Anda di halaman 1dari 8

M.

Utsumi
Department of Machine Element,
Technical Research Laboratory,
IHI Corporation,
1 Shinnakaharacho, Isogo-ku,
Yokohama, Kanagawa Prefecture 235-8501,
Japan

K. Ishida
Energy and Plant,
IHI Corporation,
1-1 Toyosu 3-chome, Koto-ku,
Tokyo 135-8710, Japan

Vibration Analysis of a Floating


Roof Subjected to Radial Second
Mode of Sloshing
In a previous paper, a cost-efficient modal analysis method for the vibration of a floating
roof coupled with nonlinear sloshing in a circular cylindrical oil storage tank is presented. This method is extended to the case in which the out-of-plane deformation of the
roof-deck caused by the radial second mode of sloshing induces an elliptical deformation
of the pontoon around the deck. First, the radial contraction of the deck is calculated
from the slope of the out-of-plane deformation of the deck, and the following two points
are confirmed: (i) the circumferential variation in this radial contraction results in the
elliptical deformation of the pontoon, and (ii) the present theoretical prediction for the
radial contraction is in good agreement with a numerical result obtained by LS-DYNA.
Based on these points, the stresses arising in the pontoon are calculated by considering
the contraction of the deck as an enforced displacement of the pontoon. Numerical results
show that (a) the elliptical deformation of the pontoon causes a large circumferential
in-plane stress, (b) reduction achieved by the increase in the thickness of the deck is
larger for the radial contraction of the deck and the stresses in the pontoon than for the
out-of-plane deformation of the deck, and (c) the radial contraction of the deck for a fixed
value of the out-of-plane deformation of the deck increases with the decrease in the
radius of the deck. DOI: 10.1115/1.3148083

Introduction

The vibration of a floating roof due to sloshing in a circular


cylindrical tank is a subject of great importance in estimating the
safety of large oil storage tanks 1,2. The oscillatory motion of a
floating roof coupled with liquid sloshing is near resonance with
low-frequency components of earthquake ground motions, and
thus can lead to serious accidents such as flood, fire, and structural
failure. Although extensive studies have been carried out on the
nonlinear sloshing problem with a free liquid surface 310, the
vibration analysis of a floating roof has been conducted using the
linearized theory under the assumption of a small amplitude sloshing 1117. However, a floating roof oscillates with finite amplitude due to the resonance behavior and thus the nonlinearity of
sloshing should be considered in the vibration analysis. In previous papers 18,19, the vibration of a floating roof subjected to
nonlinear sloshing was analyzed, and it was shown that neglecting
the nonlinearity of sloshing significantly underestimates the
stresses arising in the floating roof, even when the nonlinear effect
is small for the vertical displacement of the floating roof. This
underestimation becomes more marked when internal resonance
due to the nonlinearity of sloshing is present 19. These previous
works focused on the case in which the radial first mode of sloshing with circumferential wave number 1 is excited. However, the
radial second mode of sloshing as well as the first mode is near
resonance with low-frequency components of earthquake ground
motions. The second mode causes the out-of-plane deformation
and radial contraction of the deck 20. Because this contraction
has modal components with circumferential wave numbers 0 and
2, it causes an elliptical deformation of the pontoon around the
deck. Therefore, it is necessary to predict the stresses caused by
this elliptical deformation for estimating the safety of the floating

roof. In this paper, a method of estimating the stresses is presented


by extending the previously reported nonlinear analysis method
incorporating the radial contraction of the deck.

Analysis

2.1 Computational Model. The system to be considered is


shown in Fig. 1. The floating roof consisting of deck, pontoon,
and stiffeners is modeled as an axisymmetric elastic shell. In Fig.
1, a is the radius of the tank and h is the liquid-filling level. The
detailed parameters for the floating roof geometry are given in the
section that is dedicated to the description of the numerical examples. The analysis is performed under the assumption that the
liquid motion is inviscid, incompressible, and irrotational and that
the wall and bottom of the tank are rigid. The nonlinearity of the
boundary conditions at the interface between the liquid and the
floating roof is considered. The static position of the interface is
considered a plane expressed by z = h and 0 r a in formulating
the nonlinear boundary conditions because the variation in the z
coordinate of the static position is very small compared with the
liquid-filling level and the difference between the radii of the tank
wall and the floating roof is much smaller than the tank radius.
2.2 Variational Form of Governing Equations. To apply a
computationally efficient Galerkin method to the problem, a variational form of the governing equations is presented in this section.
From the equality of the Lagrangian density with the liquid pressure pl 21, the Lagrangian Ll of the liquid can be expressed as

Ll =
Contributed by the Pressure Vessel and Piping Division of ASME for publication
in the JOURNAL OF PRESSURE VESSEL TECHNOLOGY. Manuscript received August 27,
2008; final manuscript received February 22, 2009; published online March 30, 2010.
Review conducted by Spyros A. Karamanos.

Journal of Pressure Vessel Technology

pldV

where V is the liquid domain. The liquid pressure pl is given by

Copyright 2010 by ASME

APRIL 2010, Vol. 132 / 021303-1

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

ferential wave number of the displacement; Xmx


and Xmy are displacement components excited by the x and y components of the
excitation, respectively; Mm
and Km are the mass and stiffness
matrices that are common for Xmx
and Xmy; le is the length of the
generatrix of each shell element; s is the local coordinate defined
along the generatrix of each shell element; and r and hr are the
density and thickness of the element. Because the variations in the
velocity potential, the floating roof displacement, and the arbitrary
time-dependent function are arbitrary and independent of one another, we obtain the system of governing equations. The first
through third terms of Eq. 3 yield the condition of continuity in
the liquid domain and the boundary conditions on the liquid-tank
interface as follows:

Fig. 1 Computational model

2 = 0

pl = f
+ gz h + r cos f xt + r sin f yt
t
1
t
+ 2 + G
2

where f is the liquid density, is the velocity potential which


describes the liquid motion relative to the moving tank, g is the
gravitational acceleration, f xt and f yt are the earthquake acceleration inputs in the x and y directions, respectively, and Gt is an
arbitrary time-dependent function, which arises due to the spatial
integration of the equation of motion for the liquid. This function
is determined from the condition that the mean value of the residual for the equation of motion of the floating roof over the
liquid-roof interface vanishes. By adding the Lagrangian of the
floating roof determined by the finite element method and applying the mathematical procedures explained in the previous papers
18,22 to the calculus of variations, the variational form of the
governing equations can be obtained. In this paper, the linearized
variational form is presented by neglecting the nonlinear terms
and evaluating the liquid pressure and the kinematic condition on
the undisturbed position of the liquid-roof interface because the
effect of the nonlinearity of sloshing on the stresses caused by the
radial contraction of the deck is small as is shown in the computational results. The linear formulation simplifies the analysis. The
linearized variational principle can be written as follows:

2dV f

+ f

r=a

t
mx

mx

z=h

elem

f G

u
rdrd = 0
t

mnxtcos

m + Amnytsin m

coshmnz
coshmnh

Empxt,
Tmkp

t =
Xmyk

mkpEmpy t

p=1

where Tmkp
is the kth component of the pth eigenvector obtained
by solving the eigenvalue problem 2Mm
+ Km = 0, while
Empxt and Empyt are the modal coordinates. In terms of these
modal coordinates, the floating roof displacement at an arbitrary
position s in each element can be expressed as

rhr f xtcos

s
Empxtcos m + Empytsin mU
mp

s
vs, ,t =
Empxtsin m Empytcos mV
mp
m=0 p=1
s, ,t
w

Empxtcos m + Empytsin mWmps

are the displacement components of the floating


where u, v, and w
roof in the z, , and r directions, respectively; m is the circum021303-2 / Vol. 132, APRIL 2010

m=0 n=1

us, ,t

le

+ f xtsin f ytcos vrdsd


f ytsin w
2

r, ,z,t =

p=1

+ r cos f xt
gu

+ r sin f yt urdrd +

2.3 Differential Equations. From Eqs. 4 and 5, we express the velocity potential as follows:

t =
Xmxk

Xmx
Xmy
t Mm X my
+ Km Xmy

+ Km

z=h

=0
z=0

z=hrdrd

m=0

where Amnx and Amny are the generalized coordinates, Jm is the


mth order Bessel function of the first kind, mn is the nth positive
root of Jm
a = 0.
The kth components of Xmx
and Xmy can be expressed as

z=0rdrd

X M X

In a similar manner, the fourth term of Eq. 3 represents the


condition that on the disturbed interface between the liquid and
the floating roof, their normal velocity components are equal to
each other; the fifth to seventh terms of Eq. 3 lead to the equation of motion for the floating roof subjected to the liquid pressure
and the inertial force due to the excitation; and the last term of Eq.
3 yields the volume constant condition. Because the damping
effect is not considered in Eq. 3, modal damping terms are introduced into the ordinary differential equations to be derived
later.
Note that the terms with Mm
and Km are not the finite element
model equations for the nodal displacements Xmx and Xmy, but
modified equations with much smaller dimension size. This modification is explained in the previous work 18 and is helpful for
reducing computation time and cost.

r=aadzd

r=a

Jmmnr

z=0

u
+
t
z

= 0,

s, V s, and W
s are the modal functions of the
where U
mp
mp
mp
displacement components in the z, , and r directions,
respectively.
Substituting Eqs. 68 into the variational principle 3 leads
Transactions of the ASME

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

to ordinary differential equations for the generalized coordinates.


These equations can be expressed in the following matrix form.
M1x 1 + K1x1 = Fft
Mmx m + Kmxm = 0

9a

m = 0,2

9b

element in its generatrix direction s, circumferential direction ,


and normal direction , respectively. The positive directions of
these local coordinates s and are defined such that cosz , s is
negative and cosr , is positive.
The strain energy of the pontoon can be expressed as

where xm is the collection of Amnx, Amny, Empx, and Empy n , p


= 1 , 2 , . . . for each value of m and ft is the tank excitation vector
given by f xt , f ytt. The coefficients Mm, Km, and F are presented in the previous paper 18.
2.4 Stress Analysis for Elliptical Deformation of Pontoon.
The eigenfrequencies of the linearized system can be determined
by solving the eigenvalue problem 2Mm + Km = 0 for each
value of m, as can be seen from Eq. 9. For lower modes, the
eigenfrequencies are close to gmn tanhmnh1/2, which are the
eigenfrequencies of sloshing without the floating roof. Therefore,
the out-of-plane deformation of the deck due to the radial second
mode of sloshing with circumferential wave number1 is expressed
as follows from Eq. 8.
u = E12xtcos + E12ytsin F12r

10

s
where the modal function F12r can be determined from U
12
by transforming the local coordinate s into the global coordinate r.
This mode of the roof has one nodal circle so it is strongly excited
when the liquid oscillates in the radial second mode of sloshing.
The radial contraction of the deck due to this out-of-plane deformation can be calculated as
S,t =

b1

U=


le

16

hr/2 j=1 j 1=1

hr/2 16

jj 1u j u j 1drdsd



u1

u4
u

u5 =
v ,
s
u6
w

u2 = v ,
u3
w

u11

u7
u

u8 =
v ,

u9
w

u14

u12 =
v ,
s
u13
w
2

u10 =

2
E12y
t

1/4E212xt

+ 1/2E12xtE12ytsin 2

2
E12y
tcos

2
11

where b1 is the radius of the deck. To facilitate the discussion for


Eq. 11, let us consider the case in which the tank excitation is in
the x direction and thus E12y is not excited. In this case, the coefficients of cos 0 and cos 2 are identical and the coefficient of
sin 2 vanishes. This indicates that the pontoon around the deck
undergoes elliptical deformation. The equality between the coefficients of cos 0 and cos 2 meets the Japanese Fire Service
Law, which requires the modal components with circumferential
wave numbers 0 and 2 be considered in the earthquake-proof design of the pontoon. Furthermore, the radial contraction of the
deck determined from Eq. 11 is in good agreement with a numerical result obtained by LS-DYNA 2, as is shown in Sec. 3.
Based on these points, the contraction of the deck given by Eq.
11 is considered as an enforced displacement of the pontoon for
an approximate estimation of the stress caused by the elliptical
deformation of the pontoon.
This paper addresses the case in which the continuity condition
of the radial displacement between the deck and the pontoon is
satisfied. That is, the presented procedure does not include the
case in which the edge of the deck moves away from the pontoon
or bumps each other.
In the subsequent analysis, the change in the variation in the
strain energy due to the enforced displacement is transformed into
nonlinear functions of the generalized coordinates. These nonlinear functions are added to the right-hand sides of Eq. 9 because
Eq. 9 is derived from Hamiltons principle.
The enforced displacement components of the pontoon are
u = S cosr,s,

v = 0,

w = S cosr,

12

where cosa , b is the cosine of the angle between a and b directions and u, v, and w are the displacement components of the shell
Journal of Pressure Vessel Technology

14

The displacement components u, v, and w can be expressed in


terms of the generalized coordinates as follows:


us, ,t

Empxtcos m + Empytsin mUmps

vs, ,t

Empxtsin m Empytcos mVmps

m=0 p=1

Empxtcos m + Empytsin mWmps

15

From Eq. 13, we obtain

b1

dF12/dr2dr

2w
s2

u15 = 2 v

u16
w
2

1/4E212xt

13

where is the coordinate measured along the normal of the midplane of the shell, B jj1 are the coefficients that appear in the relations between the stress-strain and strain-displacement components, and u j are the displacement components defined as

ws, ,t

1/2u/ r2dr

elem

U =

elem

le

16

hr/2 j=1 j 1=1

hr/2 16

jj 1u j u j 1

+ u ju j1drdsd
16

We transform Eq. 16 into

U =

elem

le

16

hr/2 j=1 j 1=1

hr/2 16

jj 1u j u j 1drdsd

17

where C jj1 = B jj1 + B j1 j. By virtue of this transformation, the number of terms including variation can be decreased, thereby reducing the hand calculation effort. Using Eq. 17, the change in the
variation in the strain energy can be expressed as
U =

elem

le

hr/2 16

16

hr/2 j=1 j 1=1

jj 1u j u j 1

+ u ju j1

drdsd

18

From Eq. 12, we evaluate the changes in the displacement components defined by Eq. 14. Noting that the direction cosine values are constant within each element, we obtain

u1

u2 =

u3

S cosr,s
0

u4

S cosr,

u5 = 0 ,
0

u6

u7

u8

u9

S/cosr,s
0

u10 = 0

S/cosr,

APRIL 2010, Vol. 132 / 021303-3

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm


u11

u14

u12 = 0 ,
0

2S/2cosr,s

u15 =

u13

u16

0
2S/2cosr,

19

If the concise expression 13 is not used, the strain energy U is


given by a lengthy equation. Therefore, heavy hand calculation
effort is required for expanding U into many terms with u j,
u j, u j, and u j and expressing them in terms of the generalized coordinates through the substitution of Eqs. 11 and 15.
To overcome this difficulty, thereby presenting an efficient formulation, we introduced the concise expressions 13 and 18 using
the index j. To further enhance the efficient formulation, we express u j given by Eqs. 14 and 15 in the following form:
uj =

E
m

mptmj Umjps

j = 1 16 20

=x,y

by appropriately defining the modal functions mj and


s; for example = cos m, = sin m, and
U
mjp
m1x
m1y
s = U s. By using these circumferential modal functions,
U
m1p
mp
we can express Eq. 11 as
S,t =

Hmtm1

m=0,2 =x,y

b1

dF12/dr2dr

Radius of tank a
41.7 m
Liquid-filling level h
20.3 m
Liquid density f
887 kg/ m3
Radius of floating roof b
41.42 m
Radius of deck b1
36.08 m
Distances among compartments b2 = b3 = b4
1.78 m
Distance between deck and upper end of inner rim H1
0.272 m
Distance between deck and lower end of inner rim H2
0.272 m
Height of outer rim H
0.918 m
Slope tan1dz / dr of deck
0.002 deg
Slope tan1dz / dr of top of pontoon
4 deg
Slope tan1dz / dr of bottom of pontoon
0.002 deg
Thickness deck
0.0045 m
Thickness pontoon except inner rim
0.006 m
Thickness inner rim of pontoon
0.020 m
Radial coordinates of stiffeners
5.5+ 6i i = 0 4
Height and breadth of stiffeners
0.2 m, 0.4 m
Thickness of stiffeners
0.0045 m
Density of floating roof
7850 kg/ m3
Youngs modulus of floating roof
2.1 1011 N / m2
Poissons ratio of floating roof
0.3

21

where
2
H0xt = 1/4E212xt + E12y
t,

H0yt = 0

22
H2yt = 1/2E12xtE12yt

2
t,
H2xt = 1/4E212xt E12y

The contraction S of the deck does not contain the component


with circumferential wave number 1. Using Eq. 21, the changes
in the displacement components given by Eq. 19 can be expressed in the following form:
u j1 =

Table 1 Parameters of numerical example 120,000 m3 oil


tank

m1=0,2 1=x,y

Hm11tm1 j11U
j1

23

tion has circumferential wave number 1, the nonlinear terms of


the form E12xEm1p11 O3 are added to the appropriate components of the right-hand side of Eq. 9a.
On the other hand, B of Eq. 25 leads to the time-dependent
factors EmpHm11 with m = m1 = 0 and m = m1 = 2. From Eq. 22,
these factors have the form E212xEmp m = 0 and 2. Therefore,
the nonlinear terms proportional to E212x O2 are added to the
appropriate components of the right-hand side of Eq. 9b.
Thus, the second term of the brace in Eq. 25 is predominant to
the first term and describes the elliptical deformation of the pontoon caused by the radial contraction of the deck with modal
components of circumferential wave numbers 0 and 2.

are defined, for example, as


where U
j1
= cosr,s
U
1

b1

dF12/dr2dr

24

Substituting Eqs. 20 and 23 into Eq. 18 leads to


U =

elem

le

hr/2 16

16

hr/2 j=1 j 1=1

jj 1A

+ Bdrdsd
25

where
A=

m=0,2 =x,y

mt


m1

p1

1=x,y

Em1p11t

mjm1 j11 U
j m1 j 1 p1s
B=

E
m

=x,y

mpt

m1=0,2 1=x,y

26

Numerical Examples

Numerical calculation was conducted for a 120,000 m3 oil


storage tank. Parameters of this tank are presented in Table 1. The
detailed geometry of the floating roof used for the numerical example is shown in Fig. 2. The damping term z=h is added to
the pressure term of Eq. 3 and the constant is determined as
2
= 2mnmn mn = 0.01, mn
= gmn tanhmnh for each modal
component of given by Eq. 6. Furthermore, for the radial
displacement of the outer rim of the pontoon relative to the tank
wall, the spring and damping support constants per unit area
86,000 N / m3 and 5000 N s / m3 were taken into account. The
excitations were given as f xt = 0.24 sin12t 0 t 3T, f xt
= 0 3T t, and f t = 0 0 t, where T is the period of the
y

Hm11t

sU
mjm1 j11 U
mjp
j1

27

The factors given by Eqs. 26 and 27 respectively correspond to


the first and second terms in the bracket on the right-hand side of
Eq. 18.
From A in Eq. 25, the terms with m = m1 = 0 and m = m1 = 2
arise due to the orthogonality condition of the circumferential
modal functions. Hence, the arising time-dependent factors
HmEm1p11 have the form E12xE12xEm1p11, as can be seen from
Eq. 22. Because the generalized coordinate subjected to varia021303-4 / Vol. 132, APRIL 2010

Fig. 2 Geometry of floating roof used for numerical example

Transactions of the ASME

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 5 Responses of radial displacement of joint between deck


and pontoon and radial contraction of deck thin solid line, radial displacement at = 0 deg; dotted line, radial displacement
at = 180 deg; thick solid line, radial contraction at = 0 deg

Fig. 3 Radial contraction and out-of-plane deformation of


deck due to radial second mode of sloshing with circumferential wave number 1: a radial contraction at = 0 deg and b
Out-of-plane deformation at r = 18.9 m and = 0 deg

excitation which can be calculated as T = 2 / 12


= 2 / g12 tanh12h1/2 = 5.65 s using 12a = 5.33, a = 41.7 m,
and h = 20.3 m.
Figure 3 shows the radial contraction and out-of-plane deformation of the deck due to the radial second mode of sloshing with
circumferential wave number 1. The out-of-plane deformation is
evaluated at the position where the deformation is maximum. It
can be seen that the radial contraction shown in Fig. 3a varies
in-phase with the absolute value of the out-of-plane deformation
presented in Fig. 3b. The responses shown in Figs. 3a and 3b
reach their maximum values at t = 19.5 s after the sinusoidal excitation ends at t = 17 s. The amplitude variation in the radial contraction shown in Fig. 3a is larger than that of the out-of-plane
deformation presented in Fig. 3b because in Eq. 11 u / r is
raised to the second power.
The analysis does not include the effect of the extensional rigidity of the deck on reducing the out-of-plane deformation of the
deck because the radial contraction of the deck evaluated from the
modal function of the deck is considered as an enforced displacement of the pontoon. To check the accuracy of the relation between the radial contraction and the out-of-plane deformation of
the deck, their values that arose during the time response analysis
are plotted in Fig. 4. For the sake of comparison, the relation
predicted by a program LS-DYNA 2 is shown in Fig. 4. It can be
confirmed that the results obtained by the present analysis and
LS-DYNA are in good agreement.
Figure 5 shows the responses of the radial displacement of the
joint between the deck and the pontoon at = 0 and = , together
with the response of the radial contraction of the deck at = 0.
Within the initial time interval, during which the system under-

Fig. 4 Relation between radial contraction and out-of-plane


deformation of deck shown in Figs. 3a and 3b; , result cited
from Fig. 2.6 in Ref. 2

Journal of Pressure Vessel Technology

goes the sinusoidal excitation, the responses of the radial displacement at = 0 and = are out-of-phase to each other. The reason
for this can be explained as follows. The present analysis allows
for the radial displacement restricted by the spring and damper
supports of the seal between the pontoon and the tank wall.
Hence, the translational motion in the x direction with circumferential variation cos prevails over the radial contraction of the
deck shown by the thick line in Fig. 5. Thus, the responses of the
radial displacement at = 0 and = are out-of-phase to each
other within the initial time interval.
After the sinusoidal excitation ends at t = 17 s, the out-of-phase
relation rapidly decays due to heavy damping of the seal and thus
the radial displacement is mainly contributed by the radial contraction of the deck, whose circumferential variation is represented by cos 0 and cos 2 Eq. 11. Therefore, for times
larger than, say, t = 20 s, the responses of the radial displacement
at = 0 and = become in-phase to each other and out-of-phase
with the radial contraction of the deck at = 0, as shown in Fig. 5.
The decaying of the responses contributed by the radial contraction is slow due to the light damping property of sloshing.
We should note that from the responses of the radial displacement at = 0 and = , the radial contraction with cos 0 and
cos 2 components seems to appear after the excitation ends. This
is because while the excitation is applied, the radial contraction is
difficult to observe due to the predominant translational motion
with cos component. The radial contraction appears over the
interval 0 t, as is shown in the thick line in Fig. 5, in proportion
to the square of the out-of-plane deformation shown in Fig. 3b.
Figure 6 shows the circumferential variation in the radial displacement of the joint between the deck and the pontoon at t
= 31 s, at which the contraction of the deck reaches its local maximum. The initially excited mode with circumferential variation
cos is superposed. However, it can be confirmed that the modal
components represented by cos 0 and cos 2 with the same
magnitude are predominant the magnitude is found from the
mean and amplitude of the circumferential variation. This result
is based on the fact that the coefficients of cos 0 and cos 2 are
identical in Eq. 11.
Figure 7 illustrates the responses of the in-plane stresses that
arise in the inner rim at the joint between the deck and the pon-

Fig. 6 Circumferential variation in radial displacement at joint


between deck and pontoon t = 31 s

APRIL 2010, Vol. 132 / 021303-5

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 7 Responses of in-plane stresses arising in inner rim at


joint between deck and pontoon = 0 deg: a stress in the
vertical direction acting on the horizontal cross section of the
inner rim and b circumferential stress

toon. The circumferential position is = 0. The in-plane stresses


can be calculated as the mean of the stress values at = 0.5hr and
= 0.5hr, where hr is the thickness and the positive direction of
measured along the normal of the midplane of the shell is +r. It
can be seen from Fig. 7 that the radial contraction of the deck
results in very large circumferential in-plane stress. This numerical result is useful to explain the past structural damage due to the
elliptical deformation of the pontoon.
Conventional analysis using the program LS-DYNA is very timeconsuming and expensive, so that accurate computation is more
difficult for the stress than for the deformation. To avoid this
problem, the stress was evaluated by approximating the pontoon
shell as a curved beam subjected to the deformation computed by
LS-DYNA and evaluating the resultant moment exerted on the cross
section of the beam. The stress value predicted by these procedures also attains the order of several hundred megapascals 1,2.
The present modal analysis method requires a small amount of
computation time and cost to predict such a high stress and to
determine the thickness of the deck required for avoiding the
structural damage of the pontoon. The analysis using LS-DYNA
requires several days while the present method requires only a few
seconds.
Figure 8 shows the circumferential variations in the in-plane
stresses at t = 31 s, at which the stresses reach their local maximum. The absolute values of cos 0- and cos 2-modes are identical due to the fact that these modes are equally contained in the
radial contraction given by Eq. 11. In the stress variations illus-

Fig. 8 Circumferential variations in in-plane stresses arising


in inner rim at joint between deck and pontoon t = 31 s; thin
line, stress in the vertical direction acting on the horizontal
cross section of the inner rim; thick line, circumferential stress

021303-6 / Vol. 132, APRIL 2010

Fig. 9 Radial contraction and out-of-plane deformation of


deck due to radial second mode of sloshing with circumferential wave number 1 the case in which thickness of deck is
increased to 0.08 m: a radial contraction at = 0 deg and b
out-of-plane deformation at r = 15.0 m and = 0 deg

trated in Fig. 8, cos -mode is not included unlike the displacement variation shown in Fig. 6. This is because cos -mode of the
displacement is approximately a rigid-body mode and thus its
contributions to the stresses are very small.
In this example, the difference between the stresses evaluated
by the linear and nonlinear sloshing analyses can be neglected, in
contrast to the case in which the radial first mode of sloshing is
mainly excited 18,19. Thus, the radial contraction of the deck is
more important than the nonlinearity of sloshing.
Figures 9 and 10 show results for the case in which the thickness of the deck is increased to 0.08 m. These results are compared with Figs. 3 and 7, respectively. The increase in the thickness of the deck reduces the out-of-plane deformation of the deck,
radial contraction of the deck, and the stresses in the inner rim.
Note that the reduction is larger for the radial contraction of the

Fig. 10 Responses of in-plane stresses arising in inner rim at


joint between deck and pontoon = 0 deg; the case in which
thickness of deck is increased to 0.08 m: a stress in the vertical direction acting on the horizontal cross section of the inner rim and b circumferential stress

Transactions of the ASME

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Table 2 Parameters of numerical example presented in Fig. 12


30,000 m3 oil tank

Fig. 11 Equivalent thickness hr corresponding to distance Hr


between upper and lower plates

deck and the stresses than for the out-of-plane deformation of the
deck. This is because in Eq. 11, u / r is raised to the second
power. However, the radial contraction and therefore the stresses
do not necessarily decrease in proportion to the square of the
out-of-plane deformation of the deck because the mode function
of u varies with the thickness of the deck.
The flexural rigidity Dr of the roof-deck with thickness hr is
h3. When the deck is formed by two plates
proportional to D
r1
r

with the original thickness 0.0045 m, Dr is proportional to D


r2
Hr + 0.0045 23 Hr3, where Hr is the distance between the
plates. The relation between hr and Hr determined from the con =D
is illustrated in Fig. 11. The reduction in the outdition D
r1
r2
of-plane deformation obtained by the increase in the thickness hr
from 0.0045 m to 0.08 m can be achieved by two plates with the
original small thickness 0.0045 m, when the distance Hr is 0.13 m.
Thus, the use of double-deck-type floating roofs is effective for
avoiding the structural failure caused by the radial contraction of
the deck due to the radial second mode of sloshing.
Figure 12 shows the relation between the radial contraction and
the out-of-plane deformation of the deck for a floating roof with a
smaller radius. The parameters for this case are presented in Table
2. The radius of deck is 18.9 m and the out-of-plane deformation
is evaluated at r = 10.2 m and = 0. It can be seen from Fig. 12
that the result obtained by the present analysis agrees with a past
simulation result 2. Comparing the results shown in Figs. 4 and
12 shows that when the radius of the deck decreases, the radial
contraction of the deck for a fixed value of the out-of-plane de-

Fig. 12 Relation between radial contraction and out-of-plane


deformation of deck the case in which radius of deck is decreased; , result cited from Fig. 2.6 of Ref. 2

Journal of Pressure Vessel Technology

Radius of tank a
Liquid-filling level h
Liquid density f
Radius of floating roof b
Radius of deck b1
Compartment in pontoon is not present
Distance between deck and upper end of inner rim H1
Distance between deck and lower end of inner rim H2
Height of outer rim H
Slope tan1dz / dr of deck
Slope tan1dz / dr of top of pontoon
Slope tan1dz / dr of bottom of pontoon
Thickness deck
Thickness top and bottom of pontoon
Thickness outer rim of pontoon
Thickness inner rim of pontoon
Radial coordinates of stiffeners
Height and breadth of stiffeners
Thickness of stiffeners
Density of floating roof
Youngs modulus of floating roof
Poissons ratio of floating roof

21.35 m
21.75 m
845 kg/ m3
21.2 m
18.9 m
0.335 m
0.075 m
0.71 m
0.002 deg
3 deg
4.4 deg
0.0045 m
0.0045 m
0.006 m
0.015 m
2.75+ 3i i = 0 4
0.1 m, 0.2 m
0.0045 m
7850 kg/ m3
2.1 1011 N / m2
0.3

formation increases. The reason for this can be explained as follows. Let us express the function F12r in Eq. 11 in the following form:
F12r = F12.max fr

28

where F12.max is the maximum value of the function F12r and


fr is a function of the dimensionless radial coordinate defined
by r = r / b1. Substituting Eq. 28 into the r-integral of dF12 / dr2
in Eq. 11 gives


b1

dF12
dr

2
2
dr = F12.max

df
dr

2
= F12.max

1
b1

df
dr

dr
dr
2

dr

b1dr
29

Because the integral on the right-hand side of Eq. 29 is not


strongly dependent on the radius b1 of the deck, the radial contraction of the deck for a fixed value of the out-of-plane deformation F12.max is roughly proportional to 1 / b1. Thus, the increase in
the radial contraction with decreasing the radius is due to the fact
that the radial contraction is proportional to the squared slope of
the out-of-plane deformation.

Summary and Conclusions

A cost-efficient modal analysis method for the vibration of a


floating roof subjected to nonlinear sloshing was applied to the
case in which the pontoon undergoes elliptical deformation due to
the radial contraction of the deck caused by the radial second
mode of the sloshing. It was confirmed that i the radial contraction calculated by integrating the squared slope of the out-of-plane
deformation over the deck has modal components with circumferential wave numbers 0 and 2 that result in elliptical deformation
of the pontoon around the deck, and ii the present theoretical
prediction for the radial contraction is in good agreement with a
numerical result obtained by LS-DYNA. Based on these points, the
stress due to the elliptical deformation of the pontoon was analyzed by considering the radial contraction of the deck as an enforced displacement of the pontoon. The change in the variation in
the strain energy due to this displacement was transformed into
nonlinear functions of the generalized coordinates using the
Galerkin method, and was introduced into the ordinary differential
APRIL 2010, Vol. 132 / 021303-7

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

equations which govern the sloshing and the floating roof oscillation. Numerical results showed that a due to the elliptical deformation of the pontoon, a large circumferential in-plane stress
arises; b reduction achieved by the increase in the thickness of
the deck is larger for the radial contraction of the deck and the
stresses in the pontoon than for the out-of-plane deformation of
the deck; and c the radial contraction of the deck for a fixed
value of the out-of-plane deformation of the deck increases with
the decrease in the radius of the deck because the radial contraction of the deck is proportional to the squared slope of the out-ofplane deformation of the deck.

References
1 Yamauchi, Y., Kamei, A, Zama, S., and Uchida, Y., 2006, Seismic Design of
Floating Roof of Oil Storage Tanks Under Liquid Sloshing, Sloshing and
Fluid Structure Vibration, ASME Pressure Vessels and Piping Division Conference, Paper No. PVP2006-ICPVT-11-93280.
2 Japanese Hazardous Materials Safety Techniques Association, 2006, Report
of Study to Develop Reasonable Modification Procedures to Satisfy Technical
Standard Regarding Relatively Long Period Ground Motion for Facilities Handling Hazardous Materials, in Japanese.
3 Abramson, H. N., Chu, W. H., and Dodge, F. T., 1966, The Dynamic Behavior of Liquids in Moving Containers, H. N. Abramson, ed., Report No. NASA
SP-106.
4 Ibrahim, R. A., Pilipchuk, V. N., and Ikeda, T., 2001, Recent Advances in
Liquid Sloshing Dynamics, Appl. Mech. Rev., 54, pp. 133199.
5 Bauer, H. F., Chang, S. S., and Wang, J. T. S., 1971, Nonlinear Liquid Motion
in a Longitudinally Excited Container With Elastic Bottom, AIAA J., 9, pp.
23332339.
6 Ibrahim, R. I., and El-Sayad, M. A., 1999, Simultaneous Parametric and
Internal Resonances in Systems Involving Strong Nonlinearities, J. Sound
Vib., 225, pp. 857885.
7 Ikeda, T., and Nakagawa, N., 1997, Nonlinear Vibrations of a Structure
Caused by Water Sloshing in a Rectangular Tank, J. Sound Vib., 201, pp.
2341.
8 Ikeda, T., and Nakagawa, N., 1995, Nonlinear Vibrations of a Structure
Caused by Water Sloshing in a Cylindrical Tank, Fluid Structure Interaction

021303-8 / Vol. 132, APRIL 2010

and Structure Mechanics, PVP Am. Soc. Mech. Eng., 310, pp. 6376.
9 Peterson, L. D., Crawley, E. F., and Hansman, R. J., 1989, Nonlinear Slosh
Coupled to the Dynamics of a Spacecraft, AIAA J., 27, pp. 12301240.
10 Utsumi, M., Kimura, K., and Sakata, M., 1987, The Non-Stationary Random
Vibration of an Elastic Circular Cylindrical Liquid Storage Tank in Simulated
Earthquake Excitation Straightforward Analysis of Tank Wall Deformation,
JSME Int. J., Ser. III, 30, pp. 467475.
11 Nakagawa, K., 1955, On the Vibration of an Elevated Water Tank-II, Technol. Rep. Osaka Univ., 5, pp. 317336.
12 Kondo, H., 1978, Free Vibration Analysis for Vertical Motion of a Floating
Roof, Trans. Jpn. Soc. Mech. Eng., 44, pp. 12141223.
13 Sakai, F., Nishimura, M., and Ogawa, H., 1984, Sloshing Behavior of Floating Roof Oil Storage Tanks, Comput. Struct., 19, pp. 183192.
14 Sakai, F., Inoue, R., and Hayashi, S., 2006, Fluid-Elastic Analysis and Design
of Sloshing in Floating-Roof Tanks Subjected to Earthquake Motions, Proceedings of the ASME Pressure Vessels and Piping Division Conference, Vancouver, BC, Canada, Paper No. PVP2006-ICPVT11-93622, pp. 110.
15 Matsui, T., 2007, Sloshing in a Cylindrical Liquid Storage Tank With a Floating Roof Under Seismic Excitation, ASME J. Pressure Vessel Technol., 129,
pp. 557566.
16 Matsui, T., 2007, Sloshing in a Cylindrical Liquid Storage Tank With a
Single-Deck Type Floating Roof Under Seismic Excitation, Proceedings of
the ASME Pressure Vessels and Piping Division Conference, San Antonio, TX,
Paper No. PVP2007-26249.
17 Shimizu, S., Naito, K., and Koyama, Y., 1984, A Study on Sloshing Behaviors of Floating Roof Oil Storage Tanks During Earthquake Excited by ThreeDimensional Dynamic Simulator, Ishikawajima-Harima Eng. Rev., 24, pp.
379384.
18 Utsumi, M., and Ishida, K., 2008, Vibration Analysis of a Floating Roof
Taking Into Account the Nonlinearity of Sloshing, ASME J. Appl. Mech., 75,
p. 041008.
19 Utsumi, M., Ishida, K., and Hizume, M., 2010, Internal Resonance of a Floating Roof Subjected to Nonlinear Sloshing, ASME J. Appl. Mech., 771, p.
011016.
20 Miura, M., and Kikuchi, T., 2005, The Sloshing Simulation of Floating Roof
Tank, ASME Pressure Vessels and Piping Division Conference, Paper No.
PVP2005-71439.
21 Seliger, R. L., and Whitham, G. B., 1968, Variational Principles in Continuum Mechanics, Proc. R. Soc. London, Ser. A, 305, pp. 125.
22 Utsumi, M., 1998, Low-Gravity Propellant Slosh Analysis Using Spherical
Coordinates, J. Fluids Struct., 12, pp. 5783.

Transactions of the ASME

Downloaded 04 Apr 2010 to 124.124.247.140. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Anda mungkin juga menyukai