Dynamic Analysis of Structures PDF
Dynamic Analysis of Structures PDF
Dynamic Analysis of Structures PDF
@CivilMethod
Dynamic Analysis of
Structures
John T. Katsikadelis
Professor Emeritus of Structural Analysis
School of Civil Engineering
National Technical University of Athens
@CivilMethod
Academic Press is an imprint of Elsevier
125 London Wall, London EC2Y 5AS, United Kingdom
525 B Street, Suite 1650, San Diego, CA 92101, United States
50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom
© 2020 Elsevier Inc. All rights reserved.
No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopying, recording, or any information storage and
retrieval system, without permission in writing from the publisher. Details on how to seek
permission, further information about the Publisher’s permissions policies and our
arrangements with organizations such as the Copyright Clearance Center and the Copyright
Licensing Agency, can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing. As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical
treatment may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating
and using any information, methods, compounds, or experiments described herein. In using
such information or methods they should be mindful of their own safety and the safety of others,
including parties for whom they have a professional responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors,
assume any liability for any injury and/or damage to persons or property as a matter of products
liability, negligence or otherwise, or from any use or operation of any methods, products,
instructions, or ideas contained in the material herein.
ISBN 978-0-12-818643-5
@CivilMethod
Dedication
To my wife Efi
for her loving patience and support,
to my children Stefan and Christina,
and to
my granddaughter Katharina
for patiently enduring and sharing the years of
preparation of the book with me.
@CivilMethod
Preface
The statement of the laws of motion by Newton 334 years ago (1686)a was a
milestone in the evolution of mechanics and modern engineering. The relation
between force (cause) and motion (effect) was quantified as a relation between
the linear momentum of the body and the force exerted on it. Thus, this relation
from a subject of philosophy up to that time turned out to be a valuable tool of
science for the study of the natural world. The subsequent developments in the
sciences were rapid. Astronomy, mathematics, mechanics of fluid and deform-
able bodies, and in general, mechanics of continuous media reached their peaks
in the centuries that followed, with immense applications to all engineering dis-
ciplines. Nevertheless, the laws of motion, which were stated as an axiom
(Axiomata sive Leges Motus) by Newton because, apparently, he could not jus-
tify their derivation, was a consequence of the discoveries of great scientists
who preceded him such as Galileo,b Kepler,c Hook, etc.
The implementation of the laws of motion leads to mathematical models
described by differential equations, ordinary or partial, whose solution effort
has given a great impetus to the development of mathematics. Unfortunately,
analytical solutions are limited to simple problems such as vibrations of discrete
systems with a few degrees of freedom; linear vibrations of beams, membranes,
plates, and shells with simple geometry; and simple support conditions made
from materials, mostly with a linear behavior. These solutions, while useful
for extracting qualitative conclusions about the dynamic response of structures,
are not capable of solving realistic problems in engineering, where the geometry
and loads are complicated while the response is generally nonlinear. Although it
has been shown that Newton’s law of motion is of an integer-order derivative,d
in recent years, the fractional derivatives have been proven more suitable for
modeling the actual structures. However, the use of fractional calculus has
not been employed in mathematical physics for three centuries because the
a. I. Newton, Philosophiae Naturalis Principia Mathematica, Royal Society Press, London, 1686.
b. J.T. Katsikadelis, Derivation of Newton’s law of motion using Galileo’s experimental data, Acta
Mech. 226 (2015) 3195–3204, doi:10.1007/s00707-015-1354.
c. J.T. Katsikadelis, Derivation of Newton’s law of motion from Kepler’s laws of planetary motion,
Arch. Appl. Mech. 88 (2018) (2017) 27–38, doi:10.1007/s00419-017-1245-x.
d. J.T. Katsikadelis, Is Newton’s law of motion really of integer differential form? Arch. Appl.
Mech. 89 (2019) 639–647, doi:10.1007/s00419-018-1486-3.
xv
@CivilMethod
xvi Preface
@CivilMethod
Preface xvii
particularly after the advent of computers. Therefore, many books have appeared
on this subject. However, these books provide a means only to understand the
response of simple and mostly unrealistic structures when subjected to dynamic
loads, especially to ground motion. The principles of dynamics are illustrated by
applying them to very simple models, which cannot describe actual structures
and therefore cannot be employed for dynamic analysis and design. We should
have in mind that the dynamic analysis of actual structures requires their model-
ing, the formulation of the governing equations of motion, their solution under
any dynamic load, and the physical interpretation of the results.
In the last 30 years, almost all seismic codes have encountered earthquake
ground motion as an effective dynamic load. The advent of computers in the
early 1960s encouraged engineers to develop methods of dynamic analysis
of structures, modeled first by the FEM and later by other advanced numerical
methods. Today, these methods constitute a powerful tool for dynamic engi-
neering analysis. Thanks to the availability of cheap computer power, every
engineer can use them. The essential ingredients of a book on Dynamics of
Structures for Civil Engineers should be:
(a) The basic concepts and principles of structural dynamics as they are
applied to particles as well as rigid and deformable bodies, enabling the
student or the engineer to formulate the equations of motion of any struc-
ture, no matter how complex, once the dynamic model has been adopted.
(b) Realistic modeling of actual structures under dynamic loads.
(c) Analysis of the dynamic response of the structure represented by its
model under any specified load. The analysis should include single- and
multiple-degree-of-freedom systems for linear and nonlinear response
under any dynamic excitation.
(d) Approximation of real structures using computational methods such as the
FEM, which replaces the actual structure (distributed parameter system)
with an approximate discrete system for which analysis methods can be
applied.
(e) Effective present-day numerical methods for dynamic analysis, including the
numerical solution of eigenvalue problems and the direct solution of the equa-
tions of motion of large systems, namely, systems with a large number of
degrees of freedom such as those resulting from the employed discretization.
Students attending a course on Dynamics of Structures should be exposed at
least to the above subjects. However, not all of them can be found in a single
book. Therefore, people interested in structural dynamics should refer to more
than one book in order to retrieve the required knowledge. Apparently, these
books cannot be used as integrated textbooks in the sense described above.
The student should be acquainted with different symbols and approaches, which
complicate the acquisition of knowledge, an approach that is, at least, educa-
tionally inappropriate.
@CivilMethod
xviii Preface
@CivilMethod
Preface xix
discussed. The derivation of the equivalent element nodal quantities, that is,
mass and stiffness matrices and forces, are derived by using the Lagrange equa-
tions instead of the principle of virtual works. Although the principle of vir-
tual works offers a handy tool for the derivation of these quantities, the use of
the Lagrange equations was preferred here. The reason is that the Lagrange
equations not only offer a straightforward method for the derivation of the
equivalent nodal quantities for all types of elements, especially in complex
systems with a nonlinear response, but they also allow understanding of their
physical significance. Chapter 12 studies the free vibrations of MDOF sys-
tems without and with damping. The linear eigenvalue problem is presented
from the mathematical point of view, aiming at drawing useful conclusions
about the eigenfrequencies and the mode shapes of the physical systems.
Chapter 13 presents the numerical methods for the computation of the eigen-
frequencies and mode shapes, especially for systems with a large number of
degrees of freedom such as those derived from the application of finite ele-
ments. Chapter 14 studies the forced vibrations of MDOF systems. A large
part of this chapter is devoted to the mode superposition method. It also dis-
cusses the use of Ritz vectors to reduce the degrees of freedom. Particular
emphasis is given to the response spectrum method. The response of linear
systems when they are subjected to a synchronous and an asynchronous
motion of the supports are also discussed. This chapter ends with the presen-
tation of the numerical methods, giving the respective pseudocodes for the
time integration of linear and nonlinear systems of equations of motion.
Chapter 15 discusses the approximation of multistory buildings by skeletal
structures and presents methods of formulating their equation of motion.
Finally, Chapter 16 discusses the response of seismically isolated buildings.
This chapter is introductory to the subject and aims primarily at understand-
ing the impact of base isolation on structures.
The book is supplemented by an appendix. Therein, the basic theory of rigid
body dynamics is presented for large and small displacements and the relevant
equations are derived, which are employed in the development of the material of
the book.
In closing, the author wishes to express his sincere thanks to his former stu-
dent and coworker Dr. A. J. Yiotis for carefully reading the manuscript as well
as for his suggestions, constructive recommendations and his overall contribu-
tion to minimizing the oversights in the text. Warm thanks also belong to Dr.
Nikos G. Babouskos, also a former student and coworker of the author, not only
for his careful reading of the manuscript and his apposite suggestions for
improvement of the book but also for his assistance in checking the computer
programs and in producing the numerical results of examples therein. Finally,
thanks belong to Dr. G. Dasios, professor of mathematics at the University of
Patras as well as to his former students, Dr. G. Tsiatas, associate professor of
mathematics at the University of Patras, and Dr. P. Tsopelas, associate professor
of mechanics at NTUA, for reading certain sections of the book and making
constructive suggestions.
@CivilMethod
xx Preface
J.T. Katsikadelis
Athens
April 2020
@CivilMethod
Chapter 1
1.1 Introduction
Apart from static loads, engineering structures may be subjected to dynamic
loads, that is, loads whose magnitude as well as direction of action and/or posi-
tion vary with time. The analysis of stresses and deflections developed in a
given structure undergoing dynamic loads is the fundamental objective of the
dynamic analysis of structures. Between static and dynamic analysis of struc-
tures, there exist two substantial differences:
(a) In static analysis, the loads are assumed time-invariant, and the resulting
response is unique, at least in linear theory. On the other hand, in dynamic
analysis the loads are time-varying and the deformations and stresses
depend on time, that is, at each instant the response of the structure is
different.
(b) In dynamics analysis, the material points of the structure change position
with the time, hence they have velocity and acceleration. Inasmuch as
the structure has a mass, inertial forces are produced due to the accelerations
of the material points. These inertial forces constitute an additional loading
that cannot be ignored. To make it tangible, we consider the cantilever beam
of Fig. 1.1.1a. The beam has a mass per unit length m and a flexural rigidity
EI , both assumed constant along the length, and it is subjected to the time-
(a)
(b)
(c)
FIG. 1.1.1 Vibrating cantilever beam.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 5
The transverse deflection is a function not only of the spatial variable x but
also of time t, namely it is u ¼ u ðx, t Þ. As the element has a mass mdx, an inertial
€ arises, which, according to d’Alembert’s principle (see Section 1.5)
force m udx
opposes the motion, that is, if the positive transverse displacement u ðx, t Þ in the
beam is directed downward, the inertial force is directed upward (see Fig. 1.1.1b
and c). Similarly, due to angular acceleration ∂u€ðx, t Þ=∂x of the cross-section, an
inertial moment is also developed, which we may neglect [1]. Thus, referring to
Fig. 1.1.1b, we obtain the equation of dynamic equilibrium of the beam element
in the y direction as
€ ¼0
Q + Q + dQ + pðx, t Þdx m udx
or
∂Q ∂2 u
¼ pðx, t Þ + m 2 (1.1.1)
∂x ∂t
∂3 u
Q ¼ EI (1.1.2)
∂x 3
∂4 u ∂2 u
EI + m 2 ¼ pðx, t Þ (1.1.3)
∂x 4 ∂t
Eq. (1.1.3) is known as the equation of the dynamic equilibrium or the equa-
tion of motion of the vibrating beam. It is apparent that if we omit the inertial
term m∂2 u=∂t 2 in Eq. (1.1.3), we obtain the equation of the deflection of the
beam under static loading, that is,
d4u
EI ¼ pðx Þ (1.1.4)
dx 4
Fig. 1.1.1c shows the beam subjected to the inertial forces. These forces
resist the accelerations and they need to be accounted for in the solution. This
is the most important characteristic of the dynamic problem. Obviously, the
magnitude of the inertial forces depends on the magnitude of the acceleration.
When the produced accelerations are very small, as in the case of slow motion,
the inertial forces are very small too, and they can be neglected. In this case,
the time appears in the equation as a parameter and the response at any instant
@CivilMethod
6 PART I Single-degree-of-freedom systems
can be obtained by the static structural analysis, even though the load and the
response are time-varying. This response is pseudodynamic and is referred to
as quasistatic. The inertial forces appear in the equation of motion of the struc-
ture with the second derivatives of the displacements with respect to time.
Therefore, the equations that must be solved in dynamic analysis in order
to establish the deformations and stresses in the structure are differential
equations, contrary to static analysis where the governing equations are alge-
braic. For this reason, the solution procedure in dynamic analysis is essentially
different from that used in static analysis.
Dynamic loads can be classified into two great groups that characterize the
approach of evaluating the structural response: The deterministic dynamic loads
and the nondeterministic or random dynamic loads. In the first group are the
dynamic loads whose time variation is fully determined, regardless of the
complexity of their mathematical presentation. They are also referred to as pre-
scribed dynamic loads. They can be represented by an analytic or a generalized
function (Dirac or Heaviside) as well as numerically by a set of their values at
discrete time instances. The second group includes the loads, whose time var-
iation is not completely known but it can be defined in a statistical sense. In this
book, the dynamic response of structures only under deterministic loads is
studied.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 7
5
2
H = me sin t
2.5
0
H(t)
–2.5
T T T
–5
0 2 4 6 8 10
t
FIG. 1.2.2 Harmonic loading due to an unbalanced rotating mass.
@CivilMethod
8 PART I Single-degree-of-freedom systems
structure and, as we will see later in this book (Chapter 6), it can be reduced to an
effective dynamic load if the accelerogram of the ground motion is known
(Fig. 1.2.4).
400
–400
0 10 20 30
FIG. 1.2.4 Effective dynamic load pðt Þ ¼ m u€g ðt Þ due to seismic ground motion.
If we examine static loading closer, we will see that even what we call static
loads are actually dynamic in nature. They are applied starting from a zero value
until the final prescribed value is reached within a time span. That is, they are
time-varying, thus dynamic. However, the duration of the application of the
static load is longer than the period of vibration of the structure. This produces
negligible accelerations and consequently the response under a “static load”
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 9
@CivilMethod
10 PART I Single-degree-of-freedom systems
Massless
Massless
columns
columns
(a) (b)
(c)
FIG. 1.3.1 Systems with one degree of freedom (SDOF).
Rigid
Rig
(a) (b)
FIG. 1.3.2 Systems with two degrees of freedom (2 DOF).
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 11
The lumped mass idealization provides a simple means of reducing the num-
ber of degrees of freedom. Fig. 1.3.3 represents the discrete model of a canti-
lever column, whose mass has been localized at three points. Neglecting the
axial deformation of the column and considering plane motion, the system
has six degrees of freedom, the three transnational ui ðt Þ and the three rotational,
fi ðt Þ. If the masses are fully concentrated so that their rotational inertia can be
ignored, the inertial moments Ii f€i are zero and the number of dynamic degrees
of freedom reduces to three. Obviously, the number of degrees of freedom
increases with the number of nodal points, where the mass of the structure is
lumped. As the number of points becomes infinitely large, the discretized struc-
ture approaches the continuous system (Fig. 1.3.4).
@CivilMethod
12 PART I Single-degree-of-freedom systems
Spring
Damper
Frictionless rollers
FIG. 1.4.1 Model of a SDOF system.
Center of mass
The forces applied to the body at time t are shown in the free body diagram
of Fig. 1.4.2. These are
(a) The external load pðt Þ
(b) The elastic force fS
(c) The damping force fD
(d) The inertial force fI .
The spring force fS depends on the displacement u ðt Þ and it is generally
expressed by a nonlinear function, fS ¼ fS ðu Þ. For linear response of the struc-
ture, the force fS is proportional to the displacement and is given by
fS ¼ ku (1.4.1)
where k is the constant that represents the spring stiffness coefficient, that is, the
force required to change the length of the spring by a unit. The force fS repre-
sents the elastic force of the structure that resists the motion and tends to bring
the body to its initial undeformed position.
The damping force fD also resists the motion. It represents the energy loss
due to internal or external dissipative forces. Damping forces are complex in
nature. Their exact expression in terms of the parameters of motion and of
the geometrical and material properties of the structure is complicated and dif-
ficult to determine. The simplest form of damping is linear viscous damping.
This produces damping forces, which are the easiest to handle mathematically
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 13
and provide analytical results for the response of a system close to the experi-
mental ones. The linear viscous damping mechanism is indicated by a dashpot,
as shown in Fig. 1.4.1. In viscous damping, the resisting force is proportional to
the velocity
fD ¼ cu_ (1.4.2)
where c is a constant that can be established experimentally. Inasmuch as the
work done by this force is converted to heat, the damping force is a nonconser-
vative force. It is the force that makes the amplitude of a vibrating
structure decay.
The inertial force fI depends on the mass m of the body and its acceleration
€ It also resists the motion. It is given by Newton’s second law of motionb
u.
fI ¼ m u€ (1.4.3)
A simple example of a structure that can be modeled as SDOF is the
one-story, one-bay frame of Fig. 1.4.3. It consists of two identical weightless
columns fixed on the ground and having height h, cross-sectional moment of
inertia Ι, and modulus of elasticity E. The cross-sectional moment of inertia
of the horizontal beam is assumed infinitely large. This means that the beam
behaves like a rigid body of mass m and hence the cross sections of the columns
at the roof level cannot rotate when the frame deforms. The frame is subjected to
an external horizontal force pðt Þ, as shown in Fig 1.4.3a, which forces the frame
to move. Neglecting the axial deformation of the beam and columns, an allow-
able assumption for frames, the only possible movement is the displacement
u ðt Þ at the roof level. The rotation of the beam as a rigid body is excluded
because this would cause a change in the length of columns.
(a)
(b) (c)
FIG. 1.4.3 Two-column shear frame.
b. Actually, this form of Newton’s law of motion is attributed to L. Euler, who defined it indepen-
dently as a mechanical principle [2, 3]. This law was recently derived from Kepler’s laws of plan-
etary motion [4].
@CivilMethod
14 PART I Single-degree-of-freedom systems
Referring to Fig. 1.4.3b, we see that the elastic forces are the shear forces Q
at the top cross-sections of the columns. These forces are given by the known
relation of statics
12EI
Q¼ u ðt Þ (1.4.4)
h3
The quantity 12EI =h 3 represents the translational stiffness of the column.
This is the force required to produce a unit relative displacement between
the end cross-sections of the column. These shear forces tend to restore the
frame to the undeformed position. Therefore, they play the role of the spring
in the SDOF model with a stiffness coefficient
12EI
k ¼2 (1.4.5)
h3
The inertial force is given by fI ¼ m u€ while the damping force by fD ¼ cu. _
Another convenient model to represent the single-story frame is shown in
Fig. 1.4.3c. It consists of a mass m placed at the top of a column with transla-
tional stiffness equal to the sum of the translational stiffness coefficients of the
columns of the frame. During the motion, the top cross-sections of columns
undergo only the translational displacement u ðt Þ. Models of this type are also
suitable to idealize multistory shear frames (see Fig. 1.4.4), in which the masses
are placed at the floor levels and the girders are assumed rigid.
FIG. 1.4.4 Four-story shear frame and its model without damping.
(a) (b)
FIG. 1.4.5 Two-story, two-bay shear frame and its model without damping.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 15
Fig. 1.4.5a shows another two-story shear frame. The columns are assumed
weightless. Fig. 1.4.5b shows its dynamic model. The column 1-2 is represented
by a spring of stiffness k ¼ 12EI =h 3 . The stiffness coefficients k1 and k2 include
only the stiffness of the columns with heights h1 and h2 , respectively.
Given the dynamic model of the structure, the equation of motion of the sys-
tem is formulated. For the SDOF system, the equation of motion can be formu-
lated using Newton’s second law of motion as it is applied for the motion of a
particle
m u€ ¼ F (1.4.6)
where
F ¼ pðt Þ fS fD (1.4.7)
is the resultant of the external forces. Using Eqs. (1.4.1), (1.4.2), (1.4.7),
Eq. (1.4.6) is written
m u€ + cu_ + ku ¼ pðt Þ (1.4.8)
Eq. (1.4.8) is the equation of motion of the SDOF system. The equation of
motion represents the dynamic equilibrium of the system. It is an ordinary dif-
ferential equation of the second order with respect to the unknown variable u ðt Þ.
The solution of this equation yields the displacement as a function of time. For
MDOF systems, the number of equations of motion that must be formulated is
equal to the number of dynamic degrees of freedom. The use of Newton’s law of
motion is not always well suited to formulate the equations, especially for
MDOF systems or complex SDOF systems. It requires advanced knowledge
of the dynamics of the rigid and deformable body as well as mastering various
special methods. Generally, the equations of motion can be formulated using:
(a) d’Alembert’s principle or method of equilibrium of forces.
(b) Principle of virtual work.
(c) Hamilton’s principle.
(d) Lagrange’s equations.
These methods will be presented in the following and will be demonstrated by
appropriate examples. The acquaintance with the application of these methods
constitutes a fundamental presupposition for the analysis of the dynamic
response of structures.
@CivilMethod
16 PART I Single-degree-of-freedom systems
where F is the resultant of all external forces acting on the particle of mass m and
€ is its acceleration with respect to an inertial frame of reference.c If we consider
u
that the term m€ u is another force, known as inertial force, then Eq. (1.5.1) states
that the vector sum of all forces, external and inertial, is zero during the motion.
But this is the necessary and sufficient condition for the static equilibrium of the
particle. Thus, in a sense, the dynamic problem is reduced to a problem of statics
according to the following statement, known as d’Alembert’s principle.
The laws of static equilibrium can be applied also to a dynamic system with
respect to an inertial frame of reference if the inertial forces are considered as
applied forces on the system together with the actual external forces.
The motion of a rigid body of mass m with respect to an inertial frame of
reference X, Y ,Z is decomposed into a translational motion of its center of
mass, where the whole mass is considered to be concentrated, and a rotational
motion about it (Fig. 1.5.1).
FIG. 1.5.1 Rigid body moving with respect to the inertial. frame X,Y , Z .
c. In classical dynamics, an inertial frame of reference is a frame of reference in which a body with
zero force acting upon it is not accelerating; that is, the body is at rest or it is moving at a constant
velocity in a straight line [5].
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 17
respect to the center to mass, and H_ c is the rate of change of the angular momen-
tum Hc of the body with respect to the same point given as
ZZZ
H_ c¼ r r€rdV (1.5.3)
V
Path of P
FIG. 1.5.2 Plane body moving in the XY plane. The system of xy axes moves with P without
rotating.
@CivilMethod
18 PART I Single-degree-of-freedom systems
1 2 1
T ¼ m X_ c + Y_ c + Ic w2
2
(1.5.7)
2 2
(b) with respect to an arbitrary point P of the body (K€onig’s theorem)
1 2 1
T ¼ m X_ p + Y_ p + Ip w2 + m xc Y_ p yc X_ p w
2
(1.5.8)
2 2
We shall write now Eqs. (1.5.4a)–(1.5.4c) in terms of the displacement
vector. Apparently, the displacement vector from the beginning of the motion
is defined as
u ¼ Rðt Þ Rð0Þ ¼ u ðt Þi + v ðt Þj (1.5.9)
where
u ¼ X ðt Þ X ð0Þ, v ¼ Y ðt Þ Y ð0Þ (1.5.10)
Hence, X€ ¼ u,
€ Y€ ¼ v€. Moreover, if fðt Þ represents the change of the rota-
_ w_ ¼ f,
tion in the same time interval and set w ¼ f, € Eqs. (1.5.4a)–(1.5.4c) are
written in terms of displacements as
Fx ¼ m u€c (1.5.11a)
Fy ¼ m v€c (1.5.11b)
Mc ¼ Ic f€ (1.5.11c)
or in matrix form
€c
Fc ¼ m c U (1.5.12)
where
8 9 8 9 2 3
< Fx = < uc = m 0 0
Fc ¼ Fx , Uc ¼ vc , mc ¼ 4 0 m 0 5 (1.5.13)
: ; : ;
Mc f 0 0 Ic
are the force vector, the displacement vector, and the mass matrix of the body,
respectively.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 19
€p
Fp ¼ m p U (1.5.15)
where
8 9
>
< Fx >
=
Fp ¼ F x (1.5.16a)
>
: >
;
Mp
8 9
< up >
> =
Up ¼ vp (1.5.16b)
: >
> ;
f
2 3
m 0 my c
mp ¼ 4 0 m mx c 5 (1.5.16c)
my c mx c Ic
Note that the mass matrix is not diagonal when the point of reference is not
the center of mass.
Finally, Eqs. (1.5.7), (1.5.8) are written as
1 1
T ¼ m u_ 2c + v_ 2c + Ic f_
2
2 2
(1.5.17)
1_T _c
¼ U mc U
2 c
1 1
T ¼ m u_ 2p + v_ 2p + Ip f_ + m xc u_ p yc v_ p f_
2
2 2
(1.5.18)
1_T _p
¼ U mp U
2 p
The set of equations with reference to point P can also be derived from the
set of equations with reference to point C by transforming the displacements
and the forces from point C to P (see Section 10.7).
Example 1.5.1 Equation of motion of an elastically supported body
Consider the rigid plate of constant thickness and total mass m shown in
Fig. E1.1a. The plate is hinged at O and elastically supported at A. Formulate
@CivilMethod
20 PART I Single-degree-of-freedom systems
the equation of motion of the system for small amplitude motion using the
method of equilibrium of forces.
Solution
The only possible motion of the plate is the rotation in its plane about the point
O. Hence, the system has one degree of freedom. The motion can be described
either by the rotation fðt Þ about O or by the translational displacement of
a point, for example, the displacement u ðt Þ of point B, which is related
to fðt Þ as
d2 1 b
fIx ¼ m ðCC 0 Þx ¼ m u€ (2c)
dt 2 2 a
d2 1
fIy ¼ m 2
ðCC 0 Þy ¼ m u€ (2d)
dt 2
u€
MIc ¼ IC f€ ¼ IC (2e)
a
The quantities ðCC 0 Þx and ðCC 0 Þy are the horizontal and the vertical dis-
placements of the center of mass C due to rotation, respectively. They are
obtained from Fig. E1.1b as
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 21
(a)
(b)
(c)
FIG. E1.1 Rigid plate in Example 1.5.1.
1b
ðCC 0 Þx ¼ ðOC Þf sin b ¼ u (3a)
2a
1
ðCC 0 Þy ¼ ðOC Þf cos b ¼ u (3b)
2
The equation of motion results from the dynamic equilibrium of moments
with respect to point O. Thus, we obtain
a 2a b a
W fS fIx fIy MIc + pðt Þa ¼ 0 (4)
2 3 2 2
@CivilMethod
22 PART I Single-degree-of-freedom systems
(a)
(b)
FIG. E1.2 Frame with a rigid column in Example 1.5.2.
Solution
The only possible motion of the system is the rotation fðt Þ of the column as a
rigid body about the hinged support at point A of its base. Because the rotation is
small, we have:
sin f f, cos f 1, f2 0
Hence
h h
u ¼ L sin f ¼ Lf, d ¼ sin f f
2 2
@CivilMethod
24 PART I Single-degree-of-freedom systems
h h
ð1 2Þ ¼ sin f + L cos f f + L,
2 2
h h
ð3 4Þ ¼ L sin f + cos f Lf + ,
2 2
h h
ð5 6Þ ¼ cos f
2 2
The forces applied on the column are shown in Fig. E1.2b. These are:
The elastic moment at the corner C MS ¼ ð1:5L
6EI
Þ2
4EI
d + 1:5L f
The elastic moment due to the rotational spring MR ¼ CR f ¼ EI
L f
MIA ¼ IA f€ ¼ mL €
2
The moment of inertia of the mass m 3 f
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 25
(a)
(b)
FIG. E1.3 System with two rigid bodies in Example 1.5.3.
Solution
As the bar AF is rigid, the only possible motion is its rotation about A. Hence,
the system has a SDOF. Its motion can be described either by the angle of rota-
tion fðt Þ about the hinge at A or by the transverse displacement of any point
along the axis of the bar. We choose the upward displacement u ðt Þ at point
B as the parameter of the motion. For small amplitude motion, the forces acting
on the system are shown in Fig. E1.3b. These are:
The elastic force fS at B: As it opposes the motion, it is directed downward
and is expressed as
fS ¼ ku (1)
The damping force fD at G: It is directed also downward and is expressed as
d d
fD ¼ c ðGG 0 Þ ¼ c ð1:625u Þ ¼ 1:625cu_ (2)
dt dt
The inertial force fIK and the inertial moment MIK at the center of mass K of
the bar due the distributed mass m are
d2
Þ
fIK ¼ ðm3L ðKK 0 Þ ¼ 0:75m u€ (3)
dt 2
MIK ¼ IK f€
@CivilMethod
26 PART I Single-degree-of-freedom systems
m ∗ u€ + c ∗ u_ + k ∗ u ¼ p ∗ ðt Þ (8)
where
m ∗ ¼ 6:833m, c ∗ ¼ 5:281c, k ∗ ¼ 2k, p ∗ ðt Þ ¼ 2L
pðt Þ (9)
The quantities defined by Eq. (9) are referred to as the generalized mass, the
generalized damping, the generalized stiffness, and the generalized load,
respectively.
Once the dynamic displacement u ðt Þ is established from the solution of
Eq. (8), the vertical reaction RA can be evaluated from the dynamic equilibrium
of forces in the direction of the y axis. This yields
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 27
(b)
(a)
FIG. E1.4 Single-story shear building in Example 1.5.4.
Solution
Taking into account that the structure is symmetric with respect to the x axis,
the columns are inextensible, and the load pðt Þ acts on the axis of symmetry, the
only possible motion of the plate is the horizontal displacement u ðt Þ in the
direction of the x axis. The SDOF model of the structure is shown in Fig. E1.4b.
The total mass of the system is due to the load of the plate and to half the
weight of the columns
5 10 20 + ð4 0:3 0:3 + 2 0:3 0:2Þ 2 24
m¼ ¼ 104:285
9:81
The stiffness of the system is equal to the sum of the translational stiffness
coefficients of all columns, which are given as
12EI i
ki ¼
hi3
where Ii is the moment of inertia of the cross-section of the i column with
respect to the y axis through its mass center and hi its height. Thus, we have
@CivilMethod
28 PART I Single-degree-of-freedom systems
Columns 30 30:
0:304
12 2:1 107
k3030 ¼ 12 ¼ 2657:8kN=m
43
Columns 30 20:
0:303 0:20
12 2:1 107
k3020 ¼ 12 ¼ 1771:9kN=m
43
Therefore the stiffness of the system is
k ¼ 4 2657:8 + 2 1771:9 ¼ 14175:0kN=m
The equation of motion results from the equilibrium of the forces shown in
Fig. E1.4b. This yields
fI fS + pðt Þ ¼ 0
or
m u€ + ku ¼ pðt Þ
Substituting the numerical values for m, k and the expression for pðt Þ, the
above equation of motion becomes
Rigid
Rigid
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 29
m€
u + ku ¼ pðt Þ (3)
where
u1 m1 0 k1 k1 p1 ðt Þ
u¼ , m¼ , k¼ , pðt Þ ¼
u2 0 m2 k1 k1 + k2 p2 ðt Þ
Solution
We choose O xy as the system of reference of the motion, whose origin coin-
cides with point O at the beginning of motion. Let xi , yi represent the coordi-
nates of the center of mass of the cross-section of i column and fi the angle
between its principal x axis and the x axis. The axes xy will be referred to
as the global axes of the system while the axes xy as the local axes of the
column.
Inasmuch as the axial deformation of columns is ignored, the only possible
motion of the plate is inside its plane, which can be determined by the two
translational displacements of a point and the rotation of the plate. We study
the motion of the plate with reference to point O and let U , V represent its
translational components with respect to the global axes xy, which are
assumed fixed in the plane, and W the rotation of the plate. As a consequence
of this motion, the cross-section of the i column at the level of the plate
undergoes the displacements u i , v i , wi , with respect to its base. These displace-
ments generate elastic forces X i , Y i , M i , which act on the plate. Thus,
we define the following vectors and matrices that will be used in the subse-
quent analysis.
@CivilMethod
30 PART I Single-degree-of-freedom systems
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 31
The transformation matrix for the vector quantities related to i column from the
global axes to the local axes is given as
2 3
cos fi sin fi 0
Ri ¼ 4 sin fi cos fi 0 5
0 0 1
Hence, the vectors are transformed from one system of axes to the other as
i
D i ¼ Ri D (1a)
i ¼ Ri T D i
D (1b)
i
FiS ¼ Ri F (2a)
S
i ¼ Ri T F i
F (2b)
S S
where
2 3
cos fi sin fi 0
R ¼ 4 sin fi
i T
cos fi 0 5
0 0 1
i 1 i T
is the transpose of R . Note that R
i
¼ R because Ri is orthonormal.
The elastic forces X , Y , M are related to the displacements u i , v i , wi by
i i i
12EI y i
Xi ¼ u (3a)
h3
12EI x i
Yi ¼ v (3b)
h3
GI t i
Mi ¼ w (3c)
h
where Ix ,Iy are the principal moments of inertia of the column cross-section and
It is the torsional constant, E and G are the material constants, and h is the
height of the column.
Setting
12EI y 12EI x GI t
i
k11 ¼ i
, k22 ¼ i
, k33 ¼ (4)
h3 h3 h
Eqs. (3a)–(3c) can be written in matrix form as
8 9 2 i 38 9
< Xi = k11 0 0 < u i =
Y i ¼ 4 0 k22 i
0 5 vi
: i; i : i;
M 0 0 k33 w
or
Fi ¼ k i D i (5)
@CivilMethod
32 PART I Single-degree-of-freedom systems
The matrix
2 i
3
k11 0 0
6 7
ki ¼ 4 0 i
k22 0 5
i
0 0 k22
or
F i D
i ¼ k i (6)
where
i ¼ Ri T k i R i
k (7)
is the stiffness matrix of the column in global axes, which becomes after per-
forming the matrix multiplications
2 i 3
k11 k12 0
i
6 7
i ¼ 6 ki ki 0 7
k (8)
4 21 22 5
0 0 k
i
33
where
9
k11 ¼ k11 >
i i
cos 2 fi + k22
i
sin 2 fi >
>
>
>
>
=
k22 ¼ k11
i i
sin 2 fi + k22
i
cos 2 fi
i (9)
k12 ¼ k21 ¼ k11 sin fi cos fi >
i i
k22
i >
>
>
>
>
;
i
k 33 ¼ k33
i
Inasmuch as the plate is rigid, the displacements ui , vi , wi of the i column
of point O. The geometrical rela-
depend on the plate displacements U , V , W
tions result from the following consideration.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 33
(b) The rotation of the plate about O. Referring to Fig. E1.7 and observing that
¼ wi , we obtain
cos ai ¼ xi =ri , sin ai ¼ yi =ri , W
i
u r ¼ ri W sin ai ¼
yi W
i
v r ¼ ri W cosai ¼ xi W
Thus, we have
ui ¼ ui t + ui r ¼ U yi W (10a)
vi ¼ vi t + vi r ¼ V + xi W (10b)
wi ¼ W (10c)
The previous equations are written in matrix form as
8 9 2 38 9
< ui = 1 0 y i < U =
vi ¼ 4 0 1 xi 5 V (11)
: i; :;
w 0 0 1 W
or setting
2 3
1 0
yi
ei ¼ 4 0 1 xi 5
0 0 1
@CivilMethod
34 PART I Single-degree-of-freedom systems
X
K
P y ðt Þ Y ¼ m V€
i €
+ xc W (13b)
i¼1
K
X
ðt Þ
M xi Y yi X + M
i i €
i ¼ m xc V€ yc U€ + Io W (13c)
i¼1
or
X
K
ðt Þ
P
T i
ei F €
¼ ðec ÞT mðec ÞU (14)
i¼1
where
8 9
< P x ðt Þ >
> =
ðt Þ ¼ P y ðt Þ
P (15a)
: >
> ;
M ðt Þ
2 3
m 0 0
m ¼ 40 m 0 5 (15b)
0 0 Ic
Finally, using Eqs. (6), (12), we obtain the equation of motion
M € + K
U U ¼P
ðt Þ (16)
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 35
where
2 3
m 0 m yc
¼ ðec ÞT mec ¼ 4 0
M m m xc 5 (17a)
m yc m xc Io
X
K
¼
K ei
T T
Ri ki Ri ei (17b)
i¼1
T
d. The notation eT ¼ ðe1 Þ is employed.
@CivilMethod
36 PART I Single-degree-of-freedom systems
The off-diagonal terms cause coupling between the elastic force in one
direction and the displacement in another direction. For example, the element
k12 represents the force acting in the direction of the x axis when the plate
undergoes a unit displacement in the direction of the y axis. Similarly, the ele-
ment k31 represents the moment acting on the plate about the z axis, if the plate
undergoes a unit displacement, U ¼ 1, in the direction of the x axis. The elastic
center or center of resistance of the plate is defined as the point of the plate
where an applied force in any direction does not produce rotation. This implies
the vanishing of the elements k13 and k23 (hence also their symmetric k31 and
k32 ) in the stiffness matrix (23). This point can be established as follows.
The stiffness matrix K is transformed from point O to the sought elastic cen-
ter E ðxE , yE Þ according to Eq. (22a), if ec is replaced by eE . Namely
2 32 32 3
1 0 0 k11 k12 k13 1 0 yE
T 1
E ¼ eE
K eE
K ¼ 40 1 0 54 k21 k22 k23 54 0 1 xE 5
yE xE 1 k31 k32 k33 0 0 1
or after performing the matrix multiplications
2 E E E3
k11 k12 k13
6 7
6 E E E7
K ¼ 6 k21 k22 k23 7
E 6
7
4 5
E E E
k 31 k 32 k 33
2 3
k11 k12 k11 yE k12 xE + k13
6 7
6 7
¼ 6 k21 k22 k21 yE k22 xE + k23 7
4 5
k11 yE k21 xE + k31 k12 yE k22 xE + k32 2 2
k 13 yE + k 23 xE + k 33
The vanishing of the elements k13 and k23 yields
E E
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 37
Thus, the stiffness matrix with respect to the elastic center takes the form
2 E E 32 3
k11 k12 0 k11 k12 0
6 7
E ¼ 6 kE kE 0 76
K 7
4 21 22 54 k 21 k 22 0 5
0 0 k 13 yE + k 23 xE + k 33
0 0 k
E 2 2
33
2k12
tan 2q ¼ (26)
k 22 k11
The axes defined by angle q are referred to as the principal directions of stiff-
ness of the structure. The stiffness matrix becomes now diagonal
2 3
k^11 0 0
K^E ¼6 4 0 k^22 0 5
7
0 0 k^33
where
k^11 ¼ k11 cos 2 q + k22 sin 2 q k12 sin 2q (27a)
k^22 ¼ k11 sin 2 q + k22 cos 2 q + k12 cos 2q (27b)
k^33 ¼ k13 y2E + k23 x2E + k33 (27c)
@CivilMethod
38 PART I Single-degree-of-freedom systems
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 39
(a)
(b)
FIG. E1.8 System in Example 1.6.1.
Solution
Because the cable is inextensible, the displaced configuration of the system can
be specified either by the angle of rotation of one of the bars or by the transverse
displacement of a point on it. Thus, the system has only one degree of freedom.
If the upward transverse displacement u ðt Þ of point C is taken as the basic
parameter of the motion, then all other displacements can be expressed in terms
of it. Fig. E1.8b shows the deformed system with all forces applied to it.
The elastic forces fS1 and fS2 are due to the deformation of the springs k1 and
k2 . They are directed downward as they oppose the motion. The force fD is
due to the viscous damping mechanism and is directed upward as it also
opposes the motion. The inertia moments MIA , MIO , and MIE are due to the rota-
tion of the masses about A, O , and E, respectively. All forces are expressed in
terms of the single displacement u ðt Þ
fS1 ¼ k1 ðBB 0 Þ ¼ ku=2, fS2 ¼ k2 ðCC 0 Þ ¼ 2ku
d
fD ¼ c ðDD 0 Þ ¼ cu_
dt
ð2LÞ3 u€
m
MIA ¼ IA f€1 ¼ 2 u€
¼ 1:333mL
3 2L
ð1:5LÞ3 u€
m
MIE ¼ IE f€2 ¼ 2 u€
¼ 0:750mL
3 1:5L
ð0:8LÞ2 u€
MIO ¼ IO f€3 ¼ mL
2 u€
¼ 0:200mL
8 0:4L
@CivilMethod
40 PART I Single-degree-of-freedom systems
If point C is given a virtual displacement du, the forces ride the following
displacements
d ðCC 0 Þ ¼ du, dðBB 0 Þ ¼ du=2, dðDD0 Þ ¼ du
df1 ¼ du=2L, df2 ¼ du=1:5L, df3 ¼ du=0:4L
du ðx Þ ¼ xdf1 ¼ xdu=2L
The work done by the forces acting on the system due to the virtual displace-
ment should be set equal to zero, that is,
fS1 d ðBB 0 Þ fS2 dðCC 0 Þ fD dðDD 0 Þ MIA df1
Z L
(1)
MIE df2 MIO df3 + pðt Þdu ðx Þdx ¼ 0
0
Using the expressions for the forces and the displacements in terms of the
basic displacement derived previously, Eq. (1) yields
0:25ku 2ku cu_ 1:333mL
2 u=2L
€ 2 u=1:5L
0:750mL €
0:200mL €
2 u=0:4L + pðt ÞL=4du ¼ 0
or, inasmuch as du 6¼ 0, the expression within the square brackets should vanish.
This yields the equation of motion
m ∗ v€ + c ∗ v_ + k ∗ v ¼ p ∗ ðt Þ (2)
where
m ∗ ¼ 1:667mL, c ∗ ¼ c, k ∗ ¼ 2:25k, p ∗ ðt Þ ¼ 0:25
pðt ÞL
Example 1.6.2 Equation of motion of a rigid body assemblage
Formulate the equations of motion of the rigid body assemblage shown in-
Fig. E1.9a by using the principle of virtual displacements on the basis of small
amplitude motion.
Solution
Due to the spring k1 , the rigid bars can rotate independently from each other
about their hinged supports at A and F. Hence, the system has two degrees
of freedom. Its motion can be specified by the transverse downward displace-
ments u1 ðt Þ and u2 ðt Þ of points C and E, respectively. The forces applied to the
displaced system are shown in Fig. E1.9b. They are
The elastic force fS1 ¼ k1 ðCC 0 Þ ¼ k ðu2 u1 Þ
The elastic force fS2 ¼ k2 ðDD 0 Þ ¼ 4ku 2
The damping force fD ¼ c dtd ðBB 0 Þ ¼ c u_21
The inertial moment MIA ¼ IA f€1 ¼ IA 2a u€1 2
¼ 4ma
3 u €1
€
The inertial moment M ¼ IF f2 ¼ IF ¼
F u€2 2
8ma
u€2
I a 3
The system is given a virtual displacement pattern du1 and du2 corresponding
to the two degrees of freedom. The forces ride the following displacements
du1 du1
d ðBB 0 Þ ¼ , dðCC 0 Þ ¼ du1 , df1 ¼
2 2a
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 41
(a)
(b)
FIG. E1.9 System in Example 1.6.2.
du2
d ðDD 0 Þ ¼ 2du2 , d ðEE 0 Þ ¼ du2 , df2 ¼
a
According to the principle of virtual displacements, the work done by the
applied forces must be equal to zero, that is,
Inasmuch as the quantities du1 and du2 are arbitrary, Eq. (2) is valid
only if
2ma u_ 1
u€1 + c k ðu2 u1 Þ pðt Þ ¼ 0 (3a)
3 4
8ma
u€2 + k ð9u2 u1 Þ ¼ 0 (3b)
3
@CivilMethod
42 PART I Single-degree-of-freedom systems
Eqs. (3a), (3b) are the equations of motion of the system. In matrix form they
are written as
2 3 2 c 3( ) "
2ma #( ) ( )
6 3 0 7 u€1 0 u_ 1 k k u1 pðt Þ
4 44 5 ¼
5 u€2 +
8ma u_ 2
+
k 9k u2 0
(4)
0 0 0
3
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 43
d 2r
m F¼0 (1.7.1)
dt 2
We confine our attention to an interval of time during which the particle
moves from point 1 at t ¼ t1 to point 2 at t ¼ t2 . We consider now a varied path,
specified by rðt Þ + drðt Þ, adjacent to the actual one. We will refer to the quan-
tity drðt Þ ¼ dx ðt Þi + dy ðt Þj + dz ðt Þk as the variation of r. The only restriction is
that the two paths coincide at time t ¼ t1 and t ¼ t2 . This implies that the var-
iation dr ¼ drðt Þ vanishes at these instants, that is,
The first step to derive Hamilton’s principle is to take the inner product of
the left side of Eq. (1.7.1) with the vector dr and to integrate from time t1 to time
t2 . This gives
Z t2
d2r
m dr F dr dt ¼ 0 (1.7.3)
t1 dt 2
Integrating by parts the first term in the above integral and knowing that the
operator d acts like the differential operator [6], we obtain
Z
t2 Z t2
t2
d 2r dr dr dr
m 2 drdt ¼ m dr m d dt
t1 dt dt t1 t1 dt dt
@CivilMethod
44 PART I Single-degree-of-freedom systems
The term outside the integral is equal to zero because of Eq. (1.7.2). More-
over, we can write the integrand as
" #
dr dr 1 dr dr 1 dr 2 1 dr 2
m d ¼m d ¼m d ¼ d m ¼ dT
dt dt 2 dt dt 2 dt 2 dt
where
2
1 dr
T¼ m ¼ x_ ðt Þ2 + y_ ðt Þ2 + z_ ðt Þ2 (1.7.4)
2 dt
is the kinetic energy of the particle. Hence, the integral (1.7.3) takes the form
Z t2
ðdT + F drÞdt ¼ 0 (1.7.5)
t1
The variation dr is a virtual displacement that leads from the actual path to
the varied one. Hence the term F dr in Eq. (1.7.5) is the virtual work done
by the force Fðt Þ. Eq. (1.7.5) is a statement of Hamilton’s principle as it is
applied to a particle. This equation can be transformed into a more convenient
form if the force Fðt Þ is separated in its conservative and nonconservative
components, that is
Fðt Þ ¼ Fc ðt Þ + Fnc ðt Þ (1.7.6)
A potential function A ¼ Aðx, y, z, t Þ exists from which the conservative
force Fc ðt Þ is derived as its minus gradient
∂A ∂A ∂A
Fc ¼ i+ j+ k (1.7.7)
∂x ∂y ∂z
Hence
∂A ∂A ∂A
Fc dr ¼ dx + dy + dz
∂x ∂y ∂z
or
Fc dr ¼ dA (1.7.8)
Hence, Hamilton’s principle, Eq. (1.7.5), can be written as
Z t2 Z t2
d ðT AÞdt + dWnc dt ¼ 0 (1.7.9)
t1 t1
where
dWnc ¼ Fnc dr
represents the virtual work of the nonconservative force.
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 45
has a stationary value. In fact, it can be shown that this value is the minimum
value of the integral.
The derivation of Hamilton’s principle for a particle can be extended to
MDOF systems as well as to continuous systems. The potential energy usually
arises from the gravity field. However, it may also arise from other sources such
as electrical and magnetic fields. The strain energy U ðt Þ should be included as
an additional potential energy. Thus, we can write
Z t2 Z t2
d ðU T + AÞdt dWnc dt ¼ 0 (1.7.13)
t1 t1
@CivilMethod
46 PART I Single-degree-of-freedom systems
The next step is to remove the variation d u_ of the velocity u_ from Eq. (6).
This is achieved using integration by parts as follows:
Z t2 Z t2
du
_ udt
m ud _ ¼ _
m ud dt
t1 t1 dt
Z t2
d
¼ m u_ ðdu Þdt (7)
t1 dt
Z t2
_ tt21
¼ ½m udu €
m ududt
t1
In order that the integral in Eq. (9) is equal to zero for any time interval
½t1 , t2 , its integrand should vanish, that is,
½m u€ + cu_ + ku pðt Þdu ¼ 0
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 47
and taking into account that du1 ðt1 Þ ¼ du1 ðt2 Þ ¼ du2 ðt1 Þ ¼ du2 ðt2 Þ ¼ 0, we
obtain
Z t2 Z t2
dTdt ¼ ðm1 u€1 du1 + m2 u€2 du2 Þdt (3)
t1 t1
Moreover, it is
p
dWnc ¼ p1 ðt Þdu1 + p2 ðt Þdu2 and A ¼ 0 (4)
Introducing Eqs. (1), (3), (4) into Hamilton’s principle, Eq. (1.7.13), we
obtain Z t2
½k1 ðu1 u2 Þdu1 k1 ðu1 u2 Þdu2 + k2 u2 du2 + m1 u€1 du1
t1
+ m2 u€2 du2 p1 ðt Þdu1 p2 ðt Þdu2 dt ¼ 0
@CivilMethod
48 PART I Single-degree-of-freedom systems
or
Z t2
f½m1 u€1 + k1 ðu1 u2 Þ p1 ðt Þdu1 + ½m2 u€2 k1 u1 + ðk1 + k2 Þu2
t1
p2 ðt Þdu2 gdt ¼ 0 (5)
Because Eq. (5) is valid for any interval ½t1 , t2 , its integrand must be equal
to zero, that is,
½m1 u€1 + k1 ðu1 u2 Þ p1 ðt Þdu1 + ½m2 u€2 k1 u1 + ðk1 + k2 Þu2 p2 ðt Þdu2 ¼ 0
(6)
Inasmuch as the quantities du1 and du2 are arbitrary, Eq. (6) is valid only if
the quantities in the square brackets are equal to zero, that is,
m1 u€1 + k1 ðu1 u2 Þ p1 ðt Þ ¼ 0 (7a)
m2 u€2 k1 u1 + ðk1 + k2 Þu2 p2 ðt Þ ¼ 0 (7b)
which give the equations of motion
m1 u€1 + k1 u1 k1 u2 ¼ p1 ðt Þ (8a)
m2 u€2 k1 u1 + ðk1 + k2 Þu2 ¼ p2 ðt Þ (8b)
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 49
Solution
Inasmuch as the bars are assumed rigid, this system has only two degrees of
freedom. The displaced configuration of the system can be determined from
the two transverse displacements u1 ðt Þ and u2 ðt Þ of points B and C . Referring
to Fig. E1.11, we have
9
f1 ¼ u1 =4L =
f2 ¼ ðu2 u1 Þ=3L (1)
;
f3 ¼ u2 =3L
The displacements of points of application of the forces and the changes of
angles are expressed in terms of the basic quantities u1 and u2 as
9
EE 0 ¼ u1 =4, FF 0 ¼ u1 =2, GG 0 ¼ 3u1 =4 >>
=
HH 0 ¼ u1 + ðu2 u1 Þ=3, QQ 0 ¼ u2 =2
(2)
DfB ¼ f2 f1 ¼ ð4u2 7u1 Þ=12L >
>
;
DfC ¼ f3 + f2 ¼ ð2u2 u1 Þ=3L
The potential energy U due to the deformation of the springs is
1 1 1 1
U ¼ k1 ðEE 0 Þ + k2 ðQQ 0 Þ + k3 ðDfB Þ2 + k4 ðDfC Þ2
2 2
2 2 2 2
which by virtue of Eqs. (2) becomes
1 2 1 2 1 2
U¼ ku 1 + ku 2 + k ð4u2 7u1 Þ2 + k ð2u2 u1 Þ2
32 4 72 9
¼ 0:934ku 21 + 1:361ku 22 1:667u1 u2
Its variation is
dU ¼ k ð1:868u1 1:667u2 Þdu1 + k ð1:667u1 + 2:722u2 Þdu2 (3)
The kinetic energy consists of the kinetic energies T1 and T2 of the bars ΑΒ
and CD, and of the kinetic energy T3 of the rigid body S. Thus, we have
2
1 1 1 d 1
T ¼ IA f_ 1 + ID f_ 3 + m ðHH 0 Þ + IH f_ 2
2 2 2
(4)
2 2 2 dt 2
@CivilMethod
50 PART I Single-degree-of-freedom systems
where
ð4LÞ3 ð3LÞ3 L3
m ¼ mL,
IA ¼ m
, ID ¼ m
, IH ¼ m (5)
3 3 6
Introducing Eqs. (1), (2), (5) into Eq. (4) yields
2 1 1 1
T ¼ m u_ 21 + m u_ 22 + m ðu_ 2 + 2u_ 1 Þ2 + m ðu_ 2 u_ 1 Þ2
3 2 18 108
¼ 0:898m u_ 21 + 0:565m u_ 22 + 0:204u_ 1 u_ 2
we obtain
Z t2 Z t2
dTdt ¼ ½m ð1:796u€1 + 0:204u€2 Þdu1 + m ð0:204u€1 + 1:130u€2 Þdu2 dt
t1 t1
(6)
The nonconservative forces include the loading pðx, t Þ and the damping
forces. Their virtual work is expressed in terms of the basic quantities as follows:
Z 3L x
dWnc p
¼ pðx, t Þ 1 du2 dx
0 3L
Z 3L (7)
x x
¼ p f ðt Þ 1 du2 dx ¼ 1:5
pLf ðt Þdu2
0 L 3L
d d d
D
dWnc ¼ c1 ðFF 0 Þd ðFF 0 Þ c2 ðGG 0 ÞdðGG 0 Þ c3 ðDfB ÞdðDfB Þ
dt dt dt (8)
d
c4 ðDfC Þd ðDfC Þ
dt
Using Eq. (2) and taking into account that c1 ¼ c, c2 ¼ 3c, c3 ¼ c4 ¼ 2cL2 ,
we can write
D
dWnc ¼ cð2:840u_ 1 0:833u_ 2 Þdu1 + cð0:833u_ 1 1:111u_ 2 Þdu2
Hence, we have
dWnc ¼ cð2:395u_ 1 0:833u_ 2 Þdu1 + cð0:833u_ 1 1:111u_ 2 Þdu2 + 1:5
pLf ðt Þdu2
(9)
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 51
Finally, the potential A of the external conservative forces is due to the con-
stant axial force P . Hence it is
A ¼ P ðDD 0 Þ ¼ Pe
and
dA ¼ Pde (10)
The variation de is evaluated as follows.
Referring to Fig. E1.11, we have
e ¼ ðAD Þ ðAD0 Þ ¼ 10L 4L cos f1 3Lcos f2 3L cos f3
Therefore
de ¼ Lð4sin f1 df1 + 3 sin f2 df2 + 3sin f3 df3 Þ
(11)
¼ Lð4f1 df1 + 3f2 df2 + 3f3 df3 Þ
which is introduced into Eq. (10) to yield
7P P P 2P
dA ¼ u1 + u2 du1 + u1 u2 du2 (12)
12L 3L 3L 3L
Introducing the expressions for dU , dT , dWnc , and dA into Hamilton’s prin-
ciple, Eq. (1.7.13), we obtain the following equations of motion
7P
1:796m u€1 + 0:204m u€2 + 2:395cu_ 1 0:833cu_ 2 + 1:868k u1
12L
P
+ 1:667k + u2 ¼ 0
3L
P
0:204m u€1 + 1:130m u€2 0:833cu_ 1 + 1:111cu_ 2 + 1:667k + u1
3L
2P
+ 2:722k u2 ¼ 1:5
pLf ðt Þ
3L
or in the matrix form
" #( ) " #( )
1:796 0:204 u€1 2:395 0:833 u_ 1
m +c
0:204 1:130 u€2 0:833 1:111 u_ 2
" #( ) ( ) (13)
1:868 0:583l 1:667 + 0:333l u1 0
+k ¼
1:667 + 0:333l 2:722 0:667l u2 pLf ðt Þ
1:5
where l ¼ P=kL.
The elastic forces of the system are
fS1 ¼ k ð1:868 0:583lÞu1 + k ð1:667 + 0:333lÞu2
fS2 ¼ k ð1:667 + 0:333lÞu1 + k ð2:722 0:667lÞu2
@CivilMethod
52 PART I Single-degree-of-freedom systems
Z
1 L
∂u ðx, t Þ 2
T¼ m dx (1)
2 0 ∂t
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 53
The strain energy of the beam is obtained by integrating the strain energy
density over its volume V , namely
Z
1
U¼ sx ex dV (2)
2 V
From the beam theory we have
M ðx Þ sx ∂2 u ðx, t Þ
sx ¼ y, ex ¼ , M ðx Þ ¼ EI
I E ∂x 2
Substituting the previous equations into Eq. (2) and integrating over the
cross-section of the beam yield
Z 2 2
1 L ∂ u ðx, t Þ
U¼ EI dx (3)
2 0 ∂x 2
For the simplicity of the expressions, the differentiation with respect to time
t will be designated by an over-dot while that with respect to the spatial coor-
dinated x by a prime. Moreover, the arguments will be dropped for the same
reason. Hence, expressions (1) and (3) can be rewritten as
Z
1 L
T¼ m u_ 2 dx (4)
2 0
Z
1 L
EI ðu 00 Þ dx
2
U¼ (5)
2 0
Their variations are
Z L
dT ¼ _ udx
m ud _ (6)
0
Z L
dU ¼ EI u 00 du 00 dx (7)
0
@CivilMethod
54 PART I Single-degree-of-freedom systems
Introducing Eqs. (5), (9), (10) into Hamilton’s principle, Eq. (1.7.13), we
obtain
Z t2
Z L Z L Z L
EI u 0000 dudx _ udx
m ud _ pðx, t Þdudx dt ¼ 0 (11)
t1 0 0 0
Because Eq. (13) is valid for any interval ½t1 , t2 , the integrand must vanish,
namely
Z L
½EI u 0000 + m u€ pðx, t Þdudx ¼ 0 (14)
0
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 55
(a) (b)
FIG. 1.8.1 Simple (a) and double (b) pendulum.
@CivilMethod
56 PART I Single-degree-of-freedom systems
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 57
The variations associated with the kinetic energy and the potential energy
defined by Eqs. (1.8.2), (1.8.3), respectively, are of the form
∂T ∂T ∂T ∂T
dT ¼ dq1 + ⋯ + dqN + d q_ + ⋯ + d q_
∂q1 ∂qN ∂q_ 1 1 ∂q_ N N
∂A ∂A
dA ¼ dq1 + ⋯ + dqN
∂q1 ∂qN
Substituting these expressions into Eq. (1.8.5), integrating by parts the terms
including d q_ i and taking into account dq1 ¼ dq2 ¼ ⋯ ¼ dqN ¼ 0 at instants t1
and t2 , we obtain
Z t 2
∂T d ∂T ∂A ∂T d ∂T ∂A
dq1 + ⋯ + dqN dt ¼ 0
t1 ∂q1 dt ∂q_ 1 ∂q1 ∂qN dt ∂q_ N ∂qN
Because the time interval ½t1 , t2 as well as the virtual displacements dqi are
arbitrary, this previous equation results in the following equations
d ∂T ∂T ∂A
+ ¼ 0 ði ¼ 1, 2, …, N Þ (1.8.6)
dt ∂q_ i ∂qi ∂qi
which, using Eq. (1.8.4), become
d ∂T ∂T
¼ Qi ði ¼ 1, 2, …, N Þ (1.8.7)
dt ∂q_ i ∂qi
Eq. (1.8.6) or (1.8.7) are the Lagrange equations of motion.
When nonconservative forces act on the system in addition to the conserva-
tive forces, we can include them in Lagrange’s equations, if the work done by
the nonconservative forces riding the virtual displacements is expressed in
terms of the generalized forces, that is,
dWnc ¼ Q1 dq1 + Q2 dq2 + ⋯ + QN dqN (1.8.8)
Introducing Eq. (1.8.8) into Hamilton’s principle, Eq. (1.7.9), the Lagrange
equations (1.8.6) become
d ∂T ∂T ∂A
+ ¼ Qi ði ¼ 1, 2, …, N Þ (1.8.9)
dt ∂q_ i ∂qi ∂qi
The elastic force components, which are derivable from a potential U (strain
energy), can be also involved in Eq. (1.8.9). Noting that
U ¼ U ðq1 , q2 , …, qN Þ (1.8.10)
the associated variation is
∂U ∂U
dU ¼ dq1 + ⋯ + dqN
∂q1 ∂qN
Therefore, the components ∂U =∂qi express generalized elastic forces and
Lagrange’s equations become
@CivilMethod
58 PART I Single-degree-of-freedom systems
d ∂T ∂T ∂V
+ ¼ Qi ði ¼ 1, 2, …, N Þ (1.8.11)
dt ∂q_ i ∂qi ∂qi
where
V ¼U +A (1.8.12)
X
N
dW ¼ Qi dqi (1.8.13)
i¼1
From physical consideration, the work done by the two sets of forces is the
same. The only difference is that they are expressed in two different coordinate
systems. Therefore, we can write
X
N X
K
Qi dqi ¼ Fk dxk (1.8.15)
i¼1 k¼1
or in matrix form
QT dq ¼ FT dx (1.8.16)
where
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 59
@CivilMethod
60 PART I Single-degree-of-freedom systems
∂T
¼ m2 L1 L2 q_ 1 q_ 2 sin ðq1 q2 Þ
∂q2
∂A
¼ m2 gL2 sin q2
∂q2
Applying Eq. (1.8.6) for i ¼ 1, 2 and q1 ¼ q1 q2 ¼ q2 , we obtain the equa-
tions of motion of the double pendulum
ðm1 + m2 ÞL1 q€1 + m2 L2 q€2 cos a + q_ 2 sin a + ðm1 + m2 Þg sin q1 ¼ 0 (4a)
2
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 61
Solution
Because the rod is no more inextensional, the system has two degrees of freedom.
Its displaced configuration can be specified either by the orthogonal coordinates x
and y of the mass or by the angle of the q and the axial deformation of the rod.
The kinetic energy of the system is
1
T ¼ m x_ 2 + y_ 2 (1)
2
The potential energy of the external force (gravitational force) is
A ¼ mgy (2)
and the potential of the elastic force
1
U ¼ ke2 (3)
2
where k ¼ EA=L is the axial stiffness of the rod and e its elongation. The latter
is expressed in terms of x and y as
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
e ¼ x 2 + y2 L (4)
Introducing Eq. (4) in the expression for the axial stiffness, Eq. (3), yields
1 EA pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi 2
U¼ x 2 + y2 L (5)
2 L
Differentiating the energies, we obtain
!
d ∂T ∂A ∂U EA L
¼ m x€, ¼ 0, ¼ 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x (6)
dt ∂x_ ∂x ∂x L x 2 + y2
!
d ∂T ∂A ∂U EA L
¼ m y€, ¼ mg, ¼ 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi y (7)
dt ∂y_ ∂y ∂y L x 2 + y2
Introducing Eqs. (6), (7) into Lagrange’s equations (1.8.11), we obtain the
equations of motion of the soft pendulum
!
EA L
m x€ + 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x ¼ 0 (8a)
L x 2 + y2
@CivilMethod
62 PART I Single-degree-of-freedom systems
!
EA L
m x€ + 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi y ¼ mg (8b)
L x 2 + y2
1 1
_ 2 m y U_ W _
_ + m xc V W
T ¼ m U_ 2 + V 2 + Io W
c (1)
2 2
The potential energy U consists of the strain energy of all columns. For the i
column it is
1 h i i 2 i 2 i 2 i
Ui ¼ k11 u + k22i
v + k33i
w
2
or using matrix notation
i 8 9
k 0 0 < u i =
1 i i i
11 i 1 T
Ui ¼ u v w 0 k22 0 v i ¼ Di ki Di (2)
2 0 0 k i wi : ; 2
33
∂T
¼0 (7d)
∂V
d ∂T € m y U€
+ m xc V€
¼ Io W (7e)
_
dt ∂W c
∂T
∂W ¼0 (7f)
@CivilMethod
64 PART I Single-degree-of-freedom systems
∂U
¼ k11 U + k12 V + k13 W (8a)
∂U
∂U
¼ k21 U + k22 V + k23 W (8b)
∂V
∂U
∂W ¼ k 31 U + k 32 V + k 33 W (8c)
M € + K
U U ¼P
ðt Þ (9)
where
2 3 2 3
m 0 m yc k 11 k12 k13
6 7 6 7
¼6 0
M m mx c 7 ¼ 6 k21
K k22 k23 7
4 5, 4 5,
my c mx c Io k31 k32 k33
8 9
> P ðt Þ >
< x =
P ðt Þ ¼
P y ðt Þ
>
: >
;
y A P x ðt Þ + xA P y ðt Þ
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 65
where
dgj
aji ¼ (1.8.27)
dqi
If we assume that the constraints are frictionless, then no work is done by the
constraint forces Ri when they ride any virtual displacement dqi , that is,
X
K
Ri dqi ¼ 0 (1.8.28)
i¼1
@CivilMethod
66 PART I Single-degree-of-freedom systems
where we note that a separate equation is written for each of the constraints.
Next, we subtract the sum of equations of the form (1.8.29) from
Eq. (1.8.28) and, interchanging the order of summation, we obtain
!
XK X n
Ri lj aji dqi ¼ 0 (1.8.30)
i¼1 j¼1
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 67
Solution
The kinetic and the potential energies of the system are
1
T ¼ m x_ 2 + y_ 2
2
A ¼ mgy
U ¼0
Because the rod is rigid, the coordinates must satisfy the constraint equation
g1 ¼ x 2 + y 2 L2 ¼ 0
Differentiating the quantities T and A we obtain
d ∂T ∂T ∂A ∂g1
¼ m x€, ¼ 0, ¼ 0, ¼ 2x, Q1 ¼ px
dt ∂x_ ∂qi ∂x ∂x
d ∂T ∂T ∂A ∂g1
¼ m y€, ¼ 0, ¼ mg, ¼ 2y, Q2 ¼ py
dt ∂y_ ∂y ∂y ∂y
Applying Eq. (1.8.32) for q1 ¼ x and q2 ¼ y we obtain the equations of
motion
m x€ ¼ px + 2xl (1a)
m y€ + mg ¼ py + 2yl (1b)
x 2 + y2 l 2 ¼ 0 (1c)
Eqs. (1) must be solved for the three unknowns x, y, and l. It should be
noted that two of these equations are differential and one algebraic and therefore
special care is required for their solution. A convenient method is to differen-
tiate the constraint equation twice with respect to time and then to solve the
€ y€ and the parameter
resulting linear system of equations for the accelerations x,
l. For the problem at hand, we obtain
2T
x x€ + y y€ ¼ (2)
m
Eqs. (1a), (1b), (2) are combined and written in matrix form
2 38 9 8 9
m 0 2x < x€ = < px =
4 0 m 2y 5 y€ ¼ py mg (3)
: ; : ;
x y 0 l 2T =m
which are solved to yield
px py mg
L2 x€ + x x_ 2 + y_ 2 ¼ y 2 xy (4a)
m m
2 px 2 py mg
L y€ + y x_ + y_ ¼ xy + x
2 2
(4b)
m m
@CivilMethod
68 PART I Single-degree-of-freedom systems
xpx y py mg
L l¼
2
T (4c)
2 2
Eqs. (4a), (4b) are solved using techniques for nonlinear differential equa-
tions. Analytical solutions are in general out of the question. However, a numer-
ical solution is always feasible using the methods presented in Chapter 5. Once
the coordinates x ðt Þ, y ðt Þ and the Lagrange multiplier l have been established,
they are utilized in Eq. (1.8.31) to evaluate the constraint forces, which are the
components of the axial force of the rod. Thus, we have
∂g ∂g qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Rx ¼ l ¼ 2lx, Rx ¼ l ¼ 2ly, S ¼ R2x + R2y ¼ 2lL (5)
∂x ∂y
for all virtual displacements dxi consistent with the constraints, which are
assumed workless and bilateral.
Inasmuch as the forces are conservative, they are derivable from a potential
function V ¼ V ðx1 , x2 , …, x3N Þ, V ¼ U + A, according to the relation
∂V
Fi ¼ (1.8.35)
∂xi
Using Eq. (1.8.35), Eq. (1.8.34) is written as
X
3N
∂V
dW ¼ dxi ¼ 0 (1.8.36)
i¼1
∂xi
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 69
Substituting the previous expression for dxi into Eq. (1.8.36), we obtain
3N X
X n
∂V ∂xi
dW ¼ dqj ¼ 0 (1.8.39)
i¼1 j¼1
∂xi ∂qj
Noting that
∂V X 3N
∂V ∂xi
¼ (1.8.40)
∂qj i¼1
∂xi ∂qj
Because dqj are assumed to be independent, the virtual work is zero only if
the coefficients of dqj are zero at the equilibrium condition, that is, if
∂V
¼ 0, j ¼ 1, 2, …, n (1.8.42)
∂qj 0
The subscript zero denotes that the derivatives refer to the equilibrium
position.
Let us expand now the potential energy function V ðq1 , q2 , …, qn Þ in a Tay-
lor series about the position of equilibrium
Xn n 2
∂V 1X n X
∂ V
V ¼ V0 + dqi + dqi dqj + ⋯ (1.8.43)
i¼1
∂qi 0 2 i¼1 j¼1 ∂qi ∂qj 0
We can arbitrarily set the potential energy at the reference position equal to
zero, that is,
V0 ¼ 0 (1.8.44)
@CivilMethod
70 PART I Single-degree-of-freedom systems
If we now assume that the displacements about the equilibrium position are
small, we can neglect terms of order higher than the second in Eq. (1.8.43).
Thus, using Eqs. (1.8.41), (1.8.44) the expression for the potential energy is sim-
plified as
n 2
1X n X
∂ V
V¼ qi qj (1.8.45)
2 i¼1 j¼1 ∂qi ∂qj 0
or setting
∂2 V
kij ¼ kji ¼ (1.8.46)
∂qi ∂qj 0
The quantities kij defined by Eq. (1.8.46) are the stiffness coefficients of the
system. Thus we see that the potential energy is expressed by a homogeneous
quadratic function of the generalized coordinates qi if small motions about the
position of equilibrium are examined.
Eq. (1.8.47) is written in matrix form
1
V ¼ qT kq (1.8.48)
2
where
8 9 2 3
>
> q1 >
> k11 k12 ⋯ k1n
< = 6 k21 k22
q2 ⋯ k2n 7
q¼ , k ¼6
4 ⋮ ⋮
7 (1.8.49)
>
> ⋮> > ⋱ ⋮ 5
: ;
qn kn1 kn2 ⋯ knn
The matrix k is called the stiffness matrix of the system.
The expression for the potential energy given in Eq. (1.8.47) is an example
of a quadratic form. For a system whose reference equilibrium configuration is
stable, the potential energy V is positive for all possible values of qi , except
q1 ¼ q2 ¼ … ¼ qn ¼ 0. In this case, the function V is referred to as positive def-
inite. This condition, however, puts restrictions on the allowable values of kij . It
is clear that all diagonal elements must be positive. The necessary and sufficient
condition that V be positive definite is that
2 3
k11 k12 ⋯ k1n
k k 6 k21 k22 ⋯ k2n 7
k11 > 0, 11 12 > 0, …, 6
4 ⋮ ⋮ ⋱ ⋮ 5>0
7 (1.8.50)
k21 k22
kn1 kn2 ⋯ knn
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 71
∂xk X n
∂xk
¼ q_ (1.8.52)
∂t j¼1
∂qj j
Introducing this expression into Eq. (1.8.51) we can write the kinetic energy
in the form
1X n X n
T¼ mij q_ i q_ j (1.8.54)
2 i¼1 j¼1
X
3N
∂xk ∂xk
mij ¼ mji ¼ mk (1.8.55)
k¼1
∂qi 0 ∂qj 0
The quantities mij defined by Eq. (1.8.55) are the inertia coefficients of the
system.
Eq. (1.8.54) is written in matrix form
1
T ¼ q_ T mq_ (1.8.56)
2
where
8 9 2 3
>
> q_ 1 >
> m11 m12 ⋯ m1n
< = 6 m21 m22
q_ 2 ⋯ m2n 7
q_ ¼ , m ¼6
4 ⋮
7 (1.8.57)
>
> ⋮> > ⋮ ⋱ ⋮ 5
: ;
q_ n mn1 mn2 ⋯ mnn
The matrix m is called the mass matrix of the system. The kinetic energy is a
positive definite quadratic function because it is the sum of positive quantities,
that is, the kinetic energies of the masses of the individual particles.
@CivilMethod
72 PART I Single-degree-of-freedom systems
∂V X n
¼ kij qj
∂qi j¼1
or in matrix form
m€
q + kq ¼ pðtÞ (1.8.59)
where p(t)¼Q
The matrices m and k are symmetric. It is an advantage of the Lagrange for-
mulation of the equations of motion that it preserves the symmetry of the coef-
ficient matrices for those cases where T and V are represented by quadratic
functions of the velocities and displacements, respectively.
where cij ¼ cji are the damping coefficients of the linear viscous damping.
Apparently, we can construct a quadratic function
1X n X n
R¼ cij q_ i q_ j (1.8.61)
2 i¼1 j¼1
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 73
which yields
∂R Xn
fDj ¼ ¼ cij q_ j (1.8.62)
∂q_ j i¼1
(a) (b)
FIG. 1.9.1 Influence of the gravity load.
@CivilMethod
74 PART I Single-degree-of-freedom systems
The elongation ust of the spring under its own weight will be
ust ¼ W =k ¼ constant (1.9.2)
Further we set
u ¼ ust + uðt Þ (1.9.3)
where uðt Þ represents the vertical displacement measured from the position of
the static equilibrium.
Differentiating Eq. (1.9.3) yields
_ u€ ¼ u€
u_ ¼ u, (1.9.4)
Using Eqs. (1.9.3), (1.9.4), the equation of motion (1.9.1) becomes
m u€
+ cu_ + ku st + k u ¼ pðt Þ + W
or using Eq. (1.9.2) we obtain
m u€
+ cu_ + k u ¼ pðt Þ (1.9.5)
The conclusion drawn from Eq. (1.9.5) states that, in the study of the
dynamic response of a system undergoing small displacements, the loads due
to gravity can be neglected. Of course, the total displacements will result as
the sum of the static plus dynamic displacements. That is, the superposition
principle is valid.
1.10 Problems
Problem P1.1 The plane square rigid body B of side length L and surface
mass density g is supported by two identical inclined columns having
cross-sectional moment of inertia I , modulus of elasticity E, and negligible
mass. Derive the equation of motion neglecting the axial deformation of
the columns (Fig. P1.1).
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 75
Problem P1.2 Consider the structure of Fig. P1.2a. The square plate of constant
thickness h ¼ a=10 and mass density g is supported at its center by a flexible
column having a circular cross-section with diameter d ¼ a=10, height a, and
material constants E, n. The plate is loaded by a force P acting in the plane
of the plate at point (Aða=8, a=6Þ and in the direction ∡x, P ¼ b ¼ 30° as
shown in Fig. P1.2b. Derive the equations of motion of the plate when the mass
of the column is neglected.
(a) (b)
FIG. P1.2 Structure in problem P1.2
Problem P1.3 The semicircular rigid plate of constant thickness and total mass
m is supported as shown in Fig. P1.3. Taking into account that the support at
point O is a hinge, formulate the equation of motion of the plate using (i)
the method of equilibrium of forces, (ii) the principle of virtual displacements,
and (iii) the method of the Lagrange equations.
Problem P.1.4 Consider the system shown in Fig. P1.4. The bars AD and EG
are rigid with masses m and m=3 , respectively. The mass at end D is concen-
trated. The elastic supports at points at B, E , and D are simulated by springs
with a stiffness k while the end G is supported by a viscous damper with a
damping coefficient c. The rod CE is weightless and rigid. Derive the equation
of motion using the principle of virtual displacements.
@CivilMethod
76 PART I Single-degree-of-freedom systems
Problem P1.5 Consider the system shown in Fig. P1.5. The mass m is sup-
ported at the top of the flexible and massless column 2 3, which is supported
on the ground by means of the rigid body 1 2 of mass 2a m. The support 1 is
elastically restrained by the rotational spring CR . Formulate the equation of
motion of the structure using CR ¼ EI =2a, m ¼ m=a.
Problem P1.6 Consider the two-story frame of Fig. P1.6. The columns 1 2,
10 20 , and the beam 3 30 are rigid while the columns 2 3, 20 30 , and the
beam 2 20 are massless and flexible with cross-sectional moment of inertia
I and modulus of elasticity E. The supports at 1 and 10 are elastically restrained
by rotational springs with a stiffness CR . Formulate the equation of motion of
the structure taking CR ¼ EI =2a and m ¼ m=a.
Problem P1.7 The system of Fig. P1.7 consists of the beam AB and the rigid
body S interconnected at B. The beam AB has a negligible mass, modulus of
elasticity E , and cross-sectional moment of inertia I . The beam is fixed at A
while the rigid body is elastically restrained at C by a rotational spring with
a stiffness CR ¼ EI =10L. The total mass m is uniformly distributed. The system
is loaded by the concentrated moment M ðt Þ at point B. Derive the equation of
motion of the system using Lagrange’s equations.
10 10
Problem P1.8 The frame of Fig. P1.8 consists of the rigid beam BD of total
mass m and the two massless and flexible columns AB and CD with a
cross-sectional moment of inertia I and modules of elasticity E. The two mass-
less cables FB and GD have cross-sectional area A and cannot undertake com-
pressive force. Derive the equation of motion of the structure taking
I =A ¼ a2 =25 and m ¼ m=5a.
Problem P1.9 Consider the two-story frame of Fig. P1.9. The columns of the
frame are rigid and have a surface mass density g. Their elastic support on the
ground is simulated by the rotational springs with a stiffness CR ¼ EI =10a.
The horizontal beams are massless and flexible with a cross-sectional moment
of inertia I and modulus of elasticity E. Derive the equation of motion when the
structure is subjected to the horizontal loads pðt Þ at the beam levels.
@CivilMethod
78 PART I Single-degree-of-freedom systems
Problem P1.10 The hinge O of the soft pendulum of Fig. P1.10 is elastically
restrained by the rotational spring with a stiffness CR ¼ EAL=10. The length of
the rod is L, its cross-sectional area A, and the modulus of elasticity E. Formu-
late the equation of motion of the pendulum.
Problem P1.11 The rigid bar AB of circular cross-section and mass density
m ¼ m=a is hinged at point A (Fig. P1.11). The cables DB,FB have cross-
sectional area A and modulus of elasticity E. They are assumed massless
and are prestressed so that they can undertake compressive forces. Formulate
the equation of motion of the structure taking into account that the load P is
removed suddenly at instant t ¼ 0. Evaluate the minimum prestressing force
of the cables DB, FB so that they can undertake compressive loads.
Problem P1.12 Consider the structure of Fig. P1.12. The column AC has a
circular cross-section and a mass per unit length m ¼ m=a; it is supported by
a spherical hinge on the ground and is kept in place by three elastic cables of
cross-sectional area A and modulus of elasticity E. The cables are assumed mass-
less and are prestressed so that they can undertake compressive force. Derive the
equation of motion of the structure when it is loaded by the horizontal force P ðt Þ
acting at the top of the column in the direction ∡x,P ¼ b (Fig. P1.12b).
(a) (b)
FIG. P1.12 Structure in problem P1.12
Problem P1.13 The silo of Fig. P1.13 is supported on its fundament by four
identical columns of a square cross-section. The silo is full of material of density
g. The ground yields elastically with a subgrade constant Ks . The silo and the
fundament are rigid. Derive the equation of motion of the structure when it is
loaded by the horizontal force P ðt Þ acting at the top of the silo in the direction
(a) (b)
FIG. P1.13 Silo on elastic subgrade. (a) Vertical section. (b) Plan form.
@CivilMethod
80 PART I Single-degree-of-freedom systems
∡x,P ¼ b (Fig. P1.13b) using the following data: Side of the columns a=4;
thickness of the bottom and walls of the silo a=8; density of the material of
the silo 1:5g; and soil constant Ks ¼ EI =1500a 3 .
Problem P1.14 Consider the one-story building of Fig. P1.14. The rigid plate is
an equilateral triangular with a side a and it is supported by three columns of
height a, rectangular cross-section a=10 a=20, and modulus of elasticity E.
The columns are fixed at both ends. Derive the equation of motion of the plate
when a horizontal force P ðt Þ acts at point Að0, a=5Þ in the direction ∡x, P ¼ b.
The dead weight of the plate is included in p (kN=m2 ).
Rigid plate
Problem P1.15 The two one-story buildings of Fig. P1.15 are connected with a
beam as shown in the figure. All columns have a square cross-section with a
moment of inertia Ic ¼ 2I . The connecting beam has a square cross-section with
moment of inertia Ib ¼ I . The structure is loaded by the horizontal force F(t) at the
level of the plates as shown in Fig. P1.15b. Formulate the equations of motion
using Lagrange’s equations. Assume: Torsion constant It ¼ 2:25d 4 =16, d ¼side
length of the square cross-section of the beam.
Rigid plate
(a)
beam
(b)
FIG. P1.15 Structure in problem P1.15. (a) vertical section, (b) plan form.
Problem P1.16 The system of Fig. P1.16 consists of the block of mass m1 ,
which can slide without friction on the inclined surface, and the pendulum of
@CivilMethod
General concepts and principles of structural dynamics Chapter 1 81
length L and mass m2 , which is pivoted at the center of mass of the block. The
rod of the pendulum has a cross-sectional area A and modulus of elasticity E.
Assuming plane motion, derive the equation of motion of the system taking
EA=L ¼ 5k and m1 ¼ 5m2 .
.
FIG. P1.16 System in problem P1.16
Problem P1.18 Consider the crane of Fig. P1.18. The horizontal beam is
assumed rigid. The column is flexible with a cross-sectional moment of inertia
I and the cable axially deformable with cross-sectional area A. The mass of the
cable and column is negligible. Derive the equation of motion of the system
when it is loaded by the horizontal force pðt Þ in the plane of the structure using
I =A ¼ a 2 =100 and a common modulus of elasticity E.
@CivilMethod
82 PART I Single-degree-of-freedom systems
@CivilMethod
Chapter 2
Single-degree-of-freedom
systems: Free vibrations
Chapter outline
2.1 Introduction 83 2.3.3 Overdamped system 96
2.2 Free undamped vibrations 83 2.4 Conservation of energy in an
2.3 Free damped vibrations 91 undamped system 97
2.3.1 Critically damped system 91 2.5 Problems 99
2.3.2 Underdamped system 92 References and further reading 103
2.1 Introduction
In this chapter, the free vibrations of a single-degree-of-freedom system
(SDOF) are studied, that is, its response when it is not subjected to any external
force, pðt Þ ¼ 0, but it is excited by an initial displacement and/or initial velocity.
The dynamic model of the system is shown in Fig. 1.4.1 and the equation of
motion (1.4.8) takes the form
m u€ + cu_ + ku ¼ 0 (2.1.1)
Eq. (2.1.1) is an ordinary linear homogeneous differential equation of the
second order with constant coefficients and its solution can be obtained using
known mathematical methods. Inasmuch as we are interested in the physical
response of the system described by this equation, it is advisable to analyze
the free vibration response in two stages, first for c ¼ 0 and then c 6¼ 0. In the
first case, we speak of free undamped vibrations while in the second case we
speak of free damped vibrations. Illustrative examples analyzing the free vibra-
tions of SDOF systems are presented. The pertinent bibliography with recom-
mended references for further study is also included.
u ¼ elt (2.2.2)
where l is an arbitrary constant to be determined. Substitution of Eq. (2.2.2)
into Eq. (2.2.1) gives
2
ml + k elt ¼ 0 (2.2.3)
ml2 + k ¼ 0 (2.2.4)
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 85
u_ 0
B¼ (2.2.12b)
w
and Eq. (2.2.10) becomes
u_ 0
u ðt Þ ¼ u0 cos wt + sin wt (2.2.13)
w
Obviously, it is u ðt Þ ¼ 0, when u0 ¼ u_ 0 ¼ 0. Hence, the system is set
to motion only if it is given an initial displacement and/or an initial
velocity.
We set
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2ffi
u_ 0
r ¼ ðu 0 Þ2 + (2.2.14)
w
Inasmuch as it is
2 2
u0 u_ 0 u0 u_ 0
1, 1 and + ¼1 (2.2.16)
r rw r rw
we can set
u0 u_ 0
¼ cos q and ¼ sin q (2.2.17)
r rw
and Eq. (2.2.15) becomes
u ðt Þ ¼ rcos ðwt qÞ (2.2.18)
where
u_ 0
q ¼ tan 1 (2.2.19)
wu0
Eq. (2.2.18) states that the motion of the system is a harmonic vibration
with amplitude ju ðt Þjmax ¼ r, angular velocity w, and phase angle q. The
@CivilMethod
86 PART I Single-degree-of-freedom systems
The time required for the undamped system to complete one cycle of free
vibration is referred to as the natural period of vibration of the system, which
is denoted by T and measured in seconds. It is related to the natural frequency of
vibration through
2p
T¼ (2.2.20)
w
The inverse of the period
1
f¼ (2.2.21)
T
expresses the number of cycles that the system performs in 1 s. This is referred
to as the natural cyclic frequency. The unit of f is the hertz (Hz) (cycles per
second, cps) and it is related to w through
w
f¼ (2.2.22)
2p
The displacement versus time for a system with w ¼ 8s1 , u0 ¼ 0:05m, and
u_ 0 ¼ 1m=s is shown in Fig. 2.2.2.
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 87
0.2
(du/dt)0 T = 2p /w
0.15
0.1
0.05
u0
0
–0.05
-
–0.1
–0.15
T = 2p /w
–0.2
0 0.5 1 1.5 2
FIG. 2.2.2 Response of an undamped SDOF system.
(a) (b)
FIG. E2.1 One-story shear building in Example 2.2.1. (a) Vertical section. (b) Plan form.
@CivilMethod
88 PART I Single-degree-of-freedom systems
Solution
The motion of the slab is described by the two displacements U ,V of its
center O along the x,y axes, respectively, and its rotation W about O. The
stiffness and mass matrices of the structure can be established using
Eqs. (17a), (17b) for the single-story building in Example 1.5.6. However,
taking into account that the structure is symmetric with respect to both axes,
the three components of the motion are uncoupled and an ad hoc solution can
be readily obtained.
(i) The stiffness of the columns in the x and y directions are:
12EI 1y
kx1 ¼ ky1 ¼ ¼ 1360:8kN=m (1a)
h3
12EI 2y
kx2 ¼ ¼ 907:2kN=m (1b)
h3
12EI 2x
ky2 ¼ ¼ 403:2kN=m (1c)
h3
Hence the respective stiffnesses of the structure are
Kx ¼ 4kx1 + 2kx2 ¼ 7257:6kN=m (2a)
Ky ¼ 4ky1 + 2ky2 ¼ 6249:6kN=m (2b)
The torsional stiffness KW is equal to the moment produced by the elastic
forces of the columns for unit rotation of the slab. Referring to Fig. E2.2,
we have
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 89
fx i ¼ kxi yi W (4a)
fy i ¼ kyi xi W (4b)
GI it E
M i ¼ kWi W, kW i ¼ , G¼ (5)
h 2ð 1 + n Þ
The moment of the elastic forces of column i with respect O is
M i ¼ yi2 kxi + xi2 kyi + kWi W (6)
GI 1t GI 2
kW1 ¼ ¼ 1998:7kNm, kW 2 ¼ t ¼ 823:2kNm (8)
h h
x1 ¼ x6 ¼ 4:85m, x2 ¼ x5 ¼ 0, x3 ¼ x4 ¼ 4:85m
y1 ¼ y2 ¼ y3 ¼ 2:85m, y4 ¼ y5 ¼ y6 ¼ 2:85m
X
KW ¼ yi2 kxi + xi2 kyi + kWi ¼ 1:9663 105 kNm (9)
i¼1
The mass of the slab and its moment of inertia with respect to O are
10 53 + 103 5
Io ¼ 70=9:81 ¼ 3716:44kNs2 (11)
12
@CivilMethod
90 PART I Single-degree-of-freedom systems
rffiffiffiffiffiffi 9
Kx 1 >
w1 ¼ ¼ 4:510s > >
>
>
m >
>
rffiffiffiffiffiffi >
=
Ky 1
w2 ¼ ¼ 4:185s > (12)
m >
>
rffiffiffiffiffiffiffi >
>
>
KW 1 >
w3 ¼ ¼ 7:274s > ;
I0
Fig. E2.3 presents the displacements of the column at the upper right
corner (x1 ¼ 4:85m, y1 ¼ 2:85m)
0.2
u1(t)
0.15 v1(t)
w1(t)
0.1
0.05
–0.05
–0.1
–0.15
–0.2
0 1 2 3 4 5
t
FIG. E2.3 Displacements u ðt Þ,v ðt Þ and rotation wðt Þ of the top cross-section of column at
1 ¼ 4:85m, y1 ¼ 2:85m in Example 2.2.1.
@CivilMethod
x
Single-degree-of-freedom systems: Free vibrations Chapter 2 91
0.07
0.06 (du/dt)0
0.05
0.04
u(t)
0.03
u0
0.02
0.01
0
0 0.2 0.4 0.6 0.8 1
t
FIG. 2.3.1 Response of a system with critical damping.
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 93
or
l1,2 ¼ xw iwD (2.3.12)
where
qffiffiffiffiffiffiffiffiffiffiffiffi
wD ¼ w 1 x2 (2.3.13)
@CivilMethod
94 PART I Single-degree-of-freedom systems
dynamic response of a SDOF system for various values of the damping ratio and
u0 ¼ 1, u_ 0 ¼ 7ms1 , w ¼ 6s1 . From the plot of Eq. (2.3.19), we conclude that
the motion of an underdamped system is harmonic vibration with frequency
wD and period T ¼ 2p=wD , whose amplitude, however, decays exponentially
with time and is bounded by the envelops rexwt . Hence, wD is called the
damped frequency and the respective period the damped period of the under-
damped system.
1.5 T = 2p /wD
T T
1
re−xw t
0.5 u0
u(t)
–0.5
−re-xw t
–1
–1.5
–2
0 1 2 3 4 5
t
FIG. 2.3.2 Response of an underdamped SDOF system.
2
x=0
1.5 x=0.03
x=0.06
x=0.1
1
0.5
u(t)
–0.5
–1
–1.5
–2
0 1 2 3 4 5
t
FIG. 2.3.3 Response of an underdamped SDOF system for various values of the damping ratio.
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 95
pffiffiffiffiffiffiffi2ffi
2npx
¼ u ðt Þe 1x
from which we obtain
u ðt Þ 2npx
d ¼ ‘n ¼ pffiffiffiffiffiffiffiffiffiffiffiffi (2.3.24)
u ðt + nT Þ 1 x2
The quantity d defined by Eq. (2.3.24) is called the logarithmic decrement
and can be employed to determine the damping ratio x experimentally, when the
displacements u ðt Þ and u ðt + nT Þ between n consecutive cycles are measured.
Example 2.3.1 Free damped vibrations of a silo
Fig. E2.4 presents the idealization of a silo. It consists of two massless and inex-
tensional columns of cross-sectional area a a and a square rigid plate of mass
m. At time t ¼ 0, the silo is displaced horizontally by a force P. Then the force is
suddenly removed and the system starts to vibrate. For P ¼ 200 kN, h ¼ 5m,
a ¼ 0:3m, E ¼ 2.1 107 kN/m2, m ¼ 100kN=ms2 determine:
(i) The damping ratio x of the system, if the horizontal displacement is reduced
to u1 ¼ 0:1u0 after n ¼ 5 oscillations.
(ii) Determine the displacement at time t1 ¼ 2.
@CivilMethod
96 PART I Single-degree-of-freedom systems
Solution
(i) The only possible motion of the structure is its horizontal displacement
u ðt Þ. Hence, the system has a single degree of freedom. The stiffness k
of the structure is due to the relative displacement of the column ends,
which yield the stiffness
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 97
The arbitrary constants A,B are evaluated from the initial conditions
u ð0Þ ¼ u0 and u_ ð0Þ ¼ u_ 0 . Then, Eq. (2.3.29) becomes
u_ 0 + u0 xw
u ðt Þ ¼ exwt u0 cosh Wt + sinh Wt (2.3.30)
W
Eq. (2.3.30) has been plotted in Fig. 2.3.4 with u0 ¼ 1m, u_ 0 ¼ 10ms1 , and
w ¼ 6s1 . It becomes evident that the motion of the overdamped system is
nonoscillatory.
1.5
x =1
x =1.5
x =2
x =2.5
1
u(t)
0.5
0
0 0.5 1 1.5 2
t
FIG. 2.3.4 Response of an overdamped SDOF system for various values of the damping ratio.
@CivilMethod
98 PART I Single-degree-of-freedom systems
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 99
2.5 Problems
Problem P2.1 The structure of Fig. P2.1 consists of two identical rigid bars BA
and BC , both having line density m (mass/length). The support A is a hinge
while the support C is a simple support. The bracing rod DF has a
@CivilMethod
100 PART I Single-degree-of-freedom systems
Problem P2.4 A SDOF system of mass m and stiffness k performs free vibra-
tions. At the end of four complete cycles, the displacement is u ð0Þ=3. If the mass
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 101
Problem P2.6 The horizontal force P applied to the structure of Fig. P2.6 is
suddenly removed at t ¼ 0. Determine the motion of the structure when (i)
the cables are free of any pretension and (ii) have been prestressed to withstand
compression, and compute the minimum required pretension forces. In both
cases, the cables are assumed massless. Data: a ¼ 2m, P ¼ 100kN,
m ¼ 100 kNm1 s2 , E ¼ 2.1 107 kN/m2, I ¼ 880cm4 , A ¼ 5cm2
Problem P2.7 In the structure of Fig. P2.7, the rigid rod AB of circular cross-
section and line mass density m ¼ m=a is supported on the ground by a spher-
ical hinge and held in vertical position by three identical elastic massless cables
of cross-sectional area A and modulus of elasticity E. The cables are prestressed
@CivilMethod
102 PART I Single-degree-of-freedom systems
so that they can withstand compression. Their anchor points D,G,F form an
equilateral triangle. Determine the motion of the structure if the horizontal
force P applied at the top of the rod in the direction of the y axis is suddenly
removed at t ¼ 0. Evaluate the minimum values of the prestress forces, which
ensure the capability of the cables to withstand compression. Data: a ¼ 2m,
m ¼ 100 kN m1s2, A ¼ 5cm2 , E ¼ 2.1 108 kN/m2, P ¼ 100kN.
(a) (b)
FIG. P2.7 Structure in Problem P2.7. (A) Vertical view. (B) Plan form.
Problem P2.8 In the pendulum of Fig. P2.8, the rigid rod suspending the con-
centrated mass m has a line mass density m ¼ 2m=L. The hinge O is elastically
restrained by the rotational spring with stiffness CR ¼ kL2 =2. The rod is
displaced by an angle q0 from the vertical position and then is left to move.
Considering small displacements, derive the equation of motion and compute
its period.
Problem P2.9 The rigid silo of Fig. P2.9 is supported on its fundament by four
identical columns of square cross-sections with a side length a=8. The silo is full
of material with density g. The bottom and the walls of the silo have a thickness
a=8 and material density 1:5g. Compute the frequencies and the periods of the
structure.
@CivilMethod
Single-degree-of-freedom systems: Free vibrations Chapter 2 103
(a) (b)
FIG. P2.9 Silo on four columns. (a) Vertical section. (b) Plan form.
@CivilMethod
Chapter 3
Single-degree-of-freedom
systems: Forced vibrations
Chapter outline
3.1 Introduction 105 3.5.1 Rectangular pulse load 126
3.2 Response to harmonic loading 106 3.5.2 Triangular pulse load 128
3.2.1 Response of undamped 3.5.3 Asymmetrical triangular
systems to harmonic pulse load 131
loading 106 3.5.4 Response to piecewise
3.2.2 Response of damped linear loading 135
systems to harmonic 3.6 Response to a periodic loading 137
loading 110 3.6.1 Periodic loads 137
3.3 Response to arbitrary dynamic 3.6.2 Fourier series 138
loading—Duhamel’s integral 113 3.6.3 Response of the SDOF
3.3.1 Undamped vibrations 113 system to periodic
3.3.2 Damped vibrations 116 excitation 143
3.4 Analytical evaluation of the 3.7 Response to unit impulse 146
Duhamel integral-applications 117 3.7.1 The delta function or
3.4.1 Response to step Dirac’s delta function 146
function load 117 3.7.2 Response to unit impulse 148
3.4.2 Response to ramp 3.7.3 Response to arbitrary
function load 120 loading 151
3.4.3 Response to step function 3.7.4 The reciprocal theorem
load with finite in dynamics 151
rise time 121 3.8 Problems 152
3.5 Response to impulsive loads 125 References and further reading 157
3.1 Introduction
In this chapter, the forced vibrations of the SDOF system are studied. The
dynamic model of the system is shown in Fig. 1.4.1 and the motion of the system
is governed by Eq. (1.4.8), namely
m u€ + cu_ + ku ¼ pðt Þ (3.1.1)
where pðt Þ represents an arbitrary function of time. First, the response under a
harmonic load is examined. This type of loading is particularly important in the
dynamic analysis of structures because it allows understanding the major
differences between the static and dynamic response and identifying phenom-
ena such as resonance that are not conceived by the static consideration. More-
over, the harmonic load analysis allows studying the response of SDOF systems
under a general periodic load using the Fourier series representation of a peri-
odic load. Then, the response of a SDOF system under an arbitrary load is stud-
ied using Duhamel’s integral. Finally, the response to a unit load is discussed by
exploiting the properties of the Dirac delta function. The chapter ends by pre-
senting the dynamic reciprocal theorem. Throughout the chapter, illustrative
examples analyzing the forced vibrations of SDOF systems are presented.
The pertinent bibliography with recommended references for further study is
also included. The chapter is enriched with problems to be solved aiming at
better understanding the theoretical issues.
u ¼ uh + up (3.2.3)
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 107
@CivilMethod
108 PART I Single-degree-of-freedom systems
Obviously
ust ¼ p0 =k (3.2.11)
denotes the static displacement that would be produced by a load p0 , equal to
the amplitude of the harmonic excitation, if it were to be applied statically.
The time-dependent quantity
u ðt Þ 1
Rðt Þ ¼ ¼ b sin wt Þ
ð sin wt (3.2.12)
ust 1 b 2
is called the response ratio. It is dimensionless and expresses the number that
must multiply the static displacement at time t to obtain the respective dynamic
displacement. The response ratio provides a measure of the influence of the
dynamic loading.
The extreme value
D ¼ max jRðt Þj (3.2.13)
is referred to as the dynamic magnification factor (DMF). It is a very useful
quantity in dynamic analysis because, if it has been established for a given
loading, the extreme state of deformation and stress can be obtained by static
analysis.
Eq. (3.2.12) for b ¼ 1 takes the indeterminate form
0
Rðt Þ ¼
0
whose limit can be determined using the L’H^
opital rule. Thus, on the basis of
Eq. (3.2.6b) we obtain
b sin wt
sin wt
‘im Rðt Þ ¼ ‘im
b!1 b!1 1 b2
sin bwt b sin wt
¼ ‘im
b!1 1 b2
(3.2.14)
wt cos bwt sin wt
¼ ‘im
b!1 2b
wt cos wt sin wt
¼
2
From the latter relation, it is concluded that when b tends to 1, that is, when
the excitation frequency w of the harmonic force approaches the natural fre-
quency of the system, the dynamic displacement grows indefinitely with time,
although the amplitude of the harmonic loading is finite. This phenomenon is
called resonance. The growth of the amplitude of the displacement with time
due to resonance is shown in Fig. 3.2.1. The response is periodic with a period
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 109
2p=w. A measure of the growth rate can be obtained by taking the difference
of the amplitudes of two consecutive peaks.
The time peak occurs when
20
w t/2
15
2p /w 2p /w
10
5
R(t)
–5
p
–10 p
–15
–w t/2
–20
0 1 2 3 4 5
t
FIG. 3.2.1 Response ratio of an undamped system at resonance, D ¼ max jRðt Þj ! 1
when t ! 1 (w ¼ 7, b ¼ 1).
dRðt Þ w2 sin wt
¼ ¼0 (3.2.15)
dt 2
or
np
t¼ n ¼ 1, 2, … (3.2.16)
w
Hence the difference between consecutive peaks is
np p np
R + R ¼ p cos np
w w w (3.2.17)
¼ p
the
When the excitation force is of the cosine type, pðt Þ ¼ p0 cos wt,
employed procedure yields the particular solution
p0 1
up ¼
cos wt (3.2.18)
k 1 b2
and the general solution for zero initial conditions, u ð0Þ ¼ u_ ð0Þ ¼ 0, is
obtained as
p0 1
u ðt Þ ¼ b cos wt Þ
ð cos wt (3.2.19)
k 1 b2
@CivilMethod
110 PART I Single-degree-of-freedom systems
p0 1
+ 2
1 b2 sin wt 2xb cos wt (3.2.25a)
k 1 b2 + ð2xb Þ2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Steady-state response
p0 1
+ 2
1 b2 2xb sin wt + 1 b2 cos wt (3.2.25b)
k 1 b2 + ð2xbÞ2
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Steady-state response
It is clear that the first term in Eq. (3.2.25a) decays rapidly with time because
of the exponential term exwt , so its contribution becomes negligible after a
short time, which, of course, depends on the damping ratio. Thus, the response
of the system is governed by the second term. For this reason, we say that the
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 111
first term expresses the transient response while the second term expresses the
steady-state response of the system. This is shown in Fig. 3.2.2.
Referring to Eq. (3.2.25a) we see that the steady-state response can be writ-
ten in the form
qÞ
u ðt Þ ¼ rsin ðwt (3.2.26)
where
p0 h
2 2 2
i1=2
1 2xb
r¼ 1 b + ð2xbÞ , q ¼ tan (3.2.27)
k 1 b2
Therefore, the steady-state response of the underdamped system subjected
to a harmonic loading is an undamped free vibration.
The DMF is
r h 2 i1=2
D ¼ max jRðt Þj ¼ ¼ 1 b2 + ð2xbÞ2 (3.2.28)
p0 =k
We observe that D ¼ D ðbÞ. Consequently the maximum value of D is
obtained when
dD
¼0 (3.2.29)
db
This condition gives
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b ¼ 1 2x 2 (3.2.30)
@CivilMethod
112 PART I Single-degree-of-freedom systems
6
x=0
5
x=0.1
4
D (b)
3
x=0.2
2 x=0.3
ξ=0.7
x=0.5
1
ξ=1.0
0
0 0.5 1 1.5 2 2.5 3
Frequency ratio, b
FIG. 3.2.3 Variation of the amplitude of the response ratio Dðb, x Þ.
and
1
Dmax ¼ pffiffiffiffiffiffiffiffiffiffiffiffi (3.2.31)
2x 1 x2
Apparently, the maximum value of the dynamic factor does not occur for
b ¼ 1. Nevertheless, for a small value of x it is b 1, for example, for
x ¼ 0:05 Eq. (3.2.30) gives b ¼ 0:9975. Fig. 3.2.3 shows the variation of D ver-
sus the frequency ratio b for different values of the damping ratio x. Note that
if x ¼ 0, Dmax becomes infinite.
4
1/2x
0
R(t)
–2
–1/2x
–4
–6
0 5 10 15
t
FIG. 3.2.4 Response ratio of an underdamped system at resonance (w ¼ 5, x ¼ 0:1, b ¼ 1,
D ¼ max jRðt Þj ! 1=2x when t ! 1).
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 113
Let us study now the response of the system when b ¼ 1, which is the
conventional value for resonance. For this value of b it is w ¼ w and
Eq. (3.2.25a) becomes
p0 cos wt
u ðt Þ ¼ exwt ðAsin wD t + B cos wD t Þ (3.2.32)
k 2x
If we assume zero initial conditions, u ð0Þ ¼ u_ ð0Þ ¼ 0, Eq. (3.2.32) gives
" ! #
1 xwt x
Rðt Þ ¼ e pffiffiffiffiffiffiffiffiffiffiffiffi sin wD t + cos wD t cos wt (3.2.33)
2x 1 x2
m u€ + ku ¼ pðt Þ (3.3.1)
1 pffiffiffiffiffiffiffiffiffi
u€ + w2 u ¼ pðt Þ, w ¼ k=m (3.3.2)
m
There are several methods to obtain the solution of Eq. (3.3.2). A con-
venient, straightforward, and rather simple method for solving ordinary dif-
ferential equations with constant coefficients is the Laplace transform
method [1, 3]. This method is based on the Laplace transform, which for a func-
tion u ðt Þ, t 0 is commonly denoted by L and defined as
Z 1
L½u ðt Þ ¼ U ðs Þ ¼ u ðt Þest dt (3.3.3)
0
@CivilMethod
114 PART I Single-degree-of-freedom systems
u ðt Þ ¼ L1 ½u ðt Þ (3.3.4)
1
L½u€ + w2 L½u ¼ L½pðt Þ (3.3.6)
m
s 1 1 1
U ðs Þ ¼ u ð 0Þ + 2 u_ ð0Þ + P ðs Þ (3.3.8)
s2 +w 2 s +w 2 m s + w2
2
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 115
or
1 s 1 s 1 1 s
u ðt Þ ¼ u ð0ÞL + u_ ð0ÞL + L P ðs Þ
s 2 + w2 s 2 + w2 m s 2 + w2
(3.3.9)
From the table of the Laplace transforms we obtain
s
L1 2 ¼ cos wt (3.3.10a)
s + w2
1 1 sin wt
L 2 2
¼ (3.3.10b)
s +w w
Now we focus our attention on the last term on the right side of Eq. (3.3.9).
Its inverse Laplace transform can be obtained using the convolution theorem.
The convolution of two functions f ðt Þ and g ðt Þ denoted by f ðt Þ∗ g ðt Þ or
ðf ∗ g Þt is defined as
Z t
f ðt Þ∗ g ðt Þ ¼ f ðt τÞg ðτÞdτ
0
Z t (3.3.11)
¼ g ðt τÞf ðτÞdτ
0
@CivilMethod
116 PART I Single-degree-of-freedom systems
It obvious that the first two terms in Eq. (3.3.14) express the contribution of
the initial conditions to the motion of the system. The third term, which
expresses the contribution of the external loading, is known as the Duhamel
integral for the undamped system.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 117
s 1 l1 t
L1 ¼ l1 e l2 el2 t
ðs l 1 Þð s l 2 Þ l1 l2 (3.3.22b)
xwt
¼e cos wD t
Z t
P ðs Þ 1
L1 ¼ pðτÞexwðtτÞ sin wD ðt τÞdτ (3.3.22c)
ðs l 1 Þ ð s l 2 Þ wD 0
It is obvious that the first term in Eq. (3.3.23) expresses the contribution of
the initial conditions to the motion of the system. The last term represents the
Duhamel integral for the underdamped motion. It is obvious that Eq. (3.3.23)
for x ¼ 0 yields Eq. (3.3.14) as anticipated. The Duhamel integral can also be
derived using the method described in Section 3.7.3.
@CivilMethod
118 PART I Single-degree-of-freedom systems
is the Heaviside step function and p0 the magnitude of the load. Fig. 3.4.1 shows
the step function load applied at t ¼ 0.
For u ð0Þ ¼ u_ ð0Þ ¼ 0 the displacement is obtained from the Duhamel integral
in Eq. (3.3.14)
Z t Z t
1 p0
u ðt Þ ¼ p0 sin ½wðt τÞdτ ¼ sin ½wðt τÞd ½wðt τÞ
mw 0 mw2 0
p0
¼ ½ cos wðt τÞt0
mw2
or
p0
u ðt Þ ¼ ð1 cos wt Þ (3.4.3)
k
Taking into account that ust ¼ p0 =k represents the static displacement, the
response ratio is
u ðt Þ
R ðt Þ ¼ ¼ ð1 cos wt Þ (3.4.4)
ust
and the DMF
D ¼ max jRðt Þj ¼ 2 (3.4.5)
Eq. (3.4.5) shows that the suddenly applied load produces a maximum dis-
placement that is twice as large than the displacement that the load p0 would
produce if it were applied statically (slowly). This is an elementary but impor-
tant result that illustrates the difference between static and dynamic loading of a
structure.
When the damping is taken into account, the displacement for zero initial
conditions, u ð0Þ ¼ u_ ð0Þ ¼ 0, is obtained from Eq. (3.3.23)
Z t
p0
u ðt Þ ¼ exwðtτÞ sin ½wD ðt τÞdτ (3.4.6)
mwD 0
The evaluation of the Duhamel integral is more complicated. Nevertheless,
using MAPLE we obtain
" ! #
p0 x
u ðt Þ ¼ 1 cos wD t + pffiffiffiffiffiffiffiffiffiffiffiffi sin wD t exwt (3.4.7)
k 1 x2
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 119
2.5
x=0
2
x=0.05
x=0.15
1.5
1
R(t)
0.5
0.
–0.5
0 0.5 1 1.5 2 2.5 3 3.5 4
t/TD (TD=2p /w D)
FIG. 3.4.2 Response ratio of a system under step load.
Fig. 3.4.2 shows the plot of Eq. (3.4.8) for different values of the
damping ratio.
The extreme values of Rðt Þ occur when
" #
2
dR xwt ðwx Þ
¼e + wD sin wD t ¼ 0 (3.4.9)
dt wD
which gives
np
tn ¼ n ¼ 0,1, 2, … (3.4.10)
wD
@CivilMethod
120 PART I Single-degree-of-freedom systems
that is, the maximum displacement occurs at the first peak, where t1 ¼ p=wD .
This is shown in Fig. 3.4.2. Eq. (3.4.12) for x ¼ 0 gives D ¼ 2 as anticipated.
Fig. 3.4.4 shows the response ratio Rðt Þ ¼ u ðt Þ=ust , ust ¼ p0 =k, of the
undamped system to a ramp function load. We see that it oscillates about the
line p0 t=t1 .
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 121
30
25
20
R(t)
15
10
0
0 0.5 1 1.5 2 2.5 3
t/T
FIG. 3.4.4 Response ratio of a system under ramp function load (t1 ¼ T =10).
3.4.3 Response to step function load with finite rise time. Static load
The step function load with finite rise time is a constant load that, however, is
not applied suddenly, but rises linearly up to a value p0 within a time t1 , the rise
time, and thereafter remains constant, as shown in Fig. 3.4.5. Mathematically it
is defined by
(p
0
t t t1
pðt Þ ¼ t1 (3.4.16)
p0 t > t1
@CivilMethod
122 PART I Single-degree-of-freedom systems
u_ I ðt1 Þ p0
uII ðt Þ ¼ sin wt+ uI ðt1 Þcos wt+ ð1 cos wtÞ (3.4.18)
w k
p0
uII ðt Þ ¼ (3.4.20)
k
or
RII ðt Þ ¼ 1 (3.4.21)
which means that the motion in the constant load phase is not oscillatory but the
displacement is constant and equal to the static displacement.
Fig. 3.4.6 presents the response ratio Rðt Þ for different values of t1 .
We observe that for smaller values of t1 =T , the response is similar to that
of a step function load while for larger values, the response is similar to that of
a static load. Therefore, the loads in real structures should not be applied
suddenly but slowly rising, that is, in time much larger than the natural
period of the structure to avoid dynamic magnification effects. Due to this
property, the step function load with finite rise time is also called static
loading.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 123
FIG. 3.4.6 Response ratio of a system under a step function load with finite rise time.
Solution
The mass of the structure, its stiffness in the x direction, and the corresponding
natural frequency were evaluated in Example 2.2.1
m ¼ 356:78kN m1 s2 , k ¼ Kx ¼ 7257:6kN=m, w ¼ w1 ¼ 4:51
@CivilMethod
124 PART I Single-degree-of-freedom systems
1.5
D = 1.265
0.5
R(t)
tmax = 0.563
–0.5
–1
–1.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
t
FIG. E3.2 Response ratio in Example 3.4.1.
Fig. E3.2 shows the plot of the response ratio Rðt Þ ¼ u ðt Þ=ðP=k Þ. The DMF
was found at D ¼ max jRðt Þj ¼ 1:1265. It was determined as the maximum
value of the array used to plot Rðt Þ. Thus, we have
umax ¼ Du st ¼ 1:1265 0:02067 ¼ 0:02328m (5)
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 125
The extreme values of the shear forces and bending moments are:
Columns 30 30
max Qx ¼ kx umax ¼ 1360:8 0:02328 ¼ 31:68kN
h
max Mx ¼ max Qx ¼ 31:68 5=2 ¼ 79:20kNm
2
Columns 30 20
max Qx ¼ kx umax ¼ 907:2 0:02328 ¼ 21:12kN
h
max Mx ¼ max Qx ¼ 21:12 5=2 ¼ 52:80kNm
2
t
@CivilMethod
FIG. 3.5.1 Time variation of an impulsive load.
126 PART I Single-degree-of-freedom systems
The response of the system to an impulsive load can be analyzed using the
methods presented in previous sections for the solution of the differential equa-
tion of motion under an arbitrary loading, that is, either by solving directly the
differential equation or by evaluating Duhamel’s integral. Another method to
obtain the response is to express the pulse as the superposition of two or more
simpler pulses for which the response solution is available or simple to deter-
mine (Fig. 3.5.2). Nevertheless, the analytical methods, especially for arbitrary
impulse loads, have lost their importance because of the development of effi-
cient numerical methods. For this reason, only two simple impulsive loads
are considered, the rectangular pulse load and the triangular pulse load.
p(t )
p(t ) p 0H (t ) if t t1
p0 p(t ) 0 if t t1
O t
t1
p(t )
p0 p(t ) p0 sin t if t t1
p(t ) 0 if t t1
O t
t1
FIG. 3.5.2 Impulsive loads represented by simple functions.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 127
@CivilMethod
128 PART I Single-degree-of-freedom systems
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u_ I ðt1 Þ 2
umax ¼ + ½ u I ðt1 Þ 2 (3.5.6)
w
Using Eqs. (3.5.4a), (3.5.4b) the previous relation becomes
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p0 2p 2p 2p
umax ¼ 1 2cos t1 + cos 2 t1 + sin 2 t1
k T T T
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
p0 2p
¼ 2 1 cos t1
k T
or
2p0 t1
umax ¼ sin p (3.5.7)
k T
and the DMF D will be given by
umax t1 T
D¼ ¼ 2sin p , t1 (3.5.8)
p0 =k T 2
Eq. (3.5.8) shows that the maximum response depends only on the ratio
t1 =T . The plot of the function D ¼ D ðt1 =T Þ shown in Fig. 3.5.4 is referred
to as the displacement response spectrum or simply the response spectrum of
the impulsive load (see also Chapter 6). It is evident that the response spectrum
of a pulse load serves to determine the maximum response of the system under
this load without solving the differential equation of motion for the particular
pulse load.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 129
u_ I ðt1 Þ
uII ðt Þ ¼ sin wt+ uI ðt1 Þcos wt, t¼ t t1 0 (3.5.11)
w
where
p0 sin wt1
u I ðt1 Þ ¼ coswt1 (3.5.12a)
k t1 w
p0 w cos wt1 1
u_ I ðt1 Þ ¼ + sin wt1 (3.5.12b)
k wt1 wt1
cos wt 1
u_ I ðt Þ ¼ w sin wt + ¼0 (3.5.13)
t1 t1
@CivilMethod
130 PART I Single-degree-of-freedom systems
2
3
t1
sin 2p
p0 6 T t1 7
uI ðt1 Þ ¼ 6 4 cos 2p 7 (3.5.16a)
k t1 T 5
2p
T
2
3
t1
cos 2p
p0 w 6
6 T t1 1 7 7
u_ I ðt1 Þ ¼ + sin 2p (3.5.16b)
k 4 t1 T t1 5
2p 2p
T T
The curve RII in Fig. 3.5.6 represents the function max uII =ust versus t1 =T .
From this figure, we conclude that for t1 =T 0:4 the maximum response D of
the system to the triangular pulse load occurs in Phase II while for t1 =T > 0:4 it
occurs in Phase I.
R = maxu /ust
D(0.4)=1.0513
D(t1/T)
R = maxu /ust
t1/T
FIG. 3.5.6 DMF for the triangular pulse load.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 131
p0 sin wt
uI ðt Þ ¼ t , 0 t t1 (3.5.17)
k t1 w
u_ I ðt1 Þ
uII ðt Þ ¼ sin wet + uI ðt1 Þ cos wet
w
(3.5.18)
p0 e sin wet et
+ 1 cos wt + , 0 et ¼ t t1 < t2
k t2 w t2
u_ II ðt1 + t2 Þ
uIII ðt Þ ¼ sin wt+ uII ðt1 + t2 Þcos wt, t¼ t ðt1 + t2 Þ 0
w
(3.5.19)
@CivilMethod
132 PART I Single-degree-of-freedom systems
1 sin wt
Rðt Þ ¼ t , t t1
t1 w
1
Rðt Þ ¼ 1 + ½ sinwðt t1 Þ sin wt , t t1
wt1
sin wt t
Rðt Þ ¼ 1 cos wt + , t t1
wt1 t1
sinwt1 sinwðt t1 Þ
Rðt Þ ¼ cos wt, t t1
wt1
1 sin wt t1
Rðt Þ ¼ t , 0t
t1 w 2
2 1 t1 t1
Rðt Þ ¼ t1 t + 2sin w t sin wt , t t1
t1 w 2 2
2 t1
Rðt Þ ¼ sinwðt t1 Þ + 2sinw t sinwt , t t1
wt1 2
b sinwt
sin wt
Rðt Þ ¼
1 b2
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 133
y
k1 k2 k1
M (t ) x
O 5m
k1 k2 k1
10 m
(a) (b)
(c)
FIG. E3.3 One-story building under impulsive load.
Solution
Because the structure is symmetric with respect to both axes x and y, the only
possible motion of the slab due to the moment M ðt Þ is the rotation fðt Þ about its
center O. Hence the equation of motion of the slab is
where IO is the moment of inertia of the mass of the plate with respect to O and
Kf the torsional stiffness of the structure. These quantities have been computed
in Example 2.2.1. Thus we have
rffiffiffiffiffiffi
Kf
w¼ ¼ 7:2738 (4)
IO
2p
T¼ ¼ 0:869s (5)
w
t1 ¼ T ¼ 0:869s (6)
@CivilMethod
134 PART I Single-degree-of-freedom systems
f_ I ðt1 Þ
fII ðt Þ ¼ sin wðt t1 Þ + fI ðt1 Þcos wðt t1 Þ (13)
w
Eq. (12) for t ¼ t1 gives
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 135
–3
4 ×10
fmax = 3.594e-3
3
2
f(t)
0
tmax = 0.462
–1 t1
–2
0 0.5 1 1.5 2 2.5 3 3.5 4
t
FIG. E3.4 Response of the structure in Example 3.5.1.
The stress resultants of the columns are computed using Eqs. (4a), 4(b), (6)
of Example 2.2.1. Thus, for the upper right column we have
kx1 ¼ ky1 ¼ 1360:8kN=m, kf1 ¼ 1:9985 103 kNm
x1 ¼ 4:85m, y1 ¼ 2:35m
max Qx1 ¼ kx1 y1 fmax ¼ 11:49kN (16a)
max Qy1 ¼ ky1 x1 fmax ¼ 23:72kN (16b)
h
max Mx1 ¼ max Qx1 ¼ 28:73kNm (16c)
2
h
max My1 ¼ max Qy1 ¼ 59:30kNm (16d)
2
max Mf1 ¼ kf1 fmax ¼ 7:183kNm (16e)
@CivilMethod
136 PART I Single-degree-of-freedom systems
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 137
in which
1 xbi
Ai0 ¼ ai 2
k w
bi
Ai1 ¼ , Ai2 ¼ u ðti1 Þ Ai0 (3.5.25)
k
1
A3 ¼
i
u_ ðti1 Þ + xwAi2 Ai1
wD
Differentiating Eq. (3.5.24) with respect to time gives the velocity
u_ i ðt Þ ¼ Ai1 + exwt xwAi2 + wD Ai3 cos wD t xwAi3 + wD Ai2 sin wD t
(3.5.26)
The presented method is exact. However, numerical methods are more con-
venient to compute the response to a piecewise linear loading (see Chapter 4).
(a)
(b)
(c)
FIG. 3.6.1 Examples of periodic loads.
@CivilMethod
138 PART I Single-degree-of-freedom systems
Z T =2
1sin mw0 tdt ¼ 0 for each m (3.6.5b)
T =2
Z T =2
0 if m 6¼ n
cos mw0 t cos nw0 tdt ¼ (3.6.5c)
T =2 T=2 if m ¼ n
Z T =2
cos mw0 t sin nw0 tdt ¼ 0 for each m, n (3.6.5d)
T =2
Z T =2
0 if m 6¼ n
sin mw0 t sin nw0 tdt ¼ (3.6.5e)
T =2 T =2 if m ¼ n
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 139
Z T =2
1
a0 ¼ pðt Þdt (3.6.6a)
T T =2
Z T =2
2
an ¼ pðt Þcos nw0 tdt (3.6.6b)
T T =2
The series (3.6.2) represents the function pðt Þ, that is, converges to pðt Þ for
n ! 1, provided that it satisfies the following conditions, known as Dirichlet
conditions:
(a) The function pðt Þ has a finite number of discontinuities in one period.
(b) The function pðt Þ has a finite number of maxima and minima in one period.
(c) The function pðt Þ is absolutely integrable over a period, that is,
Z T =2
jpðt Þjdt ¼ k < 1 (3.6.7)
T =2
We shall say that that the function pðt Þ is piecewise continuous in the finite
interval ½T =2, T =2, if it satisfies conditions (a) and (b). At the points of dis-
continuity, for example, point t1 in Fig. 3.6.1c, the Fourier series converges to
the mean value
1
p t1 + p t1+ (3.6.8)
2
where p t1 and p t1+ are the left and right limits of pðt Þ at t1 .
In practice, the periodic function pðt Þ is approximated by a finite number of
terms of the Fourier series, that is, by a finite Fourier series.
Let
X
k
Sk ð t Þ ¼ a 0 + ðan cos nw0 t + bn sin nw0 t Þ (3.6.9)
n¼1
be the sum of the first k + 1 terms of the Fourier series, which will represent the
function pðt Þ in the interval ½T=2, T =2. Then we will have
X
k
pðt Þ ¼ a0 + ðan cosnw0 t + bn sin nw0 t Þ + ek ðt Þ (3.6.10)
n¼1
@CivilMethod
140 PART I Single-degree-of-freedom systems
where
ek ðt Þ ¼ pðt Þ Sk ðt Þ (3.6.11)
is the error between pðt Þ and its approximation. The mean square error Ek is
given by
Z
1 T =2
Ek ¼ ½ek ðt Þ2 dt
T T =2
Z " #2 (3.6.12)
1 T =2 Xk
¼ pðt Þ a0 ðan cos nw0 t + bn sin nw0 t Þ dt
T T =2 n¼1
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 141
(a) (b)
FIG. 3.6.2 Functions pðt Þ replacing pðt Þ in the interval ½t1 e, t1 + e.
@CivilMethod
142 PART I Single-degree-of-freedom systems
Z T
2 cos np
bn ¼ pðt Þ sin nw0 t ¼ p0 (2c)
T 0 np
Hence
p0 p0 X
N
cos np 1 p0 X
N
cos np
pðt Þ ¼ + 2 cos nw 0 t + sin nw0 t (3)
4 p n¼1 n2 p n¼1 n
Fig. E3.5 shows the graphical representation of the finite Fourier series in
Eq. (3) for various values of N with p0 ¼ 1, T ¼ 1. We observe that the conver-
gence is very slow at point t ¼ T =2 due to the Gibbs phenomenon.
N = 20 N = 100
1.2 1.2
1 1
0.8 0.8
0.6 0.6
p(t)
p(t)
0.4 0.4
0.2 0.2
0 0
–0.2 –0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t t
N = 1000 N = 2000
1.2
1 1
0.8 0.8
0.6 0.6
p(t)
p(t)
0.4 0.4
0.2 0.2
0 0
–0.2 –0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t t
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 143
Solution
The function pðt Þ within a period T ¼ 3t1 can be represented using the delta
function.a Thus we obtain
pðt Þ ¼ P ½d ðt T =3Þ d ðt 2T =3Þ (1)
Eqs. (3.6.6a)–(3.6.6c) give
Z
1 T
a0 ¼ pðt Þdt ¼ 0 (2a)
T 0
Z
2 T 2P 2np 4np
an ¼ pðt Þcos nw0 tdt ¼ cos cos (2b)
T 0 T 3 3
Z
2 T 2P 2np 4np
bn ¼ pðt Þ sin nw0 t ¼ sin sin (2c)
T 0 T 3 3
Inserting these values of the coefficients in Eq. (3.6.2) we obtain
(
1
2P X 2np 4np
pðt Þ ¼ cos cos cos nw0 t
T n¼1 3 3
)
2np 4np
+ sin sin sin nw0 t (3)
3 3
a. The delta function d ðt t0 Þ and its properties are discussed in Section 3.7.1.
@CivilMethod
144 PART I Single-degree-of-freedom systems
For the sine term bn sin nw0 t the steady-state response results from
Eq. (3.2.25a) by setting p0 ¼ bn , w ¼ nw0 , and b ¼ b n ð¼ nw0 =wÞ
bn 1
unsin ðt Þ ¼ 2 1 b2n sin nw0 t 2xbn cos nw0 t
k 1b 2
+ ð2xbn Þ 2
n
(3.6.16b)
while for the cosine term an cos nw0 t the steady-state response is obtained from
Eq. (3.2.25b) by setting p0 ¼ an , w ¼ nw0 , and b ¼ b n ð¼ nw0 =wÞ
an 1
uncos ðt Þ ¼ 2xbn sin nw0 t + 1 b 2n cos nw0 t
k 1 b 2 2 + ð2xb Þ2
n n
(3.6.16c)
The steady-state response of the damped system to periodic loading results
as the superposition of responses to individual terms of the Fourier series
X
n X
n
u ðt Þ ¼ u0 ðt Þ + unsin ðt Þ + uncos ðt Þ (3.6.17)
n¼1 n¼1
(3.6.19)
Theoretically, there is no transient response when x ¼ 0. However, for
small values of damping, which is the usual case in our structures, the
steady-state response can be obtained from Eqs. (3.6.16b), (3.6.16c) for
x 0. This yields
bn 1
unsin ðt Þ sin nw0 t (3.6.20a)
k 1 b2n
an 1
uncos ðt Þ cos nw0 t (3.6.20b)
k 1 b 2n
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 145
(a) (b)
FIG. E3.7 Response to periodic loading in Example 3.6.3.
@CivilMethod
146 PART I Single-degree-of-freedom systems
FIG. 3.7.1 Distribution of a concentrated impulsive force over the interval ½e, +e.
namely, the total force is equal to unity. If we skip the problem of determining
analytically the function pðt Þ, we may assume a priori a prescribed shape for
this function, for example,
k=2, jt j < 1=k
pk ðt Þ ¼ (3.7.2a)
0, jt j 1=k
or
k
pk ðt Þ ¼ (3.7.2b)
pð1 + k 2 t 2 Þ
where k is a positive number.
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 147
Fig. 3.7.2a and b show both functions pk defined in Eqs. (3.7.2a), (3.7.2b),
respectively. Moreover, they satisfy Eq. (3.7.1), which means that they are
equivalent to pðt Þ and can represent the actual force P.
(a) (b)
FIG. 3.7.2 Functions pk sufficiently concentrated for large values of k.
or by the relation
Z +1
dðt t1 Þf ðt Þ dt ¼ f ðt1 Þ (3.7.5)
1
@CivilMethod
148 PART I Single-degree-of-freedom systems
and
Z +1 Z e
d ðt Þdt ¼ d ðt Þdt ¼ 1 (3.7.6b)
1 e
or for e ! 0
mDu_ ¼ 1 (3.7.14)
where
Du_ ¼ u_ t1+ u_ t1 (3.7.15)
Hence the impulsive load produces an abrupt change (discontinuity) of the
t1 . If the system is at rest before the action of the impulsive load,
velocity at time
then it is u_ t1 ¼ 0 and Eq. (3.7.14) gives
1
u_ ðt1 Þ ¼ (3.7.16)
m
where t1 designates t1+ . Therefore, Eq. (3.7.16) presents the initial velocity
given to the system by the impulsive load. However, the displacement remains
continuous, which means that
Du ¼ u t1+ u t1 ¼ 0 (3.7.17)
If the elastic and damping forces are taken into account, then Eq. (3.7.12b)
is written
d u_ du
m ¼ pðt Þ c ku (3.7.18)
dt dt
which after integration over the interval ½e, +e gives
Z +e Z +e
+e +e
m ½u_ e ¼ pðt Þdt c½u e k u ðt Þdt (3.7.19)
e e
Applying the mean value theorem of integral calculus to the integral of elas-
tic force, we write Eq. (3.7.19) as
Z +e
+e +e
m ½u_ e ¼ pðt Þdt c½u e ku ðt ∗ Þ2e, e < t ∗ < e (3.7.20)
e
@CivilMethod
150 PART I Single-degree-of-freedom systems
h(t-t)
t =1
t
FIG. 3.7.3 Response to unit impulse.
1
u€ + 2xwu_ + w2 u ¼ d ðt τ Þ (3.7.23)
m
The solution of the above equation can be obtained from Eq. (3.3.23) for
zero initial conditions and pðt Þ ¼ dðt t1 Þ. Thus, we have
Z t
1
u ðt Þ ¼ dðτ t1 ÞexwðtτÞ sin ½wD ðt τÞdτ (3.7.24)
mwD 0
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 151
WI , II ¼ WII , I (3.7.28)
In dynamics of linear systems, the reciprocity is expressed by the dynamic
reciprocal theorem [10], known also as the dynamic Betti-Rayleigh theorem.
This theorem for SDOF systems reads
If two loadings pI ðt Þ and pII ðt Þ act separately on a linear dynamic system and
produce the responses uI ðt Þ and uII ðt Þ then the convolution CI , II of the loading
pI ðt Þ with the response uII ðt Þ is equal to the convolution CII , I of the loading pII ðt Þ
with the response uI ðt Þ
@CivilMethod
152 PART I Single-degree-of-freedom systems
CI , II ¼ CII , I (3.7.29)
or using definition (3.3.11) for the convolution, we may write
pI ðt Þ∗ uII ðt Þ ¼ pII ðt Þ∗ uI ðt Þ (3.7.30a)
or in integral form
Z t Z t
pI ðτÞuII ðt τÞdτ ¼ pII ðτÞuI ðt τÞdτ (3.7.30b)
0 0
3.8 Problems
Problem P3.1 A machine carrying a mass m0 is placed on the slab of the build-
ing of Fig. P3.1. The mass rotates eccentrically about the point ð2:5, 2:0Þ at a
distance s ¼ 1:0m with a frequency f ¼ 4Hz. The columns are massless, inex-
tensible, and fixed on the base while the slab is assumed uniform and rigid.
Determine the time history of the shear forces Qx , Qy , the bending moments
Mx , My , and the torsion moment Mw at the top of the columns. The material
constants are E ¼ 1:2 107 kN=m2 , n ¼ 0:2. The total load of the plate (dead
plus live) is q ¼ 12kN=m2 ; m0 ¼ m=5, where m is the total mass of the slab.
The acceleration of gravity is g ¼ 9:81m=s2 .
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 153
(a) (b)
FIG. P3.4 One-story building in problem P3.4
@CivilMethod
154 PART I Single-degree-of-freedom systems
the x direction. Plot the response ratio Rðt Þ of the structure and the stress resul-
tants at the base of the columns 1 and 3 using the following data: Height of col-
umns a ¼ 4m, cross-section of columns a=10 a=20, side length of the
triangular slab a, p0 ¼ 10kN, E ¼ 2:1 107 kN=m2 , load of the plate, including
the dead load, q ¼ 20kN=m2 , t1 ¼ p=2w, x ¼ 0:05, w ¼ 8s1 , acceleration of
gravity g ¼ 9:81m=s . 2
Problem P3.5 The structure of Fig. P3.5a consists of the rigid girder BC and
the two flexible columns AB and CD having a cross-sectional moment of inertia
I and a modulus of elasticity E. The cables FB and GC have cross-sectional area
A, cannot withstand compression, and are assumed massless. The structure is sub-
jected to the impulsive loads pðt Þ shown in Fig. P3.5b and c. Study the response of
the structure and determine the maximum error when the impulsive loads are
substituted by equivalent concentrated forces. Plot the function D ðt1 =T Þ for the
two load cases. For which value of the ratio t1 =T is the maximum error less than
2%? Data: a ¼ 1:5m, I ¼ 33,740cm4 (IPE450), E ¼ 2:1 108 kN=m2 ,
A ¼ 3cm2 , t1 ¼ 0:1, m ¼ 1:0kNm1 s2 =m, p10 ¼ 10kN. The value p20 is deter-
mined so that both loads have the same impulse.
(a)
(b) (c)
FIG. P3.5 Structure in problem P3.5
@CivilMethod
Single-degree-of-freedom systems: Forced vibrations Chapter 3 155
Problem P3.6 Consider the structure of Fig. P3.6. The rigid column AC of circu-
lar cross-section and mass per unit length m is supported by the three elastic cables
of cross-sectional area A and modulus of elasticity E. The support on the ground is
a spherical hinge. The cables have been prestressed so that they can withstand com-
pression. Three advertising panels are massless fixed at the top of the columns, as
shown in the figure. The structure is subjected to the wind blast load of Fig. P3.6c in
the y direction (see Fig. P3.6b). Determine the minimum prestress force of the
cables. The cables and the panels have negligible mass. Use the data a ¼ 5:0m,
p0 ¼ 4kN=m2 , m ¼ 0:5kNm1 s2 =m, E ¼ 2:1 108 kN=m2 , A ¼ 4cm2 ,
t1 ¼ 0:1s.
(a) (b)
(c)
FIG. P3.6 Structure in problem P3.6
@CivilMethod
156 PART I Single-degree-of-freedom systems
10
p
0
4
2
t1 t2
0
p1
–2
0 0.5 1 1.5 2
FIG. P3.7 Blast pressure in problem P3.7
Problem P3.8 Show that an impulsive load generated by the load pðt Þ
and acting over the interval ½t1 , t2 can be represented as pI ¼ ½H ðt t1 Þ
H ðt t2 Þpðt Þ, where H ðt ti Þ is the Heaviside step function. Write a
MATLAB program that constructs the impulsive load.
Problem P3.9 Determine the dynamic response of a SDOF system subjected to
the sine periodic loading of Fig. P3.9. Assume: T ¼ p=w, w ¼ 1:1w,
x ¼ 0, and
x ¼ 0:1.
@CivilMethod
Chapter 4
4.1 Introduction
The previous analysis shows that an analytical solution of the equation of
motion for an single-degree-of-freedom (SDOF) system is possible only if
the external force is described by a simple function. If the excitation force varies
arbitrarily with time or is given by a set of its values, an analytical solution is out
of the question. However, such problems can be tackled numerically by time
step integration methods for differential equations. The literature about these
methods is vast. Extensive chapters and whole books cover this subject. They
present the mathematical development of these methods, their computer imple-
mentation, and their accuracy, convergence, and stability. Several computer
packages include ready-to-use subroutines for the solution of the differential
equation of motion.
Some of these methods have been specially developed for the study of the
dynamic response of systems. A survey of these methods is given in [1, 2]. The
central difference method (CDM), Houbolt’s method, Wilson’s q-Method, and
Newmark’s method are the most well known among them [3, 4]. Nevertheless,
with the increase of cheap computer power, some of them have lost their impor-
tance while others have taken dominating places in the computational arena.
Further, adding Eqs. (4.2.1), (4.2.2) and neglecting the terms of order higher
than three, we obtain the following expression to approximate the second deriv-
ative of u ðt Þ at time t
u ðt + Dt Þ 2u ðt Þ + u ðt Dt Þ
u€ðt Þ (4.2.5)
Dt 2
Substitution of the derivatives u_ ðt Þ and u€ðt Þ into Eq. (4.1.1) with their
approximations (4.2.4) and (4.2.5) gives
u ðt + Dt Þ 2u ðt Þ + u ðt Dt Þ u ðt + Dt Þ u ðt Dt Þ
m +c + ku ðt Þ ¼ pðt Þ
Dt 2 2Dt
which is solved for u ðt + Dt Þ to yield
m
c 2m m c
+ u ð t + Dt Þ ¼ p ð t Þ k u ðt Þ u ðt Dt Þ
Dt 2 2Dt Dt 2 Dt 2 2Dt
(4.2.6)
K^ u ðt + Dt Þ ¼ P^ (4.2.7)
where
m c
K^ ¼ 2 + (4.2.8)
Dt 2Dt
2m m c
P^ ¼ pðt Þ k 2 u ðt Þ u ðt Dt Þ (4.2.9)
Dt Dt 2 2Dt
The quantities K^ and P^ are referred to as the effective stiffness and the
effective load, respectively. Obviously, Eq. (4.2.7) allows the evaluation of
the displacement at instant t + Dt, if the displacements at the two preceding
instants t and t Dt are known. Because u ð0Þ is known from the initial
@CivilMethod
162 PART I Single-degree-of-freedom systems
conditions, the procedure starts at t ¼ Dt. Obviously, this requires the value of
u ðDt Þ, which is unknown in the first instance, but it can be determined from
Eq. (4.2.2) for t ¼ 0. Thus neglecting terms of order higher than two, we have
1
u ðDt Þ u ð0Þ Dt u_ ð0Þ + Dt 2 u€ð0Þ (4.2.10)
2
In the above equation, the quantities u ð0Þ and u_ ð0Þ are known from the ini-
tial conditions while u€ð0Þ can be computed from the equation of motion,
Eq. (4.1.1), for t ¼ 0. Thus, we obtain
u€ð0Þ ¼ ½pð0Þ cu_ ð0Þ ku ð0Þ=m (4.2.11)
The stability of the CDM requires that time step Dt is less than a certain
critical value, that is, qffiffiffiffiffiffiffiffiffiffiffiffi
Dt Dtcr ¼ T 1 x2 =p (4.2.12)
where T is the period of the system (see Eq. 4.5.30). Otherwise, the procedure
“blows up” with time and the solution makes no sense. This is discussed in
Section 4.5.
Because T is usually a small number, Dt should be small, which implies
that a large number of time steps are required to solve the equation of motion.
This has been a major drawback of the method, especially in older times when
the computer capabilities in terms of memory and speed were restricted. This
fact has led researchers to develop integration methods in which the size of
the time step is not restricted by a critical value. Table 4.2.1 presents the
@CivilMethod
Numerical integration of the equation of motion Chapter 4 163
Solution
The solution is obtained using the program centr_diff_lin.m with Dt ¼ 0:01.
Fig. E4.1 gives the response of the SDOF system. Moreover, Fig. E4.2 shows
Displacement Velocity
0.2 0.5
0.1
u,t(t)
u(t)
0
0
–0.1 –0.5
0 5 10 0 5 10
t t
Acceleration Response ratio
4 2
2
1
u,tt (t)
R(t)
0
0
–2
–4 –1
0 5 10 0 5 10
t t
FIG. E4.1 Response of the SDOF system in Example 4.2.1.
@CivilMethod
164 PART I Single-degree-of-freedom systems
the displacement u ðt Þ as compared with the exact one together with the error
u ðt Þ uex ðt Þ. The exact solution was obtained by analytical evaluation of
Duhamel’s integral giving
4po ð1 2xwÞ
u ðt Þ ¼ e 10:5t
+ cos w D t + sin w D t e 1xwt
m ð1 4xw + 4w2 Þ 2wD
(1)
x 10 -4
0.2 1
u(t) computed
u(t) exact 0.8 u-uex
0.15 0.6
0.4
0.1 0.2
u(t)
0
0.05 –0.2
–0.4
0 –0.6
–0.8
-0.05 –1
0 2 4 6 8 10 0 2 4 6 8 10
t
FIG. E4.2 Computed solution and error in Example 4.2.1.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 165
FIG. 4.3.1 Variation of the acceleration, velocity, and displacement in the average acceleration
method.
τ2
u ðt + τÞ ¼ τu_ ðt Þ + ½u€ðt Þ + u€ðt + Dt Þ + C2
4
@CivilMethod
166 PART I Single-degree-of-freedom systems
Setting
Du ¼ u ðt + Dt Þ u ðt Þ (4.3.6a)
Du_ ¼ u_ ðt + Dt Þ u_ ðt Þ (4.3.6b)
Du€ ¼ u€ðt + Dt Þ u€ðt Þ (4.3.6c)
Eqs. (4.3.4), (4.3.5) are written as
Dt
Du_ ¼ ½2u€ðt Þ + Du€ (4.3.7)
2
Dt 2
Du ¼ Dt u_ ðt Þ + ½2u€ðt Þ + Du€ (4.3.8)
4
€ we obtain
Solving Eq. (4.3.8) for Du,
4
Du€ ¼ ½Du Dt u_ ðt Þ 2u€ðt Þ (4.3.9)
Dt 2
2
Du_ ¼ Du 2u_ ðt Þ (4.3.10)
Dt
@CivilMethod
Numerical integration of the equation of motion Chapter 4 167
@CivilMethod
168 PART I Single-degree-of-freedom systems
As will be shown, contrary to the CDM, the stability of the AAM does not
demand any restriction on the size of the time step Dt. The time step, however,
is influenced by the accuracy of the method and its capability to describe an
oscillatory motion. Therefore, it must be small enough. The selection of Dt
equal to 1/10 of the period of the system or of the period of the excitation force
produces accurate results.
Adhering to the steps of Table 4.3.1, a computer program called av_acc_lin.m
has been written in MATLAB for the numerical integration of the equation of
motion. The program is available on this book’s companion website. It computes
the displacement u ðt Þ, the velocity u_ ðt Þ, the acceleration u€ðt Þ, and the response
ratio Rðt Þ ¼ u ðt Þ=ðpmax =k Þ and gives their graphical representation. Moreover, it
computes the dynamic magnification factor D ¼ max jRðt Þj and the time tmax it
occurs. The user of the program is responsible for providing the function of the
excitation force.
Solution
The solution was evaluated using the program av_acc_lin.m with Dt ¼ 0:01.
Fig. E4.3 gives the graphical representation of the displacement together with
the error u ðt Þ uex ðt Þ. The computed error by the AAM is almost double the
error of the CDM. Moreover, Fig. E4.4 shows the response of the system under
the harmonic load p ¼ 2 sin 5t. Obviously, this excitation produces resonance
(w ¼ w ¼ 5).
x 10−4
0.2 2
u(t) computed
u(t) exact 1.5 u-uex
0.15
1
0.5
0.1
u(t)
0.05
–0.5
–1
0
–1.5
−0.05 –2
0 2 4 6 8 10 0 2 4 6 8 10
t t
FIG. E4.3 Computed solution and error in Example 4.3.1.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 169
Displacement Velocity
0.01 0.04
0.005 0.02
u,t(t)
u(t)
0 0
–0.005 –0.02
–0.01 –0.04
0 5 10 15 0 5 10 15
t t
Acceleration Response ratio
0.2 10
0.1 5
u,tt (t)
R(t)
0 0
–0.1 –5
–0.2 –10
0 5 10 15 0 5 10 15
t t
FIG. E4.4 Response of the system in Example 4.3.1 under harmonic load.
@CivilMethod
170 PART I Single-degree-of-freedom systems
it, we have
u€ ¼ q ðt Þ (4.4.3)
where q ðt Þ is a fictitious source, unknown in the first instance. Eq. (4.4.3) is the
analog equation of Eq. (4.4.1). It indicates that the solution of Eq. (4.4.1) can
be obtained by solving Eq. (4.4.3) with the initial conditions (4.4.2), if the q ðt Þ
is first established. This is achieved as follows.
Taking the Laplace transform of Eq. (4.4.3) we obtain
1 1 1
U ðs Þ ¼ u ð0Þ + 2 u_ ð0Þ + 2 Q ðs Þ (4.4.4)
s s s
where U ðs Þ,Q ðs Þ are the Laplace transforms of u ðt Þ,q ðt Þ, respectively. The
inverse Laplace transform of Eq. (4.4.4) gives the solution in integral from
Z t
u ðt Þ ¼ u ð0Þ + u_ ð0Þt + q ðτÞðt τÞdτ (4.4.5)
0
Thus, the IVP of Eqs. (4.4.1), (4.4.2) is transformed into the equivalent
Volterra integral equation for q ðt Þ.
Eq. (4.4.5) is solved numerically within a time interval ½0, T . The interval
½0, T is divided into N equal intervals, Dt ¼ h, h ¼ T =N , in which q ðt Þ is
assumed to vary according to a certain law, for example, constant, linear,
etc. In this analysis, q ðt Þ is assumed to be constant and equal to the mean value
in the interval h (Fig. 4.4.1). That is
qr1 + qr
qrm ¼ (4.4.6)
2
FIG. 4.4.1 Discretization of the interval ½0, T into N equal intervals h ¼ T=N .
(4.4.7)
@CivilMethod
Numerical integration of the equation of motion Chapter 4 171
where
h2
c1 ¼ (4.4.9)
2
The velocity is obtained by direct differentiation of Eq. (4.4.5) using
Leibnitz’ rule for integrals [12]. Thus, we have
Z t
u_ ðt Þ ¼ u_ ð0Þ + q ðτÞdτ (4.4.10)
0
Using the same discretization for the interval ½0, T to approximate the inte-
gral in Eq. (4.4.10), we have
X
n1
u_ n ¼ u_ 0 + c2 qrm + c2 qnm
r¼1 (4.4.11)
¼ u_ n1 + c2 qnm
where
c2 ¼ h (4.4.12)
P
n1
Solving Eq. (4.4.11) for qrm and substituting into Eq. (4.4.8) gives
r¼1
Because m 6¼ 0, the coefficient matrix in Eq. (4.4.17) is not singular for suf-
ficient small h and the system can be solved successively for n ¼ 1, 2, … to yield
the solution un and the derivatives u_ n , u€n ¼ qn at instant t ¼ nh T . For n ¼ 1,
the value q0 appears in the right side of Eq. (4.4.17). This quantity can be readily
obtained from Eq. (4.4.1) for t ¼ 0. Thus, we have
q0 ¼ ðp0 cu_ 0 ku 0 Þ=m (4.4.18)
Eq. (4.4.17) can be also written as
Un ¼ AUn1 + bpn , n ¼ 1, 2, …,N (4.4.19)
in which
8 9
< qn =
Un ¼ u_ n (4.4.20a)
: ;
un
2 31 2 3
m c k 0 0 0
6 1 7 6 1 7
6 c1 h 1 7 6 c1 0 1 7
6
A¼6 2 7 6 7 (4.4.20b)
7 6 2 7
4 1 5 4 1 5
c2 1 0 c2 1 0
2 2
2 31
m c k 8 9
6 1 7 <1=
6 c h 1 7
b¼66 2
1 7
7 :0; (4.4.20c)
4 1 5 0
c2 1 0
2
The recurrence formula (4.4.19) can be employed to construct the solution
algorithm. However, the solution procedure can be further simplified. Thus,
applying Eq. (4.4.19) for n ¼ 1, 2, … we have
U1 ¼ AUo + bp1
U2 ¼ AU1 + bp2
¼ AðAUo + bp1 Þ + bp2
(4.4.21)
¼ A2 Uo + Abp1 + bp2
⋯ ¼ ⋯⋯⋯⋯⋯⋯⋯⋯⋯
Un ¼ An Uo + An1 p1 + An2 p2 + ⋯A0 pn b
Obviously, the last of Eq. (4.4.21) gives the solution vector Un at instant
tn ¼ nh using only the known vector U0 at t ¼ 0. The matrix A and the vector
b are computed only once.
Table 4.4.1 presents the algorithm for the numerical implementation of
AEM in pseudocode-type notation so that the reader can write a computer code
in the language of his/her preference.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 173
Example 4.4.1 Response of an SDOF system using the analog equation method
Determine the response the SDOF system in Example 4.2.1 using the AEM.
Solution
The solution is obtained using the program aem_lin.m with Dt ¼ 0:01.
Fig. E4.5 gives the graphical representation of the displacement together with
the error u ðt Þ uex ðt Þ. Moreover, Fig. E4.6 shows the response of the system
under the static load p ¼ t, if 0 t 1 and p ¼ 1, if 1 < t.
@CivilMethod
174 PART I Single-degree-of-freedom systems
x 10−4
0.2 2
u(t) computed
u-uex
u(t) exact 1.5
0.15
1
0.1 0.5
u(t)
0
0.05
–0.5
–1
0
–1.5
–0.05
0 2 4 6 8 10 –2
0 2 4 6 8 10
t
× 10
-4 Displacement × 10
-4 Velocity
6 10
4 5
du(t)
u(t)
2 0
0 –5
0 5 10 0 5 10
t t
× 10
-3 Acceleration Response ratio
2 1.5
0 1
ddu(t)
R(t)
–2 0.5
–4 0
0 5 10 0 5 10
t t
FIG. E4.6 Response of the system in Example 4.4.1 under static load.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 175
@CivilMethod
176 PART I Single-degree-of-freedom systems
We observe that un + 1 can be computed if the two initial values u1 and uo
are known. Eq. (4.5.1) may also be written as
Eq. (4.5.3) is a difference equation of the second order [14] whose solution
can yield un + 1 using the recursive procedure (4.5.2). Because Eq. (4.5.3) is
linear with respect to un1 , un , and un + 1 , it is called a linear difference equa-
tion. If gn ¼ 0 the difference equation is called homogeneous while if gn 6¼ 0 it is
called nonhomogeneous. In general, an equation of the form
rn + k + a1 rn + k1 + a2 rn + k2 + … + ak rn ¼ 0
is also a solution for all n for which the difference equation is defined with
c1 ,c2 ,…, ck being arbitrary constants. Eq. (4.5.6) is the general solution of
the difference equation (4.5.4). The arbitrary constants c1 ,c2 ,…, ck are deter-
mined from the k initial conditions.
If two roots of the polynomial (4.5.5) are complex conjugate, say r1 ¼
a + ib and r2 ¼ a ib, then we can write them in exponential form
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r¼ a2 + b 2 , q ¼ tan 1 ðb=aÞ (4.5.8)
@CivilMethod
Numerical integration of the equation of motion Chapter 4 177
@CivilMethod
Numerical integration of the equation of motion Chapter 4 179
‘im ð1 + eÞ1=e ¼ e
e!0
Using this relation and taking into account that n ¼ tn =h, we obtain for a
given tn
@CivilMethod
180 PART I Single-degree-of-freedom systems
1
un ¼ c1 e2tn + (4.5.18)
2
Evidently, the first term in Eq. (4.5.17) is the exact solution. The second
term is spurious (extraneous) and results from the fact that the first-order dif-
ferential equation is substituted by a second-order difference equation. The
application of the initial conditions would give c2 ¼ 0 if the computations were
exact. In practice, however, errors are introduced, which are mainly due to the
rounding of numbers or the inaccuracy of the starting value. Therefore, the con-
stant c2 is not exactly zero and consequently, a small error is introduced in each
integration step. This is magnified because it is multiplied by the factor
ð1Þn e2tn , which increases exponentially. Because the first term of the solution
(4.5.17) diminishes exponentially, the introduced error due to the spurious solu-
tion dominates the exact solution and leads to a totally wrong result. We
describe a method as unstable if the error increases exponentially with tn .
For the first order differential equations, the one-step integration methods do
not exhibit instability for small values of h. The multistep methods, however,
which lead to difference equations of order greater than one, introduce spurious
solutions and they may be unstable either for all values of h or for a certain
region of the values of h. In order to decide whether a multistep method is sta-
ble, we work as follows.
If the multistep method leads to a difference equation of order k, we find the
roots of the characteristic equation. If ri (i ¼ 1,2, …,kÞ are these roots, the gen-
eral solution will be
One of the solutions, say rn1 , will tend to the actual solution of the differential
equation. The remaining roots are spurious. We will say that a multistep method
is strongly stable if for h ! 0 the spurious roots satisfy the condition
Because we do not know which is the actual solution, the above condition
should apply to all roots ri . Apparently, this condition ensures that the error
diminishes as n increases. On the contrary, the error increases exponentially
if jri j > 1.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 181
where
2 h 2 w2 1 hwx pn h 2
a1 ¼ , a2 ¼ , gn ¼ (4.5.21)
2ð1 + hwx Þ 1 + hwx m ð1 + hwx Þ
The characteristic equation of Eq. (4.5.20) is
r2 2a1 r + a2 ¼ 0 (4.5.22)
whose roots are
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r1, 2 ¼ a1 a21 a2
The type of root depends on the sign of the discriminant D ¼ a21 a2 , which
may be written as
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
s2 s 2 1 x2 s + 2 1 x2
D ðs Þ ¼ , s ¼ hw
4ð1 + sxÞ2
pffiffiffiffiffiffiffiffiffiffiffiffis > 0, the sign of the discriminant depends only on the factor
Because
s 2 1 x2 . Hence, we distinguish the following two cases
qffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi
2 1 x2 T
ðiÞ Dðs Þ > 0 s > 2 1 x 2
or h > ¼ 1 x2 (4.5.23)
w p
qffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi
2 1 x2 T
ðiiÞ Dðs Þ 0 s 2 1 x 2
or h ¼ 1 x 2 (4.5.24)
w p
In case (i), the characteristic equation has two real roots, r1 , r1 , The stability
of the solution requires that jr1 j < 1 and jr2 j < 1. But it can be shown that
jr2 j > 1, hence jr1 j > 1. Consequently, the solution is unstable in this case.
In case (ii), the characteristic equation has two complex conjugate roots
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r1, 2 ¼ a1 i a2 a21 (4.5.25)
This gives
sffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffi 1 sx
j r1 j ¼ j r2 j ¼ a2 ¼ 1 (4.5.26)
1 + sx
@CivilMethod
182 PART I Single-degree-of-freedom systems
@CivilMethod
Numerical integration of the equation of motion Chapter 4 183
2 31
1 2xw w2 8 9
6 7 >
> 1>
6 h2 7 < > =
6 1 7
b ¼ 6 4 0 7 0 (4.5.36b)
6 7 >
> >
4 h 5 : > ;
0
1 0
2
Applying Eq. (4.5.35) for n ¼ 1, 2, … we have
U1 ¼ AUo + b
p1
U2 ¼ AU1 + b
p2
¼ AðAUo + b
p1 Þ + b
p2
(4.5.37)
¼ A2 Uo + Ab
p1 + b
p2
⋯ ¼ ⋯⋯⋯⋯⋯⋯⋯⋯⋯
Un ¼ An Uo + An1 p1 + An2 p2 + …A0 pn b
The matrix A is known as the amplification matrix. The stability of the
method requires that An is bounded. This is true if the spectral radius of A sat-
isfies the condition [15]
rðAÞ ¼ max fjr1 j, jr2 j, jr3 jg 1 (4.5.38)
where ri (i ¼ 1, 2, 3) are the eigenvalues of the matrix A.
Using a symbolic language (here MATLAB) we find
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
4 s2 + 4s x2 1 4 s 2 4s x2 1
r1 ¼ , r2 ¼ , r3 ¼ 0 (4.5.39)
4 + 4sx + s 2 4 + 4sx + s 2
where s ¼ xw.
The type of the roots r1 , r2 depends on the sign of the discriminant
DðxÞ ¼ x2 1. Hence, we distinguish the following two cases
(i) If D ðx Þ > 0, both eigenvalues are real. It can be shown that jr1 j < 1, hence
jr2 j < 1. Therefore, the method is stable
(ii) If DðxÞ 0, the eigenvalues are complex conjugate
pffiffiffiffiffiffiffiffiffiffiffiffi
4 s 2 i4s 1 x2
r1,2 ¼
4 + 4sx + s 2
and
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð4 s 2 Þ2 + 16s2 1 x 2 ðs 2 + 4Þ2 16s 2 x2
j r1 j ¼ j r2 j ¼ ¼ 1 (4.5.40)
ð4 + 4sx + s 2 Þ ½ðs 2 + 4Þ + 4sx 2
The equality is valid for x ¼ 0.
@CivilMethod
184 PART I Single-degree-of-freedom systems
The conclusion is that the AAM is stable without imposing any constraint on
the size of the time step. We say in this case that the integration method is
unconditionally stable.
Essentially, the procedure based on the condition (4.5.38) to prove the sta-
bility of the AAM is not different from that presented in Section 4.5.3, where the
stability results from the response of the difference equation. This is shown in
what it follows.
We write Eq. (4.5.35) for tn ,tn1 , tn2
Un AUn1 ¼ b
pn
Un1 AUn2 ¼ b
pn1 (4.5.41)
Un2 AUn3 ¼ b
pn2
or in matrix form
8 9
2 3> Un > 8 9
I A 0 >
0 < >
= < pn =
4 0 I A 0 5 Un1
¼ b pn1 (4.5.42)
> Un2 > : ;
0 0 I A >
: >
; pn2
Un3
Eq. (4.5.42), beside the displacements un ,un1 ,un2 , un3 , contains the
velocities u_ n , u_ n1 , u_ n2 , u_ n3 and the accelerations u€n , u€n1 , u€n2 , u€n3 . Reor-
dering these equations and eliminating the velocities and accelerations yield the
equation
un 2a1 un1 + a2 un2 + a3 un3 ¼ c1 pn + c2 pn1 + c3 pn2 (4.5.43)
where
4 w2 h 2 4 4sx + w2 h 2
a1 ¼ , a2 ¼ , a3 ¼ 0 (4.5.44)
4a 4a
1 + xwh 1 xwh
c1 ¼ , c2 ¼ , c3 ¼ 0 (4.5.45)
a a
where
w2 h 2
a ¼ 1 + xwh +
4
Eq. (4.5.43) is a difference equation whose characteristic equation is
r3 2a1 r2 + a2 r ¼ 0 (4.5.46)
The roots of Eq. (4.5.46) are
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
4 s 2 + 4s x 2 1 4 s 2 4s x2 1
r1 ¼ , r2 ¼ , r3 ¼ 0 (4.5.47)
4 + 4sx + s 2 4 + 4sx + s 2
that is, they are identical to the eigenvalues of the matrix A.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 185
@CivilMethod
186 PART I Single-degree-of-freedom systems
or
un ¼ r n ðc1 sin nq + c2 cos nqÞ
(4.6.4)
n + c2 cos wt
¼ r n ðc1 sin wt nÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where r ¼ a 2 + b2 , q ¼ tan 1 ðb=a Þ, a ¼ Reðr1 Þ, b ¼ Imðr1 Þ, w ¼ q=h,
tn ¼ nh.
The corresponding exact solution is
qffiffiffiffiffiffiffiffiffiffiffiffi
un ¼ exwtn ðc1 sin wD tn + c2 cos wD tn Þ, wD ¼ 1 x2 , tn ¼ nh (4.6.5)
Comparison of Eqs. (4.6.4), (4.6.5) could show the accuracy of the numer-
ical scheme. Thus, the period elongation is
pffiffiffiffiffiffiffiffiffiffiffiffi
T T h 1 x2
PE ¼ ¼ 1 (4.6.6)
T q
For the amplitude decay, we can define an equivalent damping ratio x from
the relation
n
r n ¼ exwt ¼ exqn (4.6.7)
which gives
x ¼ ln r=q (4.6.8)
The difference Dx ¼ x x can be employed as a measure for the amplitude
decay. The dependence of the period elongation and amplitude decay on h=T is
shown in Figs. 4.6.1 and 4.6.2, respectively. Obviously, for small values of h=T
the scheme is accurate. Note that for x ¼ 0 it is jr2 j ¼ jr3 j ¼ r ¼ 1 and Eq. (4.6.8)
yields x ¼ 0. That is, there is no amplitude decay.
0.25
x=0
0.2 x = 0.1
Period elongation %
x = 0.2
0.15
0.1
0.05
0
0 0.05 0.1 0.15 0.2 0.25 0.3
h/T
FIG. 4.6.1 Period elongation versus h=T for different values x.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 187
−3
x 10
2
x=0
0 x = 0.1
x = 0.2
−2
Δx
−4
−6
−8
0 0.05 0.1 0.15 0.2 0.25 0.3
h/T
FIG. 4.6.2 Amplitude decay Dx ¼ x x versus h=T for different values of x.
Solution
The characteristic equation of Eq. (1) is
2 + s 2 r3 5r2 + 4r 1 ¼ 0 (2)
pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1=3
b ¼ 720a + 108a 2 + 1000 + 12 3a 104a + 100 + 27a 2 (5)
The stability condition requires that jr1 j < 1 and jr2 j ¼ jr3 j < 1. This
condition is satisfied, as is shown in Fig. E4.7.
@CivilMethod
188 PART I Single-degree-of-freedom systems
1
|r1|
0.9
|r2|
0.8 |r3|
0.7
0.6
0.5
|r2|=|r3|
0.4
0.3
0.2
0.1
| r1| (spurious)
0
1 2 3 4 5 6 7 8 9 10
h/t
FIG. E4.7 Houbolt’s method. Roots of the characteristic equation.
4.7 Problems
Problem P4.1 Find the solution of the IVPs for the difference equations
un + 1 3un ¼ 5, u0 ¼ 1
un + 2 4un + 1 + 3un ¼ 2n , u0 ¼ 0, u1 ¼ 1
Hint: The particular solution of (ii) will be sought in the form gn ¼ 2n c,
where c is a constant.
Problem P4.2 A SDOF system is subjected to pulse the load
t
pðt Þ ¼ p0 1 1 + et=t1 , 0 t t1
t1
pðt Þ ¼ 0, t > t1
Study the response of the system using
(i) The central difference method.
(ii) The average acceleration method.
(iii) The analog equation method.
Compare the results with the exact solution.
Data: u ð0Þ ¼ u_ ð0Þ ¼ 0, m ¼ 50kNm1 s2 , x ¼ 0:05, w ¼ 5s1 , p0 ¼ 100kN,
t1 ¼ 0:5s and ttot ¼ 10s.
Problem P4.3 A SDOF system is subjected to the piecewise linear load pðt Þ
shown in Fig. P4.3. Use the three discussed numerical methods to establish
its response.
@CivilMethod
Numerical integration of the equation of motion Chapter 4 189
Compare the results with the exact solution. Data: u ð0Þ ¼ u_ ð0Þ ¼ 0,
m ¼ 50kN m1 s2 , x ¼ 0:05, w ¼ 5s1 , ti ¼ 0:01i, pi ¼ 20 1 + ð1Þi ði + 5Þ=
ði + 1Þ , i ¼ 0, 2, …,100.
@CivilMethod
Chapter 5
Nonlinear response:
Single-degree-of-freedom
systems
Chapter outline
5.1 Introduction 191 5.4 The analog equation
5.2 The central difference method 195 method 203
5.3 The average acceleration 5.5 Problems 211
method 197 References and further reading 215
5.1 Introduction
The equation of motion of a vibrating system expresses the equilibrium condi-
tion of all forces applied to the system, namely the external excitation force, the
inertial force, the damping force, and the elastic force. The equilibrium condi-
tion reads
fI ðt Þ + fD ðt Þ + fS ðt Þ ¼ pðt Þ (5.1.1)
The forces fI , fD , and fS depend on the physical properties of the system.
In the systems we analyzed, the physical properties are not time-dependent
and the dependence of these forces on the cause that produces them is linear,
that is,
fI ¼ m u€ (5.1.2a)
fD ¼ cu_ (5.1.2b)
fS ¼ ku (5.1.2c)
where m, c,k are constant quantities. Systems with such a physical response are
referred to as linear systems.
In general, however, Eq. (5.1.2a)-(5.1.2c) may be of the form
fI ¼ m ðt Þu€ (5.1.3a)
_ tÞ
fD ¼ fD ðu, u, (5.1.3b)
_ tÞ
fS ¼ fS ðu, u, (5.1.3c)
That is, the mass may vary with time and the damping and elastic forces may
be nonlinear functions of u, u, _ t Þ, fS ðu, u,
_ and t, that is, fD ðu, u, _ t Þ. In this case,
Εq. (5.1.1) takes the form
_ t Þ + fS ðu, u,
m ðt Þu€ + fD ðu, u, _ t Þ ¼ pðt Þ (5.1.4)
which is a nonlinear differential equation of the second order. A dynamic sys-
tem, whose response is described by Eq. (5.1.4), is referred to as nonlinear.
Although systems with variable mass are not unusual [1], our discussion will
be limited to systems with constant mass. Besides, the forces fD and fS will
be considered of the form fD ðu_ Þ and fD ðu Þ. Thus, Eq. (5.1.4) becomes
m u€ + fD ðu_ Þ + fS ðu Þ ¼ pðt Þ (5.1.5)
We distinguish two types of nonlinearity: the geometric nonlinearity, which
is due to large displacements implying large deformations of the structure, and
the material nonlinearity, which is due to nonlinear constitutive equations (e.g.,
hyperelastic or elastoplastic materials). Of course, both types of nonlinearity
can simultaneously characterize the response of a system.
The analytical solution of the nonlinear equations of motion is a difficult and
complicated mathematical problem. Exact solutions are available only for a few
cases and for differential equations of a specific form [2, 3]. The existing solu-
tions aim rather at a qualitative study of the response of the system described by
a nonlinear equation than at offering a computational means for practical ana-
lyses. The knowledge of the nonlinear response of the single-degree-of-freedom
(SDOF) systems comes from approximate methods, and mainly from numerical
methods. Therefore, the recourse to numerical methods to solve the nonlinear
equations of motion is inevitable. The step-by-step methods play a dominant
role. The Runge-Kutta methods, usually employed for the solution of nonlinear
equations, belong to these methods [4].
The dynamic response of nonlinear systems can be studied effectively by
demanding the fulfillment of equation motion (5.1.5) at discrete time instants
Dt apart by the use of the step-by-step integration methods we discussed in
Chapter 4. These methods as developed for nonlinear equations of motion
are presented directly below while for the analytical methods, the reader is
advised to look in the vast related literature [2, 3, 5].
Example 5.1.1 Systems with a geometrical nonlinearity
Derive the equation of motion of the system shown in Fig. E5.1. The supports at
A,B and the interconnection at C are hinges. The mass m at C is concentrated.
The system is set to motion by the initial conditions u0 , u_ 0 and/or the vertical
external force pðt Þ. The bars are assumed massless.
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 193
l l
C
A φ B
u(t )
E, A E, A
m
p(t )
fI
S fS S
p(t )
FIG. E5.1 SDOF system with nonlinear response.
Solution
During the motion, the force pðt Þ, the inertial force fI , and the elastic forces of
the bars are in equilibrium
The inertial force is given by the relation
fI ¼ m u€ (1)
The total elastic force is caused by the elongation of the bars and is given by
fS ¼ 2S sin f (2)
where
EA
S¼ d (3)
l
E is the modulus of elasticity of the material of the bar, A its cross-sectional area,
and d the elongation of the bars.
The elongation of the bars at time t is
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
d¼ l 2 + u2 l (4)
and
u
sin f ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5)
l 2 + u2
@CivilMethod
194 PART I Single-degree-of-freedom systems
Hence, the elastic force in Eq. (2) by virtue of Eqs. (3)–(5) is expressed as
2 3
u6 1 7
fS ðu Þ ¼ 2EA 41 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi5 (6)
l 2
1 + ðu=l Þ
Obviously, the equation of motion is nonlinear. Fig. E5.2 shows the graph
of fS ðu Þ. Because d 2 fS =du 2 > 0 the curve is concave upward. In this system,
the slope kT increases continually, which implies that the elastic force
increases with increasing u. In this case, we say that the system exhibits hard-
ening, in contrast to other systems that exhibit softening. In the latter systems,
the curve fS ðu Þ is concave downward (d 2 fS =du 2 < 0), that is, the slope
kT decreases continually, which implies that elastic force decreases with
increasing u, for example, a system with stiffness fS ¼ 40u u 3 exhibits soft-
ening; see Fig. E5.3.
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 195
@CivilMethod
196 PART I Single-degree-of-freedom systems
where k ¼ sin 2 ðq0 =2Þ and T0 is the quarter of the period; sn represents the
sn-Jacobean elliptic function [7].
Solution
The response of the pendulum for l ¼ g, q0 ¼ 0:40p, and q_ 0 ¼ 0 is obtained using
the program centr_diff_nlin.m with Dt ¼ 0:01. It is shown in Fig. E5.5 as com-
pared with the exact one.
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 197
t t
(a) (b)
FIG. 5.3.1 Tangent and secant of (a) damping and (b) stiffness curve.
DfS ¼ fS ðt + Dt Þ fS ðt Þ
¼ fS ðu + Du Þ fS ðu Þ
(5.3.4)
df 1 d 2 fS
¼ S Du + ðDu Þ2 + ⋯
du 2 du 2
For small values of Dt, the quantities Du, Du_ are also small. Thus, neglect-
ing the nonlinear terms in Eqs. (5.3.3), (5.3.4), we obtain
DfD cT Du_ (5.3.5)
DfS kT Du (5.3.6)
Obviously, cT and kT express the slope of the tangent to the curves fD ðu_ Þ
and fS ðu Þ, respectively, at time t.
Referring to Fig. 5.3.1a, we have
fD ðt + Dt Þ ¼ fD ðt Þ + DfD
However, the exact value of DfD is
DfD ¼ Du_ tan fc
where fc is the angle of the secant. Hence, approximating DfD by Eq. (5.3.5)
introduces the error (see Fig. 5.3.1a)
eD ¼ ðcT tan fc ÞDu_ (5.3.7)
because cT is the slope of the tangent.
Similarly, the use of kT to approximate DfS introduces the error (see
Fig. 5.3.1b)
eS ¼ ðkT tanfk ÞDu (5.3.8)
The errors eD and eS cannot be avoided because u_ ðt + Dt Þ and u ðt + Dt Þ
are not known at instant t + Dt. However, as we will show, they can be kept
under a given bound, which specifies the accuracy of the solution procedure.
By virtue of Eqs. (5.3.5), (5.3.6), Eq. (5.3.2) is written in incremental form
mDu€ + cT Du_ + kT Du ¼ Dp (5.3.9)
The previous equation is of the form (4.3.11). Hence, the AAM is suitable to
solve it. Thus, using Eqs. (4.3.9), (4.3.10) to express Du€ and Du,
_ we obtain
2
Du_ ¼ Du 2u_ ðt Þ (5.3.10)
Dt
4
Du€ ¼ ½Du Dt u_ ðt Þ 2u€ðt Þ (5.3.11)
Dt 2
Then Eq. (5.3.9) becomes
2cT 4m 4m
kT + + 2 Du ¼ Dp + + 2cT u_ ðt Þ + 2m u€ðt Þ (5.3.12)
Dt Dt Dt
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 199
or
k ∗ Du ¼ Dp∗ (5.3.13)
where
2cT 4m
k ∗ ¼ kT + + 2 (5.3.14)
Dt Dt
and
4m
Dp∗ ¼ Dp + + 2cT u_ n + 2m u€n (5.3.15)
Dt
The value of Du obtained from Eq. (5.3.13) is used in Eq. (5.3.10) to eval-
_ Then we obtain
uate Du.
un + 1 ¼ un + Du (5.3.16a)
u_ n + 1 ¼ u_ n + Du_ (5.3.16b)
where
DfS DfD
ksec ¼ , csec ¼ (5.3.18)
Du Du_
and a is a small specified number defining the upper bound of the error, for
example, a ¼ 0:01. If it is
@CivilMethod
200 PART I Single-degree-of-freedom systems
When the damping force depends linearly on the velocity, then the response
is governed by Eq. (5.2.1). In this case, the previously presented incremental
method can be improved by employing an iterative procedure within each step,
which minimizes the error introduced by the tangent stiffness kT .
The starting point is Eq. (5.3.13), which we write as
k ∗ du ð1Þ ¼ Dp∗ (5.3.20)
The quantity du ð1Þ is the first approximation to Dun within the time step
from tn to tn + 1 . That is
Dunð1Þ ¼ du ð1Þ (5.3.21)
The index n denotes the number of the step that brings us from the displace-
ment un to un + 1 . Fig. 5.3.2 presents the graph of the function p∗ ðu Þ. Apparently,
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 201
the displacement du ð1Þ resulting from Eq. (5.3.20) decreases Dp∗ by dpð1Þ .
Hence, there is a remaining force
DF ð2Þ ¼ Dp∗ dpð1Þ (5.3.22)
which must be equilibrated. The change dpð1Þ is computed using Eq. (5.3.20), if
the tangential slope kT is replaced with the slope of the secant
" #
ð1Þ fS un + du ð1Þ fS ðun Þ 2c 4m
dp ¼ + + du ð1Þ
du ð1Þ Dt Dt 2
2c 4m
¼ fS un + du ð1Þ fS ðun Þ + + 2 du ð1Þ
Dt Dt
which by virtue of Eq. (5.3.14) becomes
dpð1Þ ¼ fS un + du ð1Þ fS ðun Þ + ðk ∗ kT Þdu ð1Þ (5.3.23)
The force DF ð3Þ produces the additional displacement du ð3Þ , which is com-
puted from the relation
k ∗ du ð3Þ ¼ DF ð3Þ (5.3.28)
Thus, the new approximation of Dun is
Dunð3Þ ¼ Dunð2Þ + du ð3Þ (5.3.29)
Consequently, for the i + 1 approximation it is
DF ði + 1Þ ¼ DF ðiÞ dpðiÞ (5.3.30)
dpðiÞ ¼ fS un + DunðiÞ fS un + Dunði1Þ + ðk ∗ kT Þdu ðiÞ (5.3.31)
@CivilMethod
202 PART I Single-degree-of-freedom systems
k ∗ du ði + 1Þ ¼ DF ði + 1Þ (5.3.32)
Dunði + 1Þ ¼ DunðiÞ + du ði + 1Þ (5.3.33)
Note that for i ¼ 1 it must be set Dunð0Þ ¼ 0 and DF ð1Þ ¼ Dp∗ .
The iteration procedure is terminated after I iterations, if
du ði + 1Þ
e (5.3.34)
DunðI Þ
where e is a specified small number. Then we assume that the convergence has
been achieved. The value Dun ¼ DunðI Þ is considered exact and it is used to
compute un + 1 and u_ n + 1 , u€n + 1 . Subsequently, the procedure continues to
the next step. The iterative procedure within the time step from tn to tn + 1 is
summarized in Table 5.3.2. This procedure is known as the modified
Newton-Raphson method.
5. du ði + 1Þ ¼ DRði + 1Þ =k ∗
6. If du ði + 1Þ =DunðI Þ > e set i ¼ i + 1 and go to B.2
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 203
solution procedure are presented in Table 5.4.1. Adhering to the steps of this
table, a computer program called aem_nlin.m has been written in MATLAB
for the numerical integration of the nonlinear equation of motion using the
AEM. The program is available on this book’s companion website. It computes
the displacement u ðt Þ, the velocity u_ ðt Þ, and the acceleration u€ðt Þ.
Example 5.4.1 The Duffing equation. AEM solution
Use the AEM to solve the IVP for the Duffing equation
u€ + 0:2u_ + u + u 3 ¼ pðt Þ (1)
u ð 0Þ ¼ 0 (2a)
u_ ð0Þ ¼ 1 (2b)
For
pðt Þ ¼ e0:1t ½ð0:01sin t 0:2cost sin t Þ 0:2ð0:1 sin t cos t Þ
+ sin t + e0:2t ð sin t Þ3
Eq. (1) admits an exact solution uexact ðt Þ ¼ e0:1t sin t.
Solution
The solution is obtained using the program aem_nlin.m with Dt ¼ 0:01. The
graph of the solution is shown in Fig. E5.6 as compared with the exact one.
1
Computed
0.8 Exact
error X103
0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
0 5 10 15 20 25
t
FIG. E5.6 Solution u and error u uexact in Example 5.4.1.
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 205
which is solved for the data of the problem to yield ust ¼ 0:25m.
The external force is the constant weight of the body, that is, pðt Þ ¼ mg,
while the elastic force
fS ðust + u Þ (2)
where u ¼ u ðt Þ denotes the additional displacement due to the dynamic
response. Thus, the equation of motion becomes
m u€ + fS ðust + u Þ ¼ mg (3)
with initial conditions
u0 ¼ 0:05m, u_ 0 ¼ 0 (4)
It should be noted that due to the nonlinearity of the elastic force, the super-
position of the displacements does not apply. The computed response of the sys-
tem with Dt ¼ 0:01 is shown in Fig. E5.7.
(ii) In this case the equation of motion reads
m u€ + fS ðu Þ ¼ mg (5)
with initial conditions
u0 ¼ 0, u_ 0 ¼ 0 (6)
where u ¼ u ðt Þ denotes the total dynamic displacement from the undeformed
position. The computed response of the system with Dt ¼ 0:01 is shown in
Fig. E5.8.
(iii) In this case, the IVP becomes
m u€ + fS ðu Þ ¼ mg + 10sin Wt (7)
u0 ¼ 0, u_ 0 ¼ 0 (8)
The computed response with Dt ¼ 0:01 is shown in Fig. E5.9 for the ratio
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
W=w ¼ 0:1, where w ¼ EA=lm . In all cases, the results are compared with
those obtained by the CDM
@CivilMethod
206 PART I Single-degree-of-freedom systems
u(t)
t
FIG. E5.7 Response in Example 5.4.2 (i).
u(t)
t
FIG. E5.8 Response in Example 5.4.2 (ii).
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 207
@CivilMethod
208 PART I Single-degree-of-freedom systems
(a) (b)
FIG. E5.11 System with a material nonlinearity in Example 5.4.4.
Solution
It is
fSy 69:68
uy ¼ ¼ ¼ 0:013m ¼ yield displacement
k 5360:00
um ¼ maximum displacement where the velocity changes sign
uR ¼ um uy ¼ remaining plastic deformation
The equation of motion is
m u€ + fS ðu Þ ¼ pðt Þ
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 209
@CivilMethod
TABLE E5.1 Numerical solution of elastoplastic system in
Example 5.4.4.—cont’d
u ðt Þ (cm)
Dt ¼ 0:01 Dt ¼ 0:001
t (s) AEM AAM AEM AAM
0.35 1.134 1.234 1.177 1.175
0.40 0.378 0.487 0.423 0.421
0.45 0.353 0.262 0.312 0.313
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 211
5.5 Problems
Problem P5.1 Show that the motion of the simple pendulum is governed by
the IVP
g
q€ + sin q ¼ 0, qð0Þ ¼ q0 , q_ ð0Þ ¼ q_ 0 (1)
l
Solve the equation of motion numerically when g=l ¼ 1, q0 ¼ 0:1p, q_ 0 ¼ 0
and compute the period T of the pendulum. Give the graphical representation
of the function T ¼ T ðq0 Þ, 0:1 q0 1 for the two time steps Dt ¼ 0:1 and
Dt ¼ 0:0001. Compare with the exact expression
sffiffiffi Z
l p=2 df sin ðq=2Þ
T ðq 0 Þ ¼ 4 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi , k ¼ sin ðq0 =2Þ, sin f ¼ (2)
g 0 1 k 2 sin 2 f sin ðq0 =2Þ
@CivilMethod
212 PART I Single-degree-of-freedom systems
Problem P5.3 Study the response of the system shown in Fig. P5.3, when
(i) u ð0Þ ¼ 0:05m, u_ ð0Þ ¼ 0, pðt Þ ¼ 0. The initial displacement will be taken
from the position of the static equilibrium.
(ii) u ð0Þ ¼ u_ ð0Þ ¼ 0, pðt Þ ¼ mg ð1 t=t1 ÞH ðt1 t Þ:pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(iii) u ð0Þ ¼ u_ ð0Þ ¼ 0, P ðt Þ ¼ 15sin wt,
where w ¼ kmin =m .
(iv) Plot of the dependence of the period T as a function of u0 , T ¼ T ðu0 Þ, if
u_ ð0Þ ¼ 0 and pðt Þ ¼ 0. The initial displacement will be taken from the
position of the static equilibrium.
Data: t1 ¼ 1s, ttot ¼ 5s, m ¼ 10kN m1 s2 , E ¼ 2:1 108 kN=m2 , A ¼ 1cm2 ,
L ¼ 4:0m, and g ¼ 9:81 ms2 . The cables are assumed massless to avoid the
sag due to self-weight [10–12].
Problem P5.4 Study the response of the system shown in Fig. P5.3 if the elastic
force is approximated by the first three terms of its Tailor series. Compare the
results with those in Problem P5.3 when pðt Þ ¼ mg ð1 t=t1 ÞH ðt1 t Þ and
u ð0Þ ¼ u_ ð0Þ ¼ 0.
Problem P5.5 The response of a nonlinear system is governed by the following
equation, known as the van der Pol equation
u€ m 1 u 2 u_ + u ¼ 0
Plot the solution u ðt Þ, if u ð0Þ ¼ 0, u_ ð0Þ ¼ 0:1 and for (i) m ¼ 0:2 (ii) m ¼ 1:2.
Problem P5.6 The horizontal beams of the frame in Fig. P5.6 have negligible
mass and are flexible while the shear walls having a uniform mass density with
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 213
specific weight g are assumed rigid. Their support on the ground is elastic and it
is expressed by the relation
1
MR ¼ CR f f2 (a)
4
The system is set in motion by the initial conditions u0 ¼ 5cm, u_ 0 ¼ 0. Plot
the dependence of the ratio T=T0 (T0 is the period of the structure resulting
when the nonlinear term in Eq. a is ignored) on the initial displacement and dis-
cuss the influence of the nonlinearity of the elastic supports on the response of
the structure. Data: a ¼ 5m, g ¼ 24kN=m3 , E¼2.1107kN/m2, CR ¼ EI =5a,
and cross-sectional dimensions of the beam 0:20 0:40m2 .
Problem P5.7 The buoy of Fig. P5.7 consists of two massless cones with a base
dimeter 2R and a height h. A concentrated mass m attached at the bottom of the
body keeps the buoy floating at the position shown in the figure. Study the
@CivilMethod
214 PART I Single-degree-of-freedom systems
vertical motion of the buoy if it is displaced vertically downward from the equi-
librium position by u0 . Data: m ¼ 10kNm1 s2 h ¼ 5m, R ¼ 2m,
u ð0Þ ¼ 0:30m, u_ ð0Þ ¼ 0, and specific weight of the liquid g ¼ 2kN=m3 .
Problem P5.8 The water tower of Fig. P5.8a is subjected to the blast load of
Fig. P5.8b. The response of the structuren is elastoplastic.
h The
iorestoring force in
the elastic branch is given by fS ¼ ku 1 + 1= 1 + ð10u Þ6 =2 (Fig. P5.8c).
The structure is modeled by a SDOF system. Study the response of the structure
in the interval of t 0 ¼ 1s. Data: m ¼ 50kNm1 s2 , k ¼ 2000kN=m, x ¼ 0:07,
p0 ¼ 25kN, t1 ¼ 0:1s, and yield displacement uy ¼ 0:1m.
(b)
(a)
{ }
(c)
FIG. P5.8 Water tower in problem P5.8.
@CivilMethod
Nonlinear response: Single-degree-of-freedom systems Chapter 5 215
@CivilMethod
Chapter 6
6.1 Introduction
Structural systems are often excited by the motion of their support. The
response of a structure to support excitation is dynamic even though no external
dynamic loads act on it. The seismic motion of the ground represents a typical
example of support excitation of structures. The study of the response of struc-
tures to earthquake-induced motion is a specific but very important subject of
structural dynamics. It is discussed in depth in books on earthquake engineering
as well as in books on structural dynamics, preparing engineers to design struc-
tures for earthquake-induced motion [1,2]. This book treats the dynamic
response of structures when the excitation force is known. Therefore, the dis-
cussion in this chapter is limited only to the study of the dynamic response
of the SDOF system due to support excitation. Besides, some basic concepts such
as the response spectrum concept, which facilitates the dynamic analysis of struc-
tures due to ground motion, are presented. The general problem of the support
excitation of structures will be examined later when the MDOF (multi-degree-
of-freedom) systems are studied. The transmission of vibrations from the
structure to the fundament and vice versa are also discussed. Illustrative examples
analyzing the response of SDOF systems due to ground motion are presented. The
pertinent bibliography with recommended references for further reading is also
included. The chapter is enriched with problems to be solved.
utot
m ug u
Fixed axis of reference
c m c
k k
(a) ug (t ) (b)
m
fI mutot fD cu
fS ku
(c)
FIG. 6.2.1 SDOF system subjected to ground motion (a), Dynamic model (b), Forces on the
free body (c).
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 219
max ug,tt=258.59
ug,tt (cm/s2)
t
FIG. 6.2.2 Accelerogram from Athens earthquake, Sept. 7, 1999 (Recorded by ITSAK, Α399-1.
V2, longitudinal component, max u€g ¼ 258:59cm=s2 ).
@CivilMethod
220 PART I Single-degree-of-freedom systems
Apparently, Eq. (6.2.7b) states that the deformation of the system due to
given ground acceleration u€g ðt Þ depends only on the natural frequency w, hence
on the natural period T ¼ 2p=w, and on the damping ratio x, that is,
u ¼ u ðt, T , x Þ. Consequently, two systems with the same natural period T
and the same damping ratio x will undergo the same displacement u ðt Þ under
the same ground motion, in spite of the fact that the two systems may have dif-
ferent masses or different stiffnesses.
The negative sign in the effective load pðt Þ ¼ m u€g ðt Þ affects only the
direction of the displacement and not its magnitude. In practice, this has little
significance inasmuch as the engineer is usually interested in the maximum
absolute value of u ðt Þ. Therefore, the sign can be omitted in this case. This
assumption allows us to write the Duhamel integral in the forma
1
u ðt Þ ¼ U ðt Þ (6.2.8)
wD
where
Z t
U ðt Þ ¼ u€g ðτÞexwðtτÞ sin wD ðt τÞdτ (6.2.9)
0
For a given ground motion and a fixed damping ratio, we can evaluate the
largest absolute value of the function U ðt Þ, hence of u ðt Þ, for an interval of
values of the natural period T of the damped SDOF system and plot the curves
u ðT , xÞ ¼ max t ju ðt, T , xÞj for discrete values of the damping ratio x. Fig. 6.2.3
shows the curves u ðT , xÞ ¼ U ðT , x Þ=wD , 0 < T 2 for different values of x,
when the ground motion is induced by the Athens earthquake, Sept. 7, 1999.
The respective accelerogram is shown in Fig. 6.2.2. The curves u ðT , x Þ were
obtained by direct solution of Eqs. (6.2.7a), (6.2.7b). The solution can be
obtained using either the analytic solution presented in Section 3.5.4 or numer-
ically using any of the methods presented in Chapter 4.
pffiffiffiffiffiffiffiffiffiffiffiffi
a. In Eqs. (6.2.8) and (6.2.9), we can set wD ¼ w 1 x 2 w, because in real structures the value of
the damping ratio is small (x ¼ 3% 15%, hence x2 ≪1) and the error due to this approximation is
much smaller than that due to the uncertainty of the determination of u€g ðt Þ.
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 221
u T
T (s)
FIG. 6.2.3 Deformation response spectrum u ðT , x Þ for Athens earthquake, Sept. 7, 1999.
@CivilMethod
222 PART I Single-degree-of-freedom systems
T (s)
FIG. 6.2.4 Response spectrum of the relative velocity u_ ðT , xÞ for the Athens Earthquake,
Sept. 7, 1999.
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 223
T (s)
FIG. 6.2.5 Response spectrum of the total acceleration u€tot ðT , xÞ for the Athens earthquake,
Sept. 7, 1999.
For small values of the damping ratio (say, 0 x 0:15), we may set x 2 0.
Thus, the last two equations become
Qo ¼ mwU ðt Þ (6.2.16)
max Qo ¼ mwU ðT , x Þ (6.2.17)
We observe that the quantity wU ðT , x Þ in Eq. (6.2.17) has dimensions of
acceleration. In earthquake engineering, this quantity is designated by
Spa ðT , x Þ and it is known as spectral pseudoacceleration. It is an important
quantity because it allows direct evaluation of the maximum elastic force (base
shear force) in the SDOF system from the graph of Spa ðT , xÞ.
The quantity Spv ðT , x Þ ¼ U ðT , x Þ has dimensions of velocity and it is
known as spectral pseudovelocity. Actually, the quantities Spv ðT , x Þ,
Spa ðT , x Þ are different from u_ ðT , xÞ and u€tot ðT , x Þ. Therefore, they should
not be confused. Nevertheless, it is Spa ðT , xÞ ¼ u€tot ðT , xÞ, if x ¼ 0. Indeed,
Eq. (6.2.3) for x ¼ 0 becomes
u€tot ¼ w2 u ¼ wU ðt Þ (6.2.18)
from which we obtain
u€tot ðT , x Þ ¼ Spa ðT , x Þ (6.2.19)
The deviation of the pseudoacceleration from the extreme value of the total
acceleration is small for small values of x (say 0 x 0:1). Thus we may set
Spa ðT , x Þ u€tot ðT , x Þ. This is shown in Fig. 6.2.6.
@CivilMethod
224 PART I Single-degree-of-freedom systems
(utot),tt(T,0)
Spa(T,0)
T (s)
FIG. 6.2.6 Response spectra Spa ðT , x Þ and u€tot ðT , x Þ, ðx ¼ 0Þ.
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 225
@CivilMethod
226 PART I Single-degree-of-freedom systems
Sd
(a)
(b)
Spa
(c)
FIG. 6.2.7 Response spectra for the Athens earthquake, Sept. 7, 1999 (x ¼ 0:1).
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 227
5m
z
2.5 m
x utot
1m P ug u
m c
10 m
k
x
(a) (b)
FIG. E6.1 Silo in Example 6.2.1
@CivilMethod
228 PART I Single-degree-of-freedom systems
Hence
c ¼ 2mwx ¼ 2 40:6 5:57 0:04265 ¼ 19:3kNm1 s
qffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
wD ¼ w 1 x 2 ¼ 5:57 1 0:042652 ¼ 5:56s1
2p
to ¼ 6T ¼ 6 ¼ 6:78s
wD
t (s)
FIG. E6.2 Time history of the response ratio Rðt Þ in Example 6.2.1.
@CivilMethod
230 PART I Single-degree-of-freedom systems
and
12EI k
max Q ¼ 3
max u ¼ max uII ¼ 1:781kN (6a)
h 4
h
max M ¼ max Q ¼ 8:907kNm (6b)
2
(a) (b)
FIG. E6.3 SDOF system in Example 6.3.1
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 231
Solution
The function representing the ground motion is shown in Fig. E6.3b. The prob-
lem can be solved by the following two ways:
(a) Formulating the equation of motion in terms of the total displacement, that
is, Eq. (6.3.3). This yields
m u€tot + ku tot ¼ ku 0 (1)
This implies that an effective constant excitation force pðt Þ ¼ ku 0 is
suddenly applied. The solution of Eq. (1) is obtained from Eq. (3.4.3)
for p0 ¼ ku 0 . Hence
pffiffiffiffiffiffiffiffiffi
utot ðt Þ ¼ u0 ð1 cos wt Þ, w ¼ k=m (2)
We observe that
ðutot Þmax ¼ 2u0
The relative displacement is
u ðt Þ ¼ utot ug ¼ u0 cos wt (3)
which yields an elastic force
fS ¼ ku ¼ ku 0 cos wt (4)
and
max jfS j ¼ ku 0 (5)
(b) Formulating the equation of motion in terms of the relative displacement
u ðt Þ. In this case, we have
ug ð0Þ ¼ u0 , u_ g ð0Þ ¼ 0, u€g ð0Þ ¼ 0
Consequently
pðt Þ ¼ 0, u ð0Þ ¼ utot ð0Þ ug ð0Þ ¼ u0 , u_ ð0Þ ¼ u_ tot ð0Þ u_ g ð0Þ ¼ 0
and the equation of motion becomes
m u€ + ku ¼ 0 (4)
with initial conditions u ð0Þ ¼ u0 , u_ ð0Þ ¼ 0.
The solution of Eq. (4) is obtained from Eq. (2.2.13) as
u ðt Þ ¼ u0 cos wt (5)
which is identical with that given by Eq. (3).
Example 6.3.2 The supports of the columns of the one-story frame of Fig. E6.4
are subjected to the displacements ug1 ¼ uo sin ðwt qÞ and ug2 ¼ uo sin wt.
Determine the equation of motion of the structure and give the expressions
of the relative displacement u ðt Þ and the stress resultants Q ðt Þ, M ðt Þ at the
top cross-sections of the columns. The dead load of the rigid beam is included
@CivilMethod
232 PART I Single-degree-of-freedom systems
utot (t ) utot (t )
p
EI
h
35 × 70 40 × 80
ug 1 ug 2
L
FIG. E6.4 One-story frame in Example 6.3.2
in the load p. The material of the columns is reinforced concrete. Data: specific
weight of concrete g ¼ 24kN=m3 , u0 ¼ 0:02m, x ¼ 0:05, w ¼ 2:5rad=s,
u ð0Þ ¼ 0, u_ ð0Þ ¼ 0, L ¼ 15m, h ¼ 7m, E ¼ 2:1 107 kN=m2 , and
p ¼ 200kN=m. The mass of the columns is assumed lumped at their ends.
Solution
The system has one degree of freedom. The equation of motion with respect to
the total displacement utot ðt Þ is
m u€tot + c1 u_ tot u_ g1 + c2 u_ tot u_ g2 + k1 utot ug1 + k2 utot ug2 ¼ 0 (1)
or
m u€tot + cu_ tot + ku tot ¼ pðt Þ (2)
where
c ¼ c1 + c2 , k ¼ k1 + k2 (3)
p ¼ c1 u_ g1 + c2 u_ g2 + k1 ug1 + k2 ug2 (4)
The mass of the system is
200 15 + ð0:35 0:70 + 0:40 0:80Þ ð7=2Þ 24
m¼ ¼ 310:6kNm1 s2
9:81
The stiffness of the columns
12EI 1 12 2:1 107 0:70 0:353
k1 ¼ ¼ ¼ 1837:5kN=m
h3 12 73
3EI 2 3 2:1 107 0:80 0:403
k2 ¼ ¼ ¼ 783:7kN=m
h3 12 73
Hence the stiffness of the system is
k ¼ k1 + k2 ¼ 2621:2kN=m
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 233
t
FIG. E6.5 Total displacement for different values of the phase angle q in Example 6.3.2 (a ¼ 0:2).
@CivilMethod
234 PART I Single-degree-of-freedom systems
utot(t) (m)
t
FIG. E6.6 Total displacement for different values of the allocation coefficient a in Example 6.3.2
(q ¼ 3p=4).
max|utot(t)|
a
FIG. E6.7 Extreme value max|utot ðt Þ| versus the allocation coefficient a in Example 6.3.2
(x ¼ 0:05).
From the study of the numerical results, we may draw the following
conclusion:
The percentage allocation of the damping to the two columns does not affect
significantly the dynamic response of the structure. Consequently, an arbitrary
but reasonable allocation, for example, 0:4 < a < 0:6, allows treating practical
cases of asynchronous support excitations.
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 235
@CivilMethod
236 PART I Single-degree-of-freedom systems
examined when the motion is in the steady state phase. Hence, the vertical dis-
placement is given by Eq. (3.2.26), that is,
qÞ
u ðt Þ ¼ rsin ðwt (6.4.1)
where
p0 h 2 i 1
2
r¼ 1 b2 + ð2xb Þ2 (6.4.2a)
k
2xb
q ¼ tan 1
, b ¼ w=w (6.4.2b)
1 b2
The total force transmitted to the foundation is
f ¼ fS + fD
¼ ku + cu_
(6.4.3)
qÞ + crw cos ðwt
¼ krsin ðwt qÞ
q fÞ
¼ fT sin ðwt
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
fT ¼ r k 2 + ðcwÞ2 (6.4.4)
1 cw
f ¼ tan (6.4.5)
k
Using Eqs. (6.4.2a), (6.4.2b) for the expression of r and c=k ¼ 2x=w, the
previous relations are written as
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u 1 + ð2xbÞ2
fT ¼ p0 t 2 (6.4.6)
1 b2 + ð2xb Þ2
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 237
TR=fT /p0
@CivilMethod
238 PART I Single-degree-of-freedom systems
m w2 u0 1
u ðt Þ ¼ 2 2xb cos wt
1 b2 sin wt
k 1 b + ð2xbÞ
2 2
(6.4.10)
u0
¼ 2xb3 cos wt
1 b2 b 2 sin wt
2 2 2
1 b + ð2xbÞ
and the total displacement
utot ¼ ug + u
u0
+
¼ u0 sin wt 2
2xb3 cos wt
1 b2 b 2 sin wt
1 b2 + ð2xbÞ2
u0 nh i o
¼ 2
1 b2 + ð2xb Þ2 sin wt
2xb 3 cos wt
1 b2 + ð2xbÞ2
Þ
¼ uT sin ðwt
(6.4.11)
where
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u 1 + ð2xbÞ2
uT ¼ u0 t 2 (6.4.12)
1 b2 + ð2xb Þ2
!
1 2xb3
¼ tan (6.4.13)
1 b 2 + ð2xb Þ2
From Eqs. (6.4.8), (6.4.12), we obtain
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
ðu€tot Þmax ðutot Þmax uT u 1 + ð2xbÞ2
TR ¼ ¼ ¼ ¼ t 2 (6.4.14)
u€g max ug max u0 1 b2 + ð2xb Þ2
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 239
W 30kN
m¼ ¼ ¼ 3:058kNm1 s2
g 9:81m=s2
rffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffi
k 28
w¼ ¼ ¼ 3:026s1
m 3:058
The circular frequency of the harmonic motion and the maximum
acceleration are
w ¼ 2pf ¼ 94:248s1
u€g max ¼ w2 u0 ¼ 177:653 ¼ 18:109g g ¼ 9:81ms2
hence
w
b ¼ ¼ 31:147
w
and
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u 1 + ð2xbÞ2
TR ¼ t 2 ¼ 0:0065
1 b 2 + ð2xbÞ2
For the same elastic layer, the stiffness k and the damping coefficient c
are unaltered. Consequently, the transmission ratio TR can be modified if
the ratio b is changed. But because w is prescribed, this ratio can be chan-
ged, only if the natural frequency w is changed. This is possible if the mass
of the system (instrument) is changed by an increment Dm.
Let m 0 ¼ m + Dm be the new mass and x 0 , w0 the new damping ratio and
the frequency of the system (instrument), respectively. Then we will have
rffiffiffiffiffiffi
w0 m
¼ (1)
w m0
c ¼ 2mwx ¼ 2m 0 w0 x0
@CivilMethod
240 PART I Single-degree-of-freedom systems
which yield
rffiffiffiffiffiffi
0 mw m
x ¼x 0 0 ¼x (2)
mw m0
Moreover, we have
rffiffiffiffiffiffi
0 w m0
b ¼ 0¼b (3)
w m
From Eqs. (2), (3) we obtain
x0 b0 ¼ xb ¼ 3:115
Consequently, it must be
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u u
u 1 + ð2x 0 b0 Þ
2
u 2
TR ¼ u
¼ u 1 +
ð2 3:115Þ < 0:0055
t 2 2
t 2 2
1 b0 + ð2x0 b 0 Þ 1 b0
2
+ ð2 3:115Þ2
which holds if
pffiffiffiffiffiffiffiffiffiffiffiffi
33:885 < b0 ¼ b m 0 =m or m 0 > 3:619kNm1 s2
Hence Dm 0:561kNm1 s2 .
6.5 Problems
Problem P6.1 The supports 1 and 2 of the structure shown in Fig. P6.1 are sub-
and ug2 ðt Þ ¼ u0 sin ðwt
jected to the motions ug1 ðt Þ ¼ u0 sin wt p=3Þ, respec-
tively. A plane square rigid body of side a and density g is attached to
node 3. The mass of the column and the beam is neglected. Determine the
response of the structure and the reactions at the supports as well as the forces
that produce the support excitations. Data: L ¼ 8m, a ¼ 2m, cross section of the
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 241
(ug)i
(ug)i
t
FIG. P6.2 Accelerogram in problem P6.2
Problem P6.3 The rigid vertical column AC (Fig. P6.3) of a circular cross-
section, line density m is supported on the ground by means of a spherical hinge
at A and three elastic cables of cross-sectional area A and modulus of elasticity E.
(a) (b)
FIG. P6.3 Structure in problem P6.3
@CivilMethod
242 PART I Single-degree-of-freedom systems
The cables have been prestressed so that they can undertake compressive forces.
The column carries three advertising panels of total mass 10ma, which are
arranged as in Fig. P6.3b. Their support on the column extends to a length
1:25a. The structure is subjected to ground motion in the y direction, whose accel-
erogram u€g ðt Þ is given in Problem P6.2. Determine the minimum prestressing
force of the cable GB using the results of Problem P6.2. The cables are assumed
massless. Data: a ¼ 5m, m ¼ 0:5kNm1 s2 =m, E ¼ 2:1 108 kN=m2 , and
A ¼ 4cm . 2
Problem P6.5 The vertical columns of the frame in Fig. P6.5 have specific
weight g b and are assumed rigid. The elastic support on the ground is simulated
by the rotational springs CR . The horizontal beams are flexible with cross-
sectional moment of inertia I and modulus of elasticity E while their mass
and axial deformation are assumed negligible. The structure is subjected to
of total duration ttot ¼ 3s. Give
the horizontal ground motion ug ðt Þ ¼ u0 sin wt
the graph of the response spectrum of the rotation of the structure and compute
the extreme values of the shear force and the bending moment of the beams.
Data: Cross-sectional area of the beams A ¼ a=10 a=10, CR ¼ EI =a,
E ¼ 2:1 107 kN=m, g b ¼ 24kN=m3 , u0 ¼ 0:03m, w ¼ 3, 5 and 7s1 .
@CivilMethod
Response to ground motion and vibration isolation Chapter 6 243
=∞
(a)
h
h
(b) L
@CivilMethod
244 PART I Single-degree-of-freedom systems
[4] F.B. Hildebrand, Advanced Calculus for Applications, Prentice Hall, Englewood Cliffs, NJ,
1962.
[5] R.R. Craig Jr., J. Andrew, A.J. Kurdila, Fundamentals of Structural Dynamics, second ed., John
Wiley, New Jersey, 2006.
[6] J.W. Leonard, Tension Structures, McGraw-Hill, New York, 1988.
[7] J.T. Katsikadelis, Finite deformation of cables under 3-D loading: an analytic solution, in:
D.E. Beskos, D.L. Karabalis, A.N. Kounadis (Eds.), Proc. of the 4th National Congress on Steel
Structures, Patras, May 24–25, vol. II, 2002, pp. 526–534.
@CivilMethod
Chapter 7
Damping in structures
Chapter outline
7.1 Introduction 245 7.6.1 Introduction 257
7.2 Loss of energy due to damping 246 7.6.2 The fractional derivative 258
7.3 Equivalent viscous damping 249 7.7 Measurement of damping 260
7.4 Hysteretic damping 250 7.7.1 Free vibration decay
7.5 Coulomb damping 252 method 261
7.5.1 Free vibrations with 7.7.2 Resonance amplitude
Coulomb damping 252 method 262
7.5.2 Forced vibrations with 7.7.3 Width of response curve
Coulomb damping 255 method 263
7.6 Damping modeling via 7.8 Problems 265
fractional derivatives 257 References and further reading 267
7.1 Introduction
In this chapter, the damping of structures is discussed. Damping appears in all
mechanical systems that perform vibrations. It is the dissipation of energy in a
vibrating structure. The type of energy into which the mechanical energy is
transformed depends on the system and the physical mechanism that causes
the dissipation. The energy is lost either in the form of heat or is radiated into
the environment. For example, the loss of energy in the form of heat is perceived
when an iron rod is subjected to alternating bending. The sound produced by a
body that is hit represents the loss of energy dissipated into the environment. In
the study of vibrations, we are interested in the damping related to the response
of the structure. Damping is due to different energy dissipation mechanisms act-
ing simultaneously. In spite of the age-long detailed studies on the damping of
structures, the understanding of damping mechanisms is quite primitive. A
well-known method to get rid of this problem is to use so-called viscous damp-
ing. This approach was first introduced by Rayleigh [1] via his famous dissipa-
tion function (see Section 1.8.4).
The loss of energy of a vibrating system reduces the amplitude of the free
vibration. When a system undergoing forced vibrations reaches the phase of
the steady-state response, the loss of energy is balanced by the energy input into
the system by the excitation force.
In vibrating systems, we distinguish different types of damping forces,
which may be due to the internal molecular friction, the sliding friction, or the
where T ¼ 2p= w the period of the vibration in the steady-state response. The
quantity WD depends on several factors such as temperature, frequency, or
amplitude of the vibration.
In this section, we consider the simplest form of energy loss, that is, the loss
due to viscous damping. The damping force, in this case, is given by the relation
_ Moreover, the displacement in the steady-state phase, due to the
fD ¼ cu.
t, is given by Eq. (3.2.26), namely
harmonic force pðt Þ ¼ p0 sin w
t qÞ
u ðt Þ ¼ rsin ðw (7.2.2)
where
p0 1
r¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (7.2.3)
k
2 2
1 b + ð2xbÞ2
2xb
tan q ¼ (7.2.4)
1 b2
Consequently
u_ ðt Þ ¼ r t qÞ
w cos ðw (7.2.5)
The integral (7.2.1) yields
Z T
WD ¼ cu_ 2 dt
0
Z 2p=
w
(7.2.6)
w2 r 2
¼ c cos 2 ðw
t qÞdt
0
wr2
¼ pc
@CivilMethod
Damping in structures Chapter 7 247
Of particular
pffiffiffiffiffiffiffiffiffi interest is the
pffiffiffiffiffiffienergy
ffi loss at resonance. Then it is
¼ w ¼ k=m , c ¼ 2mxw ¼ 2x km and Eq. (7.2.6) becomes
w
WD ¼ 2xpkr2 (7.2.7)
Moreover, the energy input into the system is due to the work that the force
t produces in a complete oscillation. Namely
pðt Þ ¼ p0 sin w
Z T
Wp ¼ pðt Þdu
0
Z 2p=
w
(7.2.8)
¼ t qÞ sin w
w cos ðw
p0 r tdt
0
¼ p0 rp sin q
Taking into account that (see Eq. 3.2.27)
2xb
sin q ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ¼ 2xbrk=p0 (7.2.9)
1 b2 + ð2xbÞ2
@CivilMethod
248 PART I Single-degree-of-freedom systems
Thus, it must be
WI + WD + W S ¼ Wp (7.2.13)
which by virtue of Eqs. (7.2.11), (7.2.12) yields Eq. (7.2.10).
Eq. (7.2.5) yields
u_ ðt Þ2 ¼ w
2 r2 cos 2 ðw
t qÞ
¼w r 1 sin 2 ðw
2 2
t qÞ (7.2.14)
¼w 2 r2 u 2
hence
fD2 ¼ c2 u_ ðt Þ2
(7.2.15)
¼ c2 w
2 r2 u 2
which is readily transformed into
2
fD 2 u
+ ¼1 (7.2.16)
wr
c r
@CivilMethod
Damping in structures Chapter 7 249
wr2 2pc
pc w 2pð2mwx Þ
w
wD ¼ ¼ ¼ ¼ 4pxb (7.2.19)
kr2 =2 mw2 mw2
¼ 2xb (7.2.20)
wr2
WD ¼ ceq p (7.3.1)
@CivilMethod
250 PART I Single-degree-of-freedom systems
@CivilMethod
Damping in structures Chapter 7 251
t qh Þ
u ðt Þ ¼ rh sin ðw (7.4.7)
where
p0 1
rh ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi (7.4.8)
k 2
1 b2 + 2
tan qh ¼ (7.4.9)
1 b2
Working as in Section 7.2, we can determine the geometry of the hysteresis
loop for this type of damping. Thus, we obtain from Eq. (7.4.7)
u_ ðt Þ2 ¼ r2h w
2 cos 2 ðw
t qh Þ
2 2
¼ rh w 1 sin 2 ðw t qh Þ (7.4.10)
¼w 2 r2h u 2
Moreover, it is
2
k
fD2 ¼ ðf ku Þ2 ¼ u_ 2 (7.4.11)
w
which by virtue of (7.4.10) is written as
ðf ku Þ2 ¼ 2 k 2 r2h u 2 (7.4.12)
or
2
f ku 2 u
+ ¼1 (7.4.13)
krh rh
Therefore, in the plane u, f , Eq. (7.4.13) represents a rotated ellipse
(Fig. 7.4.1). The area of the ellipse is pkr2h and expresses the loss of energy
in a cycle. We observe that Eq. (7.4.13) does not involve the excitation fre-
quency. This implies that the hysteresis loop can be determined experimentally
using a low excitation frequency, that is, quasistatic, by plotting the load-
displacement curve.
For f ¼ 0, we obtain the abscissa of the ellipse on the u axis
d ¼ rh pffiffiffiffiffiffiffiffiffiffiffiffiffi (7.4.14)
1 + 2
which can be used to evaluate the damping coefficient . Hence
d
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (7.4.15)
r2h d 2
@CivilMethod
252 PART I Single-degree-of-freedom systems
while when the body is moving to the left (Fig. 7.5.1c) the equilibrium of forces
yields the equation of motion
m u€ + ku ¼ F (7.5.4)
@CivilMethod
Damping in structures Chapter 7 253
(a)
(b) (c)
FIG. 7.5.1 Motion with Coulomb friction.
@CivilMethod
254 PART I Single-degree-of-freedom systems
Eq. (7.5.11) holds until the velocity vanishes, namely t ¼ p=w. At that
instant, the body is at the extreme left position, where the displacement is
u ðp=wÞ ¼ ðu0 2F=k Þ. The body will now start moving to the right with ini-
tial conditions
p
p
F
u ¼ u0 2 , u_ ¼0 (7.5.12)
w k w
The motion is now described by Eq. (7.5.3), which has a particular solution
F
up ðt Þ ¼ (7.5.13)
k
and general solution
F
u ðt Þ ¼ A cos wt + B sin wt (7.5.14)
k
The initial conditions (7.5.12) give
3F
A ¼ u0 , B¼0 (7.5.15)
k
and Eq. (7.5.14) becomes
F F
u ðt Þ ¼ u 0 3 cos wt (7.5.16)
k k
@CivilMethod
Damping in structures Chapter 7 255
Eq. (7.5.16) holds until the body reaches the extreme right position, namely
until the instant t ¼ 2p=w. At that time, the body has completed a full oscillation
and the displacement is u ð2p=wÞ ¼ ðu0 4F=k Þ. This solution procedure con-
tinues to obtain the response of the next oscillations. The graphical representa-
tion of the displacement versus time is shown in Fig. 7.5.2. The curve was
obtained by numerical integration of the equation of motion (7.5.6) with
m ¼ 10kNm1 s2 , T ¼ 0:5s, F ¼ 23:685kN, u0 ¼ 0:55m, u_ 0 ¼ using the pro-
gram aem_nlin.m developed in Chapter 5. The numerical results coincide with
those obtained by the above-presented analytical solution. The motion is a
vibration with a period T ¼ 2p=w, which means that the Coulomb friction does
not affect the frequency or the period of vibration. The amplitude of vibration is
reduced in each cycle by 4F=k. A consequence of this is that the envelopes of
the curve are straight lines, unlike in the cases of viscous or hysteretic damping
where the envelopes are exponential functions. The motion of the system con-
tinues until the elastic force ku becomes smaller than the force F of the friction.
Until now, no difference was made between static friction Fs ¼ ms N and
dynamic friction Fd ¼ md N . The first occurs when the body is stationary and
the second when the body moves. Generally, it is md < ms , hence the dynamic
friction coefficient md will be used in the equation of motion while the static
friction coefficient ms is used for the control of the motion. In viscous or hys-
teretic damping, theoretically, the body does not stop moving because the
amplitude of the vibration reduces exponentially. Nevertheless, real structural
systems stop after a finite time. This is due to the fact that the Coulomb friction
coexists with other forms of damping and forces the moving systems to stop.
@CivilMethod
256 PART I Single-degree-of-freedom systems
WD ¼ 4Fr (7.5.23)
We can determine an equivalent coefficient of viscous damping by equating
the loss of energy (Eq. 7.5.23) with that of the viscous damping given by
Eq. (7.3.1). Namely,
wr2 ¼ 4Fr
ceq p (7.5.24)
which gives
4F
ceq ¼ (7.5.25)
p
wr
and an equivalent damping ratio
ceq 2F
xeq ¼ ¼ (7.5.26)
2mw pkrb
@CivilMethod
Damping in structures Chapter 7 257
Eq. (7.5.28) holds if 1 ð4F=pp0 Þ2 > 0, that is, when F=p0 < p=4. Obvi-
ously, for F=p0 > p=4 r becomes imaginary and this method for determining
an equivalent damping coefficient does not apply.
The phase angle results from Eq. (7.2.4) by setting x ¼ xeq and taking the
value of r from Eq. (7.5.28). Thus, we have
4F=pp0
tan q ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (7.5.29)
1 ð4F=pp0 Þ2
The plus sign is valid when b < 1 while the minus sign is valid when b > 1.
@CivilMethod
258 PART I Single-degree-of-freedom systems
damping effects using fewer material parameters than the integer order differ-
ential models but with equal precision. This approach involves fractional dif-
ferential equations, which needs acquaintance with fractional calculus.
@CivilMethod
Damping in structures Chapter 7 259
1.2
1 D 1u
c
0.8 D 0.99999u
c
D 0.9u
0.6 c
D 0.5u
0.4 c
D 0.1u
c
0.2
D 0.00001u
c
0 u
−0.2
−0.4
−0.6
0 0.5 1 1.5 2 2.5 3
FIG. 7.6.1 The fractional derivative of the function u ¼ t t 3 =6 + t 5 =120 for various orders.
m u€ + cD aC u + ku ¼ pðt Þ (7.6.3)
m u€ + ðc + k Þu ¼ pðt Þ + cu 0 (7.6.5)
that is, it yields the elastic response with stiffness k ∗ ¼ c + k and excitation
force p∗ ¼ pðt Þ + cu 0 . The Caputo derivative is employed because, contrary
to other types of fractional derivatives, it allows the application of initial con-
ditions having a direct physical significance.
Eq. (7.6.3) represents a fractional differential equation. Analytical solutions
of such equations are difficult or impossible to obtain. This reason has recently
boosted the development of efficient numerical methods for solving fractional
differential equations [10, 11].
Fig. 7.6.2 shows the free vibration response of an DOF system for various
values of the order a
@CivilMethod
260 PART I Single-degree-of-freedom systems
FIG. 7.6.2 Free vibration response of a SDOF system for various values of the order a.
The solution of Eq. (7.6.3) has been obtained using the method presented
in [11]. The developed MATLAB program has been given the name
three_term_FD.m and is included on this book’s companion website. Note
that for a 1 and a 0 the solutions are identical to the corresponding
analytical ones.
The fractional calculus has allowed the definition of any order of fractional
derivative, real or imaginary. This fact enables us to consider the fractional
derivative to be a function of time (explicit variable-order fractional derivative)
or of some other time-dependent variable (implicit variable-order fractional
derivative). Thus, the variable-order Caputo derivative for m ¼ 1 reads
Z t
aðt Þ 1 u_ ðτÞ
DC u ð t Þ ¼ dτ ðExplicitÞ (7.6.6a)
Gð1 aðt ÞÞ 0 ðt τÞaðt Þ
Z t
aðu Þ 1 u_ ðτÞ
DC u ð t Þ ¼ dτ ðImplicitÞ (7.6.6b)
Gð1 aðu ÞÞ 0 ðt τÞaðu Þ
Z t
aðu_ Þ 1 u_ ðτÞ
DC u ð t Þ ¼ dτ ðImplicitÞ (7.6.6c)
Gð1 aðu_ ÞÞ 0 ðt τÞaðu_ Þ
The concept of a variable-order fractional derivative exhibits notable advan-
tages over the constant order derivative and it has been recently used to model
the dynamic response of actual structures [12, 13].
@CivilMethod
Damping in structures Chapter 7 261
where
ui
dn ¼ ln (7.7.3)
ui + n
This method was illustrated in Example E6.2.1.
When the damping is hysteretic, the damping force is given by Eq. (7.4.2). In
w, provided that the damping is small. Hence,
free vibrations, it can be set as w
the equivalent damping coefficient ch and the damping ratio are obtained from
the relations
k
ch ¼ (7.7.4a)
w
xh ¼ (7.7.4b)
2
The hysteretic damping coefficient is obtained by using Eq. (7.7.2), if it is
set x ¼ xh . This yields
@CivilMethod
262 PART I Single-degree-of-freedom systems
2dn
¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (7.7.5)
4p n 2 + dn 2
2
p0 h i 1
2 2
r¼ 1 b2 + ð2xb Þ2 (7.7.6a)
k
2xb
q ¼ tan 1 (7.7.6b)
1 b2
For b ¼ 1 Eq. (7.7.6b) yields q ¼ 90° regardless of the value of x. If the
employed experimental instrumentation allows the measurement of the phase
angle q, then we adjust the excitation frequency so that q ¼ 90° and we measure
the amplitude of vibration r. Besides, setting b ¼ 1 in Eq. (3.2.28) we obtain
r 1
D ¼ max jRðt Þj ¼ ¼ (7.7.7)
p0 =k 2x
which yields
@CivilMethod
Damping in structures Chapter 7 263
p0 =k
x¼ (7.7.8)
2r
This previous method requires knowledge of the stiffness k of the structure,
which is determined either from the physical characteristics of the structure or
experimentally, for example, by imposing a load and measuring the resulting
displacement.
If the measurement of the phase difference is not easy, then we measure
experimentally the amplitude of the vibration in the range of resonance. Sub-
sequently, we plot the curve Dðb Þ ¼ rðb Þ=ðp0 =k Þ (see Fig. 7.7.1) and determine
its maximum. Then Eq. (3.2.31), namely
1
Dmax ¼ pffiffiffiffiffiffiffiffiffiffiffiffi (7.7.9)
2x 1 x 2
FIG. 7.7.1 Graphical representation of the curve D ðbÞ in the range of resonance.
@CivilMethod
264 PART I Single-degree-of-freedom systems
below the resonance frequency while the other is above it (see Fig. 7.7.2). The
respective values of b are obtained from Eq. (7.7.6b). Thus, we have
2xb1
¼1 (7.7.10a)
1 b21
2xb2
¼ 1 (7.7.10b)
1 b22
which can be written as
1 b21 2xb1 ¼ 0 (7.7.11a)
1 b22 2xb2 ¼0 (7.7.11b)
FIG. 7.7.2 Response curve DðbÞ in the neighborhood of resonance to determine bI and b II .
@CivilMethod
Damping in structures Chapter 7 265
pffiffiffi
1= 2Dmax are determined. It is apparent from Fig. 7.7.2 that these values are
two, which are denoted by bI and bII . For small values of the damping ratio, it is
Dmax 1=2x. Besides, using Eq. (3.2.28) we can write
1 1 1
pffiffiffi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(7.7.13)
2 2x 2
1 b2 + ð2xb Þ2
qffiffiffiffiffiffiffiffiffiffiffiffi
b 2II ¼ 1 2x2 + 2x 1 + x2 (7.7.14b)
pffiffiffiffiffiffiffiffiffiffiffiffi
Using the binomial theorem to expand 1 + x 2 and neglecting terms of
order higher than the second, the above relations reduce to
b2I 1 2x 2x 2 (7.7.15a)
b2II 1 + 2x 2x 2 (7.7.15b)
7.8 Problems
Problem 7.1 The damping force applied to a system moving in a fluid is given
by the relation fD ¼ lu_ a , where l and a are real constants. Give:
(i) The graphical representation of the displacement u ðt Þ when the system
performs free vibrations with initial conditions u ð0Þ ¼ 0:02m, u_ ð0Þ ¼ 0.
(ii) The graphical representation of the displacement u ðt Þ when the system is
t with zero initial conditions.
subjected to the harmonic load pðt Þ ¼ p0 sin w
Data: m ¼ 10kNm1 s2 , k ¼ 1500kN=m, l ¼ 100, a ¼ 3, p0 ¼ 300kN, and
¼ 2s1 .
w
@CivilMethod
266 PART I Single-degree-of-freedom systems
Problem 7.2 The system of Fig. P7.2 is subjected to Coulomb damping. Give:
(i) The graphical representation of the displacement u ðt Þ when the system
performs free vibrations with initial conditions (a) u ð0Þ ¼ 0:40m,
u_ ð0Þ ¼ 0 and (b) u ð0Þ ¼ 0, u_ ð0Þ ¼ 2:8ms1 .
(ii) The graphical representation of the displacement u ðt Þ when the system is
t with zero initial conditions.
subjected to the harmonic load pðt Þ ¼ p0 sin w
(iii) Calculate the equivalent damping ratio xeq of the equivalent viscous
damping.
Data: m ¼ 10:132kNm1 s2 , k ¼ 1600kN=m, b ¼ w
=w ¼ 0:4, N ¼ 70kN,
m ¼ 0:32, and p0 ¼ 1:57F.
Problem P7.3 Two bodies B1 and B2 with masses m1 and m2 , respectively, are
placed on two inclined planes whose angles are f1 and f2 , as shown in
Fig. P7.3. The bodies are connected by a massless cable of length L and axial
stiffness k ¼ EA=L. The friction coefficient between the bodies and the inclined
planes is m while between the cable and the pulley it is zero. Determine the
equation of motion of the system.
Problem P7.4 The damping of a SDOF system is expressed by the Caputo frac-
tional derivative of order a ¼ 0:5. Compare the response of the system with that
of viscous damping. Data: m ¼ 10kNm1 s2 , x ¼ 0:1, k ¼ 500kN=m,
p ¼ p0 sin 5t, and p0 ¼ 2kN. Hint: Use the program three_term_FD.m avail-
able on this book’s companion website.
@CivilMethod
Damping in structures Chapter 7 267
@CivilMethod
Chapter 8
Generalized single-degree-of-
freedom systems—Continuous
systems
Chapter outline
8.1 Introduction 269 8.3.3 Free vibrations of beams 286
8.2 Generalized single-degree-of- 8.3.4 Orthogonality of the
freedom systems 275 free-vibration modes 291
8.3 Continuous systems 284 8.3.5 Forced vibrations of
8.3.1 Introduction 284 beams 293
8.3.2 Solution of the beam 8.4 Problems 295
equation of motion 285 References and further reading 297
8.1 Introduction
In this chapter, the method of global shape functions is employed to approx-
imate the response of continuous systems by SDOF systems, which we call
generalized SDOF systems. The example that follows helps in understanding
the basic ingredients of the method as well as the error introduced by the
lumped mass assumption.
In order to study the dynamic response of the frame shown in Fig. 8.1.1a, we
approximate it by the model of Fig. 8.1.1b. As mentioned, in formulating this
model, it was assumed that the mass of the columns is concentrated at their ends,
that is, half at the top and half at the foot of the column. The consequence of this
assumption is that the elastic curve of the columns has the form of an unloaded
beam fixed at both ends, whose end cross-sections undergo a relative displace-
ment u ðt Þ. This assumption is, however, not entirely correct because the mass of
the columns is actually distributed along their length. This fact, apparently, pro-
u€
duces inertial forces fI ðx, t Þ ¼ m ðx, t Þ, where uðx, t Þ is the actual elastic curve
of the columns (see Fig. 8.1.1c). Thus, the problem would be correctly
addressed if the columns were treated as systems of infinite degrees of freedom,
that is, continuous systems whose top cross-sections are connected by the rigid
beam of mass m. Such an approach, however, would be quite difficult because
the analysis leads to partial differential equations (see Section 1.1). Neverthe-
less, it is possible to approximate the system by another model, which is closer
to the actual system than that of Fig. 8.1.1b. In this model, we consider that the
mass is distributed along the length of columns, but their deformed shape is cho-
sen so that it is close to the actual one. We observe that the end cross-sections of
the columns in Fig. 8.1.1a do not rotate during the motion. Hence the shape func-
tion ðx Þ ( see Eq. (8.1.1)) representing the shape of the elastic curve should be
chosen so that it satisfies the geometrical boundary conditions at the ends of the
columns, which demand ð0Þ ¼ 0, 0 ð0Þ ¼ 0, ðh Þ 6¼ 0, and 0 ðh Þ ¼ 0. The func-
tions ðx Þ are not unique. They constitute an infinite set of functions called
geometrically admissible functions.
(a) (b)
(c)
FIG. 8.1.1 Model of a two-column frame with rigid beam.
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 271
(a) (b)
FIG. 8.1.2 Shape function (a) and and deflection curve (b) in Example 8.1.1
where sx is the normal stress on the cross-section of the column and ex the cor-
responding strain. For a linearly elastic material with modulus of elasticity E,
it is ex ¼ sx =E. Taking into account that the bending stress in a beam is
sx ¼ Mx y=I ðx Þ, we obtain
1 Mx y 2
U0 ¼ (8.1.8)
2E I ðx Þ
where I ðx Þ is the moment of inertia of the, in general, variable cross-section
and Mx the bending moment. Hence, the total elastic energy of the one
column is
Z
U
¼ U0 dV
2 V
Z (8.1.9)
1 Mx y 2
¼ dV
V 2E I ðx Þ
where V is the volume of the column. It is known from the beam theory
that
Mx ¼ EI ðx Þu00 ðx, t Þ (8.1.10)
which is introduced into Eq. (8.1.9) to yield
Z
U 1 h
EI ðx Þ½u00 ðx, t Þ dx
2
¼ (8.1.11)
2 2 0
For I ðx Þ ¼ I ¼ constant and uðx, t Þ ¼ ðx Þu ðt Þ, the previous equation
becomes
Z h
U EI
½ 00 ðx Þ dx
2 2
¼ ½u ðt Þ (8.1.12)
2 2 0
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 273
(b) Evaluation of the kinetic energy T . The kinetic energy is due to the velocity
u_ ðt Þ of the mass m of the horizontal beam as well as to the velocity u_ ðx, t Þ
of the mass m at points x of the column axis. Thus, we have
Z h
1 1
½u_ ðx, t Þ dx
2
T ¼ m u_ + 2
2
m (8.1.17)
2 0 2
By virtue of Eq. (8.1.1), the above relation is written as
Z h
1
T ¼ m u_ 2 + m u_ 2 ½ ðx Þ2 dx (8.1.18)
2 0
(c) Evaluation of the virtual work dWnc . This is due to the nonconservative
external force pðt Þ. This is
dWnc ¼ pðt Þdu (8.1.21)
Now we proceed to the derivation of the equation of motion using:
1. Hamilton’s principle. Substituting Eqs. (8.1.16), (8.1.20), (8.1.21) into
Eq. (1.7.13) and taking into account that A ¼ 0, we obtain
Z t2
EI
24 3 udu ðm + 0:742h m Þud
_ u_ pðt Þdu dt ¼ 0 (8.1.22)
t1 h
which after elimination of du_ using integration by parts yields the equa-
tion of motion
m ∗ u€ + k ∗ u ¼ p∗ ðt Þ (8.1.23)
where
24EI
m ∗ ¼ m + 0:741mh, k∗ ¼ , p∗ ðt Þ ¼ pðt Þ (8.1.24)
h3
@CivilMethod
274 PART I Single-degree-of-freedom systems
where dWex and dWin denote the virtual work of the external and inter-
nal forces, respectively.
For the system of Fig. 8.1.1, the external forces are the excitation
force pðt Þ and the inertial forces. Hence
Z h
€ 2
dWex ¼ m udu u€
m ðx, t Þd uðx, t Þdx + pðt Þdu
0
€ 0:742mhdu
dWex ¼ m udu + pðt Þdu (8.1.27)
The virtual work of the internal forces for the two columns is
obtained from the relationa
Z
dWin ¼ 2 sx dex dV (8.1.28)
V
12EI
dWin ¼ 2 udu (8.1.29)
h3
Substituting Eqs. (8.1.27), (8.1.29) into Eq. (8.1.26) yields the equation
of motion (8.1.23).
a
The expression for the strain energy of a beam due to bending is derived from the strain energy
of the beam by taking its variation ([1], Chap. 1):
Z Z
1 1
Win ¼ sx ex dV ¼ Es2 dV (a)
2 V 2 V x
which by taking the variation gives
Z Z
dWin ¼ Eex dex dV ¼ sx dex dV (b)
V V
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 275
FIG. 8.1.3 Dependence of T*/T on the ratio of the column mass over total beam mass in the frame
of Fig. 8.1.1
If the mass of the column is assumed concentrated at the ends of the col-
umns, the coefficients of the corresponding equation of motion result as
24EI
m ∗ ¼ m + mh, k∗ ¼ , p∗ ðt Þ ¼ pðt Þ (8.1.30)
h3
We observe that the generalized mass is less by 0:259mh. Fig. 8.1.3
shows the variation of the ratio T ∗ =T versus the ratio mh=m, where T ∗
is the natural period of the generalized single-degree-of-freedom system
and T the the period of the model in Fig. 8.1.1b. We observe that the lumped
mass assumption has a small influence on the natural period when the mass
of the columns with respect to the mass of the horizontal beam is small.
Illustrative examples facilitating the comprehension of all concepts are pre-
sented and the pertinent bibliography with recommended references for fur-
ther study is also included. The chapter is enriched with problems to be
solved.
@CivilMethod
276 PART I Single-degree-of-freedom systems
(a)
(b)
(c)
(d)
FIG. 8.2.1 Beam resting on foundation with nonhomogeneous elastic and damping reaction.
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 277
(a) Evaluation of the elastic energy U . This energy consists of the strain energy
of the beam due to bending and the elastic energy of the springs of Winkler’s
model. That is,
Z Z
1 L 00 2 1 L
U¼ EI ðx Þ½u ðx, t Þ dx + k ðx Þ½uðx, t Þ2 dx (8.2.3)
2 0 2 0
which yields
Z L Z L
dU ¼ EI ðx Þu00 ðx, t Þd u00 ðx, t Þdx + k ðx Þuðx, t Þd uðx, t Þdx (8.2.4)
0 0
(b) Evaluation of the kinetic energy T . This energy consists of the kinetic
energy of the distributed mass m ðx Þ and the kinetic energy of the lumped
masses mi .That is
Z h i2
1 L 1X N
1X N
0
m ðx Þ½u_ ðx, t Þ dx + mi ½u_ ðxi , t Þ + Ii u_ ðxi , t Þ
2 2
T¼
2 0 2 i¼1 2 i¼1
(8.2.7)
0
where u_ ðx, t Þ and u_ ðx, t Þ are the transverse velocity and the angular veloc-
ity, respectively, at the cross-section x. Eq. (8.2.7) gives
Z L X
N
dT ¼ m ðx Þu_ ðx, t Þdu_ ðx, t Þdx + mi u_ ðxi , t Þdu_ ðxi , t Þ
0 i¼1
(8.2.8)
X
N
0 0
+ Ii u_ ðxi , t Þdu_ ðxi , t Þ
i¼1
@CivilMethod
278 PART I Single-degree-of-freedom systems
(c) Evaluation of the virtual work dWnc of the nonconservative forces. This is
due to the damping force fD ðx, t Þ ¼ cðx Þu_ ðx, t Þ and the external force
pðx, t Þ. Hence we have
Z L Z L
dWnc ¼ cðx Þu_ ðx, t Þd uðx, t Þdx + pðx, t Þd uðx, t Þdx (8.2.11)
0 0
Z L
p∗ ðt Þ ¼ pðx, t Þ ðx Þdx (8.2.13b)
0
(d) The potential A of the conservative forces. This is due to the work of the
axial force P. Obviously, if the axial deformation is neglected, it is
A ¼ 0. But if we consider shortening of the beam due to bending, that is,
if we adopt large displacements, then the shortening is expressed by the
nonlinear term of the strain-displacement relation. Thus, according to the
nonlinear theory of elasticity, we have [1, 2]
1 ∂u 2
ex ¼ (8.2.14)
2 ∂x
Consequently
Z L
1
½u0 ðx, t Þ dx
2
A ¼ PDL ¼ P (8.2.16)
2 0
and
Z L
dA ¼ P u0 ðx, t Þd u0 ðx, t Þdx (8.2.17)
0
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 279
where
Z L
k∗ ¼ P ½ 0 ðx Þ dx
2
(8.2.19)
0
This value of the axial force is the buckling load of the structure.
A consequence of this is the vanishing of the natural frequency,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
w¼ k ∗ k∗ =m ∗ ¼ 0. Therefore, a method of determining the buckling
load is to find the value of the compressive axial load, which annuls the nat-
ural frequency. This method is known as the dynamic criterion for buckling.
The method of approximating the continuous systems by a generalized
SDOF system can be successful with regard to the displacement. However,
we should be careful when we need to determine the stress resultants from
the obtained deflection curve using the known relations
M ðx, t Þ ¼ EI u00 ðx, t Þ (8.2.24)
000
Q ðx, t Þ ¼ EI u ðx, t Þ (8.2.25)
The stress resultants resulting from the above relations may deviate con-
siderably from the actual ones. This is illustrated by the following example.
We consider the cantilever beam of Fig. 8.2.2. The function
px
ðx Þ ¼ 1 cos (8.2.26)
2L
@CivilMethod
280 PART I Single-degree-of-freedom systems
Therefore, it can be used as a shape function for the cantilever beam and
the resulting displacement is
px
uðx, t Þ ¼ 1 cos u ðt Þ (8.2.28)
2L
Eq. (8.2.25) gives
p 2 px
Q ðx, t Þ ¼ EI sin u ðt Þ (8.2.29)
2L 2L
which results in
Q ð0, t Þ ¼ 0 (8.2.30)
This result is absurd. Nevertheless, this problem can be circumvented
if the stress resultant is evaluated using the procedure described in
Example 8.2.1.
Example 8.2.1 The industrial chimney of length L ¼ 75m shown in Fig. E8.1
consists of the outer reinforced concrete shell, which supports the linings. The
thickness of the thermal insulation layer is ti ¼ 0:10m and that of the refractory
layer tr ¼ 0:10m. The chimney is fixed on the ground.
1. Determine the natural frequency of the structure using two different shape
functions: (i) the elastic curve of a cantilever with constant cross-section
under a uniformly distributed load, and (ii) the first vibration mode of a
cantilever with constant cross-section.
2. Study the dynamic response of the chimney subjected to the impulsive wind
pressure shown in Fig. E8.1b. The analysis will be done using the shape
function that produces more accurate results.
3. Compute the bending moment and shear force at the base of the chimney and
give their expressions as a function of time.
4. Compute the dynamic magnification factor D ¼ max jRðt Þj for the
displacement.
Data:
Specific weight of reinforced concrete: g b ¼ 24kN=m3
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 281
(a) (b)
FIG. E8.1 Industrial chimney and loading.
@CivilMethod
282 PART I Single-degree-of-freedom systems
1 2
ðx Þ ¼ 6x 4x3 + x4 , x ¼ x=L (3)
3
Computation of the integrals using MATLAB gives
Z L
EI ðx Þ½ 00 ðx Þ dx ¼ 2977:6218
2
k ¼
∗ (4)
0
Z L
m∗ ¼ m ðx Þ½ ðx Þ2 dx ¼ 120:9955 (5)
0
Hence
pffiffiffiffiffiffiffiffiffiffiffiffiffi
w¼ k ∗ =m ∗ ¼ 4:9608 (6)
(ii) The shape function is the first vibration mode of a cantilever with a
uniform cross-section (see Section 8.3.3.2)
1
ðx Þ ¼ ½ cosh lx cos lx 0:7341ð sinh lx sinh lx Þ, l ¼ 1:8751=L (7)
3
Z L
EI ðx Þ½ 00 ðx Þ dx ¼ 2715:2000
2
k∗ ¼ (8)
0
Z L
m∗ ¼ m ðx Þ½ ðx Þ2 dx ¼ 116:8332 (9)
0
pffiffiffiffiffiffiffiffiffiffiffiffiffi
w¼ k ∗ =m ∗ ¼ 4:8208 (10
Obviously, case (ii) is more accurate because it yields a smaller
natural frequency (see Chapter 12).
2. The dynamic response of the chimney
The outer diameter of the chimney is
d ðx Þ ¼ 6:40 0:0507x (11)
hence the wind load per unit length is given by
pðx, t Þ ¼ pw d ðx Þf ðt Þ (12)
where
8
> t
>
> 2 if 0 t t1 =2
>
< t1
f ðt Þ ¼ t (13)
>
> 2 1 if t1 =2 t t1
>
> t
: 1
0 if t1 t
hence the peak load of the generalized SDOF system is
Z L
p∗w ¼ pw d ðx Þ ðx Þdx ¼ 106:81 (14)
0
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 283
The dynamic response will be obtained from the solution of the equation
of motion
m ∗ u€ + k ∗ u ¼ p∗w f ðt Þ (15)
The solution of Eq. (15) is obtained analytically in three phases.
Phase Ι. 0 t 1. The solution is given by Eq. (3.4.15)
2p∗w sin wt
uI ¼ t ¼ 0:039337t 0:008160 sin 4:8208t (16a)
k ∗ t1 w
Phase ΙΙ. 1 t 2, et ¼ t 1. The solution is given by Eqs. (3.3.14),
(3.5.10)
u_ I ð1Þ p∗w sin wet e
uII ðt Þ ¼ sin wet + uI ð1Þcos wet + 1 cos wet + t (16b)
w k∗ w
where
uI ð1Þ ¼ 0:04745m, u_ I ð1Þ ¼ 0:03508m=s
Phase IΙΙ. 2 t, ^t ¼ t 2. The solution is given by Eq. (2.2.13)
u_ II ð2Þ
uIII ðt Þ ¼ sin w^t + uII ð2Þcos w^t (16c)
w
where
uII ð2Þ ¼ 0:01447m, u_ II ð2Þ ¼ 0:00759m=s
Fig. E8.2 shows the graphical representation of the response ratio Rðt Þ,
from which we conclude that max jRðt Þj occurs in phase ΙI. The numerical
solution gives D ¼ 1:2673 occurring at t ¼ 1:12s.
@CivilMethod
284 PART I Single-degree-of-freedom systems
The stress resultants at the base of the chimney are obtained by considering
the equilibrium of all external forces, that is, Q ð0, t Þ, M ð0, t Þ, the wind pres-
sure pðx, t Þ, and the inertia force fI ðx, t Þ. Thus, referring to Fig. E8.3, we have
Z L Z L
Q ð0, t Þ + fI ðx, t Þdx pðx, t Þdx ¼ 0
0 0
Z L Z L
M ð0, t Þ xf I ðx, t Þdx + xpðx, t Þdx ¼ 0
0 0
which yield
Q ð0, t Þ ¼ 211:6454u€ðt Þ + 337:4063f ðt Þ (17)
M ð0, t Þ ¼ 10421:7751u€ðt Þ 10870:3125f ðt Þ (18)
@CivilMethod
286 PART I Single-degree-of-freedom systems
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 287
m
fðivÞ ðx Þ l
f ðx Þ ¼ 0 (8.3.10b)
EI
If l 0, the solution of Eq. (8.3.10a) does not represent an oscillatory
motion. Therefore, l must be a positive constant, l ¼ w2 . Thus, the solution
of Eq. (8.3.10a) is (see Section 2.2)
Y ðt Þ ¼ A cos wt + B sin wt (8.3.11)
where A,B are arbitrary constants depending on the initial conditions Y ð0Þ and
Y_ ð0Þ. Thus we have
Y_ ð0Þ
Y ðt Þ ¼ Y ð0Þcos wt + sin wt (8.3.12)
w
Apparently, w is the natural frequency of the vibration, unknown in the first
instance.
Similarly, Eq. (8.3.10b) is written as
k 4 b4 ¼ 0 (8.3.16)
whose roots are
k1, 2 ¼ ib, k3, 4 ¼ b (8.3.17)
Using each of these roots in Eq. (8.3.15) yields four terms, which are added
to give the general solution
fðx Þ ¼ C10 eibx + C20 eibx + C30 ebx + C40 ebx (8.3.18)
where C 0 i ði ¼ 1, 2, 3, 4Þ are arbitrary complex constants.
Using Euler’s formula (2.2.8) and the expressions of the hyperbolic sine and
cosine, Eq. (8.3.18) becomes
fðx Þ ¼ C1 cos bx + C2 sin bx + C3 cosh bx + C4 sinh bx (8.3.19)
in which Ci ði ¼ 1, 2, 3, 4Þ are new arbitrary constants related to
C 0 i ði ¼ 1, 2, 3, 4Þ, and can be determined from the boundary (support) condi-
tions of the one-span beam.
@CivilMethod
288 PART I Single-degree-of-freedom systems
FIG. 8.3.1 Vibration modes and natural frequencies of a uniform simply supported beam.
Inasmuch as Eqs. (8.3.21a), (8.3.21b) are valid for all values of t, they are
satisfied only if
fð0Þ ¼ 0, f00 ð0Þ ¼ 0 (8.3.22a)
fðLÞ ¼ 0, f00 ðLÞ ¼ 0 (8.3.22b)
Introducing Eq. (8.3.19) into Eq. (8.3.22a) yields
C1 + C3 ¼ 0 (8.3.23a)
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 289
C1 + C3 ¼ 0 (8.3.23b)
from which we obtain
C1 ¼ C3 ¼ 0 (8.3.24)
Further, introducing Eq. (8.3.19) into Eq. (8.3.22b), we obtain
C2 sin bL + C4 sinh bL ¼ 0 (8.3.25a)
@CivilMethod
290 PART I Single-degree-of-freedom systems
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 291
Table 8.3.1 gives the first five roots of bn L of Eq. (8.3.39). They have been
obtained numerically using the function fsolve of MATLAB. Note that for n > 3
they can be obtained from the relation
p
bn L ð2n 1Þ (8.3.42)
2
The first three of these mode shapes are shown in Fig. 8.3.2 along with their
natural frequencies
n bn L
1 1.8751040688
2 4.6940911329
3 7.8547574382
4 10.995540734
5 14.137168391
FIG. 8.3.2 Vibration modes and natural frequencies of a uniform cantilever beam.
Z L
0 if r 6¼ n
fr ðt Þfn ðt Þdt ¼ (8.3.43)
0 cn if r ¼ n
for any two functions fn ,fr F, m, nN . In the language of partial differential
equations, the free-vibration modes are called the eigenfunctions of the eigenvalue
problem described by the differential equation (8.3.13) and its boundary condi-
tions. The orthogonality condition is readily proved by proceeding as follows.
The mode shapes fn ,fr satisfy Eq. (8.3.13), that is,
fðnivÞ ðx Þ b4n fn ðx Þ ¼ 0 (8.3.44a)
fðrivÞ ðx Þ b 4r fr ðx Þ ¼ 0 (8.3.44b)
Multiplication of Eq. (8.3.44a) by fr ðx Þ and integrating over the interval
½0, L gives
Z L Z L
b4n fn ðx Þfr ðx Þdx ¼ fðnivÞ ðx Þfr ðx Þdx (8.3.45)
0 0
Further, integrating the right side of the above equation twice by parts gives
Z Z L
L
00 00
00 0
L
000 L
fn ðx Þfr ðx Þdx fn ðx Þfr ðx Þ 0 + fn ðx Þfr ðx Þ 0 ¼ bn
4
fn ðx Þfr ðx Þdx
0 0
(8.3.46)
Obviously, the terms in square brackets in the above equation vanish if
either end of the beam is simply supported, fixed, or free. Thus, we have
Z L Z L
4
bn fn ðx Þfr ðx Þdx ¼ f00n ðx Þf00r ðx Þdx (8.3.47a)
0 0
which for b 4n 6¼ b4r results in the orthogonality condition for the mode shapes
Z L
fn ðx Þfr ðx Þdx ¼ 0 (8.3.49)
0
It can also be shown that the set F : ffn ðx Þg is complete, that is there is no
other function outside the set F, which satisfies the condition (8.3.49).
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 293
Using now the orthogonality condition (8.3.49) and taking into account
Eqs. (8.3.45), (8.3.14), we obtain
Mn Y€ n ðt Þ + Kn Yn ðt Þ ¼ Pn ðt Þ (8.3.52)
where
Z L Z L
Mn ¼ m f2n ðx Þdx, Kn ¼ w2n Mn , Pn ¼ fn ðx Þpðx, t Þdx (8.3.53)
0 0
denote the modal mass, the modal stiffness, and the modal force. These quan-
tities are also referred to as the generalized mass, the generalized stiffness, and
the generalized force, respectively.
The solution of Eq. (8.3.52) is given by Eq. (3.3.14), that is,
Z t
Y_ n ð0Þ 1
Y n ðt Þ ¼ sin wn t + Yn ð0Þcos wn t + Pn ðτÞsin ½wn ðt τÞdτ
wn M n wn 0
(8.3.54)
The initial conditions Yn ð0Þ, Y_ n ð0Þ for the time function result from
Eq. (8.3.50). This yields
X
∞
u ðx, 0Þ ¼ fn ðx ÞYn ð0Þ ¼ f ðx Þ (8.3.55)
n¼1
@CivilMethod
294 PART I Single-degree-of-freedom systems
Z L
f ðx Þfn ðx Þdx
Yn ð 0Þ ¼ 0
Z L
(8.3.56a)
f2n ðx Þdx
0
Similarly, we obtain
Z L
g ðx Þfn ðx Þdx
Y_ n ð0Þ ¼ 0
Z L
(8.3.56b)
f2n ðx Þdx
0
2. Determine the modal mass, modal force, and initial conditions. They are
obtained from Eqs. (8.3.53), (8.3.56a), (8.3.56b):
Z L np mL
Mn ¼ m sin 2 x dx ¼ (2)
0 L 2
Z L np L
Pn ¼ p 0 sin x dx ¼ p0 ½1 ð1Þn (3)
0 L np
Yn ð0Þ ¼ 0, Y_ n ð0Þ ¼ 0 (4)
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 295
The bending moment and the shear force are evaluated from the
expressions
M ðx, t Þ ¼ EI u 00 ðx, t Þ ¼ EI f00 ðx ÞY ðt Þ (7)
000 000
Q ðx, t Þ ¼ EI u ðx, t Þ ¼ EI f ðx ÞY ðt Þ (8)
8.4 Problems
Problem 8.1 The television tower of Fig. P8.1 is subjected to seismic ground
motion ug ðt Þ. Derive the equation of motion if the structure is approximated by
a SDOF system. The reaction moment
of the elastic ground is represented by the
nonlinear expression MR ¼ CR f + 14 f2 , where CR ¼ KI f ; If is the moment of
inertia of the planform of the fundament and K ¼ E=10h the foundation mod-
ulus with E being the modulus of elasticity of the material of the structure. The
cross-section of the flexible column, the planform of the fundament, and the
body B are circular with diameters D, Df ¼ 8D, and DB ¼ 5D, respectively.
The density of the material is r. The fundament and the body B are
assumed rigid.
Problem P8.2 The continuous beam of Fig. P8.2 rests on Winkler’s elastic
foundation with variable modulus k ðx Þ. The beam is axially subjected to the
load P. Determine the value of the load that produces buckling.
FIG. P8.2 Continous beam on nonhomogeneous Winkler’s elastic foundation in problem P8.2
@CivilMethod
296 PART I Single-degree-of-freedom systems
(a)
(b)
FIG. P8.3 Simple supported bridge in Problem P8.3
Problem P8.4 Use the method of separation of variables to solve the equation
of free flexural vibrations of the:
(a) Fixed-fixed beam.
(b) Fixed-simply supported beam
In both cases, determine the natural mode shapes and the frequency equation.
Compute the first three natural frequencies.
Problem P8.5 Analyze the free flexural vibrations of the two-span continuous
beam of Fig. P8.5 by solving the equation of motion of the beam. Hint: Consider
the continuity condition at support 2, that is, u 0 I ðL, t Þ ¼ u 0 II ð0, t Þ.
@CivilMethod
Generalized single-degree-of-freedom systems—Continuous systems Chapter 8 297
@CivilMethod
Chapter 9
9.1 Introduction
The methods we discussed in the previous chapters for solving the equation of
motion of a SDOF system were accomplished using time as an independent var-
iable or, said differently, the solution was obtained in the time domain. These
methods are either analytical, which in the general case lead to the evaluation of
the Duhamel integral, or numerical, implemented by the step-by-step integra-
tion methods. Occasionally, simpler or more convenient analytical or numerical
solutions may be possible for certain types of dynamics problems, using integral
transforms such as the Laplace transform or the Fourier transform. The integral
converts the linear differential equation into a linear algebraic equation, from
which the integral transform of the unknown function is obtained. Then the
inverse transform results in the solution in the time domain.
We have already discussed the Laplace transform in Section 3.3, where it
was employed to solve the equation of motion of a SDOF system under an arbi-
trary external excitation. The Laplace transform uses a parameter that does not
have a direct physical meaning. Instead, the parameter in the Fourier transform
has the physical meaning of frequency [1–3]. The method of analyzing dynam-
ical systems using the Fourier transform is known as the analysis in the fre-
quency domain. It plays an important role in studying the dynamic response
of linear systems, that is, systems described by linear differential equations.
@CivilMethod
Analysis in the frequency domain Chapter 9 301
loading are preceded. Then, the method extends to nonperiodic loads by intro-
ducing the Fourier integral and the Fourier transform. The chapter closes with
the presentation of the discrete Fourier transform and the fast Fourier transform.
Z T =2
2
an ¼ tdt, n ¼ 1, 2, 3…
pðt Þcos n w (9.2.2b)
T T =2
Z T =2
2
bn ¼ tdt,
pðt Þ sin n w n ¼ 1, 2, 3… (9.2.2c)
T T =2
@CivilMethod
302 PART I Single-degree-of-freedom systems
Hence
Z T =2
1 einwt einwt
bn ¼ pðt Þ dt ¼ bn (9.2.5b)
T T =2 i
The Fourier series (9.2.1) by virtue of Eqs. (9.2.3a), (9.2.3b) becomes
X1
einwt + einwt X 1
ein wt einwt
pðt Þ ¼ a0 + an + bn
n¼1
2 n¼1
2i
(9.2.6)
1X 1
1X 1
¼ ao + ðan ibn Þeinwt + ðan + ibn Þeinwt
2 n¼1 2 n¼1
n¼1
@CivilMethod
Analysis in the frequency domain Chapter 9 303
@CivilMethod
304 PART I Single-degree-of-freedom systems
where
Z T =2 Z
1 1 T
cn ¼ pðt Þeinwt dt ¼ pðt Þein wt dt (9.3.10)
T T =2 T 0
or
Z "Z #
1 T =2
2X 1 T=2
pðt Þ ¼ pðτÞdτ + ðτ t Þdτ
pðτÞcos n w (9.4.4)
T T =2 T n¼1 T =2
@CivilMethod
Analysis in the frequency domain Chapter 9 305
If we set now
2p
¼w
nw n , D n + 1 w
w¼w n ¼ (9.4.8)
T
we may write Eq. (9.4.7) as
X
1
pðt Þ ¼ lim n ÞD
P ðw w (9.4.9)
T !1
n¼1
@CivilMethod
306 PART I Single-degree-of-freedom systems
@CivilMethod
Analysis in the frequency domain Chapter 9 307
The function P ðwÞ defined by Eq. (9.4.16) is called the (direct) Fourier
transform of pðt Þ while the function pðt Þ resulting from Eq. (9.4.17) is called
the inverse Fourier transform of P ðwÞ.
In the time domain, a function will be denoted by a small letter while its
Fourier transform is by the same capital letter. The relationship between them
will be symbolized by
Þ
pðt Þ , P ðw (9.4.18)
Usually, we denote the Fourier transform of a function pðt Þ by F ½pðt Þ while
its inverse is by F 1 ½pðt Þ, namely
Þ ¼ F ½pðt Þ
P ðw (9.4.19)
pðt Þ ¼ F 1 ½P ðw
Þ (9.4.20)
The Fourier transform of the derivative of a function pðt Þ is readily estab-
lished by applying integration by parts to Eq. (9.4.16). Generally, for a function
of which the ðn 1Þ order derivatives are continuous and the nth order deriv-
ative is piecewise continuous, it can be shown that
Z 1
pðnÞ ðt Þeiwt dt ¼ ðiw
Þn P ðw
Þ (9.4.21)
1
for x 6¼ 0 and
1
h ðt τ Þ ¼ sin wðt τÞ, t > τ (9.5.3)
mw
for x ¼ 0.
The integral (9.5.1) can be also written as
Z 1
u ðt Þ ¼ pðt Þ∗ h ðt Þ ¼ pðτÞh ðt τÞdτ (9.5.4)
1
¼ ÞpðτÞeiwτ dτ
H ðw
1 (9.5.6)
Z 1
¼ H ðwÞ pðτÞeiwτ dτ
1
ÞH ðw
¼ P ðw Þ
From Eq. (9.5.6) we deduce that the Fourier transform of the response to an
arbitrary load, namely of the convolution integral, is equal to the product of the
Fourier transforms of the functions in the convolution. Hence, we may write
symbolically
Þ
pðt Þ , P ðw (9.5.7)
Þ
h ðt Þ , H ðw (9.5.8)
Þ
u ðt Þ ¼ pðt Þ ∗h ðt Þ , U ðw (9.5.9)
The Fourier transform of the function u ðt t0 Þ is obtained as
Z 1
F½uðt t0 ¼ u ðt t0 Þeiwt dt
1
Z 1
¼ u ðxÞeiwðx + t0 Þ dt
1 (9.5.10)
Z 1
¼ eiwt0 u ðx Þeiwx dt
1
¼ eiwt0 F ½u ðt Þ
@CivilMethod
Analysis in the frequency domain Chapter 9 309
@CivilMethod
310 PART I Single-degree-of-freedom systems
The integral (5) is evaluated using the method of closed line integrals in the
complex domain of b. This method yields
u ðt Þ ¼ 0, t 0
" !#
p0 x
u ðt Þ ¼ 1e xwt
cos wD t + pffiffiffiffiffiffiffiffiffiffiffiffi sin wD t , 0 < t
k 1 x2
From the last example, we observe that the dynamic analysis in the fre-
quency domain requires the evaluation of complicated integrals, even for the
simplest load cases. This problem is circumvented by applying numerical
methods for the evaluation of the Fourier transform, such as the discrete Fourier
transform (DFT), and the fast Fourier transform (FFT). These methods are dis-
cussed in the next sections.
@CivilMethod
Analysis in the frequency domain Chapter 9 311
implementation of the DFT and its application for dynamic analysis in the fre-
quency domain.
We consider the function pðt Þ of Fig. 9.6.1, which is defined in the interval
0 t ttot .
tot
1 NX1
cn ¼ s ðtk ÞDt (9.6.4)
T0 k¼0
where s ðtk Þ is the value of the integrand at instants tk ¼ kDt. That is,
s ðtk Þ ¼ pðtk ÞeinwkDt (9.6.5)
@CivilMethod
312 PART I Single-degree-of-freedom systems
1 NX
1
cn ¼ pðtk Þe2pikn=N (9.6.8)
N k¼0
then taking into account that the function pðt Þ is periodic with a period
T0 ¼ N Dt, we can write the first sum as
X
1 X
1
cn e2pikn=N ¼ cn + N e2pik ðn + N Þ=N (9.6.11)
n¼M n¼M
Substituting Eq. (9.6.11) into Eq. (9.6.10), and taking into account
Eq. (9.6.12), we obtain
X
2M
pðtk Þ ¼ cn e2pikn=N (9.6.13)
n¼0
or because 2M ¼ N 1, it follows:
X
N 1
pðtk Þ ¼ cn e2pikn=N (9.6.14)
n¼0
@CivilMethod
Analysis in the frequency domain Chapter 9 313
1 NX
1
Pn ¼ pk e2pikn=N , k ¼ 0, 1, 2, …,N 1 (9.6.15)
N k¼0
X
N 1
pk ¼ Pn e2pikn=N , k ¼ 0, 1, 2, …,N 1 (9.6.16)
n¼0
The foregoing relations express the discrete Fourier transform (DFT), direct
and inverse, respectively.
The DFT approximates numerically the continuous Fourier transform,
defined by Eqs. (9.4.16), (9.4.17). The accuracy of the DFT is very good if
Dt is selected small. However, there is a fundamental difference between the
continuous Fourier transform and the discrete Fourier transform. The first pro-
vides the exact transform of the actual function while the second assumes a peri-
odic extension of the function. This means that the discrete transform is
applicable when the interval T0 is finite. It holds only within the period. Outside
it, the two transforms are completely different unless the function happens to be
periodic.
1 NX
1
Pn ¼ pðtk Þe2pikn=N (9.7.2)
N k¼0
2. We compute the DFT of the response function h ðt Þ. This requires the con-
finement of h ðt Þ in an interval equal or smaller than T0 .
1 NX
1
Hn ¼ h ðtk Þe2pikn=N (9.7.3)
N k¼0
@CivilMethod
314 PART I Single-degree-of-freedom systems
@CivilMethod
Analysis in the frequency domain Chapter 9 315
over DFT. The first FFT algorithm was developed by Gauss in the early 19th
century [6]. Also, the contributions of Runge, Danielson, Lanczos, and others
in the early 20th century were significant. However, its use did not attract the
interest of many researchers because the calculations had to be performed by
hand. It was only with the advent of computers that the FFT came to the fore-
ground. In 1965, J. W. Cooley and J. W. Tukey published an algorithm for cal-
culating the FFT [7, 8]. This algorithm is similar to that of Gauss and others and
is named after them as the Cooley-Tukey algorithm. Today, there are several
algorithms for FFT based on this algorithm. Below we present the Sande-Tukey
algorithm that is a variation of the Cooley-Tukey.
@CivilMethod
316 PART I Single-degree-of-freedom systems
X
N 1
Pn ¼ ~k W nk
p (9.8.5)
k¼0
W ¼ e2pi=N (9.8.6)
ðNX
=2Þ1 X
N 1
Pn ¼ pk e2pikn=N +
~ ~k e2pikn=N , n ¼ 0, 1, 2, …,N 1 (9.8.7)
p
k¼0 k¼N =2
ðNX
=2Þ1 ðNX
=2Þ1
Pn ¼ ~k e2pikn=N +
p ~m + N =2 e2pinðm + N =2Þ=N
p (9.8.8)
k¼0 m¼0
or
=2Þ1
ðNX
Pn ¼ ~k + epin p
p ~k + N =2 e2pikn=N (9.8.9)
k¼0
n
We observe that eipn ¼ ðeip Þ ¼ ð1Þn . Consequently, for points with
even n this factor is equal to one while with odd n it is equal to 1. The next
step is to separate the terms of Eq. (9.8.9) into two sums corresponding to the
even and odd values of n. Hence, for even values, we have
=2Þ1
ðNX
P2n ¼ p ~k + N =2 e2pik ð2nÞ=N
~k + p
k¼0
(9.8.10)
=2Þ1
ðNX
¼ p ~k + N =2 e2pikn=ðN =2Þ
~k + p
k¼0
@CivilMethod
Analysis in the frequency domain Chapter 9 317
=2Þ1
ðNX
P2n + 1 ¼ p ~k + N =2 e2pik ð2n + 1Þ=N
~k p
k¼0
(9.8.11)
=2Þ1
ðNX
¼ p ~k + N =2 e2pik=N e2pikn=ðN =2Þ
~k p
k¼0
where n ¼ 0, 1, 2, …, ðN =2Þ 1.
By virtue of Eq. (9.8.6), Eqs. (9.8.10), (9.8.11) may be written
=21
NX
P2n ¼ p ~k + N =2 W 2kn
~k + p (9.8.12)
k¼0
=21
NX
P2n + 1 ¼ p pk + N =2 W k W 2kn
~k ~ (9.8.13)
k¼0
We can now make an important observation, which is the key to the method.
The even and the odd expressions can be considered as two DFTs of N =2 points
each. We further set
~k + p
gk ¼ p ~k + N =2 (9.8.14)
hk ¼ p~k p
~k + N =2 W k , k ¼ 0, 1, 2, …, ðN =2Þ 1 (9.8.15)
Hence
)
P2n ¼ Gn
, n ¼ 0, 1, 2, …, ðN =2Þ 1 (9.8.16)
P2n + 1 ¼ Hn
@CivilMethod
318 PART I Single-degree-of-freedom systems
DFT with two points (see Fig. 9.8.2). The total number of complex multiplica-
tion for a given DFT reduces to N log 2 N . The importance of FFT over DFT is
demonstrated in Fig. 9.8.1.
FIG. 9.8.2 Flow chart of the first substitution stage of the DFF with N points by two DFF of N =2
points each when N ¼ 8.
@CivilMethod
Analysis in the frequency domain Chapter 9 319
u(t) (m)
t
u,t(t) (m/s)
t
u,tt(t) (m/s2)
t
FIG. E9.1 Dynamic response of the SDOF system using FFT in Example 9.8.1.
@CivilMethod
320 PART I Single-degree-of-freedom systems
FIG. E9.2 Amplitude spectrum of the 1999 Athens earthquake in Example 9.8.2.
9.9 Problems
Problem P9.1 Write a computer program for the evaluation of DFT and com-
pute the DFT of the function shown in Fig. P9.1. Then compute the inverse DFT
and compare the results with the exact function.
@CivilMethod
Analysis in the frequency domain Chapter 9 321
(b)
(a)
FIG. P9.4 Water tower (a) and load (b) in problem P9.4
@CivilMethod
322 PART I Single-degree-of-freedom systems
@CivilMethod
Chapter 10
Multi-degree-of-freedom
systems: Models and equations
of motion
Chapter outline
10.1 Introduction 325 10.5 Systems with distributed
10.2 Systems with localized mass mass and distributed stiffness 341
and localized stiffness 327 10.5.2 The method of global
10.3 Systems with distributed shape functions 342
mass and localized stiffness 328 10.6 Mixed systems 347
10.4 Systems with localized mass 10.7 Transformations of the
and distributed stiffness 330 equations of motion 351
10.4.1 The method of 10.8 Problems 354
influence coefficients 334 References and further reading 358
10.1 Introduction
So far, we have studied the dynamic response of SDOF systems. We have also
shown how a system with infinite degrees of freedom can be approximated by
a SDOF system. The trustworthiness of this approximation depends on various
issues. If the actual distribution of the physical properties of the structure, that
is, mass and stiffness, and that of the external force produce deformation during
the motion similar to the assumed, then the approximation with a SDOF system
gives acceptable results. A key shortcoming of this approximation is the diffi-
culty in determining the degree of reliability of the obtained results. In general,
however, the study of the dynamic response of structures requires their model-
ing with MDOF systems, especially when the deformation shapes are compli-
cated. In engineering structures, the mass, though distributed to all its members,
is usually lumped at certain points or regions. For example, in buildings the
mass is lumped at the levels of the stories or in a water tower at the top of
the column that supports the tank. This fact allows describing the motion of
a structure with that of a MDOF system with deformation parameters the dis-
placements of the points where the dynamic characteristics (mass and moment
of inertia) are concentrated.
Fig. 10.1.1a shows a three-story frame whose horizontal beams are virtually
rigid. In this structure, the mass of the columns is negligible compared to that of
the beams. Hence, the mass is lumped at the level of the beams and the structure
can be approximated by the model of Fig. 10.1.1b. The deformation shape dur-
ing motion is shown in Fig. 10.1.1c. Obviously, its motion can be determined by
establishing the displacements u1 ðt Þ, u2 ðt Þ, and u3 ðt Þ, that is, the system has
three degrees of motion.
(a) (b)
FIG. 10.1.2 Water tower (a) and its deformed dynamic model (b).
Fig. 10.1.2a shows a water tower. With the assumption that the mass of the
column is negligible compared with that of the tank, we can model the water
tower with the system of Fig. 10.1.2b. That is, the water tower is simulated
by a flexible column, which is fixed at the ground and has a mass m at its
top with a moment of inertia Io . The mass m can move horizontally and rotate
within the plane. The determination of the motion requires the establishment of
the displacement u ðt Þ and the rotation fðt Þ of the top cross-section of the col-
umn, that is, the system has two degrees of freedom.
The MDOF systems can be categorized as follows:
1. Systems with localized mass and localized stiffness.
2. Systems with distributed mass and localized stiffness.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 327
@CivilMethod
328 PART II Multi-degree-of-freedom systems
are the mass, the damping, and stiffness matrices of the system, respec-
tively, and
u1 p1 ð t Þ
u¼ , pðt Þ ¼ (10.2.8)
u2 p2 ðt Þ
(a)
(b)
FIG. 10.2.1 System with localized mass and localized stiffness (a). Forces applied to the masses
m1 and m2 (b).
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 329
(a)
(b)
FIG. 10.3.1 System with distributed mass and localized stiffness (a). Forces acting on the two
bodies (b).
@CivilMethod
330 PART II Multi-degree-of-freedom systems
Eqs. (10.3.2), (10.3.4) are the equations of motion, which in matrix form
become
u + Cu_ + Ku ¼pðt Þ
M€ (10.3.5)
where
2 3
2 " #
0
mL 9
63 7 c 0 k kL
M¼4 , C¼ , K¼
35 4 (10.3.6)
mL 0 0
0 kL 0:5L2 k
8
represent the mass, damping, and stiffness matrices of the system and
8 9
< 1 pðt Þ =
u
u¼ , pðt Þ ¼ 2 (10.3.7)
f : ;
M ðt Þ
the displacement and load vectors.
(a) (b)
FIG. 10.4.1 System with localized mass and distributed stiffness (a). Deformed dynamic model (b).
The equations of motion will results from the motion of the plane rigid body in
its plane. We examine the motion with respect to the point O, which does not coin-
cide with the center of mass of the body. In general, the system has three degrees of
freedom, namely the horizontal displacement, the vertical displacement, and the
rotation about O. Because the column is flexible, the horizontal displacement and
the rotation are due to the bending deformation. The vertical displacement is
caused by (i) the axial deformation of the column, which is very small and thus
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 331
neglected, and (ii) the shortening of the chord of the deflection curve, which is
also neglected in the linear theory. Therefore, the parameters of the motion are
the horizontal displacement u ðt Þ and the rotation fðt Þ, Fig. 10.4.1b. In the follow-
ing, the equations of motion are derived using two different methods.
1. The method of the Lagrange equations
(i) Elastic energy: This is due to the bending deformation of the column. The
deflection curve can be set in the form
d4 1 0 0
¼ 0, 1 ð0Þ ¼ 0, 1 ð0Þ ¼ 0, 1 ðh Þ ¼ 1, 1 ðh Þ ¼ 0 (10.4.2a)
dx 4
and
d4 2 0 0
¼ 0, 2 ð0Þ ¼ 0, 2 ð0Þ ¼ 0, 2 ðh Þ ¼ 0, 2 ðh Þ ¼ 1 (10.4.2b)
dx 4
Integrating of the differential equation (10.4.2a) gives
1 3 1 2
1 ðx Þ ¼ c1 x + c2 x + c3 x + c4 (10.4.3)
6 2
After evaluation of the arbitrary constants by applying the boundary
conditions, we obtain
x 2 x 3
1 ðx Þ ¼ 3 +2 ¼ 3x2 + 2x3 , x ¼ x=h (10.4.4)
h h
Similarly, we obtain
x 2 x
2 ð x Þ ¼ h 1 ¼ hx 2 ðx 1Þ, x ¼ x=h (10.4.5)
h h
The elastic energy is given (see Eq. 8.1.11)
Z
1 h
EI ½u00 ðx, t Þ dx
2
U¼
2 0
which by virtue of Eq. (10.4.1) becomes
Z
1 h 00 2
U¼ EI 1 ðx Þu ðt Þ + 002 ðx Þfðt Þ dx (10.4.6)
2 0
Differentiating Eq. (10.4.6) with respect to u ðt Þ and fðt Þ gives
@CivilMethod
332 PART II Multi-degree-of-freedom systems
Z h
∂U 00 00
00
¼ EI 1 ðx Þu ðt Þ + 2 ðx Þfðt Þ 1 ðx Þdx
∂u 0
Z h Z
00
2 h
00 00
¼ EIu ðt Þ (10.4.7a)
1 ðx Þ dx + EI fðt Þ 2 ðx Þ 1 ðx Þdx
0 0
12EI 6EI
¼ u+ 2 f
h3 h
Z h
∂U 00 00
00
¼ EI 1 ðx Þu ðt Þ + 2 ðx Þfðt Þ 2 ðx Þdx
∂f 0
Z h Z h
00 00
00
2
¼ EIu ðt Þ (10.4.7b)
1 ðx Þ 2 ðx Þdx + EI fðt Þ 2 ðx Þ dx
0 0
6EI 4EI
¼ u+ f
h2 h
(ii) Kinetic energy: The kinetic energy with respect to point O is evaluated
from Eq. (1.5.8). Taking the origin of the coordinates at point O we have:
XP ¼ u, YP ¼ 0, w ¼ f, _ xc ¼ 0, yc ¼ a=2, IP ¼ IO and Eq. (1.5.8)
becomes
1 1 a
T ¼ m u_ 2 + IO f_ 2 m u_ f_ (10.4.8)
2 2 2
Differentiating Eq. (10.4.8) with respect to u_ and f_ gives
∂T a
¼ m u_ m f_ (10.4.9a)
∂u_ 2
∂T a
¼ IO f_ m u_ (10.4.9b)
∂f_ 2
(iv) The potential of the external conservative forces: Because there are no
conservative forces, it is A ¼ 0.
Substituting Eqs. (10.4.7a), (10.4.7b), (10.4.9a) (10.4.9b), (10.4.10a,b)
into the Lagrange equation (1.8.11) gives
a 12EI 6EI
m u€ m f€ + 3 u + 2 f ¼ pðt Þ (10.4.11a)
2 h h
a 6EI 4EI a
IO f€ m u€ + 2 u + f ¼ pðt Þ (10.4.11b)
2 h h 2
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 333
or in matrix form
M€
u + Ku ¼pðt Þ (10.4.12)
where
2 3
2 ma 3 12EI 6EI
m 6 h3 h2 7
M ¼ 4 ma 2 5, K¼6 4 6EI 4EI 5
7 (10.4.13a,b)
IO
2 h2 h
( )
u pðt Þ
u¼ , pðt Þ ¼ a (10.4.13c,d)
f pðt Þ
2
@CivilMethod
334 PART II Multi-degree-of-freedom systems
6EI 4EI
MS ¼ 2
u+ f (10.4.15b)
h h
Obviously, they are identical to those given by Eqs. (10.4.7a),
(10.4.7b).
Substituting the previous expressions for fS and MS in
Eqs. (10.4.14a), (10.4.14b) yields
a 12EI 6EI
u m f€ + 3 u + 2 f ¼ pðt Þ
m€ (10.4.16a)
2 h h
a 6EI 4EI a
IO f€ m u€ + 2 u + f ¼ pðt Þ (10.4.16b)
2 h h 2
which are identical to Eqs. (10.4.11a), (10.4.11b).
b. Equilibrium with respect to the mass center C .
The equations of motion will be obtained from Eqs. (1.5.11a),
(1.5.11b), (1.5.11c) by setting uc ¼ u af=2, Fx ¼ fS + pðt Þ,
MC ¼ MS fS a=2. Thus, we obtain
a
m€u m f€ + fS ¼ pðt Þ (10.4.17a)
2
a
Ic f€ + MS + fS ¼ 0 (10.4.17b)
2
The first of the above equations is identical to (10.4.14a). The second
equation, however, looks different from (10.4.14b). Nevertheless, multi-
plying Eq. (10.4.17a) by a=2 and adding it to (10.4.17b) gives
a 2
a a
m€ u + IC + m f€ + MS ¼ pðt Þ (10.4.18a)
2 2 2
Because IC + m ða=2Þ2 ¼ IO (Steiner’s formula), the previous equa-
tion becomes
a a
m€ u + IO f€ + MS ¼ pðt Þ (10.4.18b)
2 2
which is identical to (10.4.14b).
The above transformation of the equations of motion from the center
of mass to point O is rather occasional. A formal method to transform the
equations of motion when we change the point of reference is presented in
Section 10.7.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 335
method with the plane frame of Fig. 10.4.3. The masses are localized at the
nodes, where the external loads are applied. In the general case in which the
axial deformation of the beams and columns is not neglected, each node i
has three degrees of freedom with respect to the global system of axes X Y ,
Y axes, respectively, and a rota-
two translations ui , vi in the directions of the X,
tion fi about the Z axes. Hence, the frame of Fig. 10.4.3 with n ¼ 6 free nodes
has in total N ¼ 3n ¼ 18 degrees of freedom.
@CivilMethod
336 PART II Multi-degree-of-freedom systems
Similarly, we formulate the vector of the external nodal loads (Fig. 10.4.5)
(a) (b)
FIG. 10.4.6 Equilibrium of forces (a) or moments (b) in the direction of the displacement ui .
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 337
where kij are constants expressing the stiffness influence coefficients or simply
the stiffness coefficients. They relate the elastic force fSi to the displacement uj .
The physical meaning of the coefficient kij results by setting uj ¼ 1 and uk6¼j ¼ 0
in Eq. (10.4.25), that is, by applying a unit displacement along the degree of
freedom j, holding all other displacements zero, as shown in Fig. 10.4.7. This
yields
fSi ¼ kij (10.4.26)
Eq. (10.4.26) states that the stiffness coefficient kij expresses the elastic
force applied along the degree of freedom i for a unit displacement along the
degree of freedom j, that is, the displacement uj , while all other displacements
are zero.
Fig. 10.4.7 shows the deformation of the frame when the displacement u4 ¼ 1
is applied at node 4 while all other displacements are zero, that is, uj ¼ 0
( j ¼ 1, 2, …, N , j 6¼ 4). Obviously, the stiffness influence coefficients are equal
to the forces required to maintain the deformed shape of the frame when sub-
jected to the displacement u4 ¼ 1.
For i ¼ 1, 2, …, N , Eq. (10.4.25) yields N equations, which we write in
matrix form
8 9 2 38 9
> f > k k ⋯ k1N > > u1 >
< S1 >
> = 6 11 12 7< u2 =
>
fS2 k k ⋯ k
¼6 21 22
4 ⋯ ⋯ ⋯ ⋯ 5> ⋮ >
2N 7 (10.4.27)
> ⋮ >
>
>
: ; : >
> ;
fSN kn1 kn2 ⋯ kNN uN
@CivilMethod
338 PART II Multi-degree-of-freedom systems
or
f S ¼ ku (10.4.28)
where
8 9 8 9 2 3
> fS1 > > u1 > k11 k12 ⋯ k1N
>
> >
> > >
> >
< fS1 = < u2 = 6k k ⋯ k 7
6 21 22 2N 7
fS ¼ , u¼ , k¼6 7 (10.4.29)
> ⋮ >
> > > ⋮ >
> > 4⋯ ⋯ ⋯ ⋯ 5
>
: >
; >
: ; >
fSN uN kn1 kn2 ⋯ kNN
The vector f S with dimensions N 1 represents the vector of the elastic
forces, the vector u with dimensions N 1 represents the vector of the displace-
ments, and the matrix k with dimensions N N represents the stiffness matrix
of the structure. Obviously, this method, which explains the physical meaning
of the stiffness coefficients, is by no means suitable for their evaluation because
it requires N static analyses of the fixed structure successively for u1 ¼ 1,
u2 ¼ 1, …, uN ¼ 1. The stiffness matrix, however, can be established using
other methods of structural analysis, for example, the flexibility method or
the direct stiffness method. Anyhow, the establishment of the stiffness matrix
of a structure is a subject of the static structural analysis and it is evaluated using
the most suitable method for a particular structure.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 339
@CivilMethod
340 PART II Multi-degree-of-freedom systems
where
8 9 8 9 2 3
> fI 1 > > u€ 1 > m11 m12 ⋯ m1N
< >
> = < >
> = 6 m21 m22
fI 2 u€ 2 ⋯ m2N 7
fI ¼ , €¼
u , m¼6
4 ⋯ ⋯
7 (10.4.37)
> ⋮ > > ⋮ >
> ⋯ ⋯ 5
: >
> ; : >
;
fIN u€ N mn1 mn2 ⋯ mNN
The vector f I with dimensions N 1 represents the vector of the inertial
forces, the vector u € with dimensions N 1 represents the vector of the accel-
erations, and the matrix m with dimensions N N represents the mass matrix
or inertial matrix of the structure. The elements mij of the mass matrix are the
mass influence coefficients. Their physical meaning is analogous to that of the
stiffness and damping influence coefficients, namely mij expresses the inertial
force applied along the degree of freedom i for unit acceleration along the
degree of freedom j, that is, u€ j ¼ 1, while all other accelerations are zero.
Fig. 10.4.9 shows the mass influence coefficients at the nodes of the frame,
when u€ 4 ¼ 1, u€ j ¼ 0 (j ¼ 1, 2, …, N , j 6¼ 4).
In actual structures, the mass is distributed. The model that considers the
mass lumped at certain points of the structure, for example, at the nodes of a
frame, approximates adequately the dynamic response of the structure. When
it is assumed that the lumped mass has no geometrical dimensions, that is, it
is simulated by a material particle, then its rotational inertia is zero and the
respective influence coefficients vanish.
Writing now Eq. (10.4.24) for all directions i ¼ 1, 2, …, N , we obtain
fI 1 + fD1 + fS1 ¼ p1 ðt Þ
fI 2 + fD2 + fS2 ¼ p2 ðt Þ
(10.4.38)
… … … …
fIN + fDN + fSN ¼ pN ðt Þ
or
f I + f D + f S ¼ p ðt Þ (10.4.39)
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 341
(a) (b)
FIG. 10.5.1 Continuous systems: (a) Cantilever beam, (b) Chimney fixed on the ground.
@CivilMethod
342 PART II Multi-degree-of-freedom systems
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 343
X
N X
N
¼ kij ui duj
i¼1 j¼1
where
Z L
00 00
kij ¼ EI ðx Þ i j dx (10.5.5)
0
(ii) Kinetic energy: The kinetic energy is due to the transverse displacements
and rotations of the mass elements. Thus, we have
Z Z h i2
1 L 1 L 0
m ðx Þ½u_ ðx, t Þ dx + I ðx Þ u_ ðx, t Þ dx
2
T¼ (10.5.6)
2 0 2 0
The second term in the above expression is due to the rotation of the
cross-sections and its contribution is small. In the following development,
without limiting the generality, we omit this term for the sake of simplicity.
The variation of the kinetic energy is
Z L
dT ¼ m ðx Þu_ ðx, t Þd u_ ðx, t Þdx (10.5.7)
0
@CivilMethod
344 PART II Multi-degree-of-freedom systems
Z ! !
L X
N X
N
dT ¼ m ðx Þ i u_ i j d u_ j dx
0 i¼1 j¼1
Z !
L X
N X
N
¼ m ðx Þ i j u_ i d u_ j dx (10.5.8)
0 i¼1 j¼1
X
N X
N
¼ mij u_ i d u_ j
i¼1 j¼1
where
Z L
mij ¼ m ðx Þ i j dx (10.5.9)
0
where
Z L
00 00
cijin ¼ c s I ðx Þ i j dx (10.5.15)
0
X
N X
N
¼ cijex u_ i duj
i¼1 j¼1
where
Z L
cijex ¼ c ðx Þ i j dx (10.5.18)
0
Finally, the virtual work due the external nonconservative load pðx, t Þ is
Z L
dWnc ¼
p
pðx, t Þd uðx, t Þdx (10.5.19)
0
where
Z L
pj ðt Þ ¼ pðx, t Þ j dx (10.5.21)
0
@CivilMethod
346 PART II Multi-degree-of-freedom systems
(iv) The potential energy of the conservative forces: The potential energy of the
conservative forces is due to the constant axial force P. Thus, we have
A ¼ Pe (10.5.24)
where e is the shorting of the elastic curve due to bending and is given as
Z L Z
1 L 0 2
e¼ ex dx ¼ ½u ðx, t Þ dx (10.5.25)
0 2 0
X
N X
N
¼ P kGij ui duj
i¼1 j¼1
where
Z L
0 0
kGij ¼ i j dx (10.5.28)
0
or in matrix form
u + Cu_ + ðK PKG Þu ¼pðt Þ
M€ (10.5.30)
The matrices M, C, K represent the mass, damping, and stiffness matri-
ces, respectively. Their elements are evaluated from Eqs. (10.5.9),
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 347
(a)
(b)
FIG. 10.6.1 Mixed system (a) and motion parameters (b).
@CivilMethod
348 PART II Multi-degree-of-freedom systems
x 2 x 3
1 ðx Þ ¼ 1 3 +2 ¼ 1 3x 2 + 2x 3 (10.6.2a)
L L
x x 2
2 ðx Þ ¼ L 1 ¼ Lx ðx 1Þ2 (10.6.2b)
L L
x 2 x 3
3 ðx Þ ¼ 3 2 ¼ 3x2 2x3 (10.6.2c)
L L
x 2 x
4 ð x Þ ¼ L 1 ¼ Lx2 ðx 1Þ (10.6.2d)
L L
where
x ¼ x=h (10.6.3)
For the element 2–3 holds: u1 ¼ af1 , u2 ¼ f1 , u3 ¼ u a f2 =2, u4 ¼ f2 ,
L ¼ 3a.
For the element 4–5 holds: u1 ¼ u + a f2 =2, u2 ¼ f2 , u3 ¼ 0, u4 ¼ 0,
L ¼ 3a.
The elastic energy is given by the expression
Z Z
1 3 1 5 1
EI ½u00 ðx, t Þ dx + EI ½u00 ðx, t Þ dx + kR f1
2 2 2
U¼ (10.6.4)
2 2 2 4 2
or using Eq. (10.6.1), we obtain
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 349
Z
1 3a
f1 , f2 ¼
U u, EI 00
1 ðx Þa f1 +
00
2 ðx Þf1 +
00
3 ðx Þ u a f2 =2 + 00 2
4 ðx Þf2 dx
2 0
Z
1 3a
1
00
u + a f2 =2 + 00 2 + kR f1
2
+ EI 1 ðx Þ 2 ðx Þf2 dx
2 0 2
(10.6.5)
Differentiating the above relation yields after evaluation of the integrals
∂U
¼ k11 u + k12 f1 + k13 f2 (10.6.6a)
∂u
∂U
¼ k21 u + k22 f1 + k23 f2 (10.6.6b)
∂f1
∂U
¼ k31 u + k32 f1 + k33 f2 (10.6.6c)
∂f2
where
8 EI 10 EI
k11 ¼ , k12 ¼ k21 ¼ , k13 ¼ k31 ¼ 0
9 a3 9 a2
121 EI 17 EI 38 EI
k22 ¼ , k23 ¼ k32 ¼ , k33 ¼
36 a 9 a 9 a
Therefore the stiffness matrix of the system is
2 3
32 40a 0
EI 4
k¼ 40a 121a2 17a 2 5 (10.6.7)
36a 3
0 17a 2 152a 2
(ii) Kinetic energy: The kinetic energy results from the expression
Z Z
1 3 1 5
½u_ ðx, t Þ dx + ½u_ ðx, t Þ dx
2 2
T¼ m m
2 2 2 4
(10.6.8)
1 1 1
+ I1 f_ 12 + m u_ 2 + IC f_ 22
2 2 2
or using Eq. (10.6.1), we obtain
1 Z 3a h i2
_ f_ 1 , f_ 2 ¼
T u, 1 ðx Þa f_ 1 + 2 ðx Þf_ 1 + 3 ðx Þ u_ a f_ 2 =2 +
m _
4 ð x Þf 2 dx
2 0
Z
1 3a h i2
+ 1 ðx Þ u_ + a f_ 2 =2 + 2 ðx Þf_ 2 dx
m
2 0
2 2
1 1 1
+ I1 f_ + m u_ + IC f_
2
2 1 2 2 2
(10.6.9)
@CivilMethod
350 PART II Multi-degree-of-freedom systems
where
2712 558
m11 ¼ m, m12 ¼ m21 ¼ ma, m13 ¼ m31 ¼ 0
840 840
2294 2 675 2 1832 2
m22 ¼ ma , m23 ¼ m32 ¼ ma , m33 ¼ ma
840 840 840
Therefore, the mass matrix of the system is
2 3
2712 558a 0
m 4
m¼ 558a 2294a 2 675a 2 5 (10.6.11)
840
0 675a 2 1832a 2
(iii) Generalized forces: The virtual work of the nonconservative forces is due
to the load pðt Þ. Thus, we have
Q1 d u + Q2 df1 + Q3 df2 ¼ pðt Þd u + a f2 =2
(10.6.12)
¼ pðt Þd u pðt Þadf2 =2
from which we obtain
Q1 ¼ pðt Þ
Q2 ¼ 0
Q3 ¼ pðt Þa=2
Hence the equation of motion of the mixed system is
2 38 9
€
m6
2712 558a 0 < u >
> =
2 7 f €
4 558a 2294a 675a 5
2
1
840 : € >
> ;
0 675a 2 1832a 2 f2
2 38 9 8 9 (10.6.13)
32 40a 0 >
< u >= >
< pðt Þ >
=
EI 6 2 7
+ 4 40a 121a 2
17a 5 f ¼ 0
36a3 >
: >
1
; > : >
;
0 17a 2 152a 2 f2 pðt Þa=2
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 351
@CivilMethod
352 PART II Multi-degree-of-freedom systems
uÞT pðt Þ
dWp ¼ duT pðt Þ ¼ ðRd
(10.7.12)
uT RT pðt Þ
¼ d
and
p ¼ d
dW ðt Þ
uT p (10.7.13)
p , we obtain
Because these two virtual works are equal, dWp ¼ d W
ðt Þ ¼ RT pðt Þ
p (10.7.14)
Example 10.7.1 The equation of motion of the system shown in Fig. 10.4.1 with
respect to point O is (see Eqs. 10.4.11a, 10.4.11b):
2 ma 3
m ( pðt Þ )
6 2 7 u€ + k11 k12 u
¼
4 ma 5 € a (1)
IO f k21 k22 f pðt Þ
2 2
where
12EI 6EI 4EI
k11 ¼ , k12 ¼ k21 ¼ , k22 ¼ (2)
h3 h2 h
Transform Eq. (1) with respect to the mass center C of the body.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 353
Solution
If u and f denote the displacement of the mass center and its rotation about it,
then the transformation relations result by geometrical consideration as
a
u ¼ u + f (3)
2
f ¼ f (4)
or
" a #
u 1 u
¼ 2 (5)
f 0 1 f
Hence
"
a# " #
1 1 0
R¼ 2 , RT ¼ a (6)
1
0 1 2
Then Eq. (10.7.10) gives
2 a 32 a 3
" # m m
1 0
¼ a 6 2 74 1 2 5
M 4 5
1 a
2 m IO 0 1
2 (7)
" #
m 0
¼
0 IC
where
a2
IC ¼ m (8)
6
is the moment of inertia with respect to the center of mass C .
Further, Eq. (10.7.9) gives
" #" #2 3
1 0 k11 k12 1 a
¼ a
K 4 25
1 k21 k22
2 0 1
" # (9)
k11 k12
¼
k21 k22
where
12EI
k11 ¼ k11 ¼ 3
h
a 6EI
k 12 ¼ k 21 ¼ k11 + k12 ¼ 3 ða + h Þ (10)
a 2
2 h
12EI a 2 ah h 2
k 22 ¼ k11 + ak 12 + k22 ¼ 3 + +
2 h 4 2 3
@CivilMethod
354 PART II Multi-degree-of-freedom systems
10.8 Problems
Problem P10.1 Formulate the equations of motion of the system shown in
Fig. P10.1. Data: k1 ¼ 3k, k2 ¼ 2k, k3 ¼ k, c1 ¼ c3 ¼ c, c2 ¼ 2c, m1 ¼ m2 ¼ m,
and m3 ¼ 2m.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 355
Problem P10.3 A rigid bar of total mass m is supported by the three springs
k1 , k2 , k3 and the damper c as shown in Fig. P10.3. The circular disc of mass
0:5m at the end D of the bar has a diameter 0:2L. Formulate the equation of
motion. Data: k2 ¼ 1:5k1 and k3 ¼ 2k1 .
@CivilMethod
356 PART II Multi-degree-of-freedom systems
Problem P10.5 The beam of Fig. P10.5 consists of the flexible and massless
part AB and the rigid part BC of mass m per unit length. The beam is clamped
at A while the hinged support at C is elastically restrained by the rotational
spring CR ¼ EI =10L. The beam is loaded by a concentrated moment M ðt Þ
applied at B. Formulate the equation of motion of the structure and give the
expressions for the evaluation of the support reactions.
Problem P.10.6 The structure of Fig. P10.6 consists of the two bars AB and
BC . The bar AB is massless and flexible and is simply supported at A while
per unit length and is elastically restrained at
the bar BC is rigid with mass m
the hinged end C by the rotational spring CR ¼ EI =10L. The structure is
loaded by a concentrated moment M ðt Þ applied at B. Formulate the equation
of motion of the structure and give the expressions for the evaluation of the
support reactions.
@CivilMethod
MDOF systems: Models and equations of motion Chapter 10 357
Problem P10.9 The thin spherical tank of Fig. P10.9 has a diameter R ¼ a=3
and is full with a liquid of density g. It is supported on the ground through nine,
hinged at both ends, massless rods of cross-sectional area A and modulus of
elasticity E. The points 10 ,20 ,30 lie on a horizontal circular ring of the spherical
tank at a height a from the ground and at a distance R=3 from the center of the
sphere. The mass of the tank is neglected.
@CivilMethod
358 PART II Multi-degree-of-freedom systems
@CivilMethod
Chapter 11
11.1 Introduction
The distributed mass is the usual case we encounter in actual structures, that is,
the actual structures should be modeled as continuous systems. This modeling
leads to partial differential equations with prescribed initial and boundary con-
ditions. The solution of such equations is a difficult mathematical problem, even
for individual components of structures (e.g., beams, plates, walls, etc.). On the
other hand, the method of global shape functions, which can provide an accept-
able solution, has limited capabilities due to the difficulty in choosing the global
shape functions [1]. In practice, this method cannot cope with conventional
structures such as a multistory frame or a grid. Besides, it cannot be pro-
grammed for automatic use on a computer. Therefore, neither method is
suitable for the dynamic analysis of realistic structures in engineering. At this
dead end, the finite element method (FEM) method gives a way out.
The FEM represents a major breakthrough in the field of computational
mechanics. Extensive literature on the general formulation of the FEM [2] as
well as on its application to dynamic problems is now available [3].
In the finite element approach, the structure being analyzed (e.g., frame,
wall, plate, shell, three-dimensional deformable body, or a combination of
them) is divided into a finite number of small subregions, elements, which
are interconnected at discrete points, the nodes, where the compatibility
condition for the displacements and the equilibrium of equivalent nodal forces
are ensured. After the discretization, we assume that the field function (the dis-
placement in this case) within each element varies according to a known law,
which is expressed as a superposition of shape patterns, the shape functions,
corresponding to unit values of the element nodal displacements. After that,
the elastic and kinetic energy as well as the virtual work of the nonconservative
nodal forces in terms of the nodal displacements, viewed as generalized coor-
dinates, are established. The equivalent nodal forces (elastic, inertial, damping,
and external forces) may result by the use of the Lagrange equations or the prin-
ciple of virtual work. Finally, the compatibility of the nodal displacements and
the equilibrium of the nodal forces give the differential equations of motion of
the structure. The FEM is presented below only for skeletal structures, that is,
structures consisting of straight-line elements (trusses, frames, grids) because
surface structures (walls, plates, shells) as well as three-dimensional (3D)
deformable bodies do not fall within the scope of this book.
In the deformed state at time t, the ends of the elements are displaced to points
j 0 , k 0 and the element occupies the position j 0 k 0 (Fig. 11.2.1).
The element has four degrees of freedom, the displacements u1 ðt Þ, u2 ðt Þ of
node j and the displacements u3 ðt Þ, u4 ðt Þ of node k. These quantities constitute
the nodal coordinates of the plane truss element. A typical point x of the ele-
ment axis undergoes two displacements, the axial displacement u ðx, t Þ and the
transverse one v ðx, t Þ. We consider the element as a generalized system and
apply the Ritz method presented in Section 10.5.1. Thus, we set the axial
displacement in the form
u ðx, t Þ ¼ u1 ðt Þ 1 ðx Þ + u 3 ðt Þ 3 ðx Þ (11.2.1)
where 1 ðx Þ and 3 ðx Þ are shape functions expressing the axial deformation for
u1 ¼ 1, u3 ¼ 0 and u1 ¼ 0, u3 ¼ 1, respectively. The functions 1 ðx Þ and 3 ðx Þ
can be established as follows.
The equilibrium of the element dx (see Fig. 11.2.2) yields
dN
¼0 (11.2.2)
dx
or taking into account that
N ¼ Aðx Þsx
¼ EAðx Þex (11.2.3)
d i
¼ EAðx Þ
dx
we write Eq. (11.2.2) as
d d i
EAðx Þ ¼0 (11.2.4)
dx dx
and after integration
Z
dx
i ¼ c1 + c2 (11.2.5)
Aðx Þ
The arbitrary constants c1 and c2 are evaluated from the boundary condi-
tions. It is obvious that for a variable cross-section, the determination of the
shape functions requires the evaluation of the integral in Eq. (11.2.5). This
relation, although it gives shape functions expressing the exact static axial
deformation, is not suitable for the automation of the method because different
shape functions have to be determined for elements with a different law of
variation of the cross-section. This difficulty is surpassed if we accept the
same shape functions for all elements, regardless of the variation law of the
cross-section and, indeed, those resulting from Eq. (11.2.4) for a constant
cross-section. In that case, we obtain
i ¼ c1 x + c2
which for 1 ð0Þ ¼ 1 and 1 ð LÞ ¼ 0 gives
1 ðx Þ ¼ 1 x, x ¼ x=L (11.2.6a)
and for 3 ð 0Þ ¼ 0 and 3 ð LÞ ¼ 1
3 ðx Þ ¼ x, x ¼ x=L (11.2.6b)
The shape functions (11.2.6a), (11.26b) can also result if we arbitrarily
accept a linear law of variation of axial displacement within the element.
The assumption of linear variation of the displacement is at the expense of accu-
racy. However, the error is acceptable when the element is small and the cross-
sectional variation is not intense. This becomes clear from Example 11.2.1.
Taking into account that the element is a straight line in the deformed state,
we can write the transverse displacement asa
v ðx, t Þ ¼ u2 ð1 x Þ + u4 x
(11.2.7)
¼ u2 2 ðx Þ + u4 4 ðx Þ
But it is
x0 x + u ðx, t Þ x=L + u ðx, t Þ=L u ðx, t Þ
x0 ¼ ¼ ¼ x+ (b)
L0 L + ðu3 u1 Þ 1 + ðu3 u1 Þ=L L
be neglected.
@CivilMethod
The finite element method Chapter 11 363
Obviously, it is 2 ðx Þ ¼ 1 ðx Þ and 4 ðx Þ ¼ 3 ðx Þ.
The equivalent nodal forces in the directions of the degrees of freedom can
be established using the method of the Lagrange equations or the principle of
virtual work. Both methods are presented in the following.
@CivilMethod
364 PART II Multi-degree-of-freedom systems
Z
1 L 0 0
2
U ðu1 , u3 Þ ¼ EAðx Þ u1 1 ðx Þ + u3 3 ðx Þ dx (11.2.13)
2 0
The nodal elastic forces result from Eq. (11.2.10) for i ¼ 1, 2, 3, 4. Thus,
we obtain
∂U
fS1 ¼ ¼ k11 u1 + k13 u3 (11.2.14a)
∂u1
∂U
fS2 ¼ ¼0 (11.2.14b)
∂u2
∂U
fS3 ¼ ¼ k31 u1 + k33 u3 (11.2.14c)
∂u3
∂U
fS4 ¼ ¼0 (11.2.14d)
∂u4
where
Z L
kij ¼ EAðx Þ 0i ðx Þ 0j ðx Þdx, i, j ¼ 1, 3 (11.2.15)
0
ments, respectively, and ke the stiffness matrix of the e truss element. Hence, the
element matrix ke for the plane truss element is defined as
2 3
k11 0 k13 0
6 7
60 0 0 07
6 7
k ¼6
e
7 (11.2.18)
6 k31 0 k33 0 7
4 5
0 0 0 0
Obviously, we deduce from Eq. (11.2.15) that kij ¼ kji . Hence, the stiffness
matrix is symmetric.
For an element with a constant cross-section Aðx Þ ¼ Ae and a length Le ,
Eq. (11.2.15) is integrated analytically and Eq. (11.2.18) yields
@CivilMethod
The finite element method Chapter 11 365
2 3
1 0 1 0
e6
EA 0 0 0 07
ke ¼ e 6 7 (11.2.19)
L 4 1 0 1 05
0 0 0 0
It should be noted that Eq. (11.2.10) expresses the Castigliano theorem.
Therefore, the previous method for establishing the stiffness matrix is identical
to the so-called energy method.
(ii) Nodal inertial forces and mass matrix of the truss element
In the FEM, the equivalent inertial forces are obtained by two different assump-
tions of the mass distribution on the element: the consistent mass assumption,
which assumes a continuous distribution of the mass on the element, and the
lumped mass assumption, which lumps the mass at its nodes. The inertial mass
matrices resulting from both assumptions are derived below.
(a) Consistent mass matrix
During the motion, the infinitesimal mass m ðx Þdx undergoes the two displace-
ments u ðx, t Þ and v ðx, t Þ. Therefore, the kinetic energy of the truss element will
be given by the expression
Z n o
1 L
T¼ m ðx Þ ½u_ ðx, t Þ2 + ½v_ ðx, t Þ2 dx (11.2.20)
2 0
or using Eqs. (11.2.1), (11.2.7)
Z n o
1 L 2 2
T ðu_ 1 , …, u_ 4 Þ ¼ m ðx Þ ½u_ 1 1 ðx Þ + u_ 3 3 ðx Þ + ½u_ 2 1 ðx Þ + u_ 4 3 ðx Þ dx
2 0
(11.2.21)
The inertial forces result from Eq. (11.2.9) for i ¼ 1, 2, 3, 4. Thus, after
performing the differentiations we obtain
d ∂T ∂T
fI 1 ¼ ¼ m11 u€1 + m13 u€3 (11.2.22a)
dt ∂u_ 1 ∂u1
d ∂T ∂T
fI 2 ¼ ¼ m22 u€2 + m24 u€4 (11.2.22b)
dt ∂u_ 2 ∂u2
d ∂T ∂T
fI 3 ¼ ¼ m31 u€1 + m33 u€3 (11.2.22c)
dt ∂u_ 3 ∂u3
d ∂T ∂T
fI 4 ¼ ¼ m42 u€2 + m44 u€4 (11.2.22d)
dt ∂u_ 4 ∂u4
where
Z L
mij ¼ mi + 1, j + 1 ¼ m ðx Þ i ðx Þ j ðx Þdx, i, j ¼ 1, 3 (11.2.23)
0
@CivilMethod
366 PART II Multi-degree-of-freedom systems
Therefore, the kinetic energy of the truss element is given by the expression
1 1
T ¼ m1 u_ 1 2 + u_ 2 2 + m2 u_ 3 2 + u_ 4 2 (11.2.29)
2 2
@CivilMethod
The finite element method Chapter 11 367
The inertial forces result from Eq. (11.2.9) for i ¼ 1,2, 3, 4. Hence, after
performing the differentiations, we obtain
d ∂T ∂T
fI 1 ¼ ¼ m11 u€1 (11.2.30a)
dt ∂u_ 1 ∂u1
d ∂T ∂T
fI 2 ¼ ¼ m22 u€2 (11.2.30b)
dt ∂u_ 2 ∂u2
d ∂T ∂T
fI 3 ¼ ¼ m33 u€3 (11.2.30c)
dt ∂u_ 3 ∂u3
d ∂T ∂T
fI 4 ¼ ¼ m44 u€4 (11.2.30d)
dt ∂u_ 4 ∂u4
where
m11 ¼ m22 ¼ m1 , m33 ¼ m44 ¼ m2 (11.2.31)
Eqs. (11.2.30a)–(11.2.3d) are written in matrix form as
8 9 2 38 9
fI 1 > u€
> > 1>
>
m11 0 0 0
>
> > >
> >
>
>
< fI 2 = 6 7>> >
60 m22 0 0 7 u€2 =
<
6
¼6 7 (11.2.32)
>
>
> fI 3 >
>
> 40 0 m33 0 7 5>>
> u€3 >
>
>
> ;
: > : >
> ;
fI 4 0 0 0 m44 u€4
Therefore, the element mass matrix me is defined as
2 3
m11 0 0 0
60 0 7
6 m22 0 7
me ¼ 6 7 (11.2.33)
40 0 m33 0 5
0 0 0 m44
We observe that the lumped mass assumption results in a diagonal mass matrix.
For an element with constant mass, m ðx Þ ¼ m Eqs. (11.2.28a),
(11.2.28b) give
e Le =2
m11 ¼ m22 ¼ m33 ¼ m44 ¼ m (11.2.34)
@CivilMethod
368 PART II Multi-degree-of-freedom systems
The damping forces are obtained as generalized forces from Eqs. (11.2.8),
(1.8.8) for i ¼ 1, 2, 3, 4. Thus we have
fD1 ¼ Q1 ¼ c11 u_ 1 + c13 u_ 3 (11.2.41a)
fD2 ¼ Q2 ¼ 0 (11.2.41b)
fD3 ¼ Q3 ¼ c31 u_ 1 + c33 u_ 3 (11.2.41c)
fD4 ¼ Q4 ¼ 0 (11.2.41d)
@CivilMethod
The finite element method Chapter 11 369
The minus sign in the above relations is due to the fact that they are shifted
to the left side of Eq. (11.2.8) in order that their positive direction is the same
with that of fIi and fSi .
Eqs. (11.2.41a)–(11.2.41d) are written in matrix form as
8 9 2 38 9
> fD1 > c11 0 c13 0 > u_ 1 >
>
> > >
> > >
> >
>
< fD2 = 6 7> >
6 0 0 0 0 7< u_ 2 =
¼66 7 (11.2.42)
> fD3 > 7 u_ >
>
> >
> 4 c31 0 c33 0 5> >
> 3>>
> ;
: > > ;
: >
fD4 0 0 0 0 u_ 4
or
f eD ¼ ce u_ e (11.2.43)
where f eD , u_ e are the vectors of the nodal damping forces and nodal velocities,
respectively, and ce is the damping matrix of the e truss element. Therefore, the
element damping matrix ce is defined as
2 3
c11 0 c13 0
60 0 0 07
6 7
ce ¼ 6 7 (11.2.44)
4 c31 0 c33 0 5
0 0 0 0
Obviously, we deduce from Eq. (11.2.40) that cij ¼ cji . Hence, the damping
matrix is symmetric.
For an element with a constant cross-section, Aðx Þ ¼ Ae , Eq. (11.2.40) are
integrated analytically and Eq. (11.2.44) becomes
2 3
1 0 1 0
6 7
cs A e 6 0 0 0 0 7
ce ¼ e 6 7 (11.2.45)
L 6 4 1 0 1 0 5
7
0 0 0 0
The evaluation of the matrix ce requires knowledge of the coefficient cs , the
determination of which is practically not feasible. The damping matrix in engi-
neering praxis, as we will see in Section 12.11.3, is constructed by assuming
known values of the damping ratio for each vibration mode, which are estimated
by experiments on similar structures. Nevertheless, we presented this approach
for the completeness of FEM, which is structured through the element approach.
(iv) Equivalent nodal loads of the truss element
The truss element can also be subjected to axial external loading, which may
consist of a distributed load pðx, t Þ and a finite number of concentrated loads
Pk ðt Þ, k ¼ 1, 2, …, K , acting at the points xk of the element, Fig. 11.2.4.
@CivilMethod
370 PART II Multi-degree-of-freedom systems
Fig. 11.2.4 Loading and equivalent nodal loads of a plane truss element.
The nodal load forces are obtained from Eq. (11.2.8) for i ¼ 1, 2, 3, 4, that is,
Q1 ¼ p1 , Q2 ¼ 0, Q3 ¼ p3 , Q4 ¼ 0. Hence, the vector of the equivalent nodal load
forces for the e element is
p e ð t Þ ¼ f p 1 ð t Þ 0 p 3 ð t Þ 0g T (11.2.48)
Example 11.2.1 Compute the stiffness, mass, and damping matrices as well as
the vector of the nodal load forces of an element with variable cross-section and
compare them to those of the element with a constant cross-section equal to the
mean value of the variable cross-section.
Data: Aðx Þ ¼ A0 1 + x x2 ,
x ¼ x=L, pðx, t Þ ¼ p0 ðt Þ 1 + x x . Modulus of elasticity E, material density
2
@CivilMethod
The finite element method Chapter 11 371
@CivilMethod
372 PART II Multi-degree-of-freedom systems
Damping matrix. The elements of the damping matrix are computed from
the Eq. (11.2.40), which gives
2 3
1 0 1 0
6 0 0 0 07
cs A
ce ¼ 6 7 (9)
L 4 1 0 1 0 5
0 0 0 0
Nodal Load forces. They are computed from Eq. (11.2.47), which gives
pe ðt Þ
pe ðt Þ ¼ f 1 0 1 0 gT (10)
2
where pe ðt Þ ¼ 7p0 ðt ÞL=6 is the total load of the element.
@CivilMethod
The finite element method Chapter 11 373
Further, taking into account that fS2 ¼ fS4 ¼ 0, we can write Eq. (11.2.55)
8 9 2 38 9
> fS1 > k11 0 k13 0 > u1 >
> >
> 6 > >
> =
< 7>< > =
fS2 6 0 0 0 0 7 u2
¼6 7 (11.2.57)
>
>
> fS3 >
>
>
4 k31 0 k33 0 5> >
> u3 >
>
>
: ; : ;
fS4 0 0 0 0 u4
which is identical to Eq. (11.2.16).
(ii) Nodal inertial forces and mass matrix of the truss element
The internal inertial force of mass element m ðx Þdx is
fI ðx, t Þ ¼ m ðx Þ½u€ðx, t Þ + v€ððx, t Þdx (11.2.58)
and the virtual work produced when the end displacements are given a virtual
displacement ðdu1 , du2 , du3 , du4 Þ is
Z L
dWin ¼ m ðx Þ½u€ðx, t Þdu + v€ðx, t Þdv dx (11.2.59)
0
where
1
N¼½ 1 2 3 4¼ ½ ð1 xÞ ð1 x Þ x x (11.2.61)
L
The integral in Eq. (11.2.60) gives the mass matrix of the truss element
@CivilMethod
374 PART II Multi-degree-of-freedom systems
2 3
m11 0 m13 0
Z 6 7
L 60 m22 0 m24 7
me ¼ m ðx ÞNT Ndx ¼ 6
6m 0
7 (11.2.62)
0 4 31 m33 0 7 5
0 m42 0 m44
where
Z L
mij ¼ mi + 1, j + 1 ¼ m ðx Þ i ðx Þ j ðx Þdx, i, j ¼ 1, 3 (11.2.63)
0
@CivilMethod
The finite element method Chapter 11 375
The previous relation is readily obtained by projecting the broken line OAP
successively onto the axes Ox and Oy and taking into account that OA ¼ a1 ,
AP ¼ a2 , OC ¼ a1 , and CP ¼ a2 (see Fig. 11.2.5).
Eq. (11.2.67) is written as
a ¼L
a (11.2.68)
where
a1 a1 cos f sin f
a¼ , a ¼ , L¼ (11.2.69)
a2 a2 sin f cos f
The matrix L is referred to as the transformation matrix of the components
of a plane vector.
It can be readily shown that
L1 ¼LT (11.2.70)
that is, the matrix L is orthonormal. Thus Eq. (11.2.68) is readily inverted to
a ¼LT a (11.2.71)
!
We now consider the element e, whose longitudinal axis j k defines the
right-handed system of axes jxy. Let O xy be another arbitrarily chosen right-
handed system of axes and let f be the angle between x and x axes, that is,
the angle that the axis x sweeps when it is rotated counterclockwise to coincide
with the x axis, Fig. 11.2.6. The axes x, y are called the global axes while the
axes x, y defined by the truss element are called the local axes. Any quantity,
force, or displacement referred to the global axes will be henceforth designated
by an overbar.
Fig. 11.2.6 Nodal displacements of the truss element in global and local axes.
@CivilMethod
376 PART II Multi-degree-of-freedom systems
f e ¼ðRe ÞT f e (11.2.77)
I I
f e ¼ðRe ÞT f e (11.2.78)
D D
pe ðt Þ ¼ðR Þ p ðt Þ
e T e
(11.2.79)
@CivilMethod
Fig. 11.2.7 Nodal forces of the truss element in global and local axes.
The finite element method Chapter 11 377
¼ ðRe ÞT ke ue (11.2.80)
e T e
¼ ðR Þ k Re ue
or
f e ¼ke ue (11.2.81)
S
where
k ¼ ðRe ÞT ke Re :
e
(11.2.82)
is the element stiffness matrix with respect to the global axes.
Similarly, we obtain
f e ¼m
e ue (11.2.83)
I
f D ¼
ce ue (11.2.84)
where
e ¼ ðRe ÞT me Re
m (11.2.85)
ce ¼ ðRe ÞT ce Re (11.2.86)
ð11:2:88Þ
@CivilMethod
378 PART II Multi-degree-of-freedom systems
^f e ¼ ðae ÞT f e (11.2.90)
S S
Starting with Eq. (11.2.90) and using the transformations (11.2.87), we obtain
^f e ¼ ðae ÞT f e
S S
e T e e
¼ ða Þ k u
(11.2.91)
¼ ðae ÞT k ae u
e
¼K^ e u
where
^ e ¼ ðae ÞT ke ae
K (11.2.92)
denotes the enlarged stiffness matrix with dimensions N N . Obviously, the
places of the elements of k in K
^ e are:
e
^f e ^ e u_
¼C (11.2.94)
D
b. The matrix a is not square, therefore it cannot be inverted. The inverse transformation (11.2.89)
can be proved either by inspection or by considering the equality of the work produced by the nodal
forces in the two systems of diplacements ue and u. Indeed, according to Eq. (11.2.87) we have
f e ¼ae^f e
where f and ^f e are element force vectors with a dimension 4 and N , respectively.
e
@CivilMethod
which results in Eq. (11.2.89).
The finite element method Chapter 11 379
^e ðt Þ ¼ ðae ÞT pe ðt Þ
p (11.2.95)
where ^f e , ^f e ,
and p ^e ðt Þ denote the enlarged nodal inertial, damping matrices,
I D
and load vector, respectively, while
^ e ¼ ðae ÞT m
M e ae (11.2.96)
^e
C ¼ ða Þ e T e e
ca (11.2.97)
ðt Þ with a dimen-
Finally, all nodal loads are assembled into a load vector P
sion N , corresponding to u. Hence, the loads applied to node i will take the
places P2i1 and P2i in P
ðt Þ.
The equations of motion can be derived using the method of the Lagrange
equations with generalized coordinates the components of the vector u. Thus,
we have:
Elastic energy. The elastic energy of the e element is
1
Ue ¼ uT ^f S
e
2 (11.2.98)
1 ^e
¼ uT K u
2
and the total elastic energy is
1X Ne
^ e u
U¼ uT K
2 e¼1
1
¼ uT K u (11.2.99)
2
1X N X N
ij ui uj
¼ K
2 i¼1 j¼1
where
X
Ne
¼
K ^e
K (11.2.100)
e¼1
represents the stiffness matrix of the structure with Ne denoting the number of
all elements.
Kinetic energy. Similarly, the kinetic energy of the structure results as the
sum of the kinetic energies of all elements
1X Ne
T ^e
T¼ u_ M u_
2 e¼1
1 T
¼ u_ M u_
2
1X N X N
ij u_ i u_ j
¼ M (11.2.101)
2 i¼1 j¼1
@CivilMethod
380 PART II Multi-degree-of-freedom systems
where
X
Ne
¼
M ^e
M (11.2.102)
e¼1
uT ^f D
e
dWDe ¼ d
(11.2.103)
¼ d ^ e u_
uT C
and the total virtual work of the damping forces is
X
Ne
dWD ¼ ^ e u_
uT C
d
e¼1
¼ d u_
uT C
X
N X
N
¼ ij u_ j d
C ui (11.2.104)
i¼1 j¼1
where
X
Ne
¼
C ^e
C (11.2.105)
e¼1
u pðt Þ
¼ d T
where
X
Ne
ðt Þ +
pðt Þ ¼ P ^ e ðt Þ
p (11.2.107)
e¼1
@CivilMethod
The finite element method Chapter 11 381
X N
d ∂T
¼ Mij u€j (11.2.108b)
dt ∂u_ i j¼1
∂U X N
¼ Kij uj (11.2.108c)
∂
u i j¼1
ij u_ j + pi ðt Þ
Qi ¼ C (11.2.108d)
and the equation of motion becomes
X
N X
N X
N
ij u€j +
M ij u_ j +
C ij uj ¼ pi ðt Þ
K (11.2.109)
j¼1 j¼1 j¼1
or
M u_ + K
u€ + C u ¼
pðt Þ (11.2.110)
Obviously, Eq. (11.2.110) has the same form as that obtained using the
methods discussed in Chapter 10.
@CivilMethod
382 PART II Multi-degree-of-freedom systems
k ¼ ðRe ÞT ke Re
e
e ¼ ð Re Þ T m e Re
m
ce ¼ ðRe ÞT ce Re
pe ðt Þ ¼ ðRe ÞT pe ðt Þ
@CivilMethod
The finite element method Chapter 11 383
C
e ¼ VT CV (11.2.114b)
K
e ¼VT KV (11.2.114c)
pðtÞ ¼VT pðtÞ
e (11.2.114d)
@CivilMethod
384 PART II Multi-degree-of-freedom systems
e€ + C
eu
M eeu_ + Ke
e u ¼e
pðt Þ
which are partitioned after separating the known from the unknown
displacements
ð11:2:115Þ
or
" #
" #
" #
e ff M
M e fs e€f
u e ff C
C e fs e_ f
u e ff K
K e fs e
uf e
p f ðt Þ
+ + ¼
e e
Msf Mss u€s
e e e
Csf Css u_ s
e e e
Ksf Kss e
us e
p s ðt Þ
(11.2.116)
Performing the matrix multiplications in the previous equation yields
M u€f + C
e ff e u_ f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e (11.2.117)
M u€f + M
e sf e u€s + C
e ss e u_ f + C
e sf e u_ s + K
e ss e e sf e e ss e
uf + K us ¼ e
p s ðt Þ (11.2.118)
where it was set
pf∗ ðt Þ ¼ e
e pf ðt Þ M u€s C
e fs e u_ s K
e fs e e fs e
us (11.2.119)
Eq. (11.2.117) is the equation of motion of the supported truss. Its solution
gives the vector of the unknown displacements e uf , which is then inserted in
Eq. (11.2.118) to obtain the vector of nodal forces e
ps ðt Þ. Note that this vector con-
tains the equivalent element nodal loads contributing to the supported node. Obvi-
ously, when the support displacements do not depend on time, Eq. (11.2.119)
becomes
pf∗ ðt Þ ¼ e
e e fs e
pf ðt Þ K us (11.2.120)
@CivilMethod
The finite element method Chapter 11 385
M u€f + C
e sf e u_ f + K
e sf e e sf e e ss e
uf + K us ¼ e
ps ðt Þ (11.2.121)
Finally, when the supports do not yield, the foregoing equations are further
simplified and become
pf∗ ðt Þ ¼ e
e p f ðt Þ (11.2.122)
M u€f + C
e sf e u_ f + K
e sf e e sf e
uf ¼ e
p s ðt Þ (11.2.123)
Example 11.2.2 Formulate the equations of motion of the truss in Fig. E11.1a
when (i) It is subjected to the external loading. (ii) The support 1 is
subjected to the vertical motion ug ðt Þ. Assumed data: Nodal coordinates:
1ð0, 0Þ, 2ð3, 3:5Þ, 3ð6, 0Þ; cross-sectional areas of the bars: A1 ¼ 1:5A,
A2 ¼ A3 ¼ A; distributed load on the element 1: pðx, t Þ ¼ P ðt Þ=L1 ; damping
coefficient: cs ¼ 0, modulus of elasticity E; material density: r, lumped mass
assumption.
(a)
Fig. E11.1a Plane truss in Example 11.2.2.
Solution
(i) Equation of motion under the external loading.
(b)
Fig. E11.1b Nodal displacements of the plane truss in Example 11.2.2.
@CivilMethod
386 PART II Multi-degree-of-freedom systems
The truss has n ¼ 3 nodes, hence the free structure has N ¼ 2n ¼ 6 degrees of
freedom (see Fig. E11.2b). The numbering of the elements and the positive
direction of the local axes are shown in Fig. E11.1b.
e xj xk D
x yj yk D
y Le cos fe sinfe
1 0 6 6 0 0 0 6 1 0
2 0 3 3 0 3.5 3.5 4.61 0.651 þ0.759
3 3 6 3 3.5 0 3.5 4.61 0.651 0.759
1. Computation of ke , me , ce , pe ðt Þ, Re , e ¼ 1, 2, 3
Matrices ke . The elements have constant cross-section. Thus Eq. (11.2.19) gives
2 3 2 3
0:25 0 0:25 0 0:217 0 0:217 0
6 0 0 0 07 6 0 0 0 07
k1 ¼ EA6 7 2 6
4 0:25 0 0:25 0 5, k ¼ k ¼ EA4 0:217 0 0:217 0 5
3 7
0 0 0 0 0 0 0 0
Matrices me . The elements have constant cross section. Thus, we obtain
from Eq. (11.2.35)
2 3 2 3
4:5 0 0 0 2:305 0 0 0
6 0 4:5 0 0 7 60 2:305 0 0 7
m1 ¼ rA6 7
4 0 0 4:5 0 5, m ¼ m ¼ rA4 0
2 3 6 7
5
0 2:305 0
0 0 0 4:5 0 0 0 2:305
@CivilMethod
The finite element method Chapter 11 387
2 3
0:651 0:759 0 0
6 0:759 0:651 0 0 7
R3 ¼ 6
40
7
0 0:651 0:759 5
0 0 0:759 0:651
2. Computation of k , m
e
e , pe ðt Þ, e ¼ 1, 2, 3
Matrices k . Eq. (11.2.82) gives
e
2 3
0:25 0 0:25 0
6 0 0 0 07
k1 ¼ EA6
4 0:25 0 0:25
7,
05
0 0 0 0
2 3
0:092 0:107 0:092 0:107
6 0:107 0:125 0:107 0:125 7
k ¼ EA4
2 6 7
0:092 0:107 0:092 0:107 5
0:107 0:125 0:107 0:125
2 3
0:092 0:107 0:092 0:107
6 0:107 0:125 0:107 0:125 7
k3 ¼ EA64 0:092 0:107
7
0:092 0:107 5
0:107 0:125 0:107 0:125
e . Eq. (11.2.85) gives
Matrices m
2 3 2 3
4:5 0 0 0 2:305 0 0 0
6 0 4:5 0 0 7 60 2:305 0 0 7
m ¼ rA4
1 6 7 6
, m ¼ m ¼ rA4
2 3 7
0 0 4:5 0 5 0 0 2:305 0 5
0 0 0 4:5 0 0 0 2:305
^ e, M
3. Computation K ^ e, p
^e ðt Þ, e ¼ 1, 2, 3
The element assembly matrices ae are
2 3 2 3 2 3
1 0 0 0 0 0 1 0 0 0 0 0 0 0 1 0 0 0
6 0 1 0 0 0 07 6 07 6 07
a1 ¼ 6 7, a2 ¼ 6 0 1 0 0 0 7, a3 ¼ 6 0 0 0 1 0 7
40 0 0 0 1 05 40 0 1 0 0 05 40 0 0 0 1 05
0 0 0 0 0 1 0 0 0 1 0 0 0 0 0 0 0 1
@CivilMethod
388 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 389
M
4. Computation of the total matrices K, and total load vector pðt Þ of
the truss
2 3
0:342 0:107 0:092 0:107 0:25 0
6 7
6 0:107 0:125 0:107 0:125 0 0 7
6 7
X e 6 7
3 6 0:092 0:107 0:184 0 0:092 0:107 7
¼
K ^ ¼ EA6
K 7
6 0:107 0:125 0 0:107 0:125 7
6 0:25 7
e¼1
6 7
6 0:25 0:092 0:107 0:342 0:107 7
4 0 5
0 0 0:107 0:125 0:107 0:125
2 3
6:805 0 0 0 0 0
6 7
60 6:805 0 0 0 0 7
6 7
X e
3 60 0 4:610 0 0 0 7
¼
M ^ ¼ rA6
M 6
7
7,
60 0 0 4:610 0 0 7
e¼1 6 7
6 7
40 0 0 0 6:805 0 5
0 0 0 0 0 6:805
8 9 8 9 8 9
>
> P1 > > >
> 0:5P ðt Þ >
> >
> P1 + 0:5P ðt Þ >
>
> > > >
> > >
>
>
> P2 > >
>
>
>
> 0 >
>
>
>
> P2 >
>
>
X3 < = < = < =
ðt Þ + P ð t Þ 0 P ð t Þ
pðt Þ ¼ P ^ e ðt Þ ¼
p + ¼
> 3P ðt Þ >
> > >
> 0 >
> > 3P ðt Þ >
>
> > > > > > >
> > >
> 0:5P ðt Þ > > >
e¼1
> 0
> > >
> >
> >
> 0:5P ðt Þ >
>
: ; : ; : ;
P6 0 P6
@CivilMethod
390 PART II Multi-degree-of-freedom systems
– – –
– – –
– – –
– – –
– –
– –
e ff , M
6. Formulation of the matrices M e fs , M
e sf , M
e ss , K
e ff , K
e fs , K
e sf , K
e ss and
pf ðt Þ, e
vectors e ps ðt Þ
These matrices are obtained after partitioning M, e K,
e e pðt Þ on the basis of the
previously indicated partition. Thus, we have
2 3 2 3
4:610 0 0 0 0 0
Me ff ¼ rA4 0 4:610 0 5, M e fs ¼ M
e sf ¼ rA4 0 0 05
0 0 6:805 0 0 0
2 3
6:805 0 0
Me ss ¼ rA4 0 6:805 0 5
0 0 6:805
@CivilMethod
The finite element method Chapter 11 391
2 3 2 3
0:184 0 0:092 0:092 0:107 0:107
e ff ¼ EA6
K 4 0 0:25
7 e
0:107 5, K
6 7
fs ¼ EA4 0:107 0:125 0:125 5
0:092 0:107 0:342 0:25 0 0:107
2 3 2 3
0:092 0:107 0:25 0:342 0:107 0
e sf ¼ EA4 0:107 0:125 0
K 5, K e ss ¼ EA4 0:107 0:125 0 5
0:107 0:125 0:107 0 0 0:125
8 9 8 9
< P ðt Þ = < P1 + 0:5P ðt Þ =
e
pf ðt Þ ¼ 3P ðt Þ , e
ps ðt Þ ¼ P2
: ; : ;
0:5P ðt Þ P6
The solution of the equation of motion yields the vector euf and
Eq. (11.2.118) the nodal forces at the supports
8 9 2 38 9
<e ps1 = 0:125 0 e1 =
0:107 < u
ps2 ¼ EA4 0
e 0:092 0:092 5 ue2
: ; : ;
e
ps3 0:125 0:107 0 e3
u
Hence, the reactions at the supports are
8 9 8 9 8 9
< P1 = < e
ps1 = < 0:5P ðt Þ =
P ¼ ep 0
: 2 ; : s2 ; : ;
P6 e
ps3 0
(ii) Equation of motion due to the vertical motion ug ðt Þ of the support 1.
In this case it is
8 9 8 9
< P1 = <0 =
e
pf ðt Þ ¼ 0, e
ps ðt Þ ¼ P2 , e
us ¼ ug ðt Þ
: ; : ;
P6 0
@CivilMethod
392 PART II Multi-degree-of-freedom systems
pf∗ ðt Þ ¼ M
e u€s C
e fs e u_ s K
e fs e e fs e
us
or 8 9
< 0:107 =
pf∗ ðt Þ ¼ EA 0:125 ug ðt Þ
e
: ;
0
Hence, the equation of motion becomes
2 38 € 9 2 38 u 9
4:610 0 0 >
<u e1 >= E 0:184 0 0:092 >< e1 >
=
40 4:610 0 5 u e€2 + 4 0 0:250 0:107 5 ue2
> > ; r 0:092 0:107 0:342 > : >
0 0 6:805 : u e€3 e3
u
;
8 9
0:107 =
E<
¼ 0:125 ug ðt Þ
r: ;
0
and Eq. (11.2.123) gives the reactions
8 9 2 38 9 8 9
< P1 = e1 =
0:092 0:107 0:250 < u <0 =
P2 ¼ EA4 0:107 0:125 0 5 u
e2 + rA 6:805 u€g
: ; : ; : ;
P6 0:107 0:125 0:107 e3
u 0
8 9
< 0:107 =
+ EA 0:125 ug
: ;
0
Remark: The displacement vector u ef ¼ f u
e1 u
e2 u e3 gT can be converted to
the original vector of the free nodal displacements, that is, u ¼ f u3 u4 u5 gT
by inverting the transformations that have created uef .
@CivilMethod
The finite element method Chapter 11 393
@CivilMethod
394 PART II Multi-degree-of-freedom systems
5 ðx Þ ¼ 3x 2x3
2
(11.3.7c)
6 ðx Þ ¼ L x + x
2 3
(11.3.7d)
where x ¼ x=L.
Fig. 11.3.2 Shape functions of the transverse deformation of the beam element.
@CivilMethod
The finite element method Chapter 11 395
Fig. 11.3.3 Positive direction of the nodal forces af the plane frame element.
@CivilMethod
396 PART II Multi-degree-of-freedom systems
2 3
k11 0 0 k14 0 0
60 k22 k23 0 k25 k26 7
6 7
60 k32 k33 0 k35 k36 7
k ¼6
e
6 k41
7 (11.3.16)
6 0 0 k44 0 0 7 7
40 k52 k53 0 k55 k56 5
0 k62 k63 0 k65 k66
Obviously, we deduce from Eqs. (11.3.13a), (11.3.13b) that kij ¼ kji . Hence,
the stiffness matrix is symmetric.
For an element with a constant cross-section Aðx Þ ¼ Ae , I ðx Þ ¼ I e and
length Le , Eqs. (11.3.13a), (11.3.13b) are integrated analytically and
Eq. (11.3.16) becomes
2 3
EA EA
6 L 0 0 0 0 7
6 L 7
6 12EI 6EI 12EI 6EI 7
6 0 0 7
6 L3 L2 L3 L2 7
6 2EI 7
6 0 6EI 4EI
6EI 7
6 0 7
k ¼6
e 6 L 2 L L 2 L 7 (11.3.17)
EA EA 7
6 0 0 0 0 7
6 L L 7
6 7
6 12EI 6EI 12EI 6EI 7
6 0 3 2 0 7
6 L L L3 L2 7
4 6EI 2EI 6EI 4EI 5
0 0
L2 L L2 L
Note that the superscript e in Ae , I e , and Le has been dropped from the ele-
ments of the matrix in Eq. (11.3.17) for the sake of simplicity of the expressions.
(ii) Nodal inertial forces and mass matrix of the beam element
As in the case of the truss, the equivalent inertial nodal forces for the beam ele-
ment are obtained by two different assumptions of the mass distribution on
the element: the consistent mass assumption, which assumes a continuous dis-
tribution of the mass on the element, and the lumped mass assumption, which
lumps the mass at its nodes. The inertial mass matrices resulting from both
assumptions are derived below.
(a) Consistent mass matrix
During the motion, the infinitesimal mass m ðx Þdx undergoes the two displace-
ments u ðx, t Þ and v ðx, t Þ. Therefore, the kinetic energy of the beam element will
be given by the expression.c
c. Due to bending, the cross-sections of the beam rotate by an angle v 0 ðx, t Þ. Under this rotation the
mass element m ðx Þdx exhibits a kinetic energy rðx ÞI ðx Þv_ 0 ðx, t Þdx=2, where rðx Þ is the mass
density of the material. Therefore, Eq. (11.3.18) should also include the term
Z Z
1 L 1 L
rðx ÞI ðx Þv_ 0 ðx, t Þdx ¼ rðx ÞI ðx Þ u_ 2 02 ðx Þ + u_ 3 03 ðx Þ + u_ 5 05 ðx Þ + u_ 6 06 ðx Þ dx
2 0 2 0
@CivilMethod
However, it can be shown that the contribution of this term is small and it can be neglected.
398 PART II Multi-degree-of-freedom systems
Z L n o
1
T¼ m ðx Þ ½u_ ðx, t Þ2 + ½v_ ðx, t Þ2 dx (11.3.18)
2 0
or
€e
f eI ¼ me u (11.3.23)
@CivilMethod
The finite element method Chapter 11 399
€e are the vectors of the nodal inertial forces and the nodal acceler-
where f eI , u
ations, respectively, and me the mass matrix of the e beam element. Therefore,
the mass matrix me is defined as
2 3
m11 0 0 m14 0 0
60 m22 m23 0 m25 m26 7
6 7
60 m35 m36 7
6 m32 m33 0 7
m ¼6
e
7 (11.3.24)
6 m41 0 0 m44 0 0 7
6 7
40 m52 m53 0 m55 m56 5
0 m62 m63 0 m65 m66
Obviously, we deduce from Eq. (11.3.21) that mij ¼ mji . Hence, the mass
matrix is symmetric.
Eq. (11.3.21) are integrated
For an element with constant mass, m ðx Þ ¼ m,
analytically and Eq. (11.3.24) becomes
2 3
140 0 0 70 0 0
60 156 22L 0 54 13L 7
6 7
e 60 2
3L 27
m 6 22L 4L 0 13L 7
me ¼ 6 7 (11.3.25)
420 6 70 0 0 140 0 0 7
6 7
40 54 13L 0 156 22L 5
0 13L 3L 0 2
22L 4L2
where m e ¼ m e Le is the total mass of the e element. Note that the superscript
e
e in L has been dropped from the elements of the matrix in Eq. (11.3.25) for the
sake of simplicity of the expressions.
The shape functions employed to derive the mass matrix are the same as
those employed to derive the stiffness matrix. The mass matrix resulting in this
way is referred to as the consistent mass matrix.
(b) Lumped mass matrix
According to this assumption, the mass of the element is concentrated at its
nodes by static considerations, that is, they are obtained as the reactions of a
simply supported beam under the load m ðx Þ (see Fig. 11.2.3). Thus, we have
Z L
m1 ¼ m ðx Þð1 x Þdx (11.3.26a)
0
Z L
m2 ¼ m ðx Þxdx (11.3.26b)
0
Therefore, the kinetic energy of the beam element will be given by the
expression
1 1
T ¼ m1 u_ 1 2 + u_ 2 2 + m2 u_ 4 2 + u_ 5 2 (11.3.27)
2 2
@CivilMethod
400 PART II Multi-degree-of-freedom systems
The nodal forces result from Eq. (11.2.9) for i ¼ 1, 2, 3, 4. Thus, after
performing the differentiations, we obtain
d ∂T ∂T
fI 1 ¼ ¼ m11 u€1 (11.3.28a)
dt ∂u_ 1 ∂u1
d ∂T ∂T
fI 2 ¼ ¼ m22 u€2 (11.3.28b)
dt ∂u_ 2 ∂u2
d ∂T ∂T
fI 3 ¼ ¼0 (11.3.28c)
dt ∂u_ 3 ∂u3
d ∂T ∂T
fI 4 ¼ ¼ m44 u€4 (11.3.28d)
dt ∂u_ 4 ∂u4
d ∂T ∂T
fI 5 ¼ ¼ m55 u€5 (11.3.28e)
dt ∂u_ 5 ∂u5
d ∂T ∂T
fI 6 ¼ ¼0 (11.3.28f)
dt ∂u_ 6 ∂u6
where
m11 ¼ m22 ¼ m1 , m44 ¼ m55 ¼ m2 (11.3.29)
Eqs. (11.3.28a)–(11.3.28f) are written in matrix form as
8 9 2 38 9
> fI 1 >
>
> >
>
> 6
m11 0 0 0 0 0 > > u€1 >
>
> fI 2 >
>
> > 7>> u€2 >
>
>
<f >
> = 6
0 m 22 0 0 0 0 7>>
< >
=
6 0 0 0 0 0 0 7 u€
¼6 7
I3 3
60 (11.3.30)
>
> fI 4 >
> 6 0 0 m44 0 077>> u€4 >
>
> >
> 5> >
>
>
> >
fI 5 >
4 0 0 0 0 m 0 >
>
> u€5 >
>
>
>
: ; > 55
: ;
0 0 0 0 0 0 u€6
fI 6
Eqs. (11.3.26a), (11.3.26b)
For an element with constant mass, m ðx Þ ¼ m,
are integrated analytically to give
e Le =2
m11 ¼ m22 ¼ m33 ¼ m44 ¼ m (11.3.31)
and the mass matrix (11.3.30) becomes
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
e 60 0 07
m 6 0 0 0 7
me ¼ 6 7 (11.3.32)
2 60 0 0 1 0 07
6 7
40 0 0 0 1 05
0 0 0 0 0 0
e Le is the total mass of the e element.
where m e ¼ m
We observe that the lumped mass assumption results in a diagonal mass
matrix.
@CivilMethod
The finite element method Chapter 11 401
(iii) Nodal damping forces and damping matrix of the beam element
As in the case of the truss element, we assume only internal damping. It is there-
fore due to the deformation of the beam element, it resists it, and depends on the
velocity of the strain. If we denote by sD the axial stress due to damping, we
may write
∂ex
sD ¼ cs (11.3.33)
∂t
where cs is the coefficient of the internal damping and ex the axial strain.
The virtual work of the internal damping force in the beam element is
Z
dWnc in
¼ sD dex dV (11.3.34)
V
sx N ðx Þ M ðx Þ
ex ¼ , sx ¼ + y, N ðx Þ ¼ EAðx Þu 0 , M ðx Þ ¼ EI ðx Þv 00 ðx, t Þ
E Aðx Þ I ðx Þ
@CivilMethod
402 PART II Multi-degree-of-freedom systems
Z L
00 00
cij ¼ c s I ðx Þ i ðx Þ j ðx Þdx, i, j ¼ 2, 3,5, 6 (11.3.37b)
0
Writing Eq. (1.8.11) for the beam element, we obtain the nodal damping
forces as generalized forces for i ¼ 1, 2, …,6. Thus we have
fD1 ¼ Q1 ¼ c11 u_ 1 + c14 u_ 4 (11.3.38a)
fD2 ¼ Q2 ¼ c22 u_ 2 + c23 u_ 3 + c25 u_ 5 + c26 u_ 6 (11.3.38b)
fD3 ¼ Q3 ¼ c32 u_ 2 + c33 u_ 3 + c35 u_ 5 + c36 u_ 6 (11.3.38c)
fD4 ¼ Q4 ¼ c41 u_ 1 + c44 u_ 4 (11.3.38d)
fD5 ¼ Q5 ¼ c52 u_ 2 + c53 u_ 3 + c55 u_ 5 + c56 u_ 6 (11.3.38e)
fD6 ¼ Q6 ¼ c62 u_ 2 + c63 u_ 3 + c65 u_ 5 + c66 u_ (11.3.38f)
Eqs. (11.3.38a)–(11.3.38f) are written in matrix form as
8 9 2 38 9
>
> fD1 >
> c11 0 0 c14 0 0 > > u_ 1 >
>
> >
>
> > 6
> 7>
>
>
>
>
>
>
> f >
D2 > 6 0 c c 0 c c 7
26 > 2 >
> u_ >
> >
> > 6 22 23 25
7>
> >
>
< fD3 >
> = 6 6 0 c32 c33 0 c35 c36 7 u_ 3 =
7>
< >
¼66 7 (11.3.39)
> fD4 > 7 u_ >
>
> >
> 6 c41 0 0 c44 0 0 7> > 4> >
>
> >
> 6 7>
>
> >
>
>
> >
> 6 7>
> _ >
>
>
> fD5 >
> 4 0 c52 c 53 0 c 55 c 56 5>
> u 5 >
>
>
: ; > : >
> ;
fD5 0 c62 c63 0 c65 c66 u_ 6
or
f eD ¼ ce u_ e (11.3.40)
where f eD , u_ e are the vectors of the nodal damping forces and nodal velocities,
respectively, and ce is the damping matrix of the beam element e. Hence, the
damping matrix ce is defined as
2 3
c11 0 0 c14 0 0
6 7
6 0 c22 c23 0 c25 c26 7
6 7
6 7
6 0 c 32 c 33 0 c 35 c36 7
c ¼6
e 6 7 (11.3.41)
7
6 c41 0 0 c44 0 0 7
6 7
60 c c 0 c c 7
4 52 53 55 56 5
@CivilMethod
The finite element method Chapter 11 403
2 3
cs A cs A
6 L 0 0 0 0 7
6 L 7
6 12c I 6c I 12cs I 6cs I 7
6 0 s s
3 7
6 0
L2 7
6 L3 L2 L 7
6 2cs I 7
6 6cs I 4cs I 6cs I 7
6 0 0 2 7
6 L 2 L L L 7
ce ¼ 6 7 (11.3.42)
6 cs A cs A 7
6 0 0 0 0 7
6 L L 7
6 7
6 12c I 6c I 12cs I 6cs I 7
6 0 s
3 2
s
0 2 7
6 L3 L 7
6 L L 7
4 6cs I 2cs I 6cs I 4cs I 5
0 2
0 2
L L L L
Note that, the superscript e in Ae , I e , and Le has been dropped from the
elements of the matrix in Eq. (11.3.42) for the sake of simplicity of the
expressions.
(iv) Equivalent nodal loads of the beam element
The equivalent nodal forces can be evaluated in two ways: (a) as generalized
forces in the direction of the nodal displacements ue and (b) as static equivalent
forces. In the first case, the shape functions utilized to derive elastic forces are
employed too. The so-resulting vector of the equivalent nodal forces is referred
to as the consistent load vector.
Fig. 11.3.4 Loading and consistent equivalent nodal loads of a plane beam element.
@CivilMethod
404 PART II Multi-degree-of-freedom systems
pe ðt Þ ¼ f p1 ðt Þ p2 ðt Þ p3 ðt Þ p4 ðt Þ p5 ðt Þ p6 ðt Þ gT (11.3.46)
Consequently
pe ðt Þ ¼ f p1 ðt Þ p2 ðt Þ 0 p4 ðt Þ p5 ðt Þ 0 gT (11.3.49)
Inserting the expressions of the linear shape function in Eqs. (11.3.48a),
(11.3.48b) yields
Z L Z L
p1 ðt Þ ¼ px ðx, t Þð1 xÞdx, p4 ðt Þ ¼ px ðx, t Þxdx (11.3.50)
0 0
Z L Z L
p 2 ðt Þ ¼ py ðx, t Þð1 xÞdx, p5 ðt Þ ¼ py ðx, t Þxdx (11.3.51)
0 0
namely, the equivalent nodal loads are obtained as the reactions of a simply sup-
ported beam under the loads px ðx, t Þ and py ðx, t Þ. Hence, their name statically
equivalent nodal loads.
@CivilMethod
The finite element method Chapter 11 405
The forces resulting in this way are less accurate because the transverse
displacement is approximated by a linear polynomial instead of a cubic one.
The linear shape functions are not geometrically admissible, which means that
the conditions of the Ritz method are violated. Nevertheless, the accuracy of the
results is acceptable, especially when the elements are small. The statically
equivalent nodal load vector is particularly useful when inertial forces due
to the rotations of the cross-sections are ignored, as in the lumped mass
assumption.
(v) Geometric stiffness matrix of the beam element
When the produced rotation of the cross-sections is significant, there may be
shortening of the length of the element even if the axial deformation is ignored
(∂u=∂x ¼ 0). This shortening is due to the nonlinear term of the strain-
displacement relation [4,5], that is,
1 ∂v ðx, t Þ 2
ex ¼ (11.3.52)
2 ∂x
In the presence of an axial force N ðx Þ, the elastic energy due to its
shortening is
Z
1 L
UG ¼ N ðx Þex dx
2 0
Z (11.3.53)
1 L ∂v ðx, t Þ 2
N ðx Þ dx
2 0 ∂x
which by virtue of Eq. (11.3.3) becomes
Z
1 L 2
UG ¼ N ðx Þ u2 02 ðx Þ + u3 03 ðx Þ + u5 0 0
5 ðx Þ + u6 6 ðx Þ dx
2 0
Differentiation of the previous expression for UG with respect to ui ,
i ¼ 1, 2, …,6, results in the additional elastic nodal forces
∂UG
fG1 ¼ ¼0 (11.3.54a)
∂u1
∂UG
fG2 ¼ ¼ kG22 u2 + kG23 u3 + kG25 u5 + kG26 u6 (11.3.54b)
∂u2
∂UG
fG3 ¼ ¼ kG32 u2 + kG33 u3 + kG35 u5 + kG36 u6 (11.3.54c)
∂u3
∂UG
fG4 ¼ ¼0 (11.3.54d)
∂u4
@CivilMethod
406 PART II Multi-degree-of-freedom systems
∂UG
fG5 ¼ ¼ kG52 u2 + kG53 u3 + kG55 u5 + kG56 u6 (11.3.54e)
∂u5
∂UG
fG6 ¼ ¼ kG62 u2 + kG63 u3 + kG65 u5 + kG66 u6 (11.3.54f)
∂u6
where
Z L
kGij ¼ N ðx Þ 0i ðx Þ 0j ðx Þdx, i, j ¼ 2, 3, 5, 6 (11.3.55)
0
element is the sum ke + keG . Thus, we conclude from Eq. (11.3.59) that the stiff-
ness increases if the axial force is tensile while it decreases if it is compressive.
The vanishing of the stiffness leads to the buckling of the structure.
The axial force is given by the relationd
∂u
N ðx Þ ¼ EA
∂x (11.3.60)
EA
¼ ðu1 + u4 Þ
L
namely, it depends on the displacements. Therefore, if the geometric stiffness is
taken into account in the analysis, the problem becomes nonlinear.
u ¼ f u1 u2 u3 u4 u5 u6 gT (11.3.64b)
Substituting now Eqs. (11.3.62), (11.3.63) into Eq. (11.3.61) yields
Z L
EAðx ÞN0 a N0a dx u
T
dWina ¼ duT (11.3.65)
0
" #
∂u 1 ∂v 2
d. The exact expression for the axial force is N ðx Þ ¼ EA + [4].
∂x 2 ∂x
@CivilMethod
408 PART II Multi-degree-of-freedom systems
where
Z L
kij ¼ EAðx Þ 0i ðx Þ 0j ðx Þdx, i, j ¼ 1, 4 (11.3.71a)
0
Z L
00 00
kij ¼ EI ðx Þ i ðx Þ j ðx Þdx, i, j ¼ 2, 3, 5, 6 (11.3.71b)
0
@CivilMethod
The finite element method Chapter 11 409
Example 11.3.1 Compute the stiffness and mass matrices and the nodal load
vector
of a beam element, which is subjected to the loads px ðx, t Þ ¼ pðt Þ
1 + x x 2 and py ðx, t Þ ¼ 3pðt Þð1 x Þ, x ¼ x=L. The height of the cross-
section varies linearly, h ðx Þ ¼ h ð1 + 0:5x Þ, while its width b is constant.
The material density is r. Consider the lumped mass assumption.
Solution
i. Stiffness matrix: We set A ¼ bh, I ¼ bh 3 =12. Hence Aðx Þ=A ¼ ð1 + 0:5x Þ,
I ðx Þ=I ¼ ð1 + 0:5x Þ3 . Then we obtain from Eq. (11.3.13)
Z L
0 0
k11 ¼ EAðx Þ 1 ðx Þ 1 ðx Þdx ¼ 1:25k11
c
0
Z L
0 0
k14 ¼ EAðx Þ 1 ðx Þ 4 ðx Þdx ¼ 1:25k14
c
0
Z L
00 00
k22 ¼ EI ðx Þ 2 ðx Þ 2 ðx Þdx ¼ 2:09375k22
c
0
Z L
00 00
k23 ¼ EI ðx Þ 2 ðx Þ 3 ðx Þdx ¼ 1:7k23
c
0
Z L
00 00
k25 ¼ EI ðx Þ 2 ðx Þ 5 ðx Þdx ¼ 2:09375k25
c
0
Z L
00 00
k26 ¼ EI ðx Þ 2 ðx Þ 6 ðx Þdx ¼ 2:4875k26
c
0
Z L
00 00
k33 ¼ EI ðx Þ 3 ðx Þ 3 ðx Þdx ¼ 1:4875k33
c
0
Z L
00 00
k35 ¼ EI ðx Þ 3 ðx Þ 5 ðx Þdx ¼ 1:7k35
c
0
Z L
00 00
k36 ¼ EI ðx Þ 3 ðx Þ 6 ðx Þdx ¼ 2:125k36
c
0
Z L
00 00
k55 ¼ EI ðx Þ 5 ðx Þ 5 ðx Þdx ¼ 2:09375k55
c
0
Z L
00 00
k56 ¼ EI ðx Þ 5 ðx Þ 6 ðx Þdx ¼ 2:4875k56
c
0
Z L
00 00
k66 ¼ EI ðx Þ 6 ðx Þ 6 ðx Þdx ¼ 2:66875k66
c
0
@CivilMethod
410 PART II Multi-degree-of-freedom systems
Z L
7
m11 ¼ m22 ¼ m ðx Þð1 xÞdx ¼ rAL
0 12
Z L
8
m44 ¼ m55 ¼ m ðx Þxdx ¼ rAL
0 12
Z L
7
p1 ðt Þ ¼ px ðx, t Þð1 x Þdx ¼ pðt ÞL
0 12
Z L
7
p4 ðt Þ ¼ px ðx, t Þxdx ¼ pðt ÞL
0 12
Z L
p2 ðt Þ ¼ py ðx, t Þð1 x Þdx ¼ pðt ÞL
0
Z L
1
p5 ðt Þ ¼ py ðx, t Þxdx ¼ pðt ÞL
0 2
@CivilMethod
The finite element method Chapter 11 411
Fig. 11.3.5 Nodal displacements of the beam element in global and local axes.
Similarly, we obtain
( ) 2 3( )
u4 cosf sin f 0 u4
u5 ¼ 4 sin f cos f 0 5 u5 (11.3.73b)
u6 0 0 1 u6
Eqs. (11.3.73a), (11.3.73b) are further combined as
8 9 2 38 9
>
> u1 >>
> 6
cos f sin f 0 0 0 0 >> u1 >>
> >
>
> > > u2 >
7> >
>
< >
> u 2
= 6
sin f cos f 0 0 0 0 7>
>
< >
=
u3 6 0 0 1 0 0 0 7 u
¼6
6
7 3
(11.3.74)
> u4 >
> > 6 0 0 0 cos f sin f 0 7> u4 >
7> >
> >
> 5> >
>
> u >
>
4 0 0 0 sin f cos f 0 >
> u > >
> ;
: 5
: 5>
> ;
u6 0 0 0 0 0 1 u6
or
ue ¼ Re ue (11.3.75a)
hence
relation to ue . Their transformation obeys the same law, Eq. (11.3.75b). Thus,
we have
@CivilMethod
412 PART II Multi-degree-of-freedom systems
f e ¼ðRe ÞT f e (11.3.77)
S S
f e ¼ðRe ÞT f e (11.3.78)
I I
f e ¼ðRe ÞT f eD (11.3.79)
D
pe ðt Þ ¼ðR Þ p ðt Þ
e T e
(11.3.80)
Using Eqs. (11.3.15), (11.3.23), (11.3.40), (11.3.75a), the first three of the
foregoing equations are transformed in the global axes as
f e ¼ ke ue (11.3.81)
S
f e e €e
u
¼m (11.3.82)
I
f D ¼ c u
e _e
(11.3.83)
where
k ¼ ðRe ÞT ke Re
e
(11.3.84)
e ¼ ð Re Þ T m e Re
m (11.3.85)
u€f + C
e ff e
M u_ f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e (11.3.88)
M u€f + M
e sf e u€s + C
e ss e u_ f + C
e sf e u_ s + K
e ss e e sf e e ss e
uf + K us ¼ e
p s ðt Þ (11.3.89)
where
pf∗ ðt Þ ¼ e
e pf ðt Þ M u€s C
e fs e u_ s K
e fs e e fs e
us (11.3.90)
Apparently, Eqs. (11.3.88), (11.3.89), (11.3.90) are the counterpart of
Eqs. (11.2.117), (11.2.118), (11.2.119), which hold for the plane truss.
For the sake of convenience, we write Eq. (11.3.88) as
M€
u + C€
u + Ku ¼ pðt Þ (11.3.91)
@CivilMethod
The finite element method Chapter 11 413
Assumed data:
Coordinates of frame nodes: 1ð0, 0Þ, 2ð0, 5Þ, 3ð4, 8Þ, 4ð10, 0Þ.
Properties of the elements: A1 ¼ A, A2 ¼ A3 ¼ 2A, I1 ¼ I , I2 ¼ I3 ¼ 8I ,
I =A ¼ ðL1 =50Þ2 . Load pðx, t Þ ¼ P ðt Þ=L2 . Damping coefficient cs ¼ 0,
modulus of elasticity E, material density r. Lumped mass assumption.
Solution
The system has n ¼ 4 nodes, hence the free structure has N ¼ 3n ¼ 12 degrees
of freedom, (u3i2 , u3i1 , u3i , i ¼ 1, 2, 3, 4). The numbering of the elements and
the positive direction of the local axes are chosen as in Fig. E11.2.
1 0 0 0 0 5 5 5 1 0
2 0 4 4 5 8 3 5 0.8 0.6
3 4 10 6 8 0 8 10 0.6 0.8
1. Computation of ke , me , ce , pe ðt Þ, Re for e ¼ 1, 2, 3
Matrices ke . The elements have constant cross-section. Therefore, Eq. (11.3.17)
is employed, which gives
@CivilMethod
414 PART II Multi-degree-of-freedom systems
2 3
2500 0 0 2500 0 0
6 0 12 30 0 12 30 7
6 7
EI 66 0 30 50 7
0 30 100 7
k ¼
1
6 7
125 6 2500 0 0 2500 0 07
6 7
4 0 12 30 0 12 30 5
0 30 50 0 30 100
2 3
5000 0 0 5000 0 0
6 0 96 240 0 96 240 7
6 7
EI 66 0 240 400 7
0 240 800 7
k ¼
2
6 7
125 6 5000 0 0 5000 0 07
6 7
4 0 96 240 0 96 240 5
0 240 400 0 240 800
2 3
2500 0 0 2500 0 0
6 0 12 60 0 12 60 7
6 7
EI 6 0 60 400 0 60 200 7
k ¼
3 6 7
125 6
6 2500 0 0 2500 0 0 7
7
4 0 12 60 0 12 60 5
0 60 200 0 60 400
Matrices me . The elements have constant cross-section. Therefore,
Eq. (11.3.32) is employed for lumped mass assumption, which gives
2 3 2 3
2:5 0 0 0 0 0 5:0 0 0 0 0 0
6 0 2:5 0 0 0 0 7 6 0 5:0 0 0 0 0 7
6 7 6 7
6 0 0 0 0 0 07 7 60 0 0 0 0 07
m1 ¼ rA6 6 0 0 0 2:5 0 0 7 , m 2
¼ rA 6
6 0 0 0 5:0 0 0 7
7
6 7 6 7
4 0 0 0 0 2:5 0 5 4 0 0 0 0 5:0 0 5
0 0 0 0 0 0 0 0 0 0 0 0
2 3
10:0 0 0 0 0 0
6 0 10:0 0 0 0 07
6 7
6 0 0 0 0 0 07
m ¼ rA6
3 6 7
6 0 0 0 10:0 0 0 7 7
4 0 0 0 0 10:0 0 5
0 0 0 0 0 0
Matrices ce . It is cs ¼ 0 and Eq. (11.3.42) gives ce ¼ 0.
Vectors pe ðt Þ. To be consistent to the lumped mass assumption, the statically
equivalent nodal load vector will be employed. Therefore, Eqs. (11.3.50), (11.3.51)
give for the load-free elements 1 and 3, p1 ðt Þ ¼ p2 ðt Þ ¼ p3 ðt Þ ¼ p4 ðt Þ ¼ 0 while
for element 2, we obtain
p1 ðx, t Þ ¼ PLð2t Þ cos fð2Þ sin fð2Þ , p2 ðx, t Þ ¼ PLð2t Þ cos 2 fð2Þ
@CivilMethod
The finite element method Chapter 11 415
2. Computation of k , m
e
e , pe ðt Þ for e ¼ 1, 2, 3
Matrices k . They are computed using Eq. (11.3.84)
e
2 3
12 0 30 12 0 30
6 0 2500 0 0 2500 07
6 7
EI 6 30 0 100 30 0 50 7
k ¼
1 6 7
125 6
6 12 0 30 12 0 30 7
7
4 0 2500 0 0 2500 05
30 0 50 30 0 100
2 3
3234:6 2353:9 144:0 3234:6 2353:9 144:0
6 2353:9 1861:4 192:0 2353:9 1861:4 192:0 7
6 7
EI 6 144:0 192:0 800:0 144:0 192:0 400:0 7
k ¼
2 6 7
125 66 3234:6 2353:9 144:00 3234:6 2353:9 144:0 7 7
4 2353:9 1861:4 192:0 2353:9 1861:4 192:0 5
144:0 192:0 400 144:0 192:0 800:0
@CivilMethod
416 PART II Multi-degree-of-freedom systems
2 3
907:7 1194:2 48:0 907:6 1194:2 48:0
6 1194:2 1604:3 36:0 1194:2 1604:3 36:0 7
6 7
EI 6 48:0 36:0 400:0 48:0 36:0 200:0 7
k3 ¼ 6 7
125 6
6 907:6 1194:2 48:0 907:7 1194:2 48:0 7
7
4 1194:2 1604:3 36:0 1194:2 1604:3 36:0 5
48:0 36:0 200:0 48:0 36:0 400:0
e . They are computed using Eq. (11.3.85)
Matrices m
2 3 2 3
2:5 0 0 0 0 0 5:0 0 0 0 0 0
60 2:5 0 0 0 0 7 60 5:0 0 0 0 07
6 7 6 7
60 0 0 0 0 07 60 0 0 0 0 07
1 ¼ rA6
m 60
7, m
7 2
¼ rA 6
60
7
6 0 0 2:5 0 0 7 6 0 0 5:0 0 077
40 0 0 0 2:5 0 5 40 0 0 0 5:0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3
10:0 0 0 0 0 0
6 0 10:0 0 0 0 07
6 7
6 0 0 0 0 0 07
3 ¼ rA6
m 6 0
7
6 0 0 10:0 0 077
4 0 0 0 0 10:0 05
0 0 0 0 0 0
Vectors pe ðt Þ. They are computed using Eq. (11.3.80).
8 9 8 9
>
> 0 >
> >
> 0>>
> >
> > >
>
> 0:08P ð t Þ >
>
>
> 0>>
>
< = < >
> =
0 0
p ðt Þ ¼
2
, p ðt Þ ¼ p ðt Þ ¼
1 3
>
> 0 >
> >
> 0>>
> >
> > >
>
> 0:08P ð t Þ >
>
>
> 0>>
>
: ; : >
> ;
0 0
^ e, M
3. Computation of the enlarged matrices K ^ e, p
^e ðt Þ for e ¼ 1, 2, 3
Assembly matrices ae . By inspection, we obtain
2 3 2 3
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0
60 1 0 0 0 0 0 0 0 0 0 07 60 0 0 0 1 0 0 0 0 0 0 07
6 7 6 7
6 7 6 7
60 0 1 0 0 0 0 0 0 0 0 07 60 0 0 0 0 1 0 0 0 0 0 07
6 7 6 7
a ¼6
1
7, a ¼ 6
2
7
60 0 0 1 0 0 0 0 0 0 0 07 60 0 0 0 0 0 1 0 0 0 0 07
6 7 6 7
6 7 6 7
40 0 0 0 1 0 0 0 0 0 0 05 40 0 0 0 0 0 0 1 0 0 0 05
0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0
@CivilMethod
The finite element method Chapter 11 417
2 3
0 0 0 0 0 0 1 0 0 0 0 0
60 0 0 0 0 0 0 1 0 0 0 07
6 7
6 7
60 0 0 0 0 0 0 0 1 0 0 07
6 7
a ¼6
3
7
60 0 0 0 0 0 0 0 0 1 0 07
6 7
6 7
40 0 0 0 0 0 0 0 0 0 1 05
0 0 0 0 0 0 0 0 0 0 0 1
^ e ¼ ðae ÞT ke ae
^ e . They are computed using the relation K
Matrices K
2 3
12 0 30 12 0 30 0 0 0 0 0 0
6 0 2500 0 0 2500 0 0 0 0 0 0 07
6 7
6 7
6 30 0 100 30 0 50 0 0 0 0 0 0 7
6 7
6 12 30 0 0 0 0 0 0 7
6 0 30 12 0 7
6 7
6 0 2500 0 0 2500 0 0 0 0 0 0 07
6 7
6 30 100 0 0 0 0 0 0 7
^1 ¼ EI 6 0 50 30 0 7
K 6 7
125 6 0 0 0 0 0 0 0 0 0 0 0 07
6 7
6 0 0 0 0 0 0 0 07
6 0 0 0 0 7
6 7
6 0 0 0 0 0 0 0 0 0 0 0 07
6 7
6 0 0 0 0 0 0 0 07
6 0 0 0 0 7
6 7
4 0 0 0 0 0 0 0 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3
0 0 0 0 0 0 0 0 0 0 0 0
6 7
60 0 0 07
6 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
6 7
60 0 0 3234:6 2353:9 144:0 3234:6 2353:9 144:0 0 0 07
6 7
6 7
60 0 0 2353:9 1861:4 192:0 2353:9 1861:4 192:0 0 0 07
6 7
6 7
60 0 0 144:0 192:0 800:0 144:0 192:0 400:0 0 0 07
^ 2 ¼ EI 6
K 6
7
7
125 6 0 0 0 3234:6 2353:9 144:0 3234:6 2353:9 144:0 0 0 07
6 7
6 7
60 0 0 2353:9 1861:4 192:0 2353:9 1861:4 192:0 0 0 07
6 7
6 7
60 0 0 144:0 192:0 400:0 144:0 192:0 800:0 0 0 07
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
4 5
0 0 0 0 0 0 0 0 0 0 0 0
@CivilMethod
418 PART II Multi-degree-of-freedom systems
2 3
0 0 0 0 0 0 0 0 0 0 0 0
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 07
6 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
6 7
EI 60 0 0 0 0 0 0 0 0 0 0 07
^3 ¼
K 6 7
125 6
60 0 0 0 0 0 907:6 1194:2 48 907:6 1194:2
7
48 7
6 7
60 0 0 0 0 0 1194:2 1194:2 1604:3 36 7
6 1604:3 36 7
6 7
60 0 0 0 0 0 48 36 400 48 36 200 7
6 7
6 7
60 0 0 0 0 0 907:6 1194:2 48 907:6 1194:2 48 7
6 7
6 7
40 0 0 0 0 0 1194:2 1604:3 36 1194:2 1604:3 36 5
0 0 0 0 0 0 48 36 200 48 36 400
@CivilMethod
The finite element method Chapter 11 419
2 3
0 0 0 0 0 0 0 0 0 0 0 0
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 07
6 0 0 0 0 0 0 0 0 0 0 7
60 07
6 0 0 0 0 0 0 0 0 0 0 7
6 7
^ 60 0 0 0 0 0 0 0 0 0 0 07
M ¼ rA6
3
7
60 0 0 0 0 0 10: 0 0 0 0 07
6 7
60 0 0 0 0 0 0 10: 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 10: 0 07
6 7
40 0 0 0 0 0 0 0 0 0 10: 05
0 0 0 0 0 0 0 0 0 0 0 0
^e ðt Þ. They are computed using the relation p
Vectors p ^ e ðt Þ ¼ ðae ÞT pe ðt Þ
8 9 8 9
>
> 0 >
> >
> 0> >
>
> 0 >
> >
> 0> >
>
> >
> >
> >
>
>
> 0 >
> >
> >
>
>
> >
> >
> 0 >
>
>
> >
> >
> >
>
>
> 0 >
> >
> 0 >
>
>
> >
> >
> >
>
> 0:08P ð t Þ >
> > 0 >
>
< = < >
> =
0 0
^ ðt Þ ¼
p 2
^ ðt Þ ¼ p
, p1
^ ðt Þ ¼
3
>
> 0 >
> >
> 0> >
> >
> > >
>
>
> 0:08P ð t Þ >
>
>
>
> 0> >
>
>
> >
> >
> >
>
>
> 0 >
> >
> 0 >
>
> >
> > >
>
> 0
> >
> >0>
>
> >
>
>
> >
> >
> >
>
> >
> > >
>
: 0 ; : >
> 0 ;
0 0
K,
4. Computation of the total matrices M, pðt Þ of the frame
2 3
2:5 0 0 0 0 0 0 0 0 0 0 0
60 2:5 0 0 0 0 0 0 0 0 0 07
6 7
60 07
6 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 7:5 0 0 0 0 0 0 0 07
6 7
60 0 0 0 7:5 0 0 0 0 0 0 07
X 6 7
3 6 07
¼
M ^ e ¼ rA6 0
M
0 0 0 0 0 0 0 0 0 0 7
60 0 0 0 0 0 15: 0 0 0 0 07
e¼1 6 7
60 07
6 0 0 0 0 0 0 15: 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 10: 0 07
6 7
40 0 0 0 0 0 0 0 0 0 10: 05
0 0 0 0 0 0 0 0 0 0 0 0
@CivilMethod
X
3
¼
K ^e
K
e¼1
2 3
12 0 30 12 0 30 0 0 0 0 0 0
6 0 2500 0 0 2500 0 0 0 0 0 0 07
6 7
6 7
6 30 0 100 30 0 50 0 0 0 0 0 07
6 7
6 12 2353:9 114 3234:6 2353:9 144 07
6 0 30 3246:6 0 0 7
6 7
6 0 2500 0 2353:9 4361:4 192 2353:9 1861:4 192 0 0 07
6 7
EI 66 30 0 50 114 192 900 144 192 400 0 0 077
¼ 6 7
125 6 0 0 0 3234:6 2353:9 144 4142:28 1159:7 192 907:68 1194:2 48 7
6 7
6 0 0 2353:9 1861:4 192 156 1194:2 1604:3 36 7
6 0 1159:7 3465:7 7
6 7
6 0 0 0 144 192 400 192 156 1200 48 36 200 7
6 7
6 0 907:68 1194:2 48 907:68 1194:2 48 7
6 0 0 0 0 0 7
6 7
4 0 0 0 0 0 0 1194:2 1604:3 36 1194:2 1604:3 36 5
0 0 0 0 0 0 48 36 200 48 36 400
8 9
>
>
0 >
>
>
> >
>
>
> 0 >
>
>
> >
>
>
> 0 >
>
>
> ð Þ >
>
>
> 2P t >
>
>
> >
>
>
> 0:08P ðt Þ >
>
X3 < =
ðt Þ + 0
pðt Þ ¼ P ^ e ðt Þ ¼
p
>
> 0 >
>
> >
> 1:08P ðt Þ >
e¼1
>
> >
>
>
> >
>
> 0
> >
>
>
> >
>
> 0
> >
>
>
> >
>
>
> 0 >
>
: ;
0
The finite element method Chapter 11 421
K,
5. Modification of the matrices M, pðt Þ due to support conditions of
the frame
Referring to Fig. E11.2, the displacement vector should be rearranged as
e e1 u
uT ¼ f u e2 ue3 ue4 ue5 ue6 ue7 ue8 ue9 ue10 ue11 ue12 g
¼ f u4 u5 u6 u7 u8 u9 u12 u1 u2 u3 u10 u11 g
Hence, the matrix V is defined from the relation
8 9 2 38 9
> e1 >
u 0 0 0 0 0 0 0 1 0 0 0 0 > u4 >
> >
> > >
>u
>
> e >
> 60 07 > u5 >
> >
>
> 2 > 6 0 0 0 0 0 0 0 1 0 0 7>
> >
>
>
>
> e3 >
u >
>
>
60
6 0 0 0 0 0 0 0 0 1 0 07 >
>
>
7> u 6
>
>
>
>
> > > >
>
>
> e4 >
u > 6
>
61 0 0 0 0 0 0 0 0 0 0 07 >
7>
> u 7 >
>
>
> >
> > >
>
>
> e
u5 > >
60 1 0 0 0 0 0 0 0 0 0 07 > u8 >
>
7> >
>
< = 660 < =
e6
u 0 1 0 0 0 0 0 0 0 0 077 u9
¼6
60
>
> e
u 7 >> 6 0 0 1 0 0 0 0 0 0 0 07 > u12 >
7> >
>
>u >
> 60 > u1 >
> >
>
> e 8 >> 6 0 0 0 1 0 0 0 0 0 0 077>
> >
>
>
>u >
> 60 >
> >
>
>
> e 9 >> 6 0 0 0 0 1 0 0 0 0 0 07 >
7> u 2 >>
>
> >
> 60 > u >
> >
>
> e
u >
10 > 6 0 0 0 0 0 0 0 0 0 1 07 >
7> 3 >>
>
> >
> > >
>
>
>
: e
u 11 >
>
;
40 0 0 0 0 0 0 0 0 0 0 1 5>
>
>
: u 10 >
>
;
e12
u 0 0 0 0 0 0 1 0 0 0 0 0 u11
Applying Eqs. (11.2.114a), (11.2.114c), (11.2.114d), we obtain
@CivilMethod
422 PART II Multi-degree-of-freedom systems
e ff , M
6. Formulation of the matrices M e fs , M
e sf , M
e ss , K
e ff , K
e fs , K
e sf , K
e ss ,
e
pf ðt Þ, e
p s ðt Þ
e M,
These matrices result directly from K, e e pðt Þ taking into account the partition-
ing dictated by the separation of the free displacements from the specified
(support) ones. Thus, we have
2 3 2 3
7:5 0 0 0 0 0 0 0 0 0 0 0
60 7:5 0 0 0 0 07 60 0 0 0 07
6 7 6 7
60 0 0 0 0 0 07 7 60 0 0 0 07
6 6 7
e ff ¼ rA6 0 0 0 15: 0 0 0 7 e 6 7
M 6 7, Mfs ¼ rA6 0 0 0 0 0 7
60 0 0 0 15: 0 0 7 7 6 7
6 60 0 0 0 07
40 0 0 0 0 0 0 5 4 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3 2 3
0 0 0 0 0 0 0 2:5 0 0 0 0
60 0 0 0 0 0 07 60 2:5 0 0 0 7
6 7 6 7
Me sf ¼ rA6 0 0 0 0 0 0 0 7, M e ss ¼ rA6 0 0 0 0 0 7
6 7 6 7
40 0 0 0 0 0 05 40 0 0 10: 0 5
0 0 0 0 0 0 0 0 0 0 0 10:
2 3
3246:6 2353:9 114 3234:6 2353:9 144 0
6 2353:9 4361:4 192 2353:9 1861:4 192 07
6 7
6 114 192 900 144 192 400 0 7
EI 6 7
e
Kff ¼ 6 3234:6 2353:9 144 4142:2 1159:7 192 48 7
6
125 6 7
7
6 2353:9 1861:4 192 1159:7 3465:7 156 36 7
4 144 192 400 192 156 1200 200 5
0 0 0 48 36 200 400
2 3
12 0 30 0 0
6 0 2500 0 0 0 7
6 7
6 30 0 50 0 0 7
EI 6 7
Ke fs ¼ 6 0 0 0 907:6 1194:2 7
125 6
6 0
7
6 0 0 1194:2 1604:3 7 7
4 0 0 0 48 36 5
0 0 0 48 36
@CivilMethod
The finite element method Chapter 11 423
2 3
12 0 30 0 0 0 0
6 0 2500 0 0 0 0 07
6 7
e sf ¼ EI 6 30
K 0 50 0 0 0 07
125 46 7
0 0 0 907:6 1194:2 48 48 5
0 0 0 1194:2 1604:3 36 36
2 3
12 0 30 0 0
6 0 2500 0 0 0 7
e EI 66
7
Kss ¼ 6 30 0 100 0 0 77
125 4
0 0 0 907:6 1194:2 5
0 0 0 1194:2 1604:3
8 9
>
> 2:0P > > 8 9
>
>
>
> 0:08P > >
>
> >
> P1 >>
> >
> > >
>
< 0 = < P2 >
> =
e
pf ðt Þ ¼ 0 , e
ps ðt Þ ¼ P3
> > > >
> 1:08P >
>
> >
> > P10 >
>
> >
>
>
> >
> : ;
>
> 0 >
> P11
: ;
0
7. Equation of motion
In as much as e u€s ¼ 0 and Eq. (11.3.90) yields e
us ¼ 0, it is also e pf∗ ðt Þ ¼ e
pf ðt Þ.
The equation of motion results from Eq. (11.3.88). Thus we obtain
2 38 9
7:5 0 0 0 0 0 0 > > e€1 >
u >
60 > >
6 7:5 0 0 0 0 077>
> u >
> e€2 >
> >
60 7 > >
>
6 0 0 0 0 0 0 7><ue€3 >
=
6 7 €
rA6 0 0 0 15:0 0 0 07 u e4
6 7> € >
60 0 0 0 15:0 0 0 7>>
>ue5 >>
>
6 7>> >
40 0 0 0 0 0 0 5>> e
u€6 >
>
>
>
:€ > ;
0 0 0 0 0 0 0 e7
u
2 38 9
3246:6 2353:9 114:0 3234:6 2353:9 144:0 0 >
>ue1 >>
6 2353:9 4361:4 192:0 2353:9 1861:4 192:0 0 7>>
> e2 >
u
>
>
6 7>> >
>
6 114:0 144:0 192:0 400:0 0 7>7 > e3 >>
EI 66 192:0 900:0 < u =
7
+ 6 3234:6 2353:9 144:0 4142:2 1159:7 192:0 48:0 7 u e4
125 6 7> >
6 2353:9 1861:4 192:0 1159:7 3465:7 156:0 36:0 7> >ue5 >>
6 7>>
>
>
>
>
4 144:0 5
192:0 156:0 1200:0 200:0 > > e >
192:0 400:0 > >
:
u 6 >
;
0 0 0 48:0 36:0 200:0 400:0 e7
u
8 9
>
> 2:0P >>
>
> 0:08P > >
>
> >
>
>
> >
>
>
< 0 >
=
¼ 0
>
> >
>
>
> 1:08P >>
>
>
>
> >
>
>
> 0 >
>
: ;
0
@CivilMethod
424 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 425
where
Ktt∗ ¼ Ktt Ktq K1
qq Kqt (11.4.4a)
pt∗ ¼ pt Ktq K1
qq pq (11.4.4b)
The matrix Ktt∗ is called the statically condensed stiffness matrix while the
procedure producing it is referred to as the static condensation. Obviously,
when pq ¼ 0, pt∗ ¼ pt .
The reduction of the degrees of freedom can also be achieved when the
mass matrix is not diagonal or the inertial forces in the rotational degrees
of freedom are not negligible. In the employed technique, a subset of the coor-
dinates is selected arbitrarily as the set of active (or “master”) coordinates, say
ut , and the remaining coordinates are dependent (or “slave”) coordinates, say
uq . In this case, the mass matrix of Eq. (11.4.1) is written in the partitioned
form
Mtt Mtq
M¼ (11.4.5)
Mqt Mqq
where now the matrices Mtq , Mqt , Mqq may not vanish.
The starting point is Eq. (11.4.2b), which for pq ðt Þ ¼ 0 becomes
uq ¼ K1
qq Kqt ut (11.4.6)
In the foregoing equation ut represents the independent coordinates and
uq the dependent ones. Thus, we can write
ut I
¼ ut (11.4.7)
uq K1qq Kqt
or
u ¼ Tut (11.4.8)
where
I
T¼ (11.4.9)
K1
qq Kqt
@CivilMethod
426 PART II Multi-degree-of-freedom systems
where
K ¼TT KT
(11.4.11)
¼Ktt Ktq K1
qq Kqt
¼duTt p ∗ ðt Þ
where
p ∗ ðt Þ ¼TT pðt Þ
(11.4.16)
¼pt ðt Þ K1
qq Kqt pq ðt Þ
@CivilMethod
The finite element method Chapter 11 427
@CivilMethod
428 PART II Multi-degree-of-freedom systems
The implied partitioning of the matrices M, K, and vector pðt Þ in the foregoing
equations yields
2 3
7:5 0 0 0
60 7:5 0 0 7
Mtt ¼ rA6
40
7
0 15 0 5
0 0 0 15
2 3
3246:6 2353:9 3234:6 2353:9
EI 66 2353:9 4361:4 2353:9 1861:4 7,
7
Ktt ¼ 4
125 3234:6 2353:9 4142:2 1159:7 5
2353:9 1861:4 1159:7 3465:7
2 3
114 144 0
EI 6
6 192 192 0 7
7
Ktq ¼ 4 5
125 144 192 48
192 156 36
2 3 2 3
114 192 144 192 900 400 0
EI 4 EI
Kqt ¼ 144 192 192 156 5, Kqq ¼ 4 400 1200 200 5
125 125
0 0 48 36 0 200 400
2 3
2P ðt Þ 2 3
6 0:08P ðt Þ 7 0
pt ðt Þ ¼ 6
4 0
7, pq ðt Þ ¼ 4 0 5
5
0
1:08P ðt Þ
Applying Eqs. (11.4.4a), (11.4.4b) gives (11.4.4a)
2 3
3222:71 2389:02 3205:83 2387:19
EI 66 2389:02 4308:10 2396:65 1810:18 7,
7
Ktt∗ ¼ 4 5
125 3205:83 2396:65 4101:75 1196:10
2387:19 1810:18 1196:10 3412:98
8 9
> 2:0P ðt Þ >
>
> >
>
∗
< 0:08P ðt Þ =
pt ðt Þ ¼
>
> 0 >
>
>
: >
;
1:08P ðt Þ
@CivilMethod
The finite element method Chapter 11 429
u ¼ f u1 u2 u3 u4 gT (11.5.1)
where u1 , u3 are the translational degrees of freedom and u2 , u4 the rotational
ones. Then the reduced stiffness, mass, and damping matrices and the vectors
of the nodal forces of the element will result from the corresponding original
ones by omitting the nodal forces in these directions. Thus, for an element
with a constant cross-section, the stiffness matrix results from Eq. (11.3.17)
by omitting the first and fourth line as well as the first and fourth column. This
yields
e. The omission of the axial deformation is not always permitted because it may lead to considerable
mistakes in certain structures such as high-rise buildings, whose overall deformation pattern resem-
bles that of a cantilever when subjected to horizontal loads. Obviously, this deformation cannot be
realized when the axial deformation of the columns is neglected.
@CivilMethod
430 PART II Multi-degree-of-freedom systems
2 3
12EI 6EI 12EI 6EI
6 L 3 L2 L3 L2 7
6 7
6 6EI 4EI 6EI 2EI 7
6 2 7
6 2 L 7
ke ¼ 6 L L L 7 (11.5.2)
6 12EI 6EI 12EI 6EI 7
6 3 2 7
6 L L L3 L2 7
4 6EI 2EI 6EI 4EI 5
2
L2 L L L
@CivilMethod
The finite element method Chapter 11 431
while the geometric stiffness matrix for constant axial force results from
Eq. (11.3.59) 2 3
12 12
1 1
6 L L 7
6 4L L 7
6 1 1 7
N66 3 3 7
7
kG ¼ 6 12
e
12 7 (11.5.6)
10 6 1 7
6 L 1 L 7
4 L 4L 5
1 1
3 3
Finally, the consistent nodal load vector is obtained from Eq. (11.3.46),
which for a constant load py ðx, t Þ ¼ pe becomes
T
p e Le L L
pe ðt Þ ¼ 1 1 (11.5.7)
2 6 6
and the static equivalent nodal load vector is obtained from Eq. (11.3.49) as
p e Le
f 1 0 1 0 gT
pe ðt Þ ¼ (11.5.8)
2
In the following examples, the formulation of the equation of motion of
certain frames with elements parallel to the global axes is presented. The method
of influence coefficients is employed to establish the matrices of the structure.
Example 11.5.1 Formulate the equation of motion of the structure shown in
Fig. E11.3a. Consider the lumped mass assumption for the column. Examine
the case I0 ¼ 0. Assumed material density r ¼ m=10AL.
Solution
The structure may be viewed as a plane frame consisting of a single element, the
beam column. Obviously, the system has two free degrees of freedom in global
axes: the nodal displacement u1 and the rotation u2 of the top cross-section
(see Fig. E11.3b). Hence, the stiffness matrix of the structure is
@CivilMethod
432 PART II Multi-degree-of-freedom systems
¼6
K 4
L3 L2 7
5 (2)
6EI 4EI
L2 L
The mass matrix is
m11 0
M¼ (3)
0 m22
where
ð1Þ
11 ¼ m11 + m ¼ rAL=2 + m ¼ 1:05m and m22 ¼ I0
m
Hence, the mass matrix of the structure is
M ¼ 1:05m 0 (4)
0 I0
The vector of the nodal loads is
pðt Þ
pðt Þ ¼ (5)
M ðt Þ
Therefore, the equation of motion of the structure is
m 11 0 u€1 k11 k12 u1 pðt Þ
+ ¼ (6)
0 22
m u€2 k 21 k 22 u2 M ðt Þ
or after performing the operations
11 u€1 + k11 u1 + k12 u2 ¼ pðt Þ
m (7a)
22 u€2 + k21 u1 + k22 u2 ¼ M ðt Þ
m (7b)
@CivilMethod
The finite element method Chapter 11 433
@CivilMethod
434 PART II Multi-degree-of-freedom systems
Solution
Inasmuch as the axial deformation of the elements is neglected, the degrees of
freedom of the structure are reduced to seven, two translational, u1 , u2 , and five
rotational, u3 , u4 , u5 , u6 , u7 . The degrees of freedom are shown in Fig. E11.5.
(a) (b)
(c) (d)
(e) (f)
(g)
Fig. E11.6 Deformation patterns of the frame in Example 11.5.2 due to unit nodal displacements.
@CivilMethod
The finite element method Chapter 11 435
2 3
k11 k12 k13 k14 k15 k16 k17
6 7
6 k 21 k22 k23 k24 k25 k26 k27 7
6 7
6 k31 k32 k33 k34 k35 k36 k37 7
6 7
¼6
K 6 k41 k42 k43 k44 k45 k46
7
k47 7 (1)
6 7
6 k51 k52 k53 k54 k55 k56 k57 7
6 7
6 7
4 k 61 k62 k63 k64 k65
k66 k67 5
k 71 k 72 k 73 k 74 k 75 k76
k77
Its elements will be computed as the elastic forces at the nodes by giving a
unit value to each displacement and keeping the other displacements equal to
zero, following the method of the influence coefficients discussed in
Section 10.4.1. For the economy of the computation, we will take advantage
of the symmetry of the matrix and restrict the computations to the elements
of the main diagonal and those below it.
The elements of the 1st column of K are computed by referring to
Fig. E11.6a. The displacement u1 ¼ 1 “activates” (deforms) the elements
1 and 3. Their matching with the beam element is shown in Fig. E11.5, namely:
Element 1: u1 ¼ u1 , u2 ¼ 0, u3 ¼ 0, u4 ¼ 0
Element 3: u1 ¼ u1 , u2 ¼ 0, u3 ¼ 0, u4 ¼ 0
Then, with reference to Eq. (11.5.2), we obtain
12EI 12EI 27 EI 12EI
k11 ¼ k11 + k11 ¼
ð1Þ ð3Þ ð3Þ
+ 3 ¼ , k 21 ¼ k31 ¼ 3
ð2LÞ3 L 2 L3 L
6EI 3 EI 6EI 3 EI
k31 ¼ k41 ¼
ð1Þ ð1Þ
¼ , k 41 ¼ k21 ¼ ¼
ð2LÞ2 2 L2 ð2LÞ2 2 L2
6EI 6EI
k51 ¼ k21 ¼ 2 , k61 ¼ k41 ¼ 2 , k71 ¼ 0
ð3Þ ð3Þ
L L
The elements of the 2nd column of K are computed by referring to
Fig. E11.6b. The displacement u2 ¼ 1 activates (deforms) the elements 3, 4,
and 6. Their matching with the beam element is:
Element 3: u1 ¼ 0, u2 ¼ 0, u3 ¼ u2 , u4 ¼ 0
Element 4: u1 ¼ u2 , u2 ¼ 0, u3 ¼ 0, u4 ¼ 0
Element 6: u1 ¼ u2 , u2 ¼ 0, u3 ¼ 0, u4 ¼ 0
Then, with reference to Eq. (11.5.2), we obtain
12EI 12EI 12EI 36EI
k22 ¼ k33 + k11 + k11 ¼ 3 + 3 + 3 ¼ 3
ð3Þ ð4Þ ð6Þ
L L L L
6EI
k32 ¼ 0, k42 ¼ 0, k52 ¼ k23 ¼ 2
ð3Þ
L
6EI 6EI 6EI
k62 ¼ k43 + k21 ¼ 2 + 2 ¼ 0, k72 ¼ k21 ¼ 2
ð3Þ ð4Þ ð6Þ
L L L
@CivilMethod
436 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 437
ð2Þ
Mass matrix. The lumped mass assumption for the elements, Fig E11.7a,
concentrates the mass at the nodes of the frame as shown in Fig. E11.7b.
Obviously, inertial forces arise only in the directions of the translational degrees
of freedom u1 and u2 . Thus, we have
ð1Þ ð2Þ ð2Þ ð3Þ
m11 ¼ m1 + m1 + m2 + m1 ¼ 11:5mL
ð3Þ ð4Þ ð5Þ ð5Þ ð6Þ
m22 ¼ m2 + m1 + m1 + m2 + m1 ¼ 16:5mL
Therefore the mass matrix of the structure reads
ð3Þ
@CivilMethod
438 PART II Multi-degree-of-freedom systems
(a) (b)
Fig. E11.7 Lumped masses of the elements (a) and nodal masses of the frame (b) in Example
11.5.2.
Vector of the external nodal loads. This includes the nodal loads that are
directly applied to the nodes of the frame as well as those resulting from
the contribution of the equivalent nodal loads due to the element loads. In
our problem, only element 6 is loaded. To be consistent with the mass
matrix, the equivalent nodal load will be obtained from Eq. (11.3.51).
Hence 8 9
>
2 >
< P ðt Þ >
> =
ð6Þ p1 ðt Þ 3
p ðt Þ ¼ ¼
p3 ðt Þ >
> >
: 1 P ðt Þ >
;
3
Therefore, the vector of the external loads of the structure reads
8 9
8 9 >
> 0 >
>
P ðt Þ = >
> >
>
mL < < 0 =
33 0
Mtt ¼ , pt ðt Þ ¼ 2 , pq ðt Þ ¼ 0
2 0 23 : P ðt Þ ; >
> >
>
>
> 0 >
>
3 : ;
2P ðt ÞL
Applying Eqs. (11.4.4a), (11.4.4b) give
" #
EI 10:209 10:344 0:546
K tt ¼
∗
,
p ∗ ðt Þ ¼ P ðt Þ
t
L3 10:344 33:131 0:837
@CivilMethod
440 PART II Multi-degree-of-freedom systems
u
½ Da Db a ¼0 (11.6.4)
ub
or
Da ua + Db ub ¼0 (11.6.5)
which gives
ub ¼ D1 a
b Da u (11.6.6)
From Eq. (11.6.6), we conclude that the selection of the independent
coordinates cannot be arbitrary, but they must be chosen so that the inversion
of the matrix Db is ensured. This is accomplished, as will be shown in
Section 11.7, by the Gauss-Jordan elimination method. Generally, after select-
ing the independent coordinates, it is necessary to rearrange the elements of
the vector u to comply with the partition suggested by Eq. (11.6.3). Then
Eq. (11.6.1) is written
M u€ + K
u€ + C u ¼
pðt Þ (11.6.7)
where M, K,
C, u, and pðt Þ are the modified matrices and vectors after the
rearrangement.
Eq. (11.6.3) can be further written
I
u ¼ ua (11.6.8)
D1
b Da
or
u ¼T
ua (11.6.9)
where I is the unit matrix with dimensions L L and
I
T¼ (11.6.10)
D1b Da
@CivilMethod
The finite element method Chapter 11 441
where
^ ¼ TT MT
M (11.6.13)
1
U ¼ uT K
u
2
1 ua
¼ uTa TT KT (11.6.16)
2
1 ^
¼ uTa K
ua
2
where
K
^ ¼ TT KT (11.6.17)
is the L L transformed reduced stiffness matrix.
as M
Partitioning K in Eq. (11.6.14) we obtain from Eq. (11.6.17)
" #( )
K aa K
ab I
^¼ I
K D1
b Da
ba K
K bb D1 b Da (11.6.18)
aa K
¼K ab D Da D Da K
1 1 T
ba + D1 Da T K
bb D1 Da
b b b b
@CivilMethod
442 PART II Multi-degree-of-freedom systems
uT Cu_
¼ d
(11.6.19)
u_ a
uTa TT CT
¼ d
^ u_ a
uTa C
¼ d
Hence, we have
C
^ ¼ TT CT (11.6.20)
Similarly, we establish the transformed vector of the external loads that is,
uT pðt Þ
dWp ¼d
uTa TT pðt Þ
¼d (11.6.21)
^ ðt Þ
uTa p
¼d
Hence
^ðt Þ ¼TT pðt Þ
p
T (11.6.22)
pa ðt Þ D1
¼ b Da pb ðt Þ
@CivilMethod
The finite element method Chapter 11 443
We consider the beam element with its six degrees of freedom with respect
to global axes, as shown in Fig. 11.7.1. If we want to study only the flexural
vibrations of the frame, we must neglect the axial deformation of the elements.
This entails the imposition of constraints on the translational degrees of free-
dom. These constraints result from the fact that the projections of end displace-
ments on the element axis must be equal [7,8], that is,
@CivilMethod
444 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 445
(a) (b)
(c) (d)
(e)
Fig. 11.7.2 Frames with axial constraints.
TABLE 11.7.1 Geometrical data of the elements of the frame in Fig. 11.7.2a.
Number of
element e Node j Node k D
x D
y Le cosfe sin fe
1 1 2 0 3 3 0 1
2 2 3 4 1 4.123 0.970 0.243
Hence
2 3
0 1 0 0
6 7
D ¼ 4 0:970 0:243 0:970 0:243 5 (11.7.8)
0 0 0:447 0:894
@CivilMethod
The finite element method Chapter 11 447
TABLE 11.7.2 Geometrical data of the elements of the frame in Fig. 11.7.2b.
Number of
element e Node j Node k D
x D
y Le cosfe sinfe
1 1 2 0 3 3 0 1
2 2 3 4 1 4.123 0.970 0.243
@CivilMethod
448 PART II Multi-degree-of-freedom systems
TABLE 11.7.3 Geometrical data of the elements of the frame in Fig. 11.7.2c.
Number of
element e Node j Node k D
x D
y Le cos fe sinfe
1 1 2 0 3 3 0 1
2 2 3 4 1 4.123 0.970 0.243
Hence
2 3
0 1:000 0 0
6 7
6 0:970 0:243 0:970 0:243 7
6 7
D¼6
60 0 0:447 0:894 7
7 (11.7.16)
6 7
40 0 0 1:000 5
0 0 1:000 0
TABLE 11.7.4 Geometrical data of the elements of the frame in Fig. 11.7.2d.
Number of
element e Node j Node k D
x D
y Le cosfe sin fe
1 1 2 0 3 3 0 1
2 2 3 4 1 4.123 0.970 0.243
3 3 4 2 4 4.472 0.447 0.894
4 3 5 1 2 2.236 0.447 0.894
@CivilMethod
450 PART II Multi-degree-of-freedom systems
Hence
2 3
0 1:000 0 0
6 0:970 0:243 0:970 0:243 7
6 7
D¼6 7 (11.7.19)
40 0 0:447 0:894 5
0 0 0:447 0:894
TABLE 11.7.5 Geometrical data of the elements of the frame in Fig. 11.7.2e.
Number of
element e Node j Node k D
x D
y Le cos fe sinfe
1 1 2 0 3 3 0 1
2 2 3 4 1 4.123 0.970 0.243
3 3 4 2 4 4.472 0.447 0.894
Hence
2 3
0 1:000 0 0
6 0:970 0:243 0:970 0:243 7
6 7
D¼6
60 0 0:447 0:894 7
7 (11.7.22)
40 0 0:447 0:894 5
0 0 1 0
The Gauss-Jordan elimination gives
2 3
1 0 0 0
60 1 0 07
6 7
D¼6 60 0 1 077 (11.7.23)
40 0 0 15
0 0 0 0
the rank of which is 4. Therefore, there is no translational independent variable.
Apparently, the fifth constraint is redundant.
Example 11.7.1 Formulate the equations of motion of the frame in Example
11.3.2, when the axial constraints of the elements are taken into account.
Solution
We note that in this example, two elements of the frame are not parallel to the
global axes. Therefore, the procedure for flexural vibrations described in
Example 11.5.2 is not convenient to formulate the equation of motion of the
structure. Consequently, the general procedure for the FEM analysis of a plane
frame should be employed.
First, the mass and stiffness matrices and the vector of the nodal loads
with respect to the seven free degrees of freedom are computed. Then, the rota-
tional degrees of freedom u3 , u6 , u7 are eliminated by static condensation. This
results in a system with four translational degrees of freedom (Fig. E11.8).
@CivilMethod
452 PART II Multi-degree-of-freedom systems
Number of
Element e Node j Node k D
x D
y Le cos fe sinfe
1 1 2 0 5 5 0 1
2 2 3 4 5 5 0.8 0.6
3 3 4 6 8 10 0.6 0.8
On the base of the geometrical data shown in Table E11.3, the constraint
equations are
8 9 8 9
2 3> u1 > > 0 >
> > > >
0 1:0 0 0 < >
> = > <0>=
6 7 u2
4 0:8 0:6 0:8 0:6 5 ¼
> >
> u4 >
> >0>>
0 0 0:6 0:8 > : > ; > : >;
u5 0
hence
2 3
0 1:0 0 0
6 7
D ¼ 4 0:8 0:6 0:8 0:6 5
0 0 0:6 0:8
Using the Gauss-Jordan Elimination, we find out that the rank of the
matrix D is 3. Thus, the number of independent degrees of freedom is equal
@CivilMethod
The finite element method Chapter 11 453
@CivilMethod
454 PART II Multi-degree-of-freedom systems
(Fig. 11.8.2). The element cross-sections are symmetric with respect to the xz
plane. The nodal loads consist of forces in the z direction and of moments
about x and y axes (Fig. 11.8.1). The elements may be subjected to loads
normal to the xy plane and moments about x and y axes. Moreover, the axial
deformation of the elements is ignored. Under these assumptions, the nodes of
the grid undergo rotations about the global axes x, y and a translational
displacement normal to the x y plane. These global displacements produce
the rotations u1 , u4 about the local x axis, the rotations u2 , u5 about the local
y axis, and the translational displacements u3 , u6 in the direction of the z axis
at the ends of the element (see Fig. 11.8.2).
Thus the element displacement vector is
u ¼ f u1 u2 u3 u4 u5 u6 g (11.8.1)
The displacements u1 , u4 produce torsion of the element while the dis-
placements u2 , u3 , u5 , u6 produce bending in x z plane. The bending deforma-
tion of the element is given by Eq. (11.3.3) after changing appropriately the
role of the shape functions i ðx Þ to correspond to the nodal displacements.
Thus, we have
v ðx, t Þ ¼ u2 2 ðx Þ + u3 3 ðx Þ + u5 5 ðx Þ + u6 6 ðx Þ (11.8.2)
where now
2 ðx Þ ¼ L x 2x 2 + x 3 (11.8.3a)
3 ðx Þ ¼ 1 3x
2
+ 2x 3 (11.8.3b)
5 ðx Þ ¼ L x 2 + x3 (11.8.3c)
6 ðx Þ ¼ 3x 2x3
2
(11.8.3d)
@CivilMethod
The finite element method Chapter 11 455
Fig. 11.8.3
dM t
¼0 (11.8.5)
dx
where Mt is the twisting moment given by
dq
Mt ¼ GI t (11.8.6)
dx
in which G ¼ E=2ð1 + n Þ is the shear modulus and It the torsional constant.
@CivilMethod
456 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 457
Fig. 11.8.4 Positive directions of the nodal forces of the grid element.
The equivalent nodal forces of the grid element will be obtained using the
Lagrange equations. Fig. 11.8.4 shows the positive directions of element nodal
forces. The actions in the directions of u1 , u2 , u4 , and u5 are moments while in
the directions of u3 and u6 are forces.
(i) Nodal elastic forces and stiffness matrix of the grid element
The elastic energy is due to the torsional deformation of the element given
by the expressionf
1
Ut ¼ GI t ðq0 Þ L
2
(11.8.13)
2
as well as to the bending deformation of the element given by Eq. (10.5.2),
that is,
The second integral is equal to It while the first one is equal to zero. Indeed, the first integral is
transformed as
Z
Z h i
R¼ f2y + f2z dA + ðzfÞy + ðyfÞz dA
A
Z A
Z Z
(4)
¼ f fyy + fzz dA + ffn dA f zn y yn z dA
A A A
¼0
The foregoing result is obtained by applying the Gauss-Green theorem to the first term and the
Gauss divergence theorem to the second term and taking into account that fyy + fzz ¼ 0 and
fn ¼ zn y yn z . Hence, Eq. (3) becomes
1 1
Ut ¼ GI t ðq0 Þ L ¼ Mt q0 L
2
(5)
2 2
@CivilMethod
458 PART II Multi-degree-of-freedom systems
Z L
1
EI ½v 00 ðx, t Þ dx
2
Ub ¼ (11.8.14)
2 0
or
f eS ¼ ke ue (11.8.19)
where u are the vectors of the nodal elastic forces and nodal displacements,
f eS , e
respectively, and ke the stiffness matrix of the e grid element. Further, substitut-
ing Eqs. (11.8.3a)–(11.8.3d) and (11.8.10) into Eqs. (11.8.18a), (11.8.18b) and
performing the integration gives
@CivilMethod
The finite element method Chapter 11 459
2 3
GI t GI t
6 L 0 0 0 7 0
6 L 7
6 4EI 6EI 2EI 6EI 7
6 0 2 0 7
6 L L L L2 7
6 7
6 6EI 12EI 6EI 12EI 7
6 0 2 0 2 3 7
6 L3 L 7
ke ¼ 6
6 GI t
L L 7
7 (11.8.20)
6 GI t 7
6 L 0 0 0 0 7
6 L 7
6 6EI 7
6 0 2EI
6EI 4EI 7
6 0 7
6 L L2 L L2 7
4 6EI 12EI 6EI 12EI 5
0 3 0
L2 L L2 L3
The superscript e has been dropped from Ite , I e , and Le for the sake of the
simplicity of the expressions.
(ii) Nodal inertial forces and mass matrix of the grid element
The equivalent inertial nodal forces result again on the base of two different
assumptions regarding the mass distribution along the length of the element,
namely the consistent mass assumption and the lumped mass assumption. The
inertial mass matrices resulting from both assumptions are derived right below.
(a) Consistent mass matrix
During the motion, the infinitesimal mass m ðx Þdx of the element undergoes the
two displacements, that is, the rotation qðx, t Þ and the transverse displacement
v ðx, t Þ. Therefore, the kinetic energy of the grid element will be given by the
expression
Z
1 Ln 2 o
T¼ I0 ðx Þ q_ ðx, t Þ + m ðx Þ½v_ ðx, t Þ2 dx (11.8.21)
2 0
where I0 ðx Þ is the polar moment of inertia of the mass m ðx Þ ¼ rAðx Þ with
respect to the axis of the element, which is assumed to coincide with the
twist axis.
Using Eqs. (11.8.2), (11.8.4), Eq. (11.8.21) becomes
Z h i
1 L
T ðu_ 1 , u_ 2 , …, u_ 6 Þ ¼ I0 ðx Þ ½u_ 1 1 ðx Þ + u_ 4 4 ðx Þ2 dx
2 0
Z
1 L
+ m ðx Þ½u_ 2 2 ðx Þ + u_ 3 3 ðx Þ + u_ 5 5 ðx Þ + u_ 6 6 ðx Þ2 dx
2 0
(11.8.22)
@CivilMethod
460 PART II Multi-degree-of-freedom systems
The nodal inertial forces result from Eq. (11.2.9) for i ¼ 1, 2, …, 6. Thus,
after performing the differentiations, we obtain
8 9 2m 0 38 9
>
>
fI 1 >
>
11 0 m14 0 0 > u€1 >
>
>
> >
> 6 7>>
> >
>
>
> f >
>
I2 >
6 0 m m 0 m m 7 >
> u€2>
>
> > 6 22 23 25 26 7 > >
>
> >
> > 6 7>> >
> >
>
< fI 3 = 6 0 m32 m33 0 m35 m36 7 u€3 =
7 <
6
¼6 7 (11.8.23)
>
>
> fI 4 >
>
>
6 m41 0
6 0 m44 0 0 7 7
>
>
> u€4 >
>
>
>
> >
> 7>> >
>
> > 6 > >
> fI 5 >
>
> > 6
> 4 0 m52 m53 0 m55 m56 7 5 > u€5 >
>
> >
>
>
: ; > : >
> ;
fI 6 0 m62 m63 0 m65 m66 u€6
where
Z L
mij ¼ I0 ðx Þ i ðx Þ j ðx Þdx i, j ¼ 1, 4 (11.8.24a)
0
Z L
mij ¼ m ðx Þ i ðx Þ j ðx Þdx i, j ¼ 2, 3, 5, 6 (11.8.24b)
0
@CivilMethod
The finite element method Chapter 11 461
Z L
m1 ¼ m ðx Þð1 x Þdx (11.8.27a)
0
Z L
m2 ¼ m ðx Þxdx (11.8.27b)
0
@CivilMethod
462 PART II Multi-degree-of-freedom systems
Fig. 11.8.5 Loading and equivalent nodal loads of a plane grid element.
Z L Z L
p
dWnc ¼ mx ðx, t Þdqðx, t Þdx + pz ðx, t Þdv ðx, t Þdx (11.8.31)
0 0
where
Z L
pi ðt Þ ¼ mx ðx, t Þ i ðx Þdx, i ¼ 1, 4 (11.8.33a)
0
Z L
pi ðt Þ ¼ pz ðx, t Þ i ðx Þdx, i ¼ 2, 3, 5, 6 (11.8.33b)
0
@CivilMethod
The finite element method Chapter 11 463
@CivilMethod
464 PART II Multi-degree-of-freedom systems
tion to u . Their transformation obeys the same law, that is, Eq. (11.8.40b).
e
Thus, we obtain
f e ¼ðRe ÞT f e (11.8.42)
S S
f e ¼ðRe ÞT f e (11.8.43)
I I
f e ¼ðRe ÞT f e (11.8.44)
D D
f e e €e
u
¼m (11.8.47)
I
f e ¼e _e
D c u (11.8.48)
e e , and
where k , m ce represent the stiffness, mass, and damping matrices of the e
grid element in global axes, respectively, and are given by
k ¼ ðRe ÞT ke Re
e
(11.8.49)
@CivilMethod
The finite element method Chapter 11 465
e ¼ ð Re Þ T m e Re
m (11.8.50)
ce ¼ ð Re Þ T ce Re (11.8.51)
Note that the lumped mass matrix remains the same under this transformation.
Applying the procedure presented in Section 11.2.3 for the plane truss, we
obtain the equation of motion of the plane grid, which, if damping is taken into
account, is written as
ee
M u€ + Cu_ + Ke
ee e u ¼e
pð t Þ (11.8.52)
or after applying the support conditions
M u€f + C
e ff e u_ f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e (11.8.53)
M u€f + M
e sf e u€s + C
e ss e u_ f + C
e sf e u_ s + K
e ss e e sf e e ss e
uf + K us ¼ e
ps ðt Þ (11.8.54)
where
pf∗ ðt Þ ¼ e
e pf ðt Þ M u€s C
e fs e u_ s K
e fs e e fs e
us (11.8.55)
For convenience we will write Eq. (11.8.53) Check this as
M€
u + C€
u + Ku ¼ pðt Þ (11.8.56)
Example 11.8.1 Formulate the equation of motion of the plane grid of Fig. E11.9.
The grid is fixed at nodes 3 and 4. The element 1 is loaded by the transverse
t. A moment M ðt Þ ¼ 3p0 L2 sin w
load pðx, t Þ ¼ p0 sin w t acts at the node 2
t. The cross-section
while the support 1 is subjected to the rotation qg ðt Þ ¼ q0 sin w
of all elements is rectangular with aspect ratio h=b ¼ 2. Consider lumped mass
assumption for the elements. Data: L ¼ 3:0m, h ¼ 0:50m, a ¼ p=4,
E ¼ 2:1 107 kN=m2 , n ¼ 0:2, q0 ¼ 0:5p0 , mass density r.
Solution
The system has n ¼ 4 nodes. Hence, the free structure has N ¼ 3n ¼ 12 degrees
of freedom. The numbering of the nodes and the positive direction of the
elements are shown in Fig. E11.9
@CivilMethod
466 PART II Multi-degree-of-freedom systems
Element
number e xj xk D
x yj yk D
y Le cosfe sinfe
1 0 0 0 0 3 3 3 0 1
2 0 2.12 2.12 3 5.12 2.12 3 0.707 0.707
3 0 2.12 2.12 3 5.12 2.12 3 0.707 0.707
@CivilMethod
The finite element method Chapter 11 467
2. Computation of k , m
e
e , pe ðt Þ for e ¼ 1, 2, 3
Matrices k . They are computed using Eq. (11.8.49)
e
2 3
36:0 0 18:0 18:0 0 18:0
6 0 0 2:57 0 7
6 2:57 0 7
6 7
EI 6 18:0 0 12:0 18:0 0 12:0 7
k ¼ 36
1 6 7
L 6 18:0 0 18:0 36:0 0 18:0 77
6 7
4 0 2:57 0 0 2:57 0 5
18:0 0 12:0 18:0 0 12:0
2 3
19:29 16:71 12:73 7:71 10:29 12:73
6 16:71 19:29 12:73 10:29 7:71 12:73 7
6 7
6 7
EI 6 12:73 12:73 12:0 12:73 12:73 12:0 7
k ¼ 36
2 6 7
L 6 7:71 10:29 12:73 19:29 16:71 12:73 7 7
6 7
4 10:29 7:71 12:73 16:71 19:29 12:73 5
12:73 12:73 12:0 12:73 12:73 12:0
2 3
19:29 16:71 12:73 7:71 10:29 12:73
6 16:71 19:29 12:73 10:29 7:71 12:73 7
6 7
6 7
EI 6 12:73 12:73 12:0 12:73 12:73 12:0 7
k3 ¼ 3 6 7
L 6 7:71 10:29 12:73 19:29 16:71 12:73 7
6
7
6 7
4 10:29 7:71 12:73 16:71 19:29 12:73 5
12:73 12:73 12:0 12:73 12:73 12:0
@CivilMethod
468 PART II Multi-degree-of-freedom systems
^ e, M
3. Computation of the enlarged matrices K ^ e, p
^e for e ¼ 1, 2, 3
Assembly matrices ae .
2 3
1 0 0 0 0 0 0 0 0 0 0 0
60 1 0 0 0 0 0 0 0 0 0 07
6 7
6 7
60 0 1 0 0 0 0 0 0 0 0 07
a ¼6
1
60
7,
6 0 0 1 0 0 0 0 0 0 0 07
7
6 7
40 0 0 0 1 0 0 0 0 0 0 05
0 0 0 0 0 1 0 0 0 0 0 0
2 3
0 0 0 1 0 0 0 0 0 0 0 0
60 0 0 0 1 0 0 0 0 0 0 07
6 7
6 7
60 0 0 0 0 1 0 0 0 0 0 07
a ¼6
2
60
7
6 0 0 0 0 0 1 0 0 0 0 07
7
6 7
40 0 0 0 0 0 0 1 0 0 0 05
0 0 0 0 0 0 0 0 1 0 0 0
2 3
0 0 0 1 0 0 0 0 0 0 0 0
60 0 0 0 1 0 0 0 0 0 0 07
6 7
6 7
60 0 0 0 0 1 0 0 0 0 0 07
a ¼6
3
60
7
6 0 0 0 0 0 0 0 0 1 0 077
6 7
40 0 0 0 0 0 0 0 0 0 1 05
0 0 0 0 0 0 0 0 0 0 0 1
@CivilMethod
The finite element method Chapter 11 469
Matrices K^ e . They are computed using the relation K^ e ¼ ðae ÞT ke ae (see
Eq. 11.2.92)
2 3
36:0 0 18:0 18:0 0 18:0 0 0 0 0 0 0
6 0 2:572 0 0 2:572 0 0 0 0 0 0 07
6 7
6 18:0 0 12:0 0 0 0 0 0 0 7
6 12:0 18:0 0 7
6 7
6 18:0 0 18:0 36:0 0 18:0 0 0 0 0 0 0 7
6 7
6 0 2:572 0 0 2:572 0 0 0 0 0 0 07
6 7
6 18:0 12:0 18:0 7
EI
^1 ¼ 6 6 0 0 12:0 0 0 0 0 0 0 7
K 7
L3 6 0 0 0 0 0 0 0 0 0 0 0 07
6 7
6 0 0 0 0 0 0 0 0 0 0 0 07
6 7
6 0 0 0 0 0 0 0 07
6 0 0 0 0 7
6 7
6 0 0 0 0 0 0 0 0 0 0 0 07
6 7
4 0 0 0 0 0 0 0 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3
0 0 0 0 0 0 0 0 0 0 0 0
60 07
6 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 19:29 16:71 12:73 7:71 10:29 12:73 07
6 0 0 0 0 7
6 7
60 0 0 16:71 19:29 12:73 10:29 7:71 12:73 0 0 07
6 7
^ EI 6
60 0 0 12:73 12:73 12:0 12:73 12:73 12:0 0 0 077
K ¼ 36
2
7
L 60 0 0 7:714 10:29 12:73 19:29 16:71 12:73 0 0 07
6 7
60 10:29 7:71 12:73 16:71 07
6 0 0 19:29 12:73 0 0 7
6 7
60 0 0 12:73 12:73 12:0 12:73 12:73 12:0 0 0 07
6 7
60 07
6 0 0 0 0 0 0 0 0 0 0 7
6 7
40 0 0 0 0 0 0 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3
0 0 0 0 0 0 0 0 0 0 0 0
60 0 7
6 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0
6 0 19:29 16:71 12:73 0 0 0 7:71 10:29 12:73 77
60 0
6 0 16:71 19:29 12:73 0 0 0 10:29 7:71 12:73 7
7
6 7
^ EI 60 0 0 12:73 12:73 12:0 0 0 0 12:73 12:73 12:0 7
7
K ¼ 36
3
7
L 660 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 7
6 0 7:714 10:29 12:73 0 0 0 19:29 16:71 12:73 7
6 7
40 0 0 10:29 7:71 12:73 0 0 0 16:71 19:29 12:73 5
0 0 0 12:73 12:73 12:0 0 0 0 12:73 12:73 12:0
@CivilMethod
470 PART II Multi-degree-of-freedom systems
2 3
0 0 0 0 0 0 0 0 0 0 0 0
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
6 07
^ 2 ¼ rA6 0
M
0 0 0 0 1:5 0 0 0 0 0 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
60 0 0 0 0 0 0 0 1:5 0 0 07
6 7
60 0 0 0 0 0 0 0 0 0 0 07
6 7
40 0 0 0 0 0 0 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
2 3
0 0 0 0 0 0 0 0 0 0 0 0
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 1:5 0 0 0 0 0 0 7
^ 3 ¼ rA6
M 6
7
7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
40 0 0 0 0 0 0 0 0 0 0 0 5
0 0 0 0 0 0 0 0 0 0 0 1:5
@CivilMethod
The finite element method Chapter 11 471
P
3 P
3 P
3
¼
4. Computation of M M ¼
^ e, K ^ e , and p ¼ Pðt Þ +
K ^ e ðt Þ
p
e¼1 e¼1 e¼1
2 3
0 0 0 0 0 0 0 0 0 0 0 0
6 7
60 0 0 0 0 00 0 0 0 0 0 7
6 7
6 7
60 0 1:5 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 4:5 0 0 0 0 0 0 7
¼ rA ¼ 6
M 6
7
7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 0 0 0 7
4 5
0 0 0 0 0 0 0 0 0 0 1:5
@CivilMethod
2 3
36:0 0 18:0 18:0 0 18:0 0 0 0 0 0 0
6 0 2:572 0 0 2:57 0 0 0 0 0 0 0 7
6 7
6 18:0 0 12:0 0 7
6 12:0 18:0 0 0 0 0 0 0 7
6 7
6 18:0 0 18:0 74:57 0 18:0 7:71 10:29 12:73 7:71 10:29 12:73 7
6 7
6 0 2:572 0 0 41:14 25:46 10:29 7:71 12:73 10:29 7:71 12:73 7
6 7
6 18:0 0 12:0 18:0 25:46 12:73 12:0 12:73 12:73 12:0 7
EI
¼ 6 6 36:0 12:73 7
K 7
L3 6 0 0 0 7:71 10:29 12:73 19:29 16:71 12:73 0 0 0 7
6 7
6 0 0 0 10:29 7:714 12:73 16:71 19:29 12:73 0 0 0 7
6 7
6 0 0 12:73 12:73 12:0 12:73 12:73 12:0 0 7
6 0 0 0 7
6 7
6 0 0 0 7:71 10:29 12:73 0 0 0 19:29 16:71 12:73 7
6 7
4 0 0 0 10:29 7:714 12:73 0 0 0 16:71 19:29 12:73 5
0 0 0 12:73 12:73 12:0 0 0 0 12:73 12:73 12:0
8 9 8 9 8 9
>
> Mg >
> >
> 0 > > >
> Mg >
>
>
> >
> >
> >
> >
> >
>
>
> P >
> >
> 0 >
> >
> P >
>
>
>
2 >
> >
> >
> >
>
2 >
>
>
>
> P3 >
>
>
>
>
> 1:5 >
>
>
>
>
> t >
P3 1:5p0 sin w >
>
> > > > > >
>
>
> 3p L 2
sin
w t
>
>
> > 0 >
>
> >
>
>
>
> 27p sin
w t
>
>
>
>
> 0 >
> >
> >
> >
> 0 >
>
>
> >
> >
> >
> >
> >
>
>
> 0 >
> >
> 0 >
> >
> 0 >
>
X
3 >
<0 >
= < 1:5 >
> = >
< 1:5p sin w t
>
=
ðt Þ +
pðt Þ ¼ P ^ i ðt Þ ¼
p + t ¼
p0 sin w
0
>
> P7 >
> >
> 0 >
> >
> P7 >
>
i¼1 >
> >
> >
> >
> >
> >
>
>
> P >
> >
> 0 >
> >
> P8 >
>
>
> 8 >
> >
> >
> >
> >
>
>
> >
> >
> >
> >
> >
>
>
> P 9 >
> >
> 0 >
> >
> P9 >
>
>
> >
> >
> >
> >
> >
>
>
> P >
> >
> 0 >
> >
> P10 >
>
>
>
10 >
> >
> >
> >
> >
>
>
> >
> >
> >
> >
> >
>
>
> P 11 >
> >
> 0 >
> >
> P11 >
>
: ; : ; : ;
P12 0 P12
The finite element method Chapter 11 473
@CivilMethod
474 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 475
2 3
0 0 0 0 0 0 0 0 0
60 0 0 0 0 0 0 0 0 7
6 7
60 0 1:5 0 0 0 0 0 0 7
2 3 6 7
0 0 0 60 0 0 0 0 0 0 0 0 7
6 7
e 4
Mff ¼ rA 0 0 0 5 e
, Mss ¼ rA6 0 7
60 0 0 0 0 0 0 0 7
0 0 4:5 60 0 0 0 0 1:5 0 0 0 7
6 7
60 0 0 0 0 0 0 0 0 7
6 7
40 0 0 0 0 0 0 0 0 5
0 0 0 0 0 0 0 0 1:5
M e sf ¼ 0
e fs ¼ M
8 9 8 9
>
> Mg >
> >
> q g ðt Þ >
>
>
> >
> >
> >
>
>
> P2 >
> >
> 0 >
>
>
> >
> >
> >
>
8 9 > 3
> P 1:5p0 sin w t >
> >
> 0 >
>
>
> >
> >
> >
>
< 27 = < 7
P = < 0 =
e
pf ðt Þ ¼ 0 t, e
p0 sin w p s ð t Þ ¼ P8 es ð t Þ ¼ 0
, u
: ; >
> >
> >
> >
>
1:5 >
> P9 >
> >
> 0 >
>
>
> >
> >
> >
>
>
> P >
> >
> 0 >
>
>
>
10
>
> >
> >
>
>
> P >
> >
> 0 >
>
: 11
; : ;
P12 0
e ff ¼ 0
The equation of motion results from Eq. (11.8.53) with C
M u€f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e
Taking into account that q0 ¼ 0:5p0 , Eq. (11.8.55) gives
8 9
< 18 =
p ∗ f ðt Þ ¼ e
e e fs u
pf ðt Þ M e sf us ¼ 0
€s K t
p sin w
: ; 0
7:5
Hence, the equation of motion is
5. Static condensation
The rotational degrees of freedom u4 , u5 are statically condensed. Obviously,
the displacements should be rearranged as u6 , u4 , u5 . This defines the
modification matrix
2 3
0 1 0
VC ¼ 4 0 0 1 5
1 0 0
@CivilMethod
476 PART II Multi-degree-of-freedom systems
which yields
EI EI EI 18:0
Ktt ¼ 36 , K ¼ ½ 18:0 25:45 , K qt ¼
L3 25:45
tq
L3 L3
EI 74:57 0
Kqq ¼ 3
L 0 41:14
Mtt ¼ 4:5rA, Mtq ¼ Mqt ¼ Mqq ¼ 0
27
t, pq ¼
pt ¼ 1:5p0 sin w t
p0 sin w
0
Hence
EI
Ktt∗ ¼ Ktt Ktq K1
qq Kqt ¼ 15:91
L3
pt∗ ¼ pt Ktq K1 t
qq pq ¼ 5:017p0 sin w
which are inserted into Eq. (11.4.3) to give the final form of the equation of
motion
EI
4:5rA u€6 + 15:91 t
u6 ¼ 5:017p0 sin w
L3
@CivilMethod
The finite element method Chapter 11 477
where u, v,w are the translations in the directions of the x, y, z axes, respec-
tively, and q the rotation about the x while i ðx Þ (i ¼ 1,2, …,12) are given by
1 ðx Þ ¼ 4 ðx Þ ¼ ð1 x Þ (11.9.4a)
7 ðx Þ ¼ 10 ðx Þ ¼ x (11.9.4b)
2 ðx Þ ¼ 3 ðx Þ ¼ 1 3x
2
+ 2x3 (11.9.4c)
6 ðx Þ ¼ 5 ðx Þ ¼ L x 2x 2 + x 3
(11.9.4d)
2
12 ðx Þ ¼ 11 ðx Þ ¼ L x + x
3
(11.9.4e)
9 ðx Þ ¼ 8 ðx Þ ¼ 3x 2x3
2
(11.9.4f)
On the base of Eqs. (11.3.8), (11.3.9), (11.8.13), Eq. (11.9.2) yields
Z Z
1 L 1 L
0
GI t ½q0 ðx Þ dx
2 2
U ðu1 , …, u12 Þ ¼ EA½u ðx, t Þ dx +
2 0 2 0
Z Z (11.9.5)
1 L 00 2 1 L 00 2
+ EI z ½v ðx, t Þ dx + EI y ½w ðx, t Þ dx
2 0 2 0
Further, for the consistent mass assumption, the kinetic energy is given by
Z n o
1 L
T ðu_ 1 , …, u_ 12 Þ ¼ m ðx Þ ½u_ ðx, t Þ2 + ½v_ ðx, t Þ2 + ½w_ ðx, t Þ2 dx
2 0
Z (11.9.6)
1 L 2
+ _
I0 ðx Þ q ðx, t Þ dx
2 0
while for the lumped mass assumption by
1 1
T ðu_ 1 , …, u_ 12 Þ ¼ m1 u_ 21 + u_ 22 + u_ 23 + m2 u_ 27 + u_ 28 + u_ 29 (11.9.7)
2 2
where
Z L Z L
m1 ¼ m ðx Þð1 xÞdx, m2 ¼ m ðx Þxdx (11.9.8)
0 0
The element mass and stiffness matrices as well as the equivalent nodal
forces are obtained using the method of the Lagrange equations as described
for the plane frame and the grid element. In the following, we consider an ele-
ment with constant cross-section.
(i) Nodal elastic forces and stiffness matrix of the space frame element
For an element with a constant cross-section, the stiffness matrix having dimen-
sions 12 12 results as
e e
k k
k ¼ jje jk
e
(11.9.9)
kkj kekk
@CivilMethod
The finite element method Chapter 11 479
where
2 3
L2 A 0 0 0 0 0
6 7
60 12Iz 0 0 0 6LI z 7
6 7
6 7
60 0 12Iy 0 6LI y 0 7
E 6 7
kejj ¼ 3 6 7 (11.9.10a)
L 660 0 0
G 2
L It 0 0
7
7
6 E 7
6 7
6 7
40 0 6LI y 0 4L2 Iy 0 5
0 6LI z 0 0 0 4L2 Iz
2 3
L2 A 0 0 0 0 0
6 7
60 12Iz 0 0 0 6LI z 7
6 7
6 7
60 0 12Iy 0 6LI y 0 7
E 6 7
kkk ¼ 3 6
e 7 (11.9.10b)
L 660 0 0
G 2
L It 0 0
7
7
6 E 7
6 7
60 7
4 0 6LI y 0 4L2 Iy 0 5
0 6LI z 0 0 0 4L2 Iz
2 3
L2 A 0 0 0 0 0
6 7
6 0 12Iz 0 0 0 6LI z 7
6 7
6 7
6 0 0 12Iy 0 6LI y 0 7
E 6 7
kekj ¼ 3 6 7 (11.9.10c)
L 66 0 0 0
G 2
L It 0 0
7
7
6 E 7
6 7
6 7
4 0 0 6LI y 0 2L2 Iy 0 5
0 6LI z 0 0 0 2L2 Iz
T
kejk ¼ kekj (11.9.10d)
where
2 3
140 0 0 0 0 0
60 156 0 0 0 22L 7
e 6
6 7
m 60 0 156 0 22L 0 7 7
mejj ¼ (11.9.12a)
420 660 0 0 140rg2 0 0 7 7
40 0 22L 0 4L2 0 5
0 22L 0 0 0 4L2
2 3
140 0 0 0 0 0
60 156 0 0 0 22L 7
6
e 6
7
m 6 0 0 156 0 22L 0 7 7
mkk ¼
e
(11.9.12b)
420 660 0 0 140rg2 0 0 7 7
40 0 22L 0 4L2 0 5
0 22L 0 0 0 4L2
2 3
70 0 0 0 0 0
60 54 0 0 0 13L 7
6
e 6
7
m 6 0 0 54 0 13L 0 7 7
mkj ¼
e
(11.9.12c)
420 660 0 0 70rg2 0 0 7 7
40 0 13L 0 3L2 0 5
0 13L 0 0 0 3L2
T
mejk ¼ mekj (11.9.12d)
pffiffiffiffiffiffiffiffiffiffi
m e ¼ rAL is the total mass of the e element and rg ¼ I0 =A is the radius of
gyration of the cross-section.
(b) Lumped mass matrix
According to this assumption, the mass of the element is concentrated at its
nodes, that is, they are obtained as the reactions of a simply supported beam
under the load m ðx Þ. Thus, for an element with constant mass, the mass lumped
assumption yields
2 3
m1 0 0 0 0 0
6 0 m1 0 0 0 0 7
6 7
6 0 0 m1 0 0 0 7
mjj ¼ 6
e 6 7 (11.9.13a)
7
60 0 0 0 0 07
40 0 0 0 0 05
0 0 0 0 0 0
2 3
m2 0 0 0 0 0
6 0 m2 0 0 0 0 7
6 7
6 0 0 m2 0 0 0 7
mkk ¼ 6
e 6 7 (11.9.13b)
7
60 0 0 0 0 07
40 0 0 0 0 05
0 0 0 0 0 0
mejk ¼ mekj ¼ 0 (11.9.13c)
@CivilMethod
The finite element method Chapter 11 481
pej ¼ f p1 p2 p3 p4 p5 p6 gT (11.9.15a)
@CivilMethod
482 PART II Multi-degree-of-freedom systems
where i ðx Þ ¼ 1 x, i + 6 ðx Þ ¼ x, i ¼ 1, 2, 3, 4. Therefore
pej ðt Þ ¼ f p1 p2 p3 p4 0 0 gT (11.9.18a)
Fig. 11.9.3 Global and local systems of axes of the space frame element.
@CivilMethod
The finite element method Chapter 11 483
@CivilMethod
484 PART II Multi-degree-of-freedom systems
This is a unit vector and as a vector product, it is normal to the plane defined
by the vectors e1 and r, that is, it is in the direction of the local z axis. Its com-
ponents give the direction cosines l31 , l32 , l33 .
Finally, the direction cosines l21 , l22 , l23 will result from the components
of the vector
e2 ¼ e3 e1 (11.9.31)
Apparently, the end displacements (translations and rotations) of the
element are transformed according to Eq. (11.9.27). Thus, we have
8 9 2 38 9
< u1 = l11 l12 l13 < u1 =
u ¼ 4 l21 l22 l23 5 u2 (11.9.32a)
: 2; : ;
u3 l31 l32 l33 u3
8 9 2 38 9
< u4 = l11 l12 l13 < u4 =
u5 ¼ 4 l21 l22 l23 5 u5 (11.9.32b)
: ; : ;
u6 l31 l32 l33 u6
8 9 2 38 9
< u7 = l11 l12 l13 < u7 =
u ¼ 4 l21 l22 l23 5 u8 (11.9.32c)
: 8; : ;
u9 l31 l32 l33 u9
8 9 2 38 9
< u10 = l11 l12 l13 < u10 =
u ¼ 4 l21 l22 l23 5 u11 (11.9.32d)
: 11 ; : ;
u12 l31 l32 l33 u12
which are combined to yield
ue ¼ Re ue (11.9.33)
where
2 3
L 0 0 0
60 L 0 07
R ¼6
e
40
7 (11.9.34)
0 L 05
0 0 0 L
Obviously, the matrix Re with dimensions 12 12 represents the transfor-
mation matrix of the e space frame element. This vector is orthonormal, hence
Eq. (11.9.33) is inverted as
f e ¼ðRe ÞT f eI (11.9.37)
I
@CivilMethod
The finite element method Chapter 11 485
f e ¼ðRe ÞT f e (11.9.38)
D D
k ¼ ðRe ÞT ke Re
e
(11.9.40)
e T e
ce ¼ ð R Þ c Re (11.9.41)
e T
e ¼ ð R Þ m e Re
m (11.9.42)
Applying the procedure presented in Section 11.2.3 for the plane truss,
we obtain the equation of motion of the space frame, which, if damping is taken
into account, is written as
M u_ + K
u€ + C u ¼
pðt Þ (11.9.43)
or after applying the support conditions and partitioning
M u€f + C
e ff e u_ f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e (11.9.44)
M u€f + M
e sf e u€s + C
e ss e u_ f + C
e sf e u_ s + K
e ss e e sf e e ss e
uf + K us ¼ e
ps ðt Þ (11.9.45)
where
pf∗ ðt Þ ¼ e
e pf ðt Þ M u€s C
e fs e u_ s K
e fs e e fs e
us (11.9.46)
For convenience, we will write Eq. (11.9.44) as
M€
u + Cu_ + Ku ¼ pðt Þ (11.9.47)
@CivilMethod
486 PART II Multi-degree-of-freedom systems
Then we compute
rp ¼ xp xj e1 + yp yj e2 + zp zj e3
e1 e2 e3
¼5
2 3
e1 e2 e3
6 7
rp ¼ 6
e1 4 0:904 0:301 0:301 7
5 ¼ 0
e1 + 2:412
e2 2:412
e3
5 1 1
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
je1 rj ¼ 02 + 2:4122 + ð2:412Þ2 ¼ 3:141 (2)
@CivilMethod
The finite element method Chapter 11 487
(a)
Fig. E11.10a Space frame in Example 11.9.2.
(b)
Fig. E11.10b Translational and rotational degrees of freedom in Example 11.9.2.
Solution
The system has n ¼ 4 nodes. Thus, the free structure has N ¼ 6n ¼ 24
degrees of freedom. We set Iy ¼ I , hence A ¼ 300I , Iz ¼ 4I , and
It ¼ 0:229hb3 ¼ 2:748I . Moreover, it is G ¼ 0:4166E.
e xj xk D
x yj yk D
y zj zk D
z
@CivilMethod
488 PART II Multi-degree-of-freedom systems
2 3
7500 0 0 0 0 0
60 48 0 0 0 120 7
6 7
EI 6 0 0 12 0 30 0 7
k1jj ¼ 3 6 7,
L 660 0 0 28:62 0 0 7 7
40 0 30 0 100 0 5
0 120 0 0 0 400
2 3
7500 0 0 0 0 0
60 48 0 0 0 120 7
6 7
EI 6 0 0 12 0 30 0 7
kkk ¼ 3 6
1 7
L 66 0 0 0 28:62 0 0 7 7
40 0 30 0 100 0 5
0 120 0 0 0 400
2 3
7500 0 0 0 0 0
6 0 48 0 0 0 120 7
6 7
T
EI 6 0 0 12 0 30 0 7
6
kjk ¼ 3 6
1 7, k 1 ¼ k1
L 6 0 0 0 28:62 0 0 7 7
kj jk
4 0 0 30 0 50 0 5
0 120 0 0 0 200
k2 ¼ k3 ¼ k1
2 3 2 3
0:5 0 0 0 0 0 0:5 0 0 0 0 0
60 0:5 0 0 0 07 6 0 0:5 0 0 0 07
6 7 6 7
60 0 0:5 0 0 0 7 6 07
mjj ¼ rAL6
1 7, m1 ¼ rAL6 0 0 0:5 0 0 7
60 0 0 0 0 07 kk 60 0 0 0 0 07
6 7 6 7
40 0 0 0 0 05 40 0 0 0 0 05
0 0 0 0 0 0 0 0 0 0 0 0
T
m1jk ¼ m1kj ¼ 0 and m2 ¼ m3 ¼ m1
@CivilMethod
The finite element method Chapter 11 489
Because the loading is referred to the global axes, we obtain directly the
equivalent nodal loads pe ðt Þ
8 9 8 9 8 9
>
> 0 > > >
> 0 > > >
> 0>>
> >
> > >
> > >
>
> 2:5 >
>
>
> 2:5 >
>
>
> 0>>
>
< = >
< = < >
> =
0 0 0
pj ¼
2
Lf ðt Þ, pk ¼
2
Lf ðt Þ, pj ¼ pk ¼ pj ¼ pk ¼
1 1 3 3
> 0 >
> > >
> 0 > > >
> 0>>
> > > > > >
>
> 0 > >
>
>
> 0 > >
>
>
> 0>>
>
: ; >
: ; : >
> ;
0 0 0
Taking the positive direction of the local x to coincide with increasing node
number and the points P1 20 , P2 O, P3 30 to define the principal planes xy
of the elements 1, 2, 3, respectively, and Table E11.5, we obtain:
Element 1:
e1 rP1
e3 ¼ ¼ ½ 0 0 1 , e2 ¼ e3 e1 ¼ ½ 0:5 0:866 0
je1 rP1 j
2 1 3
2 3 L 0 0 0
0:866 0:500 0 60
4 5 6 L1 0 0 77
L ¼ 0:5
1
0:866 0 , R ¼ 4
1
0 0 L1 0 5
0 0 1
0 0 0 L1
Element 2:
e1 rP2
e1 ¼ ½ 0 0 1 , rP2 ¼ ½ 0 2:5 2:5 , e3 ¼ ¼ ½ 1 0 0
je1 rP2 j
e2 ¼ ½ 0 1 0
2 3
2 3 L2 0 0 0
0 0 1 60 L2 0 0 7
L2 ¼ 4 0 1 0 5, R2 ¼ 6
40
7
0 L2 0 5
1 0 0
0 0 0 L2
@CivilMethod
490 PART II Multi-degree-of-freedom systems
Element 3:
e1 rP3
e1 ¼ ½ 0:866 0:500 0 , rP3 ¼ ½ 0 2:5 0 , e3 ¼ ¼ ½ 0 0 1
je1 rP3 j
e2 ¼ ½0:50 0:866 0
2 3 3
2 3 L 0 0 0
0:866 0:500 0 60 L3 0 0 7
L ¼ 4 0:500 0:866 0 5, R ¼ 6
3 3
40
7
0 L3 0 5
0 0 1
0 0 0 L3
X
3 X
3 X
3
¼ ðae ÞT k ae , M
¼
e
K e ae , pðt Þ ¼ Pext +
ða e ÞT m ðae ÞT pe
e¼1 e¼1 e¼1
M u€f + K
e ff e e ff e
uf ¼ pt∗
@CivilMethod
2 38 € 9
1:0 0 0 0 0 >
> u7 >
0 >
>
> >
>
6 7>> >
60 7
1:0 0 0 0 0 7>> u8 >
€ >
>
6 >
> >
>
6 7>> >
>
60 7 <
0 1:0 0 0 0 7 u€9 =
6
rAL6 7
60 0 0 1:0 0 0 7 > u€13 >
6 7>> >
>
6 7>>
>
>
>
>
60 7
0 0 0 1:0 0 5>> >
4 > 14 >
> u€ >
>
>
> >
:€ > ;
0 0 0 0 0 1:0 u15
2 38 9
225L2 + 7:68 130L2 6:17 3:17 4:68 0:976 3:17 >
> u7 >>
6 130L2 6:17 75L2 + 13:6 7>> >
6 1:68 0:976 1:26 1:68 7>> u8 >>
>
6 7 >
< >
=
EI 6 6:17 1:68 2
300L + 7:66 3:17 1:68 300L 0:88 7 u9
2
+ 36 7
L 6 4:68 3:17 225L2 + 7:68 130L2 + 6:17 7> u13 >
7> >
0:976 3:17
6 >
> >
>
4 0:976 1:26 1:68 1:68 2
75L + 13:6 1:68 5 >
> u14 >
>
>
: >
;
3:17 1:68 300L2 0:88 3:17 1:68 300L2 u15
8 9
>
> 9:39 >
>
>
> >
>
> 16:26 2:5L >
>
>
>
< >
=
0
¼ f ðt Þ
>
> 140:61 >
>
>
> >
>
>
> 243:54 2:5L >
>
>
: >
;
0
492 PART II Multi-degree-of-freedom systems
To save space, the intermediate computed matrices are not given here
because they are very large. The reader can find them on this book’s companion
website.
Reduction of the degrees of freedom due to axial constraints. The equations
of the constraints are
@CivilMethod
The finite element method Chapter 11 493
Hence
2 3 2 3
0:5 0 0 0:866 0 0
Da ¼ 4 0 1 0 5, Db ¼ 4 0 0 1 5,
0 0 0:5 0 0:866 0
2 3
1 0 0
6 7
6 0 1 0 7
I 6 0 0 1 7
T¼ ¼ 6 7
1
Db Da 6 7
6 0:577 0 0 7
4 0 0 0:577 5
0 1 0
2 3
1:33 0 0
^ ¼ TT V
M ^TM
e tt V
^ T ¼ rAL4 0 2 0 5
0 0 1:33
2 3
23:3 3:35 1:43
^ ¼ TT
K ^TK
V e ∗V ^ T ¼ EI 4 3:35 13:5 3:35 5
tt
L3
1:43 3:35 23:3
8 9
T ∗ < 21:68 2:5L =
p ^ p ðt Þ ¼
^ðt Þ ¼ TT V 0 f ðt Þ
t
: ;
324:73 2:5L
2 38 9 2 38 9
1:33 0 0 < u€8 = EI 23:3 3:35 1:43 < u8 =
rAL4 0 2 0 5 u€9 + 4 3:35 13:5 3:35 5 u9
: € ; L3 : ;
0 1:33 u14 1:43 3:35 23:3 u14
8 9
< 21:68 2:5L =
¼ 0 f ðt Þ
: ;
324:73 2:5L
@CivilMethod
494 PART II Multi-degree-of-freedom systems
u ¼ f u1 u2 u3 u4 u5 u6 gT (11.10.1)
According to what we mentioned for the plane truss, a point on the x
local axis of the element will undergo displacements along the x, y and z axes,
that is,
u ðx, t Þ ¼ u1 1 ðx Þ + u4 4 ðx Þ (11.10.2a)
v ðx, t Þ ¼ u2 2 ðx Þ + u 5 5 ðx Þ (11.10.2b)
w ðx, t Þ ¼ u3 3 ðx Þ + u6 6 ðx Þ (11.10.2c)
@CivilMethod
The finite element method Chapter 11 495
where
1 ðx Þ ¼ 2 ðx Þ ¼ 3 ðx Þ ¼ 1 x, x ¼ x=L (11.10.3a)
4 ðx Þ ¼ 5 ðx Þ ¼ 6 ðx Þ ¼ x (11.10.3b)
Because the element does not undergo bending, the axes on the cross-section
of the rod may have any orientation. Nevertheless, it is recommended to take
them in the directions of the cross-sectional principal axes.
The elastic energy of the element is due only to the axial deformation.
Therefore, for an element with variable cross-section Aðx Þ and mass m ðx Þ, it is
Z
1 L
EAðx Þ½u 0 ðx, t Þ dx
2
U ðu1 , u2 , …, u6 Þ ¼ (11.10.4)
2 0
The kinetic energy for consistent mass assumption is
Z n o
1 L
T ðu_ 1 , u_ 2 , …, u_ 6 Þ ¼ m ðx Þ ½u_ ðx, t Þ2 + ½v_ ðx, t Þ2 + ½w_ ðx, t Þ2 dx
2 0
(11.10.5)
while for lumped mass assumption is
1 1
T ðu_ 1 , u_ 2 , …, u_ 6 Þ ¼ m1 u_ 21 + u_ 22 + u_ 23 + m2 u_ 24 + u_ 25 + u_ 26 (11.10.6)
2 2
where
Z L Z L
m1 ¼ m ðx Þð1 x Þdx and m2 ¼ m ðx Þxdx (11.10.7)
0 0
Applying the procedure developed for the plane truss, we obtain the
stiffness matrix, the mass matrix, and the vector of equivalent nodal forces
of the space truss.
(i) Nodal elastic forces and stiffness matrix of the space truss element
The relation between the elastic nodal forces and nodal displacements is
f eS ¼ ke ue (11.10.8)
in which
8 9
> fS1 >
> >
>f >
> >
>
> S2 >
>
> >
<f >
> =
S3
fS ¼
e
(11.10.9a)
> fS4 >
> >
>
> >
>
>
> >
>
> f >
: >
> S5
;
fS6
@CivilMethod
496 PART II Multi-degree-of-freedom systems
8 9
> u1 >
> >
> >
>
>
> u2 >
>
>
> >
<u >
> =
3
u ¼
e
(11.10.9b)
>
> u4 >
>
>
> >
>
>
> >
>
> u >
: >
> 5
;
u6
2 3
k11 0 0 k14 0 0
60 0 07
6 0 0 0 7
6 7
60 0 0 0 0 07
ke ¼ 6
6k
7 (11.10.9c)
6 41 0 0 k44 0 07
7
6 7
40 0 0 0 0 05
0 0 0 0 0 0
where
Z L
0 0
kij ¼ EAðx Þ i j dx, i, j ¼ 1, 4 (11.10.10)
0
(ii) Nodal inertial forces and mass matrix of the truss element
The relation between the inertial nodal forces and nodal displacements is
€e
f eI ¼ me u (11.10.12)
in which
8 9
> fI 1 >
> >
>f >
> >
>
> I2 >>
> >
<f >
> =
I3
fI ¼
e
(11.10.13a)
>
> fI 4 >
>
>
> >
>
>
> >
>
> f >
: >
> I 5
;
fI 6
@CivilMethod
The finite element method Chapter 11 497
8 9
> u€1 >
> >
> >
>
>
> u€2 >
>
>
> >
< u€ >
> =
3
€ ¼
u e
(11.10.13b)
>
> u€4 >
>
>
> >
>
>
> € >
>
> u >
: >
> 5
;
u€6
and me is the mass matrix.
For the consistent mass assumption, we obtain
2 3
m11 0 0 m14 0 0
60 m22 0 0 m25 0 7
6 7
6 0 0 m 0 0 m36 7
m ¼6
e 6 33 7 (11.10.14)
6 m41 0 0 m44 0 0 7 7
40 m52 0 0 m54 0 5
0 0 m63 0 0 m66
where
Z L
mij ¼ mi + 1, j + 1 ¼ mi + 2, j + 2 ¼ m ðx Þ i ðx Þ j ðx Þdx, i, j ¼ 1, 4
0
(11.10.15)
¼ constant the matrix
For m ðx Þ ¼ m in Eq. (11.10.14) becomes
2 3
2 0 0 1 0 0
60 2 0 0 1 07
6 7
me 6
60 0 2 0 0 17 7
m ¼
e
(11.10.16)
6 6
61 0 0 2 0 07 7
40 1 0 0 2 05
0 0 1 0 0 2
¼ rAL is the total mass of the element.
where m e ¼ mL
For lumped mass assumption, the mass matrix results as
2 3
m11 0 0 0 0 0
60 m22 0 0 0 0 7
6 7
6 0 0 m33 0 0 0 7
me ¼ 6
60
7 (11.10.17)
6 0 0 m44 0 0 7 7
40 0 0 0 m55 0 5
0 0 0 0 0 m66
where m11 ¼ m22 ¼ m33 ¼ m1 , m44 ¼ m55 ¼ m66 ¼ m2 and
Z L Z L
m1 ¼ m ðx Þð1 xÞdx, m2 ¼ m ðx Þxdx (11.10.18)
0 0
@CivilMethod
498 PART II Multi-degree-of-freedom systems
where
L 0
Re ¼ (11.10.24)
0 L
is the transformation matrix of the e element of the space truss with dimensions
6 6. The remaining procedure for formulating the equation of motion is the
same as for the plane truss.
Example 11.10.1 Formulate the equation of motion of the space truss of
Fig. E11.11. All bars have the same cross-sectional area A. The distributed load
on element 1 is pðx, t Þ ¼ ð1 x=LÞP ðt Þ=L. Assume: modulus of elasticity E,
material density r, lumped mass assumption.
Solution
@CivilMethod
500 PART II Multi-degree-of-freedom systems
@CivilMethod
The finite element method Chapter 11 501
Z L
p11 ðt Þ ¼ pðx, t Þ 1 ðx Þdx ¼ 0:333P ðt Þ,
0
Z L
p14 ðt Þ ¼ pðx, t Þ 4 ðx Þdx ¼ 0:167P ðt Þ
0
8 9
>
> 0:333 >
>
>
> >
>
>
> 0 >
>
< =
0
p ðt Þ ¼
1
P ðt Þ, p2 ðt Þ ¼ p3 ðt Þ ¼ 0
>
> 0:167 >
>
>
> >
>
>
>0 >
>
: ;
0
Computation of k , m
e
e , pe ðt Þ, e ¼ 1, 2, 3
T
k1 ¼ R1 k1 R1
2 3
0:101 0:135 0:27 0:101 0:135 0:27
6 0:135 0:18 0:135 0:18 0:36 7
6 0:36 7
6 7
EA 6
6
0:27 0:36 0:719 0:27 0:36 0:719 7
7
¼
L 6
6 0:101 0:135 0:27 0:101 0:135 0:27 7
7
6 7
4 0:135 0:18 0:36 0:135 0:18 0:36 5
0:27 0:36 0:719 0:27 0:36 0:719
k2 ¼ R2 k2 R2
T
2 3
0:101 0:135 0:27 0:101 0:135 0:27
6 0:135 0:36 0:36 7
0:135 0:18
6 0:18 7
6 7
EA 6
6
0:27 0:36 0:719 0:27 0:36 0:719 7
7
¼ 6
L 6 0:101 0:135 0:27 0:101 0:135 0:27 77
6 7
4 0:135 0:18 0:36 0:135 0:18 0:36 5
0:27 0:36 0:719 0:27 0:36 0:719
T
k3 ¼ R3 k3 R3
2 3
0:101 0:135 0:27 0:101 0:135 0:27
6 0:135 0:36 0:135 0:18 0:36 7
6 0:18 7
6 7
EA 6
6
0:27 0:36 0:719 0:27 0:36 0:719 7
7
¼ 6
L 6 0:101 0:135 0:27 0:101 0:135 0:27 77
6 7
4 0:135 0:18 0:36 0:135 0:18 0:36 5
0:27 0:36 0:719 0:27 0:36 0:719
@CivilMethod
502 PART II Multi-degree-of-freedom systems
2 3
0:5 0 0 0 0 0
6 0 0:5 0 0 0 0 7
6 7
6 7
6 0 0 0:5 0 0 0 7
1 ¼m
m 2 ¼m 3 ¼ rAL66 0 0 0 0:5 0 0 7,
7
6 7
6 7
4 0 0 0 0 0:5 0 5
0 0 0 0 0 0:5
8 9
> 0:106 >
>
> >
>
>
>
> 0:141 >
>
>
> >
< 0:283 >
> =
p1 ðt Þ ¼ P ðt Þ, p2 ðt Þ ¼ p3 ðt Þ ¼ 0
>
> 0:053 >
>
>
> >
>
>
>
>
> 0:070 >
>
>
>
: ;
0:141
Subsequently, the assembly matrices ae , ðe ¼ 1, 2, 3Þ are formulated
2 3 2 3
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0
60
6 1 0 0 0 0 0 0 0 0 0 077
60
6 0 0 0 1 0 0 0 0 0 0 077
60 0 1 0 0 0 0 0 0 0 0 07 6
7, a2 ¼ 6 0 0 0 0 0 1 0 0 0 0 0 077
a1 ¼ 6
60
6 0 0 0 0 0 0 0 0 1 0 077
60
6 0 0 0 0 0 0 0 0 1 0 077
40 0 0 0 0 0 0 0 0 0 1 05 40 0 0 0 0 0 0 0 0 0 1 05
0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1
2 3
0 0 0 0 0 0 1 0 0 0 0 0
60
6 0 0 0 0 0 0 1 0 0 0 077
60 0 0 0 0 0 0 0 1 0 0 077
a3 ¼ 6
60
6 0 0 0 0 0 0 0 0 1 0 077
40 0 0 0 0 0 0 0 0 0 1 05
0 0 0 0 0 0 0 0 0 0 0 1
X
3 X
3 X
3
¼ ðae ÞT k ae , M
¼
e
K e ae , pðt Þ ¼ Pðt Þ +
aT m ðae ÞT pðt Þe
e¼1 e¼1 e¼1
e M
Modified matrices K, e and load vector p eðt Þ due to support conditions
The equation of motion will be derived with respect to the three free degrees
of freedom of node 4. Due to the support conditions, the displacement vector
should be modified as
e e1 u
uT ¼ f u e2 u e3 ue4 ue5 ue6 ue7 ue8 ue9 ue10 ue11 u
e12 g
¼ f u10 u11 u12 u1 u2 u3 u4 u5 u6 u7 u8 u9 g
@CivilMethod
The finite element method Chapter 11 503
After partitioning the modified matrices, the equation of motion results from
Eq. (10.2.117) which in terms of the original numeration reads
2 38 9 2 38 9
1:5 0 0 < u€10 = EA 0:303 0:135 0:270 < u10 =
rAL4 0 1:5 0 5 u€11 + 4 0:135 0:539 0:359 5 u11
:€ ; L : ;
0 0 1:5 u12 0:270 0:359 2:157 u12
8 9
< 0:053 =
¼ 0:929 P ðt Þ
: ;
1:859
@CivilMethod
504 PART II Multi-degree-of-freedom systems
Fig. 11.11.2 Element e of a space frame with rigid bodies at its ends.
coordinates of the motion. This imposes the transfer of the element end
forces from its ends j and k to J and K , respectively. This is achieved by work-
ing as follows.
The positive direction of the element is from point j to point k. We denote by
Dx j ¼ xj xJ , D
y j ¼ yj yJ , D
z j ¼ zj zJ the coordinates of point j with
respect to the system J x y z (see Fig. 11.11.3). The position vector defined
by the relation
rj ¼ D
x j i + D
y j j + D
zj k (11.11.1)
will be referred to as the translation (transfer) vector or the offset vector and is
directed from point J to point j.
(i) Transfer of the nodal forces from j to J
@CivilMethod
The finite element method Chapter 11 505
The nodal element forces at the end j and those transferred to point J are
given by
ð11:11:2a; bÞ
f bJ ¼ f bj + rj f aj (11.11.3b)
After expanding the vector product using vector components, the foregoing
equations are written as
ð11:11:4Þ
or
@CivilMethod f J ¼ Tj f j (11.11.5)
506 PART II Multi-degree-of-freedom systems
where
I 0
Tj ¼ (11.11.6)
ej I
represents the transformation matrix of the end forces from point j to J .
I represents 3 3 unit matrix and
2 3
0 Dzj Dyj
ej ¼ 4 D zj 0 Dxj 5 (11.11.7)
Dyj Dxj 0
is the offset (or eccentricity) matrix, which is equal to zero in the absence of a
rigid body because rj ¼ 0. Obviously, the matrix ej is antisymmetric, namely
eTj ¼ ej (11.11.8)
Eq. (11.11.5) can be readily inverted by taking point j as a reference point
and repeating the previous procedure. This yields
f j ¼ TTj f J (11.11.9)
where
I ej I eTj
TTj ¼ ¼ ¼ (11.11.10)
0 I 0 I
ð11:11:11a; bÞ
uj ¼ TTj uJ (11.11.13)
Similar relations are obtained for the nodal displacements and forces at end
k. Namely
uk ¼ TTk uK (11.11.14)
f k ¼ TTk f K (11.11.15)
where
I eTk
TTk ¼ (11.11.16)
0 I
2 3
0 Dzk D
yk
ek ¼ 4 Dzk 0 xk 5
D (11.11.17)
Dy k Dxk 0
x k , D
The quantities D y k , Dz k are the coordinates of the point k with respect
to the system of axes K x y z (see Fig. 11.11.2). Thus, the offset vector is
rk ¼ D
x k i + D
y k j + D
zk k (11.11.18)
Eqs. (11.11.5), (11.11.15) are combined as
f 0 e ¼ Te f e (11.11.19)
where
f e ¼ fj 0e f T 0
, f ¼ J , Te ¼ j (11.11.20)
fk fK 0 Tk
The components of the vector P0 ðt Þ represent the loads that are applied
directly to the virtual nodes. But in real structures, the external loads are usually
applied to points not coinciding with the virtual nodes, therefore they must
previously be transferred to these points. The procedure is the same as that
employed for the element nodal forces. Thus, if P is the point of application
of the load with an offset vector.
rp ¼ D
x p i + D
y p j + D
zpk (11.11.29)
its transfer from P to J is performed using the relation
PJ ðt Þ ¼ Tp Pp (11.11.30)
Eq. (11.11.27) is not the equation of motion of the structure because so far
the inertial forces produced by the motion of the rigid bodies have not been
included in it. Apparently, these forces are the most important in the considered
type structures.
This is achieved if the motion of the rigid bodies is taken into account.
In Appendix, it is shown that the motion of a rigid body with respect to its center
of mass is described by Eqs. (A.2.2), (A.2.8), namely
€c
F ¼ mR (11.10.31a)
Mc ¼ H
_ c (11.10.31b)
where F is the resultant force, Mc the resultant moment with respect to the
center of mass, and Hc the angular momentum.
Introducing Eq. (A.2.14) in Eq. (11.11.31b), using (Α.2.11) and neglecting
the nonlinear terms (the products of rotations) because we examine small
displacements, we obtain the linearized Euler’s equations with respect to the
principal axes of the body
@CivilMethod
The finite element method Chapter 11 509
€c
F ¼ mR (11.11.32a)
Mc ¼ Ix w_ x + Ixy w_ y + Ixz w_ z e1
+ Iyx w_ x + Iy w_ y + Iyz w_ z e2 (11.11.32b)
+ Izx w_ x + Izy w_ y + Iz w_ z e3
The quantities Ix , Ixy , …, Iz are the moments and the products of inertia,
which are given by Eq. (Α.2.15). Using the notation of the current section,
we write the forgoing equations
€c
f c ¼ Mc u (11.11.33)
where
8 9 8 9 2 3
>
> fc1 >
> >
> uc1 >> m 0 0 0 0 0
> >
> > >
> >
> 60 0 7
> f > > u > m 0 0 0
< >
> < >
> 6 7
c2 c2
= = 60
f uc3 0 m 0 0 0 7
f c ¼ c3 , uc ¼ , Mc ¼ 6
60
7 (11.11.34)
> fc4 >
> > >
> uc4 >> 6 0 0 Ix Ixy Ixz 7
7
>
> >
> >
> >
> 40
> fc5 >
> > > u > 0 0 Iyx Iy Iyz 5
: ; : c5 >
> ;
fc6 uc6 0 0 0 Izx Izy Iz
f c , uc , and Mc the vector of the inertia forces applied at the center of mass C , the
displacement vector of the center of mass and the mass (inertia) matrix of
the body with respect to center of mass, respectively. If the point of reference
of the motion is not the center of mass, but an arbitrary point P, then Mc must
be transformed from C to P. The transformation of f c and uc is achieved by
Eqs. (11.11.5), (11.11.13). Thus, we have
f p ¼ Tc f c (11.11.35)
uc ¼ TTc up (11.11.36)
where
2 3
0 D
zc D
yc
I 0
Tc ¼ , ec ¼ 4 Dzc 0 xc 5
D (11.11.37)
ec I
Dy c D
xc 0
and D x c ¼ xc xp , D
y c ¼ yc yp , D
z c ¼ zc zp are the components of the
offset vector rc from the center of mass to the reference point P.
After that, we may write
f p ¼ Tc f c
€c
¼ Tc M c u
€p
¼ Tc Mc TTc u
Hence
Mp ¼ Tc Mc TTc (11.11.38)
of 0
Then the mass matrices of all rigid bodies are added to mass matrix M
Eq. (11.11.27) after they are enlarged by appropriate assembly matrices. There-
@CivilMethod
fore, we obtain the equation of motion
510 PART II Multi-degree-of-freedom systems
∗ u€0 + K
M 0 u0 ¼ p0 ðt Þ (11.11.39)
where now
X
Nb X
Ne
^0
e
∗¼
M ^i +
M M (11.11.40)
p
i¼1 e¼1
in which M^ i are the enlarged mass matrices of the Nb rigid bodies. The proce-
p
dure of determining M ∗ is illustrated with Example 11.11.1.
Type of structure Tj Mc
2 3 2 3
1. Plane truss 1 0 m 0 0
(structure in 4 0 1 5 40 m 05
xy-plane) D yj D xj 0 0 Iz
2 3 2 3
2. Plane frame 1 0 0 m 0 0
(bending in 4 0 1 05 40 m 05
xy-plane) D yj D xj 1 0 0 Iz
2 3 2 3
3. Grid (structure 1 0 D yj Ix Ixy 0
in xy-plane) 4 0 1 D xj 5 4 Iyx Iy 0 5
0 0 1 0 0 m
2 3 2 3
4. Space truss 1 0 0 m 0 0 0 0 0
6 0
6 1 0 7 7
60
6 m 0 0 0 0 7 7
6 0
6 0 1 7 7
60
6 0 m 0 0 0 7 7
6 0
6 D z j Dy 7
j7
60
6 0 0 Ix Ixy Ixz 7
7
4 D zj 0 D xj 5 40 0 0 Iyx Iy Iyz 5
D yj D xj 0 0 0 0 Izx Izy Iz
2 3 2 3
5. Space frame 1 0 0 0 0 0 m 0 0 0 0 0
6 0
6 1 0 0 0 077
60
6 m 0 0 0 0 7 7
6 0
6 0 1 0 0 077
60
6 0 m 0 0 0 7 7
6 0
6 D zj Dyj 1 0 077
60
6 0 0 Ix Ixy Ixz 7
7
4 D zj 0 D xj 0 1 05 40 0 0 Iyx Iy Iyz 5
D yj D xj 0 0 0 1 0 0 0 Izx Izy Iz
@CivilMethod
The finite element method Chapter 11 511
J for a plane body in a plane truss has three displacements, two translations in
the x and y directions and a rotation about the z axis. However, a joint j of the
truss has only two displacements, the translations in the x and y directions.
Therefore, we keep only the first, second, and sixth rows and the first and sec-
ond columns in the 6 6 matrix. Thus, a 3 2 transformation matrix results for
the plane truss. It should be noted that the 3 3 matrix pertaining to the grid
requires not only the deletion of rows and columns but also their rearrangement.
The inertia matrix Mc results from the inertia properties of the rigid body mov-
ing within the structure. It can also be found from that valid for the space frame
(Type 5 in Table) by deleting appropriate rows and columns.
Example 11.11.1 Formulate the equation of motion of the truss of Fig. E11.12.
Coordinates of nodes: 1ð0, 0Þ, 2ð4, 0Þ, 3ð0, 3Þ, 4ð4, 3Þ, 5ð8, 3Þ; cross-section of
bars: A1 ¼ A2 ¼ 1:5A, A3 ¼ A5 ¼ A, A4 ¼ A6 ¼ A7 ¼ 2:1A, A ¼ 27cm2 ; damp-
ing coefficient: cs ¼ 0; modulus of elasticity: E ¼ 2:1 108 kN=m2 ; accelera-
tion of gravity g ¼ 9:81m=s2 material density r ¼ 7:55kNm1 s2 =m3 ;
consistent mass assumption. The plane rigid body B has dimensions 3:0 3:0
0:02 m3 and specific weight 24kN=m3 .
Solution
The structure has n ¼ 5 nodes. We take node 1 as the point of reference of the
motion of the rigid body. The free structure has N ¼ 9 degrees of freedom, namely
the translational displacements u1 , u2 and the rotation u3 of the rigid body about
point 1 and the six displacements of nodes 2, 4, 5 (see Fig. E11.12). Thus, it is
@CivilMethod
512 PART II Multi-degree-of-freedom systems
2 3 2 3
0:420 0 0:420 0 0:20 0 0:20 0
6 0 0 0 07 6 07
6
k ¼ k ¼ EA4
4 7 7, k5 ¼ EA6 0 0 0 7
0:420 0 0:420 05 4 0:20 0 0:20 05
0 0 0 0 0 0 0 0
2 3
0:70 0 0:70 0
6 0 0 0 07
k6 ¼ EA6
4 0:70
7
0 0:70 05
0 0 0 0
2 3 2 3
2 0 1 0 1:333 0 0:666 0
6 0 2 0 17 6 0:666 7
m1 ¼ m2 ¼ rA6 7, m3 ¼ rA6 0 1:333 0 7
41 0 2 05 4 0:666 0 1:333 0 5
0 1 0 2 0 0:666 0 1:333
2 3 2 3
3:50 0 1:75 0 1:667 0 0:833 0
6 0 3:50 0 1:75 7 6 0:833 7
m4 ¼ m7 ¼ rA6 7, m5 ¼ rA6 0 1:667 0 7
4 1:75 0 3:50 0 5 4 0:833 0 1:667 0 5
0 1:75 0 3:50 0 0:833 0 1:667
2 3
2:10 0 1:05 0
6 0 2:10 0 1:05 7
m6 ¼ rA6
4 1:05
7
0 2:10 0 5
0 1:05 0 2:10
p1 ðt Þ ¼ p2 ðt Þ ¼ p3 ðt Þ ¼ p4 ðt Þ ¼ p5 ðt Þ ¼ p6 ðt Þ ¼ p7 ðt Þ ¼ 0
Computation of Re
2 3 2 3
1 0 0 0 0:8 0:6 0 0
60 1 0 07 6 0:6 0:8 0 0 7
R ¼R ¼R ¼6
1 2 3
40
7, R ¼ R ¼ 6
4 7 7
0 1 05 4 0 0 0:8 0:6 5
0 0 0 1 0 0 0:6 0:8
2 3 2 3
0:8 0:6 0 0 0 1 0 0
6 0:6 0:8 0 0 7 6 1 0 0 07
R ¼6
5
4 0
7, R ¼ 6
6 7
0 0:8 0:6 5 4 0 0 0 15
0 0 0:6 0:8 0 0 1 0
Computation of k , m
e
e , pe ðt Þ
2 3 2 3
0:375 0 0:375 0 0:250 0 0:250 0
6 0 0 0 07 6 0 0 0 07
k ¼ k ¼ EA6
1 2
4 0:375
7, k ¼ EA6
3 7
0 0:375 05 4 0:250 0 0:250 05
0 0 0 0 0 0 0 0
@CivilMethod
The finite element method Chapter 11 513
2 3 2 3
0:268 0:202 0:268 0:202 0 0 0 0
6 0:202 0:151 0:202 0:151 7 6 0 0:7 0 0:7 7
6
k ¼ k ¼ EA4
4 7 7 6
, k ¼ EA4
6 7
0:268 0:202 0:268 0:202 5 0 0 0 0 5
0:202 0:151 0:202 0:151 0 0:7 0 0:7
2 3
0:128 0:096 0:128 0:096
6 0:096 0:072 0:096 0:072 7
k5 ¼ EA64 0:128
7
0:096 0:128 0:096 5
0:096 0:072 0:096 0:072
2 3 2 3
2 0 1 0 1:333 0 0:666 0
60 2 0 17 60 1:333 0 0:666 7
m ¼m
1
¼ rA6
2 7 ¼ rA6
4 1 0 2 0 5, m
3
4 0:666 0
7
5
1:333 0
0 1 0 2 0 0:666 0 1:333
2 3 2 3
3:50 0 1:75 0 1:667 0 0:833 0
6 0 3:50 0 1:75 7 6 0 1:667 0 0:833 7
7 ¼ rA6
4 ¼ m
m 4 1:75 0
7, m 5 ¼ rA6 7
3:50 0 5 4 0:833 0 1:667 0 5
0 1:75 0 3:50 0 0:833 0 1:667
2 3
2:10 0 1:05 0
60 2:10 0 1:05 7
¼ rA6
m6
4 1:05
7
0 2:10 0 5
0 1:05 0 2:10
Computation of Te
Because point 1 has been chosen as the point of reference, only the elements 2
and 5 are affected by the motion of the rigid body. Referring to Table 11.11.1
and Eqs. (11.11.24), (11.11.25), we obtain 2 3
2 3 1 0 0 0
1 0 6 0 1 0 07
6 7
Element 2. D x 3 ¼ 0, D y 3 ¼ 3, T2j ¼ 4 0 1 5 and T2 ¼ 66 3 0 0 0 7
7
3 0 4 0 0 1 05
0 0 0 1
2 3
0:375 0 1:125 0:375 0
6 0 0 0 0 07
02 6 7
k ¼ EA6 6 1:125 0 3:375 1:125 077
4 0:375 0 1:125 0:375 0 5
0 0 0 0 0
@CivilMethod
514 PART II Multi-degree-of-freedom systems
2 3
2 3 1 0 0 0
1 0 60 1 0 07
6 7
Element 5. D x 3 ¼ 0, T5k ¼ 4 0 1 5 and T5 ¼ 6
y 3 ¼ 3, D 60 0 1 077
3 0 40 0 0 15
0 0 3 0
2 3
0:128 0:096 0:128 0:096 0:384
6 0:096 0:072 0:096 0:072 0:288 7
05 6 7
k ¼ EA6 7
6 0:128 0:096 0:128 0:096 0:384 7
4 0:096 0:072 0:096 0:072 0:288 5
0:384 0:288 0:384 0:288 1:152
0e
For the remaining elements e ¼ 1, 3, 4, 6, 7 it is k ¼ k , m
e
0e ¼ m
e , p0 e ¼ pe .
Computation of the mass matrix M1
It is convenient to express the area AB and the density rB of the plane
AB ¼ 3333:33A, rB ¼ 0:324r. This yields
body in terms of A, r. Thuswe obtain
m ¼ 21:6rA, Ic ¼ 21:6rA 32 + 32 =12 ¼ 32:4rA. Then from Table 11.11.1,
we have
2 3 2 3 2 3
21:6 0 0 1 0 0 1 0 0
Mc ¼ rA4 0 21:6 0 5 Tc ¼ 4 0 1 05 ¼ 4 0 1 05
0 0 32:4 Dy c Dxc 1 1:5 1:5 1
and by virtue of Eq. (11.11.38), we obtain
2 3
21:6 0 32:4
M1 ¼ rA4 0 21:6 32:4 5
32:4 32:4 128:9
Finally, the load P ðt Þ is transferred from point A to 1
2 3 8 9
1 0
< 0=
0
P1 ðt Þ ¼ 4 0 1 5 ¼ 1 P ðt Þ
P ðt Þ : ;
3 3 3
@CivilMethod
The finite element method Chapter 11 515
0 0 0 0 0 0 0 0 1
The assembly matrix for the plane rigid body is
2 3
1 0 0 0 0 0 0 0 0
ap ¼ 4 0 1 0 0 0 0 0 0 05
0 0 1 0 0 0 0 0 0
Note that the dimensions of a2 and a5 are 5 9 while the dimensions of ap
are 39.
The enlarged element matrices and load vectors are computed using
Eqs. (11.2.92), (11.2.95), (11.2.96), namely
^ 0 ¼ ðae ÞT m ^ 0 ¼ ðae ÞT k0 e ae , p
e e
M 0 ae , K
e e e ^ 1 ¼ ðap ÞT M1 ap
^0 ðt Þ ¼ ðae ÞT p0 ðt Þ, M
The total stiffness matrix K and load vector pðt Þ are computed using
Eqs. (11.2.100), (11.2.107), respectively, namely
X
Ne X
Ne
^ 0 , pðt Þ ¼ P
e
¼
K K ðt Þ + ^ 0 ðt Þ
p
e
e¼1 e¼1
@CivilMethod
516 PART II Multi-degree-of-freedom systems
M u€f + K
e ff e e ff e pf∗ ðt Þ
uf ¼ e
where
2 3
30:77 43:40 2:75 0 0 0
6 43:40 162:60 3:00 0 0 0 7
6 7
6 3:00 8:93 0 0 7
e ff ¼ rA6 2:75
M
0:67 7:
6 0 0 0 8:93 0 0:67 7
6 7
4 0 0 0:67 0 4:83 0 5
0 0 0 0:67 0 4:83
2 3
1:15 1:51 0:64 0:20 0 0
6 1:51 4:53 0:12 0 0 0 7
6 7
6 0:64 0:12 0:89 0:20 0:25 0 7
e ff ¼ EA6
K 7,
6 0:20 0 0:20 0:85 0 0 7
6 7
4 0 0 0:25 0 0:52 0:20 5
0 0 0 0 0:20 0:15
pf∗ ðt Þ ¼ P ðt Þf 0 3 0 1 0 5 gT , e
e uf ðt Þ ¼ f u3 u4 u6 u7 u8 u9 gT
Remark. The solution of the equation of motion gives the components of the
displacement euf ðt Þ due to the dynamic loads. However, the weight Wc ¼ 3 3
0:02 24 ¼ 4:32kN of the rigid body produces a static deformation, which
should be added to the dynamic one in order to obtain the total deformation. This
is achieved as follows.
@CivilMethod
The finite element method Chapter 11 517
11.12 Problems
Problem P11.1 A plane truss element of length L has variable cross-section
A ¼ A0 1 0:05x2 , x ¼ x=L. The element is loaded by the distributed axial load
t and the suddenly applied concentrated load P ¼ P0 at
pðx, t Þ ¼ p0 ð1 + xÞ sin w
the cross-section x ¼ 0:25L. Compute: (a) the matrices ke , me and the vector pe .
For the mass matrix consider (i) consistent mass assumption and (ii) lumped
mass assumption. Compare ke , me with those resulting when the exact shape
functions are employed. (b) Transform ke , me , pe to the
global
pffiffiffi system of axes,
if the coordinates of element end points j and k are 1 + 3 L=2, 3L=2 and
ðL=2, LÞ, respectively.
Problem P11.2 The height of the cross-section Aðx Þ ¼ bh ðx Þ of a plane frame
element of length L varies linearly so that h ðLÞ ¼ 0:5h ð0Þ. The element is
t, the distributed bend-
py ðx, t Þ ¼ p0 ð1 x Þsin w
loaded by the distributed load
ing moment mz ðx, t Þ ¼ m0 1 x + x 2 cos w t as well as by the concentrated
load Py ðL=3Þ and the concentrated moment Mz ð3L=4Þ, which are applied sud-
denly. Compute the matrices ke , me and the vector pe ðt Þ and compare them with
those resulting when the exact shape functions are employed. Consider the
lumped mass assumption for the mass matrix.
Problem P11.3 Considering only flexural vibrations, formulate the equation of
motion of the frame of Fig. P11.3. Use lumped mass assumption for the
element mass.
@CivilMethod
518 PART II Multi-degree-of-freedom systems
Problem P11.5 Find the number of the active axial constraints of the frame of
Fig. P11.5 and determine its degrees of freedom after their reduction due to
static condensation and axial constraints. The coordinates of the nodes are:
Að0, 0Þ, B ð7, 0Þ, 1ð0, 3Þ, 2ð0, 8Þ, 3ð4, 8Þ, 4ð6, 8Þ.
@CivilMethod
The finite element method Chapter 11 519
Problem P11.7 Neglecting the axial deformation of the beam elements, formu-
late the equation of motion of the frame of Fig. P11.7. The elements have a
length L ¼ 4a, cross-sectional area A, moment of inertia I , modulus of elasticity
E, and line density m. The rigid bodies B1 and B2 have dimensions a a
and 1:5a 0:5a, respectively, and unit thickness. Their material density is r.
Consider lumped mass assumption for the elements.
Problem P11.8 Formulate the equation of motion of the grid of Fig. P11.8.
Assume the data: coordinates of nodes 1ð0, 0Þ, 2ð2, 0Þ, 3ð3, 2Þ, 4ð3, 2Þ;
rectangular cross-section of all elements h b ¼ 0:50 0:30m2 ; material con-
stants E ¼ 2:1 107 , G ¼ 0:4E, r ¼ 2:4kNm1 s2 /m3; loading pðx, t Þ ¼ p0 f ðt Þ,
M ðt Þ ¼ p0 L2 f ðt Þ. Consider lumped mass assumption for the elements.
@CivilMethod
520 PART II Multi-degree-of-freedom systems
Problem P11.10 A rigid circular plate of radius R and thickness 0:1a is supported
on the ground by truss bars as shown in Fig. P11.10. The plate is horizontal
and stands off at a distance h ¼ 0:5a from the ground. The support points on the
ground lie on the vertices of an equilateral triangle. A horizontal force
P ¼ 100kN produced by a rotating mass is applied at the center of the plate.
Formulate the equation of motion of the structure. Assume: a ¼ 20m, R ¼ 4m,
cross-sectional area of the bars A; density of the plate r; modulus of elasticity
E; load of the plate q ¼ 10kN=m2 ; lumped mass assumption for the elements.
@CivilMethod
The finite element method Chapter 11 521
@CivilMethod
522 PART II Multi-degree-of-freedom systems
[7] W. Weaver Jr., P.R. Johnston, Structural Dynamics by Finite Elements, Prentice Hall,
Englewood Cliffs, NJ, 1987.
[8] W. Weaver Jr., M. Eisenberger, Dynamics of frames with axial constraints, ASCE J. Struct.
Eng. 109 (3) (1983) 773–784.
[9] S. Lipschutz, M. Lipson, Linear Algebra (Fourth Edition), Schaum’s Outline Series, McGraw-
Hill Companies, Inc., New York, 2009
[10] V.Z. Vlasov, Thin-walled Elastic Beams, second ed., Israel Program for Scientific Transla-
tions, Jerusalem, Israel, 1961.
[11] E.J. Sapountzakis, V.G. Mokos, Warping shear stresses in nonuniform torsion, Comput.
Methods Appl. Mech. Eng. 192 (2003) 4337–4353.
[12] S. Timoshenko, G.N. Goodier, Theory of Elasticity, second ed., McGraw-Hill, NY, 1951.
[13] R. Debasish, G.V. Rao, Elements of Structural Dynamics: A New Perspective (1st Edition),
John Wiley & Sons, 2012.
@CivilMethod
Chapter 12
Multi-degree-of-freedom
systems: Free vibrations
Chapter outline
12.1 Introduction 523 12.10 Solution of the vibration
12.2 Free vibrations without problem with damping 570
damping 524 12.10.1 Direct solution of
12.3 Orthogonality of eigenmodes 532 the differential
12.4 Eigenmodes of systems with equation 571
multiple eigenfrequencies 534 12.10.2 Linearization of
12.5 The linear eigenvalue the quadratic
problem 541 eigenvalue problem 575
12.5.1 The standard 12.10.3 The use of a
eigenvalue problem proportional viscous
of linear algebra 541 damping matrix 578
12.5.2 Properties of the 12.11 Construction of a
eigenvalues and proportional damping
eigenvectors 544 matrix 584
12.5.3 The generalized 12.11.1 Rayleigh damping 584
eigenvalue problem 552 12.11.2 Additional
12.6 The Rayleigh quotient 558 orthogonality
12.7 Properties of conditions: Caughey
eigenfrequencies and damping matrix 587
modes of MDOF systems 12.11.3 Construction of
without damping: theo proportional
A summary 563 damping matrix
12.8 Solution of the vibration using the modal
problem without damping 563 matrix 590
12.9 The method of mode 12.12 Problems 597
superposition 565 References and further reading 600
12.1 Introduction
In this chapter, we study the dynamic response of multi-degree-of-freedom
(MDOF) systems in the absence of external forces. Like the SDOF systems,
the MDOF systems perform oscillations caused by initial displacements and/
or initial velocities. In accordance with the SDOF systems, we call these
oscillations free vibrations of MDOF systems. In this case, it is pðt Þ ¼ 0 and the
equation of motion (see Eq. 10.2.6) becomes
u + Cu_ + Ku ¼0
M€ (12.1.1)
where M is the mass matrix, C the damping matrix, K the stiffness matrix, and
u ¼ uðt Þ the displacement vector. When damping is neglected, C ¼ 0, we talk
about free vibrations without damping or free undamped vibrations. Their study
plays an important role in the analysis of the MDOF systems because, in addi-
tion to facilitating the understanding of their dynamic response by introducing
the physical concepts of the eigenfrequencies, periods, and mode shapes, it also
serves to develop methods for solving free vibration problems with damping as
well as forced vibration problems.
As mentioned, the equation describing the free undamped vibrations results
from Eq. (12.1.1) by setting C ¼ 0. Thus, we have
M€
u + Ku ¼0 (12.1.2)
The solutions of Eqs. (12.1.1) and (12.1.2) under the specified initial con-
ditions uð0Þ and u_ ð0Þ are the subject that will be discussed in this chapter.
All presented material is illustrated by appropriately chosen examples. The per-
tinent bibliography with recommended references for further study is also
included. The chapter is enriched with problems to be solved.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 525
T€ ðt Þ
¼l (12.2.6)
T ðt Þ
or
T€ ðt Þ + lT ðt Þ ¼ 0 (12.2.7)
The constant l is unknown in the first instance and will be determined
subsequently.
Eq. (12.2.7) allows us to determine the time-dependent function T ðt Þ. To
this end, we look for solutions representing oscillatory motion. We distinguish
three cases:
(i) l ¼ 0. In this case, Eq. (12.2.7) becomes
T€ ðt Þ ¼ 0 (12.2.8)
which is integrated to yield
T ðt Þ ¼ C 1 t + C 2 (12.2.9)
where C1 , C2 are arbitrary integration constants. Obviously, the function
T ðt Þ obtained from Eq. (12.2.9) implies that Eq. (12.2.2) does not repre-
sent an oscillatory motion. Therefore, the value l ¼ 0 must be excluded.
(ii) l ¼ w2 < 0. In this case, Eq. (12.2.7) becomes
T€ ðt Þ w2 T ðt Þ ¼ 0 (12.2.10)
whose solution is (see Chapter 2)
T ðt Þ ¼ C1 ewt + C2 ewt (12.2.11)
From the foregoing expression of T ðt Þ, it follows that Eq. (12.2.2)
does not represent an oscillatory motion. Consequently, the negative
values of the constant l must also be excluded.
(iii) l ¼ w2 > 0. In this case, Eq. (12.2.7) becomes
T€ ðt Þ + w2 T ðt Þ ¼ 0 (12.2.12)
whose solution is (see Chapter 2)
T ðt Þ ¼ c cos wt + d sin wt
(12.2.13)
¼ a cos ðwt qÞ
where c,d or a,q are arbitrary integration constants.
From the foregoing expression of T ðt Þ, it follows that the displacements
obtained by Eq. (12.2.2) are bounded and the function T ðt Þ expresses a har-
monic oscillation with the natural frequency w.
Eq. (12.2.12) gives T€ ðt Þ ¼ w2 T ðt Þ, which is substituted into Eq. (12.2.3)
to yield
K w2 M bT ðt Þ ¼ 0 (12.2.14)
@CivilMethod
526 PART II Multi-degree-of-freedom systems
Because the foregoing equation must hold for all values of t > 0, it must be
K w2 M b ¼0 (12.2.15)
Eq. (12.2.15) represents a system of N linear algebraic equations, which can
be solved to determine the vector b ¼ fb1 , b 2 , …, bN gT , that is, the amplitude
of the displacements. The system of Eq. (12.2.15) is homogeneous; therefore it
has a nontrivial solution only if the determinant of the coefficient matrix van-
ishes, namely
det K w2 M ¼0 (12.2.16)
or
k11 w2 m11 k12 w2 m12 ⋯ k1N w2 m1N
k21 w2 m21 k22 w2 m22 ⋯ k2N w2 m2N
⋯ ⋯ ⋯ ⋯ ¼0 (12.2.17)
k N 1 w2 m N 1 k N 2 w2 m N 2 ⋯ kNN w2 mNN
The expansion of the determinant in Eq. (12.2.17) produces a polynomial of
the N degree with respect to w2 . Therefore Eq. (12.2.16) may hold true for its N
values w2 . These values are the eigenfrequencies or natural frequencies of the
system. Eq. (12.2.16), from which the eigenfrequencies are determined, is
referred to as the frequency equation of the system. As we will show below,
the roots of Eq. (12.2.16) are real and positive for the problems of dynamics
we examine. The eigenfrequencies are arranged in the order of magnitude,
w1 < w2 … < wN , the smallest of which is called the fundamental eigenfre-
quency. The eigenfrequencies may be all distinct or some of them may be mul-
tiple. First, we assume that the system has discrete eigenfrequencies. For each
value wi , we obtain a system of linear algebraic equations of the form (12.2.15),
which allows the determination of the corresponding vector bðiÞ . Due to the fact
that the determinant is equal to zero, the number of independent equations is
ði Þ ði Þ ði Þ
N 1, which implies that one of the components b 1 , b2 ,…,b N of the vector
ði Þ ði Þ
b can be determined arbitrarily. In this respect, we may take b 1 ¼ 1, and
write the system of Eq. (12.2.15) as
ð12:2:18Þ
ði Þ ði Þ
Partitioning the matrix A ¼ K w2i M
and the vector b as indicated in
Eq. (12.2.18) and defining the matrices
ði Þ
A11 ¼ k11 w2i m11 11 (12.2.19a)
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 527
ði Þ
A12 ¼ k12 w2i m12 ⋯ k1N w2i m1N 1ðN 1Þ (12.2.19b)
2 3
k21 w2 m21
A21 ¼ 4
ði Þ 5
⋮ (12.2.19c)
kn1 wi mN 1 ðN 1Þ1
2
2 3
k22 w2i m22 ⋯ k2N w2i m2N
A22 ¼ 4
ði Þ 5
⋮ ⋮ ⋮ (12.2.19d)
kN 2 wi mN 2 ⋯ kNN wi mNN ðN 1ÞðN 1Þ
2 2
8 ðiÞ 9
< b2 =
ði Þ
b2 ¼ ⋮ (12.2.19e)
: ðiÞ ;
bN ðN 1Þ1
@CivilMethod
528 PART II Multi-degree-of-freedom systems
or
X
N qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u¼ bi ai cos ðwi t qi Þ, ai ¼ ci2 + di2 , qi ¼ tan 1 ðdi =ci Þ (12.2.26)
i¼1
The 2N arbitrary constants ci ,di or ai ,qi are established from the initial con-
ditions using the method we will describe after discussing the properties of the
eigenvectors.
From Eq. (12.2.24), we conclude that the deformation pattern of the system
due to an eigensolution is given by a certain shape, which is expressed by the
components of the eigenvector bi multiplied each instant by the value of the
corresponding common time function T ðt Þ ¼ ai cos ðwi t qi Þ. The deformation
shape defined by vector bi is called the i-eigenmode of the free vibration.
As previously mentioned, the components of the eigenvector are determined
relative to one of its components, which can be arbitrarily chosen. Therefore,
the absolute magnitude of the eigenvector is not determined, that is, the eigen-
vectors define only directions in the N -dimensional space while their measure
remains undetermined. In view of this fact, it is appropriate to adopt a standard
scale-setting procedure for measuring it. This process is called normalization.
A common method of normalization is the use of the nondimensional
vectors
2 3 2 3
f1i b1i
6 f2i 7 1 6 7
fi ¼ 6 7 6 b2i 7
4 ⋯ 5 ¼ max b 4 ⋯ 5 (12.2.27)
ki
fNi bNi
The vectors fi (i ¼ 1,2…,N ) are called the normal modes or normal
eigenmodes of the vibration of the system. The shape of deformation defined
by the eigenmode fi is also referred to as the i-mode shape or simply the i-mode
of the free vibration.
The N normal modes can then be displayed in a single square matrix, each
column of which is a normal mode, namely
2 3
f11 f12 ⋯ f1N
6 f21 f22 ⋯ f2N 7
F ¼ ½ f1 f2 ⋯ fN ¼ 6 4⋯ ⋯ ⋯ ⋯ 5
7 (12.2.28)
fN 1 fN 2 ⋯ fNN
The matrix F is called the modal matrix.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 529
Hence
bi
fi ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi (12.2.31)
bTi Μbi
qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffi
If it is M ¼ I, then the scalar quantity bTi Ιbi ¼ bTi bi is the magnitude of
bi , that is, the normalized eigenvector fi is a unit vector. The described process
is also called orthonormalization with respect to mass.
The establishment of the natural frequencies wi and mode shapes fi can be
achieved when the flexibility matrix F ¼ K1 is known without inverting it to
obtain the stiffness matrix.
For this purpose, Eq. (12.2.15) is premultiplied by the matrix F to yield
FMw2 FK b ¼0 (12.2.32)
then taking into account that FK ¼ I, Eq. (12.2.32) becomes
1
FM 2 I b ¼0 (12.2.33)
w
which allows the establishment of the inverse of the eigenfrequencies, 1=w2i ,
and the eigenvectors bi . It should be noted that the matrix FM is not in general
symmetric.
Example 12.2.1 Compute the eigenfrequencies and mode shapes of the two-
story shear frame of Fig. E12.1a. Assume data: h1 ¼ 3:5m, h2 ¼ 4:0m,
q1 ¼ 40kN=m, q2 ¼ 50kN=m, and L ¼ 6:0m. The cross-sectional area of the
@CivilMethod
530 PART II Multi-degree-of-freedom systems
columns of the second floor is 25 25cm2 while that of the first floor is
30 30cm2 . The material of the structure is reinforced concrete with a specific
weight g b ¼ 24kN=m3 and modulus of elasticity E ¼ 2:1 107 kN=m2 . The
dead weight of the beams is included in the loads q1 ,q2 . Consider lumped mass
assumption of the columns.
Solution
(a) (b)
FIG. E12.1 Two-story shear frame.
The system has two degrees of freedom. Its dynamic model is shown in
Fig. E12.1b.
(i) Computation of the mass matrix M.
3:5
6:0 40 + 2 0:25 0:25 24
2
m11 ¼ ¼ 25kNm1 s2
9:81
3:5 4:0
6:0 50 + 2 0:25 0:25 + 0:30 0:30 24
2 2
m22 ¼
9:81
¼ 32 kNm1 s2
Hence
25 0
M¼ (1)
0 32
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 531
@CivilMethod
532 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 533
@CivilMethod
534 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 535
@CivilMethod
536 PART II Multi-degree-of-freedom systems
they do not satisfy in general the orthogonality condition between each other.
As we will see in Section 12.8, The fulfillment of the orthogonality condition is
necessary for the solution of the free vibration problem of the MDOF systems.
This can be achieved using the Gram-Schmidt orthogonalization process, which
is described right below
Let x1 ,x2 ,…, xN be a set of N linear independent but not orthogonal vectors.
These vectors constitute the basis of a space with N dimensions. We can readily
show that the vectors defined by the relations
e
x1 ¼ x1
e
xT1 x2
e
x2 ¼ x2 e
x1
x1 j2
je
e
xT1 x3 e
xT2 x3
e
x3 ¼ x3 2
e
x2e
x1 (12.4.7)
je
x1 j x2 j2
je
⋯ ¼ ⋯⋯⋯⋯⋯⋯⋯⋯⋯⋯
e
xT1 xN e
xT2 xN e
xTN 1 xN
e
xN ¼ xN e
x1 e
x2 ⋯ e
xN 1
x1 j2
je x2 j2
je xN 1 j2
je
are orthogonal to each other. From the geometrical point of view, the vector e xk
is equal to xk minus its projection on the space defined by the vectors
e
x1 ,e
x2 ,…,exk1 . For example, the vector ex2 is equal to x2 minus its projection
on the vector ex1 . This is shown in Fig. 12.4.1.
The orthogonality of the eigenvectors with respect to the mass, which is also
needed, is achieved using the following relations.
e
x1 ¼ x1
e
xT1 Mx2
e
x2 ¼ x2 e
x1
e
xT1 Me
x1
e
xT1 Mx3 e
xT2 Mx3
e
x3 ¼ x3 e
x 1 e
x2 (12.4.8)
e
xT1 Me
x1 e
xT2 Me
x2
⋯ ¼ ⋯⋯⋯⋯⋯⋯⋯⋯⋯
e
xT1 MxN e
xT MxN e
xT MxN
e
xN ¼ xN e
x1 2T x2 ⋯ TN 1
e e
xN 1
e
x1 Me
T
x1 e
x2 Me
x2 e
xN 1 Me
xN 1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 537
Solution
(i) Computation of the eigenfrequencies.
The frequency equation is given by the determinant
3826 25w2 1913 7652 + 25w2
det ½AðwÞ ¼ 1913 2285:5 8w 2
6484 16w 2 ¼0 (2)
7652 + 25w2 6484 16w2 21307:08 67w 2
@CivilMethod
538 PART II Multi-degree-of-freedom systems
For n ¼
1,we obtain
ð1Þ 1 ð1Þ
b1 ¼ and Eq. (12.4.5) gives b2 ¼ f0:3553g. Hence
0 8 9
<1 =
b1 ¼ 0 (10)
: ;
0:3553
For n ¼
2,we obtain
ð2Þ 0 ð2Þ
b1 ¼ and Eq. (12.4.5) gives b2 ¼ f0:322365g. Hence
1 8 9
< 0 =
b2 ¼ 1 (11)
: ;
0:322365
For n ¼ 3 the eigenfrequency is discrete, w3 ¼ 19:2358, the rank of the
matrix is 3 1 ¼ 2, and the matrix Að3Þ is partitioned as
ð12Þ
Thus we have
ð3Þ
A11 ¼ ½ 5424:4 (13)
ð3Þ
A12 ¼ ½ 1913 1598:4 (14)
ð3Þ 1913
A21 ¼ (15)
1598:4
ð3Þ 674:6280 563:7440
A22 ¼ (16)
563:7440 3483:9921
and Eq. (12.2.22) gives
h i1
ð3Þ ð3Þ ð3Þ ð3Þ 2:8357
b1 ¼ f1g, b2 ¼ A22 A21 ¼ . Hence
8 90
< 1 =
b3 ¼ 2:8357 (17)
: ;
0
(iii) Orthogonalization of the eigenmodes.
The computed eigenvectors b1 , b2 , and b3 are linearly independent.
This is readily verified by computing the determinant of the eigenvectors,
which apparently does not vanish
1 0 1
det ðBÞ ¼ 0 1 2:8356 ¼ 1:26936 (18)
0:3553 0:322365 0
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 539
However, they do not satisfy the orthogonality condition. They are made
orthogonal using the Gram-Schmidt process described previously. Thus
applying Eq. (12.4.8), we obtain
e1 ¼ b1 ¼ f 1 0 0:3553 gT
b
eT Mb2
b
e
b2 ¼ b2 1T e1 ¼ f 0:3868 1 0:4598 gT
b (19)
e Mb
b e1
1
e3 ¼ b3 ¼ f 1 2:8357 0 gT
b
and after their orthonormalization with respect to the mass matrix, we obtain
2 3
0:2524 0:2551 0:1058
F ¼ 4 0:0000 0:6595 0:3000 5 (20)
0:0897 0:3032 0:0000
Solution
For lumped mass assumption, the dynamic model of the system is shown in
Fig. E12.3b. The system has six degrees of freedom, which are reduced to three,
u1 , u2 ,u3 , after the static condensation of the rotations.
@CivilMethod
540 PART II Multi-degree-of-freedom systems
ð5Þ
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 541
Therefore
ð3Þ ð3Þ ð3Þ 2 ð3Þ 4 2
A11 ¼ ½3, A12 ¼ ½ 2 1 , A21 ¼ , A22 ¼
1 2 3
and Eq. (12.2.22) give
h i1
ð3Þ ð3Þ ð3Þ 1
b2 ¼ A22 A21 ¼ . Hence
1
8 9
< 1=
bð3Þ ¼ 1 (9)
: ;
1
The graphical representations of the eigenmodes are shown in Fig. E12.3c.
We readily find out that the orthogonality condition with respect to the mass
matrix is satisfied.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 543
Solution
The characteristic equation is
2 l 1
5:5 3 l ¼ l + 5l + 0:5 ¼ 0
2
(1)
@CivilMethod
544 PART II Multi-degree-of-freedom systems
which gives
1
x2 ¼ (6)
2:8978
0s ¼ðAxr ÞT x
yTr x 0s
0s
¼ ðlr xr ÞT x (12.5.13a)
0s
¼ lr xTr x
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 545
@CivilMethod
546 PART II Multi-degree-of-freedom systems
5. If the real matrix A is symmetric and positive definite, then its eigenvalues
are positive.
Consider the matrix A and an arbitrary vector x ¼ fx1 , x2 , …, xN gT . The
homogeneous polynomial of the second degree defined by
N X
X N
Pðx1 , x2 , …, xN Þ ¼ aik xi xk (12.5.24)
i¼1 k¼1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 547
or
PðxÞ ¼ xT Ax (12.5.25)
is referred to as the quadratic form of the matrix A. For a real and symmetric
matrix A, the coefficients of the polynomial PðxÞ are real and aik ¼ aki .
The quadratic form PðxÞ is called positive definite, if
T
x Ax > 0, if x 6¼ 0
PðxÞ ¼ (12.5.26)
xT Ax ¼ 0, if x ¼ 0
A matrix A is called positive definite if its quadratic form is positive.
A direct consequence of this is that the diagonal elements of A are positive.
Indeed, if we take xi ¼ 1 and xk ¼ 0, k 6¼ i, then Eq. (12.5.24) gives
Pðx1 , x2 , …, xN Þ ¼ aii > 0 (12.5.27)
Let x be an eigenvector of A. Because A is real and symmetric, the
eigenpair ðl, xÞ is real according to the previous property. To prove that
l > 0 we write Eq. (12.5.25) as
T
xT Ax ¼ AT x x
¼ ðAxÞT x (12.5.28)
¼ lxT x
Because xT x > 0, hence xT Ax > 0, the foregoing relation gives
xT Ax
l¼ >0 (12.5.29)
xT x
6. To each eigenvalue of a symmetric matrix there correspond as many line-
arly independent eigenvectors as the multiplicity of the eigenvalue.
Let A be an N N symmetric matrix. We assume that we have com-
puted the eigenvalue lk and the corresponding eigenvector xk . Note that
the eigenvalue problem has no degenerate eigenvalues because the matrix
is assumed symmetric. We construct now an N N orthonormal matrix Q,
whose first column is the vector xk , namely
Q ¼ xk Q ^ , QT Q ¼ I (12.5.30)
^
Apparently, the dimensions of Q are N ðN 1Þ. The construction of Q
is always possible because the vectors in Q provide a basis of the N dimen-
sional space defined by A. This is understood from Example 12.5.2, which
is presented below.
First we evaluate
" #
xTk
Q AQ ¼
T
A xk Q^
^
Q T
" # (12.5.31)
xTk Axk xTk AQ^
¼
^ T Axk Q
Q ^ T AQ^
@CivilMethod
548 PART II Multi-degree-of-freedom systems
^ T Axk ¼
Then taking into account that Axk ¼ lk xk , xTk xk ¼ 1, Q
^ T T ^ ^ T T
lk Q xk ¼ 0, xk AQ ¼ Q A xk ¼ 0 because it was assumed AT ¼ A,
we write the foregoing relation as
l 0
QT AQ ¼ k ^ (12.5.32)
0 A
where
^ ¼Q
A ^
^ T AQ (12.5.33)
^ ¼ c, that is, a con-
is a fully populated ðN 1Þ ðN 1Þ. If N ¼ 2, then A
stant, and the matrix Q AQ becomes diagonal
T
l 0
QT AQ ¼ k (12.5.34)
0 c
Premultiplying the previous equation by Q gives
lk 0
AQ ¼ Q
0 c
(12.5.35)
lk 0
¼ x Q ^
0 c
Hence Q ^ is another eigenvector of A and c is the other eigenvalue
regardless of whether lk is a multiple or a distinct eigenvalue. For arbitrary
N the proof is achieved by mathematical induction. Taking into account the
previous proof, the Gramm-Schmidt orthogonalization, and property (4), we
deduce that an N N symmetric matrix has N orthonormal eigenvectors.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 549
which is normalized with respect to its magnitude jx2 j ¼ 7:3186 and becomes
xT2 ¼ f 0:6832 0:1366 0:7173 g (4)
The vector xT3 ¼ f x13 x23 x33 g can be computed from the orthogonality
relations
xT1 x3 ¼ 0:3841x13 0:7682x23 + 0:5121x33 ¼ 0 (5)
xT2 x3 ¼ 0:6832x13 0:1366x23 0:7173x33 ¼ 0 (6)
Thus, by determining x13 arbitrarily, say x13 ¼ 1, the foregoing relations
yield the linear system of equations
0:7682x23 + 0:5121x33 ¼ 0:3841 (7)
0:1366x23 0:7173x33 ¼ 0:6832 (8)
from which we obtain x23 ¼ 1:0070, x33 ¼ 0:7606, hence
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 551
2 3
x1 T x1 0 ⋯ 0
6 0 x2 T x2 ⋯ 0 7
XT X ¼ XXT ¼ 6
4⋯
7
5 (12.5.46)
⋯ ⋯ 0
0 0 ⋯ xN xN
T
which,
p ffiffiffiffiffiffiffiffiffiffiffiif the eigenvectors are normalized with respect to their magnitude,
xi T xi ¼ 1, becomes
XT X ¼ XXT ¼ I (12.5.47)
Hence
X1 ¼ XT (12.5.48)
and Eq. (12.5.45) becomes
A ¼ XLXT (12.5.49)
@CivilMethod
552 PART II Multi-degree-of-freedom systems
ðA lBÞx ¼ 0 (12.5.54)
This problem is known as the generalized eigenvalue problem of linear alge-
bra. In the literature, it is also referred to as the linearized eigenvalue problem.
The study of the properties of the eigenvalues and eigenvectors of the general-
ized eigenvalue problem, Eq. (12.5.54), is facilitated if it is transformed to the
standard eigenvalue problem, Eq. (12.5.6). Thus, the properties that apply to the
standard eigenvalue problem can be transferred to the generalized eigenvalue
problem. Without excluding the generality, the discussion will be restricted
to real, symmetric, and positive definite matrices A and B because in free vibra-
tions, they represent the stiffness and mass matrices, that is, A ¼ K and B ¼ M.
Their positive definiteness results from the fact that the elastic energy U ðuÞ and
the kinetic energy T ðu_ Þ are expressed by positive definite quadratic forms,
that is,
8
> 1
< uT Ku > 0, if u 6¼ 0
U ðuÞ ¼ 2 (12.5.55)
>
: 1 uT Ku ¼ 0, if u ¼ 0
2
8
> 1
< u_ T Mu_ > 0, if u_ 6¼ 0
T ðu_ Þ ¼ 2 (12.5.56)
>
: 1 u_ T Mu_ ¼ 0, if u_ ¼ 0
2
Applying Eq. (12.5.49) to matrix B we have
eL
B¼X eX
eT (12.5.57)
where L e is the diagonal matrix of the eigenvalues of B and X e the matrix
of its eigenvectors. Obviously, when the matrix B is diagonal, as in the case
of concentrated masses, then Xe ¼ I.
Further, we can set
B ¼ SST (12.5.58)
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 553
where
pffiffiffiffi
e
S¼X Le (12.5.59)
e are positive because B
The matrix S is real on account that the elements of L
was assumed positive definite.
Substituting Eq. (12.5.59) into Eq. (12.5.54), yields
Ax ¼ lSST x (12.5.60)
We define now the vector
^ ¼ ST x
x (12.5.61)
thenb
x ¼ ST x
^ (12.5.62)
1
Premultiplying Eq. (12.5.60) by S and using Eq. (12.5.62) yield
^ lI x
A ^¼0 (12.5.63)
where
^ ¼ S1 AST
A (12.5.64)
Obviously, the matrix A ^ is real and symmetric. Therefore, according to
property 3, its eigenvalues and eigenvectors x^ are real. Moreover, taking into
^ satisfy the orthogonality condition, x
account that the vectors x ^Ti x
^j ¼ 0, i 6¼ j,
we obtain.
T
^Ti x
x ^j ¼ ST xi ST xj
¼xTi SST xj , i 6¼ j
(12.5.65)
¼xTi Bxj
¼0
which implies that the eigenvectors of the generalized eigenvalue problem are
orthogonal with respect to the matrix B.
The eigenvalue problems (12.5.54) and (12.5.63) have the same eigen-
values. Indeed, we can set I ¼ S1 SST ST and write Eq. (12.5.63) by virtue
of Eqs. (12.5.64) and (12.5.58) as
1
S ðA lBÞST x ^¼0 (12.5.66)
If PðlÞ and P ^ ðlÞ are the characteristic polynomials of the eigenvalue
problems (12.5.54) and (12.5.63), respectively, we obtain
1
b. The notation ST ¼ ST is employed.
@CivilMethod
554 PART II Multi-degree-of-freedom systems
^ ðlÞ ¼ det A
P ^ lI
¼ det S1 ðA lBÞST
(12.5.67)
¼ det S1 det ðA lBÞdet ST
¼ det S1 PðlÞ det ST
Because det S1 6¼ 0 and det ST 6¼ 0, it implies that both eigenvalue
problems have the same characteristic equation, hence the same
eigenvalues.
The spectral decomposition of matrix B requires the complete solution of
the eigenvalue problem. Therefore, the transformation of the generalized eigen-
value problem on the basis of Eq. (12.5.59) is not the most convenient one.
A usual method to determine the matrix S is the Cholesky decomposition
method, or the square root method, in which the matrix B is written in the form
of a product, that is,
B ¼ UT U (12.5.68)
where U is an upper triangular matrix. Hence
S ¼ UT (12.5.69)
9. If the matrices A and B real, symmetric, and positive definite, then the
generalized eigenvalue problem has positive eigenvalues.
This is readily proved by premultiplying Eq. (12.5.54) by xT . This
gives
xT Ax ¼ lxT Bx (12.5.70)
Because A and B are positive definite, we obtain
xT Ax
l¼ >0 (12.5.71)
xT Bx
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 555
l e2 ¼ 3:909
e1 ¼ 1:790, l (2)
Hence
e 1:790 0
L¼ (3)
0 3:909
The eigenvectors are computed from the solution of the homogeneous linear
system
e 1
2:5 l x1 0
1 e x2 ¼ 0
3:2 l
(4)
e1 ¼ 1:790 and l
for l e2 ¼ 3:909.
Thus, we obtain the matrix of the eigenvectors normalized with respect to
their magnitude
" #
0:8157 0:5784
e¼
X (5)
0:5785 0:8158
Using Eq. (12.5.59) we obtain
pffiffiffiffi
e e 1:0913 1:1435 1 0:6097 0:4323
S¼X L¼ , S ¼ (6)
0:7740 1:6129 0:2926 0:4126
and on the basis of Eq. (12.5.64)
^ ¼ S1 AST ¼ 14:3213 2:1288
A (7)
2:1288 29:2564
11. If the real and symmetric matrix A is singular, then the generalized eigen-
value problem has at least one zero eigenvalue and the corresponding
eigenvector is different from zero.
@CivilMethod
556 PART II Multi-degree-of-freedom systems
First, we will show that this property holds for the standard eigenvalue
problem Ax ¼ lx. For this purpose, we write A in the form of its spectral
decomposition,
A ¼ XLXT (12.5.74)
Because it was assumed that A is
singular,
it implies that det ðAÞ ¼ 0.
Moreover, Eq. (12.5.47) gives det XT det ðXÞ ¼ 1, hence
det ðAÞ ¼ det XT det ðLÞdet ðXÞ
¼ det ðLÞ
(12.5.75)
¼l1 l2 ⋯lN
¼0
from which we conclude that at least one of the li is zero and the eigen-
value problem for this value becomes
Axi ¼ 0 (12.5.76)
which yields xi 6¼ 0 because det ðAÞ ¼ 0.
The generalized eigenvalue problem Ax ¼ lBx is transformed to the
standard eigenvalue problem A^^ x ¼ l^
x, hence it is
det A^ ¼ l1 l2 ⋯lN (12.5.77)
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 557
12. If the real and symmetric matrix B is singular, then the generalized eigen-
value problem has at least one infinite eigenvalue.
This is shown if the eigenvalue problem is written in the form
Bx ¼ mAx (12.5.81)
where m ¼ 1=l.
13. Any vector u with a dimension N can be represented as the superposition
of the eigenvectors of the eigenvalue problem.
We showed that the set of the eigenvectors x1 ,x2 ,…, xN of an N N
symmetric matrix is linearly independent and can be employed as a base of
the N dimensional space to represent an arbitrary vector u in that space.
Thus we may set
u ¼ a1 x1 + a2 x2 + ⋯ + aN xN (12.5.82)
or
u ¼ Xa (12.5.83)
where X is the matrix of the eigenvectors and a ¼ fa1 , a2 , …, aN gT the
vector of the coefficients. The matrix X is not singular because the eigen-
vectors xi are linearly independent. Hence
a ¼ X1 u (12.5.84)
or using Eq. (12.5.48)
a ¼ XT u (12.5.85)
or
ai ¼ xi T u, i ¼ 1, 2, …,N (12.5.86)
Similarly, we can use the eigenvectors of the generalized eigenvalue
problem to represent the vector u. The establishment of the coefficients
ai in that case is established by premultiplying both sides of
Eq. (12.5.82) by xTi B and noting that xTi Bxj ¼ 0 for i 6¼ j. This yields
xTi Bu
ai ¼ (12.5.87)
xTi Bxi
If the eigenvectors xi are normalized with respect to B so that
xTi Bxi ¼ 1, then Eq. (12.5.87) gives
ai ¼ xTi Bu, i ¼ 1, 2, …,N (12.5.88)
or
a ¼ XT Bu (12.5.89)
The representation of a vector as a superposition of the eigenvectors as
in Eq. (12.5.82) is known as the expansion theorem. As we will see in
@CivilMethod
558 PART II Multi-degree-of-freedom systems
l1 rðuÞ lN (12.6.2)
and it is rðuÞ > 0, if A is positive definite and rðuÞ 0 if A is positive
semidefinite. Moreover, it is rðuÞ ¼ l1 if u ¼ x1 and rðuÞ ¼ lN if u ¼ xN .
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 559
@CivilMethod
560 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 561
X
N
S¼ aj2 (12.6.17)
j ¼1
j 6¼ k
B X C
B N2B
C
¼lk + e B lj aj lk S C
2
C + higher order terms
@j ¼1 A
j 6¼ k
uT Au
rðuÞ ¼ (12.6.21)
uT Bu
@CivilMethod
562 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 563
1
rðx2 Þ ¼ l2 ¼ w22 ¼ 370:0416, w2 ¼ 19:236, u2 ¼ (8)
1:4127
Normalizing the eigenmodes with respect to the mass, we obtain
0:1697 0:1061
F¼ (9)
0:0935 0:1499
Obviously, the obtained eigenfrequencies and mode shapes are identical to
those obtained by the method in Section 12.2.
@CivilMethod
564 PART II Multi-degree-of-freedom systems
X
N
uðt Þ ¼ fi ðci cos wi t + di sin wi t Þ (12.8.1)
i¼1
or
X
N
uðt Þ ¼ fi ai cos ðwi t qi Þ (12.8.2)
i¼1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 565
if the eigenmodes have been orthonormalized with respect to the mass, that is,
fTi Mfi ¼ 1 (see Eq. 12.2.29).
The same procedure is used to evaluate the coefficients di . Differentiating
Eq. (12.8.1) with respect to time yields
X
N
u_ ðt Þ ¼ fi wi ðci sin wi t + di cos wi t Þ (12.8.10)
i¼1
and for t ¼ 0
u_ ð0Þ ¼ f1 w1 d1 + f2 w2 d2 + ⋯ + fN wN dN (12.8.11)
Premultiplying the foregoing equation by fTi M yields
1 fTi Mu_ ð0Þ
di ¼ (12.8.12)
wi fTi Mfi
and if fTi Mfi ¼ 1
1 T
di ¼ f Mu_ ð0Þ (12.8.13)
wi i
The constants ai ,qi are evaluated from the relations
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ai ¼ ci2 + di2 , qi ¼ tan 1 ðdi =ci Þ (12.8.14)
@CivilMethod
566 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 567
The foregoing relations are further simplified if the eigenmodes are normal-
ized on the basis of Eq. (12.2.29). This gives
Mn ¼ fTn Μfn ¼ 1 (12.9.14)
Kn ¼ w2n (12.9.15)
that is, the modal masses become equal to the unity and the modal stiffnesses
equal to the square of the corresponding eigenfrequency. Thus, we have
Yn ð0Þ ¼ fTn Muð0Þ, n ¼ 1, 2, …,N (12.9.16)
Y_ n ð0Þ ¼ fTn Mu_ ð0Þ, n ¼ 1, 2, …,N (12.9.17)
The vector un ¼ fn Yn , n ¼ 1, 2, …,N expresses the contribution of the nth
eigenmode to the displacement vector u and is referred to as the nth modal
component of the displacement vector. Hence, Eq. (12.9.1) is written as
u ¼ u 1 + u2 + ⋯ + u N (12.9.18)
which implies that the displacement vector is the superposition of the modal
components or, in other words, the displacement vector is the superposition
of the eigenmodes multiplied by a time-dependent weight coefficient.
This method of solving the equation of motion for the free vibration problem
is known as the modal superposition method or the method of the superposition
of the eigenmodes. It is used not only for free undamped vibrations but also for
damped vibrations as well as for forced vibrations. The superposition is
illustrated in Fig. 12.9.1, where the mode shapes correspond to those of the
two-story shear frame in Example 12.2.1.
@CivilMethod
568 PART II Multi-degree-of-freedom systems
w1 ¼ 8:289, w2 ¼ 19:236
0:1697 0:1061
f1 ¼ , f2 ¼ (2)
0:0935 0:1499
Because the eigenmodes are orthonormalized with respect to the mass, the
modal matrices resulting from Eq. (12.9.14) are
M1 ¼ fT1 Μf1 ¼ 1, M2 ¼ fT2 Μf2 ¼ 1 (3)
Applying Eqs. (12.9.12) and (12.9.13) gives
25 0 0:03
Y1 ð0Þ ¼ ½ 0:1697 0:0935 ¼ 0:15720 (4)
0 32 0:01
25 0 0:03
Y2 ð0Þ ¼ ½ 0:1061 0:1499 ¼ 0:0316 (5)
0 32 0:01
Y_ 1 ð0Þ ¼ Y_ 2 ð0Þ ¼ 0 (6)
Thus Eq. (12.9.9) gives
Y1 ¼ 0:15720cos 8:289t (7a)
Y2 ¼ 0:0316cos 19:236t (7b)
Then Eq. (12.9.1) gives the displacement vector
u 1 ðt Þ 2:6677 0:3353
¼ 102 cos 8:289t + 102 cos 19:236t
u 2 ðt Þ 1:4698 0:4700
(8)
Example 12.9.2 The loads P1 and P2 applied to the structure in Fig.E12.4a are
removed suddenly at t ¼ 0. Determine the ratio P1 =P2 so that the structure
vibrates in the form of the second eigenmode. The axial deformation of the col-
umn is neglected. Consider lumped mass assumption. Assume a ¼ 1, m ¼ 1,
E ¼ 2:1 107 kN=m2 , and column cross-section 0:20 0:20 m2 .
(a) (b)
FIG. E12.4 Structure in Example 12.9.2 (a); Parameters of motion (b).
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 569
Solution
The system has two degrees of freedom, the horizontal displacement u and the
rotation f at the top of the column shown in Fig. E12.4b. The equation of
motion can be obtained either by applying Eqs. (A.4.8) or by considering the
structure as a plane frame consisting of one element and including a rigid body
at its free end (see Section 11.11). Here, Eqs. (A.4.8) are employed. Thus,
setting yc ¼ 0, XP ¼ u, YP ¼ 0, w_ ¼ f,€ Fx ¼ fS , Fy ¼ 0, MP ¼ MS , IP ¼ IO ,
we obtain
m u€ + fS ¼ 0 (1)
Fy ¼ mx c f€ (2)
IO f€ + MS ¼ 0 (3)
The elastic forces are applied at point O and are given by the relations
fS ¼ k11 u + k12 f
MS ¼ k21 u + k22 f
Eqs. (1) and (3) are combined in matrix form
M€
u + Ku ¼ 0 (4)
where
m 0 k11 k12
M¼ , K¼
0 Io k21 k22
Note that Eq. (2) will be used to evaluate the unknown force Fy once u,f are
established from the solution of Eq. (4).
Stiffness matrix: The stiffness matrix will result from Eq. (11.5.2) for
u3 ¼ u4 ¼ 0. Thus, we have
2 3
12EI 6EI 2 3 3a 3
6 ð4a Þ3 ð4a Þ2 7 EI
6 7 6 16 8 7
K¼6 7¼ 4 5 (5)
4 6EI 4EI 5 a 3 3a 2
a
ð4a Þ2 4a 8
Mass matrix: We have
1
m ¼ ð4a Þm + ð3a Þ10m
¼ 32a m
2
Z 2a
Io ¼ 2 dx ¼ 30ma
10mx 3
a
hence
m 0 32 0
M¼
¼ ma (6)
0 Io 0 30a 2
@CivilMethod
570 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 571
where an are 2N arbitrary constants, which are determined from the initial
conditions.
The response of the system depends on the type of roots ln . We distinguish
the following cases:
(i) Real roots
Let ln be a real root, the corresponding solution (12.10.2) is
un ¼ bn eln t (12.10.8)
Obviously, for ln > 0 the solution diverges exponentially while for
ln < 0 it converges exponentially. In both cases, the motion of the system
is not an oscillation. If the root has a multiplicity k, then the solution
(12.10.2) becomes [3]
@CivilMethod
572 PART II Multi-degree-of-freedom systems
un ¼ bn a0 + a1 t + ⋯ + t k1 eln t (12.10.9)
which again diverges or converges exponentially depending on whether
ln > 0 or ln < 0, respectively, and the motion of the system is not an
oscillation.
(ii) Complex roots.
Let lm be complex, then its complex conjugate l m is also a root.
We set
m ¼ m iwm
lm ¼ mm + iwm and l (12.10.10)
m
The respective vectors bm and b m are also complex conjugate and their
contribution to the general solution is
m e lm t
um ¼am bm elm t + am∗ b
¼emm t am bm eiwm t + am∗ b m eiwm t (12.10.11)
¼emm t am bm + am∗ b m cos wm t + i am bm a ∗ b
m m sin wm t
The constants am , am∗ are arbitrary constants, hence they can be chosen so that
they are complex conjugates, that is
am ¼ cm + id m and am∗ ¼ cm id m (12.10.12)
where cm and dm are also arbitrary constants.
Further, we set
m ¼ pm iqm
bm ¼ pm + iqm and b (12.10.13)
Substituting Eqs. (12.10.12) and (12.10.13) into Eq. (12.10.11) gives
um ¼ 2emm t ½ðcm pm dm qm Þ cos wm t ðcm qm + dm pm Þsin wm t (12.10.14)
The factor within the square bracket expresses a harmonic vibration. Its
amplitude diverges exponentially if mm > 0 and converges exponentially if
mm < 0. Finally, if mm ¼ 0, the roots are imaginary and the amplitude of vibra-
tion remains constant. The vibration with mm > 0 is known as flutter or negative
damping. The solution is obtained as a superposition of the terms given by Eq.
(12.10.14), namely
X
N
u¼ um (12.10.15)
m¼1
Eq. (12.10.14) implies that the natural components of the displacements are
real, and as would be anticipated, the general solution is real.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 573
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rim ¼ ðcm pim dm qim Þ2 + ðcm qim + dm pim Þ2 (12.10.17a)
1 ðcm qim + dm pim Þ
qin ¼ tan (12.10.17b)
cm pim dm qim
It becomes obvious from Eq. (12.10.17b) that the displacement components
in a vibration mode have different phase angles, contrary to the undamped
vibrations where the phase angles are the same within a vibration mode (see
Eq. 12.8.2).
Example 12.10.1 Determine the displacements of the frame in Example 12.2.1
in the presence of damping by direct solution of the equation of motion.
Assume:
30 5
C¼ , uð0Þ ¼ f0:03 0:01gT , u_ ð0Þ ¼ 0
5 40
Solution
The matrices M kai K were computed in Example 12.2.1. Hence, we have
25 0 3826:5 3826:5
M¼ , K¼ (1)
0 32 3826:5 9142:1
@CivilMethod
574 PART II Multi-degree-of-freedom systems
X
4
u_ ð0Þ ¼ an ln bn (8)
n¼1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 575
@CivilMethod
576 PART II Multi-degree-of-freedom systems
where
A ^ 1 K
^ ¼M ^ (12.10.28)
Eq. (12.10.27) gives 2N eigenvalues and the respective eigenvectors bn that
^ and K
satisfy the orthogonality condition with respect to M ^ (see Section 12.5.2,
property 2). The eigenvalues of the problems we encounter are complex with a
negative real part and express convergent vibrations.
^ we will have
If the eigenvectors are orthonormalized with respect to M,
^ n ¼1
fTn Mf (12.10.29)
^ n ¼ ln
fTn Kf (12.10.30)
The general solution of Eq. (12.10.23) is obtained as a superposition of the
solutions of Eq. (12.10.25), namely
X
2N
x¼ an fn eln t (12.10.31)
n¼1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 577
fn ¼ pn + iqn , n ¼ pn iqn
f (12.10.37)
Apparently, for each pair of eigensolutions, we will have
n
xm ¼ xn + x
(12.10.38)
¼ 2emn t ½ðcn pn dn qn Þcos wn t ðcn qn + dn pn Þ sin wn t
and the general solution is obtained as
X
N
x¼ xm (12.10.39)
m¼1
(12.10.42)
Y_ LY ¼ 0 (12.10.44)
@CivilMethod
578 PART II Multi-degree-of-freedom systems
or
Y_ n ln Yn ¼ 0, n ¼ 1, 2, …,2N (12.10.45)
that is, the system of the 2N coupled equations is transformed to 2N
SDOF equations with respect to the modal coordinates. Integration of
Eq. (12.10.45) gives
Yn ¼ an elt , n ¼ 1, 2, …,2N (12.10.46)
Thus the general solution is given by Eq. (12.10.40). The vector
a ¼ fa1 , a2 , …, a2N gT of the arbitrary constants is obtained from the relation
^ ð 0Þ
a ¼ FT Mx (12.10.47)
In summary, the solution of the vibration problem with damping presented
in this section can be obtained by adhering to the following steps:
1. Formulation of the matrices M^ and K
^ defined by Eq. (12.10.24).
2. Computation of the eigenvalues and eigenvectors of the eigenvalue
problem, Eq. (12.10.26).
^
3. Orthonormalization of the eigenvectors with respect to M.
4. Computation of the arbitrary constants using Eq. (12.10.34).
5. Computation of the solution using Eq. (12.10.31).
6. Separation of the vectors u and u_
X
N X
N
u_ ¼ an fðn1Þ eln t , u¼ an fðn2Þ eln t (12.10.48)
n¼1 n¼1
where fðn1Þ and fðn2Þ are the upper and lower half of fn , respectively.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 579
Using Eq. (12.9.1), which represents the transformation from the physical
coordinates to modal coordinates of the undamped vibrations, we may write
Eq. (12.10.1) as
€ + CFY_ + KFY ¼0
MFY (12.10.50)
which is premultiplied by F to yield T
€ + FT CFY_ + FT KFY ¼0
FT MFY (12.10.51)
Taking into account the orthogonality of the eigenmodes with respect to M,
K, and assuming that Eq. (12.10.49) holds, we obtain
Mn Y€ n + Cn Y_ n + Kn Yn ¼ 0, n ¼ 1, 2, …,N (12.10.52)
namely, the uncoupling of equations of motion is achieved when damping is
present.
Further, setting
Cn ¼ 2x n Mn wn (12.10.53)
and taking into account that Kn ¼ Mn w2n Eq. (12.10.52) becomes
Y€ n + 2x n wn Y_ n + w2n Yn ¼ 0, n ¼ 1, 2, …, N (12.10.54)
whose solution is given by Eq. (2.3.18), namely
_
Y n ð0Þ + Yn ð0Þxn wn
Yn ðt Þ ¼ exn wn t sin wDn t + Yn ð0Þcos wDn t (12.10.55)
wDn
The quantity xn defined by Eq. (12.10.53) expresses the damping ratio of the
n-eigenmode. The initial conditions Yn ð0Þ, Y_ n ð0Þ are evaluated from the vec-
tors uð0Þ and u_ ð0Þ using Eqs. (12.9.12) and (12.9.13). After determining
Yn ðt Þ,n ¼ 1, 2, …,N , the solution of Eq. (12.10.1) is obtained from
Eq. (12.9.1), namely
X
N
uðt Þ ¼ f n Yn ð t Þ (12.10.56)
n¼1
From the last equation, one concludes that the displacements in each mode
are in phase.
The above solution resulted by assuming that the transformation u ¼ FY
diagonalizes also the damping matrix, that is,
^
FT CF ¼ C (12.10.59)
where
2 3
2x1 w1 0 ⋯ 0 ⋯ 0
6⋯ 2x2 w2 ⋯ 0 ⋯ 0 7
6 7
6⋯ ⋯ ⋯ ⋯ ⋯ ⋯ 7
^ ¼6
C 6
7
7 (12.10.60)
6⋯ ⋯ ⋯ 2x n wn ⋯ ⋯ 7
6 7
4⋯ ⋯ ⋯ ⋯ ⋯ ⋯ 5
0 0 ⋯ 0 ⋯ 2xN wN
The damping matrix satisfying Eq. (12.10.49) is called proportional.
The damping that satisfies the orthogonality condition of the eigenmodes of
the undamped free vibrations is referred to in the literature as classical damping
versus nonclassical damping where this condition is not satisfied. Methods of
constructing proportional damping matrices will be presented in the next
section.
Example 12.10.2 Determine the displacements of the frame in Example 12.2.1
in the presence of damping by the method of linearization of the quadratic
eigenvalue problem. Assume:
30 5
C¼ , uð0Þ ¼ f0:03 0:01gT , u_ ð0Þ ¼ 0
5 40
Solution
The matrices M and K were computed in Example 12.2.1. Thus we have
25 0 3826:5 3826:5
M¼ , K¼ (1)
0 32 3826:5 9142:1
Eq. (12.10.24) gives
2 3
0
0 25 0
0 M 6 0 0 0 32 7
M^¼ ¼6 7 (2)
M C 4 25 0 30 5 5
0 32 5 40
2 3
25 0 0 0
6 7
^ ¼ M 0 ¼ 6 0 32 0
K
0 7 (3)
0 K 4 0 0 3826:5 3826:5 5
0 0 3826:5 9142:1
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 581
^ K
c. The matrices M, C,K are positive definite. However, the matrices M, ^ are not positive definite.
This justifies that the eigenvalue problem has complex eigenvalues and complex eigenvectors.
@CivilMethod
582 PART II Multi-degree-of-freedom systems
8 9
< u_ = X
2N
1 ð1Þ
¼ an f n e ln t
: u_ ;
2 n¼1
08 9 8 9 1
< +0:57263 103 = < 0:22165 =
¼e0:52764t @ cos 8:27232t + sin 8:27232t A
: 0:98528 104 ; : 0:12213 ;
08 9 8 9 1
< 0:57241 103 = < 0:63866 101 =
+e 0:69736t @ cos 19:22372t + sin 19:22372t A
: +0:98450 104 ; : +0:90551 101 ;
Note that the obtained solution by this method is identical to that obtained in
Example 12.10.1.
Example 12.10.3 Determine the displacements of the frame in Example 12.2.1
in the presence of classical damping. Assume:
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 583
30 5
C¼ , uð0Þ ¼ f0:03 0:01gT , u_ ð0Þ ¼ 0
5 40
Solution
The eigenfrequencies and mode shapes were computed in Example 12.2.1,
namely
w1 8:289 0:1697 0:1061
¼ , F¼ (1)
w2 19:236 0:0935 0:1499
The transformation (12.10.49) gives
T
0:1697 0:1061 30 5 0:1697 0:1061
FT CF ¼
0:0935 0:1499 5 40 0:0935 0:1499
1:05496 0:05712
¼ (2)
0:05712 1:39556
^ 2 ¼ 2x w2 ¼ 1:39556
C (3b)
2
which yield
qffiffiffiffiffiffiffiffiffiffiffiffi
x1 ¼ 0:0636, wD1 ¼ w1 1 x21 ¼ 8:272 (4a)
qffiffiffiffiffiffiffiffiffiffiffiffi
x2 ¼ 0:0363, wD2 ¼ w2 1 x22 ¼ 19:223 (4b)
@CivilMethod
584 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 585
@CivilMethod
586 PART II Multi-degree-of-freedom systems
2 1 3
w
1 6 wn n
7 a0 xn
4 1 5 ¼ (12.11.13)
2 a1 xm
wm
wm
from which the coefficients a0 and a1 are obtained
2wn wm ðwn xm wm xn Þ
a0 ¼ (12.11.14a)
w2n w2m
2 ð wn x n wm x m Þ
a1 ¼ (12.11.14b)
w2n w2m
If xn ¼ xm ¼ x, which is a reasonable assumption, the foregoing relations
yield
2xwn wm
a0 ¼ (12.11.15a)
wn + wm
2x
a1 ¼ (12.11.15b)
wn + wm
Once the coefficients a0 and a1 are established, the damping ratios of the
remaining eigenmodes (k 6¼ m, n) can be evaluated from the relation
1 1
xk ¼ a0 + a1 wk , k 6¼ m,n (12.11.16)
2 wk
xk = a0 /2wk
xk = xk (wk )
xk = a1wk /2
wk
FIG. 12.11.1 Damping ratio versus frequency for Rayleigh damping.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 587
@CivilMethod
588 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 589
1
Further, multiplying Eq. (12.11.25) by fTm M M1 K yields
2 1
fTm M M1 K fn ¼ l1 n fm M M K
T 1
fn (12.11.27)
which by virtue of Eq. (12.11.26) gives
2 0 n 6¼ m
fTm M M1 K fn ¼ (12.11.28)
l2
n n¼m
Repeating the previous procedure results in the orthogonality relations
1 q 0 n 6¼ m q ¼ 1, 2, …1
fm M M K fn ¼
T
(12.11.29)
lq
n n ¼ m q ¼ 1, 2, …1
Eqs. (12.11.22) and (12.11.29) are combined to
q 0 n 6¼ m q ¼ 0, 1, 2, … 1
fTm M M1 K fn ¼ (12.11.30)
lqn n ¼ m q ¼ 0, 1, 2, … 1
Therefore, the matrices defined by the relation
q
Cq ¼ M M1 K , q ¼ 0, 1, 2, … 1 (12.11.31)
satisfy the orthogonality condition. Evidently, any linear combination of k N
terms
X
k X q
C¼ aq Cq ¼ aq M M1 K (12.11.32)
q q
@CivilMethod
590 PART II Multi-degree-of-freedom systems
or because ln ¼ w2n
1 1
xn ¼ a0 + a1 wn + a2 wn + ⋯ + ak1 wn
3 2k3
(12.11.35b)
2 wn
Applying the foregoing equation for n ¼ 1, 2,…,k provides the system of
linear equations
2 3
1
6 w1 w1 w31 ⋯ w2k3
1 78 9 8 9
6 7 > a0 > > x 1 >
61 7> > > >
16 w2 w32 ⋯ w2k3 7 < a1 = < x 2 =
6 7
26 w 2
7> ⋮ > ¼ > ⋮ > (12.11.36)
6 ⋯2 ⋯ ⋯ ⋯ ⋯ 7 > > > >
6 7: ak1 ; : xk ;
41 5
wk w3k ⋯ w2k3
k
wk
which yields the coefficients ak when xk is specified.
Eq. (12.11.41) shows clearly the contribution of each eigenmode to the pro-
portional damping matrix.
When using a proportional damping constructed by the previous methods,
we should take into account the following:
1. The damping matrix C obtained from Eq. (12.11.41) is in general complete.
The same holds for the damping matrix obtained using the additional
orthogonality conditions for k > 2. Apparently, the use of a complete damp-
ing matrix increases the computational cost.
2. The linear system of Eq. (12.11.36) exhibit ill conditioning because the
coefficients 1=wn ,wn ,w3n ,⋯,w2k3
n may differ significantly. As a remedy,
it is recommended that the values of series exponents in Eq. (12.11.33)
are taken near zero. For example, it is advisable to take q ¼ 2, 1,
0, 1, 2, when k ¼ 5.
3. In Rayleigh or Caughey damping, it is possible to obtain negative values for
x n . These values should be excluded; otherwise, the response of the struc-
ture will be divergent.
4. It is inappropriate when the system consists of parts with significantly dif-
ferent damping, for example, in the case of soil-structure interaction where
soil damping is much greater than that of the structure. The same is true
when energy-absorbing dampers are incorporated into the structure or
base isolation systems are used. The resulting damping matrix does not
satisfy the orthogonality condition; therefore, the method presented in
Section 12.10.2 should be employed. Another method to cope with this
problem is to use the method of substructures with a different proportional
damping matrix for each of them.
5. The use of Rayleigh damping, which offers a reasonable solution to the
damping problem, yields increased values of the damping ratios for
higher-order eigenmodes.
@CivilMethod
592 PART II Multi-degree-of-freedom systems
Solution
To obtain reliable results, the computations were performed using arithmetic
with 10 significant figures.
The solution of the eigenvalue problem ðK w2 MÞf ¼ 0 gives
8 9 8 9 2 3
< w1 = < 2:9290 = 0:07127 0:10731 0:05133
w ¼ 3:9429 , F ¼ 4 0:10257 +0:08574 0:03682 5 (1)
: 2; : ;
w3 11:0502 +0:02068 +0:00653 0:04239
The eigenmodes are orthonormalized with respect to the mass.
(i) Additional orthogonality conditions
The damping matrix will be computed using Eq. (12.11.32) for
q ¼ 0, 1, 2
2
C ¼ a0 M + a1 K + a2 M M1 K
(2)
¼ a0 M + a1 K + a2 KM1 K
The coefficients a0 ,a1 ,a2 result from Eq. (12.11.36) for N ¼ 3
2 38 9 8 9
1=w1 w1 w31 < a0 = < 0:06 =
14
1=w2 w2 w32 5 a1 ¼ 0:08 (3)
2 : ; : ;
1=w3 w3 w33 a2 0:10
Substituting the values of the eigenfrequencies from Eq. (1) and solving
the system give
a0 ¼ 0:22227 101 , a1 ¼ 0:45470 101 , a2 ¼ 0:22666 103 (4)
and Eq. (2) gives
2 3
40:219 2:549 88:247
C ¼ 4 2:549 30:647 70:130 5 (5)
88:247 70:130 806:837
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 593
@CivilMethod
594 PART II Multi-degree-of-freedom systems
Solution
1. Global mass and stiffness matrices
The geometrical data of the truss and the element properties were com-
puted in Example 11.2.2 except for the mass of the gravity load W , which
constitutes an additional nodal mass in the directions 3, 4 of the global mass
matrix. Therefore, we have m33 ¼ m44 ¼ W =g ¼ 40:775. Hence
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 595
2 3
0 0 0 0 0 0
6 7
60 0 0 0 07
0
6 7
6 7
60 0 2000:0 0 0 07
W 6 7
M ¼ rA6 7
60 0 0 2000:0 0 0 7
6 7
6 7
60 0 0 0 0 07
4 5
0 0 0 0 0 0
and
2 3
6:805 0 0 0 0 0
60 7
6 6:805 0 0 0 0 7
X 6 7
3 60 0 2004:610 0 0 0 7
¼ MW
M + ^ e ¼ rA6
M 7
60 7
6 0 0 2004:610 0 0 7
e¼1
6 7
40 0 0 0 6:805 0 5
0 0 0 0 0 6:805
2 3
0:342 0:107 0:092 0:107 0:25 0
6 0:107 0:125 0:107 0:125 0 7
6 0 7
X 6 7
3 6 0:092 0:107 0:184 0 0:092 0:107 7
K¼ ^ e 6
K ¼ EA6 7
0:107 0:125 0:125 7
6 0 0:25 0:107 7
e¼1
6 7
4 0:25 0 0:092 0:107 0:342 0:107 5
0 0 0:107 0:125 0:107 0:125
M
2. Modification of the matrices K, due to support conditions
The truss is supported as in Example 11.2.2. Thus the matrix V that
M
reorders the matrices K, is the same. This yields
@CivilMethod
596 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 597
12.12 Problems
Problem P12.1 Compute the eigenfrequencies and eigenmodes of the truss
shown in Fig. P12.1. W1 ¼ 400 kN and W2 ¼ 250 kN are gravity dead loads.
Assumed data: Coordinates of the nodes:1ð0, 0Þ, 2ð4, 0Þ, 3ð0, 4Þ,4ð4, 4Þ,
5ð8, 4Þ; cross-sectional area of bars:A1 ¼ A2 ¼ A3 ¼ 1:5A, A4 ¼ A5 ¼ A7 ¼ A,
A6 ¼ 2A, and A ¼ 27 cm2 ; modulus of elasticity E ¼ 2:1 108 kN=m2 ; and
material density r ¼ 7:55 kNs2 =m4 . Consistent mass assumption. The rigid plane
body B has dimensions 4 4 m2 , thickness h ¼ 0:10 m, and specific weight
24 kN=m3 .
@CivilMethod
598 PART II Multi-degree-of-freedom systems
Problem P12.3 Formulate the equation of motion of the two-story shear build-
ing of Fig P12.3. Then (i) compute the eigenfrequencies and eigenmodes, and
(ii) construct a proportional damping matrix of the form C ¼ a0 M + a1 K with
x1 ¼ 1:2, x2 ¼ 0:1 and formulate the equation of motion with damping. Assumed
data: E ¼ 2:1 107 kN=m2 and g ¼ 25 kN=m3 . All columns have a square
cross-section 0:30 0:30m2 . The load of the slabs also includes their dead load.
@CivilMethod
Multi-degree-of-freedom systems: Free vibrations Chapter 12 599
Problem P12.5 Formulate the Rayleigh proportional damping matrix C for the
three-degree-of-freedom system with modal damping ratios x1 ¼ 1:5, x2 ¼ 0:10,
and x3 ¼ 0:08. Consider all different cases and compare them with that obtained
using the modal matrix. Data
2 3 2 3
1000 0 5000 120000 40000 0
6 7 6 7
M ¼ 40 500 4000 5, K ¼ 4 40000 40000 0 5,
5000 4000 77833:33 0 0 100000
8 9
< 1:120 =
w ¼ 11:553
: ;
13:541
@CivilMethod
600 PART II Multi-degree-of-freedom systems
@CivilMethod
Chapter 13
13.1 Introduction
From the study of the problem of free vibrations of MDOF systems presented in
Chapter 12, it becomes obvious that the computation of the eigenfrequencies
and eigenmodes, that is, the solution of the eigenvalue problem, plays a crucial
role in the dynamic analysis of structures. In the examples presented there, the
computation of the eigenvalues li was made by determining the roots of the
characteristic polynomial
PðlÞ ¼ a0 lN + a1 lN 1 + ⋯ + aN 1 l + aN (13.1.1)
and subsequently the eigenvectors fi were obtained from the solution of the
linear algebraic system
ðK li MÞfi ¼ 0 (13.1.2)
This method requires the computation of the coefficients of PðlÞ. As soon as
this is achieved, its roots are computed by a numerical method because an ana-
lytical determination of the roots is possible only for polynomials up to fourth
degree (N ¼ 4). The determination of the polynomial coefficients requires a large
number of computations because the roots of PðlÞ are very sensitive to small
changes of the coefficients, leading to a considerable inaccuracy of the roots.
Thus, this method is rather theoretical in nature and not suitable for solving the
eigenvalue problem, particularly for large degree-of-freedom systems.
The numerical solution of the eigenvalue problem is a difficult problem of
computational linear algebra. Great efforts have been made to develop reliable
methods for solving this problem during the last two centuries. The emergence
of computers in the 1950s gave a colossal impetus for research in this area.
Thus, since that time, many new methods have been developed for the solution
of the eigenvalue problem. Generally, these methods can be grouped into the
following three categories:
1. Transformation methods.
2. Iteration methods.
3. Determinant search method.
The transformation methods are applied when the size of the eigenvalue prob-
lem is comparatively small and the matrices are fully populated or banded with
large bandwidth. The underlying idea of these methods is the use of consecutive
transformations, which render the matrix A of the standard eigenvalue problem
ðA lIÞx ¼ 0 diagonal. These methods compute all the eigenvalues simulta-
neously while the eigenmodes are computed either by adhering to the inverse
transformations or by iteration procedures. The known transformation methods
are (i) Jacobi diagonalization, named after Carl Gustav Jacob Jacobi, who first
proposed the method in 1846, although it became widely used only in the 1950s
with the advent of computers, (ii) Givens triangularization, (iii) Householder
transformation, and (iv) QR and QL transformations.
The iteration methods are suitable for solving large eigenvalue problems
with banded or sparse matrices as those in the finite elements. Particularly, they
are applied in cases where it is necessary to compute only a few eigenvalues, the
smallest or the largest ones. The most known iteration methods that are effec-
tively applied to the solution of the vibration problem in engineering are: (i)
vector iteration, (ii) vector iteration with shift, (iii) subspace iteration, and
(iv) the Lanczos method.
Finally, the determinant search method is based on the search for the roots of
the function det ðA lBÞ. It is employed when the size of the eigenvalue
problem is comparatively large while the matrices are banded with a small
bandwidth and the computation of only a few first eigenpairs is required.
The previous methods can be combined to give the best solution for a partic-
ular problem. It should be emphasized that all methods are by nature iterative
because the solution of the eigenvalue problem is equivalent to that of determin-
ing the roots of a polynomial, which, as already mentioned, are established only
by numerical iterative methods for orders of the characteristic polynomial N > 4.
Although an iterative process is necessary for the computation of the eigen-
pair ðli , fi Þ, it should be noted that as soon as one of its members has been
computed, the other member can be computed without further iterative process.
Suppose that li is computed by an iteration method, then fi can be computed by
solving the linear system (13.1.2). On the other hand, if fi has been computed
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 603
k + 1 with respect
4. If the convergence criterion is not satisfied, we normalize x
to mass
k + 1
x
xk + 1 ¼ 1=2 (13.2.7)
Tk+ 1 M
x xk + 1
and go back to step 1.
5. If the criterion is satisfied, the procedure ends and we set
f1 xk + 1 , l1 lðk + 1Þ (13.2.8)
The normalization does not affect the convergence. However, if the vector xk + 1
is not normalized, its elements increase in each step, a fact that can cause numer-
ical problems. If the desired precision of l1 is 2d digits, then the tolerance should
be taken as e ¼ 102d . The resulting precision of the eigenvector will be d digits.
Computationally, it is more effective to formulate the previous algorithm as
follows.
We take an arbitrary vector x1 and compute the vector z1 ¼ Mx1 .
Subsequently, we compute for k ¼ 1, 2, … until convergence is achieved
K
xk + 1 ¼ zk (13.2.9)
zk + 1 ¼ M
xk + 1 (13.2.10)
Tk+ 1 zk
x
k + 1 Þ¼
rðx , lðk + 1Þ ¼ rðx
k + 1 Þ (13.2.11)
k + 1zk + 1
xT
zk + 1
zk + 1 ¼ 1=2 (13.2.12)
Tk+ 1zk + 1
x
If convergence is achieved after k iterations, we take
k + 1
x
f1 ’ 1=2 , l1 lðk + 1Þ (13.2.13)
Tk+ 1zk + 1
x
The method works provided that zT1 f1 6¼ 0.
Example 13.2.1 Compute the eigenpair ðl1 , f1 Þ of the eigenvalue problem
Kf ¼ lMf using the inverse vector iteration method, when
2 3 2 3
32 24 18 0:5 0 0
K ¼ 4 68 71 38 5, M ¼ 4 0 0:5 0 5
36 27 20 0 0 1
Solution
A convenient trial vector to start the procedure is the vector x1 ¼ f1 1 1gT . The
computed eigenpair for different values of k is shown in Table E13.1.
@CivilMethod
606 PART II Multi-degree-of-freedom systems
k l1 fT1
1 1 0.1522 0.6875 0.8672
2 13.2829 0.0327 0.6113 0.9015
3 6.4215 0.0460 0.6103 0.9015
4 6.6418 0.0438 0.6112 0.9013
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 607
ð13:2:18Þ
In the inverse vector iteration method, the starting vector x1 should not be
M-orthogonal to f1 . Now y1 should not be M-orthogonal to e1 . Apparently, the
vector
y 1 ¼ f 1 1 ⋯ 1 gT (13.2.19)
satisfies this requirement.
Applying Eq. (13.2.17) for k ¼ 1, 2, …l gives
1 1 1 T
yl + 1 ¼ ⋯ (13.2.20)
ll1 ll2 llN
Because the eigenvector does not change if it is multiplied by a number,
we can write
l + 1 ¼ ll1 yl + 1
y
( l l )T
l1 l1
¼ 1 ⋯ (13.2.21)
l2 lN
@CivilMethod
608 PART II Multi-degree-of-freedom systems
y l + 1 e1 l1
lim ¼ (13.2.24)
l!1 l e1 j
jy l2
Hence, convergence is linear and the rate of convergence is r ¼ l1 =l2 . The
smaller this ratio, the faster the rate of convergence.
In the preceding proof, we have accepted that the eigenvalues are distinct.
Let us now consider the case of multiple eigenvalues, for example,
l1 ¼ l2 ⋯ ¼ lm . Then Eq. (13.2.21) becomes
( l )T
l1 l l1
l + 1 ¼ 1 1 ⋯
y ⋯ (13.2.25)
lm + 1 lN
from which we conclude that the iteration procedure converges with rate of con-
vergence l1 =lm + 1 . Hence, we can generally state that the rate of convergence
is given by the ratio of l1 to the nearest distinct eigenvalue.
The Rayleigh quotient for the vector yl + 1 is obtained by Eq. (12.6.1) for
A ¼ L and u ¼ yl + 1 . Thus we have
yT Ly
r yl + 1 ¼ l +T 1 l + 1
yl + 1 yl + 1
which by virtue of Eq. (13.2.17) becomes
yT y
r yl + 1 ¼ T l + 1 l (13.2.26)
yl + 1 yl + 1
Using Eq. (13.2.21) to express yl and yl + 1 , and substituting into the
foregoing equation gives
X
N
l1 ðl1 =li Þ2l1
i¼1
r yl + 1 ¼ (13.2.27)
X
N
2l
ðl1 =li Þ
i¼1
which implies
lim r yl + 1 ¼ l1 (13.2.28)
l!1
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 609
¼ Sn x1
@CivilMethod
610 PART II Multi-degree-of-freedom systems
¼ Kf mMf
(13.3.8)
¼ ðl mÞMf
^
¼ lMf
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 611
Π( ) = det (K − M)
| 2 − m|
| 1 − m| | 3 − m|
1 2 3 4
m
ð13:3:10Þ
In the foregoing relation, it is jli mj=lp m < 1 for all p 6¼ i. Hence,
l + 1 ! ei as l ! 1 and the iteration vector converges to fi . Moreover, we
y
obtain li ¼ l^ i + m. The rate of convergence is determined by the element
having the largest absolute value, that is,
li m
r ¼ max (13.3.11)
p6¼i lp m
@CivilMethod
612 PART II Multi-degree-of-freedom systems
where
K ^ ¼lm
^ ¼ K mM and l (13.4.2)
Obviously, K ^ is generally not singular and can be inverted. If the
selected value of m is negative and the matrix M is diagonal, this shift is
equivalent to adding a positive number to the diagonal elements of K,
which from a physical point of view expresses restraining each degree of
freedom by a spring. The advantage of this method is that the
eigenmodes remain unchanged.
(b) Restraining the free degrees using springs
The simplest method to treat the unrestrained structure that is also com-
putationally easy to implement is by restraining the degrees of freedom cor-
responding to rigid body motion by springs with a very small stiffness.
These fictitious springs are placed between the structure and the supports
and inhibit the rigid body motion. Computationally, these springs represent
additive terms to the diagonal elements of K corresponding to the unre-
strained degrees of freedom. The addition of a small stiffness, while mak-
ing the stiffness matrix nonsingular, modifies the structure. However, the
small stiffness of the springs negligibly affects the eigenmodes and the
eigenfrequencies while at the same time the eigenmodes corresponding
to the rigid body of motion are also computed. As is anticipated, the
computed eigenfrequencies corresponding to the rigid body motion are
very small (approximately zero). This is illustrated with the example that
follows.
Example 13.4.1 Compute the mode shapes and eigenfrequencies of the plane
structure in Fig. E13.1a. The density of the rigid bodies is r while the mass of
the column is negligible.
f
′
@CivilMethod
614 PART II Multi-degree-of-freedom systems
Solution
The system has three degrees of freedom, u1 ,f, u3 (Fig. E13.1b). The equation
of motion can be formulated by different methods. Here, the method of equi-
librium of forces with respect to the center of mass C is employed. Referring
to Fig. E13.1c, we obtain
m1 u€1 + fS1 ¼ 0 (1a)
a
Ic f€ + fS2 + fS1 ¼ 0 (1b)
2
m3 u€3 + fS3 ¼ 0 (1c)
where
fS1 ¼ k11 ðu2 u3 Þ + k12 f (2a)
fS2 ¼ k21 ðu2 u3 Þ + k22 f (2b)
fS3 ¼ k31 ðu2 u3 Þ + k32 f (2c)
The coefficients kij in the previous equations are computed using
Eq. (11.5.2). Moreover, we have
1
m1 ¼ ra 2 , m3 ¼ 3ra 2 , Ic ¼ ra 4
6
Substituting Eqs. (2a), (2b), (2c) into Eqs. (1a), (1b), (1c) gives
M€u + Ku ¼ 0 (3)
where
2 3 2 3 8 9
1 0 0 1:5 2:25a 1:5 < u1 =
EI
M ¼ ra 2 4 0 a 2 =6 0 5, K ¼ 3 4 2:25a 3:875a2 2:25a 5, u ¼ f
a : ;
0 0 3 1:5 2:25a 1:5 u3
(4a--c)
It is convenient to eliminate the length a and make the matrices dimension-
less. To this end, the rotational coordinate f is replaced by the translational
coordinate u2 ¼ u1 + a2 f (see Fig. E13.1b). This is accomplished by the
transformation (see Section 10.7).
8 9 2 38 9
< u1 = 1 0 0 < u1 =
f ¼ 4 2=a 2=a 0 5 u2
: ; : ;
u3 0 0 1 u3
which yields
2 3
1:667 0:667 0
^ ¼ RT MR ¼ ra 2 6
M
7
4 0:667 0:667 0 5,
0 0 3
2 3
8:0 11:0 3:0
^ ¼ RT KR ¼ EI 6
K 4 11:0 15:5 4:5 5
7
a3
3:0 4:5 1:5
2 3
1 0 0
where R ¼ 4 2=a 2=a 0 5
0 0 1
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 615
Therefore, the eigenvalues and eigenmodes will be obtained from the solu-
tion of the eigenvalue problem
^ lM
K ^ f ¼ 0, l ¼ w2 ra5 =EI
^ ¼ 0. Hence the solution is obtained either by the inverse vector
It is det K
iteration method with shift or by restraining u3 using a spring of small stiffness.
1. The inverse vector iteration with shift
a. Computation of the first eigenpair.
We employ the algorithm presented in Section 13.3.2. We take
arbitrarily x1 ¼ f 1 2 1 gT as the starting vector, which gives
m1 ¼ rðx1 Þ ¼ 3:321. After five iterations we obtain l1 ¼ 0:868e 16 0
and f1 ¼ f 1 1 1 gT . The obtained values of the eigenpair were anticipated
because the structure has a rigid body motion.
b. Computation of the second eigenpair
We apply the vector purification method described in Section 13.3.1.
We take arbitrarily x1 ¼ f 1 1 2 gT as the trial vector. This gives
2 3
0:75 0:472 1015 0:75
f f M 6
T
7
S1 ¼ I 1T 1 ¼ 4 0:25 1:000 0:75 5,
f1 Mf1
0:25 0:482 1015 0:25
8 9
< 0:749 >
> =
e
x1 ¼ S1 x1 ¼ 0:749 , m1 ¼ rðx1 Þ ¼ 2:0:
>
: >
;
0:249
After four iterations we obtain
l2 ¼ 0:239 and f2 ¼ f 0:834 0:508 0:278 gT .
c. Computation of the third eigenpair.
We take arbitrarily x1 ¼ f 2 2 1 gT as the trial vector. This gives
2 3
0:128 0:182 0:053
f f M 6
T
7
S2 ¼ S1 2T 2 ¼ 4 0:784 1:111 0:325 5,
f2 Mf2
0:043 0:061 0:018
8 9
< 0:054 >
> =
e
x 1 ¼ S2 x 1 ¼ 0:325 , m1 ¼ rðe
x1 Þ ¼ 25:0096
>
: >
;
0:018
After five iterations we obtain
l3 ¼ 24:998 and f3 ¼ f 0:231 1:411 0:077 gT .
2. Restraining u3 using a spring of small stiffness
0
Adding the small stiffness k33 ¼ 0:0001EI =a 3 to the element k33 yields
2 3
8:0 11:0 3:0
^ ¼ EI 4 11:0 15:5 4:5 5
K
a3
3:0 4:5 1:5001
@CivilMethod
616 PART II Multi-degree-of-freedom systems
n ln fn ln fn ln fn
16 4
1 0:868 10 0:999 0:249 10 1:000 0.0000 1:000
0:999 0:999 1:000
1:000 0:998 1:000
@CivilMethod
Numerical evaluation of the eigenfrequencies and eigenmodes Chapter 13 617
13.5 Problems
Problem P13.1 The one-story building of Fig. P13.1 is subjected to the harmonic
excitation p ¼ p0 sin20t. Examine whether the structure is at resonance risk.
Assume: E ¼ 2:1 107 kN=m2 , n ¼ 0:2: Columns 1, 3, 6, 8 0:35 0:35 m2 ,
Columns 2, 4, 5, 7 0:20 0:40 m2 . The dead weight of the plate is included in q.
Problem P13.2 We consider the plane structure of Fig. P13.2. At time t the sup-
port A becomes a hinge. Compute the eigenfrequencies and eigenmodes of the
structure at t + .
Problem P13.3 We consider the plane structure of Fig. P13.3 whose base is iso-
lated. Compute the eigenfrequencies (i) if the stiffness of the isolators is small
compared with that of the columns, that is, kb =k ¼ 0:1, (ii) if the isolators break
at time t, that is, kb ¼ 0. Assume; m1 ¼ m2 , mb ¼ 0:8m1 .
@CivilMethod
FIG. P13.3 Plane structure in problem P13.3
618 PART II Multi-degree-of-freedom systems
@CivilMethod
Chapter 14
Multi-degree-of-freedom
systems: Forced vibrations
Chapter outline
14.1 Introduction 620 14.9 Comparison of mode
14.2 The mode superposition superposition method
method 620 and Rayleigh-Ritz method 672
14.3 Modal contribution in the 14.10 Numerical integration
mode superposition method 629 of the equations of
14.3.1 Modal participation 629 motions—Linear MDOF
14.3.2 Static correction systems 675
method 641 14.10.1 The central
14.3.3 Error in mode difference method
superposition method (CDM)—Linear
due to truncation of equations 675
higher modes 647 14.10.2 The average
14.4 Reduction of the dynamic acceleration method
degrees of freedom 649 (AAM)—Linear
14.4.1 Static condensation 649 equations 677
14.4.2 Kinematic constraints 650 14.10.3 The analog equation
14.5 Rayleigh-Ritz method 651 method (AEM)—
14.5.1 Ritz transformation 651 Linear equations 679
14.5.2 Approximation using 14.11 Numerical integration of the
Ritz vectors 654 equations of motions—
14.6 Selection of Ritz vectors 655 Nonlinear MDOF systems 681
14.6.1 Method of natural 14.11.1 The average
mode shapes 656 acceleration method
14.6.2 The method of (AAM)—Nonlinear
derived Ritz vectors 659 equations 681
14.7 Support excitation 663 14.11.2 The analog equation
14.7.1 Multiple support method (AEM)—
excitation 663 Nonlinear equations 684
14.7.2 Uniform support 14.12 Problems 688
excitation 665 References and further reading 692
14.8 The response spectrum
method 668
14.1 Introduction
In this chapter, we study the dynamic response of MDOF systems under the
action of external forces. As was shown in Chapter 10, the equation of motion
of a linear system with viscous damping is
u + Cu_ + Ku ¼pðt Þ
M€ (14.1.1)
where M is the mass matrix, C the damping matrix, K the stiffness matrix, u the
displacement vector, and pðt Þ the vector of the external forces in the directions
of the displacements. Besides, the system is subjected to the initial conditions
where f1 ,f2 ,…, fN are the eigenmodes of the free undamped vibrations.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 621
+ Kðf1 Y1 + f2 Y2 + ⋯ + fN YN Þ ¼ pðt Þ
(14.2.2)
Premultiplying the previous equation by fTn
and assuming that the damping
matrix is proportional, we obtain by virtue of the orthogonality properties
Mn Y€ n + Cn Y n + Kn Yn ¼Pn ðt Þ, n ¼ 1, 2,…,N
(14.2.3)
where
Mn ¼ fTn Mfn (14.2.4a)
Cn ¼ fTn Cfn (14.2.4b)
Kn ¼ fTn Kfn (14.2.4c)
Pn ¼ fTn pðt Þ (14.2.4d)
The quantities defined previously are referred to as generalized or modal
mass, generalized or modal damping, generalized or modal stiffness, and
generalized or modal force of the nth mode shape, respectively. Taking into
account that
Kn ¼ w2n Mn (14.2.5)
Cn ¼ 2x n Mn wn (14.2.6)
Eq. (14.2.3) is written as
1
Y€ n + 2xn wn Y n + w2n Yn ¼
The initial conditions Yn ð0Þ, Y n ð0Þ are computed using Eqs. (12.9.12),
(12.9.13), namely
fTn Muð0Þ
Y n ð 0Þ ¼ , n ¼ 1, 2, …, N (14.2.9a)
Mn
fTn Mu_ ð0Þ
Y n ð 0Þ ¼ , n ¼ 1, 2, …, N (14.2.9b)
Mn
@CivilMethod
622 PART II Multi-degree-of-freedom systems
5. Compute the initial conditions Yn ð0Þ, Y n ð0Þ from the natural initial condi-
tions uð0Þ, u_ ð0Þ using Eqs. (14.2.9a), (14.2.9b).
6. Compute the modal coordinates Yn ðt Þ either using Eq. (14.2.8) with analyt-
ical evaluation of the Duhamel integral or by direct numerical solution of
Eq. (14.2.7).
7. Compute the displacement vector uðt Þ from the modal contributions
fn Yn ðt Þ using the superposition relation (14.2.1).
8. Compute the nodal elastic forces f S ¼ Ku.
Example 14.2.1 Determine the dynamic response of the two-story shear frame
in Example 12.2.1 for the following two load cases:
(a) The horizontal loads p1 ¼ 190 kN and p2 ¼ 300 kN are suddenly applied
at t ¼ 0 while the structure is at rest.
(b) The ground undergoes the displacement ug ðt Þ ¼ 0:02 sin 10t.
More specifically, in case (a), compute the extreme values of the displacements,
shear forces, and bending moments at the top of the columns while in case (b),
compute the maximum shear force at the base of the frame and the maximum
overturning moment. Consider proportional damping with modal damping
ratios x1 ¼ 0:06, x 2 ¼ 0:04.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 623
Solution
The structure properties, eigenfrequencies, and mode shapes are taken from
Example 12.2.1, namely
25 0 3826:5 3826:5
M¼ , K¼ ,
0 32 3826:5 9142:1
8:289 0:1697 0:1061
w¼ , F¼
19:236 0:0935 0:1499
Case (a)
The modal matrices are computed using Eq. (14.2.4a)
25 0 0:1697
M1 ¼ ½0:1697 0:0935 ¼1
0 32 0:0935
25 0 0:1061
M2 ¼ ½ 0:1061 0:1499 ¼1
0 32 0:1499
The load vector is
190
pðt Þ ¼
300
and Eq. (14.2.4d) gives the modal loads
190
P1 ¼ ½ 0:1697 0:0935 ¼ 4:1930
300
190
P2 ¼ ½ 0:1061 0:1499 ¼ 65:1290
300
The equations of motion for the modal coordinates Y1 ðt Þ, Y2 ðt Þ result from
Eq. (14.2.7), that is
@CivilMethod
624 PART II Multi-degree-of-freedom systems
The Duhamel integral was evaluated using MAPLE (see Eq. 3.4.7). Hence,
using the data of the problem, we obtain
Y1 ðt Þ ¼ 0:06125 0:06125 cos 8:2741t + 0:36817 102 sin 8:2741t e0:4973t
Y2 ðt Þ ¼ 0:17629 0:17629cos 19:2206t + 0:70572 102 sin 19:2206t e0:7694t
u2(t)
t t
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 625
Q1(t)
Q2(t)
t t
Case (b)
The equation of motion in terms of the relative displacements u 1 ¼ u1 ug ,
u 2 ¼ u2 ug (see Chapter 6) is
€ + Cu
Mu _ + K
u ¼
pð t Þ
where
1 50
ðt Þ ¼ M
p u€g ¼ sin 10t
1 64
@CivilMethod
626 PART II Multi-degree-of-freedom systems
The solution of the foregoing equations is obtained from Eq. (14.2.8). The
Duhamel integral is evaluated using MAPLE. Thus, setting Pn ðτÞ ¼ Pn sin w τ
we can obtain
Z t
Pn
Yn ð t Þ ¼ exn wn ðtτÞ sin w
τ sin ½wDn ðt τÞdτ
Mn wDn 0
Pn 20x n wn cos 10t + w2Dn + x2n w2n 100 sin 10t
¼
Mn x2 w2 + ðw 10Þ2 x2 w2 + ðw + 10Þ2
n n Dn n n Dn
x 2 w2 1000
20x n wn cos wDn t + 10wDn + 10 n n + sin wDn t exn wn t
Pn w wDn
+ Dn
Mn x n wn + ðwDn 10Þ
2 2 2
x n wn + ðwDn + 10Þ2
2 2
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 627
t t
@CivilMethod
628 PART II Multi-degree-of-freedom systems
Solution
The load vector is
t0gT
pðt Þ ¼ f p0 sin w
Eq. (14.2.4d) gives the modal loads
t
P1 ¼ fT1 pðt Þ ¼ f11 p0 sin w
t
P2 ¼ fT2 pðt Þ ¼ f12 p0 sin w
and Eq. (14.2.7) becomes
f1n p0
Y€ n + 2xn wn Y n + w2n Yn ¼pn0 sin w
t, pn0 ¼ , n ¼ 1, 2 (1)
Mn
The steady-state response is obtained from Eq. (3.2.26) by setting u ¼ Yn
t qn Þ
Yn ðt Þ ¼ rn sin ðw (2)
where now
!
Pn0 h i 1
2 2 2x n b n
w
qn ¼ tan 1
2 2
rn ¼ 1 b n + ð 2x n b n Þ , , bn ¼
Mn w2n 1 b2n wn
(3)
The displacements of the steady state response result from Eq. (14.2.1).
t q1 Þ + f2 r2 sin ðw
u ¼ f1 r1 sin ðw t q2 Þ (4)
The eigenfrequencies and mode shapes are taken from Example 12.2.1,
namely
w
w
b1 ¼ ¼ 1:2064, b2 ¼ ¼ 0:5198
w1 w2
and Eqs. (3) give
r1 ¼ 0:02584, r2 ¼ 0:001961, q1 ¼ 0:3076, q2 ¼ 0:0569
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 629
Hence
u1 ðt Þ 0:438
¼ 102 sin ð10t + 0:3076Þ + 103
u2 ðt Þ 2:416
0:208
sin ð10t 0:0569Þ
0:294
un ¼ fn Yn ðt Þ (14.3.1b)
The exact solution is obtained when the sum includes all the modal displace-
ments. However, the economy of the analysis requires the use of a few modal
displacements corresponding to the lower eigenmodes, at least those whose
contribution is not negligible, while cutting off higher order modal displace-
ments. Besides, the truncation is necessary for another important reason. The
modeling of the structure with finite elements reveals deviations between the
discretized and the actual structure, which become pronounced in higher-order
eigenmodes. Therefore, the additional error due to truncated higher eigenmodes
does not concern us very much. In order to estimate this error, we first need to
understand the contribution of each eigenmode to the overall response.
@CivilMethod
630 PART II Multi-degree-of-freedom systems
Obviously, the resultant of the above forces equilibrates only a part pn ðt Þ of the
external force pðt Þ. Therefore, we have
pn ðt Þ ¼ ðf I Þn + ðf D Þn + ðf S Þn
¼ MY€ n + CY n + KYn fn
(14.3.2)
¼ Aðt Þfn
and consequently
X
N
pðt Þ ¼ pn ðt Þ (14.3.3)
n¼1
h i
¼ Y€ n + a0 + a1 w2n Y n + w2n Yn Mfn
(14.3.4)
¼ ðt ÞMfn
and Eq. (14.3.2) is written as
pn ðt Þ ¼ ðt ÞMfn
(14.3.5)
¼ ðt Þen
From the forgoing equation, it is evident that the vectors en ¼ Mfn
(n ¼ 1, 2, …, N ) do not change with time, therefore they provide a constant vec-
tor base to express the vector pðt Þ. Therefore, we can write
X
N
pðt Þ ¼ an ðt Þen (14.3.6)
n¼1
@CivilMethod
632 PART II Multi-degree-of-freedom systems
Y€ n + 2xn wn Y n + w2n Yn ¼ Gn f ðt Þ
(14.3.15)
or
where
Yn ¼ G n y n (14.3.17)
Eq. (14.3.16) is identical to Eq. (3.3.16) and expresses the motion of the
SDOF system when m ¼ 1, w ¼ wn , and pðt Þ ¼ f ðt Þ. The solution of
Eq. (14.3.16) gives yn ðt Þ and Eq. (14.3.1b) gives the modal displacement
un ¼ Gn fn yn ðt Þ (14.3.18)
The corresponding elastic force is
ðf S Þn ¼ Kun
(14.3.19)
¼ KGn fn yn ðt Þ
The dynamic response at time t is obtained as the static responsea when the
structure is loaded by the elastic forces. We denote by qn ðt Þ the contribution of
the nth mode to a certain quantity q ðt Þ (representing deformation or stress).
Hence qn ðt Þ will be obtained as the static response of the structure produced
by the load ðf S Þn . If qnst denotes the static response due to static load Rn , then
we may write
qn ðt Þ ¼ qnst w2n yn ðt Þ (14.3.21)
The foregoing equation leads to the conclusion that the nth mode contribu-
tion qn ðt Þ is obtained as a result of two analyses: (1) the static analysis of the
structure subjected to the external load Rn , and (2) the dynamic analysis of the
nth mode SDOF system under the excitation force f ðt Þ. Therefore, the dynamic
analysis requires the static analysis of the structure by N loadings Rn ,
n ¼ 1, 2, …,N , and the dynamic analysis of N different SDOF systems under
the same excitation force f ðt Þ.
a. The term static response is used in the sense that time is involved as a parameter and no inertial
forces or damping forces are produced. This response is called also quasistatic.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 633
Note that q st results as the static analysis of the structure subjected to the
load R. Evidently, this definition of g qn does not require the computation of
all qnst when the response is approximated by the first few mode shapes.
The quantity g qn defined by Eq. (14.3.23) is called the nth modal contribution
factor of the quantity q. This modal contribution factor has three important
properties, which show its advantage over the modal participation factor Gn
defined by Eq. (14.3.14), namely
(1) It is by definition dimensionless.
(2) It does not depend on how the modes are normalized because qnst is due to
the load Rn , which does not depend on the mode normalization method,
and the modal properties do not enter into q st .
(3) The sum of all participation factors is equal to unity
X
N
g qn ¼ 1 (14.3.24)
n¼1
@CivilMethod
634 PART II Multi-degree-of-freedom systems
The quantity
yn0
Dn ¼ (14.3.28)
ynst 0
defines the dynamic magnification factor of the nth mode. Thus, Eq. (14.3.25)
by virtue of Eqs. (14.3.27), (14.3.28) becomes
qn0 ¼ g qn q st f0 Dn (14.3.29)
Eq. (14.3.29) shows that the extreme value of the nth mode contribution to
the quantity q ðt Þ is the product of four quantities:
(1) The dimensionless modal contribution factor g qn
(2) the static value of q due to the loading R
(3) the maximum value of the function f ðt Þ
(4) the nth mode dynamic magnification factor Dn .
It is emphasized that the quantities q st and g n depend only on the spatial
distribution R of the external forces while Dn and f0 only on f ðt Þ. When the
excitation is due to seismic ground motion, the quantity w2n yn0 , according to
Eq. (6.2.21), represents the pseudoacceleration, Spa ðTn , xn Þ ¼ w2n yn0 , hence
its value can be taken from the respective response spectrum (see Fig. 6.2.6).
Example 14.3.1 The chimney of the variable cross-section shown in
Fig. E14.4a is modeled with six beam elements of constant cross-section.
Compute the modal contribution factor of (i) the base shear force Qb , (ii)
the overturning moment Mb , and (iii) the top displacement u6 when the excita-
t. Assume: m1 ¼ 6m, m2 ¼ 5m,
tion is due to ground motion ug ðt Þ ¼ uo sin w
m3 ¼ 4m, m4 ¼ 3m, m5 ¼ 2m, m6 ¼ m, I1 ¼ 16I , I2 ¼ 11I , I3 ¼ 7I , I4 ¼ 4I ,
I5 ¼ 2I , and I6 ¼ I :
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 635
Solution
The equation of motion is (see Section 14.7).
u + Cu_ + Ku ¼ M1u€g
M€ (1)
T
where 1 ¼ f 1 1 1 1 1 1 g , u€g ðt Þ ¼ t
2
w u0 sin w
1. Formulation of the matrices M, K and vector R
The structure has 12 degrees of freedom, 6 translational, and 6 rotational
(Fig. E14.4c). After the static condensation of the rotational degrees of freedom,
we obtain:
2 3
6 0 0 0 0 0
60 5 0 0 0 07
6 7
60 0 4 0 0 07
M ¼ m6 60 0 0 3 0 07
7 (2)
6 7
40 0 0 0 2 05
0 0 0 0 0 1
2 3
232:80 121:78 37:89 7:11 1:21 0:15
6 121:78 122:97 69:99 21:01 3:57 0:44 7
6 7
6 37:89 69:99 72:03 39:04 10:73 1:34 7
K ¼ k66 7 (3)
6 7:11 21:01 39:04 38:38 19:04 3:88 7 7
4 1:21 3:57 10:73 19:04 16:00 5:00 5
0:15 0:44 1:34 3:88 5:00 2:12
where
EI
k¼
ðL=6Þ3
Moreover, it is
8 9
> 6>
> >
> >
>
> 5>>
< >
> =
4
R ¼ M1 ¼ m , f ðt Þ ¼ u€g ðt Þ (4)
>
> 3>>
>
> >
>
>2>
>
: >
;
1
@CivilMethod
636 PART II Multi-degree-of-freedom systems
2 3
0:0161 0:0547 0:1123 0:1665 0:2106 0:2806
6 0:0645 0:1747 0:2478 0:1858 0:0205 0:2626 7
6 7
1 6 6 0:1469 0:2736 0:1305 0:1863 0:2569 0:1893 77
F ¼ pffiffiffiffiffi 6 7
m 6 0:2640 0:2310 0:2348 0:1831 0:3320 0:1067 7
6 7
4 0:4133 0:0662 0:2981 0:4140 0:2491 0:0498 5
0:5835 0:6010 0:4539 0:2807 0:1148 0:0178
(5)
st st st
4. Computation of Qbn , Mbn , and u6n
The base shear is obtained as the sum of all the elastic forces while the over-
turning moment as the sum of the moments of all the elastic forces with respect
to the base of the chimney. Thus, we have
X
6
st
Qbn ¼ Rkn ¼ 1T Rn
k¼1
X
6
L
st
Mbn ¼ hk Rkn ¼ hT Rn , hT ¼ f 1 2 3 4 5 6 g
k¼1
6
Gn
st
u6n ¼ f :
w2n 6n
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 637
Finally, the modal contribution factors are computed from Eq. (14.3.23):
st st st
Qbn Mbn u6n
n ¼
gQ , n ¼
gM , g un6 ¼
b b
X
6 X
6 X
6
st st st
Qbk Mbk u6k
k¼1 k¼1 k¼1
Their values are given in Table E14.1. From this table, we deduce that each
quantity requires a different number of contributing modes to achieve a spec-
ified degree of approximation. For example, four modes for the base shear, two
modes for the overturning moment, and one mode for the top displacement
should contribute in order to achieve an approximation over 90%.
Example 14.3.2 For the structure in Example 14.3.1, compute the extreme
values Qbn0 , Mbn0 , u6n0 when w ¼ 1:5w1 , m ¼ 31:056 kN m1 s2 ,
k ¼ 5486:208 kN=m, L ¼ 75 m, x1 ¼ 0:06, and x2 ¼ x3 ¼ ⋯ ¼ x6 ¼ 0:04,
u0 ¼ 1.
Solution
The extreme value qn0 of a quantity q is obtained from Eq. (14.3.29), namely
qn0 ¼ g qn q st f0 Dn
The modal contribution factors g qn (q Qb , Mb , u6 ) were computed in
Example 14.3.1 and are given in Table E14.1. It is f ðt Þ ¼ u0 w t, hence
2 sin w
. The quantities Qb , Mb are obtained from the relations:
f0 ¼ max jf ðt Þj ¼ u0 w 2 st st
X
6
Qbst ¼ Rk ¼ 1T R ¼ 652:1760 kN (1)
k¼1
@CivilMethod
638 PART II Multi-degree-of-freedom systems
X
6
Mbst ¼ hk Rk ¼ hT R ¼ 21739:2 kN m (2)
k¼1
Hence
u6st ¼ 0:2451 (3)
Therefore, the initial conditions yn ð0Þ, y_ n ð0Þ result from Eqs. (14.2.9a),
(14.2.9b) with Mn ¼ 1, n ¼ 1, 2, …,6:
T
_ g ð0Þ
f n M u fTn M1
y_ n ð0Þ ¼ Y n ð0Þ ¼ Gn ¼
wu0 , n ¼ 1, 2, …,6 (4b)
Gn
y_ n ð0Þ ¼
wu0 , n ¼ 1, 2, …,6 (5)
The eigenfrequencies wn are taken from Example 14.3.1, which for data of
this problem become
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 639
R1(t)
R2(t)
t t
R3(t)
R4(t)
t t
R5(t)
R6(t)
t t
FIG. E14.5 Time history of the response ratios and modal magnification factors in Example 14.4.2.
Eq. (6) demands the analytical evaluation of the Duhamel integral, which
leads to a complicated expression. To avoid this, the solution is obtained numer-
ically using the program called aem_lin.m described in Section 4.4 while the
modal dynamic magnification factors are obtained from the relation
Dn ¼ max ðjRn jÞt , where Rn ðt Þ ¼ yn =ðyn Þst ðn ¼ 1, 2, …, 6Þ represents the
respective response ratio within the integration interval. Their graphical
@CivilMethod
640 PART II Multi-degree-of-freedom systems
TABLE E14.2 Extreme values of Qbn0 , Mbn0 , and u6n0 in Example 14.3.2.
Solution
The vector Rn is obtained from Eq. (14.3.13), namely
Rn ¼ Gn Mfn
where
Gn ¼ fTn R ¼ lfTn fk n ¼ 1, 2, …,6
In general it is 6¼ 0. Hence, Rn 6¼ 0 and the loading pðt Þ excites all the
fTn fk
mode shapes, in contrast to the free vibrations where only the corresponding
mode shape is excited when the initial conditions are proportional to a certain
of its mode shapes (see Example 12.9.2). The mass matrix, the stiffness matrix,
and the mode shapes are taken from Example 14.3.1.
(i) R ¼ lf1
Computation of Rn and g Q n
The mode shapes have been orthonormalized with respect to the mass, hence
Mn ¼ 1, n ¼ 1, 2, …, 6. Eq. (14.3.14) gives:
pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi
G1 ¼ 0:6071l m , G2 ¼ 0:2647l m , G3 ¼ 0:1167l m
pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi
G4 ¼ 0:0537l m , G5 ¼ 0:0186l m , G6 ¼ 0:0025l m
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 641
(ii) R ¼ lf2
Computation of Rn and g Qn
Eq. (14.3.14) gives:
pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi
G1 ¼ 0:2647l m , G2 ¼ 0:5273l m , G3 ¼ 0:22218l m
pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi
G4 ¼ 0:0895l m , G5 ¼ 0:0309l m , G6 ¼ 0:0040l m
Eq. (14.3.13) gives:
lR1 =m lR2 =m lR2 =m lR2 =m lR2 =m lR2 =m
0:0256 0:1729 0:1497 0:0894 0:0391 0:0067
0:0854 0:4605 0:2752 0:0832 0:0031 0:0052
0:1555 0:5770 0:1159 0:0667 0:0317 0:0030
0:2096 0:3654 0:1565 0:0491 0:0308 0:0012
0:2188 0:0698 0:1324 0:0741 0:0154 0:0003
0:1544 0:3168 0:1008 0:0251 0:0035 0:0000
and using Eq. (14.3.23), we obtain
@CivilMethod
642 PART II Multi-degree-of-freedom systems
ttot ¼ 39 s. We observe that for ttot =Tn > 1000, hence for periods Tn < 0:039 s
or eigenfrequencies wn > 160s1 , we may ignore the dynamic effect and
assume Dn 1 for eigenmodes of order n > Nd , where Nd is the number of
eigenmodes with Tn > 0:039 s and Dn noticeably greater than the one. This
observation is the quintessence of the static correction method, presented next.
Suppose that we use N eigenmodes to approximate the dynamic response of
a quantity q ðt Þ. We split the contribution of eigenmodes into two parts. The first
Nd eigenmodes with Dn greater than one and the remaining N Nd with Dn
close to one. Then we can write
X
Nd
X
N
q ðt Þ ¼ qnst w2n yn ðt Þ + qnst w2n ynst ðt Þ (14.3.30)
n¼1 n¼Nd + 1
where yn ðt Þ is taken from the solution of Eq. (14.3.16) and ynst ðt Þ from
Eq. (14.3.26). Substituting Eq. (14.3.26) into Eq. (14.3.30) and using
Eq. (14.3.23) we obtain
X
Nd
X
N
q ðt Þ ¼ qnst wn2 yn ðt Þ + f ðt Þ qnst
n¼1 n¼Nd + 1
( ) (14.3.31)
X
Nd
2 X
N
¼q st
gqn wn yn ðt Þ + f ðt Þ gqn
n¼1 n¼Nd + 1
3.5
3 x =0.05
x =0.1
2.5 x =0.15
2
Dn
1.5
0.5
0
0 200 400 600 800 1000 1200 1400 1600
t tot /Tn
FIG. 14.3.1 Variation of the dynamic magnification factor Dn versus the ratio ttot/Tn
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 643
or
( )
X
Nd
q ðt Þ ¼ q st f ðt Þ + g qn w2n yn ðt Þ f ðt Þ (14.3.34)
n¼1
represents the static correction that should be added to the dynamic solution in
order to obtain the total dynamic response. This method is known as the static
correction method.
This static correction method is very useful when the contribution of
higher eigenmodes to a quantity q ðt Þ cannot be neglected in dynamic analysis,
but the corresponding dynamic magnification factors produced by the excita-
tion force f ðt Þ are close to one. In such cases, the combination of the dynamic
response of a few lower eigenmodes together with the static correction will
give results, which adequately approximate the response obtained by the mode
superposition method. The advantage of the static correction method offers a
significant reduction in the computational task, especially when the equations
are solved numerically. The example that follows illustrates the merits of this
method.
Example 14.3.4 The chimney in Example 14.3.1 is subjected to the wind blast
load pðt Þ ¼ Rf ðt Þ R ¼ f 1:5 6:0 15:0 25:0 40:0 60:0 gT . The time function
f ðt Þ is given in Fig. E14.6 (p0 ¼ 1 t1 ¼ 0:1). Compute the response of the base
shear force, the overturning moment, and the top displacement using the exact
mode superposition method as well as the static correction method by including
the dynamic response of the eigenmodes with Dn > 1:15. Compare the extreme
values obtained from the two methods. The equation of motion will be solved
numerically to obtain yn ðt Þ. L ¼ 75 m, x1 ¼ 0:06, x2 ¼ x3 ¼ ⋯ ¼ x6 ¼ 0:04,
m ¼ 31:056 kN m1 s2 , and k ¼ 5486:208 kN=m.
@CivilMethod
644 PART II Multi-degree-of-freedom systems
f(t)
t
FIG. E14.6 Time variation of the blast load in Example 14.3.4.
Solution
The mass matrix, stiffness matrix, eigenfrequencies, and mode shapes are taken
from Example 14.3.1 for the given values of k and m. The vectors Rn are
obtained from Eq. (14.3.13) with Mn ¼ 1. Thus, we obtain
R1 R2 R3 R4 R5 R6
5:888 9:085 8:7984 6:284 2:279 0:096
19:604 24:191 16:170 5:843 0:184 0:075
35:708 30:311 6:8147 4:685 1:853 0:043
48:124 19:197 9:196 3:454 1:796 0:018
50:225 3:6693 7:782 5:206 0:899 0:005
35:456 16:644 5:925 1:765 0:207 0:001
st st st
The quantities Qbn , Mbn , and u6n are computed from the relations
X
6
st
Qbn ¼ Rkn ¼ 1T Rn
k¼1
X
6
L
st
Mbn ¼ hk Rkn ¼ hT Rn , hT ¼ f 1 2 3 4 5 6 g
k¼1
6
Gn
st
u6n ¼ f
w2n 6n
The modal contribution factors of Qb , Mb are obtained from Eq. (14.3.23). The
computed values are given in Table E14.3. The dynamic magnification factors
are computed from the relation Dn ¼ max jRðt Þj, where Rðt Þ ¼ yn ðt Þ= ynst 0 is
t
the response ratio of the nth mode and ynst 0 ¼ f0 =w2n . For the given loading it
is f0 ¼ 1. Hence
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 645
st
y1 0 ¼ 0:1374, y2st 0
¼ 0:9870 102 , y3st 0
¼ 0:1761 102
st st
y4 0 ¼ 0:5170 103 , y5 0 ¼ 0:2096 103 , y6st 0
¼ 0:9045 104
The values max jyn ðt Þj are obtained from the solution of Eq. (14.3.16),
t
where f ðt Þ is obtained from Fig. E14.6. The initial conditions are
uð0Þ ¼ u_ ð0Þ ¼ 0, which yields yn ð0Þ ¼ y_ n ð0Þ ¼ 0. The solution of the equation
of motion is obtained numerically using the program aemlin.m developed in
Section 4.4. Thus, we obtain
We observe that the dynamic magnification factors of the last two modes
are close to one, so we can include the two higher modes in the static correc-
tion and obtain the total response from Eq. (14.3.34) with Nd ¼ 4. That is
( )
X
4
2
q ðt Þ ¼ q f ðt Þ +
st
g n wn yn ðt Þ f ðt Þ
q
n¼1
@CivilMethod
646 PART II Multi-degree-of-freedom systems
Static correction
Exact
Qb (t)
Static correction
Exact
Mb (t)
Static correction
Exact
u6(t)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 647
are given in Table E14.4. It becomes obvious that the static correction method
provides very good results.
TABLE E14.4 Extreme values max jQb j, max jMb j, max ju6 j in Example 14.3.4.
Base shear force Overturning moment Top displacement
max jQb j max jMb j max ju6 j
Static Static Static
Exact correction Exact correction Exact correction
Extreme 92.4876 91.500 3341.2 3341.4 0.0384 0.0384
value
tmax 0.634 0.634 0.620 0.620 0.751 0.751
@CivilMethod
648 PART II Multi-degree-of-freedom systems
pðt Þ (14.3.37)
n¼1
Obviously, it is e
pðt Þ 6¼ pðt Þ, when K < N . Taking into account Eq. (14.3.2),
Eq. (14.3.37) becomes
X
K
pn ðt Þ ¼ e
pðt Þ (14.3.38)
n¼1
Namely, the K < N modal forces equilibrate a part of the external force
pðt Þ. This results in an error
e¼pe
p (14.3.39)
Then the error norm eK for e can be defined as
pT e
eK ¼ (14.3.40)
pT p
This gives eK ¼ 0, if K ¼ N , and e ¼ 1, K ¼ 0.
Obviously, when the loading is given by Eq. (14.3.11), the error takes the form
RT e
eK ¼ (14.3.41)
RT R
where now
X
K
e¼R Rn
n¼1
(14.3.42)
X
K
¼R Gn Mfn
n¼1
Table E14.5 gives the mode truncation error in Examples 14.3.1 and 14.3.3.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 649
@CivilMethod
650 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 651
@CivilMethod
652 PART II Multi-degree-of-freedom systems
problem, especially for bodies of arbitrary shape [2]. In MDOF systems, the
shape functions are replaced by a set of discrete displacements, which define
admissible deformation patterns of the structure. Ritz applied this method in
1909 as an extension of Rayleigh’s method to establish the lower eigenfrequen-
cies, therefore it is also known as the Rayleigh-Ritz method.
For that reason, in the Ritz method for the MDOF systems, the displacement
vector is expressed as the superposition of a number of linearly independent
vectors representing possible deformation patterns of the structure. These vec-
tors are referred to as Ritz vectors. Thus, we can write
u ¼ y1 z1 ðt Þ + y2 z2 ðt Þ + ⋯ + yK zK ðt Þ
(14.5.2)
¼ xz
where z represent the vector of K generalized coordinates and C the matrix of
the Ritz vectors with dimensions N K ðK N Þ. The transformation defined
by Eq. (14.5.2) is known as the Ritz transformation. Apparently, the transfor-
mation u ¼ FY we met in Section 12.9 represents a special case of the Ritz
transformation when the vibration modes are used as the Ritz vectors.
Substituting the vector u from Eq. (14.5.2) into Eq. (14.1.1) gives
Mx€z + Cx z_ + Kxz ¼pðt Þ (14.5.3)
which, when premultiplied by x , gives
T
M€ e z + Kz
e z + C_ e ¼e
pðt Þ (14.5.4)
where
M e ¼ x T Cx, K
e ¼ x T Mx, C e ¼ x T Kx, e
pðt Þ ¼ x T pðt Þ (14.5.5)
Eq. (14.5.4) represents a system of K differential equations for the gener-
alized coordinates zk ðt Þ. The Ritz vectors generally do not satisfy the orthogo-
nality conditions, so consequently the matrices defined by Eq. (14.5.5) are not
diagonal; hence Eq. (14.5.4) is coupled. Nevertheless, their solution can be
obtained using the method of superposition presented in Section 14.2. Thus,
we have first to establish the eigenfrequencies w e n and eigenmodes zn from
the solution of the eigenvalue problem
Ke w e z¼0
e2 M (14.5.6)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 653
e ðt Þ
u ¼ xΖY
(14.5.8)
¼FeY
e ðt Þ
The latter equation is analogous to Eq. (14.5.2) and expresses the Ritz trans-
formation of the mode shapes of the eigenvalue problem (14.5.6). We can read-
ily show that the transformed eigenmodes f e n satisfy the orthogonality
conditions with respect to the original mass and stiffness matrices M and K,
namely
f e i ¼ 0,
e T Mf e T Kf
f e i ¼ 0, n 6¼ i (14.5.11)
n n
Indeed, we have
e i ¼ xzn T M xzi
e T Mf
f n
¼ zTn x T Mx zi
e i
¼ zTn Mz
¼0 (14.5.12)
because zn and zi are mode shapes of the eigenvalue problem (14.5.6). Simi-
larly, we prove the second orthogonality condition. The orthogonality with
respect to the damping matrix holds only if it is a proportional damping matrix
as discussed in Section 12.11. If the eigenmodes are orthonormalized with
respect to mass, we have
e n ¼ 1,
zTn Mz e n ¼w
zTn Kz e 2n (14.5.13)
A consequence of this is
f e n ¼ 1,
e T Mf e T Kf
f en ¼ w
e 2n (14.5.14)
n n
@CivilMethod
654 PART II Multi-degree-of-freedom systems
The value of the Rayleigh quotient is altered by changing the selected defor-
mation pattern, which occurs when one or more of the components of the vector
z change. The matrices Ke and Me are positive definite, hence rðzÞ has finite
values and according to Eq. (12.6.2), we have
0 < l1 rðuÞ lN < 1 (14.5.16)
We know that in the neighborhood of an eigenvalue, rðzÞ takes a stationary
value (see Section 12.6), which is a minimum near all eigenvalues except near
lN , where it is a maximum. The condition ensuring this is
∂r
¼ 0, k ¼ 1, 2, …, K (14.5.17)
∂zk
or
Te e
1 ∂ z Kz e
zT Kz ∂ zT Mz
2 ¼0 (14.5.18)
e
zT Mz ∂zk e ∂zk
zT Mz
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 655
e and M
Taking into account that the matrices K e are symmetric and that
∂z=∂zk ¼ ∂z =∂zk ¼ I, we obtain
T
e
∂ zT Kz ∂zT e e ∂z
¼ Kz + zT K
∂zk ∂zk ∂zk (14.5.20a)
e
¼ 2Kz
and similarly
e
∂ zT Mz ∂zT e e ∂z
¼ Mz + zT M
∂zk ∂zk ∂zk (14.5.20b)
e
¼ 2Mz
which proves that the stationary values of r are the eigenvalues of the trans-
formed eigenvalue problem (14.5.21). The eigenvalue problem is solved using
any of the methods presented in Chapter 13. The eigenvalues give the
eigenfrequencies
pffiffiffiffiffi
en ¼ rn , n ¼ 1, 2, …, K
w (14.5.22)
@CivilMethod
656 PART II Multi-degree-of-freedom systems
Apparently, the method of natural mode shapes requires the experience of its
user. It is clear that this method becomes difficult for complicated structures and
practically impossible for three-dimensional structures. The following example
illustrates the implementation of the method of natural mode shapes.
Example 14.6.1 Determine the eigenfrequencies and mode shapes of the chim-
ney in Example 14.3.1 using the Rayleigh-Ritz with (i) two, (ii) three, and (iii)
four Ritz vectors. Use the mode shapes of the cantilever with a constant cross-
section as Ritz vectors.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 657
Solution
The mode shapes of the cantilever with a constant cross-section are obtained
from Eq. (8.3.41). Because C1 is arbitrary, we may write
1
n ðx Þ ¼ ½ cosh an x cos an x Cn ð sinh an x sin an xÞ
2
where
cosh an + cos an
an ¼ bn L, x ¼ x=L, Cn ¼
sinh an + sin an
The values of an for n ¼ 1, 2, 3, 4 are taken from Table 8.3.1, That is
a1 ¼ 1:875, a2 ¼ 4:694, a3 ¼ 7:855, a4 ¼ 10:995
The values of the Ritz vectors are given in Table E14.6 and their graphs in
Fig. E14.8.
1. Mass and stiffness matrices
They are obtained from Example 14.3.1.
2 3
6 0 0 0 0 0
60 5 0 0 0 07
6 7
60 0 4 0 0 07
M ¼ m6 7
6 0 0 0 3 0 0 7,
6 7
40 0 0 0 2 05
0 0 0 0 0 1
2 3
232:80 121:78 37:89 7:11 1:21 0:15
6 121:78 122:97 69:99 21:01 3:57 0:44 7
6 7
6 37:89 69:99 72:03 39:04 10:73 1:34 7
K ¼ k66 7:11
7
6 21:01 39:04 38:38 19:04 3:88 7 7
4 1:21 3:57 10:73 19:04 16:00 5:00 5
0:15 0:44 1:34 3:88 5:00 2:12
@CivilMethod
658 PART II Multi-degree-of-freedom systems
e and K
(a) Computation of the matrices M e for two Ritz vectors
In this case, it is
2 3
0:05 0:23
6 0:16 0:59 7
6 7
6 0:34 0:71 7
x¼6
6 0:55 0:42 7
7
6 7
4 0:77 0:22 5
1:00 1:00
Hence
e ¼ x T Mx ¼ m 3:6987 0:8608 ,
M e ¼ x T Kx ¼ k 0:1911 0:3397
K
0:8608 5:7003 0:3397 4:5018
Thus, the reduced eigenvalue problem K e rMe z ¼ 0 becomes
0:1948 0:3305 3:6987 0:8608 z1 0 m
l ¼ , l¼r
0:3305 4:4985 0:8608 5:7003 z2 0 k
Using any of the methods presented in Chapter 12, we obtain l1 ¼ 0:0459,
l2 ¼ 0:7974. Hence
k k 1:0000 0:1257
r1 ¼ 0:0459 , r2 ¼ 0:7974 , z ¼
m m 0:0708 1:0000
The obtained results are given in Table E14.7 in juxtaposition to the exact ones.
Moreover, Tables E14.8 and E14.9 present the results obtained using three and
four Ritz vectors, respectively. We note that the use of four Ritz vectors pro-
duces very good results for the first two eigenfrequencies and mode shapes.
A practical rule is to use 2n Ritz vectors to achieve a good approximation of
n eigenfrequencies and mode shapes.
TABLE E14.7 Results obtained using two Ritz vectors in Example 14.6.1.
Rayleigh-Ritz Exact
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e1 ¼ 0:2143 k=m
w w1 ¼ 0:2030 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e2 ¼ 0:8930 k=m
w w2 ¼ 0:7570 k=m
2 3 2 3
0:0178 0:0950 0:0162 0:0547
6 0:0623 0:2421 7 6 0:0645 0:1747 7
6 7 6 7
6 7 6 0:1469 0:2736 7
e ¼ p1ffiffiffi 6 0:1526 0:2834 7
F F ¼ p1ffiffiffi 6 7
m 6 0:2740 0:1490 7 m 6 0:2640 0:2310 7
6 7 6 7
4 0:4137 0:1346 5 4 0:4133 0:0662 5
0:5639 0:4782 0:5835 0:6010
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 659
TABLE E14.8 Results obtained using three Ritz vectors in Example 14.6.1.
Rayleigh-Ritz Exact
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 1 ¼ 0:2130 k=m
w w1 ¼ 0:2030 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 2 ¼ 0:7787 k=m
w w2 ¼ 0:7570 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 3 ¼ 2:1869 k=m
w w3 ¼ 1:7928 k=m
2 3 2 3
0:0186 0:0548 0:1951 0:0161 0:0547 0:1123
6 0:0618 0:1859 0:2872 7 6 0:0645 0:1747 0:2478 7
6 7 6 7
6 0:1478 0:2910 0:0060 7 6 0:1469 0:2736 0:1305 7
e
F ¼ pffiffiffi
1 6 7 F ¼ p1ffiffiffi 6 7
m 6 0:2680 0:2149 0:2665 7 m 6 0:2640 0:2310 0:2348 7
6 7 6 7
4 0:4145 0:1058 0:1181 5 4 0:4133 0:0662 0:2981 5
0:5765 0:5563 0:3437 0:5835 0:6010 0:4539
TABLE E14.9 Results obtained using four Ritz vectors in Example 14.6.1.
Rayleigh-Ritz Exact
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 1 ¼ 0:2091 k=m
w w1 ¼ 0:2030 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 2 ¼ 0:7648 k=m
w w2 ¼ 0:7570 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 3 ¼ 1:8840 k=m
w w3 ¼ 1:7928 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 4 ¼ 4:0680 k=m
w w4 ¼ 3:3067 k=m
2 3 2 3
0:0171 0:0574 0:1166 0:2800 0:0161 0:0547 0:1123 0:1665
6 0:0623 0:1759 0:2742 0:1070 7 6 0:0645 0:1747 0:2478 0:1858 7
6 7 6 7
6 7 6 7
6 0:2745 7 6 0:1469 0:1863 7
e ¼ p1ffiffiffiffi 6 0:1482
F
0:2796 0:1056 7 ffi6
F ¼ p1ffiffiffi
0:2736 0:1305 7
m6 6 0:2616 0:2340 0:2867 0:0412 77 m6 6 0:2640 0:2310 0:2348 0:1831 77
6 7 6 7
4 0:4110 0:0838 0:2105 0:2419 5 4 0:4133 0:0662 0:2981 0:4140 5
0:5898 0:5784 0:4035 0:2215 0:5835 0:6010 0:4539 0:2807
@CivilMethod
660 PART II Multi-degree-of-freedom systems
y1
1 ¼ 1=2 (14.6.3)
y1 My1
T
The second vector is derived from the vector y2 , which is obtained as the
static displacement vector under the load R ¼ M 1 , that is, the inertial forces
associated with the first Ritz vector. Thus, we have
y2 ¼ K1 M 1 (14.6.4)
The vector y2 , which is not orthogonal to 1 , is mass-orthogonalized to it
using the Gram-Schmidt method presented in Section 12.4. Thus, taking into
account that T1 M 1 ¼ 1, the second of Eq. (12.4.8) gives
e
y2 ¼ y2 yT2 M 1 1 (14.6.5)
which is normalized with respect to mass to give
e
y2
2 ¼ 1=2 (14.6.6)
ey2 Me
T
y2
Similarly, the vector k is obtained as
yk ¼ K1 M k1 (14.6.7)
X
k 1
e
yk ¼ yk ai i (14.6.8)
i¼1
where
ai ¼ yTk M i (14.6.9)
Hence
e
yk
k ¼ 1=2 (14.6.10)
eyk Me
T
yk
The procedure is repeated until we obtain the desired number of Ritz vectors.
The set of vectors 1 , 2 , …, k is orthogonal with respect to mass and linearly
independent, thus satisfying the requirement of the Ritz method. The Ritz
vectors derived by this procedure are also called load-dependent Ritz vectors.
Example 14.6.2 Determine the eigenfrequencies and mode shapes of the chim-
ney in Example 14.3.1 using the Rayleigh-Ritz with (i) two, (ii) three, and (iii)
four derived Ritz vectors.
Solution
The mass and stiffness matrices were computed in Example 14.3.1. We adopt a
load vector proportional to the mass of the structure, that is,
R ¼ m f 6 5 4 3 2 1 gT
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 661
y1
1 ¼ 1=2 ¼ f 0:0036 0:0135 0:0289 0:0491 0:0732 0:0996 gT
y1 My1
T
a1 ¼ yT2 M 1 ¼ 0:1371
e
y2 ¼ y2 a1 1
¼ f 0:0885 0:2408 0:3182 0:2113 0:1351 0:6653 gT 103
which after normalization gives
e
y2
2 ¼ 1=2
e
y2 Me
T
y2
¼ f 0:0136 0:0371 0:0491 0:0326 0:0208 0:1026 gT
a 1 ¼ yT
3M 1 ¼ 0:00649, a 2 ¼ yT
3M 2 ¼ 0:010059
e
y3 ¼ y 3 a 1 1 a2 2
¼ f 0:3343 0:5016 0:05419 0:5528 0:3853 0:9141 g 104
T
e
y3
3¼ T 1=2
ey3 Mey3
T
¼ f 0:03003 0:04510 0:00493 0:04965 0:03476 0:08186 g
@CivilMethod
662 PART II Multi-degree-of-freedom systems
ey4
4¼ 1=2
eT
y4 Mey4
T
¼ f 0:04338 0:01131 0:04671 0:00153 0:06361 0:05861 g
Hence
2 3
0:0200 0:0760 0:1674 0:2418
6 0:0751 0:2069 0:2514 0:0631 7
6 7
1 6 0:1610 0:2733 0:0275 0:2603 7
x ¼ pffiffiffiffiffi 6 7
m6 6 0:2737 0:1816 0:2767 0:0086 7
7
4 0:4080 0:1161 0:1937 0:3545 5
0:5550 0:5715 0:4562 0:3266
Applying the Rayleigh-Ritz method as described in Example 14.6.1, we
obtain the approximate eigenfrequencies and mode shapes. The obtained results
are given in Tables E14.10–E14.12 in juxtaposition to the exact ones.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 663
Rayleigh-Ritz Exact
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 1 ¼ 0:2030 k=m
w w1 ¼ 0:2030 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 2 ¼ 0:7571 k=m
w w2 ¼ 0:7570 k=m
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
e 3 ¼ 1:9777 k=m
w w3 ¼ 1:7928 k=m
2 3 2 3
0:0162 0:0542 0:1761 0:0161 0:0547 0:1123
6 0:0645 0:1754 0:2769 7 6 0:0645 0:1747 0:2478 7
6 7 6 7
6 0:1469 0:2751 0:0643 7 6 0:1469 0:2736 0:1305 7
e
F ¼ pffiffiffi
1 6 7 F ¼ p1ffiffiffi 6 7
m 6 0:2640
6 0:2301 0:2486 7
7
m 6 0:2640 0:2310
6 0:2348 7
7
4 0:4133 0:0687 0:2048 5 4 0:4133 0:0662 0:2981 5
0:5835 0:5979 0:3804 0:5835 0:6010 0:4539
@CivilMethod
664 PART II Multi-degree-of-freedom systems
Kff ust
f + Kfs us ¼ 0 (14.7.4)
Ksf ust
f + Kss us ¼ ps
st
(14.7.5)
Assuming
that the supports do not permit rigid body motion, it is
det Kff 6¼ 0 and Eq. (14.7.4), gives
1
f ¼ Sus , S ¼ Kff Kfs
ust (14.7.6)
which is substituted into Eq. (14.7.5) to yield the support reaction due to the
static application of the displacements us
Ksf S + Kss us ¼ pst
s (14.7.7)
Moreover, if the matrix C is proportional to K, we can readily show that it
also holds
Cff u_ st _s ¼0
f + Cfs u (14.7.8)
Indeed, if C ¼ cK, where c is a constant, Eq. (14.7.8) holds true due to
Eq. (14.7.4). If C is not proportional to K, the contribution of the damping force
to the excitation force does not vanish. However, it is small compared with the
inertial force and it can be neglected.
Substituting Eq. (14.7.3) into Eq. (14.7.1) gives
Mff u€f + Cff u_ f + Kff u
f ¼ Mff u
€st
f Mfs u€s Cff u_ st _s
f + Cfs u
Kff ustf + K u
fs s (14.7.9)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 665
@CivilMethod
666 PART II Multi-degree-of-freedom systems
That is, the vector r represents the static displacements of the free nodes
produced by the vector e us .
When the supports are subjected to a uniform motion ug ðt Þ, as in the
seismic ground motion, the vector of the support displacements can be
written as
us ¼ 1ug ðt Þ (14.7.18)
and inserting it into Eq. (14.7.17) gives
r ¼ S1 (14.7.19)
T
in which 1 ¼ f 1 1 1 ⋯ 1 g is the vector with dimension s. In this case, the
vector r represents rigid body motion, therefore it is
f ¼ 1ug ðt Þ
ust (14.7.20)
Eqs. (14.7.4), (14.7.5) by virtue of Eqs. (14.7.18), (14.7.20) are written as
( )
0
^ g ðt Þ ¼
K1u (14.7.21)
pst
s
where
" #
Kff Kfs
^¼
K (14.7.22)
Ksf Kss
The matrix K^ is the stiffness matrix of the free structure, hence det K
^ ¼0
and ps ¼ 0, because Eq. (14.7.21) has a nontrivial solution. Hence, it is not
st
Example 14.7.1 Formulate the equation of motion of the bridge in Fig. E14.9
when the supports 1 and 2 are subjected to the displacements ug1 ¼ ug ðt Þ and
ug2 ¼ 2ug ðt Þ. The cross-sectional area of the cable is A and its modulus of
elasticity Es while the modulus of elasticity of the remaining structure
is E ¼ 2:1 107 kN=m2 . Assume: I1 ¼ I2 ¼ I3 ¼ 2:055 m4 , I4 ¼ 52 m4 ,
¼ 11:2 kN=m, and L ¼ 50 m. Adopt the lumped
I5 ¼ 31 m4 , Es A ¼ 0:4EI 1 , m
mass assumption.
Solution
By neglecting the axial deformations of members undergoing bending,
the structure has 9 degrees of freedom including those corresponding to
ground motion. The degrees of freedom are shown in Fig. E14.9. Their num-
bering was based on the reasoning to avoid renumbering after static conden-
sation. The lumped mass assumption is adopted. The cable is treated as a
truss member while its sag due to self-weight is neglected. Possible compres-
sion forces are encountered by appropriate prestress. The stiffness matrix
is formulated using the procedure described in Example 11.5.2. Thus,
we obtain
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 667
The displacements of free nodes are written in the form of Eq. (14.7.3), that
is, uf ¼ ust
f +uf , and Eq. (14.7.6) gives the static solution
1 1
ust
f ¼ K ff K fs u g ðt Þ
2
2:01 0:961 1
¼ u g ðt Þ
1:25 1:250 2
0:088
¼ ug ðt Þ
1:250
f is obtained from Eq. (14.7.14), that is
while u
105 0 u€1 201 134 u 1 10
+ 10 3
¼ u€ ðt Þ
0 560 u€2 134 104 u 2 674 g
where
qn ðt Þ ¼ qnst w2n yn ðt Þ (14.8.2)
Eq. (14.8.1) expresses the time history of q ðt Þ. From Eq. (14.8.2), we deduce
that the peak value qn0 of the modal response qn ðt Þ occurs simultaneously with
the extreme value of the pseudoacceleration w2n yn ðt Þ, That is
qn0 ¼ qnst max w2n yn ðt Þ (14.8.3)
t
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 669
which provides an upper bound to jq jmax . Eq. (14.8.5) expresses the absolute
sum (ABSSUM) modal combination rule. It assumes that all modal peak values
occur at the same instant. It gives a too conservative estimate of the peak value.
Therefore, it is not popular in structural design.
Another estimate is obtained from the relation
!1=2
X N X N
jq jmax rnm qn0 qm0 (14.8.6)
n¼1 m¼1
where rmn is the correlation coefficient of the peak modal values qn0 and qm0 . Its
value varies between 0 and 1, with rmn ¼ 1 if m ¼ n. The foregoing equation is
written in matrix form
qffiffiffiffiffiffiffiffiffiffiffiffiffi
jq jmax qT0 rq0 (14.8.7)
where b nm ¼ wm =wn , and rmn ¼ rnm . The above relationship results from a
stochastic process starting from the mean square response, which we do not pre-
sent here because it goes beyond the scope of this book. The interested reader
should consult the related literature, for example, [4]. The method of computing
jq jmax using Eq. (14.8.7) is known as the complete quadratic combination
(CQC) rule. Eq. (14.8.7) is greatly simplified when the eigenfrequencies are
well separated, that is, there are no multiple eigenfrequencies nor are they close
to each other. In this case, the off-diagonal elements of the correlation matrix,
rmn ðm 6¼ n Þ, are negligible and we can set r I. Then Eq. (14.8.7) becomes
qffiffiffiffiffiffiffiffiffiffiffi
jq jmax qT0 q0 (14.8.9)
or
!1=2
X
N
jq jmax 2
qn0 (14.8.10)
m¼1
@CivilMethod
670 PART II Multi-degree-of-freedom systems
Solution
The eigenfrequencies and mode shapes were computed in Example 14.3.1.
The peak modal values will be computed using Eq. (14.8.4).
st st st
Computation Qbn , Mbn , and u6n
From Example 14.3.1, we obtain for m ¼ 31:056 kN m1 s2
R1 R2 R3 R4 R5 R6
9:66 22:98 33:23 36:65 37:32 46:49
32:17 61:18 61:09 34:10 3:02 36:27
58:57 76:65 25:75 27:33 30:34 20:92
78:94 48:54 34:75 20:15 29:41 8:86
82:39 9:28 29:40 30:37 14:71 2:75
58:17 42:08 22:39 10:30 3:39 0:49
st st st
The quantities Qbn , Mbn , u6n are computed from the relations (see Example
14.3.1)
Gn
st
Qbn ¼ 1T Rn , Mbn
st
¼ hT Rn , u6n
st
¼ f
w2n 6n
The computed values are given in Table E14.13.
st st st
TABLE E14.13 Modal quantities Qbn , Mbn , u6n in
Example 14.8.1.
Mode n st
Qbn ¼ 1T Rn st
Mbn ¼ hT Rn st
u6n ¼ Gwn2 f6n
n
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 671
The peak value max t jyn ðt Þj is obtained from the solution of Eq. (14.3.16)
by setting f ðt Þ ¼ u€g ðt Þ, where u€g ðt Þ is given by the accelerogram of the spec-
ified earthquake. The computation of Spa ðx n , Tn Þ is performed numerically
using any of the methods described in Section 6.2. The computed values are
given in Table E14.14 together with the modal values of Qb , Mb , u6 .
Computation of the correlation matrix
Eq. (14.8.8) is used to compute the correlation coefficients. For the data of
the problem, it gives
3
r ¼ 102 3
1000:0 3:53 0:716 0:262 0:128 0:067
6 3:53 1000:0 7:34 1:91 0:85 0:423 7
6 7
6 0:716 7:34 1000:0 15:6 4:47 1:88 7
66 7
0:262 1:91 15:6 1000:0 29:1 7:15 7
6 7
4 0:128 0:85 4:47 29:1 1000:0 33:8 5
0:067 0:423 1:88 7:15 33:8 1000:0
Table E14.15 presents the peak values jQb jmax , jMb jmax , and ju6 jmax as com-
puted using the three modal combination rules in juxtaposition with the exact
values obtained by the response history analysis (RHA) using the relation
X 6
2
jq jmax ¼ max qn wn yn ðt Þ
st
t
n¼1
TABLE E14.14 Modal quantities Qbn0 , Mbn0 , and u6n0 in Example 14.8.1.
@CivilMethod
672 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 673
e ¼I
general diagonal, except when the Ritz vectors are load-dependent. Then M
because the Ritz vectors are orthonormalized with respect to mass.
According to Eqs. (14.3.12)–(14.3.14), we can write
X
K
e
pðt Þ ¼ f ðt Þ en M
Rn , Rn ¼ G en ¼
where G nR
T
n (14.9.1)
n¼1
The number K of the Ritz vectors that should be used in dynamic analysis
must yield a load e pðt Þ sufficiently close to pðt Þ. Hence, according to
P
K
Eq. (14.3.42), the deviation of Rn from R is
n¼1
X
K
e ¼R Rn
n¼1
(14.9.2)
X
K
¼R en M
G n
n¼1
@CivilMethod
674 PART II Multi-degree-of-freedom systems
FIG. 14.9.1 Error eK as a function of the number K of mode shapes fn and Ritz vectors c n .
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 675
@CivilMethod
676 PART II Multi-degree-of-freedom systems
^1 un k
^ n + 1 ¼ pn k
Ku ^2 un1 , n ¼ 0, 1, 2, … (14.10.7)
where
K^ ¼ 1 M+ 1 C (14.10.8a)
Dt 2 2Dt
^1 ¼ K 2 M
k (14.10.8b)
Dt 2
^2 ¼ 1 M 1 C
k (14.10.8c)
Dt 2 2Dt
1
€0
u1 u0 Dt u_ 0 + Dt 2 u (14.10.9)
2
in which u0 , u_ 0 are known from the initial conditions while u
€0 is obtained from
the equation of motion for t ¼ 0
€0 ¼ M1 p0 Cu_ 0 Ku0
u (14.10.10)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 677
1
€ ðt + τ Þ ¼ ½ u
u € ðt Þ + u
€ðt + Dt Þ 0 τ Dt (14.10.11)
2
un + 1 ¼ Du + un (14.10.12a)
u_ n + 1 ¼ Du_ + u_ n (14.10.12b)
€n + 1 ¼ D€
u €n
u+u (14.10.12c)
@CivilMethod
678 PART II Multi-degree-of-freedom systems
_ D€
The difference vectors Du, u are computed from the relations
2
Du_ ¼
Du 2u_ n (14.10.13)
Dt
4 4
u ¼ 2 Du u_ n 2€
D€ un (14.10.14)
Dt Dt
while Du from
^
KDu ¼ D^
pn (14.10.15)
where
^ ¼K+ 2 C+ 4 M
K (14.10.16)
Dt Dt 2
and
pn ¼ Dpn + ^cu_ n + m€
D^ ^ un (14.10.17)
in which
4
^c ¼ ^ ¼ 2M, Dpn ¼ pn + 1 pn
M + 2C, m (14.10.18)
Dt
The AAM, contrary to CDM, is unconditionally stable. Therefore, the time
step Dt can be chosen arbitrarily. Its size, however, is influenced by the accu-
racy of the method and its capability to describe an oscillatory motion. There-
fore, it must be small enough. The selection of Dt equal to 1/10 of the smallest
natural period of the system produces accurate results. Table 14.10.2 presents
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 679
@CivilMethod
680 PART II Multi-degree-of-freedom systems
Solution
The mass and stiffness matrices, eigenfrequencies, and mode shapes are taken
from Example 14.3.1, which for the given values of k and m give.
2 3
186:336 0 0 0 0 0
60 155:280 0 0 0 0 7
6 7
60 0 124:224 0 0 0 7
M¼6 6 7
0 0 0 93:168 0 0 7
6 7
40 0 0 0 62:112 0 5
0 0 0 0 0 31:056
2 3
1277189:2 668110:4 207872:4 39006:9 6638:3 822:9
6 668110:4 674639:0 383979:7 115265:2 19585:7 2413:9 7
6 7
6 207872:4 383979:7 395171:5 214181:5 58867:0 7351:5 7
6 7
K¼6 7
6 39006:9 115265:2 214181:5 210560:6 104457:4 21286:4 7
6 7
4 6638:3 19585:7 58867:0 104457:4 87779:3 27431:0 5
822:9 2413:9 7351:5 21286:4 27431:0 11630:7
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 681
2 3
0:0028 0:0098 0:0201 0:0298 0:0377 0:0503
6 0:0115 0:0313 0:0444 0:0333 0:0036 0:0471 7
6 7
6 0:0263 0:0491 0:0234 0:0334 0:0460 0:0339 7
F¼6
6 0:0473
7
6 0:0414 0:0421 0:0328 0:0595 0:0191 7 7
4 0:0741 0:0118 0:0534 0:0742 0:0447 0:0089 5
0:1047 0:1078 0:0814 0:0503 0:0205 0:0031
The numerical methods require the damping matrix. This is constructed as a
proportional damping matrix from the eigenfrequencies, eigenmodes, and
modal damping ratios. Thus applying Eq. (12.11.40), we obtain
C ¼ MFCF^ TM
2 3
1152:72 396:84 59:28 6:35 1:97 0:19
6 396:84 657:43 289:32 2:82 1:69 7
6 37:54 7
6 7
6 59:28 289:32 441:10 192:30 23:00 0:12 7
¼66
7
6 6:35 37:54 192:30 277:56 113:97 13:23 77
6 7
4 1:97 2:82 23:00 113:97 148:39 44:83 5
0:19 1:69 0:12 13:23 44:83 35:59
The graphical representation of Qb ðt Þ, Mb ðt Þ, and u6 ðt Þ, obtained using the
three numerical methods with Dt ¼ 0:01, are shown in Fig. E14.10. Apparently,
the obtained results are graphically identical.
@CivilMethod
682 PART II Multi-degree-of-freedom systems
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 683
C. In each step
1. Compute: ðkT Þn ¼ d ðfSi Þ=du j n ,
2. Compute: k^n ¼ ðkT Þ + 2 C + 4 2 M, Dp ¼ pn + 1 pn
n Dt Dt
^n ¼ Dp + ^cu_ n + m€
3. Compute: p ^ un
^n and p
4. Compute: Du from k ^n using the modified Newton-Raphson method
(see Table 14.11.2)
5. Compute: Du_ ¼ Dt 2
Du 2un
6. Compute: un + 1 ¼ un + Du, u_ n + 1 ¼ u_ n + Du_
u€n + 1 ¼ M1 ½pn + 1 Cu_ n + 1 f S ðun + 1 Þ
7. Set n ¼ n + 1 and check:
If tn ttot end. Else set un ¼ un + 1 , u_ n ¼ u_ n + 1 , u_ n ¼ u_ n + 1 and go to step C.1.
This equation represents a special case of Eq. (14.10.1), that is, only when the
stiffness vector is a nonlinear function of the displacements while the damping
vector is linear. The steps of the method are given in Table 14.11.1. Moreover,
Table 14.11.2 shows the steps for the modified Newton-Raphson method as
applied to MDOF systems. Following these steps, a computer program called
av_acc_nlin_MDOF.m has been written in MATLAB for the numerical integra-
tion of the nonlinear equations of motion using the average acceleration method.
The program list is available on this book’s companion website.
A. Data
ð0Þ ð0Þ
1. Read: un + 1 ¼ un , f S ¼ f S ðun Þ, DRð1Þ ¼ D^
pn
B. For each iteration Compute i ¼ 1,2,3…
1. duðiÞ ¼ k ^1 DRðiÞ
n
ði Þ ði1Þ
2. un + 1 ¼ un + 1 + duðiÞ h i
ði Þ ði Þ
3. Dp ¼ f S f S + k
ði1Þ ^n ðkT Þ duðiÞ
n
4. DRði + 1Þ ¼ DRðiÞ DpðiÞ
5. Compute ^1 DRði + 1Þ
duði + 1Þ ¼ k
n
6. If duði + 1Þ =DuðnI Þ > a Set i ¼ i + 1 and go to step Β.2
@CivilMethod
684 PART II Multi-degree-of-freedom systems
The solution procedure is similar to that for the linear systems. Thus,
Eq. (14.11.2a) for t ¼ 0 gives the initial acceleration vector
q0 ¼ M1 ½p0 Fðu_ 0 , u0 Þ, q0 ¼ u
€ (14.11.3)
Subsequently, we apply Eq. (14.11.2a) for t ¼ tn
Mqn + Fðu_ n , un Þ ¼ pn (14.11.4)
Apparently, the second and third of Eqs. (14.10.19a), (14.10.19b) are valid
in this case too, and can be written as
2 c 3 2 c 3
1
I I
1
hI I u_ n 0 I u_ n1 6 2 7 6 2 7
¼ + 4 c 5qn + 4 c 5qn1
I 0 un I 0 un1 I
2 2
I
2 2
(14.11.5)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 685
Solution
The system has two degrees of freedom, Fig. E14.11b. The equations of
motion are formulated using the method of equilibrium of forces. The forces
acting on the mass m are shown in Fig. E14.11c. The equilibrium of forces
gives
m u€ + S1 sin q1 + S2 sin q2 ¼ px (1a)
m v€ S1 cos q1 + S2 cos q2 ¼ py mg (1b)
The elastic forces S1 and S2 are obtained from the relations
l 1 l0
S1 ¼ EA (2a)
l0
l 2 l0
S2 ¼ EA (2b)
l0
@CivilMethod
686 PART II Multi-degree-of-freedom systems
where
u v d0
x¼ , y¼ , d0 ¼
l0 l0 l0
Substituting Eqs. (3a), (3b) into Eqs. (2a), (2b) gives
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
S1 ¼ EA ð1 + d 0 y Þ + x 1 2 (4a)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
S2 ¼ EA ð1 + d 0 + y Þ + x 1 2 (4b)
Moreover, we have
l v 1 + d0 y
cos q1 ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5a)
l1
ð1 + d 0 y Þ 2 + x 2
u x
sin q1 ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5b)
l1
ð1 + d 0 y Þ2 + x 2
l v 1 + d0 + y
cos q2 ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5c)
l2
ð1 + d 0 + y Þ2 + x 2
u x
sin q2 ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (5d)
l2
ð1 + d 0 + y Þ2 + x 2
Finally, substituting Eqs. (4a) and (4b), (5a)–(5d) into Eqs. (1a), (1b) yields
the equations of motion
2 3
6 1 1 7
ml 0 x€ + EAx 42 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi5 ¼ px (6a)
2 2
x 2 + ð1 + d 0 y Þ x 2 + ð1 + d 0 + y Þ
2 3
6 1 + d0 y 1 + d0 + y 7
ml 0 y€ + EA42y + qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi5 ¼ py mg
x 2 + ð1 + d 0 y Þ2 x 2 + ð1 + d0 + y Þ2
(6b)
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 687
Eqs. (6a), (6b) are solved numerically using the AAM and AEM methods.
Fig. E14.12 shows the graphs of u ðt Þ and v ðt Þ while Fig. E14.13 shows
the graphs of S1 ðt Þ and S2 ðt Þ. The results are graphically identical. The mini-
mum value of d0 is obtained iteratively until min ðS1 , S2 Þ ¼ 0. This gives
min d0 ¼ 16:5 cm.
FIG. E14.12 Time history of the displacements u(t) and v(t) in Example 14.11.1.
FIG. E14.13 Time history of the elastic forces S1(t) and S2(t) in Example 14.11.1.
@CivilMethod
688 PART II Multi-degree-of-freedom systems
14.12 Problems
Problem P14.1 The television tower modeled by the system in Fig. P14.1 is
subjected (i) to the wind pressure of Fig. E14.6 with p0 ¼ 2 kN=m2 ,
t1 ¼ 0:5 s and (ii) to the seismic ground motion ug ðt Þ ¼ 2 sin 5t. The column
is approximated by three constant elements of length l ¼ h=3. Adopting the
lumped assumption, derive the equation of motion of the structure and analyze
its response using the mode superposition method. Give the graphs of the time
history of the displacements, the base shear force Qb and overturning moment
Mb . Compute the peak value of the shear force and the bending moment at the
base cross-section of the column as well as at the cross-sections above and
beneath the body B representing a rotating restaurant. The ground is elastic.
The reaction moment of the elastic ground is represented by the expression
MR ¼ CR f, where CR ¼ KI f ; If is the moment of inertia of the planform of
the fundament and K ¼ E=10h the foundation modulus with E being the mod-
ulus of elasticity of the material of the structure. The cross-sections of the flex-
ible column, the planform of the fundament, and the body B are circular with
diameters D, Df ¼ 8D, and DB ¼ 5D, respectively. The specific weight of the
material is g. The fundament and the body B are assumed rigid. Data h ¼ 60 m,
D ¼ 2 m, E ¼ 2:1 107 kN=m2 , g ¼ 24 kN=m3 .
Problem P14.2 The shear frame in Fig. P14.2a is modeled as shown in
Fig P14.2b. Compute the modal contribution factor of (i) the base shear
force Qb , (ii) the overturning moment Mb , and (iii) the top displacement u4
when the excitation is due to ground motion ug ðt Þ ¼ uo sin w t. Assume:
m ¼ 2:0 kN m1 s2 =m; a ¼ 3:0 m; columns cross-sections: first and second
floor 0:30 0:30 m2 , third and fourth floor 0:25 0:25 m2 ; modulus of elastic-
ity E ¼ 2:1 107 kN=m2 .
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 689
(a) (b)
FIG. P14.2 Shear frame in problem P14.2 (a); dynamic model (b).
Problem P14.3 For the shear frame of Problem P14.2, derive Ritz vectors using
(i) the method of the natural mode shapes and (ii) the method of load dependent
Ritz vectors, and use them to approximate the eigenfrequencies. Then reduce
the number of equations of motion to two.
Problem P14.4 For the shear frame of Problem P14.2, compute the modal
magnification factors Dn and examine whether the static correction method
is recommended for application.
Problem P14.5 For the shear frame of Problem P14.2, compute the peak modal
values Qbn0 , Mbn0 , and u4n0 of the base shear force, overturning moment, and
top displacement, respectively. Then use the ABSSUM, CQM, and SRSS
methods to compute the peak values of these quantities. Compare them with
those resulting from the RHA (Response History Analysis).
Problem P14.6 The two one-story buildings of Fig. P14.6 are connected by a
beam as shown in Fig. P14.6. Formulate the equations of motion. Then compute
the truncation error as a function of the number of employed eigenmodes.
Use both the natural mode shapes and derived Ritz vectors and compare the
computed error. Data: column cross-sections 0:30 0:30 m2 , beam cross-
section 0:20 0:30 m2 , h ¼ 3:5 m, a ¼ 4:0 m, b ¼ 4:0 m, q ¼ 20 kN=m2 ,
b ¼ p=6, and F ðt Þ ¼ 2 cos 5t.
Problem P14.7 Consider the two-bar system of Fig. P14.7. While the system is
in static equilibrium under the gravity load W ¼ mg, the mass m is given the
additional displacements u0 , v0 . Compute the time history of the displacements
u ðt Þ,v ðt Þ in the horizontal and vertical directions, respectively, as well as the
axial forces S1 ðt Þ, S2 ðt Þ of the bars and give their graphs. Use the numerical
methods CDF, AAM, and AEM to solve the nonlinear equations of motion.
Assume: A ¼ 3:14 cm2 , l1 ¼ 4:0 m, l2 ¼ 3:0 m, m ¼ 10 kN m1 s2 ,
g ¼ 9:81 m=s2 , and E ¼ 2 108 kN=m2 . The bars are assumed massless.
Problem P14.9 The shear frame of Fig. P14.9 is supported on an elastic foun-
dation that is modeled by two nonlinear springs producing the force
FT ¼ CT ðu + u 2 =4a Þ and the moment MR ¼ CR f + f2 =4 , where
CT ¼ EI =10a 3 and CR ¼ EI =5a. Formulate the equations of motion of the
structure and determine the time history of the shear forces and bending
moments of the columns. Then compute their peak values. Data:
¼ 2 kN m1 s2 =m, a ¼ 10 m, h1 ¼ 5h2 =3, h2 ¼ 3:0 m; g ¼ 9:81 m=s2 ,
m
E ¼ 2 107 kN=m2 ; cross-section area of columns k1 0:30 0:30 m2 and k2
0:25 0:25 m2 ; p1 ðt Þ ¼ 5H ðt Þ, and p2 ðt Þ ¼ 8H ðt Þ.
@CivilMethod
Multi-degree-of-freedom systems: Forced vibrations Chapter 14 691
Problem P14.10 The steel mast of Fig. P14.10a is spherically hinged at point O
and is supported by four massless cables of cross-section A. At the top of the
mast is a steel box of side a and thickness d1 . The mast consists of a circular
steel tube with thickness d2 . Its cross-section varies linearly as shown in the fig-
ure. The external diameters at the cross-sections are D1 at the top C, D2 at B, and
D3 at the base O. The planform of the structure is shown in Fig. P14.10b. The
box is subjected to a wind blast of intensity pðt Þ in the direction b with respect to
the x axis, whose time variation is given in Fig. E14.6. The cables are pre-
stressed to undertake compressive forces. Considering small displacements,
determine the response of the structure. Data: L ¼ 60:0 m, a ¼ 5:0 m,
d1 ¼ d2 ¼ 2 cm, D1 ¼ D3 ¼ 1:0 m, D2 ¼ 1:5 m, p0 ¼ 2:0 kN=m2 , b ¼ p=6,
A ¼ 3:14 cm2 , E ¼ 2 108 kN=m2 , and g ¼ 9:81 m=s2 .
(a) (b)
FIG. P14.10 Steel mast in problem P14.10 (a); planform (b).
@CivilMethod
692 PART II Multi-degree-of-freedom systems
@CivilMethod
Chapter 15
15.1 Introduction
The load-bearing systems of modern multistory buildings are skeletal structures
made of steel, RC (reinforced concrete), or a combination of them (composite struc-
tures). The load transfer path is from slabs (plates) to beams, from beams to col-
umns, and from columns to the foundation, which may consist of individual
footings, strip footings, and raft foundations. The first modeling of a building is
the shear building. It is a simple model that is used to approximate the dynamic
response of a building. According to this model, the beam-reinforced slabs are
encountered as plane rigid bodies. The obtained response is acceptable when the
beam-slab system is very stiff, which is a usual case for buildings of reinforced con-
crete. However, in the beam-slab system, the flexible slabs are connected with the
beams either by various types of connectors, as in composite structures, or rigidly,
as in concrete structures, and they cooperate in carrying the live and dead loads. The
deformation of the slab within its plane is very small and can be neglected. Thus,
the functioning of the slab in the horizontal motion can be simulated with that of a
plane rigid body, called diaphragm. This functioning constrains the horizontal dis-
placements and the rotations about the vertical axis at the ends of the columns as
well as the axial deformation of the beams. On the other hand, this approximation
enforces the cooperation of slabs and beams in the vertical direction.
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 695
Several professional codes have been developed for the analysis of buildings
and in general of structures, such as ETABS, SAP2000, CSiBridge, STAAD,
MIDAS, etc. These codes can be used by professional engineers to efficiently
analyze structures made from various materials. However, the engineers must
be aware of the assumptions adopted by the code. It is highly essential that
the user of the code can check the correctness of the obtained results. This
presupposes a deep understanding of the static and dynamic response of the
structures. Therefore, it is recommended that these codes are employed only
by experienced engineers who master the static and dynamic structural analysis.
Effective breadth
@CivilMethod
696 PART II Multi-degree-of-freedom systems
The vertical elements are rigidly connected with the plates so that the actions
(forces and moments) can be transferred from the plates to the vertical elements
and vice versa.
The vertical elements may connect two or more, or even all, slabs of the
structure. They may be fixed, hinged, or elastically supported on the ground.
These elements will be referred to as multistory elements (MSE). The MSEs
may be columns, frames, walls, isolated or framed with beams, closed sections,
staircase cores, etc. The directions of the principal axes of the MSE may vary
from floor to floor. Nevertheless, their treatment is easier if their principal
directions are the same along the height.
Node
Node
Node
Node
Node
FIG. 15.2.2 Displacements and elastic forces at the node j of the i MSE.
The MSE is deformed due to the deformation of the structure. Its deforma-
tion can be determined by the nodal displacements, which are three for each
node, two translational in the directions of the x,y axes and one rotational about
the z axis. The displacements and elastic forces at the j node of i MSE will be
denoted by uji , vji , wij and Xji , Yji , Mji in the local axes and by u ij , v ij , w
ij and
i i i
X j , Y j , M
in the global axes, respectively.
j
Thus we define the following matrices for the displacements and elastic
forces at the j node of i MSE:
In the local axes n oT
Dij ¼ uji vji wij (15.2.1a)
T
FiSj ¼ Xji Yji Mji (15.2.1b)
In global axes n oT
i ¼ u i v i w
D i
(15.2.2a)
j j j j
n oT
i ¼ X i
F
i
Y j i
M (15.2.2b)
Sj j j
We assume that the local axes of the MSE are the same on all floors. The
transformation matrix from the global system of axes to the local one is
2 3
cos fi sin fi 0
Rij ¼ Ri ¼ 4 sin fi cos fi 0 5 (15.2.3)
0 0 1
@CivilMethod
698 PART II Multi-degree-of-freedom systems
i
FiSj ¼ Ri F (15.2.5a)
Sj
i T
i
F ¼ R FiSj (15.2.5b)
Sj
Subsequently, we formulate the total vector of displacements of the MSE
in the local and global systems of axes. To be consistent with Fig. 15.2.2,
we place the displacements in the corresponding vectors starting with the n
node and ending with the first, namely
8 i9
>
> D1 > >
< Di >
> =
Di ¼ 2
, (15.2.6a)
> >
>⋮ >
>
: >
;
Din
8 i9
>
> D >
> 1>>
< i >
> =
i
¼ D
D 2 (15.2.6b)
>
> >
>⋮ >
> >
>
: i ;
Dn
Di ¼ R i
^ iD (15.2.7a)
i T i
i ¼ R
D ^ D (15.2.7b)
where
2 3
Ri 0 ⋯ 0
60 Ri ⋯ 0 7
^i ¼ 6
R 6
7
7 (15.2.8)
4⋮ ⋮ ⋱ ⋮ 5
0 0 0 Ri
By the same reasoning, we define the total vectors of the nodal elastic forces
8 i 9
>
> FS1 >
>
< Fi >
> =
i S2
FS ¼ , (15.2.9a)
>⋮ >
> >
>
: i ; >
FSn
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 699
8 i 9
>
> >
F
> S1 >
> i >
< >
=
i ¼ FS2
F (15.2.9b)
S
>⋮ >
> >
>
> >
: i > ;
F Sn
ð15:2:11Þ
or
FiS ¼ ki Di (15.2.12)
i
The matrix k with dimensions N N ðN ¼ 3n Þ is the stiffness matrix of the
i MSE with respect to its nodal displacements. In global axes, this matrix is
transformed as
i ¼ k
F i D
i (15.2.13)
S
where
i T i i
i ¼ R
k ^ kR ^ (15.2.14)
The plates undergo translational displacements in the direction of the
y axes and rotations about the z axis. We denote the displacement vector
x,
of the j plate with respect to the O xyz with
8 9
< U j =
Uj ¼ V j , j ¼ 1, 2, …,n (15.2.15)
: ;
Wj
@CivilMethod
700 PART II Multi-degree-of-freedom systems
where
2 3
1 0 0
Tji ¼ 4 0 1 0 5 (15.2.18)
y ij xij 1
is the transformation matrix, in which xij , yij represent the coordinates of the
point i referred to the system of axes with origin the point Oj (see
Section 11.11).
Eqs. (15.2.17) are combined as
i ¼ TT U
D (15.2.19)
i
where
2 3
T1i 0 ⋯ 0
60 T2i ⋯ 0 7
Ti ¼ 6
4⋮
7 (15.2.20)
⋮ ⋱ ⋮ 5
0 0 ⋯ Tni
is the transformation matrix of the i MSE due to the transfer.
Similarly, the elastic forces are transformed to point Oj according to
Eq. (15.2.17)
oi ¼ Tj F
F i , j ¼ 1, 2, …,n
Sj i Sj
or
oi ¼ Ti F
F i (15.2.22)
S S
F oi U
oi ¼ k (15.2.23)
S
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 701
where
oi ¼ Ti k
k i TT (15.2.24)
i
where
2 3
mji 0 0
6 mji 0 7
ij ¼ 4 0
m 5, j ¼ 1, 2, …,n (15.2.26)
0 0 Iji
Eqs. (15.2.25) are combined to
i
i ¼ m
F €
iD (15.2.27)
I
where
8 i 9 2 3
> >
F mi1 0 ⋯ 0
> Ii 1 >
> >
< = 60 mi2 ⋯ 0 7
i ¼ F 6 7
F I
I2 , i ¼6
m 7 (15.2.28)
>⋮ >
> > 4⋮ ⋮ ⋱ ⋮ 5
: i >
> ;
F In
0 0 ⋯ min
where
oi ¼ Ti m
m i TTi (15.2.30)
If mj is the mass of the j plate and Ijc its moment of inertia about the vertical
axis with respect to the center of mass Cj , then the inertial force of the j plate is
c ¼ M
F €
cU
c
Ij j j , j ¼ 1, 2, …,n (15.2.31)
where
2 3
mj 0 0
c ¼6
M
7
4 0 mj 0 5, j ¼ 1, 2, …,n (15.2.32)
j
0 0 Ijc
@CivilMethod
702 PART II Multi-degree-of-freedom systems
c ¼ M
F €
cU
c (15.2.33)
I
where
8 c 9 2 c 3 8 c 9
>
> FI 1 >
> M1 0 ⋯ 0 >
> U1 >>
>
<F c => 60 c ⋯ 0 7 >
<Uc >=
c ¼ I2 c ¼6 M 7 c
F I , M 6 2
7, U ¼ 2
(15.2.34)
>
> ⋮ >
> 4⋮ ⋮ ⋱ ⋮ 5 >
> ⋮ > >
>
: c ; > : c>
> ;
FIn 0 0 c
⋯ M
U
n n
which
2 3
T1c 0 ⋯ 0 2 3
6 1 0 0
c TT , Tc ¼ 6 0
o ¼ Tc M T2c ⋯ 0 7 7 , Tj ¼ 4 0 1 0 5
M 4⋮ (15.2.36)
c
⋮ ⋱ ⋮ 5 c
ycj xcj 1
n
0 0 ⋯ Tc
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 703
M € +K
U
U ¼P
(15.2.42)
where
XK
¼M
M o+ oi
m (15.2.43a)
i¼1
X
K
¼
K oi
k (15.2.43b)
i¼1
¼ Po
P (15.2.43c)
The modelling of buildings using the concept of the MSE is not recommended
for the engineering praxis where the analysis is performed using commercial
computer codes. However, it can be used for educational purposes to analyze sim-
ple buildings aiming at understanding their dynamic response. From the examples
presented below we conclude that method of MSEs provides good results.
Example 15.2.1 Formulate the equation of motion of the one-story building of
Fig. E15.1 with respect to the center of mass O of the slab for the ground motion
ug ðt Þ in the direction b. Assume:
(a) The columns k1 , k2 , k3 , k4 have a square cross-section ða a Þ with
moments of inertia Ix ¼ Iy ¼ I .
(b) The beams b1 , b2 have a rectangular cross-section ða2a Þ, hence a
moment of inertia Ib ¼ 8I . The effective breadth of the beams as well as
their torsional stiffness are neglected.
(c) Shear modulus G ¼ 0:40E and mass of the slab r per square meter.
@CivilMethod
704 PART II Multi-degree-of-freedom systems
2 3
It contributes to the stiffness of the kx 0 0
structure with its translational k ¼ 0 0 05
2 4
stiffness kx along the local axis x. 0 0 0
2 3
It contributes to the stiffness of the 0 0 0
structure only with its torsional k ¼ 40 0 0 5
3
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 705
2 3
It contributes to the stiffness of the kx 0 0
structure with its translational k5 ¼ 4 0 ky 0 5
stiffnesses kx and ky along the local 0 0 kw
axes x and y as well as with the
torsional stiffness kw .
2 3
It contributes to the stiffness of the 0 0 0
structure with its translational k ¼ 0 ky 0 5
6 4
stiffnesses ky along the local axes y 0 0 kw
and with the torsional stiffness kw .
The translational stiffness along the
local axis x has been considered in
MSE 1.
Step 2
Formulation of the element transformation matrices Ri due to rotation of the local
axes of the MSE. From Fig. E15.1, we have f1 ¼ f3 ¼ f4 ¼ f5 ¼ f6 ¼ 0,
f2 ¼ p=2. Hence
2 3 2 3
1 0 0 0 1 0
R ¼ R ¼ R ¼ R ¼ R ¼ 4 0 1 0 5, R ¼ 4 1 0 0 5
1 3 4 5 6 2
0 0 1 0 0 1
Step 3
i of the ith element depend on the
Because the plate is rigid, the displacements D
displacements U of the point O. Hence the relation that connects them results
from Eq. (11.11.13) by setting uJ ¼ U, uj ¼ D
i , that is
i ¼ TT U
D (1)
i
where
2 3
1 0 0
Ti ¼ 4 0 1 0 5 (2)
y i xi 1
The transformation matrix is identical to that of the plane frame
(see Table 11.11.1).
Referring to Fig. E15.1 and using Eq. (2), we obtain
2 3
1 0 0
x1 ¼ 0, y1 ¼ L=2, T1 ¼ 4 0 1 05
0:5L 0 1
2 3
1 0 0
x2 ¼ L, y2 ¼ 0, T2 ¼ 4 0 1 0 5
0 L 1
@CivilMethod
706 PART II Multi-degree-of-freedom systems
2 3
1 0 0
x3 ¼ L, y3 ¼ L=2, T3 ¼ 4 0 1 05
0:5L L 1
2 3
1 0 0
x4 ¼ L, y4 ¼ L=2, T4 ¼ 4 0 1 05
0:5L L 1
2 3
1 0 0
x5 ¼ L, y5 ¼ L=2, T5 ¼ 4 0 1 05
0:5L L 1
2 3
1 0 0
x6 ¼ L, y6 ¼ L=2, T6 ¼ 4 0 1 05
0:5L L 1
Step 4
Formulation of the local stiffness matrices ki of the elements and total stiffness
of the building.
matrix K
MSE 1 (frame k1 b1 k4 )
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 707
kx is the translational stiffness of the frame of Fig. b. The stiffness matrix of the
frame with respect to its three degrees of freedom is
2 3
24 6L 6L
EI
k2 ¼ 3 4 6L 36L2 16L2 5
L
6L 16L2 36L2
and after the static condensation of the rotational degrees of freedom u2 , u3
2 3
22:61 0 0
EI
k2 ¼ 3 4 0 0 05
L
0 0 0
MSE 3, MSE 4, MSE 5, MSE 6
The stiffnesses kx , ky , kw of elements are computed from the relations
kx ¼ 3EI y =L3 , ky ¼ 3EI x =L3 , kw ¼ GI t =L
For a square cross-section a a it is: Ix ¼ Iy ¼ a 4 =12, It ¼ 0:140a 4 ¼ 1:68I .
Hence
2 3 2 3
0 0 0 3 0 0
EI EI
k3 ¼ 3 4 0 0 0 5, k4 ¼ 4 0 0 0 5
L 2 L3 2
0 0 0:672L 0 0 0:672L
2 3 2 3
3 0 0 0 0 0
EI 4 EI
5, k ¼ 4 0 3 0 5
k ¼ 3 0 3 0
5 6
L 2 L3 2
0 0 0:672L 0 0 0:672L
of the building with respect to point O is computed
The stiffness matrix K
from the relation
X
6 T
o ¼
K T i Ri k i Ri ð T i Þ T
i¼1
2 3
27:43 0 7:71L
EI 6 7
¼ 4 0 28:61 16:61L 5
L3
7:71L 16:61L 38:16L2
Step 5
Formulation of the mass matrix M of the building with respect to point O.
Because it is assumed that only the mass of the plate is taken into account,
the mass matrix will result from the inertial properties of the plate. Given that
point O coincides with the mass center of the plate, we have
m ¼ 3:36L2 r
ð2:4LÞ ð1:4LÞ3 ð2:4LÞ3 ð1:4LÞ
Io ¼ r + r ¼ 2:16L4 r
12 12
@CivilMethod
708 PART II Multi-degree-of-freedom systems
Hence
2 3
3:36 0 0
o ¼ rL2 4 0
M 3:36 0 5
2
0 0 2:16L
Step 6
The equation of motion reads
U
M €
+K
U ¼ Mb
u€g ðt Þ
where U is the vector of the relative displacement with respect to the ground
and b ¼ f cos b sin b 0 gT .
For a ¼ 0:30 m, L ¼ 3 m, E ¼ 2:1 107 kN=m2 , r ¼ 0:4 kN m1 s2 =m2
we obtain
2 3
14400:75 0 12143:25
K ¼4 0 15020:25 26160:75 5
12143:25 26160:75 180306:0
2 3
12:09 0 0
M ¼4 0 12:09 0 5
0 0 69:98
FIG. E15.2 Space frame modeling of the one-story building in Example 15.2.1.
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 709
Table E15.1 shows the eigenfrequencies obtained using the MSE modeling
and SAP2000. In both models the torsional stiffness of the beams has been
neglected.
Example 15.2.2 Formulate the equation of motion of the two-story building shown
in Fig. E15.3 with respect to the origin of the global axes O (center of column 1)
when it is subjected to the ground motion ug ðt Þ in the x direction and evaluate
the eigenfrequencies and eigenmodes. The building is made from reinforced con-
crete with material constants E ¼ 2:1 107 kN=m2 , n ¼ 0:25. The thickness of the
slabs is 15 cm, the cross-sectional area of the columns is 40 40 cm2 , and of the
beams 30 50 cm2 . The columns are assumed fixed at the ground. The loads of the
slabs are: 5 kN=m2 live load and 1 kN=m2 covering.
0:30 0:503
Ib ¼ ¼ 31:25 104 m4
12
a. Equation of motion
Step 1
Determination of the MSEs of the structure. Referring to Fig. E15.3, we distin-
guish six MSEs of the frame type, three in the direction of the x axis and three in
the direction of the y axis. The columns are treated as additional MSEs contrib-
uting with their torsional stiffness because their translational stiffnesses are
included in the frames. The stiffness matrices result using the method presented
in Example 11.5.2. The stiffness matrices of all MSEs are given below after the
static condensation of the rotational degrees of freedom.
2 3
29237:03 0 0 16737:43 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
6
k ¼6
1 7
6 16737:43 0 0 13897:72 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
2 3
44851:12 0 0 26194:78 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
6
k ¼6
2 7
6 26194:78 0 0 22307:63 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
2 3
44851:12 0 0 26194:78 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
6
k ¼6
3 7
6 26194:78 0 0 22307:63 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
2 3
41960:02 0 0 23362:92 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
6
k ¼6
4 7
6 23362:92 0 0 18828:40 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 711
2 3
41960:02 0 0 23362:92 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
k ¼6
5
6 23362:92
7
6 0 0 18828:40 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
2 3
27145:69 0 0 14736:02 0 0
6 0 0 0 0 0 07
6 7
6 0 0 0 0 0 07
6
k ¼6
6 7
6 14736:02 0 0 11490:06 0 0 77
4 0 0 0 0 0 05
0 0 0 0 0 0
2 3
0 0 0 0 0 0
60 0 0 0 0 0 7
6 7
60 0 15364:52 0 0 8642:40 7
k7, 8, …, 14 ¼ 6
60
7
6 0 0 0 0 0 7 7
40 0 0 0 0 0 5
0 0 8642:40 0 0 8642:40
Step 2
Formulation of the transformation matrices Ri due to the rotation of the local
axes. From Fig. E15.3, we have f1 ¼ f2 ¼ f3 ¼ f7 ¼ f8 ¼ ⋯ ¼ f14 ¼ 0,
f4 ¼ f5 ¼ f6 ¼ п=2. Hence
2 3 2 3
1 0 0 0 0 0 0 1 0 0 0 0
60 1 0 0 0 07 6 1 0 0 0 0 0 7
6 7 6 7
6 0 0 1 0 0 0 7 6 0 0 1 0 0 07
^ ^
R ¼R ¼R ¼6 ^ 6 7 ^ ^ ^ 6 7
7, R ¼ R ¼ R ¼ 6 0 0 0 0 1 0 7
1 2 3 4 5 6
60 0 0 1 0 07 6 7
40 0 0 0 1 05 4 0 0 0 1 0 0 5
0 0 0 0 0 1 0 0 0 0 0 1
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
6 07
^7 ¼ R
R ^ 14 ¼ 6 0
^8 ¼ ⋯ ¼ R 0 1 0 0 7
60 0 0 1 0 07
6 7
40 0 0 0 1 05
0 0 0 0 0 1
Step 3
Formulation of the transformation matrices Ti due to translation of the local
axes. Apparently, in this case, any point of the local x axis can be taken as point i.
Referring to Fig. E15.3 and applying Eq. (15.2.20), we obtain
@CivilMethod
712 PART II Multi-degree-of-freedom systems
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
60 0 1 0 0 07
MSE 1 : x1 ¼ x2 ¼ 0, y1 ¼ y2 ¼ 0, T1 ¼ 6
1 1 1 1 6 7
7
60 0 0 1 0 07
40 0 0 0 1 05
0 0 0 0 0 1
2 3
1 0 0 0 0 0
6 0 1 0 0 0 07
6 7
6 8 0 1 0 0 0 7
MSE 2 : x21 ¼ x22 ¼ 0, y21 ¼ y22 ¼ 8:0, T2 ¼ 6
6 0 0 0 1 0 07
7
6 7
4 0 0 0 0 1 05
0 0 0 8 0 1
2 3
1 0 0 0 0 0
6 0 1 0 0 0 07
6 7
6 16 0 1 0 0 0 7
MSE 3 : x1 ¼ x2 ¼ 0, y1 ¼ y2 ¼ 16:0, T3 ¼ 6
3 3 3 3 6 7
7
6 0 0 0 1 0 07
4 0 0 0 0 1 05
0 0 0 16 0 1
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
60 0 1 0 0 07
6 7
MSE 4 : x41 ¼ x42 ¼ 0, y41 ¼ y42 ¼ 0, T4 ¼ 6 7
60 0 0 1 0 07
6 7
40 0 0 0 1 05
0 0 0 0 0 1
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
60 5 1 0 0 07
MSE 5 : x1 ¼ x2 ¼ 5:0, y1 ¼ y2 ¼ 0, T5 ¼ 6
5 5 5 5 6 7
7
60 0 0 1 0 07
40 0 0 0 1 05
0 0 0 0 5 1
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
6 0 10 1 0 0 0 7
MSE 6 : x61 ¼ x62 ¼ 10:0, y61 ¼ y62 ¼ 0, T6 ¼ 6
60 0 0 1 0 07
7
6 7
40 0 0 0 1 05
0 0 0 0 10 1
The reaming MSEs are columns, which contribute only with their rotational
stiffness. Thus, we have independently from the directions of their local axes
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 713
2 3
1 0 0 0 0 0
60 1 0 0 0 07
6 7
60 0 1 0 0 07
T7 ¼ T8 ¼ ⋯ ¼ T14 ¼ 6
60
7
6 0 0 1 0 077
40 0 0 0 1 05
0 0 0 0 0 1
Step 4
Formulation of the stiffness matrix of the structure. Applying Eq. (15.2.43b)
yields the stiffness matrix of the structure with respect to the origin O of
the axes.
X
14 T i
Ko ¼ ^i k R
Ti R ^ i TT
i
2 i¼1 3
118939:27 0 1076426:90 69126:99 0 628674:79
6 61461:87 264174:87 77
6 0 111065:74 481257:02 0
6 7
6 1076426:90 481257:02 18238842:00 7
628674:79 264174:87 10509145:00 7
¼6
6 69126:99
6 0 628674:79 58512:99 0 535383:28 77
6 7
4 0 61461:87 264174:87 0 49146:87 209042:62 5
628674:79 264174:87 10509145:00 535383:28 209042:62 8827299:10
Step 5
Formulation of the mass matrix of the structure. The surface of the plates is
120:0 m2 . Neglecting the mass of the MSEs, we have
24 6
m1 ¼ m2 ¼ ð8:40 10:40 + 8:0 5:40Þ 0:15 +
9:81 9:81
1 2
¼ 127:76 kN m s
xc1 ¼ xc2 ¼ 4:17, yc1 ¼ yc2 ¼ 9:29, I1c ¼ I2c ¼ 3680:24
Hence the mass matrix with respect to the center of mass is
2 3
127:76 0 0 0 0 0
6 0 127:76 0 0 0 0 7
6 7
6 0 0 3680:24 0 0 0 7
c
M ¼6 6 7
6 0 0 0 127:76 0 0 77
4 0 0 0 0 127:76 0 5
0 0 0 0 0 3680:24
Moreover, the transformation matrix from C to O is
@CivilMethod
714 PART II Multi-degree-of-freedom systems
2 3
1 0 0 0 0 0
6 0 1 0 0 0 07
6 7
6 9:29 4:17 1 0 0 07
Tc ¼ 6
6 0
7
6 0 0 1 0 077
4 0 0 0 0 1 05
0 0 0 9:29 4:17 1
and applying Eq. (15.2.36) yields the mass matrix with respect to point O
Mo ¼ Tc Mc TTc
2 3
127:76 0 1186:89 0 0 0
6 0 127:76 532:75 0 0 0 7
6 7
6 1186:89 532:75 16928:06 0 0 0 7
¼66
7
6 0 0 0 127:76 0 1186:89 7
7
4 0 0 0 0 127:76 532:75 5
0 0 0 1186:89 532:75 16928:06
Therefore the equation of motion is
€
+ Ko U
Mo U ¼ Mo bug ðt Þ, b ¼ f 1 0 0 1 0 0 gT
w5 35.814 35.299
w6 48.829 48.332
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 715
2 3
0:0014 0:0455 0:0836 0:0025 0:0684 0:1323
6 7
6 0:0471 0:0011 0:0363 0:0768 0:0015 0:0557 7
6 7
6 7
6 0:0002 0:0003 0:0088 0:0003 0:0006 0:0139 7
6
F¼6 7
7
6 0:0018 0:0701 0:1315 0:0019 0:0440 0:0844 7
6 7
6 0:0762 0:0014 0:0561 0:0475 0:0013 0:0356 7
4 5
0:0002 0:0004 0:0139 0:0003 0:0005 0:0088
(a)
(b)
FIG. 15.3.1 Multistory building (a) with support excitation (b).
@CivilMethod
716 PART II Multi-degree-of-freedom systems
r ¼ f I I ⋯ I gT (15.3.3)
Taking into account that the elastic and damping forces are produced only
by the relative displacements, we can write the equation of motion as (see
Section 14.7.2)
€ + Cu
Mu _ + K
u ¼ Mr€
ug (15.3.4)
it is
For ground motion in direction x,
r ¼ f 1 0 0 1 0 0 ⋯ 1 0 0 gT (15.3.5)
Apparently, it is convenient to refer the displacements to the center of mass
of the plates. Then, taking into account that u c ¼ TTc u ¼ TT
, u c uc and
c T
M ¼ Tc M Tc , Eq. (15.3.4) becomes
c c
€ + Tc Cu
Tc M c u _ + Tc K u
c ¼ Tc Mc TTc r€
ug (15.3.6)
To avoid the inversion Tc , we can readily show that (see Section 11.11)
2 3
j 1 1 0 0
Tc ¼ 40 1 0 5 j ¼ 1, 2, …,n (15.3.10)
yj xcj 1
c
1 T
We recall the notation TT
c ¼ Tc (see Section 12.5.3).
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 717
@CivilMethod
718 PART II Multi-degree-of-freedom systems
are given in Table E15.4 together with the peak values of the modal
quantities.
Applying Eq. (14.8.8) gives the correlation matrix
2 3
1000:00 602:02 61:70 4:29 3:89 2:16
6 602:02 1000:00 2:51 7
6 96:66 5:08 4:59 7
6 7
6 61:70 96:66 1000:00 10:19 9:04 4:52 7
r ¼ 10 6
3
6
7
6 4:29 5:08 10:19 1000:00 811:16 70:45 7
7
6 7
4 3:89 4:59 9:04 811:16 1000:00 92:45 5
2:16 2:51 4:52 70:45 92:45 1000:00
Obviously, the off-diagonal elements are not negligible. That was antic-
ipated because there are eigenfrequencies close to each other. Therefore, the
CQC method should be used to compute the peak values. The resulting
values are given in Table E15.5 as compared with those obtained by direct
solution of the equation of motion (Response History Analysis), namely
€
+ Ko U
Mo U ¼ Mo bug ðt Þ, b ¼ f 1 0 0 1 0 0 gT
@CivilMethod
TABLE E15.4 Peak values of the modal quantities.
Hence
8 9 8 9
> U1> > 0 >
>
> > > >
> V >
> >
>
>
>
> 0 >
>
>
>
> 1 >
> >
> >
>
> =
< > >
< >
=
0 W
D ¼R
1 ^ T U
1 T ¼ , D 7
¼ ^ T U
R 7 T ¼ 1
> U 2 >
> > 0 >
1 7
> >
> >
> >
>
>
> >
> >
> >
>
>
> V >
> >
> 0 >
>
>
: ;
2 > : >
> ;
0 W2
which yields
1
u1 ¼ u 1 ¼ U 1 , u2 ¼ u 1 ¼ U 2 , w7 ¼ W (1)
1 2 1
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 721
21942:86 21942:86 8668:78 8668:78
Kab ¼
21942:85 21942:86 21942:86 21942:86
2 3
103700 26250 25600 0
6 26250 103700 0 25600 7
Kbb ¼ 6
4 25600
7
0 143522:22 26250 5
0 25600 26250 143522:22
2 3
0:163 0:148
6 0:163 0:148 7
ub ¼ Kbb Kba ua ¼ 6
1 7 u1 (2)
4 0:026 0:107 5 u2
0:026 0:107
The stress resultants at the cross-section a (see Fig. E15.4) result from the
relations
GI t
Mta ¼ W1 (3c)
L
@CivilMethod
722 PART II Multi-degree-of-freedom systems
Spa ðx n , Tn Þ
n Tn (m/sec2) Qxn0 Mxn0 Mtn0
1 0.669 3.7824 0.0159 0.0396 0.0029
2 0.616 4.1419 115.270 298.4715 1.3213
3 0.455 3.4180 2.888 7.5130 0.5139
4 0.184 5.3077 0.003 0.0064 0.0004
5 0.175 6.0243 7.5104 15.9047 0.0614
Remark The procedure applied to the solution of the previous problem is approx-
imate and is recommended only when the peak values are computed using the
pseudoacceleration spectrum provided by the earthquake codes. However, the
procedure is highly simplified if the equations of motion are solved numerically.
@CivilMethod
Dynamic analysis of multistory buildings Chapter 15 723
Actually, this is feasible only for a specified accelerogram of the ground motion.
Therefore, the earthquake codes should tend to provide a representative design
accelerogram in place of the pseudoacceleration spectrum.
15.4 Problems
Problem P15.1 Formulate the equation of motion of the two-story building in
Fig P15.1 when subjected to ground motion ug ðt Þ in the x direction and compute
the eigenfrequencies and mode shapes. The material of the structure is RC with
E ¼ 2:1 107 kN=m2 , n ¼ 0:25. Assume: The thickness of the slabs 15 cm,
cross-sectional dimensions of beams 30 50 cm2 , of columns 1, 3, 5, 7, 9
30 30 cm2 , columns 2, 8 25 40 cm2 , and columns 4, 640 25 cm2 . All col-
umns are fixed to the wall of the basement except for column 3, which is supported
elastically to the ground via an isolated footing with an elastic foundation parameter
k ¼ 30:000 kN=m3 . The beams connecting the footing to the basement walls have
dimensions 30 50 cm2 . Live load of the roof slab 5 kN=m2 and of the mezzanine
slab 10 kN=m2 .
Beam
Beam
Beam
@CivilMethod
724 PART II Multi-degree-of-freedom systems
@CivilMethod
Chapter 16
Base isolation
Chapter outline
16.1 Introduction 725 16.3.2 Reduction of the DOF
16.2 Analysis of the one-story of the superstructure
building with base isolation 727 using mode shapes 745
16.2.1 Linear response of the 16.3.3 Reduction of the
isolation systems 727 superstructure DOF
16.2.2 Modeling of nonlinear using Ritz vectors 745
response of isolation 16.3.4 Linear response of
systems 735 the isolation system 746
16.3 The multistory building with 16.3.5 Nonlinear response
base isolation 741 of the isolation
16.3.1 The equation of motion system 747
of the multistory building 16.4 Problems 753
with base isolation 741 References and further reading 753
16.1 Introduction
Base isolation, also known as seismic isolation, has always been a matter of
great interest to engineers, who have sought to isolate the structure from ground
motion in order to mitigate as far as possible the forces and deformations due to
the earthquake motion of the ground. This would ensure the integrity of the
structure and its contents as well as the protection of its users from injury or
death. Although the concept of base isolation has its roots in the distant past,
more than a century ago, it has only nowadays been a means of designing civil
engineering structures, mainly buildings and bridges, or even protecting old
works of historical value, thanks to the development of reliable low-cost base
isolation systems that guarantee the desirable performance characteristics prac-
tically unlimited. The related literature is extensive [1–4], and the research on
the subject is ongoing due to great interest.
Base isolation uses a collection of structural elements that substantially
decouple a superstructure from its substructure that is resting on a shaking
ground. All in-use isolation systems, despite their wide variation in detail,
can be categorized into two main types as they follow two basic approaches with
certain common features [3]. The first type includes the elastomeric isolators
and the second the sliding systems. The elastomeric isolators perform as a layer
with small horizontal stiffness, which is interposed between ground and super-
structure. They allow a relative displacement that can reach their height. Their
vertical stiffness is very large so that they can be considered practically unde-
formed when subjected to large vertical loads. They usually have a cylindrical
shape and are made of hard natural rubber or neoprene layers alternating with
steel plates. Inasmuch as the damping of the rubber is small, mechanical
dampers, hydraulic dampers, or lead cores are used in the elastomeric bearings
(Fig. 16.1.1) to increase damping. The latter bearings are known by the name
LRB (lead rubber bearings).
Lead core
Section
FIG. 16.1.1 Lead rubber bearing.
The second type uses sliders between the ground and superstructure. The
shear force transmitted through the interface is significantly reduced by
decreasing the friction coefficient as much as practically possible. However,
friction must not be completely eliminated to prevent movements due to wind
or minor earthquakes. In this type of bearing, the displacements are controlled
by springs, elastomeric devices, or even by curved sliding surfaces. The sliding
interface of these bearings mainly consists of stainless steel and Teflon. The
friction pendulum is a device of this type, Fig. 16.1.2.
R Articulated slider
Section
FIG. 16.1.2 Friction pendulum bearing.
The curvature of the surface provides resisting forces, which bring the struc-
ture back to its original equilibrium position after displacement. The study of
the dynamic response of structures with sliders is quite complex due to the non-
linearity of the sliding process. The isolation devices reduce stiffness and
increase damping between the ground and structure. Reducing stiffness aims
at shifting the fundamental period of the structure away from the characteristic
@CivilMethod
Base isolation Chapter 16 727
period of the ground motion. Isolated structures have large fundamental vibra-
tion periods compared with the predominant periods of most seismic motions.
Increasing the damping reduces the displacements on the isolators. However,
excessive damping may introduce forces in the structure that increase deforma-
tions in the higher modes. Fig. 16.1.3 shows the pseudoacceleration spectrum of
the Athens earthquake (see Chapter 6). As was pointed out in Section 6.2, the
maximum elastic base shear force of the building is calculated from the
relationship
max Q0 ¼ mS pa ðT , xÞ (16.1.1)
It becomes obvious from Fig. 16.1.3, that for a building with period
TF ¼ 0:6s it is Spa 425cm=s2 , while for an isolated building with TI ¼ 2s
it is Spa 50cm=s2 (see Section 16.2.1). Consequently, the peak base shear
force is considerably reduced, by 8.5 times of that before isolation.
Design spectrum
T (s)
FIG. 16.1.3 Pseudoacceleration spectrum of the Athens earthquake Spa ðT , x Þ, (x ¼ 0:05).
@CivilMethod
728 PART II Multi-degree-of-freedom systems
be isolated. More specifically, Fig. 16.2.1a shows the building resting directly
on the ground while Fig. 16.2.1b shows it mounted on a rigid isolated slab. From
now on, we will call the first structure a fixed-base structure while the second
one is an isolated structure. Although the behavior of the isolation system is
nonlinear, the assumption of linear behavior facilitates our purpose in the
first stage.
Isolation
system
Gap Base slab
(a) (b)
FIG. 16.2.1 One-story building: (a) fixed-base structure and (b) isolated structure.
Quantities referred to the fixed-base structure are provided with the subscript
F while those referred to the isolated structure with the subscript I . Moreover,
quantities referred to the isolation system are denoted with the subscript b.
The fixed-base structure is modeled as an SDOF system while the isolated
structure is modeled as a two-degree-of-freedom system (Fig. 16.2.2). Thus,
we have
rffiffiffiffiffi
k
wF ¼ (16.2.1a)
m
2p
TF ¼ (16.2.1b)
wF
c
xF ¼ (16.2.1c)
2mwF
The equations of motion of the isolated structure with respect to the relative
displacements ub and u for ground motion ug ðt Þ are obtained using the method
of equilibrium of forces
@CivilMethod
Base isolation Chapter 16 729
m 0 u€ c c u_ k k u
+ +
0 mb u€b c c + cb u_ b k k + kb ub
m
¼ u€g (16.2.2)
mb
or setting mb ¼ am, cb ¼ bc, kb ¼ gk
1 0 u€ 1 1 u_ 1 1 u
+ 2xF wF + wF
2
0 a u €b 1 1 + b u_ b 1 1 + g ub
1
¼ u€ (16.2.3)
a g
The determinant of the eigenvalue problem for the undamped vibrations is
1l 1
¼0 (16.2.4)
1 ð1 + g Þ al
where
w2
l¼ (16.2.5)
w2F
The characteristic equation of Eq. (16.2.4) is
f ðlÞ ¼ al2 ð1 + a + g Þl + g ¼ 0 (16.2.6)
which yields
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1+a+g ð1 + a + g Þ2 4ag
l1, 2 ¼ (16.2.7)
2a 2a
1 1
F¼ (16.2.8)
1 l1 1 l2
The stiffness kb of the isolation system is very small compared to the stiff-
ness k of the fixed-base structure while mb is of the same order of magnitude as
m or smaller. Thus, we may limit our study to the following two cases:
(a) g ¼ 0:1 and a ¼ 1
9
l1 ¼ 0:0487, l2 ¼ 2:0514 >
2 3>>
>
1:000 0:951 >=
w1 ¼ 0:220wF , w2 ¼ 1:432wF , F1 ¼ 4 5 (16.2.9)
>
0:951 1:000 > >
>
>
;
T1 ¼ 4:529TF , T2 ¼ 0:698TF
@CivilMethod
730 PART II Multi-degree-of-freedom systems
(a)
(b)
FIG. 16.2.3 Mode shapes and periods of the isolated structure.
The mode shapes together with the corresponding natural periods for the
examined two cases are shown in Fig. 16.2.3. The first mode has a large natural
period, which differs slightly from the mode of a rigid isolated structure,
f1 0:916gT f 1 1 gT . Obviously, only the isolation system is deformed in this
mode, whereas the structure behaves as a rigid body. Therefore, this mode is
called the isolation mode. The second mode has a small natural period and
expresses the deformation of the structure. Therefore, it is called the structural
mode. This mode, however, contributes little to the earthquake-induced forces
in the structure (see Example 16.2.1).
At this point, we make a remarkable observation. If we set l1 ¼ g=ð1 + aÞ in
Eq. (16.2.6) and take into account that g is small, we have
f ðl1 Þ ¼ g 2 0 (16.2.11)
that is, this value of l1 is the first eigenvalue as it satisfies Eq. (16.2.4). If this
value of l1 is introduced in Eq. (16.2.5), we obtain
@CivilMethod
Base isolation Chapter 16 731
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kb
w1 (16.2.12)
m + mb
3:032 3:0902
FT1 C1 F1¼c (16.2.15a)
3:0902 7:1564
2:0879 2:1799
F2 C2 F2 ¼ c
T
(16.2.15b)
2:1799 3:6716
We observe that the orthogonality condition is not satisfied. Moreover, the
off-diagonal elements are of the same order of magnitude as the diagonal ones,
hence they cannot be neglected (see Example 12.10.3). This was anticipated
because the difference in the damping between the structure and the isolation
is large. Therefore, the method of mode superposition presented in
Chapter 14 using real mode shapes cannot be applied. On the other hand, the
uncoupling of the equations of motion using complex modes as presented in
Section 12.10 is not recommended because this procedure, though effective,
is not convenient to understand the physical response of the structure. The
numerical solution of the equations of motion offers a simple and efficient alter-
native means to overcome this problem. Nevertheless, the use of the mode
superposition method with the real modes of the undamped vibrations allows
us to draw useful conclusions for the dynamic response of the isolated
single-story building, as will be shown in the next example.
Example 16.2.1 Compute the peak values of the base shear force Qb , the over-
turning moment Mb , the displacement ub , and the relative displacement u ub
as well as the peak value of the shear force at the base of the columns of the one-
story isolated building of Fig. 16.2.1 when it is subjected to the Athens 1999
earthquake using the response spectrum method. Compare the computed quan-
tities with those of the fixed-base structure. Assume: m ¼ 50kNm1 s2 ,
k ¼ 15, 000kNm1 , mb ¼ 0:7m, Tb ¼ 2s, xF ¼ 3%, xb ¼ 15%.
Solution
(i) Characteristic parameters
wF ¼ 17:32s1 , TF ¼ 0:362s
wb ¼ 2p=Tb ¼ 3:141s1 , kb ¼ w2b ðm + mb Þ ¼ 838:59kNm1 m
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x
a ¼ mb =m ¼ 0:7, g ¼ kb =k ¼ 0:056, b ¼ ð1 + aÞg b ¼ 1:541
xF
c ¼ 51:961kNm1 s1
(ii) Eigenfrequencies, mode shapes, and modal quantities.
Eqs. (16.2.7), (16.2.8) give
@CivilMethod
Base isolation Chapter 16 733
@CivilMethod
734 PART II Multi-degree-of-freedom systems
(iv) Computation of the peak values jQb jmax , jMb jmax , jub jmax , ju ub jmax ,
jQ jmax .
They are computed using the response spectrum method. First the quantities Qbn0 ,
Mbn0 , ubn0 , ðu ub Þn0 , Qn0 are computed using the relation (14.8.4), that is,
qn0 ¼ qnst Spa ðTn , xn Þ
The quantities Spa ðxn , Tn Þ are computed using the program response_
spectrum_aem.m given in Chapter 6 for the Athens earthquake and are pre-
sented together with the peak modal values in Table E16.2. For the fixed-base
structure, it is Spa ðT ¼ 0:362, x ¼ 0:03Þ ¼ 4:116cm=s2 .
TABLE E16.2 Peak modal values Qbn0 , Mbn0 , ubn0 , ðu ub Þn0 , Qn0 .
Spa ðTn , x n Þ Qbn0 Mbn0 ðu ub Þn0 Qn0
n Tn (s) xn (cm/s2) (kN) (kN m1) ubn0 (cm) (cm) (kN)
1 2.018 0.146 37.05 31.48 65.75 3.753 0.125 9.38
@CivilMethod
Base isolation Chapter 16 735
Table E16.3 shows the peak values for the isolated building obtained by the
CQC method together with the corresponding values obtained for the fixed-base
structure. The latter values are computed by treating the structure as an SDOF
system with T ¼ 0:362s and x ¼ 0:03. Obviously, the reduction of all quantities
is considerable in the isolated structure.
In obtaining the previous solution, an approximate damping matrix was used
by omitting the off-diagonal elements in order to avoid the modal superposition
method with complex eigenfrequencies and eigenmodes (see Section 12.10).
This, however, can be surpassed by solving Eq. (16.2.2) numerically. For the
data of the problem, the computed peak values are given in Table E16.4.
TABLE E16.4 Peak values obtained using a numerical solution of Eq. (16.2.2).
Isolated jQb jmax jMb jmax jub jmax ju ub jmax jQ jmax
building (kN) (kN m1) (cm) (cm) (kN)
Approximate 31.48 66.82 3.753 0.126 9.59
Numerical 31.28 95.83 3.724 0.193 14.46
@CivilMethod
736 PART II Multi-degree-of-freedom systems
(b) Slip mode. When the shear force at the interface of the bearing reaches the
maximum friction force, the bearing is forced to slide.
The damping force fD due to friction at the sliding interfaces satisfies the
condition
jfD j mW (16.2.18)
where m is the friction coefficient and W the weight applied to the bearing.
The friction coefficient can be either constant as considered in Coulomb’s
model (see Chapter 7), or depending on the sliding velocity and the bearing
pressure.
For Teflon-steel interfaces, it has been proposed as [5,6],
m ¼ mmax ðmmax mmin Þea|u_ | (16.2.19)
where mmax and mmin are the maximum and minimum values of the friction
coefficient, respectively, u_ the sliding velocity, and a a coefficient controlling
the transition between mmax and mmin . Obviously, it is m ¼ mmin if u_ ¼ 0 and
m ! mmax if u_ ! 1.
Hence the condition for nonsliding is
jfD j < mW and u_ ¼ 0 (16.2.20a)
while for sliding
fD ¼ mW sgnðu_ Þ (16.2.20b)
@CivilMethod
Base isolation Chapter 16 737
and it represents the stiffness of the isolation system. Obviously, this force
depends nonlinearly on the displacement u. Expansion of fS in Taylor’s series
with respect to u gives
W W 3W 5
fS ¼ u + 3 u3 + 5
u + O u7
R 2R 8R
1 u 2 3 u 4
6 (16.2.25)
W
¼ u 1+ + +O u
R 2 R 8 R
The approximation by a finite number of terms depends on the ratio umax =R.
For example, if it is umax ¼ 0:25m and R ¼ 1:00m, the contribution of the
second term is 3:12% and that of the third 0:14%. Hence, the response can
be reasonably approximated as linear
W
fS ¼ u (16.2.26)
R
Thus the stiffness of the bearing is
W
kb ¼ (16.2.27)
R
@CivilMethod
738 PART II Multi-degree-of-freedom systems
yielding
pffiffiffiffiffiffiffiffiffi
wb ¼ g=R (16.2.28a)
pffiffiffiffiffiffiffiffiffi
Tb ¼ 2p R=g (16.2.28b)
We observe that the natural period of the isolated structure does not depend
on the supported weight, hence on the mass of the structure. Eq. (16.2.28b) can
be employed to establish the radius of curvature R of the bearing for a specified
natural period. For example, if the desired natural period Tb ¼ 2s, the radius of
curvature should be
g
R¼ Tb2 1:00m (16.2.29)
ð2pÞ2
The Taylor’s series expansion of the restoring force, though not necessary
from the computational point of view, gives an insight into the response of the
friction pendulum bearing. Nevertheless, linearization must be done with great
caution because the nonlinear behavior of the system may be completely differ-
ent. When, for example, we keep the cubic term in the expression (16.2.25), the
equation of motion of the SDOF system takes the form
u€ + w2 u + mu 3 ¼ u€g (16.2.30)
where
pffiffiffiffiffiffiffiffiffi
w¼ g=R (16.2.31a)
g
m¼ 3>0 (16.2.31b)
2R
Eq. (16.2.30) is known as the Duffing equation with hardening. The eigen-
frequency of free vibrations depends on the amplitude of the vibration. The
is complicated,
response of the system to harmonic excitations (u€g ¼ p0 cos wt)
especially in the region of resonance where instability and jump phenomena
arise. The response becomes even more complicated in the presence of damp-
ing. Modern developments use fractional derivative models to simulate the dis-
sipation forces. However, such analyses are beyond the scope of this book.
Therefore, the interested reader is advised to look in the relevant literature.
@CivilMethod
Base isolation Chapter 16 739
The curve representing the actual variation of the restoring force fS versus
the displacement in this model is approximated by the bilinear response
(Fig. 16.2.5), where Ke is the elastic stiffness, Ky the stiffness in the yield phase,
Fy the yield force, and Uy the yield displacements
@CivilMethod
740 PART II Multi-degree-of-freedom systems
Ke Uy + Ky umax Uy
Keff ¼ (16.2.34)
umax
Subsequently, the viscous part cb can be obtained from
cb ¼ 2xeq ðmb + m Þwb (16.2.35)
where the equivalent linear viscous damping ratio xeq is calculated using
Eqs. (7.3.2), (7.3.3), which for r ¼ umax and Keff give
WD
x eq ¼ 2
(16.2.36)
2pKeff umax
The area WD of the hysteresis loop is obtained from Fig. 16.2.6.
1
WD ¼2Fd 2umax 2 2Fd 2Uy
2 (16.2.37)
¼4Fd umax Uy
This approach is permitted by the regulations (design codes) and is
very useful, at least at the stage of preliminary design. This effective modulus
is used in Eqs. (16.2.13a)–(16.2.13c) with kb ¼ Keff to compute the natural
period
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Keff 2p
wb , Tb (16.2.38)
m + mb wb
Eqs. (16.2.36)–(16.2.38) are used in a repetitive process to determine the
characteristics of the seismic isolation system at each step.
@CivilMethod
Base isolation Chapter 16 741
@CivilMethod
742 PART II Multi-degree-of-freedom systems
8 9 8 9
< un = < uj =
u¼ ⋮ with uj ¼ vj , j ¼ 1, 2, …,n (16.3.2)
: ; : ;
u1 wj
is the 3n 1 vector of the relative displacements of the fixed-base structure with
respect to the base (Fig. 16.3.2),
8 9
< ub =
u b ¼ vb (16.3.3)
: ;
wb
is the 3 1 vector of the relative displacements of the base with respect to the
ground and
8 9
< ug =
ug ¼ v g (16.3.4)
: ;
wg
is the 3 1 vector of the ground displacements. Usually, it is wg ¼ 0.
@CivilMethod
Base isolation Chapter 16 743
@CivilMethod
744 PART II Multi-degree-of-freedom systems
X
n X
n
Fa ¼ f Dj + f Sj (16.3.10)
j¼1 j¼1
@CivilMethod
Base isolation Chapter 16 745
@CivilMethod
746 PART II Multi-degree-of-freedom systems
z ¼ ZY e ðt Þ (16.3.23)
n oT
e ðt Þ ¼ Y
decouples the linear equations with Y e 1 ðt Þ, Y
e 2 ðt Þ, …, Y
e K ðt Þ being
the generalized coordinates. Thus, we obtain
" # "h i # "h i #( )
e T Mr n o n o 2 e
Y
I F €
eu 2x e
w 0 _
e u_ b + e
w 0
Y €b + j j
Y j
e
r MF r Mr + Mb
T T 0 Cb 0 Kb ub
( ) ( )
T
e Mr
0 F
+ nl ¼ €g
u (16.3.24)
fb rT Mr + Mb
where
e ¼ ΨZ
F (16.3.25)
It is convenient to use the first K eigenmodes of the superstructure as
Ritz vectors.
6
i
kx 0 i
y i kx 7
6 i¼1 7
6 X
i¼1
X 7
6 m m
7
6
Kb ¼ 6 0 i
ky xi kyi 7 (16.3.26)
7
6 i¼1 i¼1 7
6 X m Xm Xm 7
4 5
yi kxi xi kyi yi2 kxi + xi2 kyi
i¼1 i¼1 i¼1
@CivilMethod
Base isolation Chapter 16 747
2 X
m X
m 3
6 0 cxi yi cxi
7
6 i¼1 7
6 i¼1
7
6 X X 7
6 m m
7
6
Cb ¼ 6 0 i
cy i
xi cy 7 (16.3.27)
7
6 i¼1 i¼1 7
6 7
6 X m Xm Xm 7
4 2 i 5
i
yi cx i
xi cy 2 i
yi cx + xi cy
i¼1 i¼1 i¼1
Moreover, it is f nl
b ¼ 0 and the total restoring force of the isolation system is
expressed as
Rðub , u_ b Þ ¼ Cb u_ b + Kb ub (16.3.28)
In this case, it is obvious that the equation of motion (16.3.20) becomes lin-
ear, but the matrix Cb is not diagonalized by the eigenmodes of the undamped
free vibrations. The solution of the resulting equation can be obtained numer-
ically using any of the methods presented in Section 14.10.
@CivilMethod
748 PART II Multi-degree-of-freedom systems
8 m 9
> X >
>
> i
fSx >
>
>
> >
>
>
> >
>
>X
<
i¼1
m >
=
i
fb ¼
nl
fSy (16.3.31)
>
> >
>
>
>
i¼1
m >
>
>
> X >
>
>
> i
i >
>
: x i f Sy yi fSx ;
i¼1
i i
The forces fSx , fSy are obtained from Eqs. (16.3.29a), (16.3.29b) with
u ¼ u b y i wb (16.3.32a)
v ¼ v b + x i wb (16.3.32b)
Eq. (16.3.20) is now nonlinear and can be solved by the methods presented
in Section 14.11 after appropriate modification to include Eqs. (16.3.30a),
(16.3.30b).
Example 16.3.1 The two-story building of Fig. E16.1 is seismic isolated
using LRB. Study its dynamic response when it is subjected to ground motion
in the x direction due to the Athens 1999 earthquake and compute:
2 at the top of column 1, the shear force Qx , the bend-
(i) The peak values U 2 , W
ing moment Mx , and the torsion moment at the base (foot) of the same
column.
(ii) The peak values of the base shear force Qbx , overturning momentMbx , the
base torsion momentTb .
Data:
a. Modal damping of the superstructure x ¼ 0:05 for all eigenmodes.
b. The eight isolators will be of the same type with Keff ¼ 500kNnm1 and
xeq ¼ 0:113, and will be placed under the eight columns.
c. The base, which is considered undeformed, has a thickness 0:15m and a
total load 2kN=m2 .
Solution
The analysis will be done by considering a linear behavior of the damping sys-
tem. We will formulate the equation of motion (16.3.19) with respect to the cen-
ters of mass of the plates, which are located on the same vertical axis through the
point xc ¼ 4:17m, yc ¼ 9:29m.
The mass matrix of the superstructure is
2 3
127:76 0 0 0 0 0
6 7
6 0 127:76 0 0 0 0 7
6 7
6 0 0 3680:24 0 0 0 7
6 7
M ¼6
c
7
6 0 0 0 127:76 0 0 7
6 7
6 7
4 0 0 0 0 127:76 0 5
0 0 0 0 0 3680:24
@CivilMethod
Base isolation Chapter 16 749
where
2 3
1 0 0 0 0 0
6 0 1 0 0 0 07
6 7
6 9:29 4:17 1 0 0 07
Tc ¼ 6
6 0
7
6 0 0 1 0 077
4 0 0 0 0 1 05
0 0 0 9:29 4:17 1
@CivilMethod
750 PART II Multi-degree-of-freedom systems
2 3
0:940 0 0 0 0 0 0 0 0
60 1:019 0 0 0 0 0 0 0 7
6 7
60 0 7
6 0 1:379 0 0 0 0 0 7
60 0 7
6 0 0 3:413 0 0 0 0 7
6 7
C* ¼ 6 0 0 0 0 3:581 0 0 0 0 7
6 7
60 0 0 0 0 4:883 0 0 0 7
6 7
60 0 0 0 0 0 254:42 0 33:63 7
6 7
40 0 0 0 0 0 0 254:42 24:66 5
0 0 0 0 0 0 33:63 24:66 12452:77
@CivilMethod
Base isolation Chapter 16 751
2 3
88:311 0 0 0 0 0 0 0 0
6 7
6 0 103:88 0 0 0 0 0 0 0 7
6 7
6 7
6 0 7
6 0 0 190:28 0 0 0 0 0 7
6 7
6 0 7
6 0 0 0 1164:76 0 0 0 0 7
6 7
6 7
K* ¼ 6 0 0 0 0 1282:66 0 0 0 0 7
6 7
6 7
6 0 0 0 0 0 2384:40 0 0 0 7
6 7
6 7
6 0 0 0 0 0 0 4000 0 1160 7
6 7
6 7
6 0 0 0 0 0 0 0 4000 820 7
4 5
0 0 0 0 0 0 1160 820 217442
8 9
>
>
0:057 >
>
>
> >
>
>
> 15:639 >
>
>
> >
>
>
> >
>
>
> >
>
>
> 0:478 >
>
>
> >
>
>
> >
> 0:035 >
> >
> ( )
>
< >
= Y
PI ¼ 3:263 u€g , uI ¼
>
> >
> ub
>
> >
> 0:19 >
> >
>
>
> >
>
>
> >
>
>
>330:04 >
>
>
> >
>
>
> >
>
>
> >
>
>
> 0 >
>
>
: >
;
0
@CivilMethod
752 PART II Multi-degree-of-freedom systems
16.4 Problems
Problem P16.1 The two-story building of Fig. P16.1 is seismic isolated. Com-
pute the reduction of the peak values of the displacements at the top of column 1
as well as of the stress resultants Qx , Mx , Mt at the foot of the same column
when it is subjected to the Athens 1999 earthquake in the x direction. Consider
two types of isolators: (i) bilinear and (ii) Bouc-Wen. The building is made from
reinforced concrete. Assume the following data:
Modulus of elasticity E ¼ 2:1 107 kNm2 ; Poisson ratio n ¼ 0:20; specific
weight g ¼ 25kNm3 ; thickness of the slabs d ¼ 0:15m; and the isolation base
d ¼ 0:20m; columns of the first floor 0:40 0:40m2 and second floor
0:35 0:35m2 ; beams 0:30 0:50m2 ; live load of slabs p ¼ 5kNm2 ;
Bilinear isolators Kx = Ky = Ke = 900kN / m, a = 0.27, Fy = 9.6kN, Fd = 7kN,
Uy = 0.01m;
Bouc-Wen isolators Kx = Ky = Ke = 900kN / m, Fy = 9.6kN, Uy = 0.01m,
a1 = a2 = a = 0.27
@CivilMethod
754 PART II Multi-degree-of-freedom systems
[6] P. Tsopelas, M.C. Constantinou, A.M. Reinhorn, 3D-BASIS-ME: Computer Program for
Nonlinear Dynamic Analysis of Seismically Isolated Single and Multiple Structures and
Liquid Storage Tanks, Report No. NCEER 94-0010. Nat. Ctr. for Earthquake Engrg. Res,
State Univ. of New York, Buffalo, NY, 1994.
[7] J.T. Katsikadelis, A new direct time integration method for the equations of motion in struc-
tural dynamics, ZAMM Z. Angew. Math. Mech. 94 (9) (2014) 757–774, https://doi.org/
10.1002/zamm.20120024.
[8] G.C. Tsiatas, A.E. Charalampakis, A new hysteretic nonlinear energy sink (HNES), Commun.
Nonlinear Sci. Numer. Simul. 60 (2018) 1–11.
[9] Katsikadelis, J.T. A new numerical integration method for higher order ordinary differential
equations (to be published).
[10] S. Nagarajaiah, A.M. Rheinhorn, M.C. Constantinou, Nonlinear Dynamic Analysis of Three-
Dimensional Isolated Structures, 1989. Technical Report NCEER-89-0019.
[11] P. Tsopelas, S. Nagarajaiah, M.C. Constantinou, A.M. Reinhorn, Nonlinear dynamic analysis
of multiple building base isolated structures, Comput. Struct. 50 (1) (1994) 47–57.
@CivilMethod