Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

CHAPTER

8
GENERALIZED
SINGLE-
DEGREE-
OF-FREEDOM
SYSTEMS

8-1 GENERAL COMMENTS ON SDOF SYSTEMS


In formulating the SDOF equations of motion and response analysis procedures
in the preceding chapters, it has been tacitly assumed that the structure under consid-
eration has a single lumped mass that is constrained so that it can move only in a single
fixed direction. In this case it is obvious that the system has only a single degree of
freedom and that the response may be expressed in terms of this single displacement
quantity.
However, the analysis of most real systems requires the use of more complicated
idealizations, even when they can be included in the generalized single-degree-of-
freedom category. In this chapter we will discuss these generalized SDOF systems,
and in formulating their equations of motion it is convenient to divide them into two
categories: (1) assemblages of rigid bodies in which elastic deformations are limited
to localized weightless spring elements and (2) systems having distributed flexibility
in which the deformations can be continuous throughout the structure, or within some
of its components. In both categories, the structure is forced to behave like a SDOF

133
134 DYNAMICS OF STRUCTURES

system by the fact that displacements of only a single form or shape are permitted,
and the assumed single degree of freedom expresses the amplitude of this permissible
displacement configuration.
For structures in the category of rigid-body assemblages, discussed in Section
8-2, the limitation to a single displacement shape is a consequence of the assemblage
configuration; i.e., the rigid bodies are constrained by supports and hinges arranged
so that only one form of displacement is possible. The essential step in the analysis
of such assemblages is the evaluation of the generalized elastic, damping, and inertial
forces in terms of this single form of motion.
In the case of structures having distributed elasticity, considered in Section 8-3,
the SDOF shape restriction is merely an assumption because the distributed elasticity
actually permits an infinite variety of displacement patterns to occur. However, when
the system motion is limited to a single form of deformation, it has only a single degree
of freedom in a mathematical sense. Therefore, when the generalized mass, damping,
and stiffness properties associated with this degree of freedom have been evaluated,
the structure may be analyzed in exactly the same way as a true SDOF system.
From these comments it should be evident that the material on analysis of SDOF
systems, presented in the preceding chapters, is equally applicable to generalized
SDOF systems even though it was presented with reference to simple systems having
only a single lumped mass.

8-2 GENERALIZED PROPERTIES: ASSEMBLAGES OF RIGID


BODIES
In formulating the equations of motion of a rigid-body assemblage, the elastic
forces developed during the SDOF displacements can be expressed easily in terms of
the displacement amplitude because each elastic element is a discrete spring subjected
to a specified deformation. Similarly the damping forces can be expressed in terms of
the specified velocities of the attachment points of the discrete dampers. On the other
hand, the mass of the rigid bodies need not be localized, and distributed inertial forces
generally will result from the assumed accelerations. However, for the purposes of
dynamic analysis, it usually is most effective to treat the rigid-body inertial forces as
though the mass and the mass moment of inertia were concentrated at the center of
mass. The inertial-force resultants which are obtained thereby are entirely equivalent
to the distributed inertial forces insofar as the assemblage behavior is concerned.
Similarly it is desirable to represent any distributed external loads acting on the rigid
bodies by their force resultants. The total mass m and the centroidal mass moment of
inertia j of a uniform rod and of uniform plates of various shapes are summarized in
Fig. 8-1 for convenient reference.
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 135

a
 
a
2 2

L

2
j=m 
L
2

12
b

2
m
j=m
( )
a 2+ b 2
12

m = mL m = ab
L

b

2 mass
2

m = 
length

mass
= 
area

m
b

2
m j=m
( )
a 2+ b 2
16
2b

3 j=m
( )
a 2+ b 2
18
m= ab

4
ab b
m=  
2 2
b

3 Ellipse

a 2a a a
   
3 3 2 2

FIGURE 8-1
Rigid-body mass and centroidal mass moment of inertia for uniform rod and uniform
plates of unit thickness.

Example E8-1. A representative example of a rigid-body assemblage,


shown in Fig. E8-1, consists of two rigid bars connected by a hinge at E and
supported by a pivot at A and a roller at H. Dynamic excitation is provided by a
transverse load p(x, t) varying linearly along the length of bar AB. In addition,
a constant axial force N acts through the system, and the motion is constrained
by discrete springs and dampers located as shown along the lengths of the bars.
The mass is distributed uniformly through bar AB, and the weightless bar BC

x  x
a f (t)
p(x, t) = p 
Hinge Weightless, rigid bar EH
m2 , j2
A H
N
B D E G

c1 m k1 c2 k2

a 2a a a a a

FIGURE E8-1
Example of a rigid-body-assemblage SDOF system.
136 DYNAMICS OF STRUCTURES

supports a lumped mass m2 having a centroidal mass moment of inertia j2 .


Because the two bars are assumed rigid, this system has only a single
degree of freedom, and its dynamic response can be expressed with a single
equation of motion. This equation could be formulated by direct equilibration
(the reader may find this a worthwhile exercise), but because of the complexity of
the system, it is more convenient to use a work or energy formulation. A virtual-
work analysis will be employed here; although using Hamilton’s principle, as
described in Chapter 16, would be equally effective.
For the form of displacement which may take place in this SDOF structure
(Fig. E8-2), the hinge motion Z(t) may be taken as the basic quantity and all
other displacements expressed in terms of it; for example, BB 0 (t) = Z(t)/4,
DD0 (t) = 3 Z(t)/4, F F 0 (t) = 2 Z(t)/3, etc. The force components acting on
the system (exclusive of the axial applied force N, which will be discussed later)
are also shown in this figure. Each resisting force component can be expressed
in terms of Z(t) or its time derivatives, as follows:
1 1
fI1 (t) = m1 Z̈(t) = m L Z̈(t) = 2 a m Z̈(t)
2 2
1 mL L2 4
Mj1 (t) = j1 Z̈(t) = Z̈(t) = a2 m Z̈(t)
4a 4a 12 3
2
fI2 (t) = m2 Z̈(t)
3
1
Mj2 (t) = −j2 Z̈(t)
3a
hd i 1
fD1 (t) = c1 DD0 (t) = c1 Ż(t)
dt 4
fD2 (t) = c2 Ż(t)
h i 3
fS1 (t) = k1 DD0 (t) = k1 Z(t)
4
h i 1
fS2 (t) = k2 GG0 (t) = k2 Z(t)
3

The externally applied lateral load resultant is

p1 (t) = 8 p a f (t)

In these expressions, m and p denote reference values of mass and force, re-
spectively, per unit length and f (t) is a dimensionless time-dependent function
which represents the dynamic load variation.
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 137


8a p1(t) = 8 pa f (t)

3 E′
D′ F′
B′ G′ Z(t)
Mj Mj
A 1 2 H
B C D E F G
fD (t)
fS (t) 2 fI (t)
fI (t) 1 2 fS (t)
fD (t) 1 2
1

FIGURE E8-2
SDOF displacements and resultant forces.

The equation of motion of this system may be established by equating to


zero all work done by these force components during an arbitrary virtual dis-
placement δZ. The virtual displacements through which the force components
move are proportional to Z(t), as indicated in Fig. E8-2. Thus the total virtual
work may be written

δZ 4 δZ 2Z̈(t) 2
δW (t) = −2a m Z̈(t) − a2 m Z̈(t) − m2 δZ
2 3 4a 3 3
Z̈(t) Ż(t) δZ 3 3
−j2 δZ − c1 − c2 Ż(t) δZ − k1 Z(t) δZ
3a 4 4 4 4
Z(t) δZ 2
− k2 + 8p a f (t) δZ = 0 (a)
3 3 3

which when simplified becomes


"   
am 4 j2 c1
am + + m2 + 2 Z̈(t) + + c2 Ż(t)
3 9 9a 16
  #
9 k2 16
+ k1 + Z(t) − p a f (t) δZ = 0 (b)
16 9 3

Because the virtual displacement δZ is arbitrary, the term in square brackets


must vanish; thus the final equation of motion becomes
   
4 4 j2 c1
m a + m2 + 2 Z̈(t) + + c2 Ż(t)
3 9 9a 16
 
9 k2 16
+ k1 + Z(t) = p a f (t) (c)
16 9 3
This may be written in the simplified form

m∗ Z̈(t) + c∗ Ż(t) + k ∗ Z(t) = p∗ (t) (8-1)


138 DYNAMICS OF STRUCTURES

if the new symbols are defined as follows:

4 4 j2 1
m∗ = m a + m2 + 2 c∗ = c1 + c 2
3 9 9a 16
9 1 16
k∗ = k1 + k2 p∗ (t) = p a f (t)
16 9 3

These quantities are termed, respectively, the generalized mass, generalized


damping, generalized stiffness, and generalized load for this system; they have
been evaluated with reference to the generalized coordinate Z(t), which has
been used here to define the displacements of the system.
Consider now the externally applied axial force N of Fig. E8-1. As may
be seen in Fig. E8-3, the virtual work done by this force during the virtual
displacement δZ is Nδe. The displacement δe is made up of two parts, δe1 and
δe2 , associated with the rotations of the two bars. Considering the influence of
bar AE only,it is clear from similar triangles (assuming
 small deflections) that
δe1 = (Z(t) 4a) δZ. Similarly δe2 = (Z(t) 3a) δZ, thus the total displace-
ment is
7 Z(t)
δe = δe1 + δe2 = δZ
12 a
and the virtual work done by the axial force N is

7 N Z(t)
δWP = δZ (d)
12 a

Adding Eq. (d) and Eq. (a) and carrying out simplifying operations similar
to those which led to Eq. (c) shows that only one term in the equation of motion

δ e1
E″
δ e1
δZ E″
δZ
E′
E′
Z(t)
A H′ N
H″
δe
4a 3a

FIGURE E8-3
Displacement components in the direction of axial force.
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 139

is influenced by the axial force, the generalized stiffness. When the effect of the
axial force in this system is included, the combined generalized stiffness k ∗ is
7 P 9 1 7 N
k∗ = k∗ − = k1 + k2 − (e)
12 a 16 9 12 a
With this modified generalized-stiffness term, the equation of motion of the
complete system of Fig. E8-1, including axial force, is given by an equation
similar to Eq. (8-1). The last term in Eq. (e), which is directly proportional to
the axial force N, often is given the name “geometric stiffness.”
It is of interest to note that the condition of zero generalized stiffness
represents a neutral stability or critical buckling condition in the system. The
value of axial force Ncr which would cause buckling of this structure can be
found by equating k ∗ of Eq. (e) to zero:
9 1 7 Ncr
0= k1 + k2 −
16 9 12 a
Thus  27
4 
Ncr =k2 a k1 + (f)
28 21
In general, compressive axial forces tend to reduce the stiffness of a structural
system, while tensile axial forces cause a corresponding increase of stiffness.
Such loads can have a significant effect on the response of the structure to
dynamic loads, and the resulting change of stiffness should always be evaluated
to determine its importance in the given problem. It should be noted that axial
force in this and in subsequent discussions refers to a force which acts parallel
to the initial undistorted axis of the member; such a force is assumed not to
change the direction of its line of action or its magnitude with the motion of the
structure.

Example E8-2. As a second example of the formulation of the equations


of motion for a rigid-body assemblage, the system shown in Fig. E8-4 will be
considered. The small-amplitude motion of this system can be characterized by

mass
γ = 
area
(uniform)
fS (t) k a

fI (t)
1 I (t)
b
b fI (t)
 2
2
Z(t)
a

2 p(t)

FIGURE E8-4
SDOF plate with dynamic forces.
140 DYNAMICS OF STRUCTURES

the downward displacement of the load point Z(t), and all the system forces
resisting this motion can be expressed in terms of it:

b 1
fS (t) = k Z(t) fI1 (t) = γ a b Z̈(t)
a 2
b a2 + b 2 1
fI2 (t) = γ a b Z̈(t) MI (t) = γ a b Z̈(t)
2a 12 a

The equation of motion for this simple system can be written directly by ex-
pressing the equilibrium of moments about the plate hinge:

a b
fS (t) b + fI1 (t) + fI2 (t) + MI (t) = p(t) a
2 2

Dividing by the length a and substituting the above expressions for the forces,
this equation becomes

"   #
1 b2 1 b2 b2
γ ab 2
+ 1 + + 2 Z̈(t) + k 2 Z(t) = p(t)
12 a 4 4a a

Finally, it may be written

m∗ Z̈(t) + k ∗ Z(t) = p∗ (t)

in which

 
∗ γ ab b2 b2
m = 1+ 2 k∗ = k p∗ (t) = p(t)
3 a a2

8-3 GENERALIZED PROPERTIES: DISTRIBUTED


FLEXIBILITY
The example of Fig. E8-1 is a true SDOF system in spite of the complex
interrelationships of its various components because the two rigid bars are supported
so that only one type of displacement pattern is possible. If the bars could deform in
flexure, the system would have an infinite number of degrees of freedom. A simple
SDOF analysis could still be made, however, if it were assumed that only a single
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 141

flexural deflection pattern could be developed.


As an illustration of this method of approximating SDOF behavior in a flexure
system actually having infinite degrees of freedom, consider the formulation of an
equation of motion for the cantilever tower of Fig. 8-2a. The essential properties of
the tower (excluding damping) are its flexural stiffness EI(x) and its mass per unit of
length m(x). It is assumed to be subjected to horizontal earthquake ground-motion
excitation vg (t), and it supports a constant vertical load N applied at the top.
To approximate the motion of this system with a single degree of freedom, it is
necessary to assume that it will deform only in a single shape. The shape function
will be designated ψ(x), and the amplitude of the motion relative to the moving base
will be represented by the generalized coordinate Z(t); thus,

v(x, t) = ψ(x)Z(t) (8-2)

Typically the generalized coordinate is selected as the displacement of some convenient


reference point in the system, such as the tip displacement in this tower. Then the
shape function is the dimensionless ratio of the local displacement to this reference
displacement:
v(x, t)
ψ(x) = (8-3)
Z(t)

The equation of motion of this generalized SDOF system can be formulated


conveniently only by work or energy principles, and the principle of virtual work will
be used in this case.

N
Z(t)
e(t)

v t (x,t)

v(x,t) pe ff (x,t) = − m(x) v̈g (t)


Fixed reference axis

L
m(x)
EI(x)
x

(b)
vg (t) FIGURE 8-2
Flexure structure treated as a SDOF
(a) system.
142 DYNAMICS OF STRUCTURES

Since the structure in this example is flexible in flexure, internal virtual work δWI
is performed by the real internal moments M (x, t) acting through their corresponding
 2 v(x) 
virtual changes in curvature δ ∂ ∂x 2 . The virtual-work principle requires that the
external virtual work, δWE (t), performed by the external loadings acting through their
corresponding virtual displacements be equated to the internal virtual work, i.e.,

δWE = δWI (8-4)

To develop the equation of motion in terms of relative displacement v(x, t), the
base of the structure can be treated as fixed while an effective loading peff(x, t) is
applied as shown in Fig. 8-2b. The inertial loading is then given by

fI (x, t) = m(x) v̈(x, t) (8-5)

Using the full set of external forces, the external virtual work is given by
Z L Z L
δWE = − fI (x) δv(x) dx + peff(x, t) δv(x) dx + N δe (8-6)
0 0

and consistent with the above statement regarding internal virtual work,
Z L
δWI (t) = M (x, t) δv 00 (x) dx (8-7)
0

where v 00 (x) = ∂ 2 v(x)/∂x2 .


If it is assumed that damping stresses are developed in proportion to the strain
velocity, a uniaxial stress-strain relation of the form

σ = E [ + a1 ]
˙ (8-8)

may be adopted, where E is Young’s modulus and a1 is a damping constant. Then the
Euler-Bernouli hypothesis that plane sections remain plane leads to the relation

M(x, t) = EI(x) [v 00 (x, t) + a1 v̇ 00 (x, t)] (8-9)

Using this equation, the basic relations may be expressed as follows:

v(x, t) = ψ(x) Z(t) v̇ 00 (x, t) = ψ 00 (x) Ż(t)

v 0 (x, t) = ψ 0 (x) Z(t) δv(x, t) = ψ(x) δZ


(8-10)
v 00 (x, t) = ψ 00 (x) Z(t) δv 0 (x, t) = ψ 0 (x) δZ

v̈(x, t) = ψ(x) Z̈(t) δv 00 (x, t) = ψ 00 (x) δZ


GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 143

Also, by analogy with the development of Eq. (d) of Example E8-1, the expressions
for axial displacement take the form
Z L Z L
1 0 2
e(t) = [v (x, t)] dx δe = v 0 (x, t) δv 0 (x) dx (8-11)
2 0 0

Finally, expressions for the external and internal virtual work may be formulated by
using Eqs. (8-10) and (8-11):

h Z L
δWE = −Z̈(t) m(x) ψ(x)2 dx
0
Z L Z L i
− v̈g (t) m(x) ψ(x) dx + N Z(t) ψ 0 (x)2 dx δZ (8-12)
0 0
h Z L Z L i
δWI = Z(t) EI(x) ψ 00 (x)2 dx + a1 Ż(t) EI(x) ψ 00 (x)2 dx δZ
0 0

Equating Eqs. (8-12) in accordance with Eq. (8-4) yields the generalized equation of
motion
m∗ Z̈(t) + c∗ Ż(t) + k ∗ Z(t) − kG

Z(t) = p∗eff (t) (8-13)

where
Z L

m = m(x) ψ(x)2 dx = generalized mass
0
Z L

c = a1 EI(x) ψ 00 (x)2 dx = generalized damping
0
Z L

k = EI(x) ψ 00 (x)2 dx = generalized flexural stiffness (8-14)
0
Z L

kG =N ψ 0 (x)2 dx = generalized geometric stiffness
0
Z L

peff (t) = −v̈g (t) m(x) ψ(x) dx = generalized effective load
0

Combining the two stiffness terms, Eq. (8-13) can be written as

m∗ Z̈(t) + c∗ Ż(t) + k ∗ Z(t) = p∗eff (t) (8-15)

in which
k ∗ = k ∗ − kG

(8-16)

is the combined generalized stiffness.


144 DYNAMICS OF STRUCTURES

The critical buckling load can be calculated for this system by the same method
used in Example E8-1, i.e., by equating to zero the combined generalized stiffness and
solving for Ncr ; thus, one obtains
RL
EI(x) ψ 00 (x)2 dx
Ncr = 0 R L (8-17)
0
ψ 0 (x)2 dx
This SDOF approximate analysis of the critical buckling load is called Rayleigh’s
method, which is discussed in the context of vibration analysis in Section 8-5. The
value determined for the critical load depends, of course, upon the assumed shape
function ψ(x), but a very good approximation will be given by any shape that is
consistent with the geometric boundary conditions.

Example E8-3. To provide a numerical example of the formulation of


the equation of motion for a SDOF system with distributed flexibility, it will
be assumed that the tower of Fig. 8-2 has constant flexural stiffness EI and
constant mass distribution m along its length and damping in accordance with
Eq. (8-8). Also, its deflected shape in free vibrations will be assumed as
πx
ψ(x) = 1 − cos (a)
2L
which satisfies the geometric boundary conditions ψ(0) = ψ 0 (0) = 0. When
Eqs. (8-14) are applied, one obtains
Z L
πx 2
m∗ = m 1 − cos dx = 0.228 m L
0 2L
Z L 2
∗ π πx 2 a1 π 4 EI
c = a1 EI cos dx =
0 4L2 2L 32 L3
Z L 2
π πx 2 π 4 EI
k ∗ = EI 2
cos dx = (b)
0 4L 2L 32 L3
Z L
π πx 2 Nπ 2
kG∗
=N sin dx =
0 2L 2L 8L
Z L
πx 
peff(t) = −m v̈g (t) 1 − cos dx = 0.364 m L v̈g (t)
0 2L
which upon substitution into Eq. (8-13) gives the SDOF equation of motion:
   a π 4 EI   π 4 EI Nπ 2 
1
0.228mL Z̈(t) + Ż(t) + − Z(t)
32 L3 32 L3 8L
= −0.364 m L v̈g (t) (c)
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 145

In addition, the buckling load for this column subjected to tip load will be
evaluated by setting the combined stiffness equal to zero and solving for Ncr ,
with the following result:
π 2 EI
Ncr = (d)
4 L2
This is the true buckling load for an end-loaded uniform cantilever column
because the assumed shape function of Eq. (a) is the true buckled shape.
Of course, one could select a different shape function ψ(x) as long as it
satisfies the geometric boundary conditions ψ(0) = ψ 0 (0) = 0. For example, if
this function were assumed to be of the parabolic form
x2
ψ(x) = (e)
L2
the equation of motion obtained by the above procedure would be
   4 a EI   4EI 4N  mL
1
0.200 m L Z̈(t) + Ż(t) + − Z(t) = − v̈g (t)
L3 L3 3L 3
(f)
Setting the combined stiffness equal to zero, the critical load is given as
3EI
Ncr = (g)
L2
which is about 22 percent higher than the true value given by Eq. (d).

When using the Rayleigh method of buckling analysis as given by Eq. (8-17), it
should be recognized that assuming any shape other than the true buckled shape will
require additional external constraints acting on the system to maintain its equilibrium.
These additional external constraints represent a stiffening influence on the system;
therefore the critical load computed by a Rayleigh analysis using any shape other than
the true one must always be greater than the true critical load. In the above example,
it is apparent that the parabolic shape is not a good assumption for this structure, even
though it satisfies the geometric boundary conditions, because the constant curvature
of this shape implies that the moment is constant along its length. It is obvious here
that the moment must vanish at the top of the column, and any assumed shape that
satisfies this force boundary condition (i.e., one having zero curvature at the top) will
give much better results.

8-4 EXPRESSIONS FOR GENERALIZED SYSTEM


PROPERTIES
As implied by the preceding examples, the equation of motion for any SDOF
system, no matter how complex, can always be reduced to the form

m∗ Z̈(t) + c∗ Ż(t) + k ∗ Z(t) = p∗ (t)


146 DYNAMICS OF STRUCTURES

in which Z(t) is the single generalized coordinate expressing the motion of the system
and the symbols with asterisks represent generalized physical properties corresponding
to this coordinate. In general, the values of these properties can be determined
by application of either the principle of virtual work, as illustrated by the previous
examples, or Hamilton’s principle as illustrated in Chapter 16. However, standardized
forms of these expressions can be derived easily which are very useful in practice.
Consider an arbitrary one-dimensional system, as illustrated by the example in
Fig. 8-3, assumed to displace only in a single shape ψ(x) with displacements expressed

v(x,t) = ψ (x) Z(t) v1 v2


(a)
−v3
−v4
Z (t)
x
L

m(x) m1, j1 m4 , j4
(b)

x1

c2 c3
c(x) a1(x)
(c)

k1 EI (x) k2
k (x)
(d )

q (x)

(e) N

p1 (t) p3 (t)
p(x,t)

(f)

FIGURE 8-3
Properties of generalized SDOF system: (a) assumed shape; (b) mass properties; (c) damping
properties; (d ) elastic properties; (e) applied axial loading; ( f ) applied lateral loading.
GENERALIZED SINGLE-DEGREE-OF-FREEDOM SYSTEMS 147

in terms of the generalized coordinate Z(t) as given by

v(x, t) = ψ(x) Z(t)

Part of the total mass of the system is distributed in accordance with m(x) and the
remainder is lumped at discrete locations i (i = 1, 2, . . .) as denoted by mi . External
damping is provided by distributed dashpots varying in accordance with c(x) and by
discrete dashpots as denoted by the ci values, and internal damping is assumed to
be present in flexure as controlled by the uniaxial stress-strain relation of Eq. (8-8).
The elastic properties of the system result from distributed external springs varying
in accordance with k(x), from discrete springs as denoted by the ki values, and from
distributed flexural stiffness given by EI(x). External loadings are applied to the
system in both discrete and distributed forms as indicated by the time-independent
axial forces q(x) and N and the time-dependent lateral forces p(x, t) and pi (t). These
loadings produce internal axial force and moment distributions N(x) and M (x, t),
respectively.
Applying the procedure of virtual work to this general SDOF system in the same
manner as it was applied to the previous example solutions, one obtains the following
useful expressions for the contributions to the generalized properties:
Z L X X
∗ 2
m = m(x) ψ(x)2 dx + mi ψi2 + ji ψi0
0
Z L Z L X
2
c∗ = c(x) ψ(x)2 dx + a1 EI(x) ψ 00 (x) dx + ci ψi2
0 0
Z L Z L X
2
k∗ = k(x) ψ(x)2 dx + EI(x) ψ 00 (x) dx + ki ψi2 (8-18)
0 0
Z L
2
− N(x) ψ 0 (x) dx
0
Z L X
p∗ (t) = p(x, t) ψ(x) dx + pi (t) ψi (x)
0

The vectorial nature of the force and displacement quantities in the last of Eqs. (8-
18) must be carefully noted. Only components of the forces in the directions of
the corresponding assumed displacements can be included, and the positive sense of
each force component must be assigned in accordance with the positive sense of the
corresponding displacement.
The above generalized-coordinate concepts apply equally in the reduction of
two-dimensional systems to a single degree of freedom. Consider, for example, the
rectangular floor slab shown in Fig. 8-4 subjected to a distributed downward loading
148 DYNAMICS OF STRUCTURES

Z (t) w (x,y,t)
b

FIGURE 8-4
Simply supported two-dimensional slab treated as a SDOF system.

p(x, y, t). If the deflections of this slab are assumed to have the shape ψ(x, y) shown,
and if the displacement amplitude at the middle is taken as the generalized coordinate,
the displacements may be expressed
w(x, y, t) = ψ(x, y) Z(t) (8-19)
For a uniform simply supported slab, the shape function might logically be of the form
πx πy
ψ(x, y) = sin sin (8-20)
a b
but any other reasonable shape consistent with the support conditions could be used.
The generalized properties of this system can be calculated by expressions
equivalent to those presented in Eqs. (8-18) for the one-dimensional case; however,
the integrations must be carried out here in both the x and y directions. For this
specific example, the generalized mass, stiffness, and loading would be given by
Z aZ b

m = m(x, y) ψ(x, y)2 dx dy
0 0
Z Z b (h 2
a
∂ ψ(x, y) ∂ 2 ψ(x, y) i2
k∗ = D +
0 0 ∂x2 ∂y 2
)
h ∂ 2 ψ(x, y) ∂ 2 ψ(x, y)  ∂ 2 ψ(x, y) 2 i
− 2 (1 − ν) − dx dy
∂x2 ∂y 2 ∂x ∂y
Z a Z b
p∗ (t) = p(x, y) ψ(x, y) dx dy
0 0
where 
D = Eh3 12 (1 − ν 2 ) = flexural rigidity of the slab
ν = Poisson’s ratio
h = plate thickness

You might also like