ }e ωt ,
i
{ x } = { ~x }ei ω t , (4.147)
166 MECHANICAL VIBRATIONS

{ F } = { F̂ } eiω t , { q } = { q~ }ei ω t , (4.148)


2
{ ~x } = [ u ] { q~ } = ∑ q~r { u }r , (4.149)
r =1

where a hat above a letter means real amplitude and a tilde above a letter denotes
complex amplitude.
Substitute (4.148) into equations (4.145) and (4.146) to obtain

{ u }Tr { f̂ } { u }Tr { f̂ }
q~r = = . (4.150)
K r − ω 2 M r + i ω Cr (
M r ω 2r 2
− ω + i 2 ζ r ω ωr )
For unit modal masses, M r = 1 , substituting (4.150) into (4.149) yields

2
{u }Tr { f̂ } {u }r
{ ~x } = ∑
r =1 ω 2r − ω 2 + i 2 ζ rω ωr
or

⎧~x1 ⎫ 2 { u }r { u }Tr ⎧⎪ f̂1 ⎫⎪


⎨~ ⎬ = 2∑ 2
⎩ x2 ⎭ r =1 ω r − ω + i 2 ζ r ω ωr
⎨ ⎬.
⎪⎩ f̂ 2 ⎪⎭
(4.151)

Note that the dyadic product { u }r { u }Tr is a 2 by 2 square matrix.

The FRF matrix (4.143) can be expressed now in terms of the modal
parameters

{u }1{u }T1 {u }2 {u }T2


[ H (iω ) ] = + . (4.152)
ω 12 − ω 2 + i 2 ζ1ω ω 1 ω 22 − ω 2 + i 2 ζ 2ω ω 2
If we denote the jth element of the rth mode shape vector by u r ( )j then
the expression for the general FRF function h j l (iω ) in a partial fraction form is

( u 1 ) j ( u 1 )l ( u 2 ) j ( u 2 )l
h j l (iω ) = + . (4.153)
ω 12 − ω 2 + i 2 ζ1ω ω 1 ω 22 − ω 2 + i 2 ζ 2 ω ω 2

This expression shows explicitly the contribution to the response at any


frequency of each individual mode of the system. Figure 4.35 shows the polar plot
of a two-degree-of-freedom system FRF by plotting each term in the series form
expression separately (dotted lines), and then the summation of these terms (solid
line) i.e. the complete FRF.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 167

Fig. 4.35

Example 4.13
Calculate the partial fraction form of the FRF matrix for the two-degree-of-
freedom system of Example 4.12.
Solution. The mass-normalised mode shapes are
T T
{ u }1 = ⎧⎨1 2 ⎫ ⎧ 1
⎬ , {u } 2 = ⎨ −
1 ⎫
⎬ .
⎩ 6 6⎭ ⎩ 3 3⎭
The FRF matrix (4.152) becomes
1 ⎡1 2⎤ 1 ⎡ 1 − 1⎤
6 ⎢ 2 4⎥ 3 ⎢⎣ − 1 1⎥⎦
[ H (iω ) ] = ⎣ ⎦ + .
0.5 − ω 2 + i 0.05 ω 2 − ω 2 + i 0 .2 ω
The driving point FRF for the first mass is
16 13
h11 (iω ) = 2
+ 2
.
0.5 − ω + i 0.05 ω 2 − ω + i 0.2 ω
Figure 4.35 shows the Nyquist plot (solid line) which results from the
summation of the plots of each term in the series form expression (broken lines).
168 MECHANICAL VIBRATIONS

4.6.4 The Untuned Viscous Damper


The untuned viscous vibration damper (Houdaille damper) is used in
reciprocating engines to limit the amplitudes of torsional vibrations over a wide
operating range. It consists of a rigid disc free to rotate within a cylindrical cavity
filled with viscous fluid. In car engines it is located at the end of the crankshaft in
the pulley which drives the fan belt.
Consider the crankshaft of torsional stiffness K fixed at one end and with
the damper attached at the other end in a casing of rotational inertia J (Fig. 4.36).
The free rotational mass has a polar mass moment of inertia J d and is acted upon
by the damping torque which is assumed to be proportional to the relative motion
between the casing and the internal disc. When the damper is excited by a
harmonic torque M 0 cosω t , the equations of motion can be written

( )
J θ&& + K θ + c θ& − θ&d = M 0 cosω t ,
(4.154)
d d ( d )
J θ&& − c θ& − θ& = 0 ,
where θ is the casing rotation and θ d is the rotation of the internal disc. The
damping coefficient is
⎛ b R 3 R24 − R 14 ⎞
c = 2π μ ⎜ 2 + ⎟ (4.155)
⎜ h2 2h 1 ⎟
⎝ ⎠
where μ is the oil viscosity.

Fig. 4.36

Assuming solutions of the form


θ = θ 0 cosω t , θ d = θ d 0 cos (ω t − ϕ ) ,
and denoting
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 169

K ω c J
ω n2 = ,η= , ζ= , λ= d , (4.156)
J ωn 2 J d ωn J
the amplitude of the casing is described by

K θ0 η 2 + 4 ζ2
= . (4.157)
M0
(
η 2 1 −η 2 ) 2
( )
+ 4 ζ 2 η 2 + λη 2 − 1
2

For a fixed λ , the plots of K θ 0 M 0 as a function of the dimensionless


frequency η are shown in Fig. 4.37 for various values of ζ .

Fig. 4.37

The value ζ = 0 is for an undamped system with resonance frequency ωn


whose vibration amplitude is infinite at η = 1 . The value ζ = ∞ is for an undamped
system with resonance frequency K (J + J d ) = ωn 1 + λ , in which the damper
mass and the wheel move together as a single mass. The response curves for these
two extreme values of ζ cross each other at the point M of abscissa
ηM = 2 (2 + λ ) . For any other value of damping, the response curve passes
through this point. There is an optimum damping
1
ζ opt = (4.158)
2 (1 + λ )(2 + λ )

for which the peak amplitude is a minimum and is equal to the ordinate 1 + (2 λ ) of
point M .
170 MECHANICAL VIBRATIONS

Based on equations (4.156) and (4.158) a torsional vibration damper may


be designed with a flat response curve, which is effective over a wide frequency
range.

4.6.5 The Damped Vibration Absorber

A vibration absorber consists of a second mass-spring system added to a


primary mass-spring system to protect it from vibrating. The major effect of adding
the secondary system is to change from a single degree of freedom system to a
two-degree-of-freedom system. The values of the absorber physical parameters are
chosen such that the motion of the main system is a minimum. This is accompanied
by substantial motion of the added mass which “absorbs” the energy introduced in
the system by the force acting on the original mass. If damping in the secondary
system is taken into account, the deflection of the primary mass cannot be reduced
to zero, but the useful frequency range of the absorber increases, improving its
effectiveness.
Consider the harmonic response of a two-degree-of-freedom system with
undamped primary system and viscously damped absorber. Such a system is
obtained if the dashpot c1 is eliminated from the system illustrated in Fig. 4.31.

If the damping coefficient c1 is made zero, and the force f 2 is cancelled,


equations (4.136) may be rewritten as
( ) ( )
m 1 &x&1 + c2 x& 1 − x&2 + k 1 x 1 + k 2 x 1 − x2 = f1 (t ) ,
m 2 &x&2 − c2 ( x& 1 − x&2 ) − k 2 ( x 1 − x2 ) = 0 .
(4.159)

For harmonic excitation and steady-state response


~ ~
f1 = F eiω t , x 1 = X 1 ei ω t , x2 = X 2 ei ω t , (4.160)

equations (4.159) become

(k + k
1 2
~
) ~
− m 1 ω 2 + i ω c2 X 1 − ( i ω c2 + k 2 ) X 2 = F ,
(4.161)
~
( ~
)
− ( i ω c2 + k 2 ) X 1 + k 2 − m 2 ω 2 + i ω c2 X 2 = 0 .
~
The vibration amplitude of the primary system X 1 = X 1 can be expressed
as

X1 A ζ 22 + B
= (4.162)
xst C ζ 22 + D

where
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 171

(
A = 4 μ 2ψ 2θ 2 , B = μ 2 ψ 2 − θ 2 ) 2
(
, C = 4 μ 2ψ 2θ 2 1 − θ 2 − μ θ 2 ) 2
,

[ ( )( )
D = μ ψ 2 − θ 2 1 − θ 2 − μ 2ψ 2θ 2 ] 2
, xst = F k 1 , ζ 2 = c2 2 ω 2 m2 , (4.163)

μ = m2 m 1 , θ = ω ω 1 , ψ = ω 2 ω 1 , ω 1 = k 1 m 1 , ω2 = k 2 m2 .

Figure 4.38 shows the frequency response curves of X 1 xst as function of


the frequency ratio θ , for fixed values of μ and ψ , and various values of ζ 2 . The
effect of the absorber can be seen to reduce the primary system amplitude from
infinity down to a small value at θ = 1 .

When ζ 2 = 0 , X 1 xst = B D . When ζ 2 = ∞ , X 1 xst = A C . The


corresponding response curves cross each other at points R and S. All response
curves of the primary mass in the combined system, drawn for various values of
ζ 2 , pass through these two points.
By changing the frequency ratio ψ , the ordinates of these two points can
be increased or decreased. The most favourable case, for the lowest maximum
dynamic response throughout the whole frequency range, could be achieved when
the following two conditions could be fulfilled:
Firstly, arranging for these two points to have equal ordinates. This is
obtained when
ψ = 1 (1 + μ ) . (4.164)
Secondly, arranging the slope of the curve to be zero at each point, that is
the two points become the two peaks in the response curve, indicating that maxima
have been reached. Unfortunately it is not possible to have equal maxima occurring
for the same frequency ratio ψ , but if the slope at one point is zero the slope at the
other point is nearly zero. This is obtained when the response of the main system is
given by
X 1 xst = 1+ 2 μ . (4.165)

The frequency ratios at which the two peaks occur are given by

1± 2 (2 + μ )
θ2 = . (4.166)
1+ μ
The frequency ratio θ at which the trough between the two peaks occurs is
the mean of these two frequency ratios 1 (1 + μ ) = ψ 1 2 . This is in fact the tuning
frequency ratio.
172 MECHANICAL VIBRATIONS

The optimum damping coefficient resulting from the application of these


two considerations may be found by differentiating equation (4.162) with respect to
θ and equating to zero for each crossover point, giving two values of
ζ 2 ψ = c2 2 m2 ω 1 . The average of these two values is

c 2 2 m2 ω 1 = 3 μ 8 (1 + μ )3 . (4.167)

Fig. 4.38

It is also important to keep low the value of the relative motion between
the main mass and the absorber
X1 F
X1 − X 2 = (4.168)
(
2 μ k 1 θ c2 2 m 2 ω 1 )
to ensure longer fatigue life for the absorber spring.

4.6.6 Non-Proportional Viscous Damping

Consider the system of Fig. 4.30, a with non-proportional viscous


damping. The equations of free vibrations (4.121, a) are

[ m ]{ &x& } + [ c ]{ x& } + [ k ]{ x } = { 0 }, (4.169)


4. TWO-DEGREE-OF-FREEDOM SYSTEMS 173

where the matrix [c ] in not proportional with either [ m ] or [ k ] .


For computational purposes, the equation of motion (4.169) is transformed
into the first order state space form. Introducing an auxiliary equation
[ m ] { x& } − [ m ] { x& } = { 0 } . (4.170)
Equations (4.169) and (4.170) can be combined to give

⎡[ m ] [ 0 ] ⎤ ⎧{ x& }⎫ ⎡[ 0 ] − [ m ]⎤ ⎧{ x }⎫ ⎧{ 0 }⎫
+ = (4.171)
⎢[0 ]
⎣ [ m ]⎥⎦ ⎨⎩{ &x& }⎬⎭ ⎢⎣[ k ] [ c ] ⎥⎦ ⎨⎩{ x& }⎬⎭ ⎨⎩{ 0 }⎬⎭
or
[ A ] { q& } + [ B ] { q } = { 0 } , (4.172)
where the 4 to 4 matrices [ A] and [B ] are real

[ m ] [ 0 ]⎤ [ 0 ] − [ m ]⎤
[A] = ⎡⎢ ⎥ , [B] = ⎡⎢ ⎥.
⎣ [ 0 ] [ m ]⎦ ⎣[ k ] [ c ] ⎦

Assuming a solution of the form { q } = {Φ }e s t , and { x } = { u }e s t ,


equations (4.172) become
( s [ A ] + [ B ] ) {Φ } = { 0 }. (4.173)

There are four eigenvalues s r given by the solutions of

det ( s [ A ] + [ B ] ) = 0 , (4.174)
and four eigenvectors {Φ }r satisfying the eigenvalue problem

[ B ] {Φ } r = − s r [ A ]{Φ }r . (r = 1,...,4) (4.175)

Equation (4.175) can be written in the form

− [ A ] −1[ B ] {Φ } r = s r {Φ }r . (r = 1,...,4) (4.176)


where
⎡ [0 ] [I ] ⎤
− [ A ] −1 [ B ] = ⎢ ⎥, (4.177)
⎣− [ m ] [ k ] − [ m ] [ c ]⎦
−1 −1

Equation (4.171) becomes

⎛⎡ [0 ] [ I ] ⎤ ⎡[ I ] [ 0 ]⎤ ⎞⎟ ⎧ { u } ⎫
⎜⎢
⎜ − [ m ] −1 [ k ] − [ m ] −1 [ c ]⎥ − s ⎢ [ 0 ] = { 0 }. (4.178)
⎝⎣ ⎦ ⎣ [ I ]⎥⎦ ⎟⎠ ⎨⎩s{ u }⎬⎭
174 MECHANICAL VIBRATIONS

For underdamped systems, equation (4.178) has two pairs of complex


conjugate roots
s 1, 2 = −σ 1 ± iω d 1 , s 3, 4 = −σ 2 ± iω d 2 , (4.179)

where the imaginary parts, ω d 1 and ω d 2 , are the damped natural frequencies and
the real parts, σ 1 and σ 2 , are the damping factors (decay rates).

The above parameters can be related to the absolute values of the natural
frequencies and the damping ratios as follows
σr
ω r = ω d2 r + σ r2 , ζr = . r = 1, 2 (4.180)
ωr
For overdamped modes of vibration, the real eigenvalues can be written
s r , s r +1 = − σ r m τ r , (4.181)

and the relationships (4.180) become


σr
ω r = σ r2 − τ r2 , ζr = . (4.182)
ωr
The motion in an overdamped mode of vibration is aperiodic so that in the
frequency response curves there is no related resonance peak or loop in the Nyquist
plot even for relatively spaced natural frequencies.
For underdamped systems, equation (4.178) yields also two pairs of
complex conjugate eigenvectors, whose half upper part gives the mode shapes. The
relative phase differences among the degrees of freedom in a particular mode cause
the maximum mass displacements to be reached at different times.

Example 4.14
Calculate the modal parameters for the system with closely spaced natural
frequencies of Fig. 4.31, where m 1 = 100 kg , m 2 = 1 kg , k 1 = 9.9 ⋅ 106 N m ,
k 2 = 0.1 ⋅ 106 N m , c 1 = c 2 = 125 Ns m . Draw the Nyquist plots of complex
receptances.
Solution. The mass, stiffness and damping matrices are, respectively
⎡100 − 1⎤ ⎡ 250 − 125⎤
[ m ] = ⎡⎢
100 0⎤
⎥ , [ k ] = 105 ⋅ ⎢ ⎥ , [c ]= ⎢ ⎥ .
⎣ 0 1⎦ ⎣ −1 1 ⎦ ⎣− 125 125 ⎦
The eigenvalues with positive imaginary part are
λ 1= −4.785 + i ⋅ 312.659 , λ 2 = −58.965 + i ⋅ 312.685 .
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 175

Fig. 4.39

The corresponding eigenvectors are

{ u }1 = ⎧⎨ ⎫
{u } 2 = ⎧⎨ ⎫
1 1
⎬, ⎬.
⎩1.207 − i ⋅ 2.698⎭ ⎩− 8.865 − i ⋅ 35.222⎭
176 MECHANICAL VIBRATIONS

The damped natural frequencies are ω d 1 = 312.659 rad sec and


ω d 2 = 312.685 rad sec . The absolute values of the eigenvalues are

ω 1 = ω d2 1+ σ 12 = 312.695 rad sec , ω 2 = ω d2 2 + σ 2


2 = 318.196 rad sec , and the
modal damping ratios are ζ 1 = σ 1 ω 1 = 0.0153 and ζ 2 = σ 2 ω 2 = 0.1853 .

For the associated undamped system, the undamped natural frequencies


are ω 01 = 300 rad sec and ω 02 = 100 11 = 331.66 rad sec , and they are different
from the absolute values of eigenvalues of the system with non-proportional
damping. In this case, the difference of damped natural frequencies is
0.026 rad sec , hence 0.0041 Hz , while the undamped natural frequencies differ
with 5.039 Hz . Damping coupling approaches the natural frequencies, fact which
complicates their experimental determination.
The Nyquist plots of complex receptances α i j = x i f j , i , j = 1, 2 , are
shown in Fig. 4.39. The excitation frequency is marked along the curves. The
unusual shape of the direct receptance plots, with no distinct two loops, is produced
by the relative closeness of natural frequencies and the relatively high damping in
the second mode of vibration.

Example 4.15
Calculate the modal parameters for the system of Fig. 4.31, with the
physical mass , stiffness and damping values given in Table 4.1 for the following
four cases: Case I: lightly damped system with relatively separated natural
frequencies; Case II: lightly damped system with closely spaced natural
frequencies; Case III: highly damped system with relatively separated natural
frequencies; Case IV: highly damped system with closely spaced natural
frequencies.
Table 4.1
Case I II III IV
m1 = m 2 kg 0.0259 0.0259 0.0259 0.0259

k1 = k 3 N m 100 100 100 100

k2 N m 50 1 50 1

c1 Ns m 0.3 0.3 3 3

c2 Ns m 0.2 0.2 2 2

c3 Ns m 0.1 0.1 1 1

Draw the Nyquist plots of complex receptances for the four systems.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 177

Solution. The undamped natural frequencies are given by


k3 2 ⎜
⎛ k2 ⎞
2
ω 01= , 2
ω 02 = ω 01 1+ 2 ⎟.
m2 ⎜ k 3 ⎟⎠

The undamped modal vectors are the same in all cases

{ a }1 = ⎧⎨
1⎫
{ a } 2 = ⎧⎨ ⎫⎬ .
1

⎩1 ⎭ ⎩ −1 ⎭
The modal mass, stiffness and damping matrices, calculated with the
modal matrix built with the above vectors, are respectively
⎡ 2m 1 0 ⎤ ⎡k 1 + k 3 0 ⎤
[M ] = ⎢ 2m 2 ⎥⎦
, [K ] = ⎢ k 1 + k 3 + 4k 2 ⎥⎦
,
⎣ 0 ⎣ 0
⎡c + c c1 − c 3 ⎤
[C ] = ⎢c 1 − c 3 c 1 + c 3 + 4c 2 ⎥⎦
.
⎣ 1 3

For c 1 ≠ c 3 , the damping is non-proportional and the modal damping


matrix is not diagonal as the mass and stiffness matrices.
The eigenvalues are solutions of the fourth-order algebraic equation

( )
m 1s 2 + c 1 + c 2 s + k 1 + k 2 − c 2s − k 2
=0.
− c 2s − k 2 2
( )
m 2s + c 2 + c 3 s + k 2 + k 3

For underdamped modes of vibration, the complex eigenvalues can be


written under the form (4.179) while for overdamped modes of vibration, the real
eigenvalues can be written under the form (4.181).
The numerical values of the modal parameters are given in Table 4.2.

Table 4.2
Case I II III IV
ω 01 ω 02 62.13 87.86 62.13 62.75 62.13 87.86 62.13 62.75
ω1 ω2 62.24 87.69 62.16 62.71 64.28 84.92 62.16 62.71
ζ1 ζ2 0.062 0.132 0.055 0.192 0.540 1.409 0.547 1.919
ωd1 ωd 2 62.13 86.93 62.07 61.55 54.09 - 52.01 -
σ1 σ2 3.845 11.594 3.406 12.033 34.734 - 34.044 -
178 MECHANICAL VIBRATIONS

The undamped natural frequencies are different from the absolute value of
eigenvalues for the non-proportionally damped system, ω 0 r ≠ ω r . In the case of
damping coupling, the absolute values of eigenfrequencies, sometimes referred to
as resonant frequencies, are closer to each other than the corresponding undamped
natural frequencies, while the damped natural frequencies might have reversed
order. For the first mode, ω 1 > ω 01 , while for the second mode, ω 2 < ω 02 , and for
Case II, ωd 2 < ωd 1 .
The second mode is overdamped in Cases III and IV, having damping
ratios ζ r > 1 . Indeed, for the set of physical parameters corresponding to systems
with relatively spaced undamped natural frequencies, the second mode of vibration
becomes overdamped for a value c 3 ≅ 0.749 Ns m , while the first mode of
vibration becomes overdamped for values larger than c 3 ≅ 1.89 Ns m .

Fig. 4.40

Polar diagrams of the complex receptances are presented in Fig. 4.40. The
two natural frequencies are located on the diagrams. Except for Case I, the polar
plots do not exhibit two loops, one for each mode of vibration, so that the number
of degrees of freedom is not apparent from the simple inspection of the frequency
response plots. Resonance location in such cases requires adequate methods.
4. TWO-DEGREE-OF-FREEDOM SYSTEMS 179

Exercises

4.E1 For the system shown in Fig. 4.1 let k 1 = 2 k , k 2 = k , k 3 = 2k ,


m 1 = 3 m , m 2 = m . Set up the equations of motion and determine the natural
modes of vibration.
Answer: ω 1 = 0.919 k m , ω 2 = 1.776 k m , μ 1 = 0.464 , μ 2 = −6.464 .

4.E2 Using the following values: k 1 = 103 N m , k 2 = 500 N m ,


k 3 = 2 ⋅ 103 N m , m 1 = 5 kg , m 2 = 10 kg , f̂ 1 = 100 N , f̂ 2 = 0 , ω = 15 rad sec ,
determine the amplitudes of the forced vibrations of the two masses for the system
of Fig. 4.4.
Answer: X 1 = 0.16 m , X 2 = −0.32 m .

4.E3 Determine the amplitudes of the forced vibrations of the system from
Fig. 4.4, if k = 103 N m , m = 5 kg , f̂ 1 = 50 N , f̂ 2 = 100 N , ω = 10 rad sec .

Answer: X 1 = −0.5 m , X 2 = −0.8 m .

4.E4 Find the natural frequencies and associated mode shapes for the
torsional vibrations of the system shown in Fig. 4.6. The stiffnesses of the shafts
are K 1 = K , K 2 = 2 K , K 3 = 3 K , and the moments of inertia of the flywheels are
J1 = 2J , J 2 = J .

Answer: ω 1 = K J , ω 2 = 2.345 K J , μ 1 = 0.5 , μ 2 = −4 .

4.E5 Using the following values: K 1 = 107 N m rad ,


K 2 = 2 ⋅107 N m rad , K 3 = 3 ⋅ 107 N m rad , J 1 = 100 kg m 2 , J 2 = 200 kg m 2 ,
ω = 300 rad sec , determine the amplitudes of the torsional vibrations of the system
shown in Fig. 4.6 when a harmonic torque of amplitude M 0 = 105 N and
frequency ω = 300 rad sec is applied on the second disc.

Answer: Θ1 = 0.00735 rad , Θ 2 = 0.00772 rad .


180 MECHANICAL VIBRATIONS

4.E6 Calculate the natural frequencies and the mode shapes for the
flexural vibrations of the system shown in Fig. 4.15, where m 1 = 3 m , m 2 = m 3 ,
l 1 = l 2 = l and E I = const .

Answer: ω 1 = 0.522 E I m l 3 , ω 2 =1.73 E I m l3 , μ1 = 6 ,


μ 2 = −1.5 .

4.E7 Determine the normal modes for flexural vibrations of the system
shown in Fig. 4.20, where m 1 = 2 m , m 2 = m , l 1 = l 2 = l , l 3 = l 2 and f = 0 .

Answer: ω 1 = 1.477 E I m l 3 , ω 2 = 3.464 E I m l 3 , μ 1 = −1 , μ 2 = 2 .

4.E8 Find the natural frequencies and associated mode shapes for the
flexural vibrations of the massless beam shown in Fig. 4.E8.

Fig. 4.E8

Answer: ω 1 = 0.891 E I m l 3 , ω 2 = 3.688 E I m l 3 , μ 1 = 1.048 ,


μ 2 = −0.477 .

4.E9 Determine the amplitudes of the forced vibrations of the two masses
for the system of Fig. 4.E9 using the following values: E I m l 3 = 800 sec −2 ,
F0 l 3 E I = 2 mm and ω = 10 rad sec .

Fig. 4.E9

Answer: Y1 = 117.5 mm , Y2 = 25.04 mm .


4. TWO-DEGREE-OF-FREEDOM SYSTEMS 181

4.E10 In Fig. 4.E10 use coordinate x at the centre of mass and θ for the
rotation of the bar and set up the equations of motion. Determine the natural
frequencies and locate the node of the bar for each mode.

Fig. 4.E10

Answer: ω 1 = 0.707 k m , ω 2 = 1.414 k m , 1 μ 1 = 0.666 l ,


1 μ 2 = −0.333 l .

4.E11 Find the natural frequencies and draw the mode shapes for the
coupled translation and rotation vibrations of the rigid bar of Fig. 4.E11.

Fig. 4.E11

Answer: ω 1 = 0.78 k m , ω 2 = 1.281 k m , 1 μ 1 = 0.64 l , 1 μ 2 = −0.39 l .

4.E12 Calculate the amplitudes of the forced vibrations for the rigid bar of
Fig. 4.E12 using the following values: k = 103 N m , m = 5 kg , F0 = 50 N ,
l = 0.4 m and ω = 10 rad sec .

Fig. 4.E12
Answer: a x = 0.0165 m , aϕ = 0.083 rad .
182 MECHANICAL VIBRATIONS

4.E13 The results are given below of a resonance test on a structure. The
response was measured at a place some distance away from the point of application
of the excitation force and is given as values of the transfer receptance. Estimate
for the two modes of vibration recorded the modal mass, stiffness, damping ratio
and viscous damping coefficient.

Receptance Phase angle between


Frequency, magnitude, response and force,
Hz 10 −3 mm N degrees
40 10 10
41 10.9 30
41.5 10.9 50
41.8 10.4 70
42 9.5 90
42.5 8.4 110
43.5 7.1 130
44 4.9 150
50 3.0 152
60 2.8 160
68 2.0 178
70 6.4 190
72 6.9 210
73 7.2 230
73.5 6.6 250
74 5.5 270
74.5 4.3 290
75 3.0 310
77 1.8 330

Answer: ω 1 = 263 rad sec , ζ 1 = 0.0466 , K 1 = 700 N m ,


C 1 = 0.248 Ns m , M 1 = 1.015 ⋅ 10 −2 kg , ω 2 = 462 rad sec , ζ 2 = 0.0339 ,
K 2 = 1800 N m , C 2 = 0.264 Ns m , M 2 = 0.84 ⋅ 10 −2 kg .
5.
SEVERAL DEGREES OF FREEDOM

In the previous chapter, systems with two degrees of freedom have been
considered as the simplest case of, and an introduction to the general case of multi-
degree-of-freedom systems. A system has n degrees of freedom if its configuration
at any time can be represented by n independent coordinates. Usually, in the
configuration space, the coordinates are rectilinear translations and rotations, but
velocities and accelerations may be used as well.
Systems with a finite number of degrees of freedom are called discrete
systems. It is engineering practice to describe the vibration of a continuous system
with a finite number of coordinates. Each element of a discrete system is itself a
continuous system, but the lowest frequencies of the latter are much greater than
those of the idealized discrete system.
The simplest approach leads to lumped parameter systems, consisting of
lumped masses or discs, springs and dashpots. Their dynamic properties are
defined by scalar quantities. Elements can be described by stiffness and damping
matrices, relating the end forces to displacements and velocities across the element.
Another discretization technique is the finite element method which can be
regarded as a Rayleigh-Ritz method. It consists of approximating the solution of a
differential eigenvalue problem, having no known closed-form solution, by a finite
series of shape functions multiplied by undetermined coefficients. In the finite
element method, the shape functions are local low-degree polynomials and the
coefficients are the nodal displacements determined to render the Rayleigh quotient
for the system stationary. Elemental matrices are defined for each element type
which can be assembled in global mass, stiffness and damping matrices.
Having these global matrices, one can write the equations of motion.
Seeking synchronous harmonic solutions these are transformed into a
homogeneous set of algebraic equations, equivalent to an algebraic eigenvalue
problem. Solution of the eigenvalue problem yields the natural frequencies and
associated mode shapes. Alternatively, natural frequencies can be obtained by
minimization of the Rayleigh quotient.
184 MECHANICAL VIBRATIONS

The dynamic response of a discrete system can be described by


simultaneous ordinary differential equations. For a proper choice of coordinates,
known as principal or modal coordinates, the equations can be decoupled and
solved independently. The modal coordinates represent linear combinations of the
actual displacements. Conversely, the motion can be regarded as a superposition of
vibrations in the natural modes of vibration defined by the modal coordinates. The
motion in a natural mode of vibration is synchronous and harmonic at all system
coordinates.
Systems having a finite number of degrees of freedom vibrate
simultaneously in each of the several natural modes, as many natural modes as
there are degrees of freedom. It takes certain combinations of initial conditions or
applied forces to cause them to vibrate in only one mode. However, in damped
systems, the free vibration is dominated by a few lower modes, sometimes only the
lowest one, while the forced vibration can be described by a summation truncated
to the modes with resonances within the frequency range of interest, plus some
residual terms due to modes with frequencies lower or higher than the operating
range.
There is no basic difference between the low order discrete systems treated
in this chapter and the large order discrete systems treated in a next chapter except
that for the latter the solution of the eigenvalue problem requires more efficient
computational methods.

5.1 Lumped Mass Systems


Systems comprising uni-dimensional members like bars, shafts or beams
are modeled by elementary systems consisting of concentrated masses connected
by massless elastic elements. The distributed mass of such members is lumped at
arbitrary chosen discrete points, regardless of the variation of vibratory amplitude
along the element.

5.1.1 Beams with Lumped Masses

Beams with lumped masses can have either point masses with only linear
transverse displacements, or rigid discs with both translational and rotational
degrees of freedom.

5.1.1.1 Linear Displacements


Figure 5.1 shows two commonly used ways in which a uniform beam
element could be represented. Duncan’s model (Fig. 5.1, b) has all the mass
concentrated at the centre of gravity. Rayleigh’s model (Fig. 5.1, c) has half the
5. SEVERAL DEGREES OF FREEDOM 185

mass of the segment at each end. In the figure m 0 = ρ A is the mass per unit
length, where ρ is the mass density and A is the cross section area.

Fig. 5.1

Comparing the two models it can be noticed that in Duncan’s model the
element inertia to rotation with respect to the midpoint is neglected, while in
2
m0 l ⎛ l ⎞
Rayleigh’s model it is 2 ⎜ ⎟ ≠ 0 . This is why lumping using Duncan’s
2 ⎝2⎠
model generally gives higher natural frequencies, while Rayleigh’s model gives
lower natural frequencies. The difference in frequencies calculated with the two
models lowers with increasing the number of elements.
Rayleigh’s model has also advantages for stepped beams, when the
bending rigidity E I along the beam changes at the element ends, so that the model
with lumped masses has segments with constant cross section.
The discrepancy in the values of natural frequencies calculated using
Rayleigh’s model can be illustrated by two examples.
When the simply supported beam of Fig. 5.2, a is split into two segments
(Fig. 5.2, b), the ratio of the natural frequency to the true value (6.14) for the beam
with distributed mass is ω 1 ω 10 = 0.995. When the beam is split into four
segments (Fig. 5.2, c), the resulting three-degree-of-freedom system has natural
frequencies of ω 1 ω 10 = 0.98 , ω 2 ω 20 = 0.995 and ω 3 ω 30 = 0.995 .

For the cantilever beam of Fig. 5.3, a the approximation is lower. For the
one segment model (Fig. 5.3, b), the ratio of the natural frequency to the true value
(6.16) is ω 1 ω 10 = 0.7. When the beam is split into two segments (Fig. 5.3, c), the
186 MECHANICAL VIBRATIONS

ratio of the first natural frequencies is ω 1 ω 10 = 0.9. When the beam is split into
three segments (Fig. 5.2, d), the ratio of the first natural frequencies is
ω 1 ω 10 = 0.95.

Fig. 5.2 Fig. 5.3

5.1.1.2 Linear Displacements and Rotations


Figure 5.1, d shows an extension of Duncan’s model with the mass and
mass moment of inertia of a beam element concentrated at the middle.
For a uniform beam element, the total mass is
m = ρ Al = m0 l , (5.1)
and the total mass moment of inertia is
m0 l 3 I ρ A l3
J= + m0 l = + ρlI (5.2)
12 A 12
where I is the cross-section second moment of area.
For a cylindrical beam element of diameter d and length l

m0 l ⎛⎜ 3 d 2 2 ⎞⎟
J= +l . (5.3)
12 ⎜⎝ 4 ⎟

The total mass moment of inertia is composed of two parts. The first part,
2
ml 12 , is due to the element mass being distributed along the length of the beam
5. SEVERAL DEGREES OF FREEDOM 187

at the level of the neutral axis. The second part, the “rotatory inertia” ρ l I , is due
to the fact that the beam mass is also distributed away from the beam neutral axis.
This part is not activated by translation; it is only activated by rotation.
In the extension of Rayleigh’s model (Fig. 5.1, e), one-half of the mass and
a negative mass moment of inertia are lumped to the left and right ends. With this
distribution, the conservation of both the mass and total mass moment of inertia
with respect to the midpoint is satisfied. This can be checked calculating the latter
using the Huygens-Steiner theorem of parallel axes
⎛ m 0 l l 2 m 0 l3 ⎞ m 0 l3
J =2⎜ − ⎟= .
⎜ 2 4 12 ⎟ 12
⎝ ⎠
The above distinction is important in the finite element analysis. Two
diagonal lumped-mass matrices are commonly used: an element mass matrix for
translational inertia
⎡1 2 0 0 0 ⎤
⎢ 0 − l 2 12 0 0 ⎥⎥
[ ]
mt = m0 l
e ⎢
⎢0 0 12 0 ⎥
(5.4)
⎢ ⎥
⎣0 0 0 − l 2 12⎦

and an element mass matrix for rotatory inertia


⎡0 0 0 0⎤
⎢0 i 2 0 ⎥⎥
[ ]
mr = m0 l ⎢
e
⎢0 0
0
0 0⎥
, (5.5)
⎢ ⎥
⎣0 0 0 i2 ⎦

where i = I A .

5.1.1.3 Flexibility Coefficients


As shown in Section 4.3, the equations of motion for flexural systems with
lumped masses are written easier using flexibility coefficients instead of
stiffnesses.
If the linear displacements of the lumped masses are taken as the
coordinates that define the system motion, then the vector of mass displacements
{ y } is related to the vector of forces applied to the lumped masses { f } by the
flexibility matrix [δ ] as in equation (4.78)
{y } = [ δ ]{ f }. (5.6)
188 MECHANICAL VIBRATIONS

The equation of motion for free vibrations (4.82) can be written

[δ ] [ m ] { &y& } + { y } = { 0 }. (5.7)
The corresponding eigenvalue problem (4.85) is


⎜ [δ ][ m ] − 1
2
[I ]⎞⎟ { a } = {0 } . (5.8)
⎝ ω ⎠
The condition to have non-trivial solutions yields the frequency equation


det ⎜ [ δ ][ m ] − 1
[ I ]⎞⎟ = 0 (5.9)
⎝ ω2 ⎠
whose solutions are the system natural frequencies ωr .

The mode shapes are defined by the modal vectors {a }r , satisfying the
homogeneous linear set of equations
⎛ 1 ⎞
⎜ [δ ][ m ] − [I ]⎟⎟ { a }r = {0} . (5.10)
⎜ ωr 2
⎝ ⎠

Example 5.1
Calculate the natural frequencies of lateral vibration for the three-mass
beam of Fig. 5.4, a, where E I = const .
Solution. Referring to Fig. 5.4, b, the deflection at any point x due to a
concentrated load F applied a distance b from the right end, can be determined
from the equation

v(x)=
Fbx 2
6E I l
(
l − x2 − b2 . )
The flexibility coefficients are

δi j =
b j xi
6E I l
(l 2
− x 2i − b 2j . )
Because x 1 = b 3 = l 4 , x 2 = b 2 = l 2 , x 3 = b 1 = 3l 4 , the flexibility
matrix is
⎡ 9 11 7 ⎤
l3 ⎢
[δ ] = 11 16 11⎥⎥ . (5.11)
768 E I ⎢
⎢⎣ 7 11 9 ⎥⎦
5. SEVERAL DEGREES OF FREEDOM 189

Equation (5.8) can be written


⎡ 9 11 7 ⎤ ⎧ a1 ⎫ ⎧a1 ⎫
⎢11 16 11⎥ ⎪a ⎪ = λ ⎪ ⎪
⎢ ⎥ ⎨ 2⎬ ⎨a 2 ⎬
⎢⎣ 7 11 9 ⎥⎦ ⎪⎩ a3 ⎪⎭ ⎪a ⎪
⎩ 3⎭
where
λ = 768 E I ml3ω 2 .

Fig. 5.4 Fig. 5.5

The frequency equation is

λ 3 − 34λ 2 + 78 λ − 28 = 0
with roots

λ 1 = 31.5563 , λ2 = 2, λ 3 = 0.4436 .
The natural frequencies are

ω 1 = 4.933 E I ml 3 , ω 2 = 19.596 E I ml 3 , ω 3 = 41.606 E I ml3 .


The modal vectors are

⎧ 1 ⎫ ⎧1⎫ ⎧ 1 ⎫
{a }1 = ⎨1.4142⎬ , {a } 2 = ⎨ 0 ⎬ , {a } 3 = ⎪⎨− 1.4142⎪⎬ .
⎪ ⎪ ⎪ ⎪
⎪ 1 ⎪ ⎪− 1⎪ ⎪ 1 ⎪
⎩ ⎭ ⎩ ⎭ ⎩ ⎭
190 MECHANICAL VIBRATIONS

5.1.1.4 Dunkerley’s Formula


For many vibrating systems, the natural frequencies of the second and
higher modes are often considerably greater than that of the fundamental mode.
This fact allows the estimation of the fundamental frequency with simple formulae.
If

α 0 λ n + α 1 λ n −1 + .... + α n −1 λ + α n = 0
is an algebraic equation, then the sum of its roots is
n α1
∑λi = −α .
i =1 0

Considering the frequency equation (5.9), we can write


n 1 α1 n n 1
∑ =− = ∑ δ ii mi = ∑ 2 ,
α 0 i =1
i =1 ω 2i i =1 ω i i

where ω i2i is the square of the “isolated” natural frequency of the system
containing only the mass m i . This frequency is calculated based on the exact
deflected curve of the isolated one-mass system, having the same configuration as
the analysed system except for the eliminated masses.
Because ω 1 < ω 2 < . . . < ω n , all terms in the first sum except the first may
be omitted for the approximate determination of the fundamental frequency
1 n 1 n 1
≅∑ =∑
ω 21 2
i =1 ω i i =1 ω 2i i

or
1 1 1 1
≅ + + + . .. (5.12)
ω 21 ω 211 ω 222 ω 233

Thus the reciprocal of the fundamental frequency squared can be obtained


by adding the squares of the reciprocals of the isolated frequencies. Equation (5.12)
is known as Dunkerley’s formula. It was discovered experimentally and published
by S. Dunkerley (1895) then justified theoretically by R. V. Southwell (1921).
It enables the estimation of the fundamental frequency of a system without
solving the associated eigenvalue problem. The formula is valid for grounded
systems only, i.e. it cannot be applied to free-free systems. Generally it yields
lower estimates of the fundamental frequency.
5. SEVERAL DEGREES OF FREEDOM 191

For the cantilever beam of Fig. 5.3, a the true value of the fundamental
frequency (6.16) is ω 10 = 3.52 E I ml 3 . For the model with one segment (Fig.

5.3, b), the lowest natural frequency is ω 1 = 2.44 E I ml 3 , which is 30.7%


lower than the true value. When the beam is split into two segments (Fig. 5.3, c),
the first natural frequency is ω 1 = 3.098 E I ml 3 , being 12% lower than the true
one. When the beam is split into three segments (Fig. 5.2, d), the first natural
frequency is ω 1 = 3.286 E I ml 3 , which is 7.12% lower, and when it is split
into four segments, it is 4.5% lower than the true frequency.

Example 5.2
Calculate the fundamental frequency of lateral vibration for the three-mass
beam of Example 5.1 (Fig. 5.5, a) using Dunkerley’s formula.
Solution. From the flexibility matrix (5.11) we obtain
9l 3 16l 3
δ 11 = δ 33 = , δ 22 = .
768 E I 768 E I
For the single-degree-of-freedom systems with isolated masses (Fig. 5.5, b,
c, d), the squares of the natural frequencies are respectively
2 1 768 E I
ω 11 = = ,
m δ 11 9m l 3

1 768 E I
ω 222 = = ,
m δ 22 16ml 3

1 768 E I
ω 323 = = ,
m δ 33 9m l 3

so that Dunkerley’s formula (5.12) yields


1 1 1 1 9ml 3 16ml 3 9m l 3 34ml 3
≅ + + = + + = .
ω 21 ω 211 ω 222 ω 233 768 E I 768 E I 768 E I 768 E I

The estimated fundamental natural frequency is

ω 1 ≅ 4.7527 E I ml 3 ,

which is 3.6% lower than the true value calculated at Example 5.1.
192 MECHANICAL VIBRATIONS

5.1.1.5 Rayleigh’s Formula


When a beam is represented by a lumped-mass model, consisting of a
series of lumped masses m i (i = 1,..., n ) attached to a massless beam at points of
abscissae x i , Rayleigh’s formula (2.19) becomes

∫ ( ) dx 2
E I ∂2v ∂ x2
ω12 = ,n
(5.13)
∑m v i =1
2
i i

( )
where v i = v x i are static deflections at the mass locations.

If the strain energy is determined from the work done by the corresponding
1 n
lumped weights m i g , then U max = ∑ m i g ⋅ v i so that Rayleigh’s formula (5.13)
2 i =1
becomes
n
g ∑ m i vi
i =1
ω12 = n
. (5.14)
∑ m i vi2
i =1

where g is the acceleration of gravity.


As mentioned in Section 2.1.5, if the true deflection curve of the vibrating
system is assumed, the fundamental frequency found by Rayleigh’s formula will be
the correct frequency. For any other curve, the frequency determined by this
method will be higher than the true frequency.
If further accuracy is desired, a better approximation to the dynamic
deflection curve can be made by using dynamic loads instead of the static weights.
Since the dynamic load is m i ω 2 vi , which is proportional to the deflection, we can
v2 v
recalculate the deflection with the modified weights m 1 g , m 2 g , m3 g 3 .
v1 v1

Example 5.3
Calculate the fundamental frequency of lateral vibration for the three-mass
beam of Example 5.1 (Fig. 5.4, a) using Rayleigh’s formula (5.14).
Solution. Using flexibility coefficients and applying superposition, the
displacement of any mass can be calculated as the sum of the products of flexibility
coefficients at the respective location multiplied by the corresponding weights.
5. SEVERAL DEGREES OF FREEDOM 193

Equation (5.6) can be written


⎧v1 ⎫ ⎧m g ⎫ ⎧1⎫
⎪ ⎪ ⎪ ⎪ ⎪⎪
⎨v 2 ⎬ = [δ ] ⎨m g ⎬ = m g [δ ] ⎨1⎬
⎪v ⎪ ⎪m g ⎪ ⎪1⎪
⎩ 3⎭ ⎩ ⎭ ⎩⎭
wherefrom we obtain
27l 3
( )
v 1 = m g δ 11 + δ 1 2 + δ 13 =
768 E I
mg ,

38l 3
(
v 2 = m g δ 2 1 + δ 2 2 + δ 23 = ) 768 E I
mg ,

v 3 = v1.

Substitution in (5.14) yields


768EI 27 + 38 + 27 768 EI 92 768 EI
ω12 = 3 2 2 2
= 3 2902
=
ml 27 + 38 + 27 9m l 31.54348 ml 3
or
ω 1 = 4.9343 E I ml3 ,
which is only 0.02% higher than the true value calculated at Example 5.1.
If m = m0 l / 4 , then the system from Fig. 5.4, a is the same as the beam
from Fig. 5.2, c. Substituting this value in the above formula yields
EI 9.8686 E I
ω 1 = 4.9343 ⋅ 2 4
= ,
m0l l2 m0

which is lower than the true value obtained for a continuous beam (6.14)
π2 E I 9.8696 E I
ω true = 2
= ,
l m0 l2 m0

owing to the lumping process in which the total mass was not conserved.

5.1.2 Multi-Story Frames as Shear Buildings

A shear building is a structure in which there is no rotation of a horizontal


section at the level of the floors. The deflected building resembles a cantilever
beam that is deflected by shear forces only, hence the name shear building.
194 MECHANICAL VIBRATIONS

Figure 5.6, a shows a three-story shear building modelled as a frame with


three degrees of freedom. Only swaying of the frame in its plane is considered, due
to the flexural deformation of the vertical members in the plane of the frame.
To accomplish such deflection in a building, the following assumptions are
made: a) all the floors are rigid and capable of motion only in the horizontal
direction, so that joints between floor girders and stanchions are fixed against
rotation; b) all columns are inextensible and massless; and c) the mass of the
building is concentrated at the floor levels, so that the vibration of the multi-story
building is reduced to the vibration of a system with finite degrees of freedom.
It is convenient to represent the shear building solely in terms of a single
bay. The displacement of a horizontal member is resisted by elastic restoring forces
in the stanchions. If the stanchions are treated as beams of uniform cross-section,
standard methods of analysing beams with built-in ends, when one end sinks
relative to the other, show that the combined stiffness of the stanchions in flexure is

12 E ⋅ 2 I 24 E I
k= = 3 , (5.15)
l3 l
where E is the material Young’s modulus, I - the cross-sectional second moment
of area of one stanchion and l - the height of the story (length of the stanchion).

Fig. 5.6

The equations of motion can be easily obtained, using d’Alembert’s


principle, from the free body diagrams of Fig. 5.6, b by equating to zero the sum of
the forces acting on each mass. The procedure is explained for two-degree-of-
freedom systems in Section 4.1.1.
5. SEVERAL DEGREES OF FREEDOM 195

5.1.3 Torsional Systems

Torsional vibrations of multicylinder reciprocating machines are studied


with a simplified lumped mass torsional system (Fig. 5.7) obtained by lumping the
inertias of rotating and reciprocating parts from each cylinder at discrete points on
a main equivalent shaft. The parts mounted on the shafting, like the flywheel, the
torsional damper, the coupling and the driven rotor or the propeller can also be
modelled as rigid discs.

Fig. 5.7

The connecting rod is replaced by an equivalent dynamical system


comprised of two concentrated masses, one at the crank end, the other at the piston
end.
In each cylinder there is a reciprocating mass, mrec , consisting of the
piston mass and the reduced mass of the connecting rod, and a rotating mass,
consisting of the crank throw, counterweight and the reduced mass of the
connecting rod, having an inertia J rot . The average inertia per cylinder is

1
J = J rot + mrec r 2 ,
2
where r is the crank throw radius.
The torsional stiffness of crankshaft between cylinders, arising from the
combined effect of crank pin twisting and crank web bending, can be either
approximated numerically or measured experimentally. The same for the stiffness
of the crankshaft between the last cylinder and the flywheel, for the coupling and
the driven shaft. The whole system can then be reduced to a torsional model with a
series of rigid discs connected by massless flexible shafts, as in Fig. 5.7.

Example 5.4
A diesel-motor-generator set has been reduced to the equivalent system
shown in Fig. 5.8, with the J values in kg ⋅ m 2 , and with K values in
196 MECHANICAL VIBRATIONS

106 x Nm/rad : J 1 = 11.863 , J 2 = 2.011 , J 3 = J 4 = J 5 = J 6 = 0.167 ,


J 7 = 0.897 , K 1 = 1.062 , K 2 = 6.101 , K 3 = K 4 = K 5 = 3.05 , K 6 = 4.067 .
Estimate the fundamental- and second-mode natural frequencies of torsional
vibrations. Plot the corresponding mode shapes.
Solution. ω 1 = 552 rad/sec , ω 2 = 1145 rad/sec . The mode shapes are
shown in Fig. 5.8.

Fig. 5.8
Geared systems and car engine-driving train systems are modelled
likewise. Geared-branched systems can be modelled as in Section 4.2.4. It is
convenient to use an equivalent system which has no gears (or has gears that effect
no change of speed) and whose natural frequencies are the same as the frequencies
of the original system. This is possible because, while the gears effect a change of
rotary speed given by the transmission ratio, the vibration is transmitted through
the gears without change in frequency, but with proper change of amplitude.
5. SEVERAL DEGREES OF FREEDOM 197

In Section 4.2.4, the actual system with branches of different rotary speeds
was reduced to an equivalent system with all parts having the same speed, usually
equal to that of the reference shaft. Progressing away from that shaft to other
branch through a gear of transmission ratio i, the actual J’s and K’s occurring
beyond are multiplied by i 2 . If several gears are passed, the subsequent J’s and K’s
are multiplied by the squares of the respective gear ratios. For the mating gears, the
total inertia is obtained by adding to the moment of inertia of the reference gear the
moment(s) of inertia of the (two) reduced gear(s) multiplied by i 2 .
The equivalent model so obtained has the same natural frequencies as the
actual system, as well as the same position of nodes. However, in the reduced
branches, the amplitude of angular displacements is divided by − i , while the
torque in the branch is multiplied by − i .
A more straightforward approach is given in the following. In the
equivalent model, each shaft and disc has the actual speed, only the driven gear is
condensed out.

Fig. 5.9

Consider part of a geared-branched system (Fig. 5.9, a) where (the driving)


shaft 1 and (the driven) shaft 2 are connected by two gears with inertias J , J ' and
of pitch circle radii r and r' , respectively. Apart from shaft stiffnesses K 1 and
K 2 , and disc mass moments of inertia J 1 and J 2 , the angular displacements and
torques acting on each disc are shown in the figure.
The system motion is described by four torques and four angular
displacements. However, only three of each of them are independent, due to the
compatibility conditions between the mating gears. Thus it would be useful to
198 MECHANICAL VIBRATIONS

consider the torque M ' and the angular displacement θ ' as dependent variables
and eliminate them, replacing the given system by the equivalent model shown in
Fig. 5.9, b. Shaft 1 is taken as the reference shaft.
Compatibility of the angular displacements requires that
r θ = − r' θ ' ,
while compatibility of torques requires
M M′
=− or M θ =M 'θ' .
r r′
The transmissibility ratio is defined as
n2 r θ' M
i=− = =− =− . (5.16)
n1 r′ θ M′

In the original system, the stiffness matrix of shaft 2 is defined by


⎧ M ′ ⎫ ⎡ K2 − K 2 ⎤ ⎧θ ' ⎫
⎨ ⎬=⎢ ⎨ ⎬.
⎩M 2 ⎭ ⎣− K 2 K 2 ⎥⎦ ⎩θ 2 ⎭

Using the transformation equations, based on (5.16),


⎧ M ⎫ ⎡ − i 0⎤ ⎧ M ′ ⎫ ⎧θ ' ⎫ ⎡− i 0⎤ ⎧ θ ⎫
⎨ ⎬=⎢ ⎥ ⎨ ⎬, ⎨ ⎬=⎢ ⎥ ⎨ ⎬,
⎩ M 2 ⎭ ⎣ 0 1⎦ ⎩M 2 ⎭ ⎩θ 2 ⎭ ⎣ 0 1⎦ ⎩θ 2 ⎭
the stiffness matrix of shaft 2 in the equivalent model is defined as follows
⎧ M ⎫ ⎡ − i 0⎤ ⎡ K 2 − K 2 ⎤ ⎡ − i 0⎤ ⎧ θ ⎫
⎨ ⎬=⎢ ⎥⎢ ⎨ ⎬,
⎩M 2 ⎭ ⎣ 0 1⎦ ⎣− K 2 K 2 ⎥⎦ ⎢⎣ 0 1⎥⎦ ⎩θ 2 ⎭

⎧ M ⎫ ⎡i 2 K 2 i K2 ⎤ ⎧θ ⎫
⎨ ⎬=⎢ ⎥ ⎨θ ⎬ .
⎩M 2 ⎭ ⎣⎢ i K 2 K 2 ⎦⎥ ⎩ 2⎭
The mass matrix is defined likewise

⎧ M ⎫ ⎡− i 0⎤ ⎡ J ′ 0 ⎤ ⎡− i 0⎤ ⎧ θ&& ⎫
⎨ ⎬=⎢ ⎥⎢ ⎥⎢ ⎥ ⎨ ⎬,
⎩M 2 ⎭ ⎣ 0 1⎦ ⎣ 0 J 2 ⎦ ⎣ 0 1⎦ ⎩θ&&2 ⎭

⎧ M ⎫ ⎡i 2 J ′ 0 ⎤ ⎧⎪ θ&& ⎫⎪
⎨ ⎬=⎢ ⎥ ⎨ && ⎬ .
⎩M 2 ⎭ ⎣⎢ 0 J 2 ⎦⎥ ⎪⎩θ 2 ⎪⎭

In the following, two rotating discs coupled by an elastic shaft segment are
considered as a finite element. For an element of the driven shaft, the element
matrices are
5. SEVERAL DEGREES OF FREEDOM 199

[ k ] = ⎡⎢ii KK i Ke ⎤
[ m ] = ⎡⎢i 0J ′ 0⎤
2 2
e e
⎥ , ⎥ . (5.17)
e e

⎣⎢ e K e ⎦⎥ ⎣⎢ J e ⎦⎥

The element matrices for the shaft 1 in Fig.5.9 are

[ k ]= ⎡⎢−KK
1 1 − K1 ⎤
K1 ⎥⎦
, [ m ] = ⎡⎢ J0
1 1 0⎤
J ⎥⎦
.
⎣ 1 ⎣
They can be obtained from the general expressions (5.17) substituting
i = −1 . In general, for elements not adjacent to the gears, one of the end inertias is
zero.
Another approach is to condense the system matrices based on the
constraint equations expressing the compatibility of the angular displacements of
mating gears.

Example 5.5
The geared-branched system shown in Fig. 5.10, a consists of three rigid
discs having mass moments of inertia J 1 , J 5 and J 6 and three rigid gears of radii
r , r / 2 and r / 3 having mass moments of inertia J 2 , J 3 and J 4 which are
connected by three light shafts having torsional stiffnesses K 1 , K 2 and K 3 .
Derive the equations of free vibration.

Solution. In the upper branch, the speed ratio is i = −2 . In the lower


branch, the speed ratio is i = −3 . For the equivalent system shown in Fig. 5.10, b
the element matrices are calculated from equations (5.17).
The global matrices are assembled using the direct stiffness approach,
obtaining:
⎡ K1 − K1 0 0 ⎤
⎢− K K 1 + 4K 2 + 9K3 2K 2 3K 3 ⎥⎥
[K ] = ⎢⎢ 0 1 2K 2 K2 0 ⎥
⎢ ⎥
⎣⎢ 0 3K 3 0 K 3 ⎦⎥

and
⎡J 1 0 0 0⎤
⎢0 J 2 + 4J3 + 9J 4 0 0 ⎥⎥
[M ] = ⎢⎢ 0 0 J5 0⎥
.
⎢ ⎥
⎣⎢ 0 0 0 J 6 ⎦⎥
200 MECHANICAL VIBRATIONS

The global vector of angular displacements is

{Θ } = {θ 1 θ 2 θ 3 θ 4 }T .
The equations of motion can be written
[ M ]{Θ&& }+ [ K ]{Θ } = { 0 } .
They are used for the estimation of natural frequencies.

Fig. 5.10

The arrangement in Fig. 5.10 is a simplified model of a typical marine


propulsion drive, with the propeller J1 driven by a low-pressure turbine J 5 and a
high pressure turbine J 6 .
The calculated natural frequencies are compared to the excitation
frequencies. Usually the main excitation factor is the variation of the torque on
propeller due to the variation of the water forces on the blades during their rotation.
The frequency of this variation is called the blade frequency. It is the propeller
frequency (once per shaft revolution) multiplied by the number of blades.

Example 5.6
The geared-branched system shown in Fig. 5.11 has the following
parameters: J 1 = 950 , J 2 = 542 , J 3 = 406.7 , J 4 = J 5 = 6.78 , J 7 = 13.55 ,
J 6 = J 8 = J 9 = 27.12 [ kg ⋅ m ],
2
i = 76 24 , K 1 = 20.336 ⋅ 106 , K 2 = 8.135 ⋅ 106 ,
5. SEVERAL DEGREES OF FREEDOM 201

K 3 = 1.22 ⋅ 106 , K 4 = 0.407 ⋅ 106 , K 5 = 1.63 ⋅ 106 , K 6 = 2.44 ⋅ 106 [Nm/rad ] .


Calculate the natural frequencies of torsional vibrations.

Fig. 5.11

Solution. The actual system is reduced to an equivalent model in which the


two pinions are condensed out. The element matrices (5.17) are calculated for each
segment, then assembled into the global stiffness and mass matrices.
The global stiffness matrix is
⎡ K1 − K1 0 0 0 0 0 ⎤
⎢− K K1 + K 2 −K2 0 0 0 0 ⎥⎥
⎢ 1
⎢ 0 −K2 2
K2 +i K3 +i K5 2
iK 3 0 iK 5 0 ⎥
⎢ ⎥
[K ] = ⎢ 0 0 iK 3 K3 +K4 −K4 0 0 ⎥
⎢ 0 0 0 −K4 K4 0 0 ⎥
⎢ ⎥
⎢ 0 0 iK 5 0 0 K5 +K6 − K 6⎥
⎢ 0 0 0 0 0 −K6 K 6 ⎥⎦

The global mass matrix is
⎡J 1 0 0 0 0 0 0⎤
⎢0 J2 0 0 0 0 0 ⎥⎥

⎢0 0 2
J3 +i J 4 +i J5 2
0 0 0 0⎥
⎢ ⎥
[M ] = ⎢ 0 0 0 J6 0 0 0⎥
⎢0 0 0 0 J7 0 0⎥
⎢ ⎥
⎢0 0 0 0 0 J8 0⎥
⎢0 0 0 0 0 0 J 9 ⎥⎦

202 MECHANICAL VIBRATIONS

The natural frequencies have the following values: f 1 = 0 , f 2 = 14.26 ,


f 3 = 22.98 , f 4 = 35.34 , f 5 = 41.98 , f 6 = 50.52 , f 7 = 75.17 Hz .

5.1.4 Repeated Structures

Vibratory systems are frequently comprised by a number of identical


sections which are repeated several times. Examples of repeated structures are
given in Fig. 5.12 for an n-story shear building, an n-mass system with translatory
motion and an n-disc torsional system. For such systems the method of difference
equation is appropriate for the calculation of natural frequencies.

Fig. 5.12

The equation of motion for the rth mass (Fig. 5.12, a) is

m &x&r + k (xr − xr −1 ) − k (xr +1 − xr ) = 0

which, for harmonic motion xr = ar sin ω t , can be expressed in terms of


amplitudes as
⎛ ω 2m ⎞
ar +1 − 2 ⎜1 − ⎟ ar + ar −1 = 0 .
⎜ 2 k ⎟
⎝ ⎠
The solution of this equation is found by substituting

ar = e i β r
which leads to the relationship
⎛ ω 2m ⎞ e i β + e −i β
⎜1 − ⎟= = cos β ,
⎜ 2 k ⎟ 2
⎝ ⎠
5. SEVERAL DEGREES OF FREEDOM 203

which can also be written


ω 2m β
= 2 (1 − cos β ) = 4 sin 2 . (5.18)
k 2

The general solution for ar is


ar = C 1 cos β r + C2 sin β r ,

where C 1 and C 2 are determined from the boundary conditions. At r = 0 the


amplitude is zero a0 = 0 , so that C 1 = 0 . At the free end, r = n , the equation of
motion is
m &x&n + k (xn − xn −1 ) = 0
which, in terms of amplitudes, becomes
⎛ ω 2m ⎞
an −1 = ⎜1 − ⎟ an .

⎝ k ⎟⎠

Substituting from the general solution, we obtain the following relationship


for the evaluation of β :
sin β (n − 1) = [ 1 − 2 (1 − cos β )] sin β n .
This result can be reduced to the product form
⎛ 1⎞ β
2 cos β ⎜ n + ⎟ sin = 0
⎝ 2⎠ 2
which is satisfied by
β
sin =0 ,
2
⎛ 1⎞ β (2 r − 1)π
and by cos β ⎜ n + ⎟ = 0 , or = . (r = 1, . . . , n )
⎝ 2⎠ 2 2 (2 n + 1)
The natural frequencies are then available from equation (5.18) as
k β
ω=2 sin (5.19)
m 2
which leads to

ωr = 2
k
sin
(2 r − 1)π . (r = 1, . . . , n ) (5.20)
m 2 (2 n + 1)
204 MECHANICAL VIBRATIONS

Natural frequencies calculated by the method of difference equation are


always given by equation (5.19). However, for each repeated structure the quantity
β must be determined from the appropriate boundary conditions.

5.1.5 Multi-Mass-Spring-Dashpot Systems

In an one-dimensional vibrating system, every mass moves only in one


direction. A node is assigned at each lumped mass, each node having only one
degree of freedom. A blocked node is assigned at a fixed boundary.

Fig. 5.13

The four-mass model shown in Fig. 5.13, a has four degrees of freedom
and five nodes. The nodal displacements are denoted q 1 , q 2 , . . . , q 5 (Fig. 5.13, b).
The column vector { Q } = { q 1 , q 2 , . . . , q 5 }T is called the global vector of nodal
displacements and { F } = { f1 , f 2 , . . . , f 5 }T is the global vector of nodal forces. A
displacement or force has a positive value if acting along the positive q direction.
At this stage, the boundary condition q 5 = 0 is not imposed.

Table 5.1
Node
Element
1 2
1 1 2
2 2 3
3 3 4
4 4 5
5 2 4
6 3 5
5. SEVERAL DEGREES OF FREEDOM 205

The six springs are numbered according to their index. Each spring has two
nodes. The element connectivity information can be conveniently represented as in
Table 5.1. In the connectivity table, local node numbers are 1 and 2, while global
node numbers are i and j. Such a table establishes the local-global correspondence.

Referring to Fig. 5.13, c, the vector of local nodal forces { f }= { f , f


e
1 2 }T
is related to the vector of local nodal displacements { q }= { q , q }
e
1 2
T
by the
equation { f }= [ k ] { q } where the element stiffness matrix is
e e e

[ k ] = ⎡⎢−kk
e e − ke ⎤
ke ⎥⎦
. (5.21)
⎣ e

This can be established from the equilibrium and force-deflection


equations: for q 2 = 0 , f1 = − f 2 = k q 1 , and for q 1 = 0 , f1 = − f 2 = −k q 2 .

On the other hand, the global vector of nodal forces { F } is related to the
global vector of displacements { Q } by the equation { F } = [ K ] { Q } where [ K ]
is the unreduced global stiffness matrix.
The matrix [ K ] can be obtained by the direct stiffness approach. Using the
element connectivity information, the entries of each element matrix [ k ] are
e

placed in the appropriate locations in the larger [K] matrix and overlapping
elements are then summed.
The assembly of the global stiffness matrix can be explained by the
summation of element strain energies.
The strain energy in, say, spring 3, is

U3 =
1
{ q } [ k ]{ q }= 12 ⎣q
3 T 3 3
3
⎡ k
q4 ⎦ ⎢ 3
− k3 ⎤ ⎧ q3 ⎫
⎨ ⎬.
k3 ⎥⎦ ⎩ q4 ⎭
2 ⎣ − k3
Expanding the stiffness matrix at the system size we obtain
T
⎧ q1 ⎫ ⎡0 0 0 0 0⎤ ⎧ q1 ⎫
⎪q ⎪ ⎢0 ⎪ q2 ⎪⎪
0 0 0 0 ⎥⎥ ⎪
1 ⎪⎪ ⎪⎪ ⎢
[ ]{ Q }
2
q3 ⎬ = { Q } k 3
⎪ ⎪ 1 T ~
U 3 = ⎨ q3 ⎬ ⎢0 0 k3 − k3 0⎥ ⎨
2⎪ ⎪ ⎢ ⎥ 2
q ⎢0 0 − k3 k3 0⎥ ⎪ q4 ⎪
⎪ 4⎪ ⎪ ⎪
⎪⎩ q5 ⎪⎭ ⎢⎣ 0 0 0 0 0 ⎥⎦ ⎪⎩ q5 ⎪⎭
~
[ ]
where k 3 is the expanded stiffness matrix of spring 3.
206 MECHANICAL VIBRATIONS

~
[ ]
We see that elements of the matrix k 3 are located in the third and fourth
rows and columns of the [ K ] matrix. When adding spring strain energies

U = ∑U e =
1
{ Q }T [ K ] { Q },
e 2

[ ]
the elements of k e are placed in the appropriate locations of the global [ K ]
matrix, based on spring connectivity. Overlapping elements are simply added so
that
∑[ ]
[ K ] = k~ e .
e

For the system of Fig. 5.13, a, the non-reduced global stiffness matrix is
⎡ k1 − k1 0 0 0 ⎤
⎢−k k +k +k − k2 − k5 0 ⎥⎥
⎢ 1 1 2 5
[K ] = ⎢ 0 − k2 k 2 + k3 + k 6 − k3 − k6 ⎥ .
⎢ ⎥
⎢ 0 − k5 − k3 k 3 + k 4 + k5 − k4 ⎥
⎢⎣ 0 0 − k6 − k4 k 4 + k6 ⎥⎦

The boundary conditions must now be specified. Node 5 is fixed so that


q5 = 0 and should be eliminated from the displacement vector. A reduced global
stiffness matrix is obtained by eliminating, from the original stiffness matrix, the
row and column corresponding to the specified or “support” degree of freedom.
For the system of Fig. 5.13, a the reduced global stiffness matrix is
⎡ k1 − k1 0 0 ⎤
⎢− k k + k + k −k2 −k5 ⎥
[K ] = ⎢⎢ 0 1 1 − k2 5 k 2 + k3 + k 6 −k3
⎥.

2
⎢ ⎥
⎣⎢ 0 − k 5 −k3 k 3 + k 4 + k5 ⎦⎥

Together with the diagonal mass matrix


⎡m 1 0 0 0 ⎤
⎢0 m 0 0 ⎥⎥
[M ] = ⎢⎢ 0 02 m3 0 ⎥
⎢ ⎥
⎢⎣ 0 0 0 m 4 ⎥⎦

it is used to write the equations of free motion


[ M ]{Q&& }+ [ K ]{ Q } = { 0 }
5. SEVERAL DEGREES OF FREEDOM 207

and to solve the corresponding eigenvalue problem

[ K ]{Φ } = ω 2 [ M ]{Φ },
to determine the real modes of vibration of the undamped system.
For a system including dashpots, the same procedure is used to assemble
the global damping matrix [ C ] . For the model from Fig. 5.14 the same
connectivities have been taken for dashpots as for springs, though generally they
can be different.

Fig. 5.14

The global damping matrix is


⎡ c1 − c1 0 0 ⎤
⎢− c c + c + c −c2 − c5 ⎥
[C ] = ⎢⎢ 0 1 1 − c2 5 c 2 + c3 + c6 − c3
⎥.

2
⎢ ⎥
⎣⎢ 0 − c 5 − c3 c 3 + c4 + c5 ⎦⎥

The equations of free motion of the damped system can be written


[ M ]{Q&& }+ [ C ]{Q& }+ [ K ]{Q } = { 0 } .
In this case the system has complex modes of vibration. The complex
eigenvalues yield the damped natural frequencies and the modal damping ratios.
The systems with proportional damping have real modal vectors. Vibrations of
damped systems are treated as in Section 4.6.2 and are dealt with in a next chapter.

Example 5.7
Calculate the natural frequencies and mode shapes of the system shown in
Fig. 5.15. The system parameters are: m 1 = m 2 = .... = m 11 = 1 kg , k 1 = k 11 = 2421 N/m ,
k 2 = k 10 = 2989 N/m , k 3 = k 9 = 3691 N/m , k 4 = k 8 = 4556 N/m , k 5 = k 7 = 5625 N/m ,
208 MECHANICAL VIBRATIONS

k 6 = 18000 N/m .

Solution. The natural frequencies, in Hz, are 2.74, 2.95, 7.24, 7.80, 11.47,
12.13, 15.00, 15.62, 18.49, 19.32 and 28.57. There are pairs of two close natural
frequencies, one for a symmetric and one for an antisymmetric mode. This is
typical for symmetric structures.

Fig. 5.15

The first 10 mode shapes are presented in Fig. 5.16.

Fig. 5.16
5. SEVERAL DEGREES OF FREEDOM 209

Example 5.8
For the 15-dof system shown of Fig. 5.17 set up the matrices M, K and C.
Find the damped natural frequencies and the modal damping ratios.

Fig. 5.17

Solution. Values obtained using the program VIBMKC are listed in Table
5.2. Damping ratio values are multiplied by 100.

Table 5.2

Mode ω d , Hz ζ ,% Mode ωd , Hz ζ ,%
1 15.98 0.502 9 68.88 1.378
2 30.86 0.968 10 73.72 1.579
3 43.60 1.364 11 128.87 0.536
4 46.47 0.301 12 136.59 0.506
5 53.35 1.665 13 143.89 0.477
6 53.42 0.670 14 150.87 0.457
7 59.45 1.853 15 157.52 0.437
8 61.62 1.060

The five masses located on the right are one order of magnitude smaller
than the others. This produces the cluster of five higher natural frequencies. Due to
the relatively low damping values, the damped natural frequencies are
approximately equal to the undamped natural frequencies.
210 MECHANICAL VIBRATIONS

5.2 Plane Trusses

A truss structure consists of pin-jointed members. The main simplifying


assumption in trusses is that all members are connected together at their ends by
frictionless pin joints and thus cannot transmit moments from one another.
In practice, the joints of a truss are made by riveting, welding or bolting.
However, a simplified model with pin joints is a surprisingly good engineering
approximation. Truss elements can only sustain tension or compression, such
members being called ties or struts respectively. Rigid jointed structures are dealt
with in Section 5.3.
In a truss, it is required that all loads and reactions are applied only at the
joints and members have constant axial rigidity, so that they are natural finite
elements. To account for the spatial orientations of truss members, local and global
coordinate systems are used. In the following, the element stiffness and mass
matrices are calculated first in the local coordinate system, then in the global
coordinate system. The latter can be expanded to the system size, then simply
added to obtain the global stiffness and mass matrices to be used in the eigenvalue
problem or dynamic response calculations.

5.2.1 Coordinates and Shape Functions for the Truss Element

Consider a two-noded pin-jointed element in the own or local coordinate


system. Nodes are conveniently numbered 1 and 2, their coordinates in the physical
(cartesian) reference system being x 1 and x 2 respectively (Fig. 5.18, a).

Fig. 5.18

We define a natural or intrinsic reference system which permits the


specification of a point within the element by a dimensionless number

2 ⎛ x + x2 ⎞
r= ⎜x− 1 ⎟ (5.22)

x 2 − x1 ⎝ 2 ⎟⎠

so that r = −1 at node 1 and r = +1 at node 2 (Fig. 5.18, b).


5. SEVERAL DEGREES OF FREEDOM 211

Expressing the physical coordinate in terms of the natural coordinate yields


x = N1 (r ) x 1 + N 2 (r ) x 2 , (5.23)
where
1
N1 (r ) = ( 1 − r ) and N 2 (r ) = 1 ( 1 + r ) (5.24)
2 2
can be considered as geometric interpolation functions. The graphs of these
functions are shown in Figs. 5.19, a, b.

Fig. 5.19

For a two-noded bar element, a linear distribution of displacements can be


assumed. The displacement of an arbitrary point within the element can be
expressed in terms of the nodal displacements q 1 and q 2 as

u (r ) = N1 (r ) q 1 + N 2 (r ) q 2 . (5.25)

In matrix notation

{ }
2
u = ∑ N i qi = ⎣N ⎦ q e , (5.26)
i =1
where
⎣N ⎦ = ⎣N1 N 2 ⎦ and {q }= { q
e
1 q2 }T . (5.27)

In equation (5.26), { }
q e is referred to as the element displacement
vector and ⎣N ⎦ is the row vector of displacement interpolation functions also
named shape functions. It is easy to check that u = q 1 at node 1 and u = q 2 at
node 2, and that u varies linearly (Fig. 5.19, c).
212 MECHANICAL VIBRATIONS

Equations (5.23) and (5.25) show that both the element geometry and the
displacement field are interpolated using the same shape functions, which is
referred to as the isoparametric formulation.

5.2.2 Element Stiffness and Mass Matrices in Local Coordinates

In dynamic analysis, the displacement is a function of both space and time


u = u ( x ,t ) , the strains are ε = ∂ u ∂ x and the stresses are σ = Eε .
The element strain energy U e (t ) is
2
1 1 ⎛ ∂u ⎞
∫ ∫ ∫
Ee Ae
Ue = σ ε dV = Ee ε 2 Ae dx = ⎜⎜ ⎟⎟ d x . (5.28)
2
e
2
e
2
e
⎝ ∂x ⎠

The transformation from x to r in equation (5.22) yields


x 2 − x1 le
dx = dr = dr , (5.29)
2 2

where − 1 ≤ r ≤ +1 and the length of the element is l e = x 2 − x 1 .

Because
∂u ⎢ ∂ N ⎥ e
=⎢
∂x ⎣ ∂x ⎦
⎥ q and { }
∂N ∂N ∂ r 2 ∂N
∂x
= =
∂r ∂ x l e ∂r
where
∂ N1
∂r
=−
1
2
2
{ } { }
T
∂ N2 1 ⎛ ∂u ⎞ T ⎢∂N ⎥ ⎢∂N ⎥ e
and = + , we can write ⎜⎜ ⎟⎟ = q e ⎢ ∂ x ⎥ ⎢ ∂ x ⎥ q , so that the
∂r 2 ⎝ ∂x ⎠ ⎣ ⎦ ⎣ ⎦
element strain energy (5.28) becomes
+1

{q } { q }.
T
⎢∂N ⎥ ⎢∂N ⎥

1 e T 2 Ee Ae
Ue = ⎢ ⎥ ⎢ ⎥dr
e
(5.30)
2 le ⎣ ∂r ⎦ ⎣ ∂r ⎦
−1

The above equation is of the form

Ue =
1
2
{ q } [ k ] { q },
e T e e
(5.31)

[ ]
where the element stiffness matrix k e is given by
+1 +1

[k ]
T
⎢∂N ⎥ ⎢∂N ⎥ ⎧− 1 2⎫ ⎢ 1
∫ ∫
2E A 2E A 1⎥
e
= e e ⎢ ⎥ ⎢ ⎥ dr = e e ⎨ ⎬⎢ − dr
le ⎣ ∂r ⎦ ⎣ ∂r ⎦ le ⎩ 1 2 ⎭⎣ 2 2 ⎥⎦
−1 −1
or
5. SEVERAL DEGREES OF FREEDOM 213

[ k ] = El A ⎡⎢−11
e e e − 1⎤
1 ⎥⎦
. (5.32)
e ⎣
The element kinetic energy Te (t ) is
2
ρ Ae ⎛ ∂u ⎞
Te =
2 ∫ e
⎜⎜ ⎟⎟ d x .
⎝ ∂t ⎠
(5.33)

where ρ is the material mass per unit volume and ∂u ∂ t = u& is the velocity at x.
From (5.26) we obtain

{ }
u& = ⎣N ⎦ q& e , (5.34)
where { q& } is the column vector of nodal velocities.
e

Substituting (5.34) into equation (5.33) yields

Te =
1
2
{ q& }e T
ρ Ae ∫e ⎣N ⎦
T
⎣N ⎦ d x { q& }.
e
(5.35)

Equation (5.35) is of the form

Te =
1
2
{ q& } [ m ] { q& }.
e T e e
(5.36)
where
[ m ] = ρ A ∫ ⎣N ⎦
e
e
T
⎣N ⎦ d x (5.37)
e
is called the element consistent mass matrix. It is calculated using the same
procedure and shape functions as the element stiffness matrix.
Changing the variable
+1 +1

[m ]= e ρ Ae l e
2 ∫ ⎣N ⎦ ⎣N ⎦ d r =
T ρ Ae l e
8 ∫
⎡(1 − r ) 2
⎢ 2
1− r 2 ⎤
⎥dr
(1 + r ) 2 ⎥⎦
⎢⎣ 1 − r
−1 −1
or

[ m ] = ρ A6 l
e e e ⎡2 1 ⎤
⎢1 2 ⎥ . (5.38)
⎣ ⎦

5.2.3 Transformation from Local to Global Coordinates

A typical plane truss element is shown in Fig. 5.20 where both the local
coordinate system xOy and the global coordinate system XOY are drawn. Nodal
214 MECHANICAL VIBRATIONS

displacements are denoted by small letters in the local coordinate system and by
capital letters in the global coordinate system.
In the global coordinate system, every node has two degrees of freedom
(dof’s). A node whose global node index is j has associated with it dof’s 2 j − 1 and
2 j , and displacements Q2 j −1 and Q2 j .

In Fig. 5.20 we see that q 1 equals the sum of projections of Q 1 and Q 2


onto the x axis. Thus
q 1 = Q 1 cos α + Q 2 sin α . (5.39, a)
Similarly
q 2 = Q 3 cos α + Q 4 sin α . (5.39, b)

Fig. 5.20

Equations (5.39) can be written in matrix form as

{ q } = [ T ] { Q },
e e e
(5.40)

where {q }= {q
e
1 q2 }T is the element displacement vector in the local
{ } {
coordinate system, Q e = Q 1 , Q 2 , Q 3 , Q 4 }T is the element displacement vector
in the global coordinate system and

[T ] = ⎡⎢cos0 α
e sin α
0
0
cos α
0 ⎤
sin α ⎥⎦
(5.41)

is a coordinate transformation matrix.
From nodal coordinate data, denoting ( X 1 ,Y1 ) and ( X 2 ,Y 2 ) the
coordinates of nodes 1 and 2, respectively, we obtain
5. SEVERAL DEGREES OF FREEDOM 215

X 2 − X1 Y 2 − Y1
cos α =
le
, sin α =
le
, le = (X 2 − X 1 ) 2 + (Y 2 − Y1 ) 2 . (5.42)

The entries of the matrix (5.41) are calculated based on the above equations.

5.2.4 Element Stiffness and Mass Matrices in Global Coordinates

Substituting equation (5.40) into the expression (5.31) of element strain


energy in local coordinates, we get

Ue =
1
2
{ Q } [T ] [ k ][T ]{ Q }.
e T e T e e e
(5.43)

The strain energy in global coordinates can be written as

Ue =
1
2
{ Q } [ K ]{ Q },
e T e e
(5.44)

[ ]
where K e is the element stiffness matrix in global coordinates.
Comparing equations (5.43) and (5.44) we obtain the stiffness matrix in
global coordinates as

[ K ]= [T ] [ k ][T ] .
e e T e e
(5.45)

Substituting for [T ] from equation (5.41) and for [ k ] from equation


e e

(5.32) we get
⎡ c2 cs − c2 − cs⎤
⎢ ⎥
[K ]e
= e e⎢ 2
E A cs s2 − cs − s2 ⎥
l e ⎢− c − c s c 2 cs ⎥
, (5.46)
⎢ 2

⎣⎢ − c s − s cs s 2 ⎦⎥
where c = cos α and s = sin α .
Similarly, the kinetic energy in global coordinates can be written as

Te =
1
2
{ Q& } [ M ]{ Q& }.
e T e e
(5.47)

The element consistent mass matrix in global coordinates is

[ M ]= [T ] [ m ][T ] ,
e e T e e

[ ] [ ]
or, substituting for T e from equation (5.41) and for m e from equation (5.38)
216 MECHANICAL VIBRATIONS

⎡2 c2 2 c s c2 cs ⎤
⎢ ⎥
[M ]
2
ρ Ae l e ⎢2 c s 2s cs s2 ⎥
e
= . (5.48)
6 ⎢ c2 c s 2 c2 2 c s⎥
⎢ ⎥
⎢⎣ c s s 2 2 c s 2s 2 ⎥⎦
This matrix is singular (order 4 and rank 1) because the element is not
grounded. The rank deficiency is equal to the three possible rigid body motions.

5.2.5 Assembly of Stiffness and Mass Matrices

The global stiffness and mass matrices, [ K ] and [ M ] , are assembled from
[ ] [ ]
element matrices K e and M e using the element connectivity information.
The compatibility of nodal displacements, at element level with the nodal
displacements at the whole truss structure level, can be expressed by equations of
the form

{ } [ ]
Q e = T e { Q },
~
(5.49)

{ } is the element displacement vector in global coordinates, {Q } is the


where Q e
displacement vector of the truss structure and [T ] is referred to as a
~ e
full
connectivity or localisation matrix, containing ones at the dof’s of element nodes
and zeros elsewhere.
The element strain energy in global coordinates can be written in terms of
the global displacement vector, substituting (5.49) into (5.44)

Ue =
1
{ Q }T [T~ e ] T [ K e ][T~ e ]{ Q }
2
or
1
{ Q }T K~ e
Ue =
2
[ ]{ Q },
where the expanded element stiffness matrix

[ K~ ]= [T~ ] [ K ][T~ ]
e e T e e
(5.50)

has the size of the system matrices.


To illustrate this, consider the seven-bar truss shown in Fig. 5.21.
The connectivity matrix of element 4, where the two nodes are not
consecutively numbered, is
5. SEVERAL DEGREES OF FREEDOM 217

⎡0 0 1 0 0 0 0 0 0 0⎤
⎢0 0 0 1 0 0 0 0 0 0⎥⎥
[T~ ]
4
=⎢
⎢0 0 0 0 0 0 1 0 0 0⎥
⎢ ⎥
⎣0 0 0 0 0 0 0 1 0 0⎦

while, for the element 5, where the two nodes are consecutively numbered, it is
⎡0 0 0 0 1 0 0 0 0 0⎤
⎢0 0⎥⎥
[T~ ]
5
=⎢
⎢0
0
0
0
0
0
0
0
0
1
0
0
1
0
0
0
0 0⎥
.
⎢ ⎥
⎣0 0 0 0 0 0 0 1 0 0⎦

Fig. 5.21

The corresponding expanded stiffness matrices are of the form

1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
⎡ ⎤ 1 ⎡ ⎤ 1
⎢ ⎥ ⎢ ⎥
⎢ ⎥ 2 ⎢ ⎥ 2
⎢ • • • • ⎥ 3 ⎢ ⎥ 3
⎢ ⎥ ⎢ ⎥

[ K~ ] = [ K~ ] =
⎢ • • • • ⎥ 4 ⎢ ⎥ 4
4 ⎢ ⎥ 5 5 ⎢ ⎥ 5
⎢ ⎥ , ⎢
• • • •

⎢ ⎥ 6 ⎢ • • • • ⎥ 6
⎢ ⎥ ⎢ ⎥
⎢ • • • • ⎥ 7 ⎢ • • • • ⎥ 7
⎢ • • • • ⎥ 8 ⎢ • • • • ⎥ 8
⎢ ⎥ ⎢ ⎥
⎢ ⎥ 9 ⎢ ⎥ 9
⎢ ⎥ 10 ⎢ ⎥ 10
⎣ ⎦ ⎣ ⎦

We may imagine that the structure is built by adding elements one by one,
with each element being placed in a preassigned location. As elements are added to
the structure, contributions are made to the structure load carrying capacity, hence
to the structure stiffness matrix. We may add the element stiffness matrices to
218 MECHANICAL VIBRATIONS

obtain the stiffness matrix of the entire structure provided that element matrices are
of the “structure size” and operate on identical displacement vectors. Simple
addition of the expanded element stiffness matrices produces the structure stiffness
matrix.
The strain energy for the complete truss structure

U=
1
{ Q }T [ K ]{ Q }, (5.51)
2
can be calculated by simply adding element strain energies

∑U = ∑ 2 { Q } [ K ]{ Q } = 2 { Q } ∑ [ K ] { Q }.
1 ~ T 1 ~ T
U= e
e e
(5.52)
e e e

Comparing (5.51) and (5.52) we get

∑e [ ]
[ K ] = K~ e . (5.53)

The global stiffness matrix is equal to the sum of the expanded element
stiffness matrices.
Similarly, the global mass matrix [M ] is assembled from expanded
element mass matrices as

∑e [
[ M ] = M~ e = T~ e ] ∑ [ ] [ M ][T~ ].
e
T e e
(5.54)

The unreduced stiffness and mass matrices [ K ] and [ M ] are used in the
case of free-free systems. For grounded systems, they are condensed using the
boundary conditions.
The effect of lumped masses and springs can be accounted for by adding
their values along the main diagonal at the appropriate locations in the respective
matrices.

5.2.6 Equations of Motion and Eigenproblem

We define the Lagrangean L by


L =T −Π , (5.55)
where T is the kinetic energy

T=
1
2
{ Q& }T [ M ] { Q& } (5.56, a)

and Π is the potential energy


5. SEVERAL DEGREES OF FREEDOM 219

1
Π = { Q }T [ K ] { Q } − { Q } T { F } . (5.56, b)
2
In (5.56, b), { F } is the global vector of applied nodal forces.
Hamilton’s principle leads to Lagrange’s equations for undamped systems

d ⎛⎜ ∂ L ⎞
⎟ − ∂L = 0 d ⎛⎜ ∂T ⎞
⎟ − ∂Π = 0 .
or (5.57)
⎝ { }
dt ⎜ ∂ Q& ⎟ ∂ {Q }
⎠ { }
d t ⎜ ∂ Q&

⎟ ∂ {Q }

Using the rules of differentiation of a scalar with respect to a vector, we
obtain the equations of motion
[ M ] {Q&& }+ [ K ] { Q } = { F }. (5.58)
For free vibrations, the force vector is zero. Thus
[ M ] {Q&& }+ [ K ] { Q } = { 0 }. (5.59)
Looking for solutions of the form
{ Q } = {φ }sin ω t , (5.60)

where {φ } is the vector of nodal amplitudes of vibration, we obtain the


generalized eigenvalue problem

[ K ] {φ }r = ωr2 [ M ] {φ }r , (r = 1,..., n ) , (5.61)

where ωr2 are real eigenvalues equal to the natural frequencies squared and {φ }r
are real eigenvectors.
The program VIBTRUSS (written in MATLAB) outputs the natural
frequencies and mode shapes for undamped truss structures.

Fig. 5.22
220 MECHANICAL VIBRATIONS

Example 5.9
Calculate the first four natural frequencies and mode shapes for the system
shown in Fig. 5.22, a where E = 200 GPa , ρ = 7850 kg m 3 and A = 100 mm 2
for all 20 bars.
Solution. The lowest four natural frequencies, in Hz, are 48.8, 168.4, 235.9
and 336.8. The mode shapes are shown in Figs. 5.22, b to e.

5.3 Plane Frames

Frames are structures with rigidly connected members called beams.


Beams are slender members used to support transverse loading. They are connected
by rigid (nodal) joints that have determinate rotations and, apart from forces,
transmit bending moments from member to member.
In this section, we first present the finite element formulation for beams,
then extend it to plane frames. An inclined beam element will be referred to as a
frame element.

5.3.1 Static Analysis of a Uniform Beam

Beams with cross sections that are symmetric with respect to the plane of
loading are considered herein (Fig. 5.23). Transverse shear deformations are
neglected.

Fig. 5.23

The axial displacement of any point on the section, at a distance y from the
neutral axis, is approximated by
dv
u = −ϕ y = − y, (5.62)
dx
5. SEVERAL DEGREES OF FREEDOM 221

where v is the deflection of the centroidal axis at x and ϕ = v′ is the cross section
rotation (or slope) at x . Axial strains are

du d2 v
εx =
=− 2 y. (5.63)
dx dx
Normal stresses on the cross section are given by
d2 v
σ x = E ε x = −E 2 y , (5.64)
dx
where E is Young’s modulus of the material.
The bending moment is the resultant of the stress distribution on the cross
section
d2 v
M (x ) = − σ x y dA = E I z
∫ 2
= E I z v II . (5.65)
dx
A
where I z is the second moment of area of the section about the neutral axis z.
The shear force is given by
dM d3 v
T (x ) = = EIz = E I z v III . (5.66)
dx dx 3
The transverse load per unit length is
dT d4 v
p (x ) = = EIz 4
= E I z v IV . (5.67)
dx dx
The differential equation of equilibrium is
d4 v
EIz = p (x ) . (5.68)
dx4

5.3.2 Finite Element Discretization

The plane frame is divided into elements, as shown in Fig. 5.24. Each node
has three degrees of freedom, two linear displacements and a rotation. Typically,
the degrees of freedom of node i are Q3 i − 2 , Q3 i −1 and Q3 i , defined as the
displacement along the X axis, the displacement along the Y axis and the rotation
about the Z axis, respectively.
Nodes are located by their coordinates in the global reference frame XOY
and element connectivity is defined by the indices of the end nodes. Elements are
modelled as uniform beams without shear deformations and not loaded between
222 MECHANICAL VIBRATIONS

ends. Their properties are the bending rigidity E I , the mass per unit length ρ A
and the length l .

Fig. 5.24

In the following, the shape functions are established for the beam element,
then the element stiffness and mass matrices are calculated first in the local
coordinate system, then in the global coordinate system. The latter are expanded to
the structure size, then simply added to get the global uncondensed stiffness and
mass matrices. Imposing the boundary conditions, the reduced stiffness and mass
matrices are calculated which, together with the damping matrix, are used in the
dynamic analysis.

Fig. 5.25

Consider an inclined beam element, as illustrated in Fig. 5.25, a, where the


nodal displacements are also shown.
5. SEVERAL DEGREES OF FREEDOM 223

In a local physical coordinate system, the x axis, oriented along the beam,
is inclined an angle α with respect to the global X axis. Alternatively, an intrinsic
(natural) coordinate system can be used.
The vector of element nodal displacements is

{ q }= { q , q
e
1 2, q3, q 4, q5, q6 }T (5.69)

and the corresponding vector of element nodal forces can be written

{ f }= { f
e
1, f 2, f 3, f 4, f 5, f 6 }T . (5.70)

Forces f 2, f 3, f 5, f 6 and the corresponding displacements


q 2 , q 3 , q 5 , q 6 describe the element bending (Fig. 5.25, b) while axial forces
f 1 , f 4 , and displacements q 1 , q 4 , describe the element stretching (Fig. 5.25, c).
Their action is decoupled so that the respective stiffness and mass matrices can be
calculated separately.

5.3.3 Static Shape Functions of a Beam Element

For a uniform beam not loaded between ends, p = 0 and equation (5.68)
yields d 4 v d x 4 = 0 . Integrating four times, we obtain the deflection v described
by a third order polynomial

v (x ) = a 1 x 3 + a 2 x 2 + a 3 x + a 4 . (5.71)

In (5.71), the four integration constants a 1 , a 2 , a 3 , a 4 can be determined


from the geometric boundary conditions, involving the transverse displacement and
slope at each end:
x = x1 , v = q2 , d v d x = q3 , and x = x 2 , v = q5 , d v d x = q6 . (5.72)
Alternatively, the transverse displacement can be expressed in terms of the
nodal displacements as

{ }
v = ⎣N ⎦ q e , (5.73)

where ⎣N ⎦ is a row vector containing the shape functions, which are cubic
polynomials, called Hermite polynomials.
Using natural coordinates, with r = −1 at node 1 and r = +1 at node 2, the
transverse displacement can be written
224 MECHANICAL VIBRATIONS

⎛dv⎞ ⎛dv⎞
v (r ) = N1 (r ) v 1 + N 2 (r ) ⎜⎜ ⎟⎟ + N 3 (r ) v 2 + N 4 (r )⎜⎜ ⎟⎟ . (5.74)
⎝ d r ⎠1 ⎝ d r ⎠2
Because the coordinates transform by the relationship (5.22)
x1 + x 2 x 2 − x1
x= + r (5.75)
2 2
and since l e = x 2 − x 1 is the length of the element, equation (5.29) holds

le
dx = dr . (5.76)
2
Using the chain rule of differentiation
d v le d v
= , (5.77)
dr 2 dx
equation (5.74) becomes
le ⎛dv⎞ l ⎛dv⎞
v (r ) = N1 (r ) v 1 + N 2 (r ) ⎜⎜ ⎟⎟ + N 3 (r ) v 2 + N 4 (r ) e ⎜⎜ ⎟⎟ (5.78)
2 ⎝ d x ⎠1 2 ⎝ d x ⎠2
or
le l
v (r ) = N1 ⋅ q2 + N 2 ⋅ q3 + N 3 ⋅ q5 + e N 4 ⋅ q6 . (5.79)
2 2
In (5.73) the row vector of shape functions is
⎢ l ⎥
l
⎣N ⎦ = ⎢ N1 , 2e N 2 , N 3 , 2e N 4 ⎥ . (5.80)
⎣ ⎦
The Hermite shape functions are cubic polynomials which should satisfy
the boundary conditions given in Table 5.3 where primes indicate differentiation
with respect to r.
Table 5.3
N1 N ′1 N2 N ′2 N3 N ′3 N4 N ′4
r = −1 1 0 0 1 0 0 0 0
r = +1 0 0 0 0 1 0 0 1

Imposing the above conditions to cubic polynomials with four arbitrary


constants, we obtain the expressions of the beam element shape functions in natural
coordinates (5.81), graphically shown in Fig. 5.26.
5. SEVERAL DEGREES OF FREEDOM 225

N 1(r ) =
1
4 4
( )
( 1 − r ) 2 ( 2 + r ) = 1 2 − 3r + r 3 ,

N 2 (r ) =
1
4 4
( )
( 1 − r ) 2 ( 1 + r ) = 1 1 −r − r 2 + r 3 , (5.81)

N 3(r ) =
1
4
( )
( 1 + r ) 2 ( 2 − r ) = 1 2 + 3r − r 3 ,
4

N 4 (r ) = −
1
4 4
( )
( 1 + r ) 2 ( 1 − r ) = − 1 1 + r −r 2 − r 3 .

Fig. 5.26

d v le
It is easy to check that at node 1, v = q 2 and = q 3 , while at node 2,
dr 2
d v le
v = q 5 and = q6 .
dr 2

5.3.4 Stiffness Matrix of a Beam Element

The strain energy U e of a beam element is


2
⎛d 2v⎞

E Ie
Ue = ⎜ ⎟
2 ⎜ d x2 ⎟ d x . (5.82)
e
⎝ ⎠
226 MECHANICAL VIBRATIONS

Equation (5.77) yields


dv 2 d v d 2v 4 d 2 v
= and = .
d x le d r d x 2 l 2e d r 2
Substituting (5.73) we obtain

d 2v 4 ⎢ d 2 N ⎥ e
= ⎢ ⎥ q .
d x 2 l 2e ⎣⎢ d r 2 ⎦⎥
{ } (5.83)

The square of the above quantity is calculated as

2 T T
⎛ d 2v ⎞

⎜ d x2 ⎟
⎛ 2 ⎞
⎟ =⎜d v⎟
⎜ d x2 ⎟
⎛ d 2v ⎞
⎜ ⎟ = qe
⎜ d x2 ⎟
{ } T 16 ⎢ d 2 N ⎥ ⎢ d 2 N ⎥ e
⎢ ⎥ ⎢ ⎥ q ,
l 4e ⎣⎢ d r 2 ⎦⎥ ⎢⎣ d r 2 ⎦⎥
{ }
⎝ ⎠ ⎝ ⎠ ⎝ ⎠
which can be also written
2
⎛ d 2v ⎞
⎜ ⎟ = qe
⎜ d x2 ⎟
{ } T 16
l 4e
{ }
⎣N r′′ ⎦ ⎣N r′′ ⎦ q .
T e
(5.84)
⎝ ⎠

On substituting (5.76) and (5.84) into (5.82) we get the element strain
energy
+1
Ue =
1
2
{q } e T 8E I e
l 3e ∫ ⎣N r′′ ⎦ ⎣N r′′ ⎦ d r { q }
T e
(5.85)
−1

which has the form

Ue =
1
2
{ q } [ k ] { q }.
e T e
B
e
(5.86)

Comparing (5.85) with (5.86) we obtain the element stiffness matrix due to
bending
+1
[ ]=
k Be
8EI e
l 3e ∫ ⎣N r′′ ⎦ ⎣N r′′ ⎦ d r
T
(5.87)
−1
or
⎡ (N1′′)2 N1′′N 2′′ N1′′N 3′′ N1′′N 4′′ ⎤
+1 ⎢ ⎥
[ k ] = 8lEI ∫ ⎢ N 2′′ N1′′ (N 2′′ )
2
e N 2′′ N 3′′ N 2′′ N 4′′ ⎥
e
dr . (5.88)
B 3 ⎢ N ′′N ′′ N ′′N ′′ (N 3′′ )2 N 3′′N 4′′ ⎥
e −1 ⎢ 3 1 3 2 ⎥
⎢⎣ N 4′′ N1′′ N 4′′ N 2′′ N 4′′ N 3′′ (N 4′′ )2 ⎥⎦
5. SEVERAL DEGREES OF FREEDOM 227

Substituting the shape functions (5.81) and performing the integration


yields the stiffness matrix due to bending in local coordinates

⎡ 12 6l − 12 6l ⎤
⎢ 6l 4l 2 − 6l 2l 2 ⎥
[k ]
e
B
EI
= 3e ⎢
le ⎢ − 12 − 6 l 12 − 6l
⎥ .

(5.89)
⎢ 2 2⎥
⎣ 6l 2l − 6l 4l ⎦ e

5.3.5 Consistent Mass Matrix of a Beam Element

In dynamic calculations, the beam lateral deflection is a function of both


space and time, v = v (x, t ) .
The instantaneous kinetic energy of a beam element is
2
ρ Ae ⎛∂v⎞
Te =
2 ∫e
⎜⎜ ⎟⎟ d x .
⎝ ∂t ⎠
(5.90)

where ρ is the material mass per unit volume and ∂ v ∂ t = v& is the velocity at x.
From (5.73) we obtain

{ }
v& = ⎣N ⎦ q& e , (5.91)

where { q& } is the column vector of nodal velocities.


e

Substituting (5.76) and (5.91) into equation (5.90) yields

Te =
1
2
{ q& }
e T
ρ Ae ∫ ⎣N ⎦ ⎣N ⎦ d x { q& }.
T e
(5.92)
e
which has the form

Te =
1
2
{ q& } [ m ] { q& }.
e T e
B
e
(5.93)

where
+1
[ ]=
mBe
ρ Ae l e
2 ∫ ⎣N ⎦ r
T
⎣N r ⎦ d r (5.94)
−1
is the element consistent mass matrix.
On substituting the shape functions (5.81) and integrating their products
we get the beam element mass matrix in local coordinates
228 MECHANICAL VIBRATIONS

⎡ 156 22l 54 − 13l ⎤


⎢ − 3l 2 ⎥⎥
[m ]
2
ρ Ae l e ⎢ 22l 4l 13l
e
= . (5.95)
420 ⎢ 54 − 22l ⎥
B
13l 156
⎢ ⎥
⎣ − 13l − 3l 2 − 22l 4l 2 ⎦ e

This mass matrix is derived by the same approach as the stiffness matrix,
so that it is consistent with the stiffness matrix.

5.3.6 Axial Effects

The axial nodal forces are related to the nodal displacements by equation
⎧ f1 ⎫
⎨ ⎬ = kS ⎨ ⎬ [ ]
e ⎧ q1 ⎫
(5.96)
⎩ f4 ⎭ ⎩ q4 ⎭
where the stiffness matrix (5.32) is

[ k ] = EAl
e
S
e ⎡ 1 − 1⎤
⎢− 1 1 ⎥ . (5.97)
e ⎣ ⎦
Similarly, from (5.38) we obtain the element mass matrix due to stretching

[ m ] = ρ A6 l
e
S
e e ⎡2 1 ⎤
⎢1 2 ⎥ . (5.98)
⎣ ⎦

5.3.7 Frame Element Matrices in Local Coordinates

For the frame element, combining equations (5.97) and (5.89), and
arranging in proper locations we get the element stiffness matrix

⎡ EA EA ⎤
⎢ l 0 0 − 0 0 ⎥
l
⎢ 12 E I 6E I 12 E I 6E I ⎥
⎢ 0 0 − ⎥
⎢ l3 l2 l3 l2 ⎥
⎢ 6E I 4E I 6E I 2E I ⎥
⎢ 0 0 − 2
[k ]
e
=⎢ l2 l l l ⎥ .
⎥ (5.99)
⎢− EA 0 0
EA
0 0 ⎥
⎢ l l ⎥
⎢ 12 E I 6E I 12 E I 6E I ⎥
⎢ 0 − − 0 − 2 ⎥
⎢ l3 l2 l3 l
⎢ 0 6E I 2E I 6E I 4 E I ⎥⎥
0 − 2
⎢⎣ l2 l l l ⎥⎦ e
5. SEVERAL DEGREES OF FREEDOM 229

The ratio of the bending terms to the stretching terms in (5.99) is of order
(i l ) 2
, where ‘i’ is the relevant radius of gyration. For slender beams this ratio
may be as small as 1 20 or 1 50 , so the stiffness matrix may possibly be
numerically ill-conditioned.
Combining equations (5.98) and (5.95) and arranging in proper locations
we obtain the consistent mass matrix for the frame element

⎡140 0 0 70 0 0 ⎤
⎢ 0
⎢ 156 22l 0 54 − 13l ⎥⎥

[m ]
e
=
ρ Ae l e ⎢ 0
420 ⎢ 70

22l
0
4l 2
0
0
140
13l − 3l 2 ⎥
0 0 ⎥
⎥ . (5.100)

⎢ 0 54 13l 0 156 − 22l ⎥


⎢ 2 ⎥
⎣⎢ 0 − 13l − 3l 0 − 22l 4l 2 ⎦⎥ e

5.3.8 Coordinate Transformation

A frame element is shown in Fig. 5.27 both in the initial and deformed
state. For node 1, the local linear displacements q 1 and q 2 are related to the
global linear displacements Q 1 and Q 2 by the equations

q 1 = Q 1 cos α + Q 2 sin α ,
. (5.101)
q 2 = −Q 1 sin α + Q 2 cos α .

Fig. 5.27

Equations (5.101) can be written in matrix form as


230 MECHANICAL VIBRATIONS

⎧ q1 ⎫ ⎧ Q1 ⎫
⎨ q ⎬ = [ R] ⎨ Q ⎬ (5.101, a)
⎩ 2⎭ ⎩ 2⎭
where

[ R ] = ⎡⎢
c s⎤
(5.102)
⎣− s c ⎥⎦
is referred to as a rotation matrix and c = cos α and s = sin α .
The angular displacements (rotations) are the same in both coordinate
systems
q 3 = Q3 . (5.103)

Adding the similar relationships for node 2


⎧ q4 ⎫ ⎧ Q4 ⎫
⎨ q ⎬ = [R] ⎨Q ⎬ , q6 = Q6 ,
⎩ 5⎭ ⎩ 5⎭
we obtain

{ q } = [ T ] { Q },
e e e
(5.104)

where {q } is the element displacement vector in the local coordinate system,


e

{Q } is the element displacement vector in the global coordinate system and


e

⎡c s 0 0 0 0⎤
⎢− s
⎢ c 0 0 0 0⎥⎥

[T ]
e
⎢0
=⎢
0 1 0 0 0⎥
⎥ (5.105)
⎢0 0 0 c s 0⎥
⎢0 0 0 −s c 0⎥
⎢ ⎥
⎢⎣ 0 0 0 0 0 1⎥⎦
is the local-to-global coordinate transformation matrix.

5.3.9 Frame Element Matrices in Global Coordinates

Using the same procedure as in § 5.2.4, the stiffness and mass matrices of
the frame element in global coordinates are obtained as

[ K ]= [T ] [ k ][T ] ,
e e T e e
(5.106)
and

[ M ]= [T ] [ m ][T ] .
e e T e e
(5.107)
5. SEVERAL DEGREES OF FREEDOM 231

5.3.10 Assembly of Stiffness and Mass Matrices

The global stiffness and mass matrices, [ K ] and [ M ] , are assembled from
[ ] [ ]
element matrices K e and M e using element connectivity matrices T e that [ ]~
relate the nodal displacements at element level with the nodal displacements at the
complete structure level by equations of the form

{ } [ ]
Q e = T e { Q }.
~
(5.108)
The global uncondensed stiffness matrix is equal to the sum of the
expanded element stiffness matrices

∑e [ ]
[ K ] = K~ e . (5.109)

where
[ K~ ]= [T~ ] [ K ][T~ ]
e e T e e
(5.110)

Similarly, the global uncondensed mass matrix [ M ] is assembled from


expanded element mass matrices as

∑e [ ] ∑ [ ] [ ][ ]
[ M ] = M~ e = T~ e T M e T~ e .
e
(5.111)

For grounded systems the unreduced stiffness and mass matrices [ K ] and
[ M ] are condensed using the boundary conditions.
The effect of lumped masses and springs can be accounted for by adding
their values along the main diagonal at the appropriate locations in the respective
matrices. When the external loads include distributed forces, they are replaced by
kinematically equivalent nodal forces, calculated by a procedure which is
consistent with the derivation of the stiffness matrix and the mass matrix, assuming
the validity of the static shape functions.
The equations of motion can be regarded as having been derived as soon as
the mass matrix, the stiffness matrix and the force vector have been derived.

Example 5.10
Calculate the first 15 natural frequencies and mode shapes for the plane
frame shown in Fig. 5.28 where E = 207 GPa , ρ = 7810 kg m 3 , I = 271 mm 4
and A = 80.6 mm 2 for all beams. The frame is 606.9 mm wide and 606.9 mm
high and is made up of two vertical columns and two equi-spaced cross beams.
Solution. Due to symmetry, only half of the frame can be considered, with
appropriate (separate) constraints for symmetric and antisymmetric modes.
232 MECHANICAL VIBRATIONS

Each half-plane was modelled with 16 identical planar beam elements, 8


for the column and 4 for each half-cross beam.

Fig. 5.28

The lowest 15 natural frequencies, in rad sec , calculated using the


MATLAB computer program VIBFRAME, have the following values: 107.20,
377.47, 397.25, 475.73, 1099.3, 1316.2, 1504.0, 1911.6, 2061.4, 2447.5, 2695.0,
2903.7, 4171.1, 4618.3 and 4943.6.
The mode shapes are shown in Fig. 5.28.

Example 5.11
Calculate the first 15 natural frequencies and mode shapes of the planar
vibrations for the frame shown in Fig. 5.29 where E = 210 GPa ,
ρ = 7850 kg m 3 , I = 1.055 ⋅ 10 −7 m 4 , A = 3.73 ⋅ 10 −4 m 2 and l = 0.5 m .

Fig. 5.29
5. SEVERAL DEGREES OF FREEDOM 233

Fig. 5.30

Solution. Each segment of length l was modelled by 5 identical beam


elements, resulting in 85 elements for the complete frame, hence 86 nodes. With
four fixed nodes and three degrees of freedom per node, the condensed system
matrices were of order 246. The mode shapes calculated using the program
234 MECHANICAL VIBRATIONS

VIBFRAME are presented in Fig. 5.30 together with the rounded values of the
respective natural frequencies.

5.4 Grillages

Grids or grillages are planar structural systems subjected to loads applied


normally to their plane. They are special cases of tree-dimensional frames in which
each joint has only three nodal displacements, a translation and two rotations,
describing bending and torsional effects.

Fig. 5.31

5.4.1 Finite Element Discretization

The grid is divided into elements, as shown in Fig. 5.31. Each node has
three degrees of freedom, two rotations and a linear displacement. Typically, the
degrees of freedom of node i are Q3 i − 2 , Q3 i −1 and Q3 i , defined as the rotation
about the X axis, the rotation about the Y axis and the displacement along the Z
axis, respectively.
Nodes are located by their coordinates in the global reference frame XOY
and element connectivity is defined by the indices of the end nodes. Elements are
modelled as uniform bars with bending and torsional flexibility, without shear
deformations and not loaded between ends. Their properties are the flexural rigidity
E I , the torsional rigidity G I t , the mass per unit length ρ A and the length l .
Only cross sections whose shear centre coincides with the centroid are considered.
5. SEVERAL DEGREES OF FREEDOM 235

5.4.2 Element Stiffness and Mass Matrices in Local Coordinates

Consider an inclined grid element, as illustrated in Fig. 5.32, a, where the


nodal displacements are also shown.

Fig. 5.32

In a local physical coordinate system, the x axis, oriented along the beam,
is inclined an angle α with respect to the global X axis. The z axis for the local
coordinate system is collinear with the Z axis for the global system. Alternatively,
an intrinsic (natural) coordinate system can be used.
The vector of element nodal displacements is

{q }= {q , q
e
1 2, q3, q4, q5, q6 }T (5.112)

and the corresponding vector of element nodal forces can be written

{ f }= { f , f
e
1 2, f 3, f 4, f 5, f 6 }T . (5.113)

In (5.113) f 3 and f 6 are transverse forces, while f 2 and f 5 are


couples producing bending (Fig. 5.32, b). The corresponding displacements
q 3 , q 6 are translations while q 2 , q 5 are rotations. Their column vectors are
related by the flexural stiffness matrix.

Rearranging the matrix (5.89) we obtain

⎧ f2 ⎫ ⎡ 4l 2 6l 2l 2 − 6l ⎤ ⎧q 2 ⎫
⎪ ⎢ ⎥ ⎪ ⎪
⎪ f 3 ⎪⎪ EI e ⎢ 6l 12 6l − 12 ⎥ ⎪q 3 ⎪
⎨ ⎬ = ⎨ ⎬. (5.114)
⎪ f 5 ⎪ l3e ⎢ 2l 2 6l 4l 2 − 6l ⎥ ⎪q 5 ⎪
⎢ ⎥
⎪⎩ f 6 ⎪⎭ ⎢⎣− 6l − 12 − 6l 12 ⎥⎦ e ⎪⎩q 6 ⎪⎭
236 MECHANICAL VIBRATIONS

Likewise, the above nodal forces are related to the respective nodal
accelerations by the mass matrix. Rearranging the matrix (5.95) we can write

⎧ f2 ⎫ ⎡ 4l 2 22l − 3l 2 13l ⎤ ⎧q&& 2 ⎫


⎪ ⎢ ⎥ ⎪ ⎪
⎪ f 3 ⎪⎪ ρ Ae l e ⎢ 22l 156 − 13l 54 ⎥ ⎪ q&& 3 ⎪
⎨ ⎬= ⎨ ⎬. (5.115)
⎪ f5 ⎪ 420 ⎢ − 3l 2 − 13l 4l 2 − 22l ⎥ ⎪q&& 5 ⎪
⎢ ⎥
⎪⎩ f6 ⎭ ⎪ ⎢⎣ 13l 54 − 22l 156 ⎥⎦ e ⎪⎩q&& 6 ⎪⎭

The axial nodal forces f 1 , f 4 are torques and the nodal displacements
q 1 , q 4 are twist angles. They describe torsional effects so that their action is
decoupled from bending. The respective stiffness and mass matrices can be
calculated separately. The derivation of these matrices is essentially identical to the
derivation of the stiffness and mass matrices for axial effects in a frame element or
in a truss element.
The twist angle can be expressed in terms of the shape functions (5.24) as
θ (r ) = N1 (r ) q 1 + N 2 (r ) q 4
2
⎛ ∂θ ⎞

G It e
which substituted into the strain energy U e = ⎜⎜ ⎟⎟ d x yields, after the
2 ⎝ ∂x ⎠
e
+1

change of coordinates, the stiffness matrix [ ]


kte =
2 G It e
le ∫ ⎣N ′ ⎦ r
T
⎣N r′ ⎦ d r .
−1
As a consequence of this analogy, the nodal forces are related to the nodal
displacements by equation
⎧ f1 ⎫ e ⎧ q1 ⎫
⎨ ⎬ = kt ⎨ ⎬ [ ] (5.116)
⎩ f4 ⎭ ⎩ q4 ⎭
where the stiffness matrix for torsional effects is

[ k ] = GlI
t
e te ⎡ 1 − 1⎤
⎢− 1 1 ⎥ . (5.117)
e ⎣ ⎦
In (5.117), G is the shear modulus of elasticity and I t e is the torsional
constant of the cross section. For axially symmetrical cross sections the latter is the
polar second moment of area.
Similarly, the element consistent mass matrix due to torsion is

[ m ] = ρ I6 l
e
t
te e ⎡2 1 ⎤
⎢1 2 ⎥ . (5.118)
⎣ ⎦
5. SEVERAL DEGREES OF FREEDOM 237

For the grid element, combining the stiffness matrices from equations
(5.114) and (5.117), we get the stiffness matrix in local coordinates relating the
nodal forces (5.113) and the nodal displacements (5.112)

⎡ a 0 0 −a 0 0 ⎤
⎢ 0 4l 2 6l 0 2l 2
− 6l ⎥⎥

[k ]
e EI ⎢ 0
= 3e ⎢
l e ⎢− a
6l
0
12
0
0
a
6l − 12 ⎥
0 0 ⎥
⎥ (5.119)

⎢ 0 2l 2 6l 0 4l 2 − 6l ⎥
⎢ ⎥
⎢⎣ 0 − 6l − 12 0 − 6l 12 ⎥⎦ e

where a = G I t e l 2e E I e .

Combining the mass matrices from equations (5.115) and (5.118) we


obtain the consistent mass matrix for the grid element in local coordinates

⎡2b 0 0 0 ⎤
b 0
⎢0 4l 2 22l 13l ⎥⎥
0 − 3l 2

[m ]e
=
ρ Ae l e ⎢ 0 22l
420 ⎢ b

0
156
0
0
2b 0
54 ⎥
0 ⎥

− 13l
(5.120)

⎢ 0 − 3l 2 − 13l 0 4l 2 − 22l ⎥
⎢ ⎥
⎣⎢ 0 13l 54 0 − 22l 156 ⎦⎥ e

where b = 70 I t e Ae .

5.4.3 Coordinate Transformation

It is necessary to transform the matrices (5.119) and (5.120) from the local
to the global system of coordinates before their assemblage in the corresponding
matrices for the complete grid. As has been indicated, the z direction for local axes
coincides with the Z direction for the global axes, so that only the rotational
components of displacements should be converted. The transformation of
coordinates is defined by equation (5.104)

{ q } = [ T ] { Q },
e e e
(5.121)

where q e{ } is the element displacement vector (5.112) in the local coordinate


system,

{ Q }= { Q , Q
e
1 2, Q3, Q4 , Q5, Q6 }T
238 MECHANICAL VIBRATIONS

is the element displacement vector in the global coordinate system (Fig. 5.32) and

⎡c s 0 0 0 0⎤
⎢− s
⎢ c 0 0 0 0⎥⎥

[T ]
e
⎢0
=⎢
0 1 0 0 0⎥
⎥, (5.122)
⎢0 0 0 c s 0⎥
⎢0 0 0 − s c 0⎥
⎢ ⎥
⎣⎢ 0 0 0 0 0 1⎦⎥
where c = cos α and s = sin α , is the local-to-global coordinate transformation
matrix (5.105). The same transformation matrix (5.122) serves to transform the
nodal forces from local to global coordinates.

5.4.4 Element Stiffness and Mass Matrices in Global Coordinates

Using the same procedure as in § 5.2.4 and § 5.4.5, we obtain the stiffness
and mass matrices of the grid element in global coordinates as

[ K ]= [T ] [ k ][T ] ,
e e T e e
(5.123)
and

[ M ]= [T ] [ m ][T ] .
e e T e e
(5.124)
They are used to assemble the unreduced global stiffness and mass
[ ]
matrices, [ K ] and [M ] , using element connectivity matrices T e that relate the
~
nodal displacements at element level with the nodal displacements at the complete
structure level by equations of the form (5.108).
For grounded systems the unreduced matrices [ K ] and [ M ] are then
condensed using the boundary conditions. The effect of lumped masses and springs
can be accounted for by adding their values along the main diagonal at the
appropriate locations in the respective matrices.

Example 5.12
Calculate the first 9 natural frequencies and mode shapes of the plane grid
shown in Fig. 5.33, supported by four springs of stiffness k = 1000 N m each. The
system properties are E = 210 GPa , G = 81 GPa , ρ = 7900 kg m 3 , l = 0.5 m
and the diameter is d = 20 mm for all beams.
5. SEVERAL DEGREES OF FREEDOM 239

Fig. 5.33

Solution. The grid was modelled with 10 elements and 8 nodes, having
thus 24 degrees of freedom. The mode shapes calculated using the MATLAB
program VIBGRID are presented in Fig. 5.34 together with the rounded values of
the respective natural frequencies.

Fig. 5.34

The first three modes of vibration are ‘rigid body’ modes, of rocking and
bouncing of the undeformed grid on the suspension springs. Mode 4 is called ‘first
bending’ (two transverse nodal lines), mode 5 is ‘first torsion’ (one longitudinal
nodal line), mode 6 is ‘second bending’ (three transverse nodal lines), mode 7 is
‘second torsion’, mode 8 is ‘third torsion’ and mode 9 is ‘third bending’ (four
transverse nodal lines).
240 MECHANICAL VIBRATIONS

Example 5.13
The grid shown in Fig. 5.35 is fixed at points 1 and 2, and has l = 1 m ,
I = 0.785 ⋅ 10 −8 m 4 , I t = 1.57 ⋅ 10 −8 m 4 , A = 3.14 ⋅ 10 −4 m 2 , ρ = 7900 kg m 3 ,
E = 210 GPa and G = 81 GPa . Calculate the first 9 natural frequencies and mode
shapes of vibration.

Fig. 5.35

Solution. The grid was modelled with 14 elements and 8 nodes, having 18
dof’s. The mode shapes calculated using the program VIBGRID are presented in
Fig. 5.36 together with the rounded values of the respective natural frequencies.

Fig. 5.36
5. SEVERAL DEGREES OF FREEDOM 241

Mode 1 is ‘first bending’, mode 2 is ‘first torsion’, mode 3 is ‘second


bending’, mode 4 is ‘second torsion’, mode 5 is ‘third bending’ and mode 6 is a
sort of a ‘first bending’ about the longitudinal axis.

5.5 Frequency Response Functions

The preceding sections were restricted to the calculation of normal modes


of vibration for low order undamped vibrating systems. It is of interest to study the
frequency response of damped systems and to illustrate displays of their frequency
response functions (FRFs). The analysis of experimentally obtained FRFs is
presented in a next chapter devoted to Experimental Modal Analysis and
Identification of Vibrating Systems.

5.5.1 Frequency Response Function Matrix

For a viscously damped system, the Frequency Response Function (FRF)


matrix is defined as (4.143)

[ H (iω )] = [ω 2 [ m ] + iω [c ] + [ k ] ]
−1
(5.125)
where ω is the driving frequency.
The FRF matrix can be written

⎡ h 11 (iω ) h 12 (iω ) . . . h1n (iω ) ⎤


⎢ h (iω ) h (iω ) . . . h 2n (iω ) ⎥⎥
[ H (iω )] = ⎢⎢ 21. . . 22
... ... ... ⎥
(5.126)
⎢ ⎥
⎢⎣ h n1 (iω ) h n 2 (iω ) . . . h nn (iω ) ⎥⎦

where h j l (iω ) is the response measured at the coordinate j produced by a


harmonic force applied at coordinate l .
There are several frequency response functions, depending on whether the
response is a displacement, velocity or an acceleration (see § 2.4.7 ). In complex
notation, if the excitation f l ( t ) = f̂ l eiω t produces the response
i (ω t +φ )
x j ( t ) = x̂ j e , then the complex receptance can be expressed in terms of its
real and imaginary components

h j l (iω ) = = l eiφ = l cosφ + i l sinφ = h Rjl (ω ) + i h Ij l (ω ) .


fl f̂ f̂ f̂
x j x̂ j x̂ j x̂ j
242 MECHANICAL VIBRATIONS

The ratio f l xl is called a direct receptance (driving point receptance),


while the ratio f l x j ( j ≠ l ) is termed a transfer receptance or a cross receptance
depending on whether j is a coordinate at a different location or at the same point
as l but in another direction.
Note that FRFs are measurable quantities that can be obtained by
sinusoidal test techniques.

5.5.2 Frequency Response Function Plots

FRF data are complex and thus there are three quantities – the frequency
plus the real and imaginary parts (or modulus and phase) of the complex function –
to be displayed.

Fig. 5.36

The three most common graphical formats are: a) the Bode plots (a pair of
plots of modulus versus frequency and phase versus frequency); b) the Co-Quad
plots (real component versus frequency and imaginary component versus
frequency); and c) the Nyquist plots (imaginary versus real part, eventually with
marks at equal frequency increments).
5. SEVERAL DEGREES OF FREEDOM 243

Plotted quantities include motion/force-type functions like receptance,


mobility and inertance (accelerance), as well as their inverses, dynamic stiffness,
impedance and dynamic mass (see § 2.4.7 ). Scales can be linear, logarithmic and
semilogarithmic, with or without gridlines, with one or two cursors, zoom
capabilities and overlay facilities.

Fig. 5.37

Figure 5.36 shows Bode plots for the receptance FRF of a three dof
damped structure, with the displacement response measured at dofs 1, 2, and 3, due
to forcing applied at dof 1. Note the logarithmic vertical scale for the magnitude
plots, necessary to reveal the details at the lower levels of the response exhibited at
antiresonances.
Figure 5.37 shows the respective plots of the real (coincident) and
imaginary (quadrature) components of receptance, while the first column in Fig.
5.38 shows the corresponding Nyquist plots. It is seen that for lightly-damped
systems with relatively well-separated natural frequencies, the response near
resonances can be approximated by circular loops.
For comparison, mobility and inertance plots are presented in the second
and third columns of Fig. 5.38, for the same FRFs. Note their rotation by 90
244 MECHANICAL VIBRATIONS

degrees in the complex plane, due to the phase relationship between displacement,
velocity and acceleration.

Fig. 5.38

For the point receptance h11 , a three-dimensional plot is illustrated in Fig.


5.39, together with the companion coincident, quadrature, and Nyquist plots.

Fig. 5.39
5. SEVERAL DEGREES OF FREEDOM 245

For lightly damped systems with low modal density, peaks in the FRF
plots of the magnitude of either modulus or the quadrature component of
receptance, mobility or accelerance indicate resonances. A system with n degrees
of freedom, having underdamped modes of vibration, will exhibit n resonances.
Visual inspection of FRF plots helps in estimating the system order, by counting
the resonance peaks within the frequency range of interest.
As shown for two-degree-of-freedom systems (Fig. 4.40), FRF curves for
systems with high damping and/or close natural frequencies may not exhibit all
resonances as peaks. For systems having moderate damping, best resonance
location is achieved on Nyquist plots, at the points where the rate of change of
phase angle (and the arc length) with respect to frequency has a local maximum. If
marks are made on such a plot at equal frequency increments, resonance is located
at maximum spacing between successive points (as in Fig. 4.39).
Antiresonances appear as deep troughs in FRF modulus plots of the type
motion/force. In a driving point FRF, resonances and antiresonances alternate
(Foster’s reactance theorem). Between any two resonance peaks there is a deep
antiresonance through. In a transfer FRF, some or all antiresonances are replaced
by minima (shallow throughs). The number and location of antiresonances change
in FRFs plotted for different response points and directions.
Figure 5.40 is an overlay of three receptance curves calculated for three
different directions at the same point of a structure. It illustrates the coincidence of
resonance peaks and the different number and location of antiresonance troughs.
The upper curve, where resonances and antiresonances alternate, is a drive-point
receptance.

Fig. 5.40
246 MECHANICAL VIBRATIONS

Figure 5.41 illustrates the frequency response diagrams of the receptance


h 3,15 calculated for the 15-dof system of Example 5.8, shown in Fig. 5.17.

a b

c d
Fig. 5.41
Comparing the frequencies of the resonance peaks with the natural
frequencies given in Table 5.2, the cluster of five higher frequencies is first seen,
despite the low level of the FRF magnitudes, due to the logarithmic scale (Fig.
5.41, a). From the other ten resonances, only nine can be distinguished, the close
modes at 53.35 Hz and 53.42 Hz giving a single peak, while the resonances at
59.45 Hz and 73.72 Hz being almost smeared.

a b
Fig. 5.42
5. SEVERAL DEGREES OF FREEDOM 247

The appearance of resonance peaks is dependent on the frequency


resolution of response data, but for a different selection of the response
measurement and excitation points, the function h11 in Fig. 5.42, a and the
function h 6 6 in Fig. 5.42, b, some peaks are missing.

Exercises
5.E1 Find the natural frequencies and mode shapes of the system shown in
Fig. 5.E1. Let k = 106 N m , m 1 = m 4 = m 7 = 0.5 kg , m 2 = m 5 = m 9 = 1.5 kg ,
m 3 = m 6 = m 8 = 1 kg .

Fig. 5.E1
Answer: f 1 = 73.5 Hz , f 2 = 124.62 Hz , f 3 = 174.78 Hz , f 9 = 433.46 Hz .

5.E2 Calculate the natural frequencies and mode shapes of the system
shown in Fig. 5.E2, considering m j = 1 kg ( j = 1, ...,12) , k j = 100 N m
( j = 1,...,11) , k 12 = 102 N m . Compare the results with those obtained for
k 12 = 100 N m .

Fig. 5.E2
Answer: a) f 2 = 0.8238 Hz , f 3 = 0.8285 Hz , f 4 = 1.5915 Hz , f 5 = 1.5941 Hz .
248 MECHANICAL VIBRATIONS

b) f 2 = f 3 = 0.8238 Hz , f 4 = f5 = 1.5915 Hz , f 6 = f 7 = 2.2508 Hz .

5.E3 For the 10-dof system from figure 5.E3 find the damped natural
frequencies and the modal damping ratios. Calculate the undamped natural
frequencies and plot the corresponding mode shapes. System parameters:
m 1 = 1.21 , m 2 = 1.44 , m 3 = 1.69 , m 4 = 1.96 , m 5 = 2.25 , m 6 = 2.56 ,
m 7 = 2.89 , m 8 = 3.24 , m 9 = 3.61 , m 10 = 4 kg , k 1 = 1100 , k 2 = 1200 ,
k 3 = 1300 , k 4 = 1400 , k 5 = 1500 , k 6 = 1600 , k 7 = 1700 , k 8 = 1800 , k 9 = 1900 ,
k10 = 2000 N m , c j = 0.002 k j ( j = 1,...,10) .

Fig. 5.E3
Answer: f 1 = 0.718 Hz , ζ 1 = 0.45% , f 10 = 8.34Hz , ζ 10 = 5.24% .

5.E4 For the 15-dof system from figure 5.E4 find the damped natural
frequencies and the modal damping ratios.

Fig. 5.E4

Answer: f 1 = 2.65 Hz , ζ 1 = 0.81% , f 15 = 156.73Hz , ζ 15 = 0.44% .


5. SEVERAL DEGREES OF FREEDOM 249

5.E5 A ladder-type structure is fixed at both ends as shown in Fig. 5.E5.


The mass of each step is m and the lateral stiffness of each section between steps is
k. Determine the natural frequencies and plot the mode shapes.

Answer: ω 1 = 0.517 k m , ω 2 = k m , ω 3 =1.414 k m , ω 4 =1.732 k m ,

Fig. 5.E5 Fig. 5.E6

5.E6 A four-cylinder engine with flywheel is used to drive an electrical


generator. For the prediction of torsional natural frequencies, the combined system
is represented by the equivalent system shown in Fig. 5.E6. Determine the first four
natural modes of vibration and plot the mode shapes using the following values:
K 12 = K 23 = K 34 = 2.2 MN m rad , K 45 = 0.9 MN m rad , K 56 = 0.24 MN m rad ,
J 1 = J 2 = J 3 = J 4 = 0.51 kg m 2 , J 5 = 5.3 kg m 2 , J 6 = 4.4 kg m 2 .

Answer: f 2 = 46.21 Hz , f 3 = 106.11 Hz , f 4 = 291.41 Hz , f 5 = 481.45 Hz .

5.E7 An eight-cylinder diesel engine with flywheel, transmission gear and


ship propeller is represented by the torsional system shown in Fig. 5.E7. Determine
the first three eigenmodes of torsional vibrations. Use n A = 405 rpm ,
nB = 225 rpm , J 9 = 602 kg m 2 , J 10 = 91 kg m 2 , J 11 = 1953 kg m 2 ,
J 12 = 4331 kg m 2 , J 1 = J 2 = . . . = J 8 = 184 kg m 2 , K 89 = 116 MN m rad ,
K 9 ,10 = 13.5 MN m rad , K 10 ,11 = 185 MN m rad (reduced to nB ),
K 11,12 = 0.846 MN m rad , K 12 = K 23 = ... = K 78 = 84 MN m rad (G. Ziegler,
1977).
250 MECHANICAL VIBRATIONS

Fig. 5.E7
Answer: f 2 = 2.68 Hz , f 3 = 20.60 Hz , f 4 = 35.81 Hz .

5.E8 Determine the two lowest (non-zero) natural frequencies of the


torsional system shown in Fig. 5.E8, for the following values of the moments of
inertia and torsional stiffnesses: K 1 = 2 MN m rad , K 2 = 1.6 MN m rad ,
K 3 = 1 MN m rad , K 4 = 4 MN m rad , J 1 = 15 kg m 2 , J 2 = 10 kg m 2 ,
J 3 = 18 kg m 2 , J 4 = 6 kg m 2 . The speed ratio of drive shaft to axle is 4 to 1.

Fig. 5.E8

Answer: f 2 = 55.26 Hz , f 3 = 62.32 Hz .

5.E9 The gear-and-shaft system for driving the two paddles of a


commercial feed mixer is illustrated in Fig. 5.E9. Each bevel gear has 20 teeth
(20T), two of the spur gears have 36 teeth (36T) and the common spur gear has 12
teeth (12T). Determine the natural frequencies of torsional vibrations. Use
J 1 = 22.6 ⋅ 10 −3 kg m 2 , J 2 = J 3 = 5.65 ⋅10 −3 kg m 2 , J 4 = 11.3 ⋅ 10 −3 kg m 2 ,
J 5 = J 6 = 56.5 ⋅ 10 −3 kg m 2 , J 7 = J 8 = 90.4 ⋅ 10 −3 kg m 2 , l 1 = l 2 = 0.915 m ,
l 3 = 1.016 m , G = 81 GPa , d = 19 mm (James et al., 1989).
5. SEVERAL DEGREES OF FREEDOM 251

Fig. 5.E9

Answer: f 2 = 16.90 Hz , f 3 = 18.41 Hz , f 4 = 37.39 Hz , f 5 = 79.61 Hz .

5.E10 Calculate the natural frequencies and the mode shapes for the
flexural vibrations of the massless beams shown in Fig. 5.E10.

Fig. 5.E10

Answer: a) The flexibility matrix is


⎡ 32 45 40 23 ⎤

m l 3 ⎢ 45 72 68 40 ⎥⎥
[δ ] = .
3750 E I ⎢ 40 68 72 45 ⎥
⎢ ⎥
⎣ 23 40 45 32 ⎦

The natural frequencies are ω 1 = 4.4126 E I m l3 ,


ω 2 =17.611 E I m l 3 , ω 3 = 38.986 E I m l 3 , ω 4 = 64.1997 E I m l 3 .

b) ω 1 = 0.647 E I m l 3 , ω 2 = 2.482 E I m l 3 , ω 3 = 5.65 E I m l 3 .


252 MECHANICAL VIBRATIONS

5.E11 Determine the first eight natural modes of vibration of the plane
trusses from Fig. 5.E11. The pin-ended bars have cross-section Φ 6 x 1 mm and
l = 0.3 m , E = 72.7 GPa , ρ = 3100 kg m 3 . For all lumped masses m = 0.6 kg .

Fig. 5.E11
Answer: a) 20.51, 83.31, 111.49, 164.82, 215.10, 265.74, 299.27, 361.48 Hz.
b) 20.02, 79.07, 117.04, 163.05, 231.06, 272.94, 290.24, 358.46 Hz.

5.E12 For the pin-jointed framework from Fig.5.E12 determine the lowest
six natural frequencies and the corresponding mode shapes, using E = 210 GPa ,
ρ = 7800 kg m 3 , l = 2 m , A = 2000 mm 2 .

Fig. 5.E12
Answer: f 1 = 74.61 Hz , f 2 = 142.24 Hz , f 3 = 237.22 Hz , f 6 = 529.15 Hz .

5.E13 Determine the lowest 20 eigenmodes of the continuous four-span


beam from Fig. 5.E13. Comment the clustering of natural frequencies. Consider
E = 208 GPa , ρ = 7850 kg m 3 , l = 1 m and rectangular cross-section with
b = 2h = 40 mm .

Fig. 5.E13
5. SEVERAL DEGREES OF FREEDOM 253

Answer: 46.69, 54.48, 72.98, 94.33 Hz, 187.48, 203.59, 237.82, 274.84 Hz,
427.85, 452.63, 504.23, 557.66 Hz, 829.07, . . . , 1043.98 Hz.

5.E14 A two-span simply supported beam (Fig. 5.E14) has a rotational


spring at the middle support. For convenience, its stiffness is defined as
K = 500 EI l , where E I is the beam flexural rigidity and l is the total beam
length. a) Calculate the mode shapes of the flexural vibrations of the beam with
equal span lengths (Δ = 0 ) . Notice the clustering of natural frequencies. b)
Consider a disordered weakly coupled beam, taking the span lengths equal to
l (0.5 − Δ ) and l (0.5 + Δ ) , where Δ = 0.01 . Explain the mode localization in
disordered mechanical structures. c) Extend the analysis to disordered three-span
and four-span beams.

Fig. 5.E14

5.E15 A simply supported uniform beam has a length 3 m , mass per unit
length 15 kg m , and bending rigidity 4 kNm 2 with E = 200 GPa . Represent the
beam by four finite elements and calculate the first four natural frequencies
assuming the mass to be: a) concentrated at the mid point of each element; b) the
mass of each element placed half at each end point; c) distributed along the beam
(consistent mass matrix). Compare the results against those found by Rayleigh’s
method, using a “sine” function for the deflection curve, and by Southwell’s
formula. Plot the mode shapes for Case (c).
254 MECHANICAL VIBRATIONS

5.E16 Using the finite element method, find estimates for the lowest two
natural frequencies for bending vibrations of a uniform cantilevered beam. Model it
by 1, 2, 3, 4, 5 and 6 equal-length elements and compare the frequencies so
obtained with the exact values.

5.E17 A uniform steel shaft of diameter d = 40 mm and length l = 1.2 m


is supported by two bearings of stiffnesses k 1 = 4 N μ m and k 2 = 6 N μ m , as
shown in Fig. 5.E17. The two discs have masses m 1 = 400 kg and m 2 = 600 kg ,
and radii of gyration r1 = 0.3 m and r 2 = 0.5 m , respectively. Determine: a) the
natural frequencies of the flexural vibrations, and b) the first four mode shapes,
taking E = 210 GPa and ρ = 7850 kg m 3 .

Fig. 5.E17
Answer:

5.E18 The ends of the solid-steel roll shown in Fig. 5.E18 are simply
supported in short bearings. a) Estimate the first natural frequency of flexural
vibrations using Rayleigh’s method. b) Check the result against that find using a
lumped parameter model. c) Estimate the first two natural frequencies using the
program VIBFRAME, taking E = 210 GPa and ρ = 7850 kg m 3 .
5. SEVERAL DEGREES OF FREEDOM 255

Fig. 5.E18
Answer: f 1 = 36.76 Hz , f 2 = 120.43 Hz .

5.E19 Determine the first two natural frequencies for the folded beam
shown in Fig. 5.E19 and plot the mode shapes. a) Consider AB = CD = l and
BC = 0 . Model the two beams by at least three finite elements. b) Consider
AB = CD = l and BC = l 10 and compare the natural frequencies with those
determined at (a).

Fig. 5.E19
Answer: a) f 1 = f 2 = 11.71 Hz . b) f 1 = 10.45 Hz , f 2 = 10.69 Hz .

5.E20 Find the first eight natural frequencies and mode shapes for the in-
plane vibrations of the plane frame shown in Fig. 5.E20 using the following values:
A vert = 0.006 m 2 , A hor = 0.004 m 2 , A diag = 0.003 m 2 , I = 0.0756 m 4 . Material
properties: E = 75 GPa and ρ = 2800 kg m 3 .

Fig. 5.E20
256 MECHANICAL VIBRATIONS

Answer: Using 48 elements and 44 nodes (4 elements for a beam), the natural
frequencies are 45.15, 79.07, 227.72, 249.94, 365.62, 444.03, 452.83, 476.83 Hz.

5.E21 Determine the 12 lowest eigenmodes for the double cross shown in
Fig. 5.E21. Each arm is of length l = 50 m and has a square cross-section of edge
0.125 m . All the outer endpoints are simply-supported and the inner endpoints are
rigidly jointed together. The material properties are E = 200 GPa and
ρ = 8000 kg m3 . Use more than two beam elements to discretize each arm.

Fig. 5.E21
Answer: f 1 = 5.6687 Hz , f 2 = f 3 = 8.8485 Hz , f 4 = ... = f8 = 8.8856 Hz .

5.E22 Consider the system from Fig. 5.E22 as a horizontal grid and
determine the lowest four natural modes of vibration. Take l = 0.25 m , l1 = 0.2 m ,
E = 210 GPa , ν = 0.3 and ρ = 7850 kg m 3 . All beams have a square cross-
section with the edge a = 5 mm . I t = 0.141 a 4 .
5. SEVERAL DEGREES OF FREEDOM 257

Fig. 5.E22

Answer: f 1 = 10.69 Hz , f 2 = 27.31 Hz , f 3 = 44.96 Hz , f 4 = 108.65 Hz .

5.E23 Determine the first six modes of coupled bending-torsion vibrations


of the grillage from Fig. 5.E23. Consider l = 50 mm , E = 210 GPa , G = 80 GPa ,
ρ = 7800 kg m3 and circular cross-section with d = 5 mm .

Fig. 5.E23
258 MECHANICAL VIBRATIONS

Answer: The first six mode shapes are illustrated below.

5.E24 The grillage shown in Fig. 5.E24 is comprised as follows: a) a


horizontal steel shaft ABC of flexural rigidity 2.5 ⋅ 107 Nm 2 , torsional rigidity
1.92 ⋅ 10 7 Nm 2 and mass per unit length 300 kg m ; b) two horizontal steel shafts
EBD and FCG, rigidly attached to ABC, and having a flexural rigidity 5 ⋅ 107 Nm 2
and mass per unit length 500 kg m ; c) masses of 500 kg at D and E, and masses
of 250 kg at F and G. You may neglect the moments of inertia of these masses.
Consider E = 210 GPa , G = 81 GPa , ρ = 7850 kg m 3 . Find the first six natural
modes of vibration.

Fig. 5.E24

Answer: 2.35, 10.39, 13.39, 36.16, 63.25, 105.22 Hz.


6.
CONTINUOUS SYSTEMS

When the mass and elasticity are distributed throughout the vibratory
system, a description of its configuration requires an infinite number of
coordinates. Such a system may be regarded as having an infinite number of
degrees of freedom, hence an infinite number of natural frequencies and mode
shapes.
In this chapter the vibrations of bars with distributed mass and elasticity
are considered, assuming them to be composed of homogeneous, isotropic and
elastic materials. The treatment is restricted to only a few particular cases useful as
benchmark examples for the problems solved using the computer programs
mentioned in the previous chapter.

6.1 Lateral Vibrations of Thin Beams

In the analysis which follows we shall adopt the assumptions usual in


simple bending theory, that is, the Bernoulli-Euler beam theory: a) the beam is
initially straight; b) the depth of the beam is small compared to its radius of
curvature at its maximum displacement; c) plane sections remain plane at all
phases of an oscillation; d) one principal axis of inertia is perpendicular to the
direction of motion in the vibration; e) the beam is free from longitudinal force; f)
the deformation due to shearing of one cross section relative to an adjacent one is
negligible; g) the beam mass is concentrated at its neutral axis. The last two imply
that the effects of transverse shear and rotatory inertia are neglected.

6.1.1 Differential Equation of Motion

The beam lateral deflection is a function of both space and time,


v = v (x, t ) . The relationship between the bending moment and the curvature (5.65)
becomes
260 MECHANICAL VIBRATIONS

∂2 v
M (x, t ) = E I z , (6.1)
∂x2
where E is Young’s modulus for the material of the beam and I z is the second
moment of area of the cross-section about the principal axis which lies in the
neutral plane.
From the equilibrium of an infinitesimal beam element (Fig. 6.1), the
relationships between bending moment, shear force and external load are obtained.

Fig. 6.1

The shear force is given by


∂M ∂3 v
T (x, t ) = = EIz . (6.2)
∂x ∂x3
The transverse load per unit length is
∂T ∂4 v
p (x, t ) = = EIz . (6.3)
∂x ∂x4

It consists of two components, the applied external transverse load q (x , t )


and the inertial transverse load (mass per unit length times acceleration)
∂2 v
− ρ A 2 . Substituting
∂t
∂2 v
p (x, t ) = q (x , t ) − ρ A
∂ t2
in equation (6.3) we obtain the differential equation of motion

∂4 v ∂2v
EIz + ρ A = q (x ,t ) . (6.4)
∂ x4 ∂ t2
6. CONTINUOUS SYSTEMS 261

6.1.2 Natural Frequencies and Mode Shapes

For free vibrations q ( x ,t ) = 0 . We look for conditions under which v (x, t )


is a synchronous harmonic motion
v (x, t ) = V (x ) sin (ω t + φ ) . (6.5)
Substituting the solution (6.5) in equation (6.4), we obtain for q = 0

d 4V ρA
4
−ω2 V =0. (6.6)
dx EIz
Equation (6.6) is of fourth order so the general solution will contain four
constants
V ( x ) = B 1 sin α x + B 2 cos α x + B 3 sinh α x + B 4 cosh α x (6.7)

where
ρA
α 4 = ω2 . (6.8)
EIz
The four constants are determined from the end conditions. The standard
end conditions are: a) simply supported or pinned end, for which the displacement
is zero and the bending moment is zero as there is no rotational constraint; b) fixed
or clamped end, for which the displacement and slope are zero; c) free end, for
which the bending moment and shear force are zero. In terms of the function V (x )
these conditions for a uniform beam are:
a) Simply supported end:

V = 0 and d 2 V d x 2 = 0 . (6.9)
b) Clamped end:
V = 0 and d V d x = 0 . (6.10)
c) Free end:

d 2 V d x 2 = 0 and d 3V d x 3 = 0 . (6.11)
Imposing two conditions at each end we obtain a set of four homogeneous
equations. A nontrivial solution is obtained by setting the determinant of their
coefficients equal to zero. Expansion of the determinant and subsequent reduction
gives a transcendental equation in the quantity α l whose solutions yield the
natural frequencies

ωn = (α n l ) 2
EIz
. n = 1, 2, 3, . . . (6.12)
ρ A l4
262 MECHANICAL VIBRATIONS

Since equation (6.6) is homogeneous, each mode shape function Vn (x ) can


describe only relative displacements of the various parts of the beam, and not
absolute values. As for discrete systems, the process of assigning absolute values to
the function Vn (x ) is known as normalization, and it is one of multiplying it by a
suitable constant. The selection of this constant is arbitrary. We may, for example,
assign a unit value to the displacement at a certain location. For the graphical
display of several mode shapes it is useful to assign a unit value to the maximum
displacement in any mode. Another method is to normalize the mode to unit modal
mass.
The natural frequencies and mode shapes will be determined in the
following for some particular end conditions.

Fig. 6.2

Both ends simply supported


For a simply supported beam (Fig. 6.2), the end conditions are V = 0 and
d V d x 2 = 0 at x = 0 and at x = l . Substituting the conditions at x = 0 in
2

equation (6.7) we obtain


0 = B 2 + B 4 and 0 = − B 2 + B 4 .

Thus B 2 = B 4 = 0 .

Applying the conditions at x = l , we obtain


0 = B 1 sin α l + B 3 sinh α l and 0 = − B1 sin α l + B 3 sinh α l .

The only non-trivial solution is B 3 = 0 and sin α l = 0 , leading to

α l = nπ , n = 1, 2, 3, . . . (6.13)
The first four roots of equation (6.13) are

α nl = 3.142, 6.283, 9.425, 12.566 . n = 1, 2, 3, 4 .


6. CONTINUOUS SYSTEMS 263

Thus for a uniform beam with simply supported ends the natural
frequencies are given by
4
⎛ nπ ⎞ E I z
ωn2 = ⎜ ⎟ . n = 1, 2, 3, . . . (6.14)
⎝ l ⎠ ρA
and the modes of vibration are given by

v ( x, t ) = C sin
nπ x
l
(
sin ω n t + φ n , )
where C is an arbitrary constant.
The first four modes of vibration, corresponding to n = 1 , 2, 3, and 4, and
the location of the respective nodal points are given in Fig. 6.2.

Fig. 6.3

Cantilever
With the origin at the fixed end, as in Fig. 6.3, and using equations (6.10)
and (6.11), the end conditions are the following:

At x = 0 , V = 0 , 0 = B2 + B4 .

At x = 0 , d V d x = 0 . 0 = B1 + B 3 .

At x = l , d 2 V d x 2 = 0 .

0 = − B1 sin α l − B 2 cosα l + B 3 sinh α l + B 4 cosh α l .

At x = l , d 3V d x 3 = 0 .

0 = − B1 cos α l + B 2 sin α l + B 3 cosh α l + B 4 sinh α l .

The above four linear equations have a non-trivial solution if the following
determinant vanishes
264 MECHANICAL VIBRATIONS

0 1 0 1
1 0 1 0
=0
− sin α l − cos α l sinh α l cosh α l
− cos α l sin α l cosh α l sinh α l

which leads to the frequency equation


cos α l ⋅ cosh α l = −1 . (6.15)
The first five roots of equation (6.15) are
α n l = 1.875, 4.694, 7.855, 10.996, 14.137 . n = 1, 2, 3, 4, 5 . (6.16)
For n greater than five the values are given approximately by
2 n −1
α nl = π. (6.16, a)
2
Substituting B 4 = − B 2 and B 3 = − B 1 in the last two end conditions
yields
B 1 (sin α l + sinh α l ) + B 2 (cos α l + cosh α l ) = 0,
B 1 (cos α l + cosh α l ) − B 2 (sin α l − sinh α l ) = 0.

Equating to zero the determinant of the above equations yields again the
frequency equation (6.15). Expressing B 4 , B 3 and B 2 in terms of B 1 = C , we
obtain the shape of the nth mode
Vn ( x ) = C [ cosh α x − cos α x − kn ( sinh α x − sin α x ) ] (6.17)
where
cosh α l + cos α l
kn = , n = 1, 2, 3, . . . (6.18)
sinh α l + sin α l
The first four modes of vibration and the location of the respective nodal
points are given in Fig. 6.3.

Both ends free


For a free-free beam (Fig. 6.4), the end conditions are
V ′′ (0) = 0 , V ′′′ (0) = 0 , V ′′ (l ) = 0 , V ′′′ (l ) = 0 . (6.19)
Following the above procedure and using the appropriate end conditions, it
can be shown that the frequency equation is
cos α l ⋅ cosh α l = +1 . (6.20)
The first four roots of equation (6.20) are
6. CONTINUOUS SYSTEMS 265

α n l = 4.730, 7.853, 10.996, 14.137 . n = 1, 2, 3, 4 (6.21, a)


For n greater than four the values are given approximately by
2 n −1
α nl = π. (6.21, b)
2

Fig. 6.4

The mode shapes are given by


Vn ( x ) = C [ sinh α x + sin α x − kn ( cosh α x + cos α x ) ] (6.22)
where
sinh α l − sin α l
kn = , n = 1, 2, 3, . . .
cosh α l − cos α l
The first four modes of vibration and the location of the respective nodal
points are given in Fig. 6.4.
Equation (6.6) admits other two solutions which satisfy the end conditions
(6.19), namely
V (x ) = B5 = const., V (x ) = B 6 x + B7 .

These are the so-called “rigid body” modes of vibration. The former
corresponds to the vertical bouncing, the latter to the rocking in the vertical plane,
both having zero natural frequency.

Both ends clamped


For a beam with clamped ends (Fig. 6.5), the end conditions are
V (0) = 0 , V ′ (0) = 0 , V (l ) = 0 , V ′ (l ) = 0 .
The frequency equation is the same as (6.19) and so are the natural
frequencies given by α n l = 4.730, 7.853, 10.996, 14.137 , for n = 1, 2, 3, 4 , and
2 n −1
α nl = π for n greater than four.
2
266 MECHANICAL VIBRATIONS

Fig. 6.5

The mode shapes are given by


Vn ( x ) = C [ cos hα x − cosα x − kn ( sinh α x − sin α x ) ] (6.23)
where
cosh α l − cos α l
kn = , n = 1, 2, 3, . . . (6.24)
sinh α l − sin α l
The first four modes of vibration and the location of the respective nodal
points are given in Fig. 6.5.

Fig. 6.6

One end clamped and one end pinned


When the beam is clamped at one end and simply supported at the other
end (Fig. 6.6), the end conditions are
V (0) = 0 , V ′ (0) = 0 , V (l ) = 0 , V ′′ (l ) = 0 .
The frequency equation is
tanh α l = tan α l . (6.25)
The first four roots of equation (6.25) are
α n l = 3.927, 7.069, 10.210, 13.352 . n = 1, 2, 3, 4 .
For n greater than four the values are given approximately by
6. CONTINUOUS SYSTEMS 267

4n +1
α nl = π. (6.26)
4
The mode shapes are given by
Vn ( x ) = C [ cosh α x − cos α x − kn ( sinh α x − sin α x ) ] (6.27)
where
cosh α l − cos α l
kn = , n = 1, 2, 3, . . . (6.28)
sinh α l − sin α l
The first four modes of vibration and the location of the respective nodal
points are given in Fig. 6.6.

6.1.3 Orthogonality of Mode Shape Functions

As for all undamped vibrating systems, the mode shape functions of beams
with distributed mass and elasticity are orthogonal. In the general case of beams of
variable cross-section, they are orthogonal with respect to a weighting function
m (x ) , where m (x ) = ρ A( x ) is the mass per unit length. Let the bending rigidity be
E I (x ) .
Considering the modes of indices r and s, respectively, we can write

(E I Vr′′)" −ωr2 m Vr = 0 , (6.29)

(E I Vs′′)" −ω s2 m Vs = 0 . (6.30)

Multiplying equations (6.29) and (6.30) by Vs and Vr , respectively, and


integrating from 0 to l yields
l l

∫ (E I Vr′′)" Vs dx = ω r ∫ Vr Vs m dx ,
2
(6.31)
0 0

l l

∫ (E I Vs′′)" Vr dx = ω 2s ∫ Vs Vr m dx . (6.32)
0 0

Subtracting equation (6.32) from (6.31) gives

( ) ∫ V V m dx = ∫ [ (E I V ′′)" V
l l
ω 2r − ω 2s s r r s − (E I Vs′′)" Vr ] dx . (6.33)
0 0

Integrating the right side of (6.33) by parts results in


268 MECHANICAL VIBRATIONS

(ω ) ∫ V V m dx = { [V (E I V ′′)' −V (E I V ′′)' ] − E I (V ′V ′′ − V ′V ′′ ) } .
l
l
2
r − ω 2s s r s r r s s r r s 0 (6.34)
0

The right side of the equation (6.34) vanishes if, at each end of the beam, at
least one of the following pairs of boundary conditions is prescribed:
V =0 and V ′ = 0,
V =0 and E I V ′′ = 0 ,
(6.35)
V′ = 0 and (E I V ′′)' = 0,
E I V ′′ = 0 and (E I V ′′)' = 0.
Assuming that one of the above pairs of boundary conditions is applied at
each end of the beam, equation (6.34) reduces to
l

∫ m (x )Vr (x ) Vs (x )dx = 0 , r ≠s. (6.36)


0

Equations of the type (6.36) are known as the orthogonality relations for
the natural mode shapes of the beam. It is said that the functions Vr (x ) and Vs ( x )
are orthogonal (with each other) with respect to the weighting function m (x ) .
For a uniform beam, equation (6.36) becomes
l

∫ Vr (x ) Vs (x )dx = 0 , r ≠s. (6.36, a)


0

Also, for a beam with any combination of the above end conditions
(simply-supported, clamped and free)
l

∫ E I (x ) Vr′′(x ) Vs′′ (x )dx = 0 , r ≠s.


0

Equations (6.36) can be used to determine the free response for given
initial conditions.

6.1.4 Multi-Span Beams

In finding the natural frequencies of a beam on multiple supports, the


section between each pair of supports is considered as a separate beam with its
origin at left support of the section. Equation (6.7) applies to each section.
At each end of the beam the usual boundary conditions are applicable,
depending on the type of support. At each intermediate support the deflection is
6. CONTINUOUS SYSTEMS 269

zero. Since the beam is continuous, the slope and the bending moment just to the
left and to the right of the support are the same.

Example 6.1
Find the frequency equation for a two span beam with the extreme ends
simply supported and unequal span lengths.
Solution. The equation (6.7) is written for the first segment of length l 1

( )
V1 x 1 = A1 sin α x 1 + B 1 cos α x 1 + C 1 sinh α x 1 + D 1 cosh α x 1

and for the second segment of length l 2

V2 ( x2 ) = A 2 sin α x2 + B 2 cos α x2 + C 2 sinh α x2 + D 2 cosh α x2 .

The boundary conditions

V1 (0) = V1′′ (0) = 0 , V1 (l1 ) = V2 (0) = 0 , V1′ (l1 ) = V2′ (0) ,


V1′′ (l1 ) = V2′′ (0) , V2 (l 2 ) = V2′′ (l 2 ) = 0
yield the following five homogeneous algebraic equations

A1 sin α l 1 + C 1 sinh α l 1 = 0 ,
A1 cos α l 1 + C 1 cosh α l 1 − A 2 − C 2 = 0 ,
− A1 sin α l 1 + C 1 sinh α l 1 − 2 D2 = 0 ,
A 2 sin α l 2 + C 2 sinh α l 2 + D 2 ( cosh α l 2 − cos α l 2 ) = 0 ,
− A 2 sin α l 2 + C 2 cosh α l 2 + D 2 ( cosh α l 2 + cos α l 2 ) = 0.

The above set of linear equations has non-trivial solutions if the


determinant of the coefficients of A1 , C 1 , A 2 , C 2 and D 2 is equal to zero. This
condition yields the frequency equation

sinh μ α l ⋅ sinh ( 1 − μ )α l ⋅ sin α l − sin μ α l ⋅ sin (1 − μ )α l ⋅ sinh α l = 0 ,

where μ = l1 l and l = l1 + l 2 .

Table 6.1 gives the values of α nl in equation (6.12) for finding the first
five natural frequencies of uniform continuous beams on uniformly spaced
supports.
The mode shapes for the first cluster of natural frequencies plus the first
frequency in the next cluster are shown in Fig. 6.7 for beams with two, three, four
and five equal spans, respectively.
270 MECHANICAL VIBRATIONS

Fig. 6.7
6. CONTINUOUS SYSTEMS 271

Table 6.1
Beam structure Roots of characteristic equation
Nr. of Mode index
spans 1 2 3 4 5
1 3.142 6.283 9.425 12.566 15.708
2 3.142 3.927 6.283 7.069 9.425
3 3.142 3.550 4.304 6.283 6.692
4 3.142 3.393 3.927 4.461 6.283
5 3.142 3.299 3.707 4.147 4.555
6 3.142 3.267 3.550 3.927 4.304
7 3.142 3.236 3.456 3.770 4.084
8 3.142 3.205 3.393 3.644 3.927
9 3.142 3.205 3.330 3.550 3.801
10 3.142 3.205 3.299 3.487 3.707
11 3.142 3.173 3.267 3.424 3.613
12 3.142 3.173 3.267 3.393 3.550

6.1.5 Response to Harmonic Excitation

When a uniform beam is subjected to harmonic disturbing forces or


couples, they can be accounted for through the boundary conditions, as shown in
the following for a free-free beam.

Fig. 6.8

Symmetrical loading
Consider a uniform free-free beam subjected to a harmonic force
F 0 sin ω t acting at the middle (Fig. 6.8).

Assuming a steady-state solution of the form

v (x, t ) = Y (x ) sin ω t (6.37)


where ω is the excitation frequency, equation (6.4) yields
272 MECHANICAL VIBRATIONS

Y ( x ) = C 1 sin α x + C 2 cos α x + C 3 sinh α x + C 4 cosh α x (6.38)

where
ρA
α 4 = ω2 . (6.39)
EIz
The boundary conditions for half of the beam are

Y ′′ (0) = 0 , Y ′′′ (0) = 0 , Y ′ (l 2) = 0 , E I Y ′′′ (l 2) = F 0 2 . (6.40)

The amplitude of the forced vibrations at a location 0 ≤ x ≤ l 2 is

αl αl
C
F0 sinh − sin
Y (x ) = 2 2 (sin α x + sinh α x ) −
4 E I α 3 cosh αl sin αl + cos αl sinh αl
2 2 2 2
(6.41)
αl αl
(cos α x + cosh α x )G
cos + cosh
− 2 2
αl αl αl αl
cosh sin + cos sinh
2 2 2 2

The dynamic deflection at the beam midpoint is

αl αl
3 3 1 + cos cosh
⎛ l ⎞ F0 l ⎛ 2 ⎞ 2 2
Y⎜ ⎟= ⎜ ⎟ . (6.42)
⎝ 2 ⎠ 16 E I ⎝ αl ⎠ cosh αl sin αl + cos αl sinh αl
2 2 2 2

Equating to zero the denominator of equation (6.42) we obtain the


resonance frequencies. This yields
αl αl
tanh + tan = 0. (6.43)
2 2
The first three roots of equation (6.43) are

α n l 2 = 2.365, 5.498, 8.639 . n = 1, 2, 3 .


The values

(α l )1s = 4.730 , (α l ) 2 s = 10.996 , (α l ) 3s = 17.278 , (6.44)

are the roots of equation (6.20) corresponding to the symmetrical modes of


vibration.
6. CONTINUOUS SYSTEMS 273

Equating to zero the numerator of equation (6.42) we obtain the


antiresonance frequencies. This yields
αl αl
1 + cos cosh = 0, (6.45)
2 2
whose first three roots are

α n l 2 = 1.875, 4.694, 7.855 n = 1, 2, 3 .


or
(α l ) (1 )= 3.750 , (α l ) ( 2 )= 9.388 , (α l ) ( 3 )= 15.710 .

Fig. 6.9

The driving point receptance curve for the free-free beam from Fig. 6.8 is
illustrated in Fig. 6.9 using a logarithmic vertical scale.

Anti-symmetrical loading

Consider a uniform free-free beam subjected to a harmonic couple


M 0 sin ω t acting at the middle (Fig. 6.10).

Fig. 6.10
274 MECHANICAL VIBRATIONS

The lateral deflection is given by (6.38). The boundary conditions for half
of the beam are
Y ′′ (0) = 0 , Y ′′′ (0) = 0 , Y ′ (l 2) = 0 , E I Y ′′ (l 2) = − M 0 2 . (6.46)

The amplitude of the forced vibrations at a location 0 ≤ x ≤ l 2 is

αl αl
C
M0 cos + cosh
Y (x ) = 2 2 (sin α x + sinh α x ) −
4EI α 2 αl αl αl αl
cosh − cos sinh
sin
22 2 2
(6.47)
αl αl
(cos α x + cosh α x )G
sin + sinh
− 2 2
αl αl αl αl
cosh sin − cos sinh
2 2 2 2
Equating to zero the denominator of equation (6.47) we obtain the
frequency equation
αl αl
tanh − tan =0. (6.48)
2 2
The first three roots of equation (6.48) are
α n l 2 = 2.365, 5.498, 8.639 . n = 1, 2, 3 .
The values
(α l )1a = 3.927 , (α l ) 2 a = 7.069 , (α l ) 3a = 10.21 , (6.49)

are the roots of equation (6.20) corresponding to the antisymmetrical modes of


vibration.
When the beam is acted upon by a harmonic force F 0 sin ω t applied at an
arbitrary location, the equivalent loads applied at the beam middle are a force and a
couple, and the forced response may be obtained by summation of solutions (6.41)
and (6.47). The resonance frequencies are given by (6.12)

ωn = (α n l ) 2
EIz
n = 1, 2, 3, . . .
ρ A l4

where α n l is given by (6.44) and (6.49). In can be seen that

α 1s = α 1 , α 1a = α 2 , α 2 s = α 3 , α 2 a = α 4 , α 3s = α 5 , α 3a = α 6 , . . .
This is explained by writing equation (6.20) under the form
6. CONTINUOUS SYSTEMS 275

αl αl
1 − tan 2 1 + tanh 2
2 ⋅ 2 =1
α l α l
1 + tan 2 1 − tanh 2
2 2
which becomes

⎛ αl αl ⎞ ⎛ αl αl ⎞
⎜ tan + tanh ⎟ ⎜ tan − tanh ⎟ =0.
⎝ 2 2 ⎠⎝ 2 2 ⎠
Equating each factor to zero yields the frequency equations (6.43) and
(6.48).

6.2 Longitudinal Vibration of Rods

In a thin uniform rod of cross-section area A and mass per unit volume ρ ,
there are axial displacements u = u (x , t ) due to axial forces. In Fig. 6.11 a free-
body diagram of an infinitesimal element of length dx of the rod is shown.

Fig. 6.11

The equation of motion in the x-direction is


⎛ ∂N ⎞ ∂ 2u
⎜⎜ N + d x ⎟⎟ − N − (ρ Ad x ) 2 = 0 ,
⎝ ∂x ⎠ ∂t
∂N ∂ 2u
= ρA 2 , (6.50)
∂x ∂t
where u is the displacement at x.
276 MECHANICAL VIBRATIONS

The displacement at x + d x will be u + (∂u ∂ x ) d x . The axial strain is


ε x = ∂ u ∂ x , the axial stress is σ x = E ∂u ∂ x , where E is Young’s modulus of
elasticity. The strain can be expressed in terms of the axial force N as
∂u N
= . (6.51)
∂x E A
Combining equations (6.50) and (6.51) yields the differential equation of
free vibration
∂ 2 u ρ ∂ 2u
= . (6.52)
∂ x2 E ∂ t 2
The constant c = E ρ is the velocity of propagation of longitudinal
waves along the rod.
A harmonic synchronous motion is defined by a solution of the form
u (x ,t ) = U (x ) sin (ω t + φ ) (6.53)
which substituted in equation (6.52) yields
d 2U
+ β 2U = 0 (6.54)
d x2
where
ρ
β 2 = ω2 . (6.55)
E
The general solution of equation (6.52) is

U (x ) = C 1 sin β x + C 2 cos β x . (6.56)

The two integration constants C 1 and C 2 may be determined from the end
conditions. At a fixed end the displacement is zero, u = 0 . At a free end, the stress
and hence the strain is zero, ∂u ∂ x = 0 . The effect of end springs or concentrated
masses can also be accounted for.

Free-free rod
For a rod with both ends free ∂u ∂ x = 0 at x = 0 and x = l . We obtain
C 1 = 0 and, since C 2 must be nonzero in order to have vibration, the frequency
equation (characteristic equation) becomes
sin β l = 0 . (6.57)
Its solution consists of an infinite set of eigenvalues
6. CONTINUOUS SYSTEMS 277

β n l = nπ , n = 1, 2, 3, . . .
The natural frequencies (eigenfrequencies) are thus given by
π E
ωn = n , n = 1, 2, 3, . . . (6.58)
l ρ
and the mode shapes (eigenfunctions) are defined by

U n ( x ) = C cos nπ
x
,
l
where C is an undetermined amplitude.
The dynamic response of a continuous rod can be regarded as a
superposition of motions in the eigenmodes. These are synchronous harmonic
motions at the respective natural frequency with all points vibrating in phase (or
out-of-phase) with the relative spatial distribution of amplitudes defined by the
eigenfunction. In practical applications, the infinite series is truncated to the modes
with natural frequencies within the operating frequency range of the structure.

Dynamic stiffness matrix


It is of interest to derive the exact dynamic stiffness matrix for an axially
vibrating bar segment (a two-noded pin-jointed element) in local coordinates and to
compare it with the matrices obtained using static shape functions, (5.32) and
(5.38).
Denoting u 1 , u 2 , the dynamic nodal displacements and f 1 , f 2 , the
dynamic nodal forces (positive in the x-direction), as in Section 5.2.1, the end
conditions are
f1 f2
U (0) = u 1 , U (l ) = u 2 , U ′(0) = − , U ′(l ) = . (6.59)
EA EA
Using equation (6.56) we obtain

⎧u 1 ⎫ ⎡ 0 1 ⎤ ⎧C 1 ⎫ ⎧ f1⎫ ⎡ −1 0 ⎤ ⎧C 1 ⎫
⎨u ⎬ = ⎢ ⎥ ⎨ ⎬, ⎨ f ⎬ = EAβ ⎢cos β l − sin β l ⎥ ⎨C ⎬ ,
⎩ 2 ⎭ ⎣sin β l cos β l ⎦ ⎩C 2 ⎭ ⎩ 2⎭ ⎣ ⎦ ⎩ 2⎭

and, for sin β l ≠ 0 ,

⎧ f1⎫ ⎡ cotan β l − cosec β l ⎤ ⎧ u 1 ⎫


⎨ f ⎬ = EAβ ⎢− cosec β l cotan β l ⎥ ⎨u ⎬ ,
⎩ 2⎭ ⎣ ⎦ ⎩ 2⎭

so that the dynamic stiffness matrix is


278 MECHANICAL VIBRATIONS

[ k ] = EAβ ⎡⎢−cotan
e
dyn
βl
cosec β l
− cosec β l ⎤
cotan β l ⎥⎦
. (6.60)

The Taylor’s series expansions up to three terms, of the coefficients of the
dynamic matrix (6.60), are

EA ρ Al l 3 (ρ A)2
EAβ cotan β l = −ω2 −ω4 − .... ,
l 3 45 E A

EA ρ Al 7l 3 (ρ A)2
− EAβ cosec β l = − −ω2 − ω4 − .... .
l 6 360 E A
It may be seen that, in each of the above series, the first term is equal to the
corresponding stiffness coefficient of the matrix (5.32) and the second term is
( )
equal to − ω 2 times the coefficient of the consistent mass matrix (5.38), both
calculated based on static shape functions.
If the frequency dependent stiffness matrix is

[ k ] = ElA ⎡⎢−11
e − 1⎤
⎥ + ω 4 (ρ Al ) 2
l ⎡ 1 7 8⎤
45 E A ⎢⎣7 8 1 ⎥⎦
⎣ 1⎦
and the frequency-dependent consistent mass matrix is

[ m ] = ρ 6Al ⎡⎢ 12
e 1⎤
⎥ + ω 2 (ρ Al ) 2
2l ⎡ 1 7 8⎤
⎢7 8 1 ⎥⎦
⎣ 2⎦ 45 E A ⎣
it can be seen that the third term in each of the above series comes from the
frequency-dependent stiffness and mass coefficients.

6.3 Torsional Vibration of Rods

The equation of motion of a rod in torsional vibration is similar to that of


longitudinal vibration of rods presented in the preceding section. It can be written
∂ 2θ ρ ∂ 2θ
. = (6.61)
∂ x2 G ∂ t 2
where θ is the twist at x and G is the shear modulus of elasticity.
In order to calculate the natural modes of vibration, a harmonic solution is
looked for, of the form
θ (x ,t ) = Θ (x ) sin (ω t + φ ) . (6.62)
6. CONTINUOUS SYSTEMS 279

Upon substitution in equation (6.61) we obtain


d 2Θ
2
+ γ 2Θ = 0 (6.63)
dx
where
ρ
γ 2 = ω2 . (6.64)
G
The general solution of equation (6.63) is

Θ (x ) = C 1 sin γ x + C 2 cos γ x . (6.65)

The integration constants C 1 and C 2 may be determined from the end


conditions. At a fixed end the rotation angle is zero, θ = 0 . At a free end,
∂θ ∂ x = 0 (as the angle of twist due to a torque M is dθ = M dx G I t ).
The effect of end springs, concentrated discs, as well as concentrated
forces and torques can also be accounted for.

Rod with one end fixed and the other end free

At x = 0 , θ = 0 , and at x = l , ∂θ ∂ x = 0 . We obtain C 2 = 0 and the


frequency equation
cos γ l = 0 . (6.66)
Equation (6.66) has an infinite set of solutions
2n + 1
γ nl = π, n = 0,1, 2, 3, . . .
2
The natural frequencies are thus given by
2n + 1 π G
ωn = , n = 0,1, 2, 3, . . . (6.67)
2 l ρ
and the mode shapes are defined by
2n + 1 x
Θ n (x ) = C sin π
2 l
where C is an undetermined amplitude.

Dynamic stiffness matrix

By analogy with axial vibrations, the dynamic stiffness matrix of a twisted


rod segment is
280 MECHANICAL VIBRATIONS

[k ] = G I γ ⎡⎢−cotan
e
dyn
γl
cosec γ l
t
− cosec γ l ⎤
cotan γ l ⎥⎦
. (6.68)

where γ is given by (6.64) and G I t is the torsional rigidity of the beam cross-
section. This matrix relates the nodal dynamic axial torques with the nodal
dynamic twist angles.
Using Taylor series expansions as in Section 6.2, the following frequency-
dependent stiffness and mass matrices are obtained

[ k ] = GlI
e t ⎡ 1 − 1⎤ 4 l
⎢− 1 1 ⎥ + ω (ρ I t l ) 45G I
2 ⎡ 1 7 8⎤
⎢7 8 1 ⎥⎦

⎣ ⎦ t ⎣

[ m ] = ρ 6I l ⎡⎢ 12
e t 1⎤
⎥ + ω 2 (ρ I t l ) 2
2l ⎡ 1 7 8⎤
⎢7 8 1 ⎥⎦
.
⎣ 2⎦ 45G I t ⎣

6.4 Timoshenko Beams

The results obtained in section 6.1 apply to slender beams, for which rotary
inertia and deformation due to shear is neglected. The analysis of beam vibration
including these effects, usually known as the Timoshenko beam theory, is presented
in a next chapter. These effects lower the natural frequencies compared to the
values obtained from equation (6.6).
For a uniform beam with both ends simply supported, the values given by
the approximate equation (6.14) have an error of less than 5% provided
n r l < 0.08 , where r = I A is the radius of gyration and n is the mode number.
A feature of Timoshenko beams is the existence of two distinct natural
frequencies for each mode shape taken up by the centre line of the beam. For the
lower of each pair of frequencies, deformations due to bending and shear are in
phase with each other and are summed to give the total lateral deflection. In the
higher frequency case, the bending and shear deformations are out-of-phase, with
the net transverse deflection equal to their difference.
In the Timoshenko theory, the basic assumption is that plane-cross-sections
remain plane throughout vibration, but they are not perpendicular to the beam
centre line, but rotated by a shear angle which is considered constant in all points
(neglecting warping due to shear deformation). This angle is calculated using an
effective area obtained multiplying the actual cross-section area by a shear factor
which is a function of Poisson’s ratio and cross section geometry.
References

1. Ananyev, I. V., Handbook for Calculation of the Natural Vibrations of Elastic


Systems, Gostekhizdat, Moscow, 1946 (in Russian).
2. Andronov, A. A., Witt, A. A. and Chaikin, S. E., Theorie der Schwingungen,
Akademie Verlag, Berlin, 1965.
3. Babakov, I. M., Theory of Vibrations, Nauka, Moscow, 1965 (in Russian).
4. Babitsky, V. I., Theory of Vibro-Impact Systems and Applications, Springer,
Berlin, 1998.
5. Beards, C. F., Engineering Vibration Analysis with Application to Control
Systems, Edward Arnold, London, 1995.
6. Benaroya, H., Mechanical Vibration, Analysis, Uncertainties and Control,
Prentice Hall International, Upper Saddle River, NJ, 1998.
7. Bendat, J. S. and Piersol, A. G., Random Data: Analysis, Measurement and
Procedures, Wiley Interscience, New York, 1986.
8. Beranek, L. L., Noise and Vibration Control, McGraw-Hill, New York, 1971.
9. Biderman, V. A., Applied Theory of Mechanical Vibrations, Vyshaya Shkola,
Moscow, 1972 (in Russian).
10. Biezeno, C. B. and Grammel, R., Technische Dynamik, Springer, Berlin, 1939.
11. Bishop, R. E. D., Vibration, Cambridge University Press, 1965.
12. Bishop, R. E. D., Gladwell, G. M. L. and Michaelson, S., The Matrix Analysis
of Vibration, Cambridge Univ. Press, Cambridge, 1965.
13. Bishop, R. E. D. and Johnson, D., The Mechanics of Vibration, Cambridge
Univ. Press, 1960.
14. Bisplinghoff, R., Ashley, H. and Halfman, R., Aeroelasticity, Addison-Wesley,
Cambridge, 1955.
15. Blake, M. P. and Mitchell, W. S., Vibration and Acoustic Measurement
Handbook, Spartan, New York, 1972.
16. Blevins, R.D., Formulas for Natural Frequency and Mode Shape, Van
Nostrand Reinhold, New York, 1979.
282 MECHANICAL VIBRATIONS

17. Bogoliubov, N. N. and Mitropolsky, Yu. A., Asymptotic Methods in the Theory
of Nonlinear Oscillations, 3rd ed., Fizmatghiz, Moscow, 1963 (in Russian).
18. Braun, S., Rao, S. S. and Ewins, D. J. (eds), Encyclopedia of Vibration,
Academic Press, London, 2002.
19. Broch, J. T., Mechanical Vibration and Shock Measurements, 2nd ed., Brüel &
Kjær, Nærum, 1980.
20. Buzdugan, Gh., Fetcu, L. and Radeş, M., Vibration of Mechanical Systems,
Editura Academiei, Bucureşti, 1975 (in Romanian).
21. Buzdugan, Gh., Fetcu, L. and Radeş, M., Mechanical Vibrations, 2nd ed.,
Editura didactică şi pedagogică, Bucureşti, 1982 (in Romanian).
22. Buzdugan, Gh., Mihăilescu E. and Radeş, M., Measurement of Vibrations,
Editura Academiei, Bucureşti, 1979 (in Romanian).
23. Buzdugan, Gh., Mihăilescu, E. and Radeş, M., Vibration Measurement,
Martinus Nijhoff Publ., Dordrecht, 1986.
24. Bykhovsky, I. I., Fundamentals of Vibration Engineering, Mir Publishers,
Moscow, 1972.
25. Church, A.. H., Mechanical Vibrations, Wiley, New York, 1957.
26. Clough, R. W. and Penzien, J., Dynamics of Structures, McGraw Hill, New
York, 1975.
27. Collar, A. R. and Simpson, A., Matrices and Engineering Dynamics, Ellis
Horwood, Chichester, 1987.
28. Crandall, S. H. and Mark, W. D., Random Vibration in Mechanical Systems,
Academic Press, New York, 1963.
29. Den Hartog, J. P., Mechanical Vibrations, Dover, New York, 1985 (First
edition Mc-Graw-Hill, New York, 1934).
30. Dimarogonas, A. D., Vibration Engineering, West Publ. Co., St. Paul, MN,
1976.
31. Dimentberg, F. M., Shatalov, K. T. and Gusarov, A. A., Vibrations of
Machines, Mashinostroyeniye, Moscow, 1964 (in Russian).
32. Dincă, F. and Teodosiu, C., Nonlinear and Random Vibration, Editura tehnică,
Bucureşti, 1969 (in Romanian).
33. Ewins D. J., Modal Testing: Theory, Practice and Application, 2nd ed.,
Research Studies Press, Taunton, 2000.
34. Fertis, D. G., Mechanical and Structural Vibrations, Wiley, New York, 1995.
35. Filippov, A. P., Vibration of Mechanical Systems, Naukova Dumka, Kiev, 1965
(in Russian).
REFERENCES 283

36. Frazer, R. A., Duncan, W. J. and Collar, A. R., Elementary Matrices, AMS
Press, New York, 1983 (Reprint of C.U.P. edition of the original 1938 text).
37. Gasch, R. and Knothe, K., Strukturdynamik, Band 1: Discrete Systeme. Band 2:
Kontinua und ihre Diskretisierung, Springer, Berlin, 1987.
38. Gatti, P. L. and Ferrari, V., Applied Structural and Mechanical Vibrations, E &
FN SPON, London, 1999.
39. Geradin, M. and Rixen, D., Mechanical Vibrations. Theory and Application to
Structural Dynamics, 2nd ed., Wiley, New York, 1997.
40. Hamburger, L. and Buzdugan, Gh., Theory of Vibrations and Its Applications
in Machines, Editura tehnică, Bucureşti, 1958 (in Romanian).
41. Hansen, H. M. and Chenea, P. F., Mechanics of Vibrations, Wiley, N.Y., 1952.
42. Harris, C. M. (ed.), Shock and Vibration Handbook, 4th ed., McGraw-Hill,
New York, 1996 (First edition, 1961).
43. Hayashi, C, Nonlinear Oscillations in Physical Systems, McGraw-Hill, New
York, 1964.
44. Heylen, W., Lammens, S. and Sas, P., Modal Analysis Theory and Testing, K.
U. Leuven, 1997.
45. Hohenemser, K. and Prager, W., Dynamik der Stabwerke, Springer, Berlin,
1933.
46. Holzweissig, F. and Dresig, H., Lehrbuch der Maschinendynamik, 3rd ed.,
Fachbuchverlag, Leipzig, 1992.
47. Hunt, J. B., Dynamic Vibration Absorbers, Mechanical Engineering
Publications, London, 1979.
48. Hurty, W. C. and Rubinstein, M. F., Dynamics of Structures, Prentice-Hall,
Englewood Cliffs, NJ, 1964.
49. Ibrahim, R. A., Parametric Random Vibration, Research Studies Press Ltd.,
Taunton, 1998.
50. Inman, D. J., Engineering Vibration, 2nd ed., Prentice Hall, Upper Saddle
River, NJ, 1994.
51. Inman, D. J., Vibration with Control, Measurement and Stability, Prentice Hall,
Englewood Cliffs, NJ, 1989.
52. Irretier, H., Schwingungstechnik, 2nd ed., University of Kassel, 1988.
53. Jacobsen, L. S. and Ayre, R. S., Engineering Vibrations, McGraw-Hill, New
York, 1958.
284 MECHANICAL VIBRATIONS

54. James, M. L., Smith, G. M., Wolford, J. C. and Whaley, P. W., Vibration of
Mechanical and Structural Systems with Microcomputer Applications, Harper
& Row, New York, 1989.
55. Kauderer, H., Nichtlineare Mechanik, Springer, Berlin, 1958.
56. Kelly S. G., Fundamentals of Mechanical Vibrations, 2nd ed., McGraw-Hill
Higher Education, New York, 2000.
57. Kelly S. G., Schaum’s Outline on Mechanical Vibrations, McGraw-Hill
Engineering CS, New York, 1996.
58. Ker Wilson, W. T., Practical Solution of Torsional Vibration Problems, 2nd
ed., Vol. I-V, Wiley, New York, 1956-1965.
59. Ker Wilson, W. T., Vibration Engineering: A Practical Treatment, Ch. Griffin
& Co., London, 1959.
60. Klotter, K., Technische Schwingungslehre, Band I: Einfache Schwingungen,
Band II: Technische Schwingungslehre, Springer, Berlin, 1960, 1978.
61. Koloušek, V., Dynamik der Baukonstruktionen, VEB Verlag für Bauwesen,
Berlin, 1962.
62. Korenev B. G. and Rabinovitch, I. M. (eds), Structural Dynamics Handbook,
Stroyizdat, Moscow, 1972 (in Russian).
63. Korenev, B. G. and Reznikov, L. M., Dynamic Vibration Absorbers. Theory
and Technical Applications, Wiley, New York, 1993.
64. Krämer, E., Maschinendynamik, Springer, Berlin, 1984.
65. Lalanne M., Berthier, P. and Der Hagopian, J., Mechanical Vibrations for
Engineers, Wiley, New York, 1983.
66. Lancaster, P., Lambda Matrices and Vibrating Systems, Pergamon Press,
Oxford, 1966.
67. Lee, K. Ch., Vibrations of Shells and Rods, Springer, Berlin, 1999.
68. Leissa, A. W., Vibration of Plates, NASA SP-160, 1969.
69. Leissa, A. W., Vibration of Shells, NASA SP-288, 1973.
70. Levy, S. and Wilkinson, J. C., The Component Element Method in Structural
Dynamics, Mc-Graw-Hill, New York, 1976.
71. Lorenz, H., Grundbau Dynamik, Springer, Berlin, 1960.
72. Lutes, L. D. and Sarkani S., Stochastic Analysis of Structural and Mechanical
Vibrations, Prentice Hall, Upper Saddle River, NJ, 1997.
73. Lyon, R. H., Random Noise and Vibration in Space Vehicles, SVIC,
Washington, DC, 1967.
REFERENCES 285

74. Macduff, J. N. and Currieri, J. R., Vibration Control, McGraw-Hill, New York,
1958.
75. Maia, N. M. M. and Silva, J. M. M., Theoretical and Experimental Modal
Analysis, Research Studies Press, Taunton, 1997.
76. Magnus, K., Schwingungen, Teubner, Stuttgart, 1976.
77. Marguerre, K. and Wölfel, H., Mechanics of Vibration, Sijthoff & Noordhoof,
Alphen aan den Rijn, 1979.
78. Mazet, R., Mécanique vibratoire, Librairie Polytechnique Ch. Béranger, Paris,
Liège, 1955.
79. McCallion, H., Vibration of Linear Mechanical Systems, Wiley, New York,
1973.
80. McConnell, K. G., Vibration Testing, Wiley, New York, 1995.
81. McLachlan, N. W., Theory of Vibrations, Dover, New York, 1951.
82. Meirovitch, L., Analytical Methods in Vibrations, Macmillan, New York, 1967.
83. Meirovitch, L., Elements of Vibration Analysis, McGraw-Hill, New York,
1975.
84. Meirovitch, L., Fundamentals of Vibrations, McGraw-Hill International
Edition, Boston, 2001.
85. Meirovitch, L., Principles and Techniques of Vibrations, Prentice-Hall
International, Upper Saddle River, N. J., 1997.
86. Minorsky, N., Nonlinear Oscillations, Van Nostrand, New York, 1962.
87. Mobley, R. K., Vibration Fundamentals, Newnes, Boston, 1999.
88. Morris, J., The Strength of Shafts in Vibration, Crosbie Lockwood & Co.,
London, UK, 1929.
89. Morrow, C. T., Shock and Vibration Engineering, Wiley, New York, 1963.
90. Morse, P. M., Theoretical Acoustics, McGraw-Hill, New York, 1968.
91. Morse, P. M., Vibration and Sound, McGraw-Hill, New York, 1948.
92. Mottershead, J. E. (ed.), Modern Practice in Stress and Vibration Analysis,
Pergamon Press, Oxford, 1989.
93. Müller, P. C. and Schiehlen, W. O., Lineare Schwingungen, Akademische
Verlaggesellschaft, Wiesbaden, 1976.
94. Myklestad, N. O., Fundamentals of Vibration Analysis, McGraw-Hill, New
York, 1956.
286 MECHANICAL VIBRATIONS

95. Nestorides, E. J., A Handbook on Torsional Vibration, Cambridge Univ. Press,


1958.
96. Newland, D. E., Mechanical Vibration Analysis and Computation, Longman
Scientifical & Technical, Harlow, Essex, 1989.
97. Newland, D. E., Random Vibrations and Spectral Analysis, Longman, London,
1975.
98. Norris, C. H., Hansen, R. J., Holley, M. J., Biggs, J. M., Namyet, S. and
Minami, J. K., Structural Design for Dynamic Loads, McGraw-Hill, New York,
1959.
99. Osiński Z. (ed.), Damping of Vibrations, A.A. Balkema, Rotterdam, 1998.
100. Palmov, V., Vibrations of Elasto-Plastic Bodies, Springer, Berlin, 1998.
101. Panovko, Ya. G., Elements of the Applied Theory of Elastic Vibrations, Mir
Publishers, Moscow, 1971.
102. Panovko, Ya. G., Fundamentals of the Applied Theory of Vibration and
Shocks, Mashinostroyeniye, Moscow, 1976 (in Russian).
103. Panovko, Ya. G., Internal Friction During the Vibration of Elastic Systems,
Fizmatghiz, Moscow, 1960 (in Russian).
104. Paz, M., Structural Dynamics. Theory and Computation, Van Nostrand
Reinhold Comp., New York, 1980.
105. Pestel, E. C. and Leckie, F. A., Matrix Methods in Elastomechanics, McGraw-
Hill, New York, 1963.
106. Pilkey, W.D. and Pilkey, B., Shock and Vibration Computer Programs:
Reviews and Summaries, SVIC, Washington, DC, 1975.
107. Prentis, J. M., Dynamics of Mechanical Systems, 2nd ed., Ellis Horwood,
Chichester, 1980.
108. Preumont, A., Vibration Control of Active Structures. An Introduction, 2nd
ed., Kluwer Academic Publ., 2002.
109. Pusey, H. C., Volin, R. H. and Showalter, J. G., An International Survey of
Shock and Vibration Technology, SVIC, Washington, DC, 1979.
110. Radeş, M., Identification of Vibrating Systems, Editura Academiei, Bucureşti,
1979 (in Romanian).
111. Rao, J. S., Dynamics of Plates, Marcel Dekker, New York, 1999.
112. Rao, S. S., Mechanical Vibrations, 3rd ed., Addison-Wesley, 1995.
113. Rayleigh, Lord (Strutt, J. W.), The Theory of Sound, Macmillan, London,
1894 (Reprinted by Dover, 1945).
REFERENCES 287

114. Richart, F. E., Jr., Hall, J. R., Jr. and Woods, R. D., Vibrations of Soils and
Foundations, Prentice-Hall, Englewood Cliffs, NJ, 1970.
115. Robson, J. D., Random Vibration, Edinburgh Univ. Press, UK, 1964.
116. Rocard, Y., Dynamique générale des vibrations, 3rd ed., Masson, Paris, 1960.
117. Roseau, M., Vibrations des systèmes mécaniques, Masson, Paris, 1984.
118. Ruzicka, J. E. and Derby, T. F., Influence of Damping in Vibration Isolation,
SVIC, Washington, DC, 1971.
119. Salter, J. P., Steady-State Vibration, Kenneth Mason, Havant, 1969.
120. Scanlan, R. H. and Rosenbaum, R., Aircraft Vibration and Flutter, Macmillan
Co., New York, 1951.
121. Sevin, E. and Pilkey, W. D., Optimum Shock and Vibration Isolation, SVIC,
Washington, DC, 1971.
122. Seto, W. W., Theory and Problems of Mechanical Vibrations, Schaum, New
York, 1964.
123. Smith J. D., Gear Noise and Vibration, Marcel Dekker, New York, 1999.
124. Snowdon, J. C., Vibration and Shock in Damped Mechanical Systems, Wiley,
New York, 1968.
125. Snowdon, J. C. and Ungar, E. E., Isolation of Mechanical Vibration, Impact
and Noise, AMD Vol.1, ASME, New York, 1973.
126. Sólnes, J., Stochastic Processes and Random Vibrations. Theory and Practice,
Wiley, New York, 1997.
127. Soong, T. and Grigoriu, M., Random Vibration of Mechanical and Structural
Systems, Prentice Hall, Englewood Cliffs, NJ, 1993.
128. Sorokin, E. S., Theory of Internal Damping in Vibrations of Elastic Systems,
Gosstroyizdat, Moscow, 1960 (in Russian).
129. Steinberg, D. S., Vibration Analysis for Electronic Equipment, 3rd ed., Wiley,
New York, 2000.
130. Stoker, J. J., Nonlinear Vibrations in Mechanical and Electrical Systems,
Wiley, New York, 1950.
131. Szemplińska-Stupnicka, W., The Behaviour of Nonlinear Vibrating Systems,
Vol.2, Advanced Concepts and Applications to Multi-Degree-of-Freedom
Systems, Kluwer, 1990.
132. Thomson, W. T., Theory of Vibration with Applications, Prentice-Hall,
Englewood Cliffs, NJ, 1972; 4th ed., Stanley Thornes Publ., Cheltenham, 1998.
133. Thomsen, J. J., Vibrations and Stability. Order and Chaos, McGraw-Hill,
New York, 1997.
288 MECHANICAL VIBRATIONS

134. Timoshenko, S., Young, D. H. and Weaver, W. Jr., Vibration Problems in


Engineering, 4th ed., Wiley, New York, 1974 (First edition, 1928).
135. Tong, K. N., Theory of Mechanical Vibration, Wiley, New York, 1960.
136. Tongue, B. H., Principles of Vibration, Oxford University Press, 2002.
137. Tse, F. S., Morse, I. E. and Hinkle, R. T., Mechanical Vibrations: Theory and
Applications, Allyn and Bacon, Boston, MA, 1963, 2nd ed., 1978.
138. Tuplin, W. A., Torsional Vibrations, 2nd ed., Sir Issac Pitman and Sons,
London, UK, 1966.
139. Ulbrich, H., Maschinendynamik, Teubner, Stuttgart, 1996.
140. Vernon, J. B., Linear Vibration Theory: Generalized Properties and
Numerical Methods, Wiley, New York, 1967.
141. Vierck, R. K., Vibration Analysis, International Textbook Co., Scranton, PA,
1967.
142. Voinea, R. and Voiculescu, D., Mechanical Vibrations, Institutul Politehnic
Bucureşti, 1979 (in Romanian).
143. Voinea, R., Voiculescu, D. and Ceauşu, V., Mechanics, Editura didactică şi
pedagogică, Bucureşti, 1975 (in Romanian).
144. Wagner, K. W., Lehre von den Schwingungen und Wellen, Dietrich Verlag,
Wiesbaden, 1957.
145. Warburton, G. B., The Dynamical Behaviour of Structures, Pergamon Press,
Oxford, 1964.
146. Walker, D. N., Torsional Vibration of Turbomachinery, McGraw-Hill
Engineering Reference, New York, 2004.
147. Wallace, R. H., Understanding and Measuring Vibrations, Wykeham
Publications, London, 1970.
148. Yablonsky, A. A. and Noreyko, S. S., Course on Vibration Theory, 3rd ed.,
Vysshaya Shkola, Moscow, 1975 (in Russian).
149. Yorish Yu. I., Vibration Measurement, Mashghiz, Moscow, 1963 (in
Russian).
150. Yu, Y.-Y., Vibrations of Elastic Plates, Springer, Berlin, 1995.
151. Ziegler, G., Maschinendynamik, Carl Hanser Verlag, München, Wien, 1977.
Index

Absorber, dynamic vibration 117 − critical 36


Accelerance 53, 56 − energy dissipated by 47
Accelerometer 65 − equivalent viscous 51
Additional mass 51 − factor 49
Amplitude 13 − hereditary 68
− complex 53 − hysteretic 48
− distortions 65 − optimum 169
− of forced vibration 24, 124 − quadratic 97
− resonant 45, 49 − ratio 37, 158
Antiresonance 31, 117 − structural 48
Aperiodic motion 39 − velocity squared 97
Apparent mass 54 − viscous 36
Dashpot 36
Base excitation 32
Decrement, logarithmic 39
Beam element 223
Degrees of freedom 6
− influence coefficient 187 Diagram 22
− lumped masses 184 − Bode 55, 242
− multi-span 268
− displacement-force 46
− Rayleigh’s formula 192 − Nyquist 55, 244
Beats 25, 154
Difference equation 202
Bode diagram 55, 242
Discretisation 221
Branched system 128, 197
Distortions 65
Characteristic equation 108 Duhamel’s integral 23
Complex algebra method 52 Dunkerley’s formula 190
Concentrated mass 184 Dynamic absorber 117
Consistent mass matrix 213, 227 − stiffness 54
Continuous systems 259 − stresses 125
Coordinate transformation 213, 229
Eigenfrequencies 108
Coulomb damping 92
Eigenvalues 133
Coupled pendulums 151
Eigenvectors 133
− translation 106 Energy dissipated by damping 47
− translation and rotation 145 − kinetic 17, 213, 227
Coupling 146
− method 17
Critical damping 36
− strain 205, 225, 236
Critical speed 33
Equivalent viscous damping 51
Cubic stiffness 81
Excitation, base 31
D’Alembert’s principle 106, 132 − harmonic 23, 114
Damped systems 156 − mass 23
Damping 10 − with unbalanced rotating masses 30
− coefficient 157 Flexibility coefficients 130, 187
− Coulomb 92
290 MECHANICAL VIBRATIONS

− matrix 23 Mass addition effect 51


Flexural system 130 Mass-spring system 11
Forced vibration, damped 42 Matrix, damping 157
− − undamped 22 − flexibility 131
Frame element 228 − mass 107, 212, 227, 235
Free body diagram 106 − stiffness 107, 212, 225, 235
Free-free beam 21 Mechanical impedance 54
Free vibration 11, 135, 153 Mobility 53, 56
− − damped 35, 157 Modal analysis 114, 165
− − undamped 11, 107 − coordinates 112
Frequency 12 − damping coefficients 158
− damped 37, 159, 174 − mass 113, 135
− equation 108, 264, 266 − matrix 113
− fundamental 192 − stiffness 113
− natural 108 − vectors 109
− peak 45 Mode orthogonality 111
− resonant 45 − natural 108
− response 163 − normal 109
− − curves 26 − shape 109
− − function 163 Multi-span beam 268
− − − matrix 163, 241 Multi-storey frame 193
− undamped 12
Natural coordinates 210
Geared system 126, 196 − frequency 108
Grillage 234 − − closely spaced 176
− − damped 37, 159, 174
Half-power points 49
− − fundamental 19
Harmonic excitation 23, 161
− − undamped 12
− motion 13
Non-linear systems 79
Houdaille damper 168
Non-proportional damping 172
Hysteresis loop 47
Normal coordinates 108
Hysteretic damping 48
− modes 109
Inertance 53 Normalisation 108
Influence coefficients 130, 187 Nyquist diagram 55, 244
Initial conditions 13
Orthogonality 111, 267
Instruments, vibration measuring 62
Oscillations 5
Isochrones 84
Overdamped modes 174
Isolation, vibration 32
− system 39
Jump phenomena 89
Pendulum 151
Kinetic energy 17, 213, 227 Percent of critical damping 37
Period of vibration 13
Lagrange’s equations 152, 219 Phase angle 43
Lateral vibration of beams 259 − distortion 66
Logarithmic decrement 39 − shift 14, 62
Longitudinal vibration of rods 275 Plane frame 220
Loss factor 42 Potential energy 17
Lumped mass system 184
Principal coordinates 113
Magnification factor 26 Proportional damping 157
INDEX 291

Rayleigh’s method 18, 192


− quotient 113
Receptance 53, 241
Repeated structure 202
Residues 58
Resonance 27
− acceleration through 29
− curve 45
− frequency 45
Rigid body modes 123, 265
Rod, longitudinal vibration of 275
− torsional vibration of 278
Rotating shaft 66
Seismic instruments 62
Shape functions 210, 223
Skeleton curve 83
Spring combinations 14
State space equations 173
Stiffness coefficients 13
− matrix 107, 212, 225, 235
Strain energy 205, 212, 225, 236
Structural damping 48
Transfer function 162
Torsional stiffness 16, 120
− system 15, 119, 195
Transmissibility 32, 61
Truss element 210
Undamped natural frequency 12, 176
Underdamped system 37
Vibration absorber 170
− damper 36
− free 8, 11, 135, 153
− forced 8, 22, 42
− parametric 8
Vibrometer 65
Viscous damper, untuned 168
− damping 36
− − non-proportional 172
Whirling shaft 66
Work due to damping 47

You might also like