You are on page 1of 32

17

Systems with Distributed Mass


and Elasticity

PREVIEW

So far in this book we have focused on discretized systems, typically with lumped masses;
such a system is an assemblage of rigid elements having mass (e.g., the floor diaphragms of
a multistory building) and massless elements that are flexible (e.g., the beams and columns
of a building). A major part of this book is devoted to lumped-mass discretized systems,
for two reasons. First, such systems can effectively idealize many classes of structures,
especially multistory buildings. Second, effective methods that are ideal for computer im-
plementation are available to solve the system of ordinary differential equations governing
the motion of such systems. However, a lumped-mass idealization, although applicable,
is not a natural approach for certain types of structures, such as a bridge (Fig. 2.1.2e),
a chimney (Fig. 2.1.2f), an arch dam (Fig. 1.10.2), or a nuclear containment structure
(Fig. 1.10.1).
In this chapter we formulate the structural dynamics problem for one-dimensional
systems with distributed mass, such as a beam or a tower, and solutions are presented
for simple systems (e.g., a uniform beam and a uniform tower). The solutions presented
for these simple cases provide insight into the dynamics of distributed-mass systems that
have an infinite number of DOFs and how they differ from lumped-mass systems with a
finite number of DOFs. The chapter ends with a discussion of why this infinite-DOF ap-
proach is not feasible for practical systems, pointing to the need for discretized methods for
distributed-mass systems. The results presented for the simple systems provide the exact
solution against which results from discretized methods can be compared (Chapter 18).

697
698 Systems with Distributed Mass and Elasticity Chap. 17

17.1 EQUATION OF UNDAMPED MOTION: APPLIED FORCES

In this section we develop the equation governing the transverse vibration of a straight
beam without damping subjected to external force. Figure 17.1.1a shows such a beam
with flexural rigidity E I (x) and mass m(x) per unit length, both of which may vary with
position x. The external forces p(x, t), which may vary with position and time, cause
motion of the beam described by the transverse displacement u(x, t) (Fig. 17.1.1b). The
equation of motion to be developed will be valid for support conditions other than the
simple supports shown and for beams with intermediate supports.

p(x,t) p dx

V ∂M
m(x), EI(x) M M+ dx
∂x
x
L
• • ∂V
V+ dx
(a) ∂x
∂2u
u u(x,t) fI = m dx
∂t2

x dx

(b) (c)

Figure 17.1.1 System with distributed mass and elasticity: (a) beam and applied force;
(b) displacement; (c) forces on element.

The system has an infinite number of DOFs because its mass is distributed. There-
fore, we consider a differential element of the beam, isolated by two adjoining sections.
The forces on the element are shown in Fig. 17.1.1c, where an inertia force has been in-
cluded following D’Alembert’s principle (Section 1.5.2); V(x, t) is the transverse shear
force and M(x, t) is the bending moment. Equilibrium of forces in the y-direction gives
∂V ∂ 2u
= p−m 2 (17.1.1)
∂x ∂t
Without the inertia force this equation is the familiar relation between the shear force in
a beam and external transverse force. The inertia force modifies the external force in
recognition of the dynamics of the problem. If the inertial moment associated with angular
acceleration of the element is neglected, rotational equilibrium of the element gives the
standard relation
∂M
V= (17.1.2)
∂x
We now use Eqs. (17.1.1) and (17.1.2) to write the equation governing the trans-
verse displacement u(x, t). With shear deformation neglected, the moment–curvature
Sec. 17.2 Equation of Undamped Motion: Support Excitation 699

relation is
∂ 2u
M = EI (17.1.3)
∂x2
Substituting Eqs. (17.1.3) and (17.1.2) into Eq. (17.1.1) gives
 
∂ 2u ∂2 ∂ 2u
m(x) 2 + 2 E I (x) 2 = p(x, t) (17.1.4)
∂t ∂x ∂x
This is the partial differential equation governing the motion u(x, t) of the beam subjected
to external dynamic forces p(x, t). To obtain a unique solution to this equation, we must
specify two boundary conditions at each end of the beam and the initial displacement
u(x, 0) and initial velocity u̇(x, 0).

17.2 EQUATION OF UNDAMPED MOTION: SUPPORT EXCITATION

Consider two simple cases: a cantilever beam subjected to horizontal base motion (Fig.
17.2.1a) or a beam with multiple supports subjected to identical motion in the vertical
direction (Fig. 17.2.1b). The total displacement of the beam is
u t (x, t) = u g (t) + u(x, t) (17.2.1)
where the beam displacement u(x, t), measured relative to the support motion u g (t), results
from the deformations of the beam.

u(x,t)
u(x,t)
L ug(t)
x

x

(b)
ug(t)

(a)

Figure 17.2.1 (a) Cantilever beam subjected to base excitation; (b) continuous beam
subjected to identical motion at all supports.

For these simple cases of a beam excited by support motion the derivation of the
equation of motion is only slightly different than that for applied forces. Recognizing that
the inertia forces are now related to the total accelerations and that external forces p(x, t)
do not exist, Eq. (17.1.1) becomes
700 Systems with Distributed Mass and Elasticity Chap. 17

∂V ∂ 2ut ∂ 2u d 2u g
= −m 2 = −m 2 − m 2 (17.2.2)
∂x ∂t ∂t dt
wherein Eq. (17.2.1) has been used to obtain the second half of the equation. Substituting
Eqs. (17.1.3) and (17.1.2) into Eq. (17.2.2) gives
 
∂ 2u ∂2 ∂ 2u
m(x) 2 + 2 E I (x) 2 = −m(x)ü g (t) (17.2.3)
∂t ∂x ∂x
By comparing Eqs. (17.2.3) and (17.1.4), it is clear that the deformation response u(x, t)
of the beam to support acceleration ü g (t) will be identical to the response of the system
with stationary supports due to external forces = −m(x)ü g (t). The support excitation can
therefore be replaced by effective forces (Fig. 17.2.2):
peff (x, t) = −m(x)ü g (t) (17.2.4)

üg üg üg

m(x)
x

=
= peff(x,t)
x = −m(x)üg(t) peff(x,t) = −m(x)üg(t)

üg(t)
Stationary base
Stationary supports

(a) (b)

Figure 17.2.2 Effective forces peff (x, t).

This formulation can be generalized to include the possibility of different motions of


the various supports of a structure. Such multiple support excitation may exist in several
practical situations (Section 9.7), but is not included in this chapter because it is usually
not possible to analyze such practical problems as infinite-DOF systems. They are usually
discretized by the finite element method (Chapter 18) and analyzed by extensions of the
procedures of Section 13.5.

17.3 NATURAL VIBRATION FREQUENCIES AND MODES

For the case of free vibration, Eqs. (17.1.4) and (17.2.3) become
 
∂ 2u ∂2 ∂ 2u
m(x) 2 + 2 E I (x) 2 = 0 (17.3.1)
∂t ∂x ∂x
Sec. 17.3 Natural Vibration Frequencies and Modes 701

We attempt a solution of the form


u(x, t) = φ(x)q(t) (17.3.2)
Then
∂ 2u ∂ 2u
= φ(x)q̈(t) = φ (x)q(t) (17.3.3)
∂t 2 ∂x2
where overdots denote a time derivative and primes denote an x derivative; thus q̇(t) ≡
dq/dt, q̈(t) = d 2 q/dt 2 , and φ (x) = d 2 φ/d x 2 . Substituting Eq. (17.3.3) in Eq. (17.3.1)
leads to
 
m(x)φ(x)q̈(t) + q(t) E I (x)φ (x) = 0
which, when divided by m(x)φ(x)q(t), becomes
−q̈(t) [E I (x)φ (x)]
= (17.3.4)
q(t) m(x)φ(x)
The expression on the left is a function of t only and the one on the right depends only on
x. For Eq. (17.3.4) to be valid for all values of x and t, the two expressions must therefore
be constant, say ω2 . Thus the partial differential equation (17.3.1) becomes two ordinary
differential equations, one governing the time function q(t) and the other governing the
spatial function φ(x):
q̈ + ω2 q = 0 (17.3.5)
 
E I (x)φ (x) − ω2 m(x)φ(x) = 0 (17.3.6)
Equation (17.3.5) has the same form as the equation governing free vibration of an SDF
system with natural frequency ω. For any given stiffness and mass functions, E I (x) and
m(x), respectively, there is an infinite set of frequencies ω and associated modes φ(x)
that satisfy the eigenvalue problem defined by Eq. (17.3.6) and the support (boundary)
conditions for the beam.
For the special case of a uniform beam, E I (x) = E I and m(x) = m, and Eq. (17.3.6)
becomes
E I φIV (x) − ω2 mφ(x) = 0 or φIV (x) − β 4 φ(x) = 0 (17.3.7)
where
ω2 m
β4 = (17.3.8)
EI
The general solution of Eq. (17.3.7) is (see Derivation 17.1)
φ(x) = C1 sin βx + C2 cos βx + C3 sinh βx + C4 cosh βx (17.3.9)
This solution contains four unknown constants, C1 , C2 , C3 , and C4 , and the eigenvalue pa-
rameter β. Application of the four boundary conditions for a single-span beam, two at each
end of the beam, will provide a solution for β and hence for the natural frequency ω [from
Eq. (17.3.8)] and for three constants in terms of the fourth, resulting in the natural mode of
702 Systems with Distributed Mass and Elasticity Chap. 17

Eq. (17.3.9). This procedure is illustrated next by two examples: a simply supported beam
and a cantilever beam. Results are also available for other boundary conditions but are not
included in this book.

17.3.1 Uniform Simply Supported Beam

The natural frequencies and modes of vibration of a uniform beam simply supported at
both ends are determined next. At x = 0 and x = L, the displacement and bending
moment are zero. Thus, using Eqs. (17.3.2), (17.1.3), and (17.3.9) at x = 0 gives

u(0, t) = 0 ⇒ φ(0) = 0 ⇒ C2 + C4 = 0 (17.3.10a)


M(0, t) = 0 ⇒ E I φ (0) = 0 ⇒ β 2 (−C2 + C4 ) = 0 (17.3.10b)

These two equations give C2 = C4 = 0 and the general solution reduces to

φ(x) = C1 sin βx + C3 sinh βx (17.3.11)

Then at x = L,

u(L , t) = 0 ⇒ φ(L) = 0 ⇒ C1 sin β L + C3 sinh β L = 0 (17.3.12a)


M(L , t) = 0 ⇒ E I φ (L) = 0 ⇒ β 2 (−C1 sin β L + C3 sinh β L) = 0 (17.3.12b)

Adding these two equations after dropping β 2 from the second equation gives

C3 sinh β L = 0

Since sinh β L cannot be zero (otherwise, ω will be zero, a trivial solution implying no
vibration at all), so C3 must be zero. This leads to the frequency equation:

C1 sin β L = 0 (17.3.13)

This equation can be satisfied by selecting C1 = 0, which gives φ(x) = 0, a trivial solution.
Therefore, sin β L must be zero, from which

β L = nπ n = 1, 2, 3, . . . (17.3.14)

Equation (17.3.8) then gives the natural vibration frequencies:



n2π 2 E I
ωn = n = 1, 2, 3, . . . (17.3.15)
L2 m
The natural vibration mode corresponding to ωn is obtained by substituting Eq. (17.3.14)
in Eq. (17.3.11) with C3 = 0 as determined earlier:
nπ x
φn (x) = C1 sin (17.3.16)
L
Sec. 17.3 Natural Vibration Frequencies and Modes 703

m, EI
x
L
• •
φ1(x) = sin(πx/L)
π2 EI
ω1 =
L2 m

φ2(x) = sin(2πx/L)
4π2 EI
ω2 =
L2 m

φ3(x) = sin(3πx/L)
9π2 EI
ω3 =
L2 m
• Figure 17.3.1 Natural vibration modes and
• frequencies of uniform simply supported
• beams.

The value of C1 is arbitrary; C1 = 1 will make the maximum value of φn (x) equal to unity.
These natural modes are shown in Fig. 17.3.1.
For a simply supported uniform beam, we have determined an infinite series of
modes each with its vibration frequency. Equations (17.3.15) and (17.3.16) and Fig. 17.3.1
tell us that the first mode is a half sine wave and that its frequency ω1 = π 2 (E I /m L 4 )1/2 .
The second mode is a complete sine wave with frequency ω2 = 4ω1 ; the third is one and a
half sine waves with frequency ω3 = 9w1 ; and so on.

17.3.2 Uniform Cantilever Beam

In this section the natural vibration frequencies and modes of a uniform cantilever beam
are determined. At the clamped end, x = 0, the displacement and slope are zero. Thus
Eq. (17.3.9) gives

u(0, t) = 0 ⇒ φ(0) = 0 ⇒ C2 + C4 = 0 ⇒ C4 = −C2 (17.3.17a)


 
u (0, t) = 0 ⇒ φ (0) = 0 ⇒ β(C1 + C3 ) = 0 ⇒ C3 = −C1 (17.3.17b)

At the free end, x = L, of the cantilever the bending moment and shear are both zero.
Thus, from Eqs. (17.3.9) and after using Eq. (17.3.17), we obtain
704 Systems with Distributed Mass and Elasticity Chap. 17

M(L , t) = 0 ⇒ E I φ (L) = 0


⇒ C1 (sin β L + sinh β L) + C2 (cos β L + cosh β L) = 0 (17.3.18a)

V(L , t) = 0 ⇒ E I φ (L) = 0
⇒ C1 (cos β L + cosh β L) + C2 (− sin β L + sinh β L) = 0 (17.3.18b)
Rewriting Eqs. (17.3.18a) and (17.3.18b) in matrix form yields
    
sin β L + sinh β L cosβ L + cosh β L C1 0
= (17.3.19)
cos β L + cosh β L − sin β L + sinh β L C2 0
Equation (17.3.19) can be satisfied by selecting both C1 and C2 equal to zero, but this
would give a trivial solution of no vibration at all. For either or both of C1 and C2 to be
nonzero, the coefficient matrix in Eq. (17.3.19) must be singular (i.e., its determinant must
be zero). This leads to the frequency equation:
1 + cos β L cosh β L = 0 (17.3.20)
No simple solution is available for β L, so Eq. (17.3.20) is solved numerically to obtain
βn L = 1.8751, 4.6941, 7.8548, and 10.996 (17.3.21)
for n = 1, 2, 3, and 4. For n > 4, βn L  (2n − 1)π/2. Equation (17.3.8) then gives the
first four natural frequencies:
   
3.516 E I 22.03 E I 61.70 E I 120.9 E I
ω1 = ω2 = ω3 = ω4 =
L2 m L2 m L2 m L2 m
(17.3.22)
Corresponding to each value of βn L, the natural vibration mode is
 
cosh βn L + cos βn L
φn (x) = C1 cosh βn x − cos βn x − (sinh βn x − sin βn x)
sinh βn L + sin βn L
(17.3.23)

x

φ3(x) φ4(x)
φ2(x)

L m, EI φ1(x) • • •

φ

3.516 EI 22.03 EI 61.70 EI 120.9 EI


ω1 = ω2 = ω3 = ω4 =
L2 m L2 m L2 m L2 m

Figure 17.3.2 Natural vibration modes and frequencies of uniform cantilever beams.
Sec. 17.3 Natural Vibration Frequencies and Modes 705

where C1 is an arbitrary constant. To arrive at Eq. (17.3.23), C2 is expressed in terms of C1


from Eq. (17.3.18a) and substituted in the general solution, Eq. (17.3.9), and Eq. (17.3.17)
is used. The first four natural vibration modes are shown in Fig. 17.3.2.

Derivation 17.1
The solution of the fourth-order ordinary differential equation (17.3.7) is of the form

φ(x) = Aeax (a)

where A is an arbitrary constant. Substituting for φ(x) and its fourth derivative in Eq. (a)
yields the characteristic equation

a 4 − β 4 = 0 or (a 2 − β 2 )(a 2 + β 2 ) = 0 (b)

which gives a = ±β and a = ±iβ. Thus the general solution of Eq. (17.3.7) is

φ(x) = A1 eiβx + A2 e−iβx + A3 eβx + A4 e−βx (c)

Equation (c) can be rewritten as Eq. (17.3.9) because

e±βx = cosh βx ± sinh βx e±iβx = cos βx ± i sin βx (d)

17.3.3 Shear Deformation and Rotational Inertia

In the preceding derivation of the equation of motion for the transverse vibration of a
beam, the inertial moment associated with rotation of the beam sections was ignored in
Eq. (17.1.2), and only the deflection associated with bending stress in the beam was in-
cluded in Eq. (17.1.3), thus ignoring the deflection due to shear stress in the beam. The
analysis of beam vibration, including both the effects of rotational inertia and shear defor-
mation, is called the Timoshenko beam theory.
The following equation governs such free vibration of a uniform beam with m(x) =
m and E I (x) = E I :
  4
∂ 2u ∂ 4u E ∂ u m 2r 2 ∂ 4 u
m 2 + E I 4 − mr 1 +
2
+ =0 (17.3.24)
∂t ∂x κG ∂ x 2 ∂t 2 κG A ∂t 4

where G is the modulus of rigidity, r = I /A is the radius of gyration of the beam cross
section, A is the area of cross section, and κ is a constant that depends on the cross-sectional
shape and accounts for the nonuniform distribution of shear stress across the section. The
constant κ is derived for various cross-sectional shapes in textbooks on solid mechanics
(e.g., κ is 56 for rectangular cross section and 109
for circular cross section).
Consider a beam with both ends simply supported. Assuming a solution of the form
u(x, t) = C sin(nπ x/L) sin ωn t, which satisfies the necessary end conditions, the fre-
quency equation is obtained. Denoting a natural frequency of the beam by ωn if shear and
706 Systems with Distributed Mass and Elasticity Chap. 17

rotational inertia effects are included, by ωn if these effects are neglected [Eq. (17.3.15)],
and defining n = ωn /ωn , this frequency equation can be written as
nπr
2  E
 nπr
4 E
(1 − 2n ) − 2n 1+ + 4n =0 (17.3.25)
L κG L κG

If it is assumed that nr /L << 1, the (nπr /L)4 term may be dropped and Eq. (17.3.25)
reduces to
1
ωn = ωn (17.3.26)
1 + (nπr /L)2 (1 + E/κG)

which implies that ωn < ωn . The correction due to rotational inertia is represented by
the term (nπr /L)2 in the denominator, whereas the shear deformation correction appears
as (nπr /L)2 (E/κG). Thus the correction term for shear deformation is E/κG times
larger than the rotational inertia correction term. For steel beams of rectangular cross
section, E/κG is approximately 3.2. Values of n = ωn /ωn are plotted in Fig. 17.3.3
using the solution of Eq. (17.3.25), a quadratic equation in 2n , for three values of E/κG;
these results are valid for all natural frequencies because n is included in the abscissa
scale. Similar results are presented in Fig. 17.3.4 for the first five natural frequencies
of a beam with E/κG = 3.2. Also included is the approximate value of the frequency

1.0

0.8

0.6
ω′n / ωn

E/κG = 1
0.4 2
4 3

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0
nr / L

Figure 17.3.3 Influence of shear deformation and rotational inertia on natural frequencies
of simply supported beams.
Sec. 17.4 Modal Orthogonality 707

1.0

n=1
0.8

n=2
0.6
ω′n / ωn

n=3
0.4 n=4
Exact
n=5
Approximate
0.2

0.0
0.0 0.05 0.10
r/L

Figure 17.3.4 Influence of shear deformation and rotational inertia on natural frequencies
of simply supported beams.

given by Eq. (17.3.26). When nr /L < 0.2 the error in the approximate equation is less
than 5%.
Shear deformation and rotational inertia have the effect of lowering the natural fre-
quencies, as shown in Figs. 17.3.3 and 17.3.4. For a fixed value of the slenderness ratio
L/r of the beam, the frequency reduction due to shear deformation and rotational inertia
increases with the E/κG value and with mode number. The latter observation implies that
while the corrections due to shear deformation and rotational inertia may be unimportant
for the fundamental natural frequency, they could be significant for the higher frequencies.
For a fixed value of E/κG and mode number, the frequency reduction increases with r /L,
implying its significance for less slender or stubby beams. From the results presented one
can estimate whether these corrections need to be included in a particular problem. For
earthquake response analysis of many practical structures, these corrections are not signif-
icant, but it is important to realize that these corrections do exist. If significant, they can be
included in the finite element formulation for practical structures which are not amenable
to solution as infinite-DOF systems.

17.4 MODAL ORTHOGONALITY

In this section we derive the orthogonality properties of natural vibration modes of systems
with distributed mass and elasticity. For convenience, the derivation is restricted to single-
span beams with hinged, clamped, or free ends and without any lumped mass at the ends,
although the final result applies in general.
708 Systems with Distributed Mass and Elasticity Chap. 17

The starting point for this derivation is Eq. (17.3.6), which governs the natural fre-
quencies and modes; for mode r ,
 
E I (x)φr (x) = ωr2 m(x)φr (x) (17.4.1)
Multiplying both sides by φn (x) and integrating from 0 to L gives
L L
 

φn (x) E I (x)φr (x) d x = ωr 2
m(x)φn (x)φr (x) d x (17.4.2)
0 0

The left side of this equation is integrated by parts; applying this procedure twice leads to
L L L
 
    
φn (x) E I (x)φr (x) d x = φn (x)[E I (x)φr (x)] − φn (x)[E I (x)φr (x)]
0 0 0
L
+ E I (x)φn (x)φr (x) d x (17.4.3)
0

It is easy to see that the quantities enclosed in {· · ·} are zero at x = 0 and L if the ends of the
beam are free, simply supported, or clamped. This is true at a clamped end because φ = 0
and φ = 0, at a simply supported end because φ = 0 and the bending moment is zero (i.e.,
E I φ = 0), and at a free end because the bending moment is zero (i.e., E I φ = 0) and the
shear force is zero [i.e., (E I φ ) = 0]. With the quantities in {· · ·} set to zero, Eq. (17.4.3)
substituted in Eq. (17.4.2) gives
L L
 
E I (x)φn (x)φr (x) d x = ωr 2
m(x)φn (x)φr (x) d x (17.4.4)
0 0

Similarly, starting with Eq. (17.3.6) written for mode n, multiplying both sides by
φr (x), integrating from 0 to L, and using integration by parts twice leads to
L L
E I (x)φn (x)φr (x) d x = ωn2 m(x)φn (x)φr (x) d x (17.4.5)
0 0

Subtracting Eq. (17.4.4) from Eq. (17.4.5) gives


L
(ωn2 − ωr2 ) m(x)φn (x)φr (x) d x = 0
0

Therefore, if ωn = ωr ,
L
m(x)φn (x)φr (x) d x = 0 (17.4.6a)
0

and this substituted in Eq. (17.4.2) leads to


L
 
φn (x) E I (x)φr (x) d x = 0 (17.4.6b)
0

Equations (17.4.6a) and (17.4.6b) are the orthogonality relations for the natural vibration
Sec. 17.5 Modal Analysis of Forced Dynamic Response 709

modes. If a system has repeated frequencies, modes φn (x) still exist such that any two
modes, n = r , satisfy the orthogonality relations even if ωn = ωr .

17.5 MODAL ANALYSIS OF FORCED DYNAMIC RESPONSE

We now return to the partial differential equation (17.1.4), which is to be solved for a given
applied force p(x, t). Assuming that the associated eigenvalue problem of Eq. (17.3.6) has
been solved for the natural frequencies and modes, the displacement due to each mode is
given by Eq. (17.3.2) and the total displacement by


u(x, t) = φr (x)qr (t) (17.5.1)
r =1

Thus the response u(x, t) has been expressed as the superposition of the contributions of
the individual modes; the r th term in the series of Eq. (17.5.1) is the contribution of the r th
mode to the response.
We will see next that Eq. (17.1.4) can be transformed to an infinite set of ordinary
differential equations, each of which has one modal coordinate qn (t) as the unknown. Sub-
stituting Eq. (17.5.1) in Eq. (17.1.4) gives

 ∞
  
m(x)φr (x)q̈r (t) + E I (x)φr (x) qr (t) = p(x, t)
r =1 r =1

Now we multiply each term by φn (x), integrate it over the length of the beam, and inter-
change the order of integration and summation to get
∞ L ∞
 L
 
q̈r (t) m(x)φn (x)φr (x) d x + qr (t) φn (x) E I (x)φr (x) d x
r =1 0 r =1 0
L
= p(x, t)φn (x) d x
0
By virtue of the orthogonality properties of modes given by Eq. (17.4.6), all terms in each
of the summations on the left side vanish except the one term for which r = n, leaving
L L L
 

q̈n (t) m(x) [φn (x)] d x + qn (t)
2
φn (x) E I (x)φn (x) d x = p(x, t)φn (x) d x
0 0 0
This equation can be rewritten as
Mn q̈n (t) + K n qn (t) = Pn (t) (17.5.2)
where

L L  
Mn = m(x) [φn (x)] d x
2
Kn = φn (x) E I (x)φn (x) d x
0 0
L
Pn (t) = p(x, t)φn (x) d x (17.5.3)
0
710 Systems with Distributed Mass and Elasticity Chap. 17

If each end of the beam is free, hinged, or clamped, Eq. (17.4.3) and subsequent discussion
gives an alternative equation for K n :
L
 2
Kn = E I (x) φn (x) d x (17.5.4)
0

The generalized mass Mn and generalized stiffness K n for the nth mode are related:

K n = ωn2 Mn (17.5.5)

This relation can be derived by writing Eq. (17.4.1) for the nth mode, multiplying both
sides by φn (x), integrating over 0 to L, and utilizing the definitions of Mn and K n . The term
Pn (t) in Eq. (17.5.2) is called the generalized force for the nth mode. Equa-
tion (17.5.2) governs the nth modal coordinate qn (t), and the generalized properties Mn ,
K n , and Pn (t) depend only on the nth mode φn (x).
Thus we have an infinite number of equations like Eq. (17.5.2), one for each mode.
The partial differential equation (17.1.4) in the unknown function u(x, t) has been trans-
formed to an infinite set of ordinary differential equations (17.5.2) in unknowns qn (t).
Recall that the same equations (12.3.3), N in number, were obtained for N -DOF systems.
For applied dynamic forces defined by p(x, t), the motion u(x, t) of the system can
be determined by solving the modal equations for qn (t). The equation for each mode is
independent of the equations for all other modes and can therefore be solved separately.
Furthermore, each modal equation is of the same form as the equation of motion for an SDF
system. Thus the results obtained in Chapters 3 and 4 for the response of SDF systems to
various dynamic forces—harmonic force, impulsive force, etc.—can be adapted to obtain
solutions qn (t) for the modal equations.
Once the qn (t) have been determined, the contribution of the nth mode to the dis-
placement u(x, t) is given by
u n (x, t) = φn (x)qn (t) (17.5.6)
The total displacement is the combination of the contributions of all the modes:

 ∞

u(x, t) = u n (x, t) = φn (x)qn (t) (17.5.7)
n=1 n=1

The bending moment and shear force at any section along the length of the beam are
related to the displacements u(x) as follows:
 
M(x) = E I (x)u  (x) V(x) = E I (x)u  (x) (17.5.8)
These static relationships apply at each instant of time with u(x) replaced by u n (x, t),
which is given by Eq. (17.5.6). Thus the contribution of the nth mode to the internal forces
is given by
 
Mn (x, t) = E I (x)φn (x)qn (t) Vn (x, t) = E I (x)φn (x) qn (t) (17.5.9)
Sec. 17.5 Modal Analysis of Forced Dynamic Response 711

Combining the contributions of all modes gives the total internal forces:

 ∞

M(x, t) = Mn (x, t) = E I (x)φn (x)qn (t) (17.5.10a)
n=1 n=1


 ∞
  
V(x, t) = Vn (x, t) = E I (x)φn (x) qn (t) (17.5.10b)
n=1 n=1

Example 17.1
Derive mathematical expressions for the dynamic response—displacement and bending
moments—of a uniform simply supported beam to a step-function force po at distance ξ from
the left end (Fig. E17.1). Specialize the results for the force applied at midspan.

po
po

m, EI 0 t
ξ
L
• •
(a) (b)

Figure E17.1

Solution
1. Determine the natural vibration frequencies and modes.

n2π 2 EI nπ x
ωn = φn (x) = sin (a)
L2 m L
2. Set up the modal equations. Substituting φn (x) in Eq. (17.5.3a) gives Mn , which is
substituted in Eq. (17.5.5) together with ωn2 to get K n :

mL n4π 4 E I
Mn = Kn = (b)
2 2L 3
Substituting p(x, t) = po δ(x − ξ ), where δ(x − ξ ) is the Dirac delta function centered at ξ ,
in Eq. (17.5.3c) gives
Pn (t) = po φn (ξ ) (c)

Then the nth modal equation is

Mn q̈n (t) + K n qn (t) = po φn (ξ ) (d)


712 Systems with Distributed Mass and Elasticity Chap. 17

3. Solve the modal equations. Equation (4.3.2) describes the response of an SDF system
to a step force. We will adapt this solution to Eq. (d) by changing the notation u(t) to qn (t)
and noting that (u st )o = po φn (ξ )/K n . Thus

po φn (ξ ) 2 po L 3 φn (ξ )
qn (t) = (1 − cos ωn t) = 4 (1 − cos ωn t) (e)
Kn π E I n4
Substituting Eq. (e) in Eq. (17.5.7) and noting that φn (x) is known from Eq. (a), we obtain the
displacement response u(x, t).
4. Specialize for ξ = L/2. Substituting ξ = L/2 in Eq. (e) and the latter in Eq. (17.5.7)
gives

2 po L 3  φn (L/2) nπ x
u(x, t) = (1 − cos ωn t) sin (f)
π4E I n4 L
n=1

where 
  0 n = 2, 4, 6, . . .
L
φn = 1 n = 1, 5, 9, . . . (g)
2
−1 n = 3, 7, 11, . . .
from Eq. (a) and Fig. 17.3.1. Substituting Eq. (g) in Eq. (f) gives

2 po L 3 1 − cos ω1 t πx 1 − cos ω3 t 3π x
u(x, t) = 4 sin − sin
π EI 1 L 81 L

1 − cos ω5 t 5π x 1 − cos ω7 t 7π x
+ sin − sin + ··· (h)
625 L 2401 L

The displacement at midspan is


   
L 2 po L 3 1 − cos ω1 t 1 − cos ω3 t 1 − cos ω5 t 1 − cos ω7 t
u ,t = + + + + ··· (i)
2 π4E I 1 81 625 2401

The coefficients 1, 81, 625, 2401, and so on, in the denominator suggest that the first-mode
contribution is dominant and that the series converges rapidly.
The bending moments are obtained by substituting Eq. (h) in Eq. (17.5.9a):

2 po L 1 − cos ω1 t πx 1 − cos ω3 t 3π x
M(x, t) = − sin − sin
π2 1 L 9 L

1 − cos ω5 t 5π x 1 − cos ω7 t 7π x
+ sin − sin + ··· (j)
25 L 49 L

The bending moment at midspan is


   
L 2 po L 1 − cos ω1 t 1 − cos ω3 t 1 − cos ω5 t 1 − cos ω7 t
M ,t =− + + + + ··· (k)
2 π2 1 9 25 49

This series with n 2 in the denominator converges slowly compared to Eq. (i) with n 4 in the
denominator. This difference implies that higher modes contribute more significantly to forces
Sec. 17.5 Modal Analysis of Forced Dynamic Response 713

than to displacements, a result consistent with the conclusions of Chapters 12 and 13 for
discretized systems.

Example 17.2
A simply supported bridge with a single span of length L has a deck of uniform cross section
with mass m per feet length and flexural rigidity E I . A single wheel load po travels across
the bridge at uniform velocity of v, as shown in Fig. E17.2a. Neglecting damping, determine
an equation for the deflection at midspan as a function of time.

n=2
n=1
n=3
Pn
po po

v
t
x m, EI L/v
L
• • -po
(a) (b)

0 50 100 150 200


Location of load, ft
u(L/2)÷(p0 /2454.7)

0 1 2 2.5
0 Time t, sec
0.5 1 mode
1
10 modes (c)
1.5

Figure E17.2

The properties of a prestressed concrete box girder elevated freeway connector are
L = 200 ft, m = 11 kips/g per foot, I = 700 ft4 , and E = 576,000 kips/ft2 . If v = 55 mph,
determine an equation for the deflection at midspan as a function of time. Also determine the
maximum value of deflection over time.
Solution We assume that the mass of the wheel load is small compared to the bridge mass,
and it can be neglected.
1. Determine the natural vibration frequencies and modes.

n2π 2 EI nπ x
ωn = φn (x) = sin (a)
L2 m L
2. Determine the generalized mass and stiffness. Substituting φn (x) in Eq. (17.5.3a)
gives Mn , which is substituted in Eq. (17.5.5) together with ωn2 to get K n :
mL n4π 4 E I
Mn = Kn = (b)
2 2L 3
714 Systems with Distributed Mass and Elasticity Chap. 17

3. Determine the generalized force. A load po traveling with a velocity v takes time
td = L/v to cross the bridge. At any time t the position is as shown in Fig. E17.2a. Thus the
moving load can be described mathematically as

po δ(x − vt) 0 ≤ t ≤ td
p(x, t) = (c)
0 t ≥ td

where δ(x − vt) is the Dirac delta function centered at x = vt. Substituting Eq. (c) in
Eq. (17.5.3c) gives

L
Pn (t) = 0
po δ(x − vt)φn (x)d x 0 ≤ t ≤ td
0 t ≥ td

po φn (vt) 0 ≤ t ≤ td
=
0 t ≥ td

po sin(nπ t/td ) 0 ≤ t ≤ td
= (d)
0 t ≥ td

This generalized force is shown in Fig. E17.2b; for the nth mode, it consists of n half-cycles
of the sine function.
4. Set up modal equations.

Mn q̈n (t) + K n qn (t) = Pn (t) (e)

where Mn , K n , and Pn (t) are given by Eqs. (b) and (d). Pn (t) represents n half-cycles of
a sine function. To solve these modal equations, we first determine the response of an SDF
system to such an excitation.
5. Response of SDF system to n half-cycles of p(t) = po sin ωt. The equation of motion
is

po sin(nπ t/td ) t ≤ td
m ü + ku = (f)
0 t ≥ td

During t ≤ td , the force is the same as the harmonic force p(t) = po sin ωt considered earlier
with frequency:

nπ nπ v
ω= = (g)
td L

The response is given by Eq. (3.1.6b), which is repeated here for convenience, with (u st )o ≡
po /k:
 
u(t) 1 ω
= sin ωt − sin ωn t t ≤ td (h)
(u st )o 1 − (ω/ωn )2 ωn

After the force ends (i.e., t ≥ td ) the system vibrates freely with its motion described by
Eq. (4.7.3). The displacement u(td ) and velocity u̇(td ) at the end of the excitation are
Sec. 17.5 Modal Analysis of Forced Dynamic Response 715

determined from Eq. (h):


u(td ) −ω/ωn
= sin ωn td (i1)
(u st )o 1 − (ω/ωn )2

u̇(td ) ω  
= (−1)n − cos ωn td (i2)
(u st )o 1 − (ω/ωn ) 2

Substituting Eqs. (i) in Eq. (4.7.3) gives


u(t) ω/ωn  
= (−1)n sin ωn (t − td ) − sin ωn t t ≥ td (j)
(u st )o 1 − (ω/ωn )2
6. Solve modal equations. The solution of Eq. (f) is given by Eqs. (h) and (j). We will
adapt this solution to the modal equations (e) by changing the notation u(t) to qn (t) and noting
that
L po Pno 2 po
td = (u st )o ≡ = =
v k Kn m Lωn2
where ωn and ω are given by Eqs. (a) and (g), respectively. The results are
 
2 po 1 nπ vt nπ v
qn (t) = sin − sin ωn t t ≤ L/v (k)
m L ωn2 − (nπ v/L)2 L ωn L

2 po 1 nπ v  
qn (t) = (−1)n sin ωn (t − L/v) − sin ωn t t ≥ L/v (l)
m L ωn − (nπ v/L) ωn L
2 2

This solution is valid provided that ωn = nπ v/L or Tn = 2L/nv. Note that when specialized
for n = 1, Eqs. (k) and (l) reduce to Eqs. (g) and (h) of Example 8.4.
7. Determine the total response. The displacement response of the beam is given by
Eq. (17.5.7):


u(x, t) = φn (x)qn (t) (m)
n=1

where φn (x) is given by Eq. (a) and qn (t) by Eqs. (k) and (l).
8. Determine deflection at midspan. Substituting x = L/2 in Eq. (m) gives


u(L/2, t) = φn (L/2)qn (t) (n)
n=1

where
  
0 n = 2, 4, 6, . . .
L
φn = 1 n = 1, 5, 9, . . . (o)
2
−1 n = 3, 7, 11, . . .
Equation (o) indicates that the even-numbered modes, the antisymmetric modes, do not con-
tribute to the midspan deflection.
716 Systems with Distributed Mass and Elasticity Chap. 17

9. Numerical results. For the given prestressed concrete structure and vehicle speed:
11
m= = 0.3416 kip-sec2 /ft2
32.2
E I = 576,000 × 700 = 4.032 × 108 kip-ft2

π2 4.032 × 108
ω1 = = 8.477 rad/sec
2002 0.3416
T1 = 0.74 sec

ωn = n 2 ω1

v = 55 mph = 80.67 ft/sec


πv
= 1.267
L
L 200
td = = = 2.479 sec
v 80.67
Because the duration of the excitation td = L/v is greater than Tn /2 for all n, the maximum
response occurs while the moving load is on the bridge span. This phase of the response is
given by Eqs. (k) and (n):

 ∞  
2 po φn (L/2) 1.267n
u(L/2, t) = sin 1.267nt − sin 8.477n 2 t
(0.3416)200 (8.477n 2 )2 − (1.267n)2 8.477n 2
n=1

  
po φn (L/2) 0.1495
= sin 1.267nt − sin 8.477n 2 t (p)
2454.7 n 4 (1 − 0.02234/n 2 ) n
n=1

Equation (p) is valid for 0 ≤ t ≤ 2.479 sec. The values of midspan deflection calculated
from Eq. (p) at many values of t are shown in Fig. E17.2c; the maximum deflection u o =
1.1869 po /2454.7 is attained when the moving load is near midspan. This deflection is only
17% larger than the static deflection (= po L 3 /48E I ) due to a stationary load po at midspan,
implying that the dynamic effects of a moving load are small.
Also shown is the result considering only the contribution of the first vibration mode
[i.e., the n = 1 term in Eq. (p)]. (Recall that this was the result obtained in Example 8.4.) It is
clear that the response contributions of higher modes are negligible.

17.6 EARTHQUAKE RESPONSE HISTORY ANALYSIS

As shown earlier, when the excitation is acceleration ü g (t) of the supports, the equation of
motion for a beam is the same as for applied force p(x, t) except that this force is replaced
by peff (x, t) given by Eq. (17.2.4). Thus the modal analysis procedure of Section 17.5 can
readily be extended to the earthquake problem.
From Eq. (17.2.4) the effective earthquake forces are

peff (x, t) = −m(x)ü g (t) (17.6.1)


Sec. 17.6 Earthquake Response History Analysis 717

The spatial distribution of these forces is defined by m(x). This force distribution can
be expanded as a summation of inertia force distributions sn (x) associated with natural
vibration modes (see Section 12.8):

 ∞

m(x) = sr (x) =
r m(x)φr (x) (17.6.2)
r =1 r =1

Premultiplying both sides by φn (x), integrating over the length of the beam, and utilizing
modal orthogonality with respect to the mass distribution, Eq. (17.4.6a), leads to
L
L nh

n = where Ln =
h
m(x)φn (x) d x (17.6.3)
Mn 0

The contribution of the nth mode to m(x) is


sn (x) =
n m(x)φn (x) (17.6.4)
Observe that these modal expansion equations for a distributed-mass system are similar to
the corresponding equations (13.2.2) and (13.2.4) for lumped-mass systems.
For uniform cantilever towers with mass m per unit length the modal expansion
of Eq. (17.6.2) is as shown in Fig. 17.6.1. The functions sn (x) were evaluated from
Eq. (17.6.4) using Eqs. (17.6.3) and (17.5.3a) and the φn (x) given by Eqs. (17.3.23) and
(17.3.21).

1.566m -0.868m 0.509m -0.364m

= + + + + • • •

-m 0 m -m 0 m -m 0 m -m 0 m -m 0 m
s(x) s1(x) s2(x) s3(x) s4(x)

Figure 17.6.1 Modal expansion of effective earthquake forces for uniform towers.

Returning now to the modal analysis procedure of Section 17.5, p(x, t) in


Eq. (17.5.3c) is replaced by peff (x, t) given by Eq. (17.6.1), to obtain
Pn (t) = −L nh ü g (t) (17.6.5)
Substituting Eq. (17.6.5) in Eq. (17.5.2), dividing by Mn , and using Eqs. (17.5.5) and
(17.6.2a) gives the modal equations of an undamped tower subjected to earthquake excita-
tion:
q̈n + ωn2 qn = −
n ü g (t) (17.6.6)
718 Systems with Distributed Mass and Elasticity Chap. 17

For classically damped systems, Eq. (17.6.6) becomes


q̈n + 2ζn ωn q̇n + ωn2 qn = −
n ü g (t) (17.6.7)
where ζn is the damping ratio for the nth mode. This is the same as Eq. (13.1.7) de-
rived earlier for N -DOF systems, except that L nh and Mn that enter into
n are now given
by Eqs. (17.6.3) and (17.5.3), respectively. As shown in Section 13.1.3, the solution of
Eq. (17.6.7) is
qn (t) =
n Dn (t) (17.6.8)
where Dn (t) is the deformation response of the nth-mode SDF system. This is an SDF
system with vibration properties—natural frequency ωn and damping ratio ζn —of the nth
mode of the distributed-mass system. Thus qn (t) is readily available once the SDF system
response has been determined by the methods of Chapter 6. The contribution of the nth
mode to the earthquake response of the tower can be expressed in terms of Dn (t). Substi-
tuting Eq. (17.6.8) in Eqs. (17.5.6) and (17.5.9) gives the displacements, bending moments,
and shear forces due to the nth mode:
u n (x, t) =
n φn (x)Dn (t) (17.6.9)
 
Mn (x, t) =
n E I (x)φn (x)Dn (t) Vn (x, t) =
n E I (x)φn (x) Dn (t) (17.6.10)
Alternatively, as in Chapter 13 for an N -DOF system, the internal forces can be deter-
mined from the equivalent static forces associated with displacements u n (x, t) computed
from dynamic analysis. To derive an equation for these forces, we introduce a familiar
equation from elementary beam theory relating deflections u(x) to applied forces f (x).
For a uniform beam
E I u IV (x) = f (x) (17.6.11)
where E I is the flexural rigidity and u IV = d 4 u/d x 4 . The more general version of this
equation applicable to nonuniform beams with flexural rigidity E I (x) is
 
E I (x)u  (x) = f (x) (17.6.12)
Replacing u(x) by the time-varying displacements u n (x, t) from Eq. (17.6.9) gives
 
f n (x, t) =
n E I (x)φn (x) Dn (t) (17.6.13)
which, by using Eq. (17.4.1) rewritten for the nth mode, becomes
f n (x, t) = sn (x)An (t) (17.6.14)
where sn (x) is given by Eq. (17.6.4), and An (t), the pseudo-acceleration response of the
nth-mode SDF system, is given by Eq. (13.1.12), which is repeated:
An (t) = ωn2 Dn (t) (17.6.15)
Observe the similarity between Eqs. (17.6.14) and (13.2.7) for a lumped-mass system. At
Sec. 17.6 Earthquake Response History Analysis 719

any time instant the contribution rn (t) of the nth mode to any response quantity r (t)—
deflection, shear force, or bending moment at any location of the beam—is determined by
static analysis of the beam subjected to external forces f n (x, t), and can be expressed as
rn (t) = rnst An (t) (17.6.16)
The modal static response rnst is determined by static analysis of the tower due to external
forces sn (ξ ) (Fig. 17.6.2). As shown in Fig. 17.6.1, these forces due to the fundamental
mode all act in the same direction, but for the second and higher modes they will change
direction as one moves up the tower.

sn(ξ) = Γnm(ξ)φn(ξ)
L
ξ
x
Figure 17.6.2 Static problem to be solved

to determine modal static responses.

The modal static responses are presented in Table 17.6.1 for five response quantities:
the shear V(x) at location x, the bending moment M(x) at location x, the base shear
Vb = V(0), the base moment Mb = M(0), and deflection u(x). The first four equations
come from static analysis of the system in Fig. 17.6.2. The result for u(x) is obtained by
comparing Eqs. (17.6.9) and (17.6.16) and using Eq. (17.6.15). Parts of the equations for
Vbn
st
and Mstbn are obtained by substituting Eq. (17.6.4) for sn (ξ ), using Eq. (17.6.3) for L nh ,
and defining
L

Mn∗ =
n L nh h ∗n = nh L θn = xm(x)φn (x) d x (17.6.17)
Ln 0

TABLE 17.6.1 MODAL STATIC RESPONSES

Response, r Modal Static Response, rnst


L
V(x) Vnst (x) = x
sn (ξ ) dξ
L
M(x) Mstn (x) = x
(ξ − x)sn (ξ ) dξ
L
Vb Vbn
st
= 0
sn (ξ ) dξ =
n L nh = Mn∗
L
Mb Mstbn = 0
ξ sn (ξ ) dξ =
n L θn = h ∗n Mn∗
u(x) u stn (x) = (
n /ωn2 )φn (x)

Observe the similarity between the equations in Table 17.6.1 and those for a lumped-mass
system in Table 13.2.1. The approach symbolized by Eq. (17.6.16) to determine shear and
720 Systems with Distributed Mass and Elasticity Chap. 17

moment is preferable over Eq. (17.6.10) because it avoids computation of the second and
third derivatives of the mode shapes; obviously, both methods will give identical results.
The base shear Vbn (t) and base moment Mbn (t) due to the nth mode are obtained
by specializing Eq. (17.6.16) for Vb and Mb and substituting for Vbn st
and Mstbn from
Table 17.6.1:
Vbn (t) = Mn∗ An (t) Mbn (t) = h ∗n Vbn (t) (17.6.18)
Because Eq. (17.6.18) is identical to Eqs. (13.2.12b) and (13.2.15b) for lumped-mass sys-
tems, following Section 13.2.5, Mn∗ and h ∗n may be interpreted as the effective modal mass
and effective modal height for the nth mode. Observe that Eq. (17.6.17a and b) is iden-
tical to Eq. (13.2.9a) for lumped-mass systems; the definitions of Mn , L nh , and L θn differ,
however, between distributed-mass and lumped-mass systems.
The sum of the effective modal masses over all modes is equal to the total mass of
the tower:
 ∞ L
Mn∗ = m(x) d x (17.6.19)
n=1 0

and the sum of the first moments about the base of the effective modal masses Mn∗ located
at heights h ∗n is equal to the first moment of the distributed mass about the base:
 ∞ L
∗ ∗
h n Mn = xm(x) d x (17.6.20)
n=1 0

These relations can be proven in the same manner as the analogous equations (13.2.14)
and (13.2.17) for a lumped-mass system. In particular, Eq. (17.6.19) can be proven by
integrating Eq. (17.6.2) over the height of the tower and using Eq. (17.6.3b). Similarly,
Eq. (17.6.20) can be derived from the modal expansion of forces xm(x).
The effective modal masses Mn∗ and effective modal heights h ∗n for a uniform can-
tilever tower are shown in Fig. 17.6.3; note that h ∗n are plotted without their algebraic signs.
Mn∗ and h ∗n were determined from Eq. (17.6.17) using the known modes (Fig. 17.3.2). Ob-
serve that the sum of Mn∗ for the first four modes gives 90% of the total mass of the tower.

0.613mL
L

=
0.726L

0.188mL
m(x) = m
0.209L

0.127L

0.090L

0.065mL
0.033mL
•••
u¨ g(t) ¨ug(t)

Mode 1 2 3 4

Figure 17.6.3 Effective modal masses and effective modal heights.


Sec. 17.7 Earthquake Response Spectrum Analysis 721

Combining the response contributions of all the modes the earthquake response of
the system:

 ∞

r (t) = rn (t) = rnst An (t) (17.6.21)
n=1 n=1

where Eq. (17.6.16) has been used for rn (t). This nth-mode contribution to the response
can be determined from the modal static response (Table 17.6.1) and An (t), the pseudo-
acceleration response of the nth-mode SDF system, just as for N -DOF systems (Fig. 13.1.1).

17.7 EARTHQUAKE RESPONSE SPECTRUM ANALYSIS

The peak response of a distributed-mass system, such as a cantilever tower, can be es-
timated from the earthquake response (or design) spectrum by procedures analogous to
those developed in Chapter 13, Part B for lumped-mass systems.
The exact peak value of the nth-mode response rn (t) is

rno = rnst An (17.7.1)

where An ≡ A(Tn , ζn ) is the ordinate of the pseudo-acceleration spectrum corresponding


to natural period Tn and damping ratio ζn . Alternatively, rno may be viewed as the result of
static analysis of the tower subjected to external forces

f no (x) = sn (x)An (17.7.2)

which are the peak values of the equivalent static forces f n (x, t) defined in Eq. (17.6.14).
The peak value ro of the total response r (t) can be estimated by combining the modal
peaks rno according to one of the modal combination rules presented in Section 13.7.2.
Because the natural frequencies of transverse vibration of a beam are well separated, the
SRSS combination rule is satisfactory. Thus
∞ 1/2

ro  2
rno (17.7.3)
n=1

Example 17.3
A reinforced-concrete chimney, 600 ft high, has a uniform hollow circular cross section with
outside diameter 50 ft and wall thickness 2 ft 6 in. (Fig. E17.3a). For purposes of earthquake
analysis, the chimney is assumed clamped at the base and its mass and flexural stiffness are
computed from the gross area of the concrete (neglecting the reinforcing steel). The elastic
modulus for concrete E c = 3600 ksi, and its unit weight is 150 lb/ft3 . Modal damping ratios
are estimated as 5%. Determine the displacements, shear forces, and bending moments due
to an earthquake characterized by the design spectrum of Fig. 6.9.5 scaled to a peak ground
acceleration of 0.25g. Neglect shear deformations and rotational inertia.
722 Systems with Distributed Mass and Elasticity Chap. 17

A4 = 0.601g
A3 = 0.678g
A2 = 0.678g
(a) 0.75
• • (b)


0.50

A/g
50′

A1 = 0.124g
• •
600′
0.25


•• 2′ - 6″
0

0 T4 T3 T2 1 2 3 T1 4
Tn, sec

Figure E17.3a, b

Solution
1. Determine the chimney properties.

[252 − (22.5)2 ]0.15


m=π = 1.738 kip-sec2 /ft2
32.2
π 4 
E I = (3600 × 144) 25 − (22.5)4 = 5.469 × 1010 kip-ft2
4
2. Determine the natural vibration periods and modes. Equation (17.3.22) gives the nat-
ural frequencies of vibration, and the corresponding periods, in seconds, are T1 = 3.626, T2 =
0.5787, T3 = 0.2067, T4 = 0.1055, and so on. The natural modes, given by Eq. (17.3.23)
with βn L defined by Eq. (17.3.21), were evaluated numerically for many values of x and are
shown in Fig. 17.3.2, normalized to unit value at the top.
3. Compute the modal properties. With the mode shapes known, the properties Mn ,
L nh , L θn , Mn∗ , and h ∗n were obtained by numerically evaluating their respective integrals, and
are presented in Table E17.3.
4. Read the design spectrum ordinates. The design spectrum of Fig. 6.9.5 scaled to a
peak acceleration of 0.25g is shown in Fig. E17.3b, wherein the pseudo-acceleration ordinates
corresponding to the first four periods are noted: An /g = 0.124, 0.678, 0.678, and 0.601.
5. Compute the displacements. The peak displacements u no (x) due to the nth mode
are given by Eq. (17.7.1), where the modal static response rnst becomes u st n (x) given in Ta-
ble 17.6.1. Substituting known values of
n , φn (x), ωn2 , and An leads to u no (x), n = 1, 2,
3, and 4, shown in Fig. E17.3c. At each location x these peak modal displacements are com-
bined according to the SRSS rule, Eq. (17.7.3), to obtain an estimate of the total displacements
u o (x), which are also shown. Observe that the total displacements are due primarily to the first
mode.
Sec. 17.7 Earthquake Response Spectrum Analysis 723

25.0 -1.93 0.144 -0.024 25.1

-25 0 25 -25 0 25 -25 0 25 -25 0 25 -25 0 25


u1o u2o u3o u4o Total uo
(c) Lateral displacements, in.
10.86 -32.9 19.29 -12.24

-40 0 40 -40 0 40 -40 0 40 -40 0 40


f1o f2o f3o f4o
(d) Equivalent static forces, kips

-6 0 6 -6 0 6 -6 0 6 -6 0 6 -6 0 6
V1o V2o V3o V4o Total V o
(e) Shears, 103 kips

-1.5 0 1.5 -1.5 0 1.5 -1.5 0 1.5 -1.5 0 1.5 -1.5 0 1.5
M1o M2o M3o M4o Total Mo
(f) Bending moments, 106 kip-ft

Figure E17.3c–f
724 Systems with Distributed Mass and Elasticity Chap. 17

TABLE E17.3 MODAL PROPERTIES

Mode Mn /m L L nh /m L
n L θn /m L 2

1 0.2500 0.3915 1.5660 0.2844


2 0.2500 −0.2170 −0.8679 −0.0454
3 0.2500 0.1272 0.5089 0.0162
4 0.2498 −0.0909 −0.3637 −0.0082

6. Determine the modal expansion of m(x), Eq. (17.6.2). With φn (x) known from
Fig. 17.3.2 and
n from Table E17.3, the functions sn (x) are determined from Eq. (17.6.4).
Actually, these were presented in Fig. 17.6.1.
7. Compute the equivalent static forces for the nth mode. These forces f no (x) are
determined from Eq. (17.7.2) using the sn (x) of Fig. 17.6.1 and the An values of Fig. E17.3b.
The results for the first four modes are shown in Fig. E17.3d.
8. Compute the shears and bending moments. For each mode the peak values of shears
and bending moments at location x are computed by static analysis of the chimney subjected to
forces f no (x). The resulting shears and bending moments due to the first four modes are shown
in Fig. E17.3e and f. At each section x, these modal responses are combined by the SRSS
rule [Eq. (17.7.3)] to obtain an estimate of the total forces, which are also shown. Observe
that the first two modes contribute significantly to the total response, with the second-mode
contribution more significant for the shears than for moments.
9. Compare with Rayleigh’s method. It is of interest to compare the results above
considering response in four modes with the approximate solution using Rayleigh’s method
(Example 8.3). The approximate analysis predicts the displacements reasonably well but not
the bending moments or shears. There are two reasons for the larger errors in forces: (a) The
approximate results differ from the exact response due to the first mode because the assumed
shape function in Rayleigh’s method is an approximation to this mode; this discrepancy in-
troduces larger errors in forces than in displacements. (b) The second and higher modes,
whose response contributions to forces are more significant than they are for displacements,
are neglected in Rayleigh’s method.

17.8 DIFFICULTY IN ANALYZING PRACTICAL SYSTEMS

It is evident that the dynamic response of systems with distributed mass and elasticity can
be determined by the modal analysis procedure once the natural vibration frequencies and
modes of the system have been determined. Both examples solved in Section 17.3 involved
uniform beams, and we found the natural frequencies and modes analytically, although the
frequency equation for the cantilever had to be solved numerically. This classical approach
is rarely feasible if the flexural rigidity E I or mass m vary along the length of the beam,
several intermediate supports are involved, or the system is an assemblage of several mem-
bers with distributed mass. In this section we identify some of the difficulties in obtaining
analytical solutions for the above-mentioned systems.
Consider a single-span beam with mass m(x) and flexural stiffness E I (x). To de-
termine the natural frequencies and modes, we need to solve Eq. (17.3.6), which can be
Sec. 17.8 Difficulty in Analyzing Practical Systems 725

rewritten as
E I (x)φIV (x) + 2E I  (x)φ (x) + E I  (x)φ (x) − ω2 m(x)φ(x) = 0 (17.8.1)
Because the coefficients E I (x), E I  (x), E I  (x), and m(x) of this fourth-order differential
equation vary with x, an analytical solution is rarely feasible for ω2 and φ(x). Therefore,
it is not practical to use the classical approach for practical problems in which E I (x) and
m(x) may be complicated functions.
In finding the natural frequencies and modes of a beam on multiple supports, the
uniform segment between each pair of supports is considered as a separate beam with its
origin at the left end of the segment. Equation (17.3.9) applies to each segment, there is
one such equation for each segment, and the necessary boundary conditions are:

1. At each end of the beam the usual boundary conditions are applicable, depending on
the type of support.
2. At each intermediate support the deflection is zero, and since the beam is continuous,
the slope and the moment just to the left and to the right of the support are the same.

This process quickly becomes unmanageable because of the four constants in Eq. (17.3.9),
which must be evaluated in each segment. An analytical solution is rarely feasible for
ω2 and φ(x), especially if the span lengths vary and m(x) and E I (x) vary within each
segment, as would often be the case for a multispan bridge.
Consider the two-member frame shown in Fig. 17.8.1. Each member is axially rigid
and has uniform properties—flexural rigidity and mass—as indicated; however,
they may differ from one member to the other. Each member is considered as a separate
beam with its origin at one end. Equation (17.3.9) applies to each uniform member, there
is one such equation for each member, and the necessary end and joint conditions are:

1. At the supports of the frame the usual boundary conditions are applicable, depending
on the type of support, resulting in four equations for the frame of Fig. 17.8.1:
φ(1) (L 1 ) = 0 φ(1) (L 1 ) = 0 φ(2) (0) = 0 φ(2) (0) = 0

φ(1)(x1)
L1
• •
x1

EI1, m1

L2 EI2, m2
φ(2)(x2)

x2

Figure 17.8.1
726 Systems with Distributed Mass and Elasticity Chap. 17

2. At the joint the end displacements of the joining members should be compatible; this
condition for axially rigid members gives
φ(1) (0) = 0 φ(2) (L 2 ) = 0
3. At the joint the end slopes of the joining members should be compatible; thus
φ(1) (0) = φ(2) (L 2 )
4. At the joint the bending moments should be in equilibrium; thus
E I1 φ(1) (0) + E I2 φ(2) (L 2 ) = 0

A simple two-member frame requires setting up these eight conditions and the evaluation
of eight constants. The process becomes unmanageable for a frame with many members.
It should now be evident that the classical procedure to determine the natural frequen-
cies and modes of a distributed-mass system with infinite number of DOF, is not feasible
for practical structures. Such problems can be analyzed by discretizing them as systems
with a finite number of DOFs, as discussed in the next chapter.

FURTHER READING

Clough, R. W., and Penzien, J., Dynamics of Structures, McGraw-Hill, New York, 1993, Chapters
17–19.
Humar, J. L., Dynamics of Structures, 2nd ed., A. A. Balkema Publishers, Lisse, The Netherlands,
2002, Chapters 14–17.
Stokey, W. F., “Vibration of Systems Having Distributed Mass and Elasticity,” Chapter 8 in Shock
and Vibration Handbook (ed. C. M. Harris), McGraw-Hill, New York, 1988.
Timoshenko, S., Young, D. H., and Weaver, W., Jr., Vibration Problems in Engineering, Wiley, New
York, 1974.

PROBLEMS

17.1 Find the first three natural vibration frequencies and modes of a uniform beam clamped at
both ends. Sketch the modes. Comment on how these frequencies compare with those of a
simply supported beam.
17.2 Find the first three natural vibration frequencies and modes of a uniform beam clamped at
one end and simply supported at the other. Sketch the modes.
17.3 Find the first five natural vibration frequencies and modes of a uniform beam free at both
ends. Sketch the modes. Comment on how these frequencies compare with those for a beam
clamped at both ends. (Hint: The first two modes are rigid-body modes.)
Chap. 17 Problems 727

17.4 A weight W is suspended from the midspan of a simply supported beam as shown in
Fig. P17.4. If the wire by which the weight is suspended suddenly snaps, describe the
subsequent vibration of the beam. Specialize the general result to obtain the deflection at
midspan. Neglect damping.

L
• •
m, EI

Wire

W Figure P17.4

17.5 Derive mathematical expressions for the displacement response of a simply supported uni-
form beam to the force distribution shown in Fig. P17.5; the time variation of the force is a
step function. Express the displacements u(x, t) in terms of the natural vibration modes of
the beam. Identify the modes that do not contribute to the response. Specialize the general
result to obtain the deflection at midspan. Neglect damping.

p(t)

m, EI
L
• • Figure P17.5

17.6 Derive mathematical expressions for the displacement response of a simply supported uni-
form beam to the force distribution shown in Fig. P17.6; the time variation of the force is a
step function. Express the displacements u(x, t) in terms of the natural vibration modes of
the beam. Identify the modes that do not contribute to the response. Specialize the general
result to obtain the deflection at quarter span. Neglect damping.

p(t)

m, EI
p(t)
L
• • Figure P17.6

17.7 Solve Problem 8.25 considering all natural vibration modes of the bridge.
17.8 Solve Problem 8.26 considering all natural vibration modes of the bridge.
17.9 Prove that Eq. (17.6.19) is valid for a vertical cantilever beam.
17.10 Prove that Eq. (17.6.20) is valid for a vertical cantilever beam.
728 Systems with Distributed Mass and Elasticity Chap. 17

17.11 A free-standing intake–outlet tower 200 ft high has a uniform hollow circular cross section
with outside diameter 25 ft and wall thickness 1 ft 3 in. Assume that the tower is clamped
at the base and that its mass and flexural stiffness are computed from the gross area of the
concrete (neglecting reinforcing steel). The elastic modulus for concrete is 3600 ksi, and
its unit weight is 150 lb/ft3 . Modal damping ratios are estimated as 5%. Determine the top
displacement, base shear, and base overturning moment due to an earthquake characterized
by the design spectrum of Fig. 6.9.5 scaled to the peak ground acceleration of 13 g. Neglect
shear deformations and rotational inertia.

You might also like