Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

Chapter 1

General concepts and principles


of structural dynamics
Chapter outline
1.1 Introduction 3 1.7 Hamilton’s principle 42
1.2 Types of dynamic loads 6 1.8 Lagrange’s equations 54
1.3 Dynamic degrees of freedom 9 1.8.1 Derivation of
1.4 Dynamic model and Lagrange’s equations 54
formulation of the equation 1.8.2 Lagrange multipliers 64
of motion of SDOF systems 11 1.8.3 Small displacements 68
1.5 Derivation of the equations 1.8.4 Raleigh’s dissipation
of motion using d’Alembert’s function 72
principle 15 1.9 Influence of the gravity loads 73
1.6 Principle of virtual 1.10 Problems 74
displacements 38 References and further reading 82

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-

Dynamic Analysis of Structures. https://doi.org/10.1016/B978-0-12-818643-5.00001-7


© 2020 Elsevier Inc. All rights reserved. 3
4 PART I Single-degree-of-freedom systems

varying position-dependent distributed loading pðx, t Þ (Fig. 1.1.1a). We can


derive the equation of motion by considering the dynamic equilibrium of a
beam element of length dx (Fig. 1.1.1b). In static consideration, the element
would be in equilibrium under the action of the external load pðx, t Þdx and
the internal forces, that is, the shear force Q, the bending moment M at the left
cross-section of the element, and the shear Q + dQ and moment M + dM at
its right cross-section.

(a)

(b)

(c)
FIG. 1.1.1 Vibrating cantilever beam.
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

From the Euler-Bernoulli beam theory we have

∂3 u
Q ¼ EI (1.1.2)
∂x 3

Substituting Eq. (1.1.2) into Eq. (1.1.1) yields

∂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
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.

1.2 Types of dynamic loads


As already mentioned, dynamic loads are time-varying. Such loads are those
due to an unbalanced rotating machinery, the motion of vehicles on structures,
blast loads, wind loads, etc. The motion of the support of a structure, even in the
absence of external dynamic loads, produces a dynamic response, too. This is
the case of seismic ground motion (see Chapter 6).

FIG. 1.2.1 Periodic load.

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.
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.

From the analytical point of view, it is convenient to divide the deterministic


loads into two basic categories, periodic and nonperiodic loads. Periodic loads
are those whose time variation profile repeats continually at regular time inter-
vals T . Mathematically, they can be represented by a periodic function
pðt Þ ¼ pðt + nT Þ (1.2.1)
where n is a natural number. The time interval T is called the period of the load.
A general type of periodic load is shown in Fig. 1.2.1, which also identifies the
period of the load. A usual type of periodic load is the harmonic load caused by
an unbalanced rotating machine (Fig. 1.2.2); H ¼ mew2 cos wt is the horizontal
component of the centrifugal force. Loads that do not show any periodicity are
called nonperiodic loads. They may be of long duration, such as those resulting
from an earthquake. Nonperiodic loads of short duration are called impulsive
loads. A windblast striking a building and the pressure of a bomb explosion
on a structure are typical impulsive loads (Fig. 1.2.3). The earthquake produces
a special type of load, which is due to the excitation of the support of the
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).

FIG. 1.2.3 Nonperiodic load due to explosion.

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”
General concepts and principles of structural dynamics Chapter 1 9

could be considered as a special case of the dynamic problem with negligible


accelerations, that is, quasistatic.

1.3 Dynamic degrees of freedom


The displacement method is the most suitable method for the dynamic analysis
of structures. In this method, the unknowns are the displacements. For structures
with distributed mass, the displacements are functions of the spatial coordinates
and of the time as well. Their dynamic response is described by partial differen-
tial equations of the hyperbolic type, which must be solved in order to determine
these displacements. The solutions of such equations belong to the most difficult
problems of mathematics. The available solutions refer to simple structures, for
example, beams with a constant cross-section, which are not adequate to study
the dynamic response of engineering structures. Thanks to the development of
modern computational methods such as FEM, BEM and Meshless Methods, the
actual structure is approximated by discrete models in which the mass is local-
ized at a finite number of points (nodal points). These models are adequate to
represent the effects of all significant inertial forces of a structure. At each
instant, the deformed configuration of the structure is determined from the dis-
placementsa of the nodal points, which are functions only of time. The response
of the discretized structure is governed by ordinary differential equations, which
are easy to solve analytically or at least numerically. The number of the indepen-
dent nodal displacements required to determine the deformed shape of the mov-
ing structure is called the number of degrees of freedom. It is apparent that
continuous systems have an infinite number of dynamic degrees of freedom.
Structures with one degree of freedom are called single-degree-of-freedom
(SDOF) systems. Accordingly, we have two-degree-of-freedom (2 DOF) sys-
tems, three-degree-of-freedom (3 DOF) systems, and generally multi-degree-
of-freedom (MDOF) systems. Fig. 1.3.1 shows SDOF systems. Fig. 1.3.1a rep-
resents the idealization of a silo. It consists of two massless columns and a square
rigid plate of mass m. With the assumption that the axial deformation of the col-
umns is negligible, the horizontal displacement u ðt Þ is adequate to completely
determine the motion of the system. Hence, the system has one degree of free-
dom. Likewise, under the same assumptions for the columns, the motion of
the water tower of Fig. 1.3.1b can be determined from the angle fðt Þ.
Fig. 1.3.1c represents the typical model of a SDOF. Fig. 1.3.2a represents the
model of the two-story shear frame. To determine its motion, it is necessary to
establish the two independent horizontal displacements u1 ðt Þ and u2 ðt Þ.
Fig. 1.3.2b represents a cantilever column with a mass atop. This may be
considered as the idealization of a water tower. The whole mass is lumped at
the top while the column is massless. During the motion, the mass undergoes
horizontal displacement u ðt Þ and rotation fðt Þ. These two geometrical quanti-
ties are independent. Hence, the system has two degrees of freedom and thus

a. The term displacent denotes both translation and rotation.


10 PART I Single-degree-of-freedom systems

two differential equations of motion are necessary to determine these displace-


ments. If, however, the mass m of the system is assumed to be concentrated at a
point, its rotational inertia I is equal to zero. Hence, the inertial moment I f€ðt Þ is
zero, too, and one of the equations of motion becomes algebraic. This permits
the elimination of the rotational displacement, leading to only one equation of
motion for u ðt Þ. Consequently, the system has only one dynamic degree of free-
dom, even though it has two static degrees of freedom. Apparently, the number
of static degrees of freedom is not necessarily equal to the number of dynamic
degrees of freedom. Concluding, we can say that in an MDOF system, the num-
ber of dynamic degrees of freedom is equal to the number of independent dif-
ferential equations of motion that must be formulated to establish the dynamic
response of the system.

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).
General concepts and principles of structural dynamics Chapter 1 11

FIG. 1.3.3 Multi-degree-of-freedom (MDOF) system.

FIG. 1.3.4 System with infinite degrees of freedom. Continuous system.

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).

1.4 Dynamic model and formulation of the equation


of motion of SDOF systems
The modeling of the real structure plays a fundamental role in the dynamic anal-
ysis of structures. It is the most difficult task in dynamic analysis because in this
stage of analysis, the experience and theoretical background of the engineer
intervene critically in approximating the structural response.
12 PART I Single-degree-of-freedom systems

Spring

Damper

Frictionless rollers
FIG. 1.4.1 Model of a SDOF system.

Fig. 1.4.1 shows a simple dynamic model of a SDOF system. It consists of a


rigid body of mass m constrained to move along the x axis in the plane of the
paper, a weightless spring connecting the mass to the firm support, and a damper.

Center of mass

FIG. 1.4.2 Forces applied to the free body.

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
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].
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.
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.

1.5 Derivation of the equations of motion using


d’Alembert’s principle
Actually, d’Alembert’s principle is a different interpretation of Newton’s
second law of motion. Suppose that we write it in the form
F  m€
u¼0 (1.5.1)
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 .

If R ¼ X ðt Þi + Y ðt Þj + Z ðt Þk is the position vector of a particle A of


the body with respect to the inertial system of axes XYZ and
r ¼ x ðt Þe1 + y ðt Þe2 + z ðt Þe3 the position of the same point with respect to
the nonrotating system of axes xyz through the center of mass C (see
Fig. 1.5.1), then the equations of motion of the body can be written as
€c
F ¼ mR (1.5.2a)
_c
Mc ¼ H (1.5.2b)
where F ¼ Fx i + Fy j + Fz k is the resultant of the external forces,
€ c ¼ X€ ci + Y€ c j + Z€c k is the acceleration of the center of mass,
R
Mc ¼ Mx e1 + My e2 + Mz e3 is the resultant moment of the external forces with

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].
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

in which r ¼ rðx, y, z Þ is the mass density of the body.


Eq. (1.5.2a) is the equation of the translational motion while Eq. (1.5.2b) is
the equation of the rotational motion.
For a plane body moving in its plane, Eqs. (1.5.2a), (1.5.2b) become (see
Appendix)
Fx ¼ m X€ c (1.5.4a)
Fy ¼ m Y€ c (1.5.4b)
Mc ¼ Ic w_ (1.5.4c)
where w is the angular velocity of the rotational motion about the center of mass
and Ic the polar moment of inertia of the body about the same point.

Path of P

FIG. 1.5.2 Plane body moving in the XY plane. The system of xy axes moves with P without
rotating.

It is often convenient to study the motion with reference to an arbitrary point


P, which is not the center of mass of the body (see Fig. 1.5.2). Special attention
should be paid in this case because Eqs. (1.5.2a), (1.5.2b) take the form
€ p ¼ m€rc
F  mR (1.5.5a)
€p¼H
Mp  rc  m R _p (1.5.5b)
where Rp is the position vector of point P moving with the body and rc the posi-
_ p are the moment of the
tion vector of the center of mass with respect to P. Mp , H
external forces and the rate of change of the angular momentum with respect to
P, respectively.
18 PART I Single-degree-of-freedom systems

When small displacements are considered, as in the theory of linear vibra-


tions, Eqs. (1.5.5a), (1.5.5b) become (see Appendix).
 
Fx ¼ m X€ p  yc w_ (1.5.6a)
 
Fy ¼ m Y€ p + xc w_ (1.5.6b)
 
Mp ¼ Ip w_ + m xc Y€ p  yc X€ p (1.5.6c)
The kinetic energy of a plane body moving in its plane is given
(a) with respect to the center of mass

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.
General concepts and principles of structural dynamics Chapter 1 19

Similarly, Eqs. (1.5.6a)–(1.5.6c) are written


 
Fx ¼ m u€p  yc f€ (1.5.14a)
 
Fy ¼ m v€p + xc f€ (1.5.14b)
 
Mp ¼ Ip f€ + m xc v€p  yc u€p (1.5.14c)
or in matrix form

€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
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

u ðt Þ ¼ a tan fðt Þ  afðt Þ, u ðt Þ ¼ BB 0 cos f  BB 0 (1)

because we have assumed small displacements. Moreover, AA0 ¼ BB 0 =2 ¼ u=2


The applied forces are shown in the free body diagram in Fig. E1.1b. These
are:
The weight of the body:
W ¼ mg (2a)
The spring force:
2
fS ¼ k ðAA0 Þ ¼ ku (2b)
3
The inertial force at the center of mass along x

d2 1 b
fIx ¼ m ðCC 0 Þx ¼ m u€ (2c)
dt 2 2 a

The inertial force at the center of mass along y

d2 1
fIy ¼ m 2
ðCC 0 Þy ¼ m u€ (2d)
dt 2

The inertial moment about the center of mass

u€
MIc ¼ IC f€ ¼ IC (2e)
a

The external force


pðt Þ (2f)

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
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
22 PART I Single-degree-of-freedom systems

which by virtue of Eq. (2) becomes


" #
a 2 b2
m + + IC
2 2 4 W
u€ + ku ¼ + pðt Þ (5)
a2 9 2
Using Steiner’s formula, we have
" #
a 2 b2 a 2 + b2
IO ¼ m + + IC ¼ m (6)
2 2 3

Hence, Eq. (5) becomes


a 2 + b2 4 W
m 2
u€ + ku ¼ + pðt Þ (7)
3a 9 2
Eq. (7) can be also obtained if we consider the motion with reference to point
O and employ Eq. (1.5.14c) for
u€ 2a a
u€O ¼ v€O ¼ 0, f€ ¼ IO , MO ¼ fS + W + pðt Þa, MIo ¼ IO f€
a 3 2
The weight W can be eliminated from Eq. (7), if the total displacement u ðt Þ
is expressed as the sum of the static displacement ust caused by the weight plus
the additional dynamic displacement uðt Þ, as shown in Fig. E1.1c, that is,
u ðt Þ ¼ ust + uðt Þ (8)
The static equilibrium of moments with respect to point O, when the plate is
loaded only by the weight yields (see Fig. E1.1c)
4 a
kau st ¼ W (9)
9 2
Noting that u€st ¼ 0 because ust is a constant, and using Eqs. (8), (9), Eq. (7)
becomes
m ∗ u€
 + k ∗ u ¼ p ∗ ðt Þ (10)
where
a 2 + b2 4
m∗ ¼ m 2
, k ∗ ¼ k, p ∗ ðt Þ ¼ pðt Þ
3a 9
Eq. (10) has the form of Eq. (1.4.8) and represents the equation of motion of
the system. The quantities m ∗ , k ∗ , having dimensions of mass and translational
stiffness, respectively, are referred to as the generalized mass and the general-
ized stiffness of the SDOF system.
If the rotation fðt Þ, measured from the position of static equilibrium, is
taken as the parameter of motion in place of uðt Þ, the equation of motion results
from Eq. (10) using the relation u ¼ fa. Thus, we have
4
IO f€ + ka2 f ¼ apðt Þ (11)
9
General concepts and principles of structural dynamics Chapter 1 23

Example 1.5.2 Equation of motion of a frame with a rigid column


Formulate the equation of motion of the plane frame shown in Fig. E1.2a for
small amplitude motion. The mass of the horizontal beam CD is negligibly small
while the column of height L and nonnegligible width h ¼ L=4 is assumed rigid
 The elastic stiffness of the ground is simulated by the
with total mass m ¼ mL.
spring CR while its damping by the two dashpots with damping parameters c.

(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
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

The elastic shear force at the beam end C QS ¼ ð1:5L


12EI
Þ3
6EI
d + ð1:5L Þ2
f
The damping forces fD ¼ c h2 f_
The external load pðt Þ
The equilibrium of moments with respect to point A yields

MIA + MS + MR + QS  ð3  4Þ + 2fD  ð5  6Þ  pðt Þ  ð1  2Þ ¼ 0

which after substituting their exressions becomes


 
mL2 € cL2 _ 79EI pL 28EI 2
f+ f+  f+ f ¼ Lpðt Þ (1)
3 32 18L 8 9L

Further, linearizing (f2  0) gives


 
mL2 € cL2 _ 79EI pL
f+ f+  f ¼ Lpðt Þ (2)
3 32 18L 8

If the displacement u ¼ Lf at the level of the beam is taken as the parameter


of the motion, the equation of motion becomes
 
m cL 79EI p
u€ + u_ +  u ¼ pðt Þ (3)
3 32 18L3 8L

Example 1.5.3 Equation of motion of a system of rigid bodies


The rigid body assemblage shown in Fig. E1.3a consists of the rigid bar AF of
total mass m hinged at A, and the rigid square plane body supported rigidly at F.
The dynamic excitation of the bar is due to the uniformly distributed transverse
load pðt Þ. The motion is constrained by a spring at B and the damper at G. For-
mulate the equation of motion of the system for small amplitude displacements
using the method of equilibrium of forces. The mass per unit length of the bar is
m ¼ m=3L and the surface mass density of the body is g ¼ 2m=L2
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€
26 PART I Single-degree-of-freedom systems

or taking into account that


ð3LÞ3  m  ð3LÞ3

IK ¼ m ¼ ¼ 0:75mL2
12 3L 12
u u u€
f ¼ ¼ 0:5 , f€ ¼ 0:5
2L L L
we obtain
MIK ¼ 0:375mLu€ (4)
The inertial force fIG and the inertial moment MIG at the center of mass G of
the rigid body due to the mass gL2 =2:
 
L d2 2m L2 d 2
fIG ¼ gL 2
ðGG 0 Þ ¼ 2 ð1:625u Þ ¼ 1:625m u€ (5)
2 dt L 2 dt 2
" # 
€ LðL=2Þ3 ðL=2ÞL3 u€
M I ¼ IG f ¼ g
G
+ 0:5 ¼ 0:052mLu€ (6)
12 12 L

The external load is 2Lpðt Þ.


The equilibrium of the moments about A yields the equation of motion of the
system. Thus, we have

fS  ð2LÞ  fD  ð3:25LÞ  fIK  ð1:5LÞ  fIG  ð3:25LÞ


(7)
MIK  MIG + pðt Þ  ð2LÞ  L ¼ 0

or inserting Eqs. (1)–(6) into Eq. (7) we obtain

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

RA + pðt Þ2L  fIK  fIG  fS  fD ¼ 0

or using Eqs. (1)–(3), (5) we obtain

RA ¼ ku + 1:625cu_ + 2:375m u€  2L


pðt Þ
General concepts and principles of structural dynamics Chapter 1 27

Example 1.5.4 Equation of motion of a single-story shear building


Formulate the equation of motion of the single-story building shown in Fig. E1.4a.
The damping is neglected. The columns are fixed on the ground, are inextensible,
and their mass is assumed to be lumped at their ends. Moreover, the roof plate
is assumed rigid. The material of the structure is reinforced concrete having spe-
cific weight g ¼ 24kN=m3 and modulus of elasticity E ¼ 2:1  107 kN=m2 . The
total load of the plate (dead and live) is 20kN=m2 . The force pðt Þ acts in the direc-
tion of the x axis and is given by pðt Þ ¼ 20sin 13t. The acceleration of gravity is
g ¼ 9:81m=s2 and the dimensions of the rectangular cross-sections of columns are
k1 : 30  30cm2 and k2 : 30  20cm2 .

(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
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

5:21u€ + 708:75u ¼ sin 13t

Example 1.5.5 Equation of motion of a two-story shear frame


Formulate the equations of motion of the two-story shear frame shown in
Fig. E1.5a using the method of equilibrium of forces. The damping is ignored.
Solution
The system has two degrees of freedom because the girders are rigid and the
axial deformation of columns is ignored. The model of the structure is shown
in Fig. E1.5b. The masses are lumped at the story levels. The motion of the
system can be fully determined from the horizontal displacements u1 ðt Þ and
u2 ðt Þ of the masses m1 and m2 , respectively.

Rigid

Rigid

(a) (b) (c)


FIG. E1.5 Two-story shear frame in Example 1.5.5.
General concepts and principles of structural dynamics Chapter 1 29

The equations of motion result from the dynamic equilibrium of forces


applied to the masses m1 and m2 . These forces are shown in the free body dia-
grams of Fig. E1.5c. Thus, we obtain

m1 u€1 + k1 ðu1  u2 Þ ¼ p1 ðt Þ (1)

m2 u€2  k1 ðu1  u2 Þ + k2 u2 ¼ p2 ðt Þ (2)

Eqs. (1), (2) are written in matrix form as

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

Example 1.5.6 Equation of motion of a general single-story shear building


The rigid horizontal plate is supported by K columns as shown in Fig. E1.6. The
columns are fixed on the ground as well as on the plate. Their principal axes have
arbitrary directions in the xy plane. Formulate the equation of motion of the
plate when the plate is loaded by the horizontal load P  ðt Þ through the point A.

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.
30 PART I Single-degree-of-freedom systems

FIG. E1.6 General single-story shear building in Example 1.5.6.

(a) In global axes:


8 9
< U =
 ¼ V displacements of point O of the plate
U
:;
W
8 9
< ui =
 i ¼ vi displacements of i column
D
: i;
w
8 i9
< X =
 i ¼ Y i elastic forces of i column
F S
: i;
M
(b) In local axes
8 9
< ui =
D ¼ v i displacements of i column
i
: i;
w
8 9
< Xi =
FiS ¼ Y i elastic forces of i column
: i;
M
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)
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

is the stiffness matrix of i column.


Eq. (5) is transformed in global axes using Eqs. (2b), (1a). Thus, we have
 
 i ¼ Ri T F i
F S S
 T
¼ Ri k i D i
 T
i
¼ Ri k i R i D

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.
General concepts and principles of structural dynamics Chapter 1 33

FIG. E1.7 Displacements of point i due to rotation of plate.

The point i of the plate undergoes translational displacements due to


(a) The translational displacements of point O
 i
u t ¼ U
 i
v ¼ V
t

(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
34 PART I Single-degree-of-freedom systems

we can write Eq. (11) as


 i ¼ ei U
D  (12)
The matrix e defined by Eq. (12) is referred to as the translation matrix or
i

transformation matrix of the i column.


The equations of motion result from Eqs. (1.5.14a)–(1.5.14c) if they are
employed for point O, with P  O, up ¼ U , vp ¼ V , f ¼ 
W. The external force
 ðt Þ ¼ P x ðt Þ, P y ðt Þ T . Thus,
F is equal to the sum of all elastic forces FiS plus P
we have
X
K  
P x ðt Þ  X ¼ m U€
i €
  yc W (13a)
i¼1

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

where M  ðt Þ with respect


 ðt Þ ¼ xA P y ðt Þ  yA P x ðt Þ is the moment of the force P
to O, m is the mass of the plate, and Io its moment of inertia with respect to O.
Eqs. (13a)–(13c) are written in matrix form as
8 9 2 38 i 9 2 32 32 38 9

< P x ðt Þ >
> = XK 1 0 0 > < X >= 1 0 0 m 0 0 yc >
1 0  < U >
=
6 7 6 76 76 7
P y ðt Þ  4 0 1 0 5 Y i
¼ 4 0 1 0 54 0 m 0 54 0 1 xc 5 V€
>
:  >
; i¼1 >
: i> ; >
: € >
;
M ðt Þ y i xi 1
 M y c xc 1
 0 0 Ic 0 0 1 W

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)
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    

K ei
T T
Ri ki Ri ei (17b)
i¼1

are the mass and stiffness matrices of the structure, respectively.


The equation of motion (16) can be transformed with reference to the center
of mass by working as follows.
Using Eq. (12), we relate the displacements of the center of mass C to the
displacements of O. Hence, we have
U 
 c ¼ ec U (18)
which can be inverted to give
 ¼ ðec Þ1 U
U c (19)
We can readily show that
2 3
1 0 yc
ðec Þ1 ¼ 4 0 1 xc 5 (20)
0 0 1
Substituting Eq. (19) into Eq. (16) and premultiplying it byðec ÞT , we
obtaind
€ + K
c
mU  cU  ðt Þc
c ¼P (21)
where
 c ¼ ðec ÞT K
K  ðec Þ1 (22a)

 c ðt Þ ¼ ðec ÞT P
P  ðt Þ (22b)
Eq. (22a) represents the transformed stiffness matrix of the structure from
point O to the center of mass C .
The stiffness matrix with respect to an arbitrary point O is, in general, a full
3  3 matrix, namely
2 3
k11 k12 k13
 ¼6
K
7
4 k21 k22 k23 5 (23)
k31 k32 k33

T
d. The notation eT ¼ ðe1 Þ is employed.
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

k11 yE  k12 xE + k13 ¼ 0


k21 yE  k22 xE + k23 ¼ 0
from which we establish the coordinates of E

 
k 11 k 13
k k
xE ¼
21 23 (24a)
k 11 k 12
k21 k22


k 12 k13
k k23
yE ¼
22 (24b)
k 11 k12
k21 k22
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

The coupling is now limited between the translational displacements and


the corresponding elastic forces. They can also be decoupled if the matrix is
transformed into a new system of axes x 0 Ey 0 , resulting from xE
 y by rotation
through an angle q and demanding the off-diagonal elements to vanish. The
stiffness matrix in the new system is obtained using Eq. (7) for ki ¼ K
 E and
2 3
cos q sin q 0
R ¼ 4  sin q cos q 0 5
0 0 1
Thus, we have
K  ER
^ E ¼ RT K (25)
or after performing the matrix multiplications
2   3
k11  k22
6 k cos 2 q + k sin 2 q  k sin 2q 
sin 2q + k 12 cos2q 0 7
6 11 22 12 2 7
^ 6  7
K ¼ 6 k 11  k 22
E 7
6
4 sin 2q + k12 cos 2q k11 sin q + k22 cos q + k12 sin 2q 0
2 2 7
5
2
0 0 k13 y2 + k23 x 2 + k33
E E

The vanishing of the off-diagonal elements yields

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)
38 PART I Single-degree-of-freedom systems

The previous analysis allows us to draw the following conclusions:


(a) In static analysis, the concepts of the elastic center and the principal direc-
tions permit the uncoupling of the three equations of static equilibrium and
give a better insight into the deformation of the structure. These concepts,
however, do not have any advantage in the dynamic analysis because, in
general, the center of mass does not coincide with the elastic center and
therefore the equations of motion remain coupled through the second deriv-
atives of the displacements. Apparently, this fact complicates the study of
the dynamic analysis.
(b) The equations of motion can be decoupled with respect to the physical dis-
placements U , V , W only if the center of mass coincides with the elastic
center. This uncoupling should not be confused with that achieved via
modal coordinates (see Chapter 12).

1.6 Principle of virtual displacements


D’Alembert’s principle allows the application of the principle of virtual dis-
placements to formulate the equations of motion of structural systems, espe-
cially for complex assemblages comprising a number of interconnected
particles or rigid bodies of finite size. The principle of virtual displacements
or virtual works can be expressed as follows:
The necessary and sufficient condition for the dynamic equilibrium of a
system is the vanishing of the total work done by the set of all externally applied
forces (actual and inertial) when the system is subjected to a virtual displace-
ment, that is, a displacement pattern compatible with the geometrical con-
straints of the system.
Thus, the equations of motion of the dynamic system can be derived by first
identifying all forces (imposed external forces, elastic forces, damping forces,
and inertial forces) acting on the masses. Then by introducing a virtual displace-
ment corresponding to each degree of freedom, the equations of motion are
obtained by setting the virtual work produced by all forces equal to zero. A
major advantage of this approach is that the contribution of the work done
by the reactions of nonyielding bilateral supports as well as the internal inter-
action forces on the separated masses do not appear explicitly in the equations.
Moreover, the quantities we have to deal with are scalar and not vectorial, thus
they can be algebraically manipulated. An important provision for the applica-
tion of the principle of virtual displacements is that the masses of the system are
subjected to small displacements. This implies that the geometry of the structure
remains essentially unchanged after the action of the displacements.
Example 1.6.1 Equation of motion of a complex SDOF system
Formulate the equation of motion of the system shown in Fig. E1.8a by using the
principle of virtual displacements for small amplitude motion. It is assumed that
the cable can undertake compression.
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
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
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,

MIA df1  fD dðBB 0 Þ + fS1 dðCC 0 Þ  fS1 dðEE 0 Þ  fS2 dðDD 0 Þ


(1)
MIF df2 + pðt Þd ðCC 0 Þ ¼ 0

Introducing the expressions of the forces and virtual displacements into


terms of the basic displacements in Eq. (1) yields

 

2ma u_ 1
 u€1  c + k ðu2  u1 Þ + pðt Þ du1
3 4

(2)
8ma
+  u€2  k ðu1  u2 Þ  8ku 2 du2 ¼ 0
3

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
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

1.7 Hamilton’s principle


The development of dynamics and generally of mechanics has been accom-
plished through two different approaches. The first is based on Newton’s laws
of motion. These laws deal with the motion of a body under the action of forces
acting on it. The involved quantities are (i) the imposed forces, which may be
externally applied forces, forces of interaction between the masses, and reactions
of constraints and (ii) the momentum or the quantity of motion as defined by
Newton. Because both quantities are vector quantities, this approach of mechan-
ics is called vectorial mechanics. The analysis of complicated systems by direct
application of Newton’s laws of motion becomes increasingly difficult. The prin-
cipal reason is that the equations are vectorial in nature and the forces and accel-
erations are often difficult to determine. Moreover, the reactions of the constraints
and the interaction forces between bodies must be explicitly accounted in the
equations of motion and have to be evaluated even when there is no interest to
evaluate them. In addition, each problem seems to require its own particular insi-
ghts and there are no general procedures for obtaining the equations of motion.
The second approach is based mainly on the work of Lagrange and Hamilton
and is called analytical mechanics. In this approach, the involved quantities are
scalar functions, and therefore the fundamental equations, in contrast to vectorial
mechanics, do not depend on the choice of the coordinates. Also, it is not neces-
sary to include explicitly the forces of the constraints and the interaction forces. It
will be shown that this approach circumvents to some extent the difficulties found
in the direct application of Newton’s law of motion to complicated systems. Fur-
thermore, the equations of motion are formulated in a standard convenient form.
Analytical dynamics is based on Hamilton’s principle and Lagrange’s equations.
Hamilton’s principle is presented in this section. Lagrange’s equations are pre-
sented in the next section resulting directly from Hamilton’s principle.
One of the most important principles of dynamics is Hamilton’s principle,
named after the famous Irish mathematician and physicist Sir William Rowan
Hamilton (1805–65). Inertial and elastic forces are not explicitly involved in
this principle; instead, variations of the kinetic and potential energy are utilized.
This formulation has the advantage of dealing only with purely scalar quanti-
ties. In the procedure of virtual displacements, even though the works them-
selves are scalar quantities, vector quantities, displacements, and forces are
utilized to represent them. Hamilton’s principle is presented here for discrete
parameter systems.
General concepts and principles of structural dynamics Chapter 1 43

FIG. 1.7.1 Particle moving in space.

Consider a particle of mass m moving in space under the action of a force


Fðt Þ as shown in Fig. 1.7.1. If r ¼ rðt Þ ¼ x ðt Þi + y ðt Þj + z ðt Þk represents the
position vector of the particle at time t, then according to Newton’s second
law of motion, the Newtonian path of the particle is governed by the differential
equation

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,

drðt1 Þ ¼ drðt2 Þ ¼ 0 (1.7.2)

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
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.
General concepts and principles of structural dynamics Chapter 1 45

In the absence of nonconservative forces, Fnc ¼ 0, Eq. (1.7.9) becomes


Z t2
d ðT  AÞdt ¼ 0 (1.7.10)
t1

The scalar quantity


L¼T A (1.7.11)
is termed the Lagrangian or the kinetic potential. We should emphasize that
Hamilton’s principle depends upon the energies of the system and is invariant
under the coordinate transformation.
Eq. (1.7.10) states that of all possible paths of motion of the particle during
an interval of time from t1 to t2 , the actual path is that for which the integral
Z t2
Ldt ¼ 0 (1.7.12)
t1

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

Hamilton’s principle is rather utilized to derive the equations of motion of


continuous systems. The equations of motion of discrete parameter systems can
result directly from Lagrange’s equations.
Example 1.7.1 Equation of motion of the SDOF system
Formulate the equation of motion of the SDOF system shown in Fig. 1.4.1 using
Hamilton’s principle.
Solution
The potential energy is due to the strain energy stored in the spring during defor-
mation. It is expressed in terms of the spring stiffness coefficient k and the dis-
placement u as
1
U ¼ ku 2 (1)
2
The kinetic energy is due to the motion of the mass m and is given as
1
T ¼ m u_ 2 (2)
2
46 PART I Single-degree-of-freedom systems

The damping force fD ¼ cu,_ as a dissipative force, is nonconservative. The


virtual work of this force is
D
dWnc _
¼ fD du ¼ cudu (3)
The negative sign results from the fact that fD is opposite to the virtual dis-
placement du.
The external force is also treated as nonconservative and it does the
virtual work
p
dWnc ¼ pðt Þdu (4)
Because no conservative external forces act on the system, it is A ¼ 0.
The variations dU and dT are obtained from Eqs. (1), (2)
_ u_
dU ¼ kudu, dT ¼ m ud (5)
Introducing Eqs. (3)–(5) into Hamilton’s principle, Eq. (1.7.13), yields
Z t2 Z t2
_ u_ Þdt 
ðkudu  m ud _ + pðt Þdu dt ¼ 0
½cudu (6)
t1 t1

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

According to Hamilton’s principle it holds


du ðt1 Þ ¼ du ðt2 Þ ¼ 0
Thus, the term outside the integral vanishes and Eq. (7) becomes
Z t2 Z t2
_ udt
m ud _ ¼ €
m ududt (8)
t1 t1

Eq. (6) by virtue of Eq. (8) is written as


Z t2
½m u€ + cu_ + ku  pðt Þdudt ¼ 0 (9)
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
General concepts and principles of structural dynamics Chapter 1 47

Moreover, because du is arbitrary, it must be


m u€ + cu_ + ku  pðt Þ ¼ 0
or
m u€ + cu_ + ku ¼ pðt Þ (10)
which is the equation of motion.
Example 1.7.2 Equation of motion of a two-story shear frame
Formulate the equations of motion of the frame in Example 1.5.5 using Hamil-
ton’s principle.
Solution
Referring to Fig. E1.5b and c in Example 1.5.5, we have
1 1
U ¼ k1 ðu1  u2 Þ2 + k2 u22
2 2
1 1
T ¼ m1 u_ 21 + m2 u_ 22
2 2
Their variations are
dU ¼ k1 ðu1  u2 Þðdu1  du2 Þ + k2 u2 du2
(1)
¼ k1 ðu1  u2 Þdu1  k1 ðu1  u2 Þdu2 + k2 u2 du2
dT ¼ m1 u_ 1 d u_ 1 + m2 u_ 2 d u_ 2 (2)
Integrating by parts the variation dT in the interval ½t1 , t2  yields
Z t2 Z t2
dTdt ¼ ðm1 u_ 1 d u_ 1 + m2 u_ 2 d u_ 2 Þdt
t1 t1
Z t2
¼ ½m1 u_ 1 du1 + m2 u_ 2 du2 tt21  ðm1 u€1 du1 + m2 u€2 du2 Þdt
t1

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
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)

Example 1.7.3 Equation of motion of a complex MDOF system


The system shown in Fig. E1.9 consists of the three rigid bars AB,BC , CD con-
nected by hinges at points B and C , and it is supported by a roller at point D and
a hinge at point A. The relative rotations of the bars at the hinges B and C are
restrained by moment-resisting rotational springs with stiffness coefficients
k3 ¼ k4 ¼ 4kL2 and by the rotational dashpots with damping coefficients
c3 ¼ c4 ¼ 2cL2 . In the transverse direction, the motion is restrained by the
two springs at points E and Q with stiffness coefficients k1 ¼ k, k2 ¼ 2k, and
the two dashpots at points F and G with damping coefficients c1 ¼ c and
c2 ¼ 3c. A constant axial force P is applied at point D. The system is set in
motion by the transverse load pðx, t Þ ¼ ðpx=LÞf ðt Þ, linearly distributed along
the bar CD. The mass per unit length of the bars AB and CD is m  while the
bar BC is massless and supports the rigid body S at H having surface mass

density g ¼ m=L. Assuming small amplitude displacements, formulate the
equations of motion of the system using Hamilton’s principle.

FIG. E1.10 System in Example 1.7.3.


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

FIG. E1.11 Deformed configuration of the system.

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
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

and its variation

dT ¼ m ð1:796u_ 1 + 0:204u_ 2 Þd u_ 1 + m ð0:204u_ 1 + 1:130u_ 2 Þd u_ 2


Rt
Using integration by parts in the integral t12 dTdt and taking into
account that
  d
du1 ðt1 Þ ¼ du1 ðt2 Þ ¼ du2 ðt1 Þ ¼ du2 ðt2 Þ ¼ 0 and d u_ ¼ d du
dt ¼ dt ðdu Þ

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)
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
52 PART I Single-degree-of-freedom systems

They may become zero if the system of equations



 
1:868  0:583l 1:667 + 0:333l u1 0
k ¼ (14)
1:667 + 0:333l 2:722  0:667l u2 0
has a nontrivial solution. This occurs if
 
 +1:868  0:583l 1:667 + 0:333l 
 
 1:667 + 0:333l +2:722  0:667l  ¼ 0 (15)

Expanding the determinant yields


0:27797l2  1:7227l + 2:3058 ¼ 0
from which we obtain
l1 ¼ 1:9555 l2 ¼ 4:2419
The obtained values of l specify two critical values, Pcr1 ¼ 1:9555kL and
Pcr2 ¼ 4:2419kL, of the compressive axial force for which the structure exhibits
no resistance to deformation, that is, it has no stiffness and the structure buckles.
Therefore, these critical loads are the buckling loads of the structure (first and
second). The condition for buckling is the vanishing of the determinant of the
stiffness matrix, Eq. (15). The resulting equation is called the buckling equation.
It is apparent that if the axial force is tensile, the determinant cannot vanish for
real values of the parameter l. Concluding, we can state that tensile axial forces
increase the stiffness of the structure while compressive axial forces reduce it
and may lead to buckling.
Example 1.7.4 Equation of motion of the elastic cantilever beam
Derive the equation of motion of the cantilever beam shown in Fig. E1.12.
Solution
The mass of the beam element is equal to mdx. Its kinetic energy is
mdx ½∂u ðx, t Þ=∂t 2 =2, which is integrated along the beam length to yield the
kinetic energy of the beam, namely

FIG. E1.12 Cantilever in Example 1.7.4.

Z  
1 L
∂u ðx, t Þ 2
T¼ m dx (1)
2 0 ∂t
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

Integrating twice by parts the integral representing dU yields


Z L
EI u 0000 dudx  ½EI u 000 du 0 + ½EI u 00 du 0 0
L L
dU ¼ (8)
0

The boundary conditions of the beam are


At x ¼ 0 u ¼ u 0 ¼ 0, hence du ¼ du 0 ¼ 0
At x ¼ L M ¼ EI u 00 ¼ 0 Q ¼ EI u 000 ¼ 0
Therefore, the quantities outside the integral vanish and Eq. (8) becomes
Z L
dU ¼ EI u 0000 dudx (9)
0
54 PART I Single-degree-of-freedom systems

Because no conservative loads act on the system, it is A ¼ 0. Moreover, the


virtual work of the external load is
Z L
dWnc ¼
p
pðx, t Þdudx (10)
0

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

Interchanging the integration in the second term and performing integration


by parts with respect to time, yield
Z t2
Z L Z L
Z t2
_ udx
m ud _ dt ¼  €
½m ududt  + ½mudu tt21 dx
t1 0 0 t1
Z t
Z L 2
(12)
¼ €
½m ududx  dt
t1 0

On the basis of Eq. (12), Eq. (11) becomes


Z t 2
Z L
½EI u 0000 + m u€  pðx, t Þdudx dt ¼ 0 (13)
t1 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

Moreover, because du is arbitrary, Eq. (14) is valid only if


EI u 0000 + m u€  pðx, t Þ ¼ 0 (15)
which yields the equation of motion of the cantilever
EI u 0000 + m u€ ¼ pðx, t Þ (16)
Apparently, Eq. 16 is identical to that obtained in Section 1.1.

1.8 Lagrange’s equations


1.8.1 Derivation of Lagrange’s equations
In a system with N degrees of freedom, the displaced configuration can be deter-
mined from a set of coordinates, which take a certain value at each instant. The
system of coordinates for the analysis of a given mechanical system is not nec-
essarily unique. Many coordinate systems are possible. Furthermore, the number
of coordinates may vary, but it cannot be less than N . Anyhow, if the number of
coordinates is greater than N , then additional equations, referred to as equations
General concepts and principles of structural dynamics Chapter 1 55

of constraint, must relate the coordinates so that the number of coordinates is


equal to the number of degrees of freedom plus the independent equations of
constraint. The requirement that the equations of motion hold together with
the equations of constraint complicates the solution. For this reason, we seek,
if possible, to choose N independent coordinates, which can specify the config-
uration of the system. For example, we consider the simple pendulum shown in
Fig. 1.8.1a. The rod is rigid and weightless. Its length is L and it can rotate freely
about the hinge at O, such that the motion is confined in a single vertical plane.
The position of the mass can be specified by the angle q between the vertical
axis y and the rod. Hence, the system has a SDOF. However, the displaced
configuration can also be determined by the coordinates ðx, y Þ, which represent
the position of the mass m within the xy plane. These coordinates, however,
are not independent because they must satisfy the constraint equation
x 2 + y 2 ¼ L2

(a) (b)
FIG. 1.8.1 Simple (a) and double (b) pendulum.

Similarly, the configuration of the double pendulum of Fig. 1.8.1b can be


specified by the two angles q1 and q2 . Hence the system has two degrees of free-
dom. On the other hand, the position of the masses m1 , m2 can be determined by
the coordinates ðx1 , y1 Þ and ðx2 , y2 Þ, which, however, are not independent
because they must satisfy the following two constraint equations
x12 + y12 ¼ L21

ðx2  x1 Þ2 + ðy2  y1 Þ2 ¼ L22


The quantities q in the simple pendulum or q1 , q2 in the double pendulum,
which would determine the configuration of the system, could be considered as
coordinates in a more general sense. Any set of quantities that serves to specify
the configuration of the system is referred to as generalized coordinates. The
geometrical significance of the generalized quantities is not always cognizable.
For systems in motion, the generalized coordinates vary with time and are trea-
ted as algebraic variables. The process of obtaining one set of generalized coor-
dinates from another is known as a coordinate transformation.
56 PART I Single-degree-of-freedom systems

We consider now a transformation from a set of N generalized coordinates


q1 ðt Þ, q2 ðt Þ,…, qN ðt Þ to a set K of ordinary (for example, Cartesian) coordinates
x1 ,x2 ,…, xK ðK  N Þ. The transformation equations are of the form
x1 ¼ x1 ðq1 , q2 , …, qN Þ
x2 ¼ x2 ðq1 , q2 , …, qN Þ
(1.8.1)
⋯ ⋯
xK ¼ xK ðq1 , q2 , …, qN Þ
For example, the transformation equations of the generalized coordinates
q1 , q2 to the ordinary coordinates x1 , x1 , y2 , y2 of the double pendulum are
x1 ¼ L1 sin q1
y1 ¼ L1 cos q1
x2 ¼ L1 sin q1 + L2 sin q2
y2 ¼ L1 cos q1  L2 cos q2
The kinetic energy of a system with K degrees of freedom may also depend
on the generalized coordinates q1 , q2 , …, qN beside the generalized velocities
q_ 1 , q_ 2 ,…, q_ N , that is,
T ¼ T ðq1 , q2 , …, qN , q_ 1 , q_ 2 , …, q_ N Þ (1.8.2)
In conservative systems, the potential energy A depends only on the posi-
tion, namely, it is
A ¼ Aðq1 , q2 , …, qN Þ (1.8.3)
The work done by the forces derivable from the potential energy A, when the
generalized coordinates qi are given a virtual displacement dqi , is expressed as
dA ¼ Q1 dq1 + Q2 dq2 + ⋯ + QN dqN
where
∂A ∂A ∂A
Q1 ¼  ¼ , Q2 ¼  ¼ , …, QN ¼  (1.8.4)
∂q1 ∂q2 ∂qN
The quantity Qi dqi represents the work done through the displacement dqi .
Inasmuch as the quantity Qi may or may not represent a force, it is referred to as
generalized force. Hence, if qi represents a translational displacement then Qi is
a force, whereas if qi represents a rotation then Qi is a moment. In some prob-
lems, the quantities dqi may represent surfaces, volumes, etc. Therefore, the
nature of the corresponding Qi is defined so that the quantity Qi dqi has the phys-
ical dimension of work.
Lagrange’s equations may be derived by direct application of Hamilton’s
principle. Thus, when the applied forces are conservative, we write
Z t2
ðdT  dAÞdt ¼ 0 (1.8.5)
t1
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
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)

is the total potential energy of the system.


The N generalized forces Qi can be evaluated from the set of K actual forces
Fk associated with the set of Cartesian coordinates. For this purpose, we con-
sider the work done by the set of forces Qi when the coordinates qi are given an
increment, that is, a virtual displacement dqi

X
N
dW ¼ Qi dqi (1.8.13)
i¼1

If dxk represent the ensuing virtual displacements of the coordinates xk then


the set of forces Fk do the work
X
K
dW ¼ Fk dxk (1.8.14)
k¼1

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

Q ¼ fQ1 Q2 ⋯ QN gT , dq ¼ f dq1 dq2 ⋯ dqN gT (1.8.17a)

F ¼ f F1 F2 ⋯ FK gT , dx ¼ f dx1 dx2 ⋯ dxK gT (1.8.17b)


The relation between dxk and dqi results from the transformation equations
(1.8.1) by considering the variation of dxk . Thus, we have
∂xk ∂xk
dxk ¼ dq1 + ⋯ + dqN (1.8.18)
∂q1 ∂qN
or in matrix form
dx ¼ Jdq (1.8.19)
General concepts and principles of structural dynamics Chapter 1 59

where J is the Jacobian matrix of the transformation (1.8.1), that is,


2 ∂x ∂x ∂x1 3
1 1

6 ∂q1 ∂q2 ∂qN 7
6 7
6 ∂x ∂x ∂x2 7
6 2 2
⋯ 7
6 7
J ¼ 6 ∂q1 ∂q2 ∂qN 7 (1.8.20)
6 7
6 ⋮ ⋮ ⋱ ⋮ 7
6 7
4 ∂xK ∂xK ∂xK 5

∂q1 ∂q2 ∂qN
Substituting Eq. (1.8.19) into (1.8.16) yields
QT dq ¼ FT Jdq (1.8.21)
from which we obtain
QT ¼ F T J (1.8.22)
or
Q ¼ JT F (1.8.23)
Apparently, Eq. (1.8.23) represents the sought relation between Q and F.
Example 1.8.1 Equation of motion of the double pendulum
Formulate the equations of motion of the double pendulum shown in
Fig. 1.8.1b.
Solution
Because the bars are inextensional, the displaced configuration of the moving
system can be specified by the generalized coordinates q1 and q2 . Referring to
Fig. 1.8.1b, the Cartesian coordinates of the masses m1 and m2 are expressed in
terms of q1 and q2 by the geometrical relations
x1 ¼ L1 sin q1
y1 ¼ L1 cos q1
(1)
x2 ¼ L1 sin q1 + L2 sin q2
y2 ¼ L1 cos q1  L2 cos q2
The kinetic energy of the system is
1   1  
T ¼ m1 x_ 21 + y_ 21 + m2 x_ 22 + y_ 22
2 2
which by virtue of Eqs. (1) becomes
1 1
T ¼ ðm1 + m2 ÞL21 q_ 1 + m2 L1 L2 q_ 1 q_ 2 cos ðq1  q2 Þ + m2 L22 q_ 2
2 2
(2)
2 2
The potential energy is
A ¼ m1 gy 1 + m2 gy 2
60 PART I Single-degree-of-freedom systems

or using Eqs. (1)


A ¼ ðm1 + m2 ÞgL1 cos q1  m2 gL2 cos q2 (3)

Differentiating Eqs. (2), (3) yields


∂T
¼ ðm1 + m2 ÞL21 q_ 1 + m2 L1 L2 q_ 2 cos ðq1  q2 Þ
∂q_ 1
 
d ∂T
¼ ðm1 + m2 ÞL21 q€1 + m2 L1 L2 q€2 cos ðq1  q2 Þ
dt ∂q_ 1
 
 m2 L1 L2 q_ 2 sin ðq1  q2 Þ q_ 1  q_ 2
∂T
¼ m2 L1 L2 q_ 1 q_ 2 sin ðq1  q2 Þ
∂q1
∂A
¼ ðm1 + m2 ÞgL1 sin q1
∂q1
∂T
¼ m2 L1 L2 q_ 1 cos ðq1  q2 Þ + m2 L22 q_ 2
∂q_ 2
 
d ∂T  
¼ m2 L1 L2 q€1 cos ðq1  q2 Þ  m2 L1 L2 q_ 1 sin ðq1  q2 Þ q_ 1  q_ 2 + m2 L22 q€2
_
dt ∂q 2

∂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

L1 q€1 cos a + L2 q€2  L1 q_ 1 sin a + g sin q2 ¼ 0


2
(4b)
where
a ¼ q1  q2

Example 1.8.2 Equation of motion of the “soft” pendulum


Formulate the equations of motion of the simple pendulum shown in Fig. E1.13,
taking into account the axial deformation of the rod (soft pendulum). The unde-
formed length of the rod is L, its cross-sectional area A, and the modulus of
elasticity of the material E.
General concepts and principles of structural dynamics Chapter 1 61

FIG. E1.13 “Soft” pendulum in Example 1.8.2.

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
62 PART I Single-degree-of-freedom systems

!
EA L
m x€ + 1  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi y ¼ mg (8b)
L x 2 + y2

Example 1.8.3 Equation of motion of a general one-story shear building


Formulate the equations of motion of the one-story building in Example 1.5.6
using the method of Lagrange’s equations.
Solution
The system has three degrees of freedom. We choose the displacements U , V of
 about the z axis as the gen-
point O in the plane of the plate and its rotation W
eralized coordinates of the system, namely

q1 ¼ U , q2 ¼ V , q3 ¼ W
Because the point O is not the center of mass of the plate, the kinetic energy
is given by the expression (1.5.18) (K€onig’s theorem)

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

Taking into account that (see Eqs. 1a, 12 of Example 1.5.6)


 i T    
 ¼ Ri ei U,
D i ¼ Ri D  D ¼U T e i T Ri T
Eq. (2) is written as
1  T i 
Ui ¼ U kU
2
where
   
i ¼ ei T Ri T ki Ri ei
k
Thus, we have
XK
1
U¼ 
Ui ¼ UT KU
i¼1
2
(3)
1  + 2k23 V W
 2 + 2k12 U V + 2k13 U W 

¼ k11 U + k22 V + k33 W
2 2
2
General concepts and principles of structural dynamics Chapter 1 63

ij are the elements of the matrix


where k
X
K X
K    

K i ¼
k ei
T T
Ri k i Ri e i (4)
i¼1 i¼1

Because there are no external conservative forces, it is


A¼0 (5)
The generalized forces result from Eq. (1.8.20). In this case it is
T
F ¼ P x ðt Þ P y ðt Þ 0
The transformation relations between the displacements u, v, w of a point
 y of the plate and the displacements of point O are
x,
 v ¼ V + xW,
u ¼ U  yW,  w¼W


which yields the Jacobian matrix, Eq. (1.8.20)


2 ∂u ∂u ∂u 3
6 ∂U ∂V ∂W
7 2 3
6 7 1 0 0
6 ∂v ∂v ∂v 7
J¼6
6 ∂U
7 ¼ 4 0 1 05
6 7
∂V ∂W 7
4 5 y x 1

∂w ∂w ∂w
∂U ∂V ∂W

Thus, for point A we have
T
Q ¼ JT F ¼ P x ðt Þ P y ðt Þ 
y A P x ðt Þ + xA P y ðt Þ (6)
Differentiation of Eq. (1) yields
 
d ∂T €
¼ m U€
  m yc W (7a)
dt ∂U_
∂T
¼0 (7b)
∂U
!
d ∂T €
¼ m V€
 + mx c W (7c)
dt ∂ V

∂T
¼0 (7d)
∂V
 
d ∂T €  m y U€
 + m xc V€
¼ Io W (7e)
_
dt ∂W c

∂T
∂W ¼0 (7f)

Moreover, Eq. (3) yields


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)

Introducing Eqs. (6)–(8) into the Lagrange


 equations (1.8.11) with N ¼ 3,
 ¼ 0, yields the equation of motion
and taking into account that A U , V , W
of the structure

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


As was anticipated, Eq. (9) is identical to Eq. (16) of Example 1.5.6.

1.8.2 Lagrange multipliers


Lagrange’s equations result as a direct application of Hamilton’s principle pro-
vided that the energies (kinetic and potential) as well as the virtual work of the
nonconservative forces can be expressed in terms of the generalized coordinates
and velocities, as is indicated in Eqs. (1.8.2), (1.8.3), (1.8.8), (1.8.10).
Lagrange’s equations apply to linear as well as to nonlinear systems.
In certain cases, it is impossible or it is not convenient to choose N indepen-
dent coordinates. Then, we choose K > N coordinates and we introduce
n ¼ K  N constraint equations, which in general have the form
g1 ðq1 , q2 , …, qK Þ ¼ 0
g2 ðq1 , q2 , …, qK Þ ¼ 0
(1.8.24)
… …
gn ðq1 , q2 , …, qK Þ ¼ 0
We distinguish two approaches to derive the equations of motion:
General concepts and principles of structural dynamics Chapter 1 65

(a) Eq. (1.8.24) can be solved in terms of n ¼ K  N coordinates. Then the


redundant coordinates can be eliminated from Eqs. (1.8.2), (1.8.3),
(1.8.8), (1.8.10), and the equations of motion are formulated using
Lagrange’s equations (1.8.11). If the constraint equations are linear, the
technique presented in Section 1.8.1 can also be employed to reduce the
number of equations to N .
(b) Eq. (1.8.24) cannot be solved in terms of n ¼ K  N coordinates. In this
case, the equations of motion can be derived by using the method of
Lagrange multipliers.
The second approach preserves the symmetry of the problem because there are
no preferred coordinates while others are eliminated. Though the method of
Lagrange multipliers deals with more coordinates than the degrees of freedom
of the system, quite often this procedure results in simpler equations.
To illustrate the method of Lagrange multipliers, we consider the variations
of the constraint functions given by Eq. (1.8.24)
dg1 dg1
dg1 ¼ dq1 + ⋯ + dqK ¼ 0
dq1 dqk
dg2 dg2
dg2 ¼ dq1 + ⋯ + dqK ¼ 0 (1.8.25)
dq1 dqk
… …
dgn dgn
dgn ¼ dq1 + ⋯ + dqK ¼ 0
dq1 dqk
which we write as
X
K
aji dqi ¼ 0 ðj ¼ 1, 2, …, n Þ (1.8.26)
i¼1

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

Now we multiply Eq. (1.8.26) by a factor known as the Lagrange multiplier.


Thus, we obtain
XK
lj aji dqi ¼ 0 ðj ¼ 1, 2, …, n Þ (1.8.29)
i¼1
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

from which, noting that dqi are arbitrary, we conclude that


X n Xn
dgj
Ri ¼ lj aji ¼ lj (1.8.31)
j¼1 j¼1
dqi

The constraint forces Ri constitute additional generalized forces, which


must be included in Lagrange’s equations. Thus Eq. (1.8.11) become
  X n
d ∂T ∂T ∂V dgj
 + ¼ Qi + lj ði ¼ 1, 2, …, K Þ (1.8.32)
dt ∂q_ i ∂qi ∂qi j¼1
dqi

What we have accomplished by this procedure is the inclusion of the con-


straint reactions in the equations of motion as additional generalized forces.
Therefore, the number of unknowns becomes K + n, namely the K generalized
coordinates Qi ðt Þ and the n functions lj ðt Þ. The available equations are also
K + n, that is Eq. (1.8.24) plus Eq. (1.8.32).
Eq. (1.8.32) can be derived from Hamilton’s principle if the potential energy
of the external forces is replaced by
Xn
A ¼A lj gj (1.8.33)
j¼1

The function A is referred to as the reduced potential energy. For a more


advanced formulation, including nonholonomic constraints and a dynamic
treatment of the Lagrange multipliers, see Ref. [7].
Example 1.8.4 Equation of motion of the simple pendulum using Lagrange
multipliers
Formulate the equations of motion of the simple pendulum shown in Fig. E1.14
in terms of the Cartesian coordinates x ðt Þ, y ðt Þ, assuming that the rod is weight-
less and rigid.

FIG. E1.14 Pendulum in Example 1.8.4.


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
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

1.8.3 Small displacements


So far, we have not imposed any restrictions on the magnitude of the displace-
ments in our discussion on Lagrange’s equations. Therefore, the derived equa-
tions of motion hold equally for small and large displacements and they are in
general nonlinear differential equations. For small displacements, however,
about the position of static equilibrium, the equations of motions are highly sim-
plified as they become linear. This is very important in structural dynamics,
where we usually deal with small displacements and the resulting linear differ-
ential equations can be readily solved and predict the response of the structure.

1.8.3.1 Potential energy and stiffness matrix


We shall consider a system of particles that is in static equilibrium under the
action of a set of conservative forces. If its configuration is specified by ordinary
coordinates x1 ,x2 , …, x3 N , then the condition for static equilibrium is that the
virtual work done by the applied forces Fi is zero, that is,
X
3N
dW ¼ Fi dxi ¼ 0 (1.8.34)
i¼1

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
General concepts and principles of structural dynamics Chapter 1 69

If the number of degrees of freedom is n < 3N , dx’s are not independent. It


is possible to find n independent generalized coordinates by considering trans-
formation equations, that is,
x1 ¼ x1 ðq1 , q2 , …, qn Þ
x2 ¼ x2 ðq1 , q2 , …, qn Þ
(1.8.37)
… …
x3N ¼ x3N ðq1 , q2 , …, qn Þ
Then we have
X
n
∂xi
dxi ¼ dqj (1.8.38)
j¼1
∂qj

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

and interchanging the order of summation, Eq. (1.8.39) becomes


X
n
∂V
dW ¼  dqj ¼ 0 (1.8.41)
j¼1
∂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)
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

we can write Eq. (1.8.45) in the form


1X n X n
V¼ kij qi qj (1.8.47)
2 i¼1 j¼1

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
General concepts and principles of structural dynamics Chapter 1 71

1.8.3.2 Kinetic energy and mass matrix


The kinetic energy is
 2
1X3N
∂xk
T¼ mk (1.8.51)
2 k¼1 ∂t

Differentiating Eq. (1.8.37) with respect to time yields

∂xk X n
∂xk
¼ q_ (1.8.52)
∂t j¼1
∂qj j

Consequently, it is about the position of the static equilibrium


 2 Xn X n    
∂xk ∂xk ∂xk
¼ q_ i q_ j (1.8.53)
∂t 0 i¼1 j¼1
∂qi 0 ∂qj 0

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

where it was set

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.
72 PART I Single-degree-of-freedom systems

The equations of motion are obtained by applying Eq. (1.8.11). Differenti-


ating the kinetic energy, Eq. (1.8.54), and the potential energy, Eq. (1.8.47),
yields
∂T
¼0
∂qi
  X n
d ∂T
¼ mij q€j
dt ∂q_ i j¼1

∂V X n
¼ kij qj
∂qi j¼1

Substituting into Lagrange’s equations, we obtain the following equations of


motion
X
n X
n
mij q€j + kij qj ¼ Qi ði ¼ 1, 2, …, n Þ (1.8.58)
j¼1 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.

1.8.4 Raleigh’s dissipation function


The dissipative forces arising in a mechanical system are nonconservative
forces. Therefore, they are not derivable from a potential function. They are
involved in the Lagrange equations with their virtual work. When the dissipa-
tive forces are due to such sources as air resistance or internal friction, they are
usually assumed to depend linearly on the velocities along with the physical
coordinates and opposed to the motion, that is
X
n
fDj ¼  cij q_ j (1.8.60)
i¼1

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
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

If these forces are introduced as generalized forces in Lagrange’s equations


(1.8.11) and shifted to the left side, we obtain
 
d ∂T ∂T ∂V ∂R
 + + ¼ Qi ði ¼ 1, 2, …, n Þ (1.8.63)
dt ∂q_ i ∂qi ∂qi ∂q_ i
The function R defined by Eq. (1.8.61) is known as Raleigh’s dissipation
function.
The equations of motion, Eq. (1.8.58), become now
X
n X
n X
n
mij q€j + cij q_ j + kij qj ¼ pi ði ¼ 1, 2, …, n Þ (1.8.64)
j¼1 j¼1 j¼1

where pi denotes the nonconservative external forces.


Eq. (1.8.64) is written in matrix form
q + cq_ + kq ¼ p
m€ (1.8.65)
The matrix c is called the damping matrix of the system.

1.9 Influence of the gravity loads


We consider the system of Fig. 1.9.1a, which can move in the vertical direction.
Apparently, the weight of the body must be added to the external forces. If the
vertical displacement from the undeformed position is designated by u ¼ u ðt Þ,
the equation of motion will read
m u€ + cu_ + ku ¼ pðt Þ + W (1.9.1)

(a) (b)
FIG. 1.9.1 Influence of the gravity load.
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).

FIG. P1.1 Structure in problem P1.1


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

FIG. P1.3 Structural model in problem P1.3

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.
76 PART I Single-degree-of-freedom systems

FIG. P1.4 System in problem P1.4

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.

FIG. P1.5 Structure in problem P1.5

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.

FIG. P1.6 Two-story frame in problem P1.6


General concepts and principles of structural dynamics Chapter 1 77

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.

FIG. P1.7 System in problem P1.7

10 10

FIG. P1.8 Frame in problem P1.8

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.
78 PART I Single-degree-of-freedom systems

FIG. P1.9 Two-story frame in problem P1.9

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.

FIG. P1.10 Soft pendulum in problem P1.10

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

FIG. P1.11 Structure in problem P1.11


General concepts and principles of structural dynamics Chapter 1 79

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.
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

FIG. P1.14 One-story building in problem P1.14

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
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 P1.17 Derive the equation of motion of the system shown in


Fig. P1.17. The mass of the case is m2 . Use m2 ¼ 5m1 , k2 ¼ 3k1 , c2 ¼ 1:5c1 .

FIG. P1.17 System in problem P1.17

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.
82 PART I Single-degree-of-freedom systems

FIG. P1.18 Crane in Problem P1.18.

References and further reading


[1] S. Timoshenko, D.H. Young, W. Weaver Jr., Vibration Problems in Engineering, fifth ed., John
Wiley, New York, 1990.
[2] R.L. Coelho, On the deduction of Newton’s second law, Acta Mech. 229 (5) (2018) 2287–2290.
[3] Euler, L. (1912). Decouverte d’un Nouveau Principe de Mecanique, Memoires de l’academie
des sciences de Berlin 6, 185–217, in: Opera Omnia, Series II, vol. 5, 81–108, Leipzig.
[4] J.T. Katsikadelis, Derivation of Newton’s Law of Motion from Kepler’s Laws of Planetary
Motion, Arch. Appl. Mech. 88 (2018) 27–38, https://doi.org/10.1007/s00419-017-1245-x.
[5] D.T. Greenwood, Principles of Dynamics, Prentice-Hall, Englewood Cliffs, NJ, 1965.
[6] J.T. Katsikadelis, The Boundary Element Method for Engineers and Scientists, Academic Press,
Elsevier, Oxford, UK, 2016.
[7] S. Natsiavas, E. Paraskevopoulos, A set of ordinary differential equations of motion for con-
strained mechanical systems, Nonlinear Dyn. 79 (2015) 1911–1938.
[8] E.N. Strømmen, Structural Dynamics, Springer Series in Solid and Structural Mechanics,
Springer, New York, 2014.

You might also like