Sie sind auf Seite 1von 20

EARTHQUAKE ENGINEERING & STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2013; 42:1637–1656


Published online 11 March 2013 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/eqe.2291

Dynamic analysis of flexible rectangular fluid containers subjected


to horizontal ground motion

S. Hashemi1, M. M. Saadatpour1 and M. R. Kianoush2,*,†


1
Department of Civil Engineering, Isfahan University of Technology, Iran
2
Department of Civil Engineering, Ryerson University, Toronto, Canada

SUMMARY
In this paper, an analytical method is proposed to determine the dynamic response of 3-D rectangular liquid
storage tanks with four flexible walls, subjected to horizontal seismic ground motion. Fluid–structure inter-
action effects on the dynamic responses of partially filled fluid containers, incorporating wall flexibility, are
accounted for in evaluating impulsive pressure. The velocity potential in which boundary conditions are
satisfied is solved by the method of separation of variables using the principle of superposition. The impul-
sive pressure distribution is then computed. Solutions based on 3-D modeling of the rectangular containers
are obtained by applying the Rayleigh–Ritz method using the vibration modes of flexible plates with suitable
boundary conditions. Trigonometrical functions that satisfy boundary conditions of the storage tank such
that the flexibility of the wall is thoroughly considered are used to define the admissible vibration modes.
The analysis is then performed in the time domain. Moreover, an analytical procedure is developed for
deriving a simple formula that evaluates convective pressure and surface displacements in a similar rigid
tank. The variation of dynamic response characteristics with respect to different tank parameters is investi-
gated. A mechanical model, which takes into account the deformability of the tank wall, is developed. The
parameters of such a model can be obtained from developed charts, and the maximum seismic loading can
be predicted by means of a response spectrum characterizing the design earthquake. Accordingly, a simpli-
fied but sufficiently accurate design procedure is developed to improve code formulas for the seismic design
of liquid storage tanks. Copyright © 2013 John Wiley & Sons, Ltd.

Received 26 March 2012; Revised 29 November 2012; Accepted 7 February 2013

KEY WORDS: rectangular tank; fluid–structure interaction; Rayleigh-Ritz; impulsive response; convective
response; seismic design

1. INTRODUCTION

Liquid storage tanks are important components of lifeline and industrial facilities. They also play an
important role in the rescue work after an earthquake. On the basis of observations from previous
earthquakes, it is concluded that liquid storage tanks can be subjected to large hydrodynamic
pressures during earthquakes. Consequently, high stresses can cause buckling failure in the steel
tanks. In concrete tanks, because of the large inertial mass of concrete, the stresses could be large
and result in cracking, leakage, or even collapse of the structure. The poor performance of some of
these structures in past earthquakes has led engineers and researchers to study this problem and to
improve the behaviors of these structures.
There are some numerical and a few analytical methods that have been used for dynamic analysis of
concrete rectangular liquid storage tanks. Hoskins and Jacobsen [1] published the first report on
analytical and experimental observations of rigid rectangular tanks under a simulated horizontal

*Correspondence to: M. R. Kianoush, Department of Civil Engineering, Ryerson University, Toronto, Canada.

E-mail: kianoush@ryerson.ca

Copyright © 2013 John Wiley & Sons, Ltd.


1638 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

earthquake excitation. Housner [2, 3] developed the most commonly used analytical model for
estimating the dynamic response of a rigid rectangular tank. In his model, hydrodynamic pressures
induced by seismic excitation were separated into impulsive and convective components using
lumped mass approximation.
The 1964 Alaska earthquake [4] caused the first large-scale damage to tanks of modern design of its
time and initiated many investigations into the dynamic characteristics of flexible containers. In
addition, evaluations by both digital computers and various associated numerical techniques have
significantly enhanced solution capacity. Several studies were carried out to investigate the dynamic
interaction between the deformable wall in the tank and the liquid and showed that the seismic
response of a flexible tank may be substantially greater than that of a similarly rigid tank.
A 3 DOF model of the ground-supported cylindrical tank was developed by Haroun [4]; the
application of which resulted in design charts used to estimate sloshing, impulsive, and rigid masses.
For evaluating the effective masses, only the fundamental natural mode of vibration of the
deformable tank was considered.
For rectangular tanks, Haroun [5] presented a very detailed method of analysis on the typical system
of loadings. The hydrodynamic pressures were calculated by a classical potential flow approach. The
formula of hydrodynamic pressures only considered the rigid wall condition. This may be due to the
fact that rectangular fluid containers are usually made of reinforced or prestressed concrete and may
be considered quite rigid dynamically. Nevertheless, there are containers of this type for which
flexibility must be taken into account in their dynamic response analysis, such as very large
reinforced concrete structures used for the storage of nuclear spent fuel assemblies or prestressed
concrete water tanks [6].
Some numerical methods that consider wall flexibility have been used for the dynamic analysis of
rectangular liquid storage tanks. Dogangun et al. [7] and Dogangun and Livaoglu [8] investigated
the seismic response of liquid-filled rectangular storage tanks by using the three-dimensional
Lagrangian fluid finite element. Park et al. [9] and Koh et al. [10] studied the seismic response of
rectangular tanks with four flexible walls by using a three-dimensional coupled boundary element–
finite element method. Ghaemmaghami and Kianoush [11] investigated the dynamic behavior of
concrete rectangular tanks by using the FEM in 2-D space, in which the coupled fluid–structure
equations are solved using a direct integral method.
Moreover, there are a few analytical methods that have been used for dynamic analysis of
rectangular liquid storage tanks. Kim et al. [12] studied the dynamic behavior of 3-D rectangular
flexible fluid containers by using the Rayleigh–Ritz method. In their study, only the two walls,
orthogonal to the direction of the applied ground motion, were assumed to be flexible, whereas the
other two walls remained rigid. They considered the vibration modes of a simple beam with hinged
or fixed boundary conditions. Chen and Kianoush [13] proposed a simplified method using the
generalized SDOF system to study the dynamic response of 2-D rectangular tanks. The prescribed
vibration shape functions representing the first five mode shapes for the cantilever beam boundary
condition were studied. In the analytical methods that have so far been used for dynamic analysis of
rectangular liquid storage tanks, neither the effect of sloshing nor the effect of wall flexibility has
been appropriately considered. Moreover, the current design approach is inaccurate, as it does not
fully consider all the major parameters affecting the response.
In this paper, an analytical method is presented to investigate the dynamic response of flexible 3-D
rectangular liquid storage tanks with flexible walls on all four sides, subjected to horizontal seismic
ground motion. Fluid–structure interaction effects on the dynamic responses of partially filled fluid
containers, incorporating wall flexibility, are accounted for in evaluating impulsive pressure. The
velocity potential in which boundary conditions are satisfied is solved by the method of separation
of variables using the superposition principle. On the basis of the achieved solution, the impulsive
pressure distribution is computed. Solutions based on three-dimensional modeling of the rectangular
containers are obtained by applying the Rayleigh–Ritz method using the vibration modes of flexible
plates with suitable boundary conditions. Trigonometrical functions that satisfy boundary conditions
of the storage tank such that the flexibility of the wall is thoroughly considered are used to define
the admissible vibration modes. The analysis is then performed in the time domain. The influence of
tank flexibility on the convective response components is generally small and may be neglected.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1639

Therefore, an analytical procedure is developed for evaluating convective pressure and surface
displacements in a similar rigid tank.
The numerical and analytical methods that have so far been used for dynamic analysis of flexible
rectangular fluid containers are rather complex and cannot be easily used by practicing engineers.
The main objective of the final part of this study is to close the gap between analytical studies and
practical design applications. This is to provide the practicing engineers with a simple and
sufficiently accurate tool for estimating seismic design forces in rectangular tanks.
Seismic loads in most common codes and standards currently used for the design of tanks are based
on the mechanical model developed by Housner [2, 3] for rigid tanks with some modifications. To
develop this model, Housner considered a lamina of fluid of unit thickness, subjecting the walls to a
given horizontal acceleration. He considered the fluid to be constrained by thin, massless vertical
membranes and free to move in the x direction. In the present study, the assumptions underlying the
proposed analytical procedures used to develop a mechanical model equivalent to a rectangular tank
are the basic equations of motion based on the theory of fluid dynamics. Such assumptions are
considered to be more realistic, and the results are expected to be more accurate than those from
previous studies on the basis of equivalent mechanical model. A new mechanical model equivalent
to the rectangular liquid storage tank, which takes into account the deformability of the tank wall, is
developed in this study. The parameters of such a model can be obtained from developed charts.
The maximum seismic loading can be predicted by means of a response spectrum characterizing the
design earthquake. Accordingly, a simplified but sufficiently accurate design procedure is developed
to improve code formulas for the seismic design of liquid storage tanks.

2. FLUID MOTION IN RECTANGULAR TANKS

A rectangular tank with four flexible vertical walls of uniform thickness ts and a horizontal rigid bottom
is partially filled with incompressible and nonviscous liquid of depth HL, as shown in Figure 1. The
side lengths and height of the tank are 2Lx, 2Ly, and Hs, respectively. The walls of the tank are
considered as thin plates made of linearly elastic, homogeneous, and isotropic material and are
assumed to perform transverse bending deflection but no inplane deformation. The motion of the
liquid is assumed to be frictionless and irrotational so that the velocity distribution of the liquid may
be represented as a gradient of the velocity potential:

@Φ @Φ @Φ
υx ¼ ; υy ¼ ; υz ¼ (1)
@x @y @z

Figure 1. 3-D model of rectangular container.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1640 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

According to the theory of fluid dynamics, the liquid velocity potential should satisfy the Laplace equation:

@ 2Φ @ 2Φ @ 2Φ
þ þ ¼0 (2)
@x2 @y2 @z2
And the hydrodynamic pressure at any point and time is given by


pðx; y; z; tÞ ¼ —r l (3)
@t

in which rl is the mass density of the liquid. Considering the walls of the tank to be impermeable and no
cavitation on the liquid–wall interface, the liquid adjacent to the wall must move with it by the same
velocity. When the tank is subjected to an earthquake in the x direction, €u gx , the consistent conditions
between the liquid motion and the tank wall deflection can be given as
. . .
@ Φ .. @ w .. @ Φ .. ¼ @ w ..
x¼TLx ¼ þ _
u ; y ¼ TL @t . y ¼ TLy (4)
@t. x ¼ TLx
gx @y. y
@x.

where w (x, y, z, t) is the relative tank wall deflection. The rigid bottom condition of the tank is
.

.
. ¼0 (5)
@z. z 0
¼

Assuming that the liquid makes a small amplitude oscillation, the linearized sloshing condition at
the free liquid surface (z = HL) is

@ 2Φ @Φ
þg ¼0 (6)
@t 2 @z

3. SOLUTION OF VELOCITY POTENTIAL

The coupling between liquid sloshing modes and wall vibrational modes is weak; consequently, for the
analysis, it is sufficient to consider the two uncoupled systems separately. This includes the liquid–wall
system and the free surface gravity waves in a similar rigid tank [4, 10, 12, 14, 15].
The solution for Ф will be expressed as the sum of the impulsive component, Фi, and the convective
component, Фc. The impulsive component of the solution satisfies the actual boundary conditions
along the tank wall and bottom and the condition of zero hydrodynamic pressure at z = HL , whereas
the convective component corrects for the difference between the actual boundary condition at
z = HL and the one considered in the development of the impulsive solution.
Using the superposition method to solve the velocity potential, one can assume that
y
Φi ¼ Φx þ Φ þ Φgx (7)
f f

where Φxf satisfies the flexible wall conditions in the x direction, the rigid wall conditions in the y
y
direction, and the no-sloshing condition on the free liquid surface. Φ f satisfies the flexible wall
conditions in the y direction, the rigid wall conditions in the x direction, and the no-sloshing
condition on the free liquid surface. Φgx is the hydrodynamic pressure in a similar rigid tank when it
is subjected to an earthquake in the x direction.
For the tank that is subjected to an earthquake in the x direction, using Eqs. (2), (4), and (5), and the
condition of zero hydrodynamic pressure at z = HL , on the basis of the method of separation of

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1641

variables and considering symmetry of both the liquid and the tank about the x–z plane and
antisymmetry of them about the y–z plane (Figure 2), the impulsive velocity potential can be
analytically given in an infinite series form:
1 1
. . . . Z L Z H
4 cos zjz cosðbiyÞ sinh xijx y L
@w
XX · . . . .
Φf ¼
x
xij HL L̄ y cosh xij Lx 0 0 @t ðLx; y; z; tÞ cos zj z cosðbiyÞdzdy (8)
i¼0 j¼0

. . . 0 . . .
X X1 4 cos
1 z z sin b x cosh x 0ijy Z Lx Z HL
y
j
i @w . . . . . .
Φf ¼ 0 · . 0 . x; Ly; z; t cos zjz sin b 0i x dzdx (9)
i¼0 j¼0
x ijHLLx sinh x ijLy 0 0 @t

Φgx ¼ Igx u_ gx ðtÞ (10)

in which

zj ¼ ð 2j þ 1Þp=2HL (11)

0
bi ¼ pi=Ly ; b i ¼ ð2i þ 1Þp=2Lx (12)

xij2 ¼ bi2 þ zj2 ; x0 ij2 ¼ b0 i 2 þ zj2 (13)

.
2Ly i¼0
(14)
L̄ y ¼ y i 6¼ 0
L

X1 4ð—1Þj cosz z sinhx x


j 0j
Igx ¼ (15)
j¼0
ð2j þ 1Þpx 0j coshx 0jLx

(a) (b)

(c)

Figure 2. Relation of the tank wall with the line supported rectangular plate and the relation of the plate
with the beam: (a) deflected shape of the tank wall, (b) equivalent model of the tank wall using symmetry,
and (c) a unit strip equivalent to (b).

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1642 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

The hydrodynamic pressure in a similar rigid tank can be obtained by

@Φgx
pgx ¼ —rl
¼ —rl Igx €u gx ðtÞ (16)
@t

4. ADMISSIBLE FUNCTIONS

It is important to develop a set of suitable admissible functions to describe the vibration of the tank wall
in the Rayleigh–Ritz method. For this purpose, the transverse motion of the flexible walls is expressed
as a linear combination of admissible functions:

X
M XN
wðx; y; z; tÞ ¼ fmnðtÞΨH ΨV (17)
m n
m¼1 n¼1

where coefficients fmn(t) equal the generalized coordinates to be determined; ΨV n equals the nth
vibration mode of a cantilever beam that is appropriate for wall displacement in a vertical
direction; and ΨHm equals the mth eigenfunction of vibrating beams, which boundary conditions are
similar to those of the wall in the horizontal direction. Because of the symmetry of both the liquid
and the tank about the x–z plane and antisymmetry of them about the y–z plane (Figure 2(a)),
only a quarter of the tank with the appropriate boundary conditions is considered. As the tank
wall can only provide transverse deflection but no inplane deformation, the vibration of the tank
wall can be equivalent to the vibration of a line supported rectangular plate whose plane is shown
in Figure 2(c). The connecting line of the adjacent walls corresponds to the internal simply
support condition, which prevents the transverse motion of the plate but offers no resistance to
the rotation.
The general form for eigenfunctions of the vibrating beam of two parts of the beam shown in
Figure 2(c) can be expressed as
.
XmðxÞ y ¼ Ly; 0⩽x⩽Lx
ΨHm ðx; yÞ ¼ YmðyÞ x ¼ Lx; 0⩽y⩽Ly
(18)

Boundary conditions of the beam shown in Figure 2(c) should be satisfied. By equating the
determinants of the homogeneous boundary conditions to zero, eigenvalues km for the mth mode are
determined in two sets of equations. The first set of equations, which governs for most modes of the
tank, is
.0 . . 0 .
— cotðkmÞ þ cothðkmÞ þ tan k m þ tanh k m ¼ 0 (19)

The second set of equations, which may govern for some modes of the tank, which has particular
lengths, is
. 0 .
sinðkmÞ ¼ cos k m ¼ 0 (20)

where km and k0 m are the shape parameters, which are in the relationship:

0
Ly
k m ¼ km (21)
Lx
Eigenfunctions corresponding to the first set of equations (Eq. (19)) are the following:

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1643

. . . .
kmx k mx
Xm ðxÞ ¼ sin
Lx
= sinðk Þ — sinh
m
Lx
= sinhðk Þ
m (22)

. 0 . . . . 0 . . .
k my k my
YmðyÞ ¼ cos
Ly
= 0
cos k m — cosh
Ly
= cosh k 0
m (23)

And eigenfunctions corresponding to the second set of equations (Eq. (20)) are the following:
. .
k mx
XmðxÞ ¼ sin
Lx
cosðkm Þ = (24)

. 0 . . .
k my
YmðyÞ ¼ cos
Ly
= sin k 0
m (25)

For the cantilever beam condition that is appropriate for wall displacement in the vertical direction, the
vibration functions for the nth mode is
. . . . . . . . ..
g z g z g z g z
ΨVn ¼ cosh Hn — cos Hn — sn sinh Hn — sin Hn (26)
L L L L

in which

sn ¼ cosðg Þ þ coshðg Þ (27)


sinðgnÞ þ sinhðg nÞ ;
n n

and the eigenvalues (gn) for the nth mode are determined using Eq. (28):
cosðgnÞ· coshðgnÞ þ 1 ¼ 0 (28)

5. ANALYSIS METHOD

5.1. Impulsive response


Using Eq. (17), one can rewrite the transverse motion of flexible wall as

w ¼ Γf (29)

in which
. .
Γ ¼ ΨH ΨV ΨH ΨV ⋯ΨH ΨV ΨH ΨV ⋯ΨH ΨV ⋯ΨH ΨV ΨH ΨV ⋯ΨH ΨV (30)
1 1 1 2 1 N 2 1 2 N M 1 M 2 M N

fT ¼ ½f11ðtÞ f12ðtÞ⋯ f1N ðtÞ f21ðtÞ ⋯ f2N ðtÞ⋯ fM1ðtÞ fM2ðtÞ⋯ fMN ðtÞ] (31)

The strain energy of the wall is expressed by


1 T
EP ¼ f Kf ⇒ dEP ¼ dfT Kf (32)
2
The kinetic energy of the wall is given by
Z Z . .
rsts Hs Ly @w .2 . 2.
.
. Z Hs Z Lx . . .
EK1 ¼ dy dz þ rsts @w ..y þ Ly dx dz (33)
2 0 0 þ u_ gx . þ 2 0 0 @t ¼
@t x¼ Lx

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1644 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

. Z Z .
dEK1 ¼ d f_ T Ms f_ þ rs ts u_ gx
Hs Ly T
Γ j dydz (34)
0 0 x¼þLx

where rs, Ms, and K are the mass density, mass matrix, and stiffness matrix of the flexible wall that is
modeled as a plate, respectively. The kinetic energy of motion of the liquid in the storage tank can be
written as
1 Z Hl Z Ly Z Lx . 2 2
.
EK2 ¼ rl 2 þ υy þ υz dx dy dz (35)
2 0 υ x
0 0

Using the Green principle and the boundary conditions to Eq. (4), one can obtain
Z
dEK2 ¼ rl Φi dw_ dA (36)
A

. Z .
_ T _
dEK2 ¼ df Mliquid f þ rl Igx u_ gx Γ dA
T
(37)
A

in which
. . . .
1 1Z 4rl cos zj z cosðbi yÞ sinh xij x Z LZ
XX y HL . .
Mliquid ¼ Γ T
. . dA ΓðLx; y; z; tÞ cos zjz cosðbiyÞdzdy
·
xij Hl L̄ y cosh xij Lx 0 0
i¼0 j¼0 A
. . . 0 . . 0 .
1 Z
Z Z HL . . . . . 0 .
X1 X 4rl cos zj z sin b i x cosh x ij y Lx
b x
þ Γ T 0 · . 0 . dA Γ x; L y ; z; t cos zjz sin
i dzdx
i¼0 j¼0
x ij Hl Lx sinh x ij L y 0 0
A
(38)
where A is the area of a quarter of the tank wall shown in Figure 2(b), which is in contact with liquid. A
system of ordinary differential equations can be obtained by application of Hamilton’s principle:
Z t
dðEK — EPÞdt ¼ 0 (39)
0

Integration by parts to these terms results in


.. . .
d f T Ms þ Mliquid €f þ K f þ Ix €
u gx ¼ 0 (40)

where
Z Hs Z Ly Z
I x ¼ r st s T
Γ dxdz þ rl IgxΓT dA (41)
0 0 A

Therefore, a matrix equation, which governs the earthquake response of the undamped liquid–wall
system, is determined:
. .
Ms þ Mliquid €f þ Kf ¼ —Ix €u gx (42)

Mliquid is the added mass matrix because of the effect of the liquid. This matrix equation of motion can
be solved by the mode-superposition method. By equating the determinant of the left hand of Eq. (42)
to zero, frequencies of impulsive modes of tank and eigenvectors pertinent to them are obtained.
K
X
f¼ fk q k (43)
k¼1

fk is the modal vector pertinent to the kth mode of vibration of the tank. K = M × N is the number of
modes. Introducing damping into Eq. (42) and using Eq. (43) result in:

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1645

€q k þ 2zf ok q_ k þ o2 qk ¼ —b €u gx (44)
k k

where zf, ok, and bk are damping ratio, frequency, and modal participation factor of the kth impulsive
mode, respectively. The complete time history of qk(t) for each mode (k = 1,2,..,K) and its time
derivatives can be computed by a step-by-step application (Interpolation of Excitation [16]). Once
they are obtained, the displacement, acceleration, and the impulsive pressure can be calculated using
Eqs. (43), (17), and (3).

5.2. Convective response


The convective pressure can be evaluated with reasonable accuracy by considering the tank wall to be
rigid. Rather than evaluating the convective component of the pressure, pc, from the associated
component of the velocity potential Фc function, it is convenient to evaluate it directly by requiring
that pc satisfy Laplace’s equation. Solution of Laplace’s equation such that it satisfies the boundary
conditions may be expressed as

X
1 . . . .
pcðx; y; z; tÞ ¼ rlLx DjAjðtÞ sin ajx cosh ajz (45)
j¼0

in which
lj ¼ ð2j þ 1Þp=2 (46)

aj ¼ lj=Lx (47)

h . .i
j
Dj ¼ —2ð—1Þ = cosh ljm (48)
l2j

where
m ¼ HL=Lx (49)
and
Z t h i
AjðtÞ ¼ ojc €u gx ðtÞ sin ocjðt — tÞ dt (50)
0

where Aj(t) represents the instantaneous pseudo acceleration of an undamped SDOF system, which has
a circular natural frequency ocj equal to that of the jth sloshing mode of vibration of the liquid and is
excited by a base acceleration €u gx ðtÞ.
. .
ojc 2
¼ gaj tanh ajHl (51)

Damping of convective responses (zc) can be introduced into Eq. (50) easily. The first sloshing
frequency determined using Eq. (51) is equal to its quantity measured using Eurocode-8 2006. The
hydrodynamic pressure at any point and time is given by:

pðx; y; z; tÞ ¼ pf þ pc þ pgx (52)

where pc, pf, and pgx are, respectively, convective pressure, pressure due to flexibility, and pressure on a
similar rigid tank that can be evaluted using Eqs. (45), (53), and (16), respectively.
. .
y
@ Φx þ Φ
f f
pf ¼ —rl (53)
@t

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1646 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

5.3. Surface displacements


The vertical displacement, y(x,y,t), of an arbitrary point at the liquid surface is determined from
pðx; y; tÞjz¼HL ¼ rlg yðx; y; tÞ; (54)

which effectively states that the hydrodynamic pressure at z = HL is equal to the weight of the liquid
column above. The contribution of the inertia of the oscillating liquid is neglected in this expression.
The following relationship can be obtained using Eqs. (45) and (54):
1X j
. .
2ð—1Þ AjðtÞ
yðx; y; tÞ ¼ —Lx sin ajx (55)
j¼0 l2j g

Vertical surface displacement of the liquid near the wall x = —Lx can be determined from
X
1
2Aj ðt Þ
yðLx; y; tÞ ¼ Lx (56)
j¼0 l2j g

when the surface displacement near the wall is maximum, using Eq. (51) and by means of a response
spectrum, one can conclude
AjðtÞ⩽Sac j ¼ 1; 2; ⋯ (57)

in which Sac is the maximum of A0(t) or the spectral acceleration corresponding to the natural
.
j j¼0 c

frequency oc ¼ oc. and damping z . Therefore,

X1
2
y max⩽ S ac L 2 (58)
x
g j¼0 l j

where ymax is the maximum surface displacement near the wall. On making use of identities (Standard
Math Tables [17]):

1
X 2 tanh ðxÞ
lim ¼1 (59)
2 ¼ x!0
lj x
j¼0

Equation (58) results in

y max⩽ ðSac=gÞLx (60)

That is the same as ACI 350.3 [21] formula used to determine the vertical surface displacement.

6. CONVERGENCE AND VALIDITY VERIFICATION

In this section, some numerical examples are presented to investigate the convergence and validity of
the present method. A computer program was first written to determine the validity of the theoretical
formulation in computing seismic responses of the flexible storage tank. The dimensionless natural
frequencies (Ω) of wall vibration are expressed as

. . 1=2
Ω ¼ rso 2tsH4s =D (61)

where D is the flexural rigidity of the tank wall:


Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1647

. . ..
D ¼ E ts3= 12 1 — n2 (62)

The first six dimensionless natural frequencies of wall vibration (impulsive modes) of the empty
cubic tank (2Lx = 2Ly = Hs) and a cubic tank fully filled with liquid, which has a mass ratio of the
liquid to the tank wall equal to 5, are presented in Table I and compared with those presented by
Kim and Dickinson [18] and Zhu and Liu [19]. Five terms of the static beam functions are used in
each direction. Good agreements are observed for all cases.
The present method is now applied to investigate seismic response characteristics of 3-D rectangular
fluid storage tanks. For validity verification, typical dimensions of those used by Koh et al. [10] are
selected; height of the wall, Hs = 10 m; wall thickness, ts = 1 m; water depth, HL = 9 m; length of the
short side wall, 2Lx = 20 m; and length of the long side wall, 2Ly = 50 m, and typical material
properties for the concrete tanks: density, rs = 2400 kg/m3; Young’s modulus, E = 2.1 × 1010 N/m2;
and the Poisson’s ratio, □ = 17. The N-S component of the 1940 El Centro earthquake records is
used as an input motion in the x direction. The tank is assumed to be fixed to the ground and to
have 3% structural damping. The time history of the resultant force acting on the long side wall of
the tank model is presented in Figure 3 and can be compared with those presented by Koh et al.
The comparison shows that they are in good agreement and the maximum of the resultant force
predicted by the present method (9460 kN) is very close to that obtained by Koh et al. (9244 kN).
The accuracy of the present method can be further verified by pressure distribution over the long
side wall when the resultant forces reach their peak values. The predictions by the proposed method
are compared with the results from the coupled 3-D BEM-FEM [10], a Lagrangian fluid finite
element [8], and the results obtained by using the Eurocode-8[8] as shown in Figure 4.
It is worthy here to note that hydrodynamic pressure at the time when the equivalent resultant force
is maximum can be much amplified in the middle length of the wall in rectangular tanks and shows the
distribution of 2-D spatial variation over the surface of the wall as shown in Figure 5.

Table I. Comparison on the first six impulsive frequencies of the empty and filled cubic tank.

Cubic tank
Empty cubic tank filled with liquid
Dimensionless
natural Zhu and Kim and Zhu and
frequency Present study Liu (2007) Dickinson (1987) Present study Liu (2007)

Ω1 17.619 17.632 17.577 8.212 8.48


Ω2 36.049 36.045 36.104 18.011 18.175
Ω3 52.076 52.133 52.069 26.742 26.212
Ω4 71.206 71.208 71.463 38.484 37.494
Ω5 74.377 74.576 74.579 42.441 42.931
Ω6 106.297 107.531 107.13 60.47 58.016

15000
Resultant force, kN

10000
5000
0
5000
10000
15000
0 1 2 3 4 5 6 7 8 9 10
Time, sec

Figure 3. Time history of resultant force acting on the long side wall.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1648 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

Depth (m)
Rigid Tank
6 present work
Koh et al. (1998)
Eurocode 8
Dogangun and Livaglu (2004)
9
0 10 20 30 40
Hydrodynamic Pressure (kPa)

Figure 4. Comparison of pressure distributions along the height of the middle cross section of the long side wall.

Figure 5. Pressure distribution on the long side wall.

0.5
Height, m

0.5 Present Work


Ghaemmaghami (2010)
Koh et al.(1998)
1
0 2 4 6 8 10
Time, sec

Figure 6. Time history of liquid surface elevation at the middle cross section of the long side wall.

Time histories of the sloshing motion at the middle cross section of the long side wall are presented
and compared with those using the indirect BEM-FEM [10] and FEM [20] in Figure 6. The predictions
by the proposed method are virtually identical to those of mentioned references.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1649

7. SEISMIC DESIGN FORCES

The most common standards and codes currently used for the design of tanks are ACI 350.3 [21],
Eurocode-8 [22], and New Zealand standard [23]. Seismic loads in these standards are based on the
mechanical model derived by Housner [2, 3] for rigid tanks with some modifications. The design
procedure considers two response modes of the contained liquid: (1) The impulsive response in
which the portion of the liquid accelerates with the tank walls and (2) the convective response
caused by the portion of the liquid sloshing in the tank. It should be noted that in these codes, the
importance of the effects of wall flexibility has been recognized and the corresponding increase in
the acceleration coefficients has been adopted. However, the effect of wall flexibility has not been
thoroughly accounted for by a reasonable and accurate method.
The main objective of this part of the study is to devise a practical approach, which would allow,
from an engineering point of view, a simple, fast, and sufficiently accurate estimate of the seismic
response of rectangular storage tanks.

7.1. Hydrodynamic base shear


The instantaneous hydrodynamic base shear Q(t) is given by:
Z Ly Z Hl
QðtÞ ¼ 2 pjx¼Lx dzdy (63)
—Ly 0

The hydrodynamic base shear due to wall deformation relative to the ground Qf (t) using Eq. (53)
can be expressed as
.
Z Ly Z Hl . XK
f
Qf ðtÞ ¼ 2 .
. dzdy ¼ Q k€
u kðtÞ (64)
—Ly 0 p f. k¼1
x¼Lx
where

"X . . . .Z L Z H
1 sin zj Hl tanh zjLx y L . .
Q kf ¼ 8rl zj 2H l ΓðLx; y; zÞ cos zjz dzdy
j¼0 0 0
0 . . . 0 . 1#
X1 X 1 sin zj H sin b L Z Z . . . .
Γ .x; Ly; z . cos zj z sin b ix dzdxA fk bk
L i x L H
þ @16rl 0
2H L z
x L 0

i¼0 j¼0 x ij L x j 0 0

(65)
and uk(t) is the solution of the differential equation:

€u kðtÞ þ 2zf ok u_ kðtÞ þ o2 uk kðtÞ ¼ —€u gx ðtÞ (66)

For the applicable tank, all frequencies of impulsive modes are large. Acceleration spectral
spectrums accepted in codes show that the maximum accelerations, which correspond to significant
f
modes, are close to each other. Therefore, the mode that has the maximum Q k is the predominant
mode. The dimensionless fundamental natural frequency (Ωf) that corresponds to the predominant
mode of vibration is displayed for tanks completely filled with water in Figure 7 for two different
values of ts/HL assuming normal density concrete is used. Using Eq. (61), one can determine the
fundamental natural frequency (of) that is pertinent to Ωf for different values of the aspect ratio.
Figure 7 shows that the first mode is predominant when Lx ⩽ 1.2Ly; otherwise, another mode could
become predominant. Therefore, in general, for considering the seismic response, it is better that the
effect of all modes is considered. The value of €u f ðtÞ corresponding to of can be obtained by the
solution of Eq. (66) for the predominant mode. It is assumed that the maximum accelerations
corresponding to significant modes are equal to the maximum of €u f ðtÞ . Therefore, Eq. (64) can be
rewritten approximately:

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1650 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

(a)

(b)

Figure 7. Dimensionless fundamental natural frequency (Ωf) for tanks completely filled with water (a) ts/
HL = 0.05 and (b) ts/HL = 0.1.

Qf ðtÞ ¼ mf €u f ðtÞ (67)

f
where mf is obtained by combining Q k for all modes (k = 1,2,..,K) by the complete quadratic
combination method [24]. The value of mf can be approximated using Eq. (68) obtained by curve
fitting the numerical results shown in Figure 8. The figure shows that mf is independent of
wall thickness.

. . .. . . ..
mf 2Lx 2Lx
¼
tanh 0:866
HL
= 1:732
HL
(68)
ml

where ml is the total mass of liquid. Investigations show that when Ly/HL < 0.65, a correction factor for
mf, cf should be considered:

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1651

0.5

ts/HL=0.05
0.4
ts/HL=0.1
Approximte(Eq.68)
0.3

mf /mL
0.2

0.1

0
0.5 2 3.5 5 6.5 8 9.5
2Lx/HL
(a)
1.4
Exact
1.2 Approximate(Eq.69)
Cf

0.8
0 1 2 3 4 5
Ly/HL
(b)

Figure 8. Values of mf /mL and correction factor: (a) factors mf /mL versus ratio 2Lx/HL and (b) its correction
factor versus ratio Ly/HL.
. .
cf ¼ 1:25-0:71 Ly=HL — 0:3 (69)

Using Eqs. (52), (63), and (67), one may express the base shear as

X
1
QðtÞ ¼ mcjAj ðtÞ þ mf €u f ðtÞ þ mr€u gx ðtÞ (70)
j¼0

in which
. . . .
mr X1 j tanh lj=m sin lj
ð—1Þ
2 (71)
¼ j¼0 lj3=m
ml
. .
mcj 2 tanh ljm
(72)
¼ lj3m
ml
where mr, mf, and mcj
are equivalent masses corresponding to forces associated with ground motion,
wall deformation relative to the ground, and liquid sloshing, respectively.
Figure 9 shows mr obtained using Eq. (71) is in good agreement with that determined by ACI 350.3.
To determine the base shear, one can consider only the first mode of convective response while the
next modes are negligible:

QðtÞ ¼ mc €u c ðtÞ þ mf €u f ðtÞ þ mr€u gx ðtÞ (73)

.
j j¼0 c 0

where mc ¼ m.
c and €u ðtÞ ¼ A ðtÞ is the absolute acceleration of an SDOF system, which has a
.
circular natural frequency oc ¼ ocj .j¼0 . Because the base shear and moment due to wall
deformability are proportional to the relative acceleration of the wall, one can rearrange Eq. (73) to
estimate the maximum seismic loads by means of a response spectrum:
Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1652 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

0.8

Analytical
0.6

mr/mL
ACI

0.4

0.2

0
0.5 1.5 2.5 3.5 4.5 5.5 6.5 7.5
2Lx/HL

Figure 9. Comparison of analytical and ACI solutions for mr/mL versus 2Lx/HL.

. . . .
€ gx ðtÞ þ mr — mf €u gx ðtÞ
QðtÞ ¼ mc €u c ðtÞ þ mf €u f ðtÞ þ u (74)

And subsequently, the maximum base shear can be estimated by


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
.. 2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

2 . . . 2
jQ maxj ¼ ðmcsacÞ þ mf saf þ mr — mf (75)
.. . max
€u gx
in which Sac and Saf are the spectral accelerations corresponding to the natural frequencies oc and of,
respectively.

7.2. Overturning moment


The overturning moment M(t) induced on a section of the tank immediately above its base is given by
Z Ly Z Hl
MðtÞ ¼ 2 pjx¼Lx zdzdy (76)
—Ly 0

The overturning moment due to wall deformation relative to the ground, Mf(t), applied to the bottom
of the wall using Eq. (53) can be expressed as
.
Z Ly Z Hl . XQ
f
Mf ðtÞ ¼ 2 .. zdzdy ¼ M k€
u kðtÞ (77)
—Ly 0 pf
.x ¼Lx k¼1
where:
. . . .Z L Z HL .
" . .
1 . R zj Hl tanh zj Lx y

M kf ¼ X 8rl 3 ΓðL ; y; zÞ cos z z dzdy


zj Hl 0 0
x j
j¼0
0 . . . 0
. # (78)
X 1 X 1 R zj H0L sin b i Lx Z Lx Z HL . . . .
@16rl Γ .x; Ly; z . cos zj z sin b i x dzdx fkbk
0
þ 2 H L z2
i¼0 j¼0 x ij L x j 0 0

in which R is defined as a function:

RðÞ ¼ cosðÞ þ ðÞ sinðÞ — 1 (79)


f
mfhf can be determined by combining M k for all modes (k = 1,2,..,K) using complete quadratic
combination method. The value of hf is obtained by dividing it by mf for different values of the
aspect ratio and is shown in Figure 10. A proper function is fitted with this curve approximately:

hf =HL ¼ 0:58 — 0:12 tanh½2:5ðLx=HL — 0:25Þ] (80)


Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1653

0.7

0.6

0.5

hf / HL
0.4

0.3 Exact
0.2 Approximate

0.1

0
0.25 1 1.75 2.5 3.25
Lx/HL

Figure 10. Comparison of exact and approximate solutions for hf /HL versus Lx/HL.

The results show that the effect of variation of thicknesses or Ly/HL on hf/HL is negligible. Using
Eqs. (52) and (76), one can express M(t) as
X
MðtÞ ¼ 1 mc hc Aj ðtÞ þ mf hf €u f ðtÞ þ mr hr€u gx ðtÞ (81)
j j
j¼0

where hjc , hf, and hr are, respectively, the heights at which the jth convective component of the liquid
mass, mjc , equivalent mass corresponding to force associated with ground motion, mr, and equivalent
mass corresponding to force associated with wall deformation relative to the ground, mf, are
considered to be concentrated:
. . .. . . . .
hr X 1 j
ð—1Þ tanh lj=m R lj X1 j
ð—1Þ tanh lj=m sin lj

HL
¼ j¼0 l j4 j¼0 lj 3 = (82)

. . . .
hc csch l m — cot h l m
j j j
þ1 (83)
¼ l jm
HL
The numerical values of hr are determined from Figure 11 that are compared with those obtained
using ACI 350.3. One may consider only the first mode of convective response to determine the
overturning moment:

MðtÞ ¼ mc hc €u c ðtÞ þ mf hf €u f ðtÞ þ mr hr €u gx ðtÞ (84)


.
j j¼0 c L c L

where hc ¼ h. .
c
Table II shows that h /H and m /m obtained from the present work are in good
agreement with those determined by ACI 350; however, the present formulas are easier. Using
Eqs. (73) and (84), one can develop a mechanical model that is equivalent to the rectangular liquid

0.5

0.4

0.3
hr/HL

0.2 Analytical
ACI
0.1

0
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7 7.5 8
2Lx /HL

Figure 11. Comparison of analytical and ACI solutions for hr/HL versus 2Lx/HL.
Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1654 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

Table II. Comparison of factors hC/HL and mC/mL versus ratio 2Lx/HL for rectangular tanks.
hC/HL mC/mL

2Lx/HL ACI Analytical ACI Analytical

1 0.709 0.708 0.263 0.257


2 0.583 0.582 0.485 0.473
4 0.524 0.524 0.695 0.677
6 0.511 0.511 0.765 0.743
8 0.506 0.506 0.793 0.771

storage tank, which is shown in Figure 12. The maximum overturning moment applied to the bottom of
the wall is given by

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
.2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
.. . . 2ffiffi
2 . ..
jM maxj ¼ ðmchcsacÞ þ mf hf saf þ mr hr — mf hf €u gx max (85)

7.3. Numerical example


Consider the tank discussed in the preceding example is full of water and its impulsive and convective
responses are damped 5% and 0.5%, respectively. The maximum ground acceleration of the N-S
. .
component of the 1940 El Centro earthquake records is €u gx max ¼ 0:313g. The fundamental natural
.
frequency of sloshing, oc ¼ oc .j 0 ¼ 1:19rad= sec, is obtained from Eq. (51), and consequently, the
j ¼
spectral acceleration corresponding to it for a damping ratio of zc = 0.5% can be found from spectral
response (Sac = 0.049g). Table II shows that the convective mass ratio and its elevation are
mc/ml = 0.473 and hc = 0.582HL = 5.82m, respectively. The fundamental impulsive frequency of
liquid–wall vibration for values of Lx/HL = 1, Ly/HL = 2.5, and ts/HL = 0.1 is determined from
Figure 7; Ωf = 2.87 and of = 24.9rad/sec. The spectral acceleration for a damping ratio of zf = 5% is
Saf = 0.799g. The remaining parameters can be obtained from Eqs. (68) and (80), mf/ml = 0.271,
hf = 0.464HL = 4.64m, and from Figures 9 and 11: mr/ml = 0.542 , hr = 0.4HL = 4m. Using Eqs. (60),
(75) and (85), one can determine the earthquake response values and compare these with those from
other studies as shown in Table III. It is worth noting that the ACI 350.3 hydrodynamic pressure
values are calculated assuming the Importance Factor (I) and the Response Modification Coefficient
(R) are set equal to 1. This study shows that the approximations used by the proposed mechanical
model are in good agreement with the exact results determined by the present analytical method.
Moreover, the results of the proposed analytical model are in good agreement with the results of the
studies performed by Moslemi and Kianoush [25], Ghaemmaghami and Kianoush [26], Chen and
Kianoush [27], and Kianoush et al. [28].

Figure 12. Mechanical model of flexible tank.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
DYNAMIC ANALYSIS OF FLEXIBLE RECTANGULAR TANKS 1655

Table III. Comparison on the impulsive earthquake response obtained by different methods.

Analytical method Mechanical


Variable ACI 350.3-06 (proposed) model (proposed)

Maximum surface displacement 0.49 0.456 0.49


near the wall, ymax, (m)
Maximum base shear, Qmax , (N) 42.67 × 106 19.32 × 106 22.99 × 106
Maximum overturning 160.7 × 106 86.13 × 106 103.4 × 106
moment, Mmax, (N, m)

8. CONCLUSIONS

Analytical solution methods are developed that can be used for the analysis of the dynamic behavior of
partially filled rectangular fluid containers under horizontal ground excitations.
In concrete liquid storage tanks, the natural frequencies of significant sloshing modes are much
lower than the fundamental impulsive frequencies. Consequently, the effect of the surface waves
on the impulsive frequencies and the effect of the tank wall flexibility on the sloshing frequencies
can be ignored. In such cases, the studies on the impulsive response of the tank wall vibration and
the sloshing response of the liquid oscillation can be separately considered without significant
errors. Using the combination of the superposition method and the method of separation of
variables, one can derive the exact analytical solution of the impulsive velocity potential. The
coefficients in the impulsive velocity potential are determined by the contact conditions
between the liquid and the tank wall, which are represented by integrals including the tank wall
dynamic deflection.
Solutions based on three-dimensional modeling of the rectangular containers are obtained by
applying the Rayleigh–Ritz method using the vibration modes of flexible plates with suitable
boundary conditions. Trigonometrical functions that satisfy boundary conditions of the storage tank
such that the flexibility of the wall can be thoroughly considered are used to define the admissible
vibration modes. The analysis is then performed in the time domain.
In this study, an analytical procedure is developed to derive a simple formula that estimates the
time–history response of the convective pressure and surface displacements in a similar rigid
tank. Also, through a logical mathematical formulation, it is shown that the maximum vertical
surface displacement of the liquid near the wall is always less than those predicted by the
current practice.
The validity and convergence of the proposed methods are confirmed through numerical
examples. The predictions by the proposed method are in good agreement with the results
from a Lagrangian fluid finite element, FEM, and the results obtained by using the coupled BEM-
FEM.
In the proposed analytical method that is used for dynamic analysis of rectangular liquid storage
tanks, not only the effect of wall flexibility but also the effect of sloshing characteristics is
appropriately considered. Therefore, the predictions by the proposed method are considered to be
more accurate than the results obtained from the methods that are used in current practice.
The results of this study show that hydrodynamic pressure distributions for assuming rigid and
flexible walls differ from each other in magnitude and in shape. The hydrodynamic pressures in the
middle of the wall for flexible storage tanks are generally larger than for rigid storage tanks.
Moreover, hydrodynamic pressure varies not only in the vertical direction but also in the horizontal
direction over the wall surface.
A mechanical model, which takes into account the deformability of the tank wall, is developed, and
its parameters can be obtained from charts.
The maximum seismic response of a deformable rectangular tank can therefore be estimated by
means of a response spectrum. Comparison with the exact solution of the problem confirms the
validity of the method. It is recommended that the effect of wall flexibility on hydrodynamic
pressures should be considered as design rules in design codes and standards.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe
1656 S. HASHEMI, M. M. SAADATPOUR AND M. R. KIANOUSH

REFERENCES
1. Hoskins LM, Jacobsen LS. Water pressure in a tank caused by simulated earthquake. Bulletin of the Seismological
Society of America 1934; 24:1–32.
2. Housner GW. Dynamic pressure on accelerated fluid containers. Bulletin of the Seismological Society of America
1957; 47:5–35.
3. Housner GW. The dynamic behavior of water tanks. Bulletin of the Seismological Society of America 1963; 53:381–387.
4. Haroun MA. Vibration studies and tests of liquid storage tanks. Earthquake Engineering and Structural Dynamics
1983; 11:179–206.
5. Haroun MA. Stress analysis of rectangular walls under seismically induced hydrodynamic loads. Bulletin of the
Seismological Society of America 1984; 74(3):1031–1041.
6. Luft RW. Vertical accelerations in prestressed concrete tanks. Journal of Structural Engineering 1984; 110(4):706–714.
7. Dogangun A, Durmus A, Ayvaz Y. Static and dynamic analysis of rectangular tanks using the lagrangian fluid finite
element. Computers and Structures 1996; 59(3):547–552.
8. Dogangun A, Livaoglu R. Hydrodynamic pressures acting on the walls of rectangular fluid containers. Structural
Engineering and Mechanics 2004; 17(2):203–214.
9. Park JH, Koh HM, Kim J. Fluid-structure interaction analysis by a coupled boundary element-finite element method
in time domain. In Proceedings of the 7th International Conference on Boundary Element Technology, BETECH/92.
Computational Mechanics Publications. Southampton, June, Albuquerque, USA, 1992; 227–243.
10. Koh HM, Kim JK, Park JH. Fluid-structure interaction analysis of 3D rectangular tanks by a variationally coupled
BEM-FEM and comparison with test results. Earthquake Engineering and Structural Dynamics 1998; 27:109–124.
11. Ghaemmaghami AR, Kianoush MR. Effect of wall flexibility on dynamic response of concrete rectangular
liquid storage tanks under horizontal and vertical ground motions. Journal of Structural Engineering, ASCE 2010;
136(4):441–451.
12. Kim JK, Koh HM, Kwahk IJ. Dynamic response of rectangular flexible fluid containers. Journal of Engineering
Mechanics 1996; 122(9):807–817.
13. Chen JZ, Kianoush MR. Generalized SDOF system for seismic analysis of concrete rectangular liquid storage tanks.
Engineering Structures 2009; 31:2426–2435.
14. Veletsos AS, Tang Y. Rocking response of liquid storage tanks. Journal of Engineering Mechanics, ASCE 1987;
113:1774–1792.
15. Haroun MA. Dynamic Analyses of Liquid Storage Tanks. EERL 80-4. California Inst. of Tech.: Pasadena, CA, 1980.
16. Clough RW, and Penzien J. Dynamics of Structures. McGraw-Hill Book: New York, 1997.
17. Beyer WH. Standard Mathematical Tables (24th edn). CRC Press: Cleveland, Ohio, 1976.
18. Kim CS, Dickinson SM. The flexural vibration of line supported rectangular plate system. Journal of Sound and
Vibration 1987; 114:129–142.
19. Zhou D, Liu W. Hydroelastic vibrations of flexible rectangular tanks partially filled with liquid. International Journal
for Numerical Methods in Engineering 2007; 71:149–174.
20. Ghaemmaghami AR. Dynamic time-history response of concrete rectangular liquid storage tanks. PhD thesis, Dept.
of Civil Engineering, Ryerson University, Toronto, Canada, 2010.
21. ACI 350.3, seismic design of liquid-containing concrete structures and commentary (350.3-06). American Concrete
Institute, Farmington Hills, MI, USA, 2006.
22. Eurocode-8, Design of Structures for Earthquake Resistance-Part 4: Silos, Tanks and Pipelines. European Committee
for Standardization, 2006; 65 pp.
23. New Zealand standard. Structural design actions-part 5: earthquake actions-New Zealand-commentary(NZS 1170.5
Supp 1:2004), 2004.
24. Kiureghian AD. Structural response to stationary excitation. Journal of Engineering Mechanics Division, ASCE
1980; 106(6):1195–1213.
25. Moslemi M, Kianoush MR. Parametric study on dynamic behavior of cylindrical ground-supported tanks. Engineering
Structures, Elsevier 2012; 42:214–230.
26. Ghaemmaghami A, Kianoush MR. Effect of earthquake frequency content on seismic behavior of concrete rectangular
liquid tanks using the finite element method incorporating soil-structure interaction effects. Engineering Structures,
Elsevier 2011; 33(7):2186–2200.
27. Chen JZ, Kianoush MR. Seismic response of concrete tanks for liquid containing structures. Canadian Journal of
Civil Engineering 2005; 32:739–752.
28. Kianoush MR, Mirzabozorg H, Ghaemian M. Dynamic analysis of rectangular liquid containers in three-dimensional
space. Canadian Journal of Civil Engineering 2006; 33:501–507.

Copyright © 2013 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2013; 42:1637–1656
DOI: 10.1002/eqe

Das könnte Ihnen auch gefallen