Вы находитесь на странице: 1из 27

Journal of Constructional Steel Research 53 (2000) 307–333

www.elsevier.com/locate/jcsr

Seismic behaviour of cylindrical steel liquid


storage tanks
F.H. Hamdan
Engineering Structures Section, Department of Civil Engineering, Imperial College of Science,
Technology and Medicine, London SW7 2BU, UK

Received 5 May 1998; received in revised form 2 June 1999; accepted 10 June 1999

Abstract

This paper presents a review on the behaviour and design guidelines of cylindrical steel
liquid storage tanks subjected to earthquake motion. Field observations during past earthquakes
are presented and then used together with finite element analyses and published experimental
results to assess the accuracy of current design guidelines, with special emphasis on EUROC-
ODE 8 (EC8), Earthquake Resistant Design of Structures, Part 4: Tanks, Silos and Pipelines
(2nd draft, December 1993). The various phenomena addressed in this paper include sloshing
and required free board, base shear and overturning moment applied at the base of the tank
and the buckling strength of tanks. An important aim of this paper is to determine the areas
where current design guidelines need further development.  2000 Elsevier Science Ltd. All
rights reserved.

Keywords: Liquid storage tanks; Seismic analysis; Seismic design; Guidelines; Fluid structure; Interaction
sloshing; Buckling

1. Introduction

Observations from available field reports on the structural performance of tanks


during recent earthquakes indicate that steel tanks, rather than concrete tanks, are
more susceptible to damage and eventual collapse. There is a wide variety of code
guidelines for the earthquake resistant design of steel liquid storage tanks. A common
feature to all these codes is the clear distinction between the treatment of anchored
and unanchored tanks. While there is a consensus in the codes on the treatment of
several aspects of the phenomenon, various other aspects remain controversial or
unresolved (e.g. vertical earthquake motion, effect of tank roof, redistribution of
stresses during tank uplift, etc…). This paper is an attempt at assessing the various

0143-974X/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 3 - 9 7 4 X ( 9 9 ) 0 0 0 3 9 - 5
308 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Nomenclature
a parameter indicating quality of construction
dmax maximum wave height (m)
fs fundamental sloshing frequency (Hz)
g gravitational acceleration
h fill height (m)
pi hydrodynamic pressure (kg/m2)
p0 applied pressure magnitude at z=0, q=0 (N/m2)
r radial coordinate (m)
t shell thickness (m)
Be(Ts) elastic, damping adjusted, normalised response spectrum for period
Ts
D diameter (m)
H height (m)
I Importance factor
M moment at base of tank (N-m)
Pi impulsive rigid hydrodynamic pressure (N/m2)
Pd impulsive flexible hydrodynamic pressure (N/m2)
Ps convective hydrodynamic pressure (N/m2)
Q shear at base of tank (N)
R radius (m)
S soil amplification parameter
V volume of liquid (l)
W weight of tank (N)
Z Seismic zoning coefficient
l imperfection reduction factor
q angle measured from axis of earthquake excitation (radians)
sb buckling strength of an imperfect tank (N/m2)
spr buckling strength of a perfect tank (N/m2)
sy yield stress (N/m2)
sz axial stress (N/m2)
sq hoop stress (N/m2)
t shear stress (N/m2)

codes by comparing their predictions with field observations, finite element analyses
and published experimental results.
The remainder of this paper is divided into four sections. First, field observations
during past earthquakes are highlighted with particular emphasis on the performance
of steel versus concrete tanks. Next, the behaviour of steel tanks is compared to code
predictions and finite element analyses. The final two sections present the concluding
remarks and areas where future work should be oriented.
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 309

2. Summary of field observations during past earthquakes

This section documents the performance of tanks during past earthquakes. Many
tanks were not designed according to current design procedures, and therefore the
damage may not be an indicator of the accuracy of the current design criteria. How-
ever, if the tank geometry and content, soil type, strong ground motion records and
the tank condition after the earthquake are known, it then becomes possible to assess
the accuracy of the code equations in determining the applied earthquake load and
the required load carrying capacity. The complicated deformed configurations of
liquid storage tanks and the interaction between the fluid and the structure result in
a wide variety of possible failure mechanisms. This section discusses the different
collapse modes and summarises the findings of various field studies [1–10]. In the
next section, the accuracy of the EC8 [11] code, amongst others, will be assessed
by applying it to predict the behaviour of various tanks which have experienced
earthquake loading and the performance of tanks under controlled laboratory loading.

2.1. “Elephant foot” and “Diamond shape” buckling

Elephant foot buckling is an outward bulge just above the tank base which usually
occurs in tanks with a low height to radius ratio. In an experimental study conducted
by Niwa and Clough in 1982 [12], it was concluded that the elephant foot buckle
mechanism results from the combined action of vertical compressive stresses
exceeding the critical stress and hoop tension close to the yield limit. However,
Rammerstorfer et al. [13] attributed the bulge formation to three components; the
third being the local bending stresses due to the restraints at the tank base. The
elephant foot buckle often extends around the circumference of the tank. Haroun and
Bhatia [10] attributed this phenomenon to a “reverse” in the direction of earthquake
acceleration which leads to the development of buckling in a region opposite to the
location of the initial buckle. Fig. 1 shows an elephant foot buckle for a tank
destroyed by an earthquake (Wyllie et al. [14]).

Fig. 1. Elephant foot buckle mechanism.


310 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Diamond shape buckling is an elastic buckling phenomenon due to the presence


of high axial compressive stresses. Niwa and Clough [12] concluded that the rocking
motion which develops at the base of unanchored tanks generates very high com-
pressive axial stresses surrounding the contact zone which in turn lead to the diamond
shape buckle. Diamond shape buckles may also occur well above the base of the
tank where the hydrodynamic pressure, which leads to an increase in the elastic
buckling load, is small as compared to its magnitude at the tank base. Fig. 2 shows
a diamond shape buckle developed in an experimental tank model subjected to earth-
quake motion (Niwa and Clough [12]).
Table 1 summarises some of the buckling damage and collapse of steel tanks
during past earthquakes. It can be seen that unanchored tanks are more prone to
buckling. Furthermore, there is very little published data on the seismic-induced
buckling of concrete tanks.

2.2. Uplifting of tanks

Both anchored and unanchored tanks may undergo local uplift when the magnitude
of the overturning moments exceeds a critical value. As a result, a strip of the base
plate is also lifted from the foundation. Although uplift does not necessarily result
in the collapse of the tank, its consequences include serious damage to any piping
at the connection with the tank and an increase in the axial stress acting on the tank
wall which remains in contact with the ground. Table 2 provides a list of some
of the tanks that underwent uplift and damage to anchors during past earthquakes.
Examination of Tables 1 and 2 indicates that the uplift of unanchored tanks may
lead to buckling.

2.3. Damage and collapse of tank roofs

Sloshing waves of high amplitude often cause damage to the roofs of tanks and
render them temporarily unserviceable. As a consequence, liquid spillage over the

Fig. 2. Shape buckle mechanism.


F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 311

Table 1
Buckled and collapsed steel tanks during past earthquakes

Tank Radius R Height h Number % Full Thick. Damage Earthquake Type of


name (m) (m) anchors (mm) product

AL-1 3.05 12.2 0 100 4.766 ia Alaska 1964 *d


AL-2 4.27 12.2 0 100 4.766 iib Alaska 1964 *d
AL-3 4.575 10.675 0 100 4.766 ia Alaska 1964 *d
AL-4 6.1 12.963 0 100 7.116 ia Alaska 1964 *d
AL-5 6.405 12.2 0 100 5.846 ia Alaska 1964 *d
Sesnon 14.02 12.16 0 95 N/A ia San Fernando, Water
1971
Wash 15.24 5.565 12 50 17.5 iiic San Fernando, Water
1971
SJ1/A 3.75 12.667 29 100 6.375 ia San Juan, 1977 Wine
SJ2/ 3.83 10.97 27 100 4.596 ia San Juan, 1977 Wine
SJ4/ 2.6 12.0 16 100 4.68 iib San Juan, 1977 Wine
IP-5 7.289 10.63 0 73 N/A ia Imperial Valley, Gasoline
1979
IP13 6.283 13.27 0 89 6.93 ia Imperial Valley, Gasoline
1979
GV1 1.983 6.096 8 100 1.981 iib Greenville, 1980 Wine
GV3 1.448 5.337 2 100 1.981 ia Greenville, 1980 Wine
T-400 8.38 7.32 0 75 N/A ia Coalinga, 1983 Oil
T-262 10.668 10.668 0 100 N/A iiic Coalinga, 1983 Oil
3978 6.48 8.53 0 100 N/A ia Loma Prieta, 1989 Lube oil
A 6.48 8.53 0 82 N/A ia Loma Prieta, 1989 Lube oil
Nor 1 9.14 9.0 0 98 N/A ia Northridge, 1994 Water
BG 15.24 11.67 0 94.5 N/A ia Northridge, 1994 Water
ANC 5.64 8.63 0 88.5 10 ia Northridge, 1994 Water

a
i: Buckled.
b
ii: Collapsed.
c
iii: Buckling in upper tank walls.
d
*: specific gravity =0.8.

roof may either result in fires or in the loss of water supply used in putting out fires.
Table 3 documents some of the damage to tank roofs during past earthquakes. It
can be seen that the most commonly reported damage to above ground prestressed
concrete tanks is due to sloshing.

3. Design guidelines for anchored and unanchored tanks

The accuracy of the codes is assessed by contrasting the predicted values with
those obtained from field observations, published experimental results and finite
element analyses. The following design guidelines are considered in this study.
312 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Table 2
Uplifting and damage to anchors during past earthquakes

Tank Radius R Height H Number % Full Thickness Damage Earthquake Type of


name (m) (m) anchor (mm) product

Wash 15.24 5.565 12 50 17.5 iib San Fernando, Water


1971
SJ3 2.87 13.54 18 100 4.879 ia San Juan, 1977 Wine
SJ5/c 2.89 8.14 16 100 4.913 ia San Juan, 1977 Wine
SJ1/A 3.75 12.0 29 100 6.375 iib San Juan, 1977 Wine
GV1 1.983 6.096 8 100 1.981 iib Greenville, 1980 Wine
GV3 1.448 5.337 2 100 1.981 iib Greenville, 1980 Wine
GV4 1.983 4.88 2 100 2.667 ia Greenville, 1980 Wine
GV5 1.448 3.05 2 100 1.524 iib Greenville, 1980 Wine
20002 8.38 12.42 0 N/A N/A iiic Loma Prieta, 1989 Diesel
oil
3956 6.48 8.53 0 100 N/A iiic Loma Prieta, 1989 Gasoline

a
i: fracture of all anchors.
b
ii: fracture of anchors.
c
iii: uplift.

3.1. American Petroleum Institute (API) 650 [15] — 1993 and AWWA 1984 [16]

Both of these codes are based on one of the earliest codes to develop a systematic
approach for treating the seismic design of liquid storage tanks, based on a paper
by Wozniak and Mitchell [17]. This design procedure offers a concise way of
determining the maximum applied axial stresses for anchored as well as unanchored
tanks. Minimum anchorage requirements are also presented. Only one horizontal
earthquake component is considered and the effects of vertical excitation and soil-
structure interaction are not taken into account. Elastic buckling is addressed while
elasto-plastic buckling (elephant foot buckle) is not accounted for. Finally the flexi-
bility of the tank is also ignored when calculating the hydrodynamic pressures.

3.2. ASCE recommendations — 1984 [18]

This set of design recommendations account for tank flexibility, soil-structure


interaction and vertical excitation. Its main limitation is that it does not account for
the redistribution and change of magnitude of the axial and hoop stresses due to
uplift of the base plate of the tank. As the codes mentioned previously, it suggests
a direct superposition of the impulsive and convective effects. In addition, it does
not account for the presence of two horizontal earthquake components. Finally it
suggests a rule for the addition of the vertical and horizontal hydrodynamic pressures
using the Square Root of the Sum of the Squares (SRSS) Method. Again elasto-
plastic buckling is not accounted for.
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 313

Table 3
Damage and buckling of tank roofs during past earthquakes

Tank Radius R Height H Anchors % Full Thick. Damage Earthquake Type of


name (m) (m) (mm) product

Sesnon 14.02 12.16 0 95 N/A ia San Fernando, Water


1971
Granada 8.38 10.0 0 73 N/A iib San Fernando, Water
1971
Res1* 7.62 3.05 0 83 N/A iib San Fernando, Water
1971
Res5* 25 4.57 0 88 N/A iiic San Fernando, Water
1971
IP-5 7.289 10.63 0 72.6 N/A ivd Imperial Valley, Gasoline
1979
IP-16 7.32 12.16 0 83 8.33 ve Imperial Valley, Gasoline
1979
T-5 21.34 14.63 0 100 N/A vif Coalinga, 1983 Oil
T-9 21.34 4.87 0 33 N/A ve Coalinga, 1983 Oil
T-FL 18.3 13.11 0 93.5 N/A vif Coalinga, 1983 Oil
BG 15.24 11.67 0 94.5 N/A ve Northridge, 1994 Water
MH 7.925 9.6 0 94 N/A viig Northridge, 1994 Water
Susanai 30.5 13.41 100 N/A viiih Northridge, 1994 Water

a
i: roof uplift.
b
ii: roof collapse.
c
iii: cracks in concrete roof.
d
iv: leak and dents.
e
v: damage to roof.
f
vi: buckling of roof.
g
vii: overflow.
h
viii: chipping of roof.
i
prestressed concrete tanks.

3.3. Austrian design recommendations [19]

The design recommendations of this code include the effects of tank flexibility,
three dimensional earthquake motion and soil-structure interaction. It also provides
the most rational approach for determining the redistribution of axial stresses due to
the uplift of the tank base plate. It offers different scenarios for superimposing the
horizontal, vertical and static effects according to the failure mode under consider-
ation. The main limitation is that the formulae given for the redistribution of axial
stresses due to uplift do not extend to all geometries.

3.4. Eurocode 8 (EC8), Part 4 [11]

This code takes into account soil-structure interaction, tank flexibility, vertical
excitation and the effects of base uplift on the axial stress distribution in the tank
314 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

walls. It is based on the code above, except for the parameters shown in Table 4
which summarises the code recommendations.

3.5. New Zealand guidelines [20]

These guidelines take into account wall flexibility and vertical excitation. In
addition it is the first code to account for the reduction in the buckling load due to
the tensile hoop stresses approaching the yield stress (elasto-plastic buckling). How-
ever, the redistribution of axial stresses due to uplift is not accounted for.

3.6. Japanese guidelines [21–23]

In these codes the effects of tank flexibility and vertical excitation are taken into
account, however the issues of soil-structure interaction, uplift and the effects of
large tensile hoop stresses, close to the yield limit, are not adequately considered.

4. Assessment of code guidelines

The phenomena examined in this paper are liquid sloshing displacement, hydrody-
namic pressure acting on the tank walls, shear force and overturning moment acting
at the base of the tank, the buckling strength of tanks and the effects of uplift on
the axial stresses acting on the tank wall in contact with the ground. These are
discussed in Sections 4.1, 4.2, 4.3, 4.4 and 4.5 respectively.

4.1. Sloshing wave height

The determination of the maximum sloshing wave height requires a value for
the ground acceleration corresponding to that particular mode of deformation of the
contained liquid. Two different methods may be used to determine the value of the
ground acceleration. The first is to use the peak acceleration of an event as the design
ground acceleration which is a prerequisite for using the elastic response spectrum

Table 4
summary of the different strategies of the codes

Property Content Impulsive Horizontal Horizontal Tank SSI Elasto- Effect of


code damping + excitation + vertical flexibility plastic uplift on
Convective buckling

API650 0.5% Direct 1 direction N/A N/A N/A N/A N/A


ASCE 0.5% Direct 1 direction SRSS Yes Yes N/A N/A
Austrian 0% SRSS 2 directions Direct Yes Yes N/A Axial stress
EC8 0.5% SRSS 1 direction SRSS Yes Yes N/A Axial stress
NZ 0.5% SRSS 1 direction SRSS Yes Yes Yes Axial stress
Japanese 0.5% SRSS 1 direction SRSS Yes N/A N/A N/A
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 315

specified by the code (EC8, part 1-1 [24], Section 4.2.2: Elastic Response Spectrum).
An alternative approach would be to obtain the response spectrum of an event, and
using a specific period, get a corresponding value for the spectral acceleration. How-
ever, this often has to be done in the absence of any records of ground motion in
the vicinity of the tanks involved. In this study the former option will be adopted.
In EC8 [11] the frequency corresponding to the first sloshing mode is given as:

fs⫽
2p冉
1 1.84g tanh(1.84H/R)
R 冊1/2
(Hz) (1)

where H is the liquid height (m), and R is the tank radius (m).
One of the main requirements of a safe design is the accurate estimation of the
required freeboard. This is important to safeguard against roof damage and its conse-
quences such as spilling of contents over the roof. In addition, pressures due to
fluid impact during sloshing contribute to the overturning moment and therefore may
increase the possibility of uplift and buckling. According to the expression given in
EC8, and considering the contribution of the first mode alone, the maximum magni-
tude of the wave height, which occurs at q =0° and r = R, is:
dmax⫽0.84R(Sbe(Ts)) (m) (2)
where r is the radial coordinate, q is the circumferential angle, dmax is the maximum
wave height, Ts is the period corresponding to the first sloshing natural frequency,
fs, S is the soil amplification factor, and be(Ts) is the elastic, damping adjusted, nor-
malised response spectrum, corresponding to 0.5% liquid damping.
It can be seen that EC8 provides a rational approach for determining the maximum
sloshing wave height as it is a function of the sloshing period, the sub-soil classi-
fication and the peak ground acceleration. However, EC8 does not consider the pres-
ence of two horizontal earthquake components which may lead to higher wave
heights. The same approach is adopted in the ASCE and New Zealand design guide-
lines. The Austrian code uses a similar approach but with a response spectrum corre-
sponding to a 0% liquid damping.
The API650 and the AWWA codes are based on the paper by Wozniak and Mitch-
ell [17] which suggest the following formula for the determination of the maximum
wave height:
dmax⫽1.124ZIC2T 2s tanh(4.77 H/D) (ft)冑 (3)
where Z is the seismic zoning coefficient, I is the importance factor, and C2 is the
convective response spectral coefficient defined as:
S S
0.3 for Ts⬍4.5 sec or 1.35 2 for Tsⱖ4.5 sec, and
Ts Ts
S is the soil foundation deformability coefficient.
It should be noted that while Eq. (3) is a function of the sloshing period and the
sub-soil classification, it is not a function of the peak ground acceleration. It is there-
fore possible that Eq. (3) will give identical results for different earthquakes or design
316 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Table 5
Predicted wave height in tanks during the 1979 Imperial Valley earthquake (mm)

Tank Freeboard/observation API (650) Japan EC8

IP-5 3985/No damage 413 1110 688


IP-13 1673/No damage 380 1030 694
IP-16 2491/Damage 413 1110 691

spectrums and liquid damping ratios. This detracts from the rationality of the
above approach.
The Japanese code equation for the maximum sloshing wave height is given by:
dmax⫽0.42DKh2 (m) (4)
where
Kh2⫽0.15v1v4
v1 is the zoning coefficient, and
4.5
v 4⫽
Ts

It should be noted that Eq. (4) is a function of the tank diameter D and sloshing
period Ts, however it does not account for soil conditions or ground acceleration.
Tables 5 and 6 compare the predicted wave height of various unanchored tanks
using the above codes with field observations during the 1979 Imperial Valley earth-
quake and 1983 Coalinga earthquake respectively. Examination of Tables 5 and 6
shows that current design guidelines could not have predicted the damage which
occurred to tank IP-16. This may be attributed to the assumption of small amplitude
wave motion which is adopted in all of the codes reviewed in this study.
Maximum free surface wave heights predicted using the code equations are also
compared to field observations of tanks during the Northridge earthquake, as shown

Table 6
Predicted wave height in tanks during the 1983 Coalinga earthquake (mm)

Tank Freeboard/observation API (650) Japan EC8

T-FL 915/Damage 460 1637 1215


T-5 3000/No damage 377 1657 926
T-9 549/Damage 393 1737 1185
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 317

in Tables 7 and 8. As can be seen from Table 8, in many cases the available freeboard
was less than that required by any of the codes. This may be due to the fact that
some of these tanks were of old design. However, in several cases design according
to API-650 or EC8 would have resulted in an unsafe design (it should be noted that
site conditions of the soil were not known). Furthermore, the difference between the
various codes is very large, which suggests that there are parts of the problem which
have not yet been understood.
The experimental wave heights measured by Manos [25] are reproduced in Table
9, along with those predicted by the various codes. The spectral accelerations
presented by Manos, shown between brackets in column 1, were used in determining
the EC8 values. It can be seen that the values predicted by the API(650) and the
Japanese codes are not a function of the spectral acceleration, whereas those pre-
dicted by EC8 are. Examination of the results shows that the sloshing wave height
predicted by EC8 is consistently unconservative for the unanchored flexible foun-
dation case while being sometimes unconservative for the unanchored rigid foun-

Table 7
Predicted wave heights (m) in tanks during the Northridge earthquake

Area/tank ID Ts (sec) EC8 API-650 Japan

Soil A Soil B Soil C

NOR1 4.59 0.66 1.00 1.08 0.69 1.21


NOR2 3.66 0.70 1.05 1.13 0.56 1.01
NOR3 3.66 0.70 1.05 1.13 0.56 1.01
NOR4 3.66 0.70 1.05 1.13 0.56 1.01
NOR5 4.75 0.66 0.99 1.07 0.69 1.25
NOR6 2.58 0.60 0.91 0.98 0.4 0.72
NOR7 5.91 0.60 0.90 0.98 0.69 1.41
Larwin 4.51 0.55 0.82 0.89 0.69 1.23
MM1 4.73 0.50 0.75 0.81 0.69 1.17
MM2 5.40 0.46 0.70 0.75 0.69 1.25
Zalzah 4.96 0.56 0.84 0.91 0.69 1.31
MulHolland 4.21 0.58 0.87 0.94 0.65 1.14
Beverly Glen 6.13 0.53 0.79 0.85 0.69 1.51
Cold Water 6.13 0.53 0.79 0.85 0.69 1.51
Granada High 4.33 0.58 0.87 0.94 0.67 1.18
Susana 10.9 0.39 0.58 0.63 0.68 1.84
Tajunga 8.04 0.43 0.64 0.70 0.68 1.61
A 5.03 0.73 1.08 1.17 0.69 1.2
B 6.64 0.48 0.69 0.74 0.66 1.00
C 5.65 0.53 0.79 0.86 0.67 0.98
AG-1 2.00 0.56 0.84 1.08 0.31 0.56
NOR8 3.11 0.79 1.19 1.28 0.48 0.86
NOR9 3.46 0.79 1.19 1.29 0.59 0.96
NOR10 4.75 0.79 1.18 1.28 0.69 1.31
NOR11 4.94 0.77 1.17 1.27 0.69 1.35
NOR12 2.24 0.59 0.88 0.96 0.35 0.62
318 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Table 8
Summary of comparison between code range prediction and observed freeboard

Area/tank ID Predicted maximum wave Freeboard (m) Damage


height (m)

Newhall Area
NOR4 0.56 to 1.13 0.02 ia
NOR5 0.66 to 1.25 0.25 ia
NOR7 0.60 to 1.41 0.15 ia
Los Angeles
MulHolland 0.58 to 1.14 0.6 iib
Beverly Glen 0.53 to 1.51 0.7 iiic
Cold Water 0.53 to 1.51 0.7 iiic
Susana 0.39 to 1.84 0.45 ia

a
i: roof damage.
b
ii: overflow.
c
iii: roof collapse.

dation case. The results also show the variation of the experimental data with the
boundary condition at the base of the tank. This is in contrast to current European
design practice where it is assumed that the base boundary condition does not affect
the maximum sloshing wave height, which is determined assuming a rigid boundary
condition. This becomes especially important since the presence of a flexible foun-
dation appears to increase the maximum sloshing wave height [26].
The next step is to carry out finite element analyses to determine the accuracy of
the code in predicting the maximum free surface wave height. The results will be
validated by comparison with field observations during past earthquakes. Although
many tanks of known dimensions suffered damage or roof failure, only those which
were not full, and where the freeboard at the time of the event is known, can be
used for comparison purposes.
The frequencies of the fundamental sloshing and fundamental tank deformation
modes are clearly separated. In this study, the effect of initial imperfections is not
taken into account, and it is assumed that the tank wall will deform in its fundamental
circumferential mode. Therefore, to determine the maximum sloshing wave height
at the liquid free surface, the tank is assumed rigid and a series of dynamic analyses
of the fluid domain are carried out. Two cases will be considered, corresponding to
the Mulholand (MH) and Beverly Glen (BG) tanks. The geometry of these tanks is
shown in Table 3, while the predicted wave heights and the available free board are
shown in Tables 7 and 8 respectively. Two sets of earthquake records were used,
one corresponding to the longitudinal component of the 1940 El-Centro earthquake
and another to the transverse, longitudinal and vertical motion recorded at Santa
Monica station during the 1994 Northridge earthquake.
Preliminary studies using two-dimensional geometry were initially undertaken
with the objective of validating the sloshing formulation and investigating any
insights that may be gained from such an analysis. The Mulholand and Beverly Glen
tanks were modelled using an 8×6 and a 10×6 eight noded plane strain fluid finite
Table 9
Maximum wave height (mm)

Record Measured API (650) Japan EC8

ia iib iiic ia iib iiic ia iib iiic ia iib iiic

El Centro (0.07 g) 134.4 137.2 142.2 215 215 215 161.8 161.8 161.8 134 134 134
Pacoima (0.025 g) 31 50.8 63.5 215 215 215 161.8 161.8 161.8 31 31 31
Parkfield (0.105 g) 155.7 151.1 172.7 215 215 215 161.8 161.8 161.8 155.7 155.7 155.7

a
i: anchored, rigid foundation
b
ii: unanchored, rigid foundation
c
iii: unanchored flexible foundation
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333
319
320 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

elements respectively [26]. In both cases a uniform mesh was used. All runs are
made assuming a rigid base.
The Mulholand tank was also analysed using a three dimensional fluid mesh con-
sisting of fifteen and twenty noded fluid finite elements. Fig. 3 shows the free surface
displacement at r = R and q=0°, while Fig. 4 shows contours of the free surface
displacement (for the three dimensional case) at a time corresponding to the occur-
rence of the peak displacement in Fig. 3. In both cases the transverse component of
the El-Centro earthquake record was used as loading in the x direction. As expected,
Fig. 4 shows that the maximum displacement occurs at q=0° and r = R, while q=90°
the free surface displacement is negligible. Also, the free surface displacement is
symmetric about the y axis.
Fig. 5 shows the free surface contours of the Mulholand tank when subjected to
the three components of the Northridge earthquake. The displacements are no longer
symmetric and the maximum value of the displacements does not occur at r = R.
The maximum free surface displacement obtained from the three-dimensional finite
element analysis lies in the range of the code predictions. However it is higher than
that predicted by the EC8 code. Future runs are needed to investigate the validity
of the code suggestions regarding the superposition of the liquid response due to the
various components of the earthquake record. However, at this stage it can be con-
cluded that the three-dimensional nature of the earthquake must be taken into account
when determining the maximum free surface displacement.

Fig. 3. Response of the Mullholand tank subjected to the El-Centro earthquake record.
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 321

Fig. 4. Contours of the maximum free surface vertical displacements under uni-directional horizontal
earthquake excitation.

4.2. Hydrodynamic pressures

EC8 divides the hydrodynamic pressure into three different components due to
the movement of the liquid assuming the tank shell is rigid (impulsive-rigid), the
movement of the liquid due to the flexibility of the tank (impulsive-flexible) and the
movement of the liquid due to free surface sloshing (convective).

4.2.1. Impulsive-rigid hydrodynamic pressures


The contribution of this pressure to the total pressure is given by
pi(r,z,q)⫽Ci(r,z)r Rcos q (kg/m2) (5)
where

Ci(r,z)⫽0.81a 冘
冉 冊 冉 冊
j−1
−1 2 I1 j
pr
2H
cos j
pz
j⫽1,3,5%
j
lj jI1⬘(lj ) 2H
p
lj ⫽j ,
2a
z is the axial coordinate, and I1(.) is the modified Bessel function of the first kind.
322 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Fig. 5. Contours of the maximum free surface vertical displacements under three dimensional earth-
quake excitation.

4.2.2. Impulsive-flexible hydrodynamic pressures


This component is given by
pd(r,z,q)⫽Cd(r,z)r Rcosq (kg/m2) (6)
where

Cd(r,z)⫽2 冘
冉 冊
I1 j
pr
2H
冉 冊
F(z/H)cos j
pz
,
j
lj I1⬘(lj ) 2H

冕 冉 冊
1
pz
F(z/H)⫽ f(z/H)cos j ∂(z/H), and
2H
0

f(z/H) is the normalised mode of vibration of the shell.

4.2.3. Convective hydrodynamic pressures


This component is given by
ps(r,z,q)⫽Cs(r,z)r Rcosq (kg/m2) (7)
where J1(.) is Bessel function of the first kind, and
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 323

Cs(r,z)⫽
0.84
J 1.84
J1(1.8413) 1 冉 冊
r cosh(1.84z/R)
R cosh(1.84a)
.

The peak value of any one component of the hydrodynamic pressure is obtained
from the following equation:
|Pk(r,z,q,t)|max⫽pk(r,z,q)Sagbe(Tk)(N/m2) (k⫽i,d,s) (8)
where ag is the horizontal peak ground acceleration in m/sec2. The above formulae
may be applied to determine the hydrodynamic pressures in any of the two horizontal
directions. A similar treatment is adopted for the determination of the hydrodynamic
pressure due to vertical excitation.
Due to size limitations it is not possible to present the formulae of all the codes
in this section. However, the areas where the codes differ include the effects of the
flexibility of the tank on the impulsive pressure, the superposition of the impulsive
and convective components of the pressure, the superposition of the pressure due to
the horizontal and vertical excitations, the number of horizontal earthquake compo-
nents considered in the design and the effects of soil structure interaction. The treat-
ment of the above factors by the various codes is summarised in Table 4.
Hydrodynamic pressures determined from the above equations are compared to
those measured by Clough [2], for the El-Centro record with a pga =0.5 g. Table
10 shows the results obtained and the contribution of each pressure component, for
two different magnitudes of the peak ground acceleration (pga). In determining the
values for the impulsive-flexible hydrodynamic pressure the mode shape was
assumed to consist of one-half sine wave, with a maximum normalised displacement
equal to one. In both instances, the code underestimates the magnitude of the press-
ure. Before any conclusions are drawn, the effects of the choice of superposition
method (in this case SRSS) and response spectrum (actual versus code specified
elastic response spectrum) must clearly be investigated further.

4.3. Base shear and overturning moment

An accurate prediction of the base shear and overturning moment is essential for
judging the safety of tanks against shell buckling and uplift. The EC8 code gives
the following expressions for the base shear and overturning moment:
|Qk|max⫽mkSagbe(Tk)(k⫽s,i,d) (N) (9)

Table 10
Predicted vs. measured hydrodynamic pressures (N/m2)

Component pi pd ps Total

Measured – – – 8294
Predicted (pga=0.5 g) 1962 855 1244 2475
Predicted (pga=0.84 g) 3297 2413 3511 5387
324 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

|Mk|max⫽mkhkSagbe(Tk)(k⫽s,i,d) (N⫺m) (10)


where mk is the mass associated with a liquid component, hk is the corresponding
height.
Due to size limitations it is not possible to summarise the formulae of all the
codes in this section. The areas where the codes differ have been outlined in the
previous section and the approach of each code is summarised in Table 4.
Maheri and Severn [27] measured the base shear in a water filled storage tank
subjected to the Parkfield earthquake record (pga=0.25 g). The tank diameter is 0.153
m, the tank height is equal to 1.33 m and the thickness is 2.286 mm. Table 11
compares the base shear determined by the EC8 equation with the experimental and
other design procedures [28]. It can be seen that all methods underestimate the meas-
ured base shear. However, it should be noted that the dynamic similitude require-
ments were not satisfied in this experimental study. The contributions of the separate
components were not given in the other study and, therefore, the accuracy of the
expressions for the separate contributions cannot be assessed.

4.4. Elastic and elasto-plastic buckling strength of anchored tanks

In the EC8 code stability checks have to be carried out for two possible failure
modes: elastic and elasto-plastic buckling.

4.4.1. Elastic buckling


The axial stress required to cause buckling in a cylindrical shell structure is a
function of the internal pressure, the amplitude of imperfections, shell thickness and
the circumferential variation of the axial stress. Imperfections tend to decrease the
buckling strength of a perfect shell. A reduction factor, l, can usually be calculated
based on the amplitude of imperfections.
First evaluating the following non-dimensional, a, parameter is defined:
sy
a2 ⫽ (11)
lspr
where l is a reduction factor due to imperfections which can be estimated as:

Table 11
Predicted vs. measured base shear (N)

Component Qi Qd Qs Total

Measured 570
Haroun and Housner 443
Maheri and Severn 487
EC8 426 64 44 433
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 325

冤冪 冥
2
d 1+ ⫺1
l⫽1⫺1.24 d
t 1.24
t

冪t
d 0.06 R

t a
a is a parameter indicating the quality of the construction: a=1 for normal construc-
tion; a=1.5 for quality construction; a=2.5 for very high quality construction; spr is
the buckling strength of a perfect tank defined as: 0.6 Et/R, E is the young modulus
of the material, t is the thickness of the tank shell, and sy is the yield stress of
the material.
A buckling stress parameter, so, is defined according to the following formula:
a2ⱖ2⇒so⫽lspr (N/m2) (12a)

a2ⱕ2⇒so⫽sy 1⫺
a2
4 冉 冊 (12b)

Eqs. (12a) and (b) reflect the reduction in the buckling stress due to imperfections.
Eq. (12a) applies for thin shells, while Eq. (12b) is for thick shells. Denoting by sb
the maximum vertical membrane stress, the following inequality should be satisfied:
sb sd
ⱕ0.19⫹0.81 (N/m2) (13)
spr spr
where

冪冉1−冉1−5冊 冉1−冉s 冊 冊冊ⱕs


2 2
s̄ so
sd⫽spr pr (N/m2) (14)
pr

pR
s̄⫽ ⱕ5
tspr
where p is the lowest hydrodynamic pressure. Eq. (13) reflects the increase in the
buckling strength due to the axial stresses being induced by bending action rather
than uniform axial loading. Circumferential variation of the buckling stress decreases
the probability of coincidence of the maximum axial stress with the maximum imper-
fection amplitude, and therefore increases the buckling strength. In Eq. (14) the
internal hydrodynamic pressure decreases the effects of imperfections and therefore
increases the buckling stress.
According to EC8, when checking against elastic buckling, the contribution of the
vertical component of the earthquake motion to the internal pressure should be
ignored. Furthermore, since the internal pressure increases the buckling strength it
326 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

is possible that elastic buckling will occur well above the base of the tank where
both the shell thickness and the hydrodynamic pressure are smaller than the corre-
sponding ones at the base of the tank. Indeed, this is often observed in the pattern
of diamond shape buckles in steel tanks.
It can be seen from the above discussion that EC8 provides a rational approach
to determine the elastic buckling strength of tanks which takes into account imperfec-
tions, internal pressure and non-uniform axial stresses. This is based on the New
Zealand [19] code which first adopted this approach. All the other codes account
for the effects of internal pressure, while ignoring the effects of bending.

4.4.2. Elasto-plastic buckling


The bottom of the tank is usually subjected to a bi-axial stress state consisting of
axial membrane compression and circumferential hoop tensile stress. The EC8 for-
mula for elasto plastic collapse is based on the work of Rotter [29,30], and is given
by the following relationship:

sbⱕspr 1⫺冋 冉 冊 册冉
pR
tsy
2
1⫺
1
1.12+s1.5冊冋
s+sy/36
s+1 册
(N/m2) (15)

where s is a non-dimensional factor given by


R/t
s⫽
400

In contrast to the case of elastic buckling, the effect of the internal pressure in
Eq. (15) is to reduce the buckling strength of tanks. In this case the hydrodynamic
pressure due to all the components of the earthquake excitation must be considered
to get a conservative design. The New Zealand code was the first to adopt this
relationship based on the work of Rotter. The Austrian design guidelines are based
on the same expression, while all the remaining codes do not account for elasto-
plastic buckling. The above relationship was developed for pinned boundary con-
ditions at the tank base. However, fixed boundary conditions at the base result in a
higher buckling strength. Furthermore, at the boundary between two shell courses
with different thickness, conditions close to a fixed tank base prevail. Therefore the
above equation may be used for all these cases. Fig. 6 shows the elastic and elasto-
plastic buckling strengths of tanks as calculated by the formulae above.

4.4.3. Finite element analyses


A complete description is provided by Haroun and Bhatia [10] of an unanchored
tank that underwent Elephant Foot Buckling at the American National Can Company.
The tank is 11.28 m (37 ft) in diameter, 9.75 m (32 ft) in height, filled with water
to a depth of 8.63 m (28.3 ft) and has a thickness of 10 mm (3/8 in). It is supported
on a concrete pad. Additional parameters are: thickness of bottom plate 13 mm (1/2
in), yield stress of steel =250 MPa (36 ksi), unit weight of steel =7.7 × 105 N/m3
(0.2835 lb/in3), E =2.1 × 105 MPa (30 × 106 psi); and unit weight of water =9.81
Fig. 6. Elastic and elasto-plastic buckling strengths of anchored steel cylindrical tanks.
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333
327
328 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Table 12
Summary of code predictions

Property code API650 AWWA ASCE Austrian EC8- Haroun and


Flexible Housner
[31]

Qmax (MN) 1.04 1.5 3.1 3.73 3.7 5.78


Mmax (N m) 0.87 E07 0.98 E07 0.26 E08 0.12 E08 0.22 E08 0.23 E08
Mallowable (N m) 0.54 E07 0.54 E07 0.54 E07 N/A N/A 0.54 E07
szmax (MPa) 10.1 11.1 15.8 12.6 23.1 24.35
sqmax (MPa) N/A N/A 78 143 125 130
sqyield (MPa) N/A N/A 250 250 250 250

× 104 N/m3 (0.03611 lb/in3). The spectra of the free field ground motion recorded
at Arleta station is the closest to the site of the tank. In view of the complete set of
parameters, this tank will be used as a case study for assessing the accuracy of the
code. Table 12 provides a summary of the various code predictions for the above
tank properties.
The general purpose finite element package Abaqus is used to simulate static tilt
analyses. The objectives of the analyses are: (1) to determine whether an anchored
base would have prevented damage to the tank; (2) model the uplift behaviour of
the tank and compare the results with those predicted by the codes; and (3) give
insights into the effect of uplift on the axial and hoop stresses at the base of the
tank. The tank and base plate are modelled with thin shell elements. The static tilt
condition is simulated by applying a distributed load which varies both in the axial
and circumferential directions (Fig. 7).
The distribution of the pressure in the axial and circumferential directions is
defined as:

冉 冊
z2
p⫽p0 1⫺ 2 cosq
h
(16)

The overturning moment about the centre of the base, M, and the base shear, Q,
may be computed from:

gr7.tif

Fig. 7. Pseudo hydrodynamic pressures applied at the tank walls.


F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 329

冕冕
H 2⌸

M⫽ pRz cos q∂q∂y⫽ H2p0R (17)
4
0 0

冕冕
H 2⌸
2⌸
Q⫽ pR cos q∂q∂y⫽ Hp0R (18)
3
0 0

As a first step an anchored tank was analysed, taking into account geometric and
material nonlinearities. The results show that the anchored tank can clearly withstand
pressures higher than those that led to the collapse of the unanchored one. The magni-
tude of the pressure at the base, P0, was taken as 71820 N/m2. Hence the total
moment applied is 8.27 MN which is well above that which led to the collapse of
the unanchored tank. Fig. 8 shows the deformed shape of the tank.

4.5. Redistribution of stresses during uplift

Many tanks are unanchored, particularly those of old construction. This is due to
the high cost of a concrete foundation and corresponding anchors. These tanks seem
to be more susceptible to damage and collapse than their anchored counterparts. The
API 650 and the AWWA suggest an increase in the axial force in the tank wall
which remains in contact with the ground during uplift. However, this increase is
independent of the magnitude of uplift and is not always conservative. The ASCE
recommendations acknowledge the possible increase in the axial stresses however,
it is argued that this will be counterbalanced by the reduction in the acceleration

Fig. 8. Deflected shape of anchored tank.


330 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

amplitude due to the higher period of the tank vibration mode in its uplifted position.
The Austrian design guidelines provide the most comprehensive regulations for
assessing the increase in the axial force as a function of the magnitude of the uplift
displacement. However, the design charts are limited to a narrow range of tank thick-
ness and soil-stiffness. Furthermore, it can be used in the presence of vertical exci-
tation with only one horizontal earthquake component. The EC8 guidelines are based
on the Austrian procedure and have the same limitations. While this is a step in the
right direction, there is a clear need for more research in this area.

5. Concluding remarks

5.1. Sloshing

1. The code formulas used to predict the maximum free surface displacement vary
significantly with the soil type considered, while no allowance is made for the
boundary condition at the tank base.
2. Very little insights can be gained from two dimensional analysis of liquid stor-
age tanks.
3. The maximum free surface displacement obtained from the three-dimensional
finite element analysis lies in the range of the code predictions. However, it is
higher than those predicted by EC8 for a similarly anchored tank.
4. From Figs. 4 and 5 it seems that a superposition of the maximum free surface
displacements corresponding to the various earthquake components in order to
obtain the maximum free surface displacement may be erroneous. The combined
presence of the three earthquake components in a finite element analysis seems to
significantly alter the distribution and magnitude of the free surface displacements.
Interaction between the different responses corresponding to the various earth-
quake components must be considered.
5. EC8, Clause 1.4.3.2: Contents Damping — Page 11, recommends “the value of
x=0.5% may be adopted for water and other fluids unless otherwise mentioned”.
Other researchers [19] have adopted a damping value of 0.0%. No justification
is given for the recommended value, which may be less conservative than a damp-
ing value of 0.0%. Furthermore all the code equations for determining the
maximum sloshing displacement are based on small amplitude sloshing theory,
while field observations indicate the occurrence of large sloshing waves at the
liquid free surface during an earthquake.

5.2. Hydrodynamic pressures, overturning moment and base shear

It was not possible to find any conclusive evidence concerning the accuracy of
the code equations used for determining the hydrodynamic pressures, overturning
moment and base shear. Experimental studies should be undertaken with the aim of
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 331

measuring the separate contributions, of each component of fluid motion, to the total
magnitude of the pressure, base shear and over turning moment.

5.3. Effects of tank-base support on axial compressive stresses

The formulae given by the Austrian design guidelines and EC8 regarding the effect
of base uplift on the magnitude and distribution of the axial stresses around the tank
base should be extended to a wider range of tank geometries.

5.4. Effect of tank-base support on hoop stresses

At the moment there are no design guidelines regarding the effect of uplifting on
the distribution of the hoop stresses. These are considered unaltered by the uplift
mechanism of the tank. However, recently, preliminary finite element analyses by
Haroun and Bhatia [10] have shown that the hoop stresses are increased in the pres-
ence of uplift of the base plate.

5.5. Effect of roof strength and stiffeners on the buckling load

While the EC8 formulae for the buckling strength provide a rational approach for
determining the stability of tanks, they do not explicitly account for the presence of
any stiffeners. Furthermore, the effect of the roof presence on the buckling strength
should be further investigated.

5.6. Presence of two earthquake components

EC8, states that “Tanks shall be designed assuming the concurrent presence of
one horizontal and one vertical component of the seismic motion”.
It has been shown [19] that under the combined effect of two horizontal earthquake
components, the maximum horizontal pressure and associated stresses need not occur
in the direction of application of one of the horizontal ground motion. Hence some
form of superposition of the two horizontal earthquake components becomes neces-
sary. More recently, due to the high levels of vertical earthquake excitation recorded
in Kobe and Northridge, researchers are re-examining the ratio of vertical to horizon-
tal components and the frequency content assumptions [32].

5.7. Effects of tank inertia

EC8 recommends that “For steel tanks, the inertia forces acting on the shell due
to its own mass are small in comparison with the hydrodynamic forces, and can
normally be neglected”. The above statement is not always true. In the presence of
a roof and corresponding snow loading, it may become necessary to include the
inertia effects due to the tank (ASCE, 1984) [18].
332 F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333

Acknowledgements

I would like to thank Professor Roger Hobbs, Head of Engineering Structures


Section at the Civil Engineering Department of Imperial College. This work is part
of a study funded by EPSRC Grant Number GR/K88644.

References

[1] Hanson RD. Behaviour of liquid storage tanks. In: The great Alaska Earthquake of 1964, Committee
on the Alaska Earthquake, Division of the Earth Sciences, National Research Council, National
Academy of Sciences, Washington D.C., 1973:331–39.
[2] Clough DP. Experimental evaluation of seismic design methods for broad cylindrical tanks, Report
UCB/EERC-77/10, 1977.
[3] Steinbrugge KV. Earthquake damage and structural performance in the United States. In: Wiegel
RL, editor. Earthquake Engngr, Chapter 9, 1971:183–226.
[4] Haroun MA. Behaviour of unanchored oil storage tanks: Imperial Valley earthquake. ASCE J Techn
Topics in Civ Egngr 1983;109:23–40.
[5] Moore TA. The response of cylindrical liquid storage tanks to earthquakes. Proc of the Conf on
Large Earthquakes in New Zealand: Anticipation Precaution Reconstruction, 1981:117–31.
[6] Manos GC, Clough RW. Tank damage during the May 1983 Coalinga earthquake. Earthq Egngr
Struct Dyna 1985;13:449–66.
[7] Haroun MA, Badawi HS, Abou-Izzeddine W. Performance of liquid storage tanks during the 1989
Loma Prieta earthquake. Proc 3rd Conf Lifeline Earthq Egngr, 1991:1152–61.
[8] Goltz JD. The Northridge, California earthquake of January 17, 1994, General Reconnaissance
Report.
[9] Hall JF. Northridge earthquake of January 17, 1994 Reconnaissance report Vol 1, Suppl. C. Earthq.
Spectra 11, EERI report.
[10] Haroun MA, Bhatia H. Analysis of tank damage during the 1994 Northridge Earthquake. Proceedings
of the 4th US Conference on Lifeline Earthquake Engineering, ASCE, New York, 1995:763–70.
[11] Eurocode 8. Earthquake resistant design of structures, part 4: tanks, silos and pipelines, 2nd draft,
December 1993.
[12] Niwa A, Clough RW. Buckling of cylindrical liquid-storage tanks under earthquake loading. Earthq
Engngr and Struct Dyna 1982;10:107–22.
[13] Rammerstorfer FG, Scharf K, Fisher FD. Storage tanks under earthquake loading. ASME Applied
Mechanics Review 1990;43(11):261–82.
[14] Wyllie LA, Bolt B, Durkin ME, Gates JH, McCormick D, Smith PD, Abrahamson M, Castro G,
Escalante L, Luft R, Olson RS, Vallenas J. The Chile Earthquake of March 3, 1985. Earthq Spectra
1986;2:373–409.
[15] American Petroleum Institute API650. Welded Steel Tanks for Oil Storage, 1993.
[16] AWWA D100-84. AWWA Standard for Welded Steel Tanks for Water Storage, 1984.
[17] Wozniak RC, Mitchell WW. Basis of seismic design provisions for welded steel oil storage tanks.
Proceedings of Refining Department, API, Washington D.C., 1978:485–501.
[18] ASCE. Guidelines for the seismic design of oil and gas pipeline systems. Technical Council on
Lifeline Earthquake Engineering, Committee on Gas and Liquid Fuel Pipelines, ASCE, New
York, 1984.
[19] Fischer FD, Rammerstorfer FG, Scharf K. Earthquake resistant design of anchored and unanchored
liquid storage tanks under three-dimensional earthquake excitation. In: Schueller GI, editor. Struc-
tural dynamics, chapter 5. Heidelberg, FRG: Springer Verlag, 1991:317–71.
[20] Priestley MJN, Wood JH, Davidson BJ. Seismic design of storage tanks. Bull Of the New Zealand
National Society for Earthquake Engineering 1986;19(4):272–84.
[21] Sakai F, Ogawa H. Seismic resistant design of liquid storage tanks in Japan, 1983:4-17-1–18.
F.H. Hamdan / Journal of Constructional Steel Research 53 (2000) 307–333 333

[22] The Ministry of International Trade and Industry. Standard of Seismic Design of High Pressure Gas
Facilities, The Official Gazette, Extra No. 93, Japan, 1981.
[23] Standard for Seismic Design of Oil Storage Tanks, The Fire Defence Agency of the Ministry of
Home Affairs. Notification Specifying the Details of Technical Standard on the Regulations for
Dangerous Objects, Notification No. 119, Japan, 1983.
[24] Eurocode 8, Design Provisions for Earthquake Resistance of Structures-Part 1-1: General rules-
Seismic actions and general requirements for structures, May 1994.
[25] Manos GC. Influence of the earthquake input characteristics on the dynamic response of a broad
cylindrical liquid storage tank model. Proc of the European Conf on Earthq Engngr 1986;7:81–8.
[26] Hamdan FH. An assessment of EUROCODE 8 Part 4: Design of Liquid Storage Tanks, The Fifth
SECED International Conference on European Seismic Design. In: Balkema AA, Elnashai AS, edi-
tors. Chester, UK, 26–27 October, 1995:521–9.
[27] Hamdan FH. Tank damage during the 1994 Northridge Earthquake: Code comparison and finite
element analyses, Structural Dynamics — Eurodyn’96 Conference, In: Balkema AA, Augusti G, et
al, editors. Florence, Italy, 5–8 June, 1998:105–12.
[28] Maheri MR, Severn RT. Hydrodynamic consideration for seismic design of cylindrical structures.
Proc of the Conf on Civil Engngr Dyna, Organised by the University of Bristol and SECED, Univer-
sity of Bristol, 24–25 March, 1988:297–313.
[29] National Aeronautics and Space Administration, Buckling of Thin-Walled Circular Cylinders, NASA
SP-8007, 1968.
[30] Rotter JM. Local inelastic collapse of pressurised thin cylindrical steel shells under axial com-
pression, Research Report, School of Civil and Mining Engineering, University of Sydney, 1985.
[31] Haroun MA, Housner GW. Earthquake response of deformable liquid storage tanks. ASME J Appl
Mech 1981;48:411–8.
[32] Elnashai AS, Papzoglou AJ. Significance of vertical earthquake motion for RC buildings, Structural
Dynamics — Eurodyn’96 Conference, In: Balkema AA, Augusti G, et al, editors. Florence, Italy,
5–8 June, 1996:129–36.

Вам также может понравиться