Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2007; 71:149–174


Published online 14 November 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nme.1921

Hydroelastic vibrations of flexible rectangular tanks partially


filled with liquid

Ding Zhou∗, † and Weiqing Liu


College of Civil Engineering, Nanjing University of Technology, Nanjing 210009, People’s Republic of China

SUMMARY
In this paper, the three-dimensional vibratory characteristics of flexible rectangular tanks partially filled
with liquid are studied. The surface waves of the liquid are taken into account in the analysis. Both the
bulging modes of the tank-wall vibration and the sloshing modes of the liquid oscillation are investigated.
The vibrating modes of the liquid–tank system are divided into four distinct categories: double symmetric
modes (SS); antisymmetric–symmetric modes (AS); symmetric–antisymmetric modes (SA) and double
antisymmetric modes (AA). Each of these categories is separately investigated. The velocity potential of
the liquid is analytically deduced by using a combination of the superposition method and the method
of separation of variables. According to the liquid–tank interface conditions and the orthogonality of
trigonometric functions, the coefficients in the solution of liquid velocity potential are expressed in the
integral forms including the tank–wall dynamic deflection. A set of reasonable static beam functions is
constructed as the admissible functions of the tank-wall vibration. The eigenfrequency equation of the
liquid–tank system is derived by using a combination of the Rayleigh–Ritz method and the Galerkin
method. Convergence study demonstrates the high accuracy and small computational cost of the proposed
approach. Finally, some numerical results are presented for the first time. Copyright q 2006 John Wiley
& Sons, Ltd.

Received 20 April 2006; Revised 11 September 2006; Accepted 26 September 2006

KEY WORDS: hydroelastic vibration; eigenfrequency; rectangular tank; liquid–tank interaction; super-
position method; Rayleigh–Ritz method; Galerkin method; semi-analytical method

1. INTRODUCTION

It is well known that storage tanks can be heavily damaged by liquid dynamic pressure under
earthquake loads. Some typical cases of tank damages have been reported [1–3]. Therefore, it is
important to understand the dynamic characteristics of liquid–tank interaction.

∗ Correspondence to: Ding Zhou, College of Civil Engineering, Nanjing University of Technology, Nanjing 210009,
People’s Republic of China.

E-mail: dingzhou57@yahoo.com

Copyright q 2006 John Wiley & Sons, Ltd.


150 D. ZHOU AND W. LIU

Numerous investigations on liquid–structure interaction can be found in literature, using dif-


ferent methods such as finite element methods [4, 5], boundary element methods [6–8], analyti-
cal methods [9–11] and semi-analytical methods [12–16]. There is no doubt that the numerical
methods such as finite element, boundary element or their combination can resolve the liquid–
structure interaction especially for structures with complex geometry, however, the modelling,
code preparation and numerical computation generally require a long time. Analytical methods
have the advantage of the best accuracy with the minimum computational cost. However, they
can be obtained only for some special cases. Semi-analytical methods possess both the advan-
tage of small computation cost over numerical methods and the advantage of wide applicability
over the exact analytical methods. Therefore, semi-analytical methods had a frequent applica-
tion in the analysis of liquid–tank interaction especially when the tank has a regular geometric
shape.
A close scrutiny among the literature on the dynamic analysis of liquid–tank interaction re-
veals that to date, most studies are concerned with cylindrical tanks [17–20]. Few studies are
concerned with the hydroelastic analysis of rectangular tanks. Although cylindrical tanks have
had a widest application in various types of tanks, in practical engineering the rectangular tanks
can also find their applications such as the storage of nuclear spent fuel assemblies, water tanks
and dams–reservoir system, etc. Housner [21, 22] and Haroun [23] studied the seismic response
of rigid rectangular tanks filled with liquid. Bauer [24] studied the hydroelastic vibration of
a simply supported elastic bottom in a rectangular tank with rigid walls. Soedel and Soedel
[11] studied the free and forced vibration of a simply supported rectangular plate connected
to a large reservoir of liquid at four edges and supporting a liquid with free sloshing surface.
Dogangun et al. [25] used finite element method to analyse the static and hydroelastic re-
sponse of rectangular tank. Kim et al. [26] studied the dynamic response of rectangular tanks
partially filled with liquid using the Rayleigh–Ritz method. In their analysis, two opposite tank-
walls parallel to the direction of the applied ground motion were considered to be rigid and
the other two opposite tank-walls were considered to be flexible, and the effect of the surface
waves of the liquid was neglected. Furthermore, Koh et al. [27] studied the seismic response
of rectangular tanks with four flexible walls by using the coupled boundary element–finite ele-
ment method. In their analysis, the effect of the surface waves of the liquid has been taken into
account.
In this paper, a semi-analytical method is presented for the hydroelastic characteristic analysis
of rectangular tanks with four flexible walls. The effect of the liquid surface waves is taken into
account in the analysis. The analytical method is applied to obtain the solution of the liquid velocity
potential. A set of suitable admissible functions [28, 29] is constructed as the approximate solution
of the tank-wall dynamic deflection. The Rayleigh–Ritz method combining with the Galerkin
method is applied to derive the eigenfrequency equation of the liquid–tank system, which includes
both the bulging modes of the tank-wall and the sloshing modes of the free liquid surface. According
to the symmetry of the liquid–tank structure, the vibration modes of the liquid–tank system are
divided into four distinct categories: double symmetric modes (SS); antisymmetric–symmetric
modes (AS); symmetric–antisymmetric modes (SA) and double antisymmetric modes (AA). Each
of these categories can be separately investigated. This greatly reduces the computational cost.
In the present analysis, the bulging modes of the tank-wall vibration and the sloshing modes of
the liquid oscillation can be distinctly distinguished so that the effect of various parameters on
these two types of modes can be clearly discussed. Some valuable results have been reported for
the first time.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 151

2. BASIC EQUATIONS

A rectangular tank with four flexible vertical walls of uniform thickness h and a horizontal rigid
bottom is partially filled with incompressible and non-viscous liquid of depth H , as shown in
Figure 1(a). The side lengths and height of the tank are 2a, 2b and c, 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 in-plane 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 (x, y, z, t):
* * *
vx = , vy = , vz = (1)
*x *y *z
According to the theory of fluid dynamics, the liquid velocity potential should satisfy the Laplace
equation:
2 2 2
*  *  * 
+ 2 + 2 =0 (2)
*x 2 *y *z
Considering the walls of the tank to be impermeable and no cavitation on the liquid–wall interface,
the consistent conditions between the liquid motion and the tank-wall deflection w(x, y, z, t) can
be given as
   
*  *w  *  *w 
= , = (3)
*x x=±a *t x=±a *y  y=±b *t  y=±b

Assuming that the liquid makes a small amplitude oscillation, the linearized sloshing condition at
the free liquid surface is
2
* 1 * 
+ =0 on z = H (4)
*z g *t 2

a
z,ζ

2b
H H

c y,η
b
c
x,ξ
o

(a) 2a (b)

Figure 1. The rectangular tank partially filled with liquid and its co-ordinate system: (a) the sketch of the
tank; and (b) the quarter of the tank and its co-ordinates.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
152 D. ZHOU AND W. LIU

The rigid bottom condition of the tank is



* 
=0 (5)
*z z=0
It is noteworthy that because of the symmetry of both the liquid and the tank about x–z plane and
y–z plane, respectively, the vibration modes of the liquid–tank system can be classified into four
distinct categories: i.e. double symmetric modes (SS), antisymmetric (about y–z plane)–symmetric
(about x–z plane) modes (AS), symmetric (about y–z plane)–antisymmetric (about x–z plane)
modes (SA) and double antisymmetric modes (AA). Each of these categories can be separately
investigated. Thus, four sets of eigenfrequency equation with small size can be obtained. In such
a case, only a quarter of the liquid–tank system, as shown in Figure 1(b), should be considered
and the boundary conditions of the liquid at the sides x = 0 and y = 0 should be presented. For
the SS vibration, one has
 
*  * 
= 0, =0 (6)
*x x=0 *y  y=0

For the AS vibration, one has


 
*  * 
= 0, =0 (7)
*t x=0 *y  y=0

For the SA vibration, one has


 
*  * 
= 0, =0 (8)
*x x=0 *t  y=0

For the AA vibration, one has


 
*  * 
= 0, =0 (9)
*t x=0 *t  y=0

3. SOLUTION OF VELOCITY POTENTIAL

When the liquid–tank system makes free vibration, the frequency of the liquid motion is identical
with the frequency of the tank vibration, therefore
(x, y, z, t) = jejt (x, y, z),
w(x, y, z, t) = ejt W (x, y, z) (10)

where  is the natural frequency of the liquid–tank system and j = −1. Using the superposition
method to solve the velocity potential, one can assume that
(x, y, z) = z (x, y, z) + x (x, y, z) +  y (x, y, z) (11)
where z (x, y, z) satisfies the zero velocity conditions on all the four walls. x (x, y, z) satisfies
the flexible wall conditions in the x direction, however, the rigid wall conditions in the y direction
and the no sloshing condition on the free liquid surface.  y (x, y, z) satisfies the flexible wall

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 153

conditions in the y direction, however, the rigid wall conditions in the x direction and the no
sloshing condition on the free liquid surface. Namely,
 
*z  *z 
= 0, =0 (12a)
*x x=±a *y  y=±b
  
*x  *x  *x 
= W |x=±a , = 0, =0 (12b)
*x x=±a *y  y=±b *t z=H
  
* y  * y  * y 
= 0, = W | y=±b , =0 (12c)
*x x=±a *y  y=±b *t z=H

For the simplification in expression, the following dimensionless co-ordinates and parameters are
introduced:
 = x/a,  = y/b,  = z/H,  = a/b
(13)
 = H/b,  = c/b,  = z/c = H /c = /
The solutions of the velocity potential corresponding to every vibration modes can be analytically
given in an infinite series form by using the method of separation of variables. For the SS vibration,
one has

∞ 
∞ 1
z (, , ) = b ci j cos(i ) cos(j ) cosh(qi j ) (14a)
i=0 j=0 cosh qi j

qi j =  i 2 /2 + j 2 (14b)

∞ 
∞ 2
m Ilm
x (, , ) = 2b cos[(l + 0.5) ] cos(m ) cosh(slm ) (15a)
l=0 m=0 slm sinh slm

slm = m 2 /2 + (l + 0.5)2 /2 (15b)
 1 1
2
Ilm = W |=1 cos(m ) cos[(l + 0.5) ] d d (15c)
0 0

∞ 
∞ 1
n Iln
 y (, , ) = 2b cos[(l + 0.5) ] cos(n ) cosh(rln ) (16a)
l=0 n=0 rln sinh rln

rln =  n 2 + (l + 0.5)2 /2 (16b)
 1 1
1
Iln = W |=1 cos(n ) cos[(l + 0.5) ] d d (16c)
0 0

where ci j are the unknown constants and

0 = 1, i =2 for i  = 0 (17)

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
154 D. ZHOU AND W. LIU

For the AS vibration, one has



∞ 
∞ 1
z (, , ) = b ci j sin[(i + 0.5) ] cos( j ) cosh(qi j ) (18a)
i=0 j=0 cosh qi j

qi j =  (i + 0.5)2 /2 + j 2 (18b)

∞ 
∞ 2
m Ilm
x (, , ) = 2b cos[(l + 0.5) ] sin[(m + 0.5) ] cosh(slm ) (19a)
l=0 m=0 slm sinh slm

slm = (m + 0.5)2 /2 + (l + 0.5)2 /2 (19b)
 1 1
2
Ilm = W |=1 sin[(m + 0.5) ] cos[(l + 0.5) ] d d (19c)
0 0

∞ 
∞ 1
n Iln
 y (, , ) = 2b cos[(l + 0.5) ] cos(n ) sinh(rln ) (20a)
l=0 n=0 rln sinh rln

rln =  n 2 + (l + 0.5)2 /2 (20b)
 1 1
1
Iln = W |=1 cos(n ) cos[(l + 0.5) ] d d (20c)
0 0

For the SA vibration, one has



∞ 
∞ 1
z (, , ) = b ci j cos(i ) sin[( j + 0.5) ] cosh(qi j ) (21a)
i=0 j=0 cosh qi j

qi j =  i 2 /2 + ( j + 0.5)2 (21b)

∞ 
∞ 2
m Ilm
x (, , ) = 2b cos[(l + 0.5) ] cos(m ) sinh(slm ) (22a)
l=0 m=0 slm cosh slm

slm = m 2 /2 + (l + 0.5)2 /2 (22b)
 1 1
2
Ilm = W |=1 cos(m ) cos[(l + 0.5) ] d d (22c)
0 0

∞ 
∞ 1
Iln
 y (, , ) = 4b cos[(l + 0.5) ] sin[(n + 0.5) ] cosh(rln ) (23a)
l=0 n=0 rln sinh rln

rln =  (n + 0.5)2 + (l + 0.5)2 /2 (23b)
 1 1
1
Iln = W |=1 sin[(n + 0.5) ] cos[(l + 0.5) ] d d (23c)
0 0

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 155

For the AA vibration, one has



∞ 
∞ 1
z (, , ) = b ci j sin[(i + 0.5) ] sin[( j + 0.5) ] cosh(qi j ) (24a)
i=0 j=0 cosh qi j

qi j =  (i + 0.5)2 /2 + ( j + 0.5)2 (24b)

∞ 
∞ 2
m Ilm
x (, , ) = 2b cos[(l + 0.5) ] cos[(m + 0.5) ] sinh(slm ) (25a)
l=0 m=0 slm cosh slm

slm = (m + 0.5)2 /2 + (l + 0.5)2 /2 (25b)
 1 1
2
Ilm = W |=1 sin[(m + 0.5) ] cos[(l + 0.5) ] d d (25c)
0 0

∞ 
∞ 1
Iln
 y (, , ) = 4b cos[(l + 0.5) ] sin[(n + 0.5) ] sinh(rln ) (26a)
l=0 n=0 rln sinh rln

rln =  (n + 0.5)2 + (l + 0.5)2 /2 (26b)
 1 1
1
Iln = W |=1 sin[(n + 0.5) ] cos[(l + 0.5) ] d d (26c)
0 0

It should be mentioned that when determining the unknown coefficients in Equations (15a), (16a),
(19a), (20a), (22a), (23a), (25a) and (26a), the orthogonal relations among the trigonometric
functions have been utilized. In the following numerical computations, all the series in the above
equations will be truncated: i and j up to I + 1 and J + 1, respectively, m and n up to M + 1 and
N + 1, respectively, and l up to L + 1.

4. RAYLEIGH–RITZ–GALERKIN METHOD

It is obvious that for the liquid–tank interaction considered in the present study, to obtain an
exact solution is impossible, even for an empty flexible rectangular tank. Therefore, approximate
numerical methods have to be used. The high accuracy and small computational cost can be
achieved by using the following semi-analytical approach: a combination of the Rayleigh–Ritz
method and the Galerkin method.

4.1. Rayleigh quotient


The maximum kinetic energy Tl of the liquid in the tank is

2 ∗ ∗ 1
Tl =  Tl , Tl = l (∇)2 dv (27)
2 Vl

where l defines the liquid mass density, Vl defines the liquid domain and ∇ defines the gradient
operator. Considering Equation (2) and applying the Green’s theorem [30] to the harmonic function

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
156 D. ZHOU AND W. LIU

, the volume integral in Equation (27) can be transformed into the surface integral around the
liquid surface Sl . Therefore, one has

1
Tl∗ = l  ∇Qn ds (28)
2 Sl

where n is the unit normal vector on the liquid surface, whose positive direction is outward the
liquid. Substituting the boundary conditions of the velocity potential at  = 0,  = 0 and  = 0,
respectively, into Equation (28), one has
       
∗ 1 l ab 1 1 *  1 laH 1 1 * 
Tl =  d d +  d d
2 H 0 0 * =1 2 b 0 0 * =1
 1 1  
1 l bH * 
+  d d (29)
2 a 0 0 * =1

The maximum kinetic energy T p of the tank-wall is


 1 1  1 1
1 1
Tp =  2
T p∗ , T p∗ = p hac (W |=1 ) d d +
2
p hbc (W |=1 )2 d d (30)
2 0 0 2 0 0

where p is the mass density of the tank-wall.


The maximum potential energy Ul of the liquid is
    2    
1 l abg 1 1 *  1 l ab2 1 1 * 
Ul = d d =  d d (31)
2 H2 0 0 * =1 2 H 0 0 * =1

The maximum potential energy U p of the tank-wall is [31]


⎧  2 2  2   2 2
 
1 aD 1 1 ⎨  4 * W
2
 2 * W* W * W
Up = +2 +
0 ⎩  
3 2 2 2
2 c 0 * * * *2
⎡  2 ⎤⎫

2 2 2 ⎬
⎣* W * W * W
2
− 2(1 − ) − ⎦ d d
 *2 *2 ** ⎭
=1

 1  1 ⎨  2 2  2 2
  2 2
1 bD 4 * W 2 * W * W * W
+  + 2 +
2 c3 0 0 ⎩ *2 *2 *2 *2
⎡  2 ⎤⎫
* W ⎦⎬
2 2 2
* W* W
− 2(1 − )2 ⎣ − d d (32)
*2 *2 ** ⎭
=1

where D is the flexural rigidity of the tank-wall.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 157

The Rayleigh quotient of the liquid–tank system is


Ul + U p
2 = ∗ (33)
Tl + T p∗
Substituting Equations (29) and (31) into the above equation, the Rayleigh quotient can be rewritten
as the following simplified formulation:
Up
2 =  1 1       (34)
1 laH *   1 l bH 1 1 * 
 d d +  d d + T ∗
2 b 0 0 *  =1 2 a 0 0 *  =1
p

Considering Equation (11) and the boundary conditions of each parts of the velocity potential, one
has
       y 
1 laH 1 1 *  1 laH 1 1 x * 
 d d =  d d
2 b 0 0 * =1 2 b 0 0 * =1
   
1 laH 1 1 * y 
+ y d d
2 b 0 0 * =1
   y 
1 laH 1 1 z * 
+  d d (35a)
2 b 0 0 * =1
       x 
1 l bH 1 1 *  1 l bH 1 1 x * 
 d d =  d d
2 a 0 0 * =1 2 a 0 0 * =1
   
1 bH 1 1 *x 
+ l y d d
2 a 0 0 * =1
   
1 bH 1 1 *x 
+ l z d d (35b)
2 a 0 0 * =1

4.2. Eigenfrequency equation


Assume that the dynamic deflection function W (, , ) of the tank-wall can be written in the
form of

∞  ∞ f q () for  = 1
W (, , ) = cqs Rs () (36)
q=1 s=1 gq () for  = 1
where cqs are the unknown constants. In the following section, the admissible functions f q (),
gq () and Rs () will be reasonably developed.
Substituting Equation (36) into Equation (34), one can observe that all the right-hand terms in
Equation (34) contain the unknown constants cqs . Minimizing the Rayleigh quotient with respect to
these constants and truncating q up to I +1 and s up to J +1, a set of algebraic equation is given as
K pp C − 2 [(M pp + M pl )C + M pl C] = 0 (37)

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
158 D. ZHOU AND W. LIU

in which,
C = [c11 , c12 , . . . , c1J , c21 , c22 , . . . , c2J , . . . , c I 1 , c I 2 , . . . , c I J ]T
(38a)
C = [c11 , c12 , . . . , c1J , c21 , c22 , . . . , c2J , . . . , c I 1 , c I 2 , . . . , c I J ]T

pp (2,2) (0,0) (0,0) (2,2) (0,2) (2,0) (2,0) (0,2)


K qsq s = 4 /3 E qq Fss + E qq Fss + 2 /(E qq Fss + E qq Fss )

(1,1) (1,1) (2,2) (0,0) (0,0) (2,2) (0,2) (2,0)


+ 2(1 − )2 /E qq Fss + 4 E qq Fss + E qq Fss + 2 (E qq Fss

(2,0) (0,2) (1,1) (1,1) pp (0,0) (0,0) (0,0) (0,0)


+ E qq Fss ) + 2(1 − )2 E qq Fss , Mqsq s = E qq Fss + E qq Fss

q, q = 1, 2, . . . , I, s, s = 1, 2, . . . , J (38b)

ph l abH
2 = 2 c 4 , = (38c)
D p h(a + b)c
where  is the dimensionless natural frequency, which is called as eigenfrequency in the following
analysis and
 1
(i, j)
E qq = [di f q ()/di ][d j f q ()/d j ] d
0
 1
(i, j)
E qq = [di gq ()/di ][d j gq ()/d j ] d (39)
0
 1
(i, j) i j
Fss = [di Rs ()/d ][d j Rs ()/d ] d, i, j = 0, 1, 2
0
pp pp
The first two subscripts in K qsq s and Mqsq s determine their row locations and the later two
subscripts determine the column locations in matrices K pp and M pp . In the following analysis,
the similar subscript descriptions will be used in all the matrices such as M pl and M pl without
further explanation. It is obvious that the matrices M pl and M pl are related to the types of the
vibration modes. For the SS modes, one has

pl 
L 
M
em
Mqsq s = 2(1 + ) Pmq Pmq Q ls Q ls
s
l=0 m=0 lm tanh slm
  L M N  
1    1 2
+2 1 + (−1) m+n
em en +
 l=0 m=0 n=0 (n )2 + slm
2 (m )2 + rln
2

× (Pmq P nq + Pmq P nq )Q ls Q ls


L 
N 1
+ 2(1 + ) P nq P nq Q ls Q ls (40a)
l=0 n=0 rln tanh rln

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 159

  L
pl 1  l qi j tanh qi j
M qsi j = 1 + (−1)l [(−1)i P jq + (−1) j Piq ]Q ls for i  = j  = 0
 l=0 (l ) + qi j
2 2
(40b)
 
pl 1
M qs00 = 1 + (P 0q + P0q )Q 0s , i = 0, 1, 2, . . . , I , j = 0, 1, 2, . . . , J

where
 1  1
Pmq = f q () cos(m ) d, P nq = gq () cos(n ) d
0 0
  (40c)
1 1
Q ls = Rs () cos[(l + 0.5) ] d, Q ls = Rs () cos(l ) d
0 0

For the AS modes, one has

pl 
L 
M 1
Mqsq s = 4(1 + ) Pmq Pmq Q ls Q ls
s
l=0 m=0 lm tanh slm
  L M N 
1    1
+4 1 + (−1) m+n
n
 l=0 m=0 n=0 (n ) + slm
2 2

2
+ (Pmq P nq + Pmq P nq )Q ls Q ls
[(m + 0.5) ]2 + rln
2


L 
N
n
+ 4(1 + ) tanh rln P nq P nq Q ls Q ls (41a)
l=0 n=0 rln

  L
pl 1  l qi j tanh qi j
M qsi j = 1 + (−1)l [(−1)i P jq + (−1) j Piq ]Q ls (41b)
 l=0 (l )2 + qi2j
 1  1
Pmq = f q () sin[(m + 0.5) ] d, P nq = gq () cos(n ) d (41c)
0 0

For the SA modes, one has

pl 
L 
M
m
Mqsq s = 4(1 + ) tanh slm Pmq Pmq Q ls Q ls
l=0 m=0 slm

  
1 
L 
M 
N 1
+2 1 + (−1)m+n m
 l=0 m=0 n=0 [(n + 0.5) ]2 + slm
2


2
+ (Pmq P nq + Pmq P nq )Q ls Q ls
(m )2 + rln
2

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
160 D. ZHOU AND W. LIU


L 
N 1
+ 4(1 + ) P nq P nq Q ls Q ls (42a)
r
l=0 n=0 ln tanh rln
 
pl 1 
L
l qi j tanh qi j
M qsi j = 1+ (−1)l [(−1)i P jq + (−1) j Piq ]Q ls (42b)
 l=0 (l )2 + qi2j
 1  1
Pmq = f q () cos(m ) d, P nq = gq () sin[(n + 0.5) ] d (42c)
0 0

For the AA modes, one has

pl L  M 1
Mqsq s = 4(1 + ) tanh slm Pmq Pmq Q ls Q ls
l=0 m=0 slm

  
1 
L 
M 
N 1
+4 1 + (−1) m+n
 l=0 m=0 n=0 [(n + 0.5) ]2 + slm
2


2
+ (Pmq P nq + Pmq P nq )Q ls Q ls
[(m + 0.5) ]2 + rln
2

L  N 1
+ 4(1 + ) tanh rln P nq P nq Q ls Q ls (43a)
l=0 n=0 rln

  L
pl 1  el qi j tanh qi j
M qsi j = 1 + (−1)l [(−1)i P jq + (−1) j Piq ]Q ls (43b)
 l=0 (l )2 + qi2j
 1  1
Pmq = f q () sin[(m + 0.5) ] d, P nq = gq () sin[(n + 0.5) ] d (43c)
0 0

Equation (37) cannot be solved until C is given. Substituting Equation (11) into the sloshing
equation (4) and utilizing the orthogonality relations among the trigonometric functions can provide
a set of additional Galerkin equation. Therefore, an expanding eigenfrequency equation increasing
the dimension of the associated eigenvalue problem from (I × J ) × (I × J ) to (I × J + I × J ) ×
(I × J + I × J ) is obtained:
 pp    pp  
K 0 C M + M pl M pl C
−  2
=0 (44)
K pl Kll C 0 Mll C

in which,
D
= 3
(45)
g p hc

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 161

It is clear that  reflects the effect of the gravity on the eigenfrequency. The matrices K pl , Kll and
Mll are also related to the types of the vibration modes. For the SS modes, one has
 
pl 4L (−1)i+l (−1) j+l
K i jsq = 2 (l + 0.5) P jq + Piq Q ls (46a)
l=0 (i )2 + r 2jl ( j )2 + sil2
qi j
K illji j = − tanh qi j ii  j j (46b)
ij

Millji j = − ii  j j (46c)
 ij

where


1, i = j, ⎨ 1, i = 0 and j = 0

i j = ij = 2, i = 0 or j = 0 (47)
0, i  = j, ⎪

4, i  = 0 and j  = 0
For the AS modes, one has
 
pl 
4L (−1)i+l (−1) j+l
K i jsq = 2 (l + 0.5) P jq + Piq Q ls (48a)
l=0 (i + 0.5 )2 + r 2jl ( j )2 + sil2
qi j
K illji j = − tanh qi j ii  j j (48b)
j

Millji j = −   (48c)
 j ii j j

For the SA modes, one has


 
pl 
4L (−1)i+l (−1) j+l
K i jsq = 4 (l + 0.5) P jq + Piq Q ls (49a)
l=0 (i )2 + r 2jl ( j + 0.5 )2 + sil2
qi j
K illji j = − tanh qi j ii  j j (49b)
i

Millji j = −   (49c)
 i ii j j

For the AA modes, one has


 
pl 
4L (−1)i+l (−1) j+l
K i jsq = 8 (l + 0.5) P jq + Piq Q ls (50a)
l=0 (i )2 + r 2jl ( j )2 + sil2

K illji j = −qi j tanh qi j ii  j j (50b)



Millji j = − ii  j j (50c)


Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
162 D. ZHOU AND W. LIU

It is obvious that if the liquid surface waves are neglected, the eigenfrequency equation is
degenerated into

K pp C − 2 (M pp + M pl )C = 0 (51)

In this case, only the bulging modes of the tank-wall vibration are considered.
It should be pointed out that for preventing the numerical overflow in calculation, all the
hyperbolic functions appeared in the integrals have been expanded into the cosine series. Moreover,
considering the fact that the convergent rate of the summing series in Equations (46a), (48a), (49a)
and (50a) is lower than that in other equations, the number of the summing series used in Equations
(46a), (48a), (49a) and (50a) is taken as four times of that used in other equations.

5. 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. As the tank-wall can only provide transverse deflection but
no in-plane deformation, the vibration of the tank-wall can be equivalent to the vibration of a line
supported rectangular plate as shown in Figure 2(a) where only a quarter of the tank is considered.
The connecting line of the adjacent walls corresponds to the internal line support which prevents
the transverse motion of the plate but offers no resistance to the rotation. Letting  = x/(a + b),
the co-ordinate of the line support is 1 = 1/(1 + ).
Taking a unit width strip from the rectangular plate in the  direction [28], as shown in
Figure 2(b), which is acted by a set of static sinusoidal series loads, the transverse deflection u()
of the strip satisfies the following differential equation:

d4 u 

= P1 ( − 1 ) + Q i (i )4 sin(i ) (52)
d4 i=1

z,  Unit width strips


a The line support

c c

o x,ξ
b a
b
(a)
u v
b a c
x,ξ z, ζ
o o
(b) (c)

Figure 2. The relation of the tank-wall with the line supported rectangular plate and the relations of the
plate with the beams: (a) the equivalent model of the tank-wall; (b) a unit width strip taken from the plate
in the x direction; and (c) a unit width strip taken from the plate in the z direction.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 163

where P1 is the dimensionless reaction of the internal point support corresponding to the internal
line support of the rectangular plate, ( − 1 ) is the Dirac delta function. Q i are the amplitude
coefficients of the sinusoidal series loads.
The general solution of Equation (52) can be easily given:


u() = Q i u i ()
i=1 (53)
u i () = a0i + a1i  + a2i 2 + a3i 3 + P1i ( − 1 )3 U ( − 1 )/6 + sin(i )

where both aki (k = 0, 1, 2, 3) and P1i are the unknown constants, which can be determined by the
boundary conditions of the beam and the zero deflection condition at the internal point support.
U ( − 1 ) is the Heaviside function. The boundary conditions of the beam corresponding to those
of the rectangular plate are also related to the symmetries of the vibration modes considered. For
the SS modes, one has
3
du/d = 0, d3 u/d = 0 at  = 0 (54a)
3
du/d = 0, d3 u/d = 0 at  = 1 (54b)

For the AS modes, one has


2
u = 0, d2 u/d = 0 at  = 0 (55a)
3
du/d = 0, d3 u/d = 0 at  = 1 (55b)

For the SA modes, one has


3
du/d = 0, d3 u/d = 0 at  = 0 (56a)
2
u = 0, d2 u/d = 0 at  = 1 (56b)

For the AA modes, one has


2
u = 0, d2 u/d = 0 at  = 0 (57a)
2
u = 0, d2 u/d = 0 at  = 1 (57b)

For whatever types of vibration modes, u() should satisfy the zero deflection condition at the
internal point support  = 1 :
u(1 ) = 0 (58)
Substituting Equation (53) into Equations (54)–(58), the unknown constants aki (k = 0, 1, 2, 3) and
P1i can be uniquely determined. For the SS modes, one has
2
a0i = i 1 {1 − 1 [1 − (−1)i ]/2} + (i )3 1 {1 − (1 − 1 )2 [1 − (−1)i ] − 2/3}/4 + sin(i 1 )
a1i = −i , a2i = i [1 − (−1)i ]/2 − (i )3 {1 − (1 − 1 )2 [1 − (−1)i ]}/4 (59)
a3i = (i )3 /6, P1i =− (i )3 [1 − (−1)i ]

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
164 D. ZHOU AND W. LIU

For the AS modes, one has


2 2
a0i = i 1 [1 − 2(i )2 1 /3] + (1 − 1 )1 {3i + (i )3 [1 + 1 − 1 /2]
2
+ 3 sin(i 1 )/(1 − 1 )}/2(21 − 1 − 1) − sin(i 1 )
a1i = −i
2
a2i = −(i )3 /2 − (1 − 1 ){3i + (i )3 [1 + 1 − 1 /2] (60)
2
+ 3 sin(i 1 )/(1 − 1 )}/2(21 − 1 − 1)
a3i = (i )3 /6
2 2
P1i = {3i + (i )3 [1 + 1 − 1 /2] + 3 sin(i 1 )/(1 − 1 )}/2(21 − 1 − 1)
For the SA modes, one has
a0i = 0
a1i = {(i )(−1)i [2 + (i )2 (1 − 1 )2 ]1 − 2 sin(i 1 )}/4(3 − 21 ) − sin(i 1 )/1
a2i = 0 (61)
2
a3i = −{(i )(−1) [2 + (i ) (1 − 1 ) ]1 − 2 sin(i 1 )}/41 (3 − 21 )
i 2 2

2
P1i = 3{(i )(−1)i [2+(i )2 (1−1 )2 ]1 −2 sin(i 1 )}/21 (3−21 ) + (i )3 (−1)i
For the AA modes, one has
a0i = 0, a1i =− (2 − 1 )/21 (1 − 1 ), a2i = 0
2 2
(62)
a3i = sin(i 1 )/21 (1 − 1 ), P1i =− 3 sin(i 1 )/1 (1 − 1 )2

Similarly, taking a unit width strip from the rectangular plate in the  direction [29], as shown
in Figure 2(c), which is also acted by a set of static sinusoidal series loads, the deflection v()
satisfies the following differential equation:
d4 v 

4
= Q j ( j )4 sin( j ) (63)
d j=1

The general solution of the above equation can be easily given:




j j j j
v() = Q j v j (), v j () = b0 + b1  + b2 2 + b3 3 + sin( j ) (64)
j=1

For the open rectangular tank, the boundary conditions of the strip is
dv
v = 0, =0 at  = 0 (65a)
d
d2 v d3 v
2
= 0, = 0 at  = 1 (65b)
d d3

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 165

Substituting Equation (64) into the above two equations can uniquely determine the unknown
j
constants bk (k = 0, 1, 2, 3). Therefore, one has
2 3
v j () = − j  − ( j )3 (−1) j  /2 + ( j )3 (−1) j  /6 + sin( j ) (66)

u i () (i = 1, 2, 3, . . .) and v j () ( j = 1, 2, 3, . . .) are called as the static beam functions. It is


obvious that there is the co-ordinate relation:

⎨ /(1 + ) for 01/(1 + )
= (67)
⎩ 1/(1 + ) + /(1 + ) for 1/(1 + )1

Therefore, the admissible functions of the tank-wall can be taken as


f q () = u q () where  = 1/(1 + ) + /(1 + )
(68)
gq () = u q () where  = /(1 + ), Rs () = vs ()

6. NUMERICAL RESULTS

In this section, some numerical results are presented by solving the eigenvalue equation (44). The
calculation is carried out on the Pentium IV microcomputer and the double-precision program has
been coded by the Fortran language. The integrals in equations (39), (40c), (41c), (42c) and (43c)
are numerically evaluated by the piecewise Gauss numerical integration of 24 points. The terms in
the calculation are taken as I = J = 5, L = M = N = 30 and I = J = 10, respectively. As a typical
example, the vibratory characteristics of a cubic tank partially filled with liquid are analysed in
detail.

6.1. Convergence and comparison study


It is obvious that the accuracy of the results is concerned with the size of the truncated matrices
and the term number of the summing series. The first six eigenfrequencies of the empty tank for
two groups of different size parameters are given in Table I and compared with those presented by
Kim and Dickinson [32] using the orthogonally generated one-dimensional polynomial functions
as the admissible functions of the tank-wall vibration. Five terms of the static beam functions are
used in each direction. Good agreements are observed for all cases.
In the following study, a cubic tank fully filled with liquid ( = 1;  = 2;  = 2) is taken to show
the convergence. The mass ratio of the liquid and tank-wall is = 5. The rigidity parameter of the
tank-wall is  = 2. The convergence study on the first six eigenfrequencies of the bulging modes
with respect to the term number of the static beam functions are given in Table II. It is seen that the
bulging eigenfrequencies monotonically decrease with the increase of the term number of the static
beam functions. The results show rapid convergence of the present method. The convergence study
on the first six eigenfrequencies of the bulging modes with respect to the term number of summing
series in each of matrices and the number of the Galerkin equation are given in Tables III and IV,
respectively. It is seen from Table III that the bulging eigenfrequencies monotonically decrease
with the increase of the term number of the summing series. However, it is seen from Table IV
that the bulging eigenfrequencies monotonically increase with the increase of the number of the

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
166 D. ZHOU AND W. LIU

Table I. The comparison study on the first six eigenfrequencies of the empty rectangular tank.
Mode type 1 2 3 4 5 6
a : b : c=1 : 1 : 2
SS 12.757 24.038 33.076 40.032 72.661 76.953
Kim and Dickinson [32]∗ 12.689 24.014 33.066 40.145 72.516 77.055
AS (SA) 17.632 36.045 52.133 71.208 74.576 107.531
Kim and Dickinson [32]∗ 17.577 36.104 52.069 71.463 74.579 107.13
AA 41.984 63.118 63.536 80.713 103.212 116.747
Kim and Dickinson [32]∗ 41.703 63.016 64.130 81.820 103.29 118.57

a : b : c = 1 : 1.5 : 2.5
SS 11.676 28.272 31.131 45.439 70.389 75.551
Kim and Dickinson [32]∗ 11.624 28.219 31.142 45.571 70.223 75.048
AS 13.850 31.878 64.226 70.690 85.304 91.957
Kim and Dickinson [32]∗ 13.818 31.937 63.684 70.590 85.026 92.913
SA 22.918 41.983 42.697 59.315 81.363 96.310
Kim and Dickinson [32]∗ 22.810 42.316 42.789 60.051 81.616 97.489
AA 35.763 55.525 78.996 94.245 98.171 130.022
Kim and Dickinson [32]∗ 35.622 55.589 79.109 94.562 98.746 131.42
∗ Six terms of orthogonal polynomial functions are used in each direction.

Table II. The convergence study on the first six eigenfrequencies of bulging modes with
respect to the term number of the static beam functions.
I×J 1 2 3 4 5 6
SS modes
2×2 6.911 7.015 17.840 18.702
3×3 6.866 7.006 17.490 18.667 39.296 44.220
4×4 6.860 6.993 17.477 18.654 39.002 43.062
5×5 6.855 6.992 17.434 18.645 38.879 42.888

AS (SA) modes
2×2 8.533 18.253 26.860 40.375
3×3 8.506 18.213 26.383 37.900 43.706 65.044
4×4 8.489 18.196 26.237 37.660 43.076 58.907
5×5 8.480 18.175 26.212 37.494 42.931 58.016

AA modes
2×2 24.232 27.798 38.339 43.684
3×3 24.229 27.545 37.994 41.741 64.664 66.186
4×4 24.222 27.376 37.954 41.546 64.406 65.911
5×5 24.215 27.347 37.856 41.352 64.387 65.854

Galerkin equation. This means that the Ritz solutions as shown in Tables II and III provide the
upper bounds (downward convergence) of bulging eigenfrequencies while the Galerkin solutions
as shown in Table IV provide the lower bounds (upward convergence) of bulging eigenfrequencies.
In general, five terms of the static beam functions and 10 terms of the Galerkin equation in each
direction, and 30 terms of summing series in each of matrices can provide the satisfactory accuracy,
which results in a quite small computational cost.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 167

Table III. The convergence study on the first six eigenfrequencies of bulging modes
with respect to the term number of summing series.

L=M=N 1 2 3 4 5 6
SS modes
10 6.860 7.002 17.510 18.736 39.160 43.167
20 6.857 6.993 17.446 18.659 38.922 42.931
30 6.855 6.992 17.434 18.645 38.879 42.888

AS (SA) modes
10 8.497 18.262 26.243 37.670 43.190 58.054
20 8.485 18.188 26.217 37.522 42.971 58.022
30 8.480 18.175 26.212 37.494 42.931 58.016

AA modes
10 24.265 27.358 38.002 41.480 64.818 66.329
20 24.223 27.349 37.862 41.372 64.454 65.928
30 24.215 27.347 37.856 41.352 64.387 65.854

Table IV. The convergence study on the first six eigenfrequencies of bulging modes
with respect to the number of the Galerkin equation.

I=J 1 2 3 4 5 6
SS modes
5 6.852 6.984 17.429 18.640 38.876 42.886
10 6.855 6.992 17.434 18.645 38.879 42.888

AS (SA) modes
5 8.476 18.171 26.210 37.492 42.929 58.015
10 8.480 18.175 26.212 37.494 42.931 58.016

AA modes
5 24.215 27.347 37.856 41.352 64.387 65.854
10 24.265 27.358 38.002 41.448 64.818 66.329

The convergence study on eigenfrequencies of the sloshing modes is not listed here
because the convergence rate of the sloshing eigenfrequencies is much more rapid than that
of the bulging eigenfrequencies. No difference could be found if the term numbers in the
sloshing modes are the same as those used in the convergence study for the bulging modes.
A comparison of the first six sloshing eigenfrequencies with those in the rigid tank is given in
Table V. It iswell known that the sloshing  eigenfrequency of liquid in the rigid rectangular tank is
 = ( /2) (i/)2 + j 2 tanh( /2 (i/)2 + j 2 ) where i = 0, 2, 4, . . . and j = 0, 2, 4, . . . corre-
sponds to the SS mode; i = 1, 3, 5, . . . and j = 1, 3, 5, . . . corresponds to the AA mode;
i = 0, 2, 4, . . . and j = 1, 3, 5, . . . corresponds to the SA mode; i = 1, 3, 5, . . . and j = 0, 2, 4, . . .
corresponds to the AS mode. It is seen from Table V that the flexibility of the tank decreases the
eigenfrequencies of the sloshing modes, especially for the fundamental sloshing eigenfrequencies.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
168 D. ZHOU AND W. LIU

Table V. The comparison study on the first six eigenfrequencies of sloshing modes.
1 2 3 4 5 6
SS modes
1.760 1.769 2.105 2.501 2.505 2.649
(1.772) (1.772) (2.108) (2.507) (2.507) (2.650)

AS (SA) modes
1.244 1.873 2.167 2.379 2.544 2.800
(1.251) (1.874) (2.171) (2.380) (2.545) (2.802)

AA modes
1.489 2.227 2.229 2.581 2.830 2.830
(1.490) (2.229) (2.229) (2.582) (2.830) (2.830)

Note: Data in parentheses are sloshing eigenfrequencies of the liquid in the rigid tank.

Table VI. The effect of the ratio of liquid-depth to tank-height


on the first six bulging eigenfrequencies.
H/c 1 2 3 4 5 6
SS modes
0.5 11.924 18.690 24.088 27.281 54.578 56.144
0.6 11.014 14.734 22.661 26.793 49.916 53.082
0.7 9.937 11.837 22.353 26.024 45.385 48.608
0.8 8.891 9.754 22.041 23.960 43.735 46.305
0.9 7.950 8.211 20.783 20.905 42.807 45.767
1.0 6.967 7.118 17.490 18.677 38.893 42.898

AS (SA) modes
0.5 15.956 24.974 39.523 54.582 57.795 76.056
0.6 14.146 23.544 35.260 51.482 56.588 69.757
0.7 12.358 22.998 32.553 46.572 54.780 65.497
0.8 10.865 22.134 30.513 43.511 51.299 62.583
0.9 9.636 20.489 28.421 41.210 47.765 60.305
1.0 9.068 18.211 26.228 37.507 42.945 58.019

AA modes
0.5 36.292 42.564 48.847 65.787 84.529 90.926
0.6 32.774 36.960 48.536 64.604 78.259 82.544
0.7 30.052 33.332 48.012 60.133 72.453 78.351
0.8 27.941 30.860 45.743 53.609 70.758 77.528
0.9 26.108 28.998 42.062 47.316 69.676 73.968
1.0 24.231 27.358 37.849 41.368 64.394 65.863

6.2. Parametric effect study


The parameter effects on eigenfrequencies of the liquid–tank system are investigated for a cubic
tank. Taking the mass ratio of the liquid and tank-wall = 5 and the rigidity parameter  = 1, the
effect of the liquid depth on the first six bulging eigenfrequencies is given in Table VI when the
tank is partially filled with the liquid. One can see that the bulging eigenfrequencies decrease with
the increase of the liquid depth.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 169

The effect of the liquid depth on the first two sloshing eigenfrequencies of the SS modes
and on the fundamental sloshing eigenfrequencies of the AS(SA) modes and AA modes are
given in Figures 3–5, respectively. It is shown that the effect of the tank-wall flexibility on
sloshing eigenfrequencies increases with the increase of the liquid depth, especially on the
fundamental sloshing eigenfrequency of the SS modes. Moreover, from Figure 3, one can see
that the tank-wall flexibility tends to separate the duplicate sloshing eigenfrequencies in the

2.51

2.5

2.49

2.48

2.47

2.46

2.45
0.5 0.6 0.7 0.8 0.9 1
H/c

Figure 3. The effect of the ratio of liquid-depth to tank-height on the first two sloshing eigenfrequencies of
SS modes, , the first sloshing eigenfrequency; , the second sloshing eigenfrequency; 䊏, the fundamental
duplicate sloshing eigenfrequencies of the liquid in the rigid tank.

1.77

1.76

1.75

1.74
Ω1

1.73

1.72

1.71

1.7

1.69
0.5 0.6 0.7 0.8 0.9 1
H /c

Figure 4. The effect of the ratio of liquid-depth to tank-height on the fundamental


sloshing eigenfrequency of AS(SA) modes, , fundamental sloshing eigenfrequency;
䊏, fundamental sloshing eigenfrequency of the liquid in the rigid tank.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
170 D. ZHOU AND W. LIU

2.11

2.105

2.1

Ω1
2.095

2.09

2.085

2.08
0.5 0.6 0.7 0.8 0.9 1
H/ c

Figure 5. The effect of the ratio of liquid-depth to tank-height on the fundamental sloshing eigenfre-
quency of AA modes, , fundamental sloshing eigenfrequency; 䊏, fundamental sloshing eigenfrequency
of the liquid in the rigid tank.

Table VII. The effect of the rigidity parameter on the


first six sloshing eigenfrequencies.
 1 2 3 4 5 6
SS modes
0.5 3.433 3.522 4.187 4.958 5.004 5.291
0.6 3.162 3.219 3.828 4.538 4.570 4.832
0.7 2.939 2.982 3.547 4.208 4.232 4.475
0.8 2.757 2.791 3.320 3.940 3.959 4.187
0.9 2.604 2.633 3.131 3.718 3.733 3.948
1.0 2.475 2.499 2.972 3.529 3.542 3.745

AS (SA) modes
0.5 2.448 3.739 4.313 4.751 5.086 5.588
0.6 2.243 3.415 3.941 4.338 4.644 5.105
0.7 2.082 3.162 3.652 4.017 4.300 4.728
0.8 1.951 2.959 3.418 3.759 4.022 4.424
0.9 1.842 2.792 3.224 3.544 3.792 4.172
1.0 1.750 2.647 3.060 3.363 3.598 3.958

AA modes
0.5 2.973 4.448 4.457 5.162 5.657 5.660
0.6 2.715 4.062 4.069 4.712 5.165 5.167
0.7 2.515 3.761 3.767 4.363 4.782 4.783
0.8 2.353 3.519 3.524 4.081 4.473 4.474
0.9 2.219 3.318 3.322 3.848 4.218 4.219
1.0 2.105 3.149 3.152 3.650 4.001 4.002

rigid tank with the increase of the liquid depth and the bifurcating is more remarkable for
the fundamental duplicate sloshing eigenfrequencies. Such a case can also be found in the AA
modes.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 171

30
SS
AS(SA)
25 AA

20

Ω1
15

10

5
0.5 0.6 0.7 0.8 0.9 1
σ

Figure 6. The effect of rigidity parameter  on the fundamental bulging eigenfrequencies.

Table VIII. The effect of the liquid–tank mass ratio on the


first six bulging eigenfrequencies.
1 2 3 4 5 6
SS modes
5.0 7.344 7.428 18.879 19.516 40.651 44.127
6.0 6.820 6.958 17.730 18.424 38.423 41.799
7.0 6.381 6.546 16.764 17.502 36.551 39.827
8.0 6.017 6.343 15.948 16.711 34.939 38.124
9.0 5.736 6.221 15.240 16.020 33.533 36.632
10.0 5.467 5.861 14.624 15.411 32.288 35.309

AS (SA) modes
5.0 8.921 19.090 26.849 38.879 44.988 58.150
6.0 8.333 17.952 25.102 36.581 42.578 54.366
7.0 7.869 17.003 23.659 34.662 40.525 51.236
8.0 7.458 16.194 22.440 33.024 38.749 48.591
9.0 7.110 15.493 21.392 31.605 37.191 46.318
10.0 6.771 14.878 20.480 30.358 35.811 44.337

AA modes
5.0 24.778 27.584 39.413 43.686 66.699 69.408
6.0 23.276 25.537 37.258 40.810 63.387 65.756
7.0 22.016 23.888 35.422 38.499 60.543 62.634
8.0 20.941 22.522 33.833 36.537 58.062 59.920
9.0 20.009 21.367 32.440 34.846 55.870 57.531
10.0 19.191 20.374 31.206 33.370 53.914 55.406

The effect of the rigidity parameter  on the first six sloshing eigenfrequencies is given in
Table VII where the ratio of liquid-depth to tank-height is taken as H/c = 0.95. It is shown
that the sloshing eigenfrequencies increase with the decrease of rigidity parameter . The ef-
fect of rigidity parameter  on the fundamental bulging eigenfrequencies is given in Figure 6.
It is shown that the bulging eigenfrequencies decrease slowly with the increase of rigidity

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
172 D. ZHOU AND W. LIU

2.5
SS
SA(AS)
2.3 AA

Ω1
2.1

1.9

1.7
5 6 7 8 9 10
τ

Figure 7. The effect of liquid–tank mass ratio on the fundamental sloshing eigenfrequencies.


parameter . It should be mentioned that the actual bulging frequency ( = (D/ p hc4 )) should
increase with the increase of the tank rigidity D because the dimensionless bulging eigenfrequency
has been taken in the analysis. It can be seen from Figure 6 and Table VII that the increase of the
sloshing eigenfrequencies with the decrease of rigidity parameter  is much quicker than that of
the bulging eigenfrequencies. Therefore, with the decrease of the rigidity parameter the fundamen-
tal sloshing eigenfrequencies will be close to the fundamental bulging eigenfrequencies. In such
a case, a strong liquid–tank interaction occurs.
The effect of liquid–tank mass ratio on the first six bulging eigenfrequencies is given in
Table VIII where the ratio of liquid-depth to tank-height is also taken as H/c = 0.95. It is shown
that the bulging eigenfrequencies monotonically decrease with the increase of the liquid–tanks
mass ratio. The effect of liquid–tank mass ratio on the fundamental sloshing eigenfrequencies is
given in Figure 7. It is shown that the sloshing eigenfrequencies slowly decrease with the increase
of the liquid–tank mass ratio. It can be seen form Figure 7 and Table VIII that the increase of
the bulging eigenfrequencies with the increase of liquid–tank mass ratio is much quicker than
that of the sloshing eigenfrequencies. Therefore, with the decrease of liquid–tank mass ratio, the
fundamental bulging eigenfrequencies will be close to the fundamental sloshing eigenfrequencies.
In such a case, a strong liquid–tank interaction also occurs.

7. CONCLUSIONS

A semi-analytical method is developed to analyse the three-dimensional vibratory characteristics


of flexible rectangular tanks partially filled with liquid. Using a combination of the superposition
method and the method of separation of variables, the exact analytical solution of the liquid
velocity potential is derived. The coefficients in the liquid velocity potential are determined by the
contact conditions between the liquid and tank-wall, which are represented by integrals including
the tank-wall dynamic deflection. By constructing a set of reasonable admissible functions, the
Rayleigh–Ritz method is used to obtain the bulging eigenfrequency equation. A group of additional
sloshing eigenfrequency equation is provided by utilizing the Galerkin method to the surface wave

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
HYDROELASTIC VIBRATIONS OF FLEXIBLE RECTANGULAR TANKS 173

equation. The convergence study confirms the high accuracy and small computational cost of the
present method.
From the numerical examples, it is observed that the decrease of the tank-wall rigidity results
in the increase of the sloshing eigenfrequencies, while the increase of the liquid–tank mass ratio
results in the decrease of the bulging eigenfrequencies. Therefore, either the small tank-wall rigidity
or the large liquid–tank mass ratio will result in a strong liquid–tank interaction. Moreover, it
can be seen that the bulging eigenfrequencies increase when the surface waves are taken into
account and the sloshing eigenfrequencies decrease when the tank-wall flexibility is taken into
account. If the fundamental sloshing eigenfrequencies are far away from the fundamental bulging
eigenfrequencies, the effect of the surface waves on the bulging eigenfrequencies and the effect of
the tank-wall flexibility on the sloshing eigenfrequencies can be ignored. In such a case, the studies
on the bulging modes of the tank-wall vibration and the sloshing modes of the liquid oscillation
can be separately considered without significant errors.
In the present study, the investigation is focused on the linear hydroelastic vibration of flexible
rectangular tanks. The method can be extended to the study of nonlinear vibration where the
Galerkin equation from liquid surface waves should be nonlinear.

REFERENCES
1. Nielsen R, Kiremidjian AS. Damage to oil refineries from major earthquakes. Journal of Structural Engineering
(ASCE) 1986; 112:1481–1491.
2. Hanson RD. Dynamic analyses of liquid storage tanks, in the great Alaska earthquake of 1964. National Academy
of Sciences 1973; 7.
3. Jennings PC. Engineering Features of the San Fernando Earthquake. EEERI-71-02, 434-470, California Institute
of Technology: Pasadena, 1971.
4. Zienkiewicz OC, Bettess P. Fluid-structure dynamic interaction and wave forces, an introduction to numerical
treatment. International Journal for Numerical Methods in Engineering 1978; 13:1–16.
5. Marcus MS. A finite-element method applied to the vibration of submerged plates. Journal of Ship Research
1978; 22:94–99.
6. Jablonski AM, Humar JL. Three-dimensional boundary element reservoir model for seismic analysis of aech and
gravity dams. Earthquake Engineering and Structural Dynamics 1990; 19:359–376.
7. Wolf DH, Bachman H. Hydroelastic-stiffness matrix based on boundary elements for time-domain dam-reservoir-
soil analysis. Earthquake Engineering and Structural Dynamics 1988; 16:417–432.
8. Humar JL, Jablonski AM. Boundary element reservoir model for seismic analysis of gravity dams. Earthquake
Engineering and Structural Dynamics 1988; 16:1129–1156.
9. Nagaya K, Takeuchi J. Vibration of a plate with arbitrary shape in contact with a fluid. Journal of the Acoustical
Society of America 1984; 75:1511–1518.
10. Kwak MK. Hydroelastic vibration of rectangular plates. Journal of Applied Mechanics (ASME) 1996; 63:110–115.
11. Soedel SM, Soedel W. On the free and forced vibration of a plate supporting a freely sloshing surface liquid.
Journal of Sound and Vibration 1994; 171:159–171.
12. Zhou D, Cheung YK. Vibration of vertical rectangular plate in contact with water on one side. Earthquake
Engineering and Structural Dynamics 2000; 29:693–710.
13. Cheung YK, Zhou D. Coupled vibratory characteristics of a rectangular container bottom plate. Journal of Fluids
and Structures 2000; 14:339–357.
14. Ambili M, Paidoussis MP, Lakis AA. Vibration of partially filled cylindrical tanks with ring-stiffeners and flexible
bottom. Journal of Sound and Vibration 1998; 213:259–299.
15. Amabili M. Ritz method and substructuring in the study of vibration with strong fluid-structure interaction.
Journal of Fluids and Structures 1997; 11:507–523.
16. Rashed AA, Iwan WD. Dynamic analysis of short length gravity dams. Journal of Engineering Mechanics
(ASCE) 1985; 111:1067–1083.
17. Rammerstorfer FG, Sharf K, Fisher FD. Storage tanks under earthquake loading. Applied Mechanics Review
(ASME) 1990; 43:261–282.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme
174 D. ZHOU AND W. LIU

18. Veletsos AS, Tang Y. Dynamics of vertically excited liquid storage tanks. Journal of Structural Engineering
(ASCE) 1985; 112:1228–1246.
19. Haroun MA, Housner GW. Earthquake response of deformable liquid storage tanks. Journal of Applied Mechanics
(ASME) 1981; 48:411–418.
20. Chiba M. Free vibration of a partially liquid-filled and partially submerged, clamped-free circular cylindrical
shell. Journal of the Acoustical Society of America 1996; 100:2170–2180.
21. Housner GW. Dynamic pressure on accelerated fluid containers. Bulletin of Seismic Society of America 1957;
47:15–35.
22. Housner GW. The dynamic behavior of water tanks. Bulletin of Seismic Society of America 1963; 53:381–389.
23. Haroun MA. Stress analysis of rectangular walls under seismically induced hydrodynamic loads. Bulletin of
Seismic Society of America 1984; 74:1031–1041.
24. Bauer HF. Hydroelastic vibrations in a rectangular container. International Journal of Solids and Structures 1981;
17:639–652.
25. Dogangun A, Durmus A, Ayvaz Y. Static and dynamic analysis of rectangular tanks by using the Lagrangian
fluid finite element. Computers and Structures 1996; 59:547–552.
26. Kim JK, Koh HM, Kwahk IJ. Dynamic response of rectangular flexible fluid containers. Journal of Engineering
Mechanics 1996; 122:807–817.
27. Koh HM, Kim JK, Park JH. Fluid-structure interaction analysis of 3-D rectangular tanks by a variationally
coupled BEM-FEM and comparison with test results. Earthquake Engineering and Structural Dynamics 1998;
27:109–124.
28. Zhou D, Cheung YK. Free vibration of line supported rectangular plates using a set of static beam functions.
Journal of Sound and Vibration 1999; 223:231–245.
29. Zhou D. Natural frequencies of rectangular plates using a set of static beam functions in Rayleigh–Ritz method.
Journal of Sound and Vibration 1996; 189:81–87.
30. Lamb H. Hydrodynamics. Dover: New York, 1945.
31. Leissa AW. The free vibration of rectangular plates. Journal of Sound and Vibration 1973; 31:257–293.
32. Kim CS, Dickinson SM. The flexural vibration of line supported rectangular plate system. Journal of Sound and
Vibration 1987; 114:129–142.

Copyright q 2006 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2007; 71:149–174
DOI: 10.1002/nme

You might also like