Вы находитесь на странице: 1из 24

Journal of Earthquake Engineering, 12:124, 2008 Copyright A.S. Elnashai & N.N.

. Ambraseys ISSN: 1363-2469 print / 1559-808X online DOI: 10.1080/13632460701672714

Journal 1559-808X 1363-2469 UEQE of Earthquake Engineering, Vol. 0, No. 0, Dec 2007: pp. 00 Engineering

A Static Nonlinear Procedure for the Evaluation of the Elastic Buckling of Anchored Steel Tanks Due to Earthquakes
JUAN C. VIRELLA, LUIS E. SUREZ, and LUIS A. GODOY
Department of Civil Engineering and Surveying, University of Puerto Rico at 5 Mayagez, Mayagez, Puerto Rico

J. C. xxxx Virella, L. E. Suarez, and L. A. Godoy

Ground-supported steel tanks experienced extensive damage in past earthquakes. The failure of tanks in earthquakes may cause severe environmental damage and economic losses. This study deals with the evaluation of the elastic buckling of above-ground steel tanks anchored to the foundation due to seismic shaking. The proposed nonlinear static procedure is based on the capacity 10 spectrum method (CSM) utilized for the seismic evaluation of buildings. Different from the standard CSM, the results are not the base shear and the maximum displacement of a characteristic point of the structure but the minimum value of the horizontal peak ground acceleration (PGA) that produces buckling in the tank shell. Three detailed finite element models of tank-liquid systems with height to diameter ratios H/D of 0.40, 0.63, and 0.95 are used to verify the methodology. The 1997 15 UBC design spectrum and response spectra of records of the 1986 El Salvador and 1966 Parkfield earthquakes are used as seismic demand. The estimates of the PGA for the occurrence of first elastic buckling obtained with the proposed nonlinear static procedure were quite accurate compared with those calculated with more elaborate dynamic buckling studies. For all the cases considered, the proposed methodology yielded slightly smaller values of the critical PGA for the first elastic 20 buckling compared to the dynamic buckling results. Keywords Nonlinear Analysis of Tanks; Seismic-Induced Buckling; Steel Tanks; Nonlinear Static Procedure; Buckling of Cylindrical Tanks

1. Introduction
There is ample evidence of cases of buckling failure or other severe damage of ground- 25 supported steel storage tanks during earthquakes [Cooper and Wachholz, 1999; Jain et al., 2001; Suzuki, 2002]. Practically every major earthquake provides further evidence of the vulnerability of these structures. For instance, a recently published report by an EERI reconnaissance team [2003], describes the buckling of two ground supported steel storage tanks during the 2003 San Simeon earthquake in California. In addition to the loss of the 30 structure, the failure of the tanks may have other severe consequences, such as long-term contamination of the soil when there is loss of liquid content and loss of the storage capacity of the plant for a long time. In some cases, water storage tanks are part of fire-fighting systems, and so the interruption of the water supply could result in more damage due to the fires that might follow an earthquake. 35 Early investigations on the seismic behavior of anchored liquid storage tanks allowed us to gain significant insight into the effect of the hydrodynamic fluid-structure interaction
Received 6 March 2006; accepted 27 August 2007. Address correspondence to Luis E. Surez, Department of Civil Engineering and Surveying, University of Puerto Rico at Mayagez, Mayagez, Puerto Rico, 00681-9041; E-mail: Isuarez@uprm.edu

J. C. Virella, L. E. Surez, and L. A. Godoy

on the seismic response. Jacobsen [1949] was the first researcher that studied the seismic behavior of rigid cylindrical tanks. Housner [1963], Haroun and Housner [1981], Veletsos and Yang [1977], and Veletsos [1984] reported that a circular cylindrical tank containing liquid develops a cantilever beam type mode when subjected to horizontal excitations. Housner [1963] evaluated the hydrodynamic response of a tank-liquid system as the contribution of two different components, the convective mode and the impulsive mode. The liquid in the upper portion of the tank vibrates with a long period sloshing motion while the rest moves rigidly with the tank, with an impulsive mode. Several studies have addressed the behavior and design of anchored and unanchored steel cylindrical tanks subjected to earthquakes, and clearly established the importance of the buckling behavior of the tanks in their seismic response [Rammerstorfer et al., 1990; Lau et al., 1996; Liu and Schubert, 1999; Handam, 2000]. The American Lifelines Alliance [2001] has collected different modes of failure that have been observed in tanks during past earthquakes. These failure modes include shell buckling, roof damage, anchorage failure, tank support system failure, foundation failure, hydrodynamic pressure failure, connecting pipe failure, and manhole failure. Among the modes of failure stated, the most relevant for the present study is the shell buckling. Sakai and Isoe [1989] and Rammerstorfer et al. [1988] pointed out that the most important types of damage for tanks subjected to earthquakes are elastic-plastic buckling and elastic buckling of the tank shell. Ito et al. [2003] described the buckling patterns of thin-walled liquid storage tanks as bending buckling, shear and compression buckling, and nonlinear ovaling vibration. The bending buckling occurs at the lower part of the cylindrical shell and it includes the diamond shape buckling and elephant foot buckling (an elastic-plastic buckling type), which is the most frequently encountered buckling mechanism in tanks during earthquakes. The shear and compression buckling is an elastic type of buckling, occurring near the middle of the tank. The nonlinear ovaling vibration is a buckling type which is typically encountered at the top of the tank. Rammerstorfer et al. [1990] described this phenomenon as buckling due to external pressure. Design codes for ground supported storage tanks, such as API 650 [1991] and AWWA-D100 [1984], prescribe the buckling evaluation of steel tanks using an allowable stress formulation based on test results of cylindrical thin shells subjected to uniform compression. This is clearly a different load configuration than the actual earthquake loading that induces buckling. Therefore, it would be of great importance to have a methodology to evaluate the seismic buckling of tank structures for a specific seismic demand that can take into account the hydrodynamic response of the tank-liquid system based on three dimensional models. Buckling analyses of structures can be performed by using static and dynamic buckling techniques. Computational dynamic buckling analyses considering geometric and material nonlinearity have been performed for tank structures subjected to earthquakes. Using experimental or computational methods, Liu and Lam [1983], Natsiavas and Babcock [1987], Nagashima et al. [1987], Redekop et al. [2002], and Morita et al. [2003] reported the occurrence of buckling at the top of tanks induced by horizontal base excitations. In their experimental dynamic buckling studies, Natsiavas and Babcock [1987] considered a tall tank with an open top undergoing a horizontal harmonic base acceleration. Dynamic and static buckling experimental studies of tall tanks with roof were carried out by Nagashima et al. [1987] and Morita et al. [2003]. Nagashima et al. [1987] considered, in addition to the horizontal component, a vertical harmonic base acceleration, while Morita et al. [2003] used both harmonic excitations along with a simulated earthquake excitation. The shell buckling at the top of steel tanks observed after earthquakes has been frequently attributed to the sloshing component of the hydrodynamic response of the

45

50

55

60

65

70

75

80

85

90

xxxx

3 Q1

tank-liquid system [Malhotra, 2000]. Both Ito et al. [2003] and Natsiavas and Babcock [1987] proved that this buckling mode comes mostly from the impulsive action of the hydrodynamic response of the liquid. Although the sloshing action contributes to the occurrence of this type of buckling, it is not the main cause. Because the seismic response is of dynamic nature, dynamic buckling studies can predict more accurately the buckling behavior of a structure subject to an earthquake. However, dynamic buckling analyses are very expensive and thus the buckling evaluation must frequently rely on static analyses. Fischer and Rammerstorfer [1982] and Redekop et al. [2002] used static buckling analyses for the seismic evaluation of steel tanks, and presented the results in terms of the critical spectral acceleration that induces buckling. A simplified nonlinear analysis for the seismic design of liquid storage tanks was proposed by Malhotra [2000] based on the nonlinear static procedure for buildings [ATC-40, 1996]. The design forces were obtained by considering two degrees of freedom beam models, in which the impulsive and convective action were accounted for. Malhotra [2000] studied unanchored and partially unanchored tank-liquid systems and therefore the elastic acceleration-deformation spectrum (demand spectrum) was reduced based on the viscous damping of the tank structure and the hysteretic damping due to the plastic behavior of the base plate. This study presents a methodology to calculate the seismic buckling response of above ground steel tanks, using a three-dimensional model of the tank structure and a procedure based upon the capacity and demand spectra. The seismic buckling response is due to a horizontal earthquake excitation. Only the impulsive component of the hydrodynamic response of the tank-liquid system is considered, i.e., sloshing effects, are ignored. A non-linear static procedure (NLSP) based on the capacity spectrum method (CSM) for building structures [ATC-40, 1996] is proposed, for the evaluation of the first buckling mode of the tank structure (buckling at the top). For any buckling seismic demand, the NLSP renders the spectral acceleration associated with the first buckling mode of the tank. Knowing this quantity, it is possible to obtain the horizontal peak ground acceleration at the critical state and thus the results can be readily compared with dynamic buckling studies.

95

100

105

110

115

2. Nonlinear Static Procedure for Steel Tanks

120

A nonlinear static procedure (NLSP) based on the CSM can be implemented for the seismic evaluation of ground-supported steel tanks, provided that certain modifications are introduced in the CSM to account for the differences between the seismic response of tanks and buildings. One of the main differences is that the fundamental mode for the 125 horizontal base excitation of tank-liquid systems does not always correspond to the first (i.e., highest period) natural mode, as it is usually the case for buildings. Usually the mode shapes of tank-liquid systems that are associated to the highest natural periods correspond to cos(m,n,q) modes that are characterized by a series of circumferential waves (n) and axial half-waves (m). The fundamental modes for the seismic response of tank-liquid 130 systems are those with the largest participation factors to a horizontal base motion. For horizontal motions, the fundamental modes tend to cantilever beam mode shapes [Barton and Parker, 1987]. A more detailed explanation of this concept is presented by Virella et al. [2006a]. Thus, to apply the concepts of the CSM to tanks filled with liquid, the procedure needs to be modified by using the fundamental mode instead of the first mode. 135 Notice also that the CSM applied to buildings is a sort of global collapse analysis; in our case, the CSM is used to study a local behavior. Thus, although there is a small change in periods when local buckling occurs near the top of the tank, this change is restricted to those periods of higher modes. The fact that the buckling is localized does not diminish its

J. C. Virella, L. E. Surez, and L. A. Godoy

importance because the tanks buckling capacity is substantially reduced due to the intro- 140 duction of an imperfection. 2.1. Development of the Capacity Curve To develop the capacity curve for the seismic evaluation of steel tanks, a methodology similar to the load level 3 of the ATC-40 [1996] for buildings is applied, in which the lateral forces (Fx) vary according to the product of the story masses and the first mode shape of the structure. Certain modifications are introduced to adapt the methodology for tank-liquid systems. These capacity curves were obtained by considering only the impulsive mode of the tank-liquid system, ignoring the sloshing effects. The force distribution for the level 3 approach used in the tank structure is produced by the impulsive hydrodynamic pressure distribution due to the fundamental mode of the tank-liquid-system under a horizontal excitation, including the flexibility of the tank. In previous studies, such as those by Veletsos and Shivakumar [1997] and Fischer and Rammerstorfer [1982], an impulsive modal pressure distribution formulation was proposed for the seismic analysis of flexible tanks. Although in principle this formulation could be applied to obtain the capacity curves, in the present study the impulsive pressure distribution for the fundamental mode proposed by Virella et al. [2006a] was used instead. The latter distribution was obtained from threedimensional finite element models in which the liquid was represented with acoustic FE and the tank was modeled with shell elements. Notice that the mass of the liquid and the impulsive pressure are correlated because the added mass of the liquid is calculated in terms of the impulsive pressure. The calculation of the added mass and impulsive pressure is explained in Virella et al. [2005]. The capacity curve is obtained from a step-by-step nonlinear static buckling analysis in which geometric and material nonlinearity are considered. Material nonlinearity is considered by means of an elastic-plastic model with the von Mises yield criterion in which the maximum value reached by the von Mises stresses is the yield stress, sy = 248 MPa. The pressure distribution used in the analysis is normalized for a base acceleration of 1 g. The buckling load estimate corresponds to the sum of any initial loadls applied to the structural system before the static buckling step, plus the loads applied in the buckling step multiplied by a load factor l, i.e., 145

150

155

160

165

170

Applied loads to estimate buckling = initial loads + l (loads applied in second step) (1)
The initial loads are the hydrostatic pressure and the self-weight of the tank, which were defined in ABAQUS [2002] in an initial step by means of a static analysis. Then, a step- 175 by-step nonlinear analysis was performed, in which the hydrodynamic impulsive pressure was introduced. The load factor, l, obtained from the buckling analysis then corresponds to the fraction of the impulsive pressure required to buckle the tank. The Modified Riks Method is used for the step-by-step nonlinear analyses in ABAQUS [2002], from which the nonlinear response of the structure is obtained, by 180 applying increments of load and evaluating equilibrium. In the modified Riks algorithm [Riks, 1979], the load factor l is an additional unknown and the arc length along the static equilibrium path of the load-displacement space is used to measure the progress of the solution. As the arc length is found, the static equilibrium equations are solved stepby-step, considering load increments, whereas the stiffness of the system decreases 185

xxxx

successively in buckling problems. The magnitudes of the applied loads in Eq. (1) are found for each step, and so are the displacements. The process continues until a maximum predefined number of total steps is reached, or a predefined displacement is found at a prescribed point in the shell. With the values of l and the displacement at a point of the shell where buckling occurred, a graph of the load multiplier l vs. the radial (i.e., out of plane) 190 displacements () is prepared. We consider the critical load multiplier lcrit as the last load multiplier l found before a negative load increment l occurs. Figure 1(a) shows a l vs. curve. The l vs. curve is equivalent to the pushover or capacity curve (i.e., base shear vs. roof displacement) used for buildings in the CSM. Since the applied pressure distribution, i.e., the loads that multiplied l in Eq. 1 was normalized for a base acceleration of 1 g, the 195 critical load factor lcrit can be associated to the critical spectral acceleration for the tank, as explained in the next section. 2.2. Procedure to Find the Performance Point 2.2.1. The Capacity Spectrum. The participation factor PFf and the modal mass coeffi- 200 cient af for the fundamental mode (instead of those for the first mode) will be used to transform the capacity curve to modal coordinates. These two modal parameters are obtained from eigenvalue analyses of the tank-liquid systems [Virella et al., 2006a]. Because for the analysis the impulsive pressure is normalized at an acceleration of 1 g, the load factor l in the capacity curve is a multiplier of that acceleration. Then the load 205 factor is analogous to the base shear to weight of the structure ratio for buildings (V/W). The l vs. curve is then transformed into a spectral acceleration (Sa) vs. spectral displacement (Sd) curve as shown in Fig. 1. The spectral acceleration Sa can be expressed as:

Sa =

l . af

(2)

Also, the spectral displacement is obtained from a critical node in the zone of buckling as 210

Sd =

crit PFf fcrit , f

(3)

where crit is the radial displacement at the critical node where buckling occurs and fcrit, f is the modal coordinate of the fundamental mode shape at the location of the critical node.

crit Load Factor, Established for first negative obtained

crit Radial displacement Crit Node,

(a)

(b)

FIGURE 1 Capacity curve for tank structures. (a) l vs. format. (b) Sa vs. Sd format.

J. C. Virella, L. E. Surez, and L. A. Godoy

(a)

(b)

FIGURE 2 Demand Spectrum for tank structures. (a) Sa vs. T format. (b) Sa vs. Sd format. 2.2.2. The Ground Response Spectrum. To define the seismic demand, and as recom- 215 mended by Fischer and Rammerstorfer [1982], a response spectrum for a damping ratio of 2% is used instead of the usual 5% damping response spectrum adopted for buildings (see Fig. 2a). In addition to a design spectrum typical of those prescribed in seismic codes, ground response spectra associated with accelerograms of historic earthquakes are used in 220 this article. The latter permits an easier comparison of the results obtained from the CSM with those from nonlinear time history analysis. The response spectrum is transformed to the spectral acceleration (Sa) vs. spectral displacement (Sd) format ADRS spectra as illustrated in Fig. 2. 2.2.3. Performance Point. The primary interest for the seismic evaluation of anchored tanks is to determine the first buckling mode. Thus, contrary to the CSM for buildings, there is no need to reduce the elastic demand curve (i.e., the response spectrum). It should be pointed out that steel tanks in the nuclear industry are designed for an elastic seismic spectrum [Sugiyama et al., 2003]. Then, the performance objective here is not to find the nonlinear deformation of tank-liquid systems, but to obtain instead the horizontal peak ground acceleration for which these structures would buckle elastically during a seismic event. In other words, instead of finding a performance point for reduced demand spectra, the quantity sought is the maximum demand that can be resisted elastically by the tankliquid system until buckling occurs. The evaluation of the response of the steel tanks is done first using a smooth design spectrum as seismic demand. In this study, the elastic design spectrum of the UBC code [1997] was chosen to show the implementation of the methodology. The 1997 UBC spectrum is defined by two seismic coefficients, Ca and Cv, that depend on the seismic zone and soil profile. To implement the CSM, the parameters Ca and Cv are varied until the demand spectrum intersects the capacity spectrum at the point where the first buckling of the tank occurs. Proceeding in that way, and depending on which zone of the design spectrum is intersected by the capacity spectrum, either the critical coefficients Ca or Cv can be found. If the seismic coefficient, Ca is obtained from the procedure, it is directly related to the horizontal peak ground acceleration (PGA) that the tank-liquid system can resist. If on the other hand the coefficient Cv is found, then there are several PGAs associated with this parameter, depending on the seismic zone and soil profile. It should be noticed that the spectral accelerations for high periods do not correlate well with the PGA. However, practically all anchored steel tanks have relatively small fundamental periods (i.e., less than 0.4). Thus we are always in the plateau of the design spectrum, which is defined in terms of Ca in the UBC-97 spectrum used in the article. As it is well known, in many codes (such as the UBC-97) the plateau is assumed to be 225

230

235

240

245

250

xxxx

Sa elastic Spectral Acceleration, Sa Sa-crit P.P Performance point P.P. Demand spectra reduced to increase to reach Sa-crit

Sd crit Spectral Displacement, Sd

FIGURE 3 Location of performance point. 2.5 times the PGA. Moreover, in our analyses the natural periods of the tanks remain equal to the elastic periods, at least until first local buckling occurs. When a response spectrum corresponding to the accelerogram of a historic or artificial 255 earthquake is used, the PGA that causes the tank to buckle elastically can be obtained indirectly as follows. The performance point is obtained first as the critical spectral acceleration in the capacity spectrum associated with the occurrence of the first buckling in the system. The elastic response spectrum is then calculated by scaling the original accelerogram with an initial trial scale factor. By scaling up or down, this response spectrum in the ADRS 260 format until it intersects the demand spectrum precisely at the performance point, one can obtain the scale factor associated with the critical PGA. Multiplying the PGA of the original accelerogram by this scale factor one can obtain the minimum PGA that induces the first buckling in the tank-liquid system. This process is illustrated graphically in Fig. 3.

3. Numerical Examples
3.1. Tank Models

265

The NLSP is applied for the seismic evaluation of three tanks with different geometries. The detailed description of the geometries and models used can be found in Virella et al. [2005]. The finite element models of these tanks are identified as Models A, B, and C, and 270 they correspond to tanks with height to diameter ratios (H/D) of 0.40, 0.63, and 0.95, respectively (see Fig. 4). These cylindrical tanks have a clamped condition at the base, with a cone roof supported by roof rafters. The tapered thicknesses of the tanks illustrated in Fig. 4 were chosen for serviceability conditions using the API 650 provisions [1991], without taking any seismic design considerations into account. The bottom plate of the 275 tank was not modeled since only anchored tanks are considered in this study and our primarily interest lies on the buckling of the cylinder shell. Because the tanks are assumed to be anchored, this condition will prevent the deflections of the plate and thus it is not necessary to include it in the FE model. The liquid height of 90% the height of the cylinder was used for all the tank models. 280 The rationale for considering that the tanks were filled up to a height of 0.9 times the height H of the cylinder is based on the fact that design codes for such as API 650 [1991] always require a freeboard, so that if a sloshing wave is formed, it will not come in contact with the tank roof. The amplitude of the sloshing wave needs to be computed for a specific

J. C. Virella, L. E. Surez, and L. A. Godoy


slope = 3V:16H
tr

2.858 m
t = 0.0079 m t = 0.0079 m t = 0.0079 m t = 0.0095 m t = 0.0127 m

2.416 m 2.416 m 2.416 m 2.416 m 2.425 m D = 30.48 m

H = 12.191m

(a)
slope = 3V:16H
tr

2.858 m
t = 0.0079 m t = 0.0095 m t = 0.0095 m t = 0.0127 m t = 0.0127 m t = 0.0159 m t = 0.0191 m t = 0.0191 m

2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.425 m D = 30.48 m

H = 19.337 m

(b)
slope = 3V:16H 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.416 m 2.425 m D = 30.48 m
tr t = 0.0079 m t = 0.0095 m t = 0.0095 m t = 0.0127 m t = 0.0127 m t = 0.0159 m t = 0.0191 m t = 0.0191 m t = 0.0222 m t = 0.0254 m t = 0.0254 m t = 0.0286 m

2.858 m

H = 29.103 m

(c)

FIGURE 4 Tank models with cone roof supported by rafters; t = shell thickness, roof thickness tr = 0.00635 m (for roof with rafters). (a) Tank with H/D = 0.40; (b) Tank with H/D = 0.63; (c) Tank with H/D = 0.95. seismic demand, but in previous studies [Virella et al., 2005, 2006a, b] we have found 285 adequate to use 10% of the height of the cylinder as freeboard. The finite element package ABAQUS [2002] was used to perform the computations. The finite element meshes of the tank models consisted of S3R triangular elements for the roof, S4R quadrilateral elements for the cylinder, and B31 beam elements for the rafters in the roof. The S4R element is a four node, doubly curved shell element with reduced 290

xxxx

integration, hourglass control and finite membrane strain formulation. The S3R element is a three node degenerated version of the S4R with finite membrane strain formulation. The B31 is a two node linear beam element in space. All of these elements are described in more detail in Hibbit et al. [2002]. The finite element meshes consisted of 9,262 elements for Model A and 10,942 elements for Models B and C. 295 3.2. Capacity Curves The capacity curves were obtained using the level 3 loading distribution as described previously. The fundamental modes for the impulsive component of the hydrodynamic response of tank-liquid systems subjected to horizontal base motions were obtained in a previous study by Virella et al. [2006a]. Figures 57 show the fundamental impulsive 300

(a)
14.0 12.0 10.0 8.0 Height [m] 6.0 H HL 4.0
3 2 Xg
mode: n = 1 Meridian in Figure

Xg
U-radial variation with circumference

2.0 0.0 3.0 2.0 1.0

Smooth deformed shape of cylinder 2.0 3.0

0.0 1.0 Normalized U-radial

(b)

FIGURE 5 Fundamental mode for Model A. (a) 3D view; (b) Deformed shape in the meridian with maximum displacements.

10

J. C. Virella, L. E. Surez, and L. A. Godoy

(a)
25.0
mode: n = 1 Meridian in Figure

20.0

Xg U-radial variation with circumference

Height [m]

15.0

10.0 H HL 5.0
3 2 Xg

0.0 3.0 2.0 1.0 0.0 1.0 Normalized U-radial

Smooth deformed shape of cylinder 2.0 3.0

(b)

FIGURE 6 Fundamental mode for Model B. (a) 3D view; (b) Deformed shape in the meridian with maximum displacements. modes for Models A, B, and C, and Table 1 presents the modal parameters that characterize the fundamental mode of each of the tank-liquid systems. The pressure distributions for the fundamental impulsive mode of the three tankliquid systems at the meridian in the direction of the excitation were obtained from Virella et al. [2006a] and they are presented in Fig. 8. 305 A step-by-step static nonlinear analysis was performed for all the models to find the elastic buckling. For Model A, buckling occurred at a region extending from about half the height up to the top of the cylinder, while for Models B and C buckling occurred at a region in the shell near the top of the cylinder. Figure 9 shows the buckling modes found. The node used to calculate the capacity curves for the systems is also indicated in the 310 figure. The capacity curves for the tank-liquid systems are presented in Fig. 10. As indicated in the figure, elastic buckling occurred for load factors equal to = 0.556 for Model A, = 0.476 for Model B, and = 0.370 for Model C. Therefore, the load factor of the tank decreased with an increase in the H/D ratio. The capacity curves are transformed into capacity spectra with the help of Eqs. (2) 315 and (3). The modal parameters of the fundamental mode needed for the transformation are presented in Table 1.

xxxx

11

(a)
35.0 30.0 25.0 Height [m] 20.0 15.0 H HL 10.0
3 2 Xg
mode: n = 1 Meridian in Figure

Xg U-radial variation with circumference

5.0 0.0 1.0 1.0 Normalized U-radial

Smooth deformed shape of cylinder 3.0 5.0

5.0

3.0

(b)

FIGURE 7 Fundamental mode for Model C. (a) 3D view; (b) Deformed shape in the meridian with maximum displacements. TABLE 1 Modal parameters for the fundamental modes of the tank-liquid systems Model A B C H/D 0.40 0.63 0.95 Tf[s] 0.212 0.200 0.196 af 0.64 0.79 0.77 PFf 1888 1.5545 1.582 fcrit,f 0.0003803 0.67349 0.8678 PFf* fcrit,f 0.718 1.047 1.373

3.3. Demand Spectra This section presents the elastic design spectrum and the ground response spectra from earthquakes records used for the seismic buckling evaluation of the steel tanks. The design spectrum 320 is the 1997 UBC elastic spectrum presented in Fig. 11. This spectrum is for a site with a zone factor (Z) of 0.30, a soil profile type SB (rock), an importance factor (I) of 1.0, and a damping ratio of 2%. The seismic coefficients Ca and Cv are equal to 0.39 and 0.37, respectively.

12

J. C. Virella, L. E. Surez, and L. A. Godoy


25.0 14.0
mode: n = 1 Meridian in Figure Xg mode: n = 1 Meridian in Figure

12.0 10.0 8.0 H HL 6.0 4.0

20.0

Xg U-radial variation with circumference

U-radial variation with circumference

15.0 Height [m] Height [m] 0.0

H HL

10.0

5.0
2 3 Xg

2 Xg

2.0 0.0 3.0

0.0 3 0.0 1.0 2.0 1.0 1.0

2.0

1.0

Normalized pressure (a)


35.0
mode: n = 1 Meridian in Figure

Normalized pressure (b)

30.0 25.0 20.0 HL 15.0 10.0 Height [m] 0.0

Xg U-radial variation with circumference

2 3 Xg

5.0

0.0 3.0 2.0 1.0 1.0

Normalized pressure (c)

FIGURE 8 Pressure distributions at the meridian with maximum displacements for the fundamental mode. (a) Model A, (b) Model B, (c) Model C.

In addition, the accelerograms from the 1986 El Salvador earthquake and from the 1966 Parkfield earthquake in California shown in Fig. 12 were also used for the seismic 325 buckling evaluation of the tanks. The 1986 El Salvador earthquake had a moment magnitude of 5.6, an epicentral distance of 4.3 km, a focal depth of 7.3 km, and a total duration of 9.04 s. The accelerogram used was recorded in rock at the Geotechnical Investigation Center (CIG) station, and it has a PGA of 0.69 g. The 1966 Parkfield earthquake had a magnitude of 5.5, a focal depth of 6 km, an epicentral distance of 330 27 km, and a total duration of about 30 s. This accelerogram was also recorded in rock and has a PGA of 0.27 g. The 2% damping ratio response spectrum was computed for both accelerograms. The response spectra were smoothed by using a running average with 5 points. The smooth spectra are presented in Fig. 13. 335

xxxx

13

Critical node

Critical node

(a)

(b)

Critical node

(c)

FIGURE 9 First buckling mode for the tank-liquid systems, from step-by-step static nonlinear analyses. (a) Model A, (b) Model B, (c) Model C. Notice that the methodology does not have restrictions regarding the specific ground motion that can be used in the analysis (regarding their duration, frequency content, etc.). However, for the validation of the present approach there was a limitation in the results available that could be used for a comparison. The results of dynamic buckling analyses performed by Virella et al. [2006b] were used for comparison. These analyses were based 340 on two earthquake ground motions: El Salvador (1986) and Parkfield (1966). 3.4. Nonlinear Static Procedure This section discusses the evaluation of the elastic buckling for Models A, B, and C for the seismic demand defined in the previous section. First, the performance point is found for the 1997 UBC elastic spectrum, and then results are presented by considering the response 345 spectra from the earthquake accelerograms. The first sets of results are obtained for the 1997 design spectrum shown in Fig. 11. Both the capacity and demand spectrum were plotted in the ADRS format and the iterative procedure mentioned in a previous section was applied to find the performance point. The performance point for Models A, B, and C subjected to the seismic demand represented by 350 the 1997 UBC design spectrum is presented in Fig. 14. It can be observed that the critical spectral acceleration (whose values are shown in the figures) diminishes with the H/D ratio. The coordinates (Sa, Sd) of the performance point, the physical radial displacement , and the seismic coefficient Ca for the three tanks are presented in Table 2. Notice that the seismic coefficient Ca decreases with the H/D ratio of the tanks while the radial 355 displacement at the zone of buckling increases with H/D. The next sets of results correspond to the response spectra of the actual earthquake records shown in Fig. 13. The performance points for the tank models A, B, and C

14

J. C. Virella, L. E. Surez, and L. A. Godoy


0.60 0.50 Model A, crit = 0.556 Model B, crit = 0.476 Model C, crit = 0.370

Load factor,

0.40 0.30 0.20 0.10

0.00 0.002

0.003

0.008

0.013

0.018

0.023

Radial displacement [m] (a)


0.60 0.50 Model A, crit = 0.556 Model B, crit = 0.476 Model C, crit = 0.370

Load factor,

0.40 0.30 0.20 0.10 0.00 0.000

0.005

0.010

0.015

0.020

0.025

Radial displacement [m] (b)

FIGURE 10 Capacity curves for the tank-liquid systems. (a) Radial displacements from original configuration; (b) Radial displacements from static equilibrium position.

1.2 1.0 0.8 Sa, g 0.6 0.4 0.2 0.0 0.0 0.5 1.0 T, s 1.5 2.0 2.5

FIGURE 11 The 1997 UBC elastic spectrum for 2% damping ratio.

subjected to the 1986 El Salvador and 1966 Parkfield records are displayed in Figs. 15 and 16. The peak ground acceleration PGA that will induce the first buckling mode is illus- 360 trated in the figures. It was found that in these cases also the PGA associated with the first elastic buckling mode decreases with H/D. Tables 3 and 4 summarize the results of the seismic evaluation of the tank-liquid systems subjected to the El Salvador and Parkfield earthquake records. Note that as shown in Tables 3 and 4, the ratio between the maximum displacement at the buckling zone and the shell thickness (crit /tshell) increases with H/D. 365

xxxx
0.60 0.40

15

Acceleration, g

0.20 0.00 0.20 0.0 0.40 0.60 0.80 2.0 4.0 6.0 8.0 10.0

time, s (a)

0.30 0.20

Acceleration, g

0.10 0.00

0.0 0.10
0.20 0.30

5.0

10.0

10.5

20.0

20.5

30.0

time [s] (b)

FIGURE 12 Earthquake records used in this study. (a) 1986 El Salvador earthquake, PGA = 0.69 g; (b) 1966 Parkfield earthquake, PGA = 0.27 g. The critical horizontal PGAs for all the tank models, calculated with the NLSP are presented in Fig. 17. The results obtained by using the 1997 UBC design spectrum are included in the same figure. The seismic coefficient Ca is related to the PGA (is the spectral acceleration corresponding to a period zero). Regardless of the seismic event considered, the PGA that induces the first elastic buckling is very close for Models B 370 (H/D = 0.63) and C (H/D = 0.95). For Model A (H/D = 0.40), the critical PGA is significantly smaller for the Parkfield record than for the other two ground motions. 3.5. Comparisons with Dynamic Buckling Analyses The PGAs that cause elastic buckling obtained from the NLSP is an approximation to the true values. Therefore, it is important to verify the accuracy of the proposed numerical 375 procedure. In this study, the accuracy of the results were verified by a comparison with full dynamic time history buckling analysis using finite element models that account for geometric and material nonlinearities and the presence of liquid. To obtain the values of the critical PGA in Fig. 18 for a given tank model, between 80 and 100 computer hours are required in a Pentium IV computer with a 3.0 GHz processor. This lengthy process was 380 carefully undertaken to obtain the most accurate possible results until an experimental test program can be implemented. These comprehensive dynamic analyses can be used to provisionally validate the proposed NLSP. In theory, the method could also be verified by comparing its results with those from experimental works by other researchers. However, this is not a straightforward task because there are very few experimental studies on the 385 dynamic buckling of liquid storage steel tanks subjected to earthquakes. The dynamic buckling analysis of the tank liquid systems was described in detail in an article by Virella et al. [2006b]. The critical PGA obtained from the analysis was

16
2.5 2.0 1.5 1.0 0.5 0.0

J. C. Virella, L. E. Surez, and L. A. Godoy

Sa, g

0.0

0.2

0.4

0.6

0.8 T, s

1.0

1.2

1.4

1.6

(a)
1.2 1.0 0.8 Sa, g 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 T, s 1.0 1.2 1.4 1.6

(b)

FIGURE 13 Response spectra from earthquake accelerograms for 2% damping ratio. (a) 1986 El Salvador earthquake record; (b) 1966 Parkfield earthquake record. calculated by Virella et al. [2006b] for the same tank-liquid systems considered in this article. The accelerograms for the 1986 El Salvador and 1966 Parkfield earthquakes were used for the comparison. As shown in Figs. 18 and 19, the PGA obtained from the proposed NLSP and the values predicted by the more accurate dynamic buckling analyses are reasonably similar for all the tank-liquid systems. Notice also that the PGAs for the first elastic buckling predicted by the procedure proposed in this paper are either about the same or smaller than those calculated from dynamic buckling analyses. This occurred for all the tank-liquid systems and for both accelerograms. Therefore, it seems that the modified CSM procedure not only predicted well the elastic buckling acceleration of the tanks, but also provided conservative estimates when the differences were larger, at least for the tank geometries investigated in this paper. Notice that the PGA for the first elastic buckling mode obtained for both the UBC 97 design spectrum and the response spectra of the two accelerograms, the PGA for the first buckling mode decreased with the H/D ratio of the tank. This trend is the same as that observed for other types of buckling modes induced by dynamic loads, as reported by Rammerstorfer et al., [1990, p. 275]. The deformed shapes for the first dynamic buckling mode for models A, B, and C subjected to the seismic demand from the 1986 El Salvador accelerogram are presented in Fig. 20. The static and dynamic buckling analyses yielded similar buckling modes as it

390

395

400

405

xxxx
1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.00 P.P. - first buckling PSAcrit = 0.87g

17

Sa, g

0.05

0.10 Sd, m (a)

0.15

0.20

0.70 0.60 0.50

P.P. - first buckling PSAcrit. = 0.60g

Sa, g

0.40 0.30 0.20 0.10 0.00 0.00 0.02 0.04 0.06 Sd, m (b) 0.6 0.5 0.4 P.P. - first buckling PSAcrit. = 0.48g 0.08 0.10 0.12

Sa, g

0.3 0.2 0.1 0.0 0.00

0.02

0.04 Sd, m (c)

0.06

0.08

0.10

FIGURE 14 Performance point for the 1997 UBC elastic spectrum obtained with the NLSP. (a) Model A, (b) Model B, (c) Model C.

TABLE 2 Results for the tank models for the seismic demand from the 1997 UBC elastic spectrum Model A B C H/D 0.40 0.63 0.95 Critical Sa [g] 0.87 0.60 0.48 Critical Sd [cm] 0.98 1.66 1.54 -radial [cm] 0.70 1.76 2.11 Ca [g] 0.35 0.24 0.19

18

J. C. Virella, L. E. Surez, and L. A. Godoy


1.2 1.0 0.8 Sa, g 0.6
Reduced spectrum

P.P.- first buckling

0.4 0.2 0.0 0.000

PGA = 0.33g Capacity spectrum

0.025

0.050

0.075 Sd, m
(a)

0.100

0.125

0.150

0.70 0.60 0.50


P.P.- first buckling

Sa, g

0.40 0.30 0.20 0.10 0.00 0.00


Reduced spectrum PGA = 0.21g Capacity curve

0.02

0.04

0.06

0.08

0.10

Sd, m
(b)

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.00 0.02 0.04 Sd, m
(c)
Capacity spectrum Reduced spectrum PGA = 0.195g P.P.- first buckling

Sa, g

0.06

0.08

0.10

FIGURE 15 Performance point for the 1986 El Salvador accelerogram, obtained with the NLSP. (a) Model A, (b) Model B, (c) Model C. can be verified by comparing Figs. 9 and 20. For all the tank-liquid systems, buckling occurred near the top of the cylinder as indicated in these figures. 410 A few comments may help to better understand the significance of the results presented. The buckling at the top is very common on anchored tanks and it usually occurs before the better known elephant foot buckling. As a matter of fact, the computer model used in our study does take into account the possibility of occurrence of elephant foot buckling, but this only occurs in an advanced post-critical state and the buckling at the top 415 occurs first. The first buckling occurrence, which is the object of study of this paper, is not

xxxx
1.0 0.8 Sa, g 0.6 0.4 0.2 0.0 0.000
Reduced spectrum PGA = 0.23g Capacity spectrum P.P. - first buckling

19

0.010

0.020 Sd, m (a)

0.030

0.040

1.0 0.8 0.6 Sa, g 0.4 0.2 0.0 0.00


Reduced spectrum PGA = 0.22g Capacity curve P.P.- first bucling

0.02

0.04 Sd, m (b)

0.06

0.08

0.8 0.7 0.6 0.5 Sa, g 0.4 0.3 0.2 0.1 0.0 0.00 0.02 0.04 Sd, m (c) 0.06 0.08
P.P.- first buckling Reduced spectrum PGA = 0.18g Capacity spectrum

FIGURE 16 Performance point for the 1966 Parkfield accelerogram, obtained with the NLSP. (a) Model A, (b) Model B, (c) Model C. TABLE 3 Horizontal peak ground acceleration and radial displacements for the tank models using the NLSP and the 1986 El Salvador accelerogram Model A B C Accelerogram PGA [g] 0.69 0.69 0.69 Scale factor Critical PGA [g] 0.48 0.30 0.28 0.330 0.210 0.195 crit. [cm] 0.70 1.76 2.11 crit/tshell 0.89 1.85 2.22

20

J. C. Virella, L. E. Surez, and L. A. Godoy

TABLE 4 Horizontal peak ground acceleration and radial displacements for the tank models using the NLSP and the 1966 Parkfield accelerogram Model A B C Accelerogram PGA [g] Scale factor Critical PGA [g] crit. [cm] 0.269 0.269 0.269 0.85 0.81 0.68 0.228 0.217 0.182 0.70 1.76 2.11 crit/tshell 0.89 1.85 2.22

Static buckling - NLSP 0.40 1997 UBC spectrum 0.35 Static buckling - NLSP 0.30 El Salvador (1986) 0.25 0.20 0.15 Static buckling - CSM 0.10 Parkfield (1966) 0.05 0.00 0.00 0.20 0.40 0.60 0.80 1.00 1.20

PGA [g] or Ca

H/D

FIGURE 17 Horizontal peak ground acceleration PGA obtained with the NLSP.
0.35 0.30 0.25 PGA [g] 0.20 0.15 0.10 0.05 0.00 0.00 0.20 0.40 0.60 H/D 0.80 1.00 1.20 Static buckling-NLSP El Salvador (1986) Dynamic buckling El Salvador (1986)

FIGURE 18 Comparison of the NLSP and dynamic buckling results, for the accelerogram from the 1986 El Salvador earthquake. necessarily the damaged shape that is usually observed on tanks after an earthquake because after the elastic buckling occurs the post-buckling behavior and plasticity take place and the final deformed shape of the structure may be different from the initial elastic buckling deformed shape. 420 Notice also that unlike silo structures, which are built with a constant thickness, storage tanks are constructed with a tapered wall. The wall thickness at the bottom of the structure can be about twice the thickness at the top courses. As a consequence, buckling will usually initiate at the thinner zones of the cylindrical shell, before occurring at the lower stiffer parts. 425

4. Conclusions
This article introduced a method for the evaluation of the buckling of steel tanks subjected to a horizontal seismic base motion. The method is based upon a static buckling analysis

xxxx
0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0.00

21

Dynamic buckling Parkfield (1966)

PGA [g]

Static buckling-NLSP Parkfield (1966) 0.20 0.40 0.60 H/D 0.80 1.00 1.20

FIGURE 19 Comparison of the NLSP and dynamic buckling results, for the accelerogram from the 1966 Parkfield earthquake.

(a)

(b)

(c)

FIGURE 20 First buckling mode for the tank-liquid systems obtained from dynamic buckling analyses [Virella et al., 2006a]. (a) Model A, (b) Model B, (c) Model C.

22

J. C. Virella, L. E. Surez, and L. A. Godoy

instead of the much more expensive dynamic buckling analyses. The proposed method is 430 based on concepts similar to those used in the CSM applied for the seismic evaluation of buildings. For a given response spectrum, the method allows one to calculate the critical horizontal PGA that induces buckling in the tank shell. The results were compared with those obtained from full dynamic buckling analyses to assess the accuracy of the results. Three typical steel tanks with decreasing height to diameter ratios, identified as 435 models A, B, and C, were used to provide numerical examples of the implementation of the proposed procedure. The PGAs that induce the first elastic buckling in the tank shells obtained with the NLSP were similar but lower than those from the dynamic buckling studies, in all cases considered. Therefore, the method is able to provide conservative estimates of the critical peak ground acceleration for the tank geometries considered in 440 this article. In addition, similar first buckling modes were observed by using both static and dynamic buckling analyses. In the three tank models, buckling occurred at a region near the top of the cylinder. The numerical results show that the PGA for the first elastic buckling mode decreases with the H/D ratio, regardless of the seismic demand considered.

Acknowledgments
J. C. Virella was supported by a PR-EPSCOR post-doctoral fellowship grant EPS-0223152 for this research. L. A. Godoy thanks the support of the Mid-America Earthquake Center. We would also like to thank the two anonymous reviewers for their contribution which improved the explanation of several concepts presented in the article.

445

References
ABAQUS Standard User Manual [2002] Version 6.4. Hibbit, Karlsson and Sorensen, Inc., Pawtucket, Rhode Island. American Lifelines Alliance [2001] Seismic fragility formulations for water systems (ASCE) Part 1 - Guidelines, Part 2 - Appendices, pp. 6079. API 650 [1991] Welded steel tanks for oil storage, American Petroleum Institute, Standard 650, Washington, D.C. ATC-40 [1996] Seismic evaluation and retrofit of concrete buildings, Applied Technology Council, Redwood City, California. AWWA D100 [1984] AWWA standard for welded steel tanks for water storage, American Water Works Association, Denver, Colorado. Barton, D. C. and Parker, J. V. [1987] Finite element analysis of the seismic response of anchored and unanchored liquid storage tanks, Earthquake Engineering and Structural Dynamics 15(3), 299322. Cooper, T. W. and Wachholz, T. P. [1999] Optimizing post-earthquake lifeline system reliability, Proceedings of the 5th U.S. Conference on Lifeline Earthquake Engineering (ASCE), Seattle, Washington, 16, 878886. EERI Reconnaissance Team [2003] The San Simeon, California, earthquake, December 22, 2003, EERI Learning from Earthquakes Reconnaissance Report, Earthquake Engineering Research Institute, Oakland, California, 2344. Fischer, D. F. and Rammerstorfer F. G. [1982] The stability of liquid-filled cylindrical shells under dynamic loading. In: Buckling of Shells, Ramm, E., Ed., Springer, New York, 569597. Handam, F. H. [2000] Seismic behavior of cylindrical steel storage tanks, Journal of Construction Steel Research 53, 307333. Haroun, M. A. and Housner, G. W. [1981] Earthquake response of deformable liquid storage tanks, Journal of Applied Mechanics (ASME) 48(2), 411418. Hibbit H. D., Karlsson, B. I., and Sorensen, P. [2002] ABAQUS Theory Manual, Version 6.4, Pawtucket, Rhode Island.

450

455

460

465

470

475

xxxx

23

Housner, G. W. [1963] The dynamic behavior of water tanks, Bulletin of the Seismological Society of America 53(2), 381389. Ito, T., Morita, H., Hamada, K., Sugiyama, A., Kawamoto, Y., Ogo, H., and Shirai, E. [2003] Investigation on buckling behavior of liquid storage tanks under seismic excitation (1st report: Investigation on elephant foot bulge), Proceedings of the Pressure Vessels and Piping Conference (ASME), Cleveland, Ohio, 466, 193201. Jacobsen, L. S. [1949] Impulsive hydrodynamics of fluid inside a cylindrical tank and of fluid surrounding a cylindrical pier, Bulletin Seismological Society of America 39(3), 189203. Jain S. K., Lettis, W. R., Murty, C. V. R., and Bardet, J. [2002] Bjuh, India Earthquake of January 26, 2001, Reconnaissance Report, Earthquake Spectra, Supplement A to volume 18, 257295. Lau, D. T., Marquez, D., and Qu, F. [1996] Earthquake resistant design of liquid storage tanks, Proceedings of the Eleventh World Conference on Earthquake Engineering, Acapulco, Mexico, Paper no. 1293. Liu, W. K. and Lam, D. [1983] Nonlinear analysis of liquid filled tank Journal of Engineering Mechanics (ASCE) 109(6), 13441357. Liu, He. and Schubert, D. H. [1999] Seismic behavior of anchored, partially anchored and unanchored liquid-storage tanks, Proceedings of the 8th Canadian Conference on Earthquake Engineering, Vancouver, Canada, 329334. Malhotra, P. K. [2000] Practical nonlinear seismic analysis of tanks, Earthquake Spectra 16 (2), 473492. Morita, H., Ito, T., Hamada, K., Sugiyama, A., Kawamoto, Y., Ogo, H., and Shirai, E. [2003] Investigation on buckling behavior of liquid storage tanks under seismic excitation: 2nd report Investigation on the nonlinear ovaling vibration at the upper wall, Proceedings of the Pressure Vessels and Piping Conference (ASME), Cleveland, Ohio, 466, 227234. Nagashima, H., Kokubo, K., Takayanagi, M., Saitoh, K., and Imaoka, T. [1987] Experimental study on the dynamic buckling of cylindrical tanks (Comparison between static buckling and dynamic buckling), JSME International Journal 30 (263), 737746. Natsiavas, S. and Babcock, C. D. [1987] Buckling at the top of a fluid-filled tank during base excitation, Journal of Pressure Vessel Technology (ASME) 109, 374380. Rammerstorfer, F. G., Scharf K., Fisher F. D., and Seeber, R. [1988] Collapse of earthquake excited tanks, Res Mechanica Journal 25, 129143. Rammerstorfer, F. G., Scharf, K., Fisher, F. D., and Seeber, R. [1990] Storage tanks under earthquake loading, Applied Mechanics Reviews 43(11), 261282. Redekop, D., Mirfakhraei, P., and Muhammad, T. [2002] Nonlinear analysis of anchored tanks subject to equivalent seismic loading, Proceedings of the Pressure Vessels and Piping Conference (ASME), Vancouver, Canada, 442, 157163. Riks, E. [1979] An incremental approach to the solution of snapping and buckling problems, International Journal of Solids and Structures 15, 524551. Sakai, F. and Isoe, A. [1989] Computation and experiment on base-uplift behavior of cylindrical oil storage tanks during earthquakes, Proceedings of the Pressure Vessel and Piping Conference (ASME), Honolulu, Hawaii, 157, 914. Sugiyama, A., Setta, K., Kawamoto, Y., Hamada, K., Morita, H., Ito, T., Ogo, H., and Shirai, E. [2003] Investigation on buckling behavior of liquid storage tanks under seismic excitation (3rd report; proposed design procedure considering dynamic response reduction), Proceedings of the Pressure Vessels and Piping Conference (ASME), Cleveland, Ohio, 466, 235242. Suzuki, K. [2002] Report on damage to industrial facilities in the 1999 Kocaeli earthquake, Turkey, Journal of Earthquake Engineering 6(2), 275296. Uniform Building Code [1997] Structural Engineering Design Provisions 2, International Conference of Building Officials, Whittier, California. Veletsos, A. S. [1984] Seismic response and design of liquid storage tanks, Guidelines for the seismic design of oil and gas pipeline systems, Technical Council on Lifeline Earthquake Engineering (ASCE), New York, 255370, 443461. Veletsos, A. S. and Shivakumar, P. [1997] Tanks containing liquids or solids, in: Computer analysis and design of earthquake resistant structures: A Handbook, In: Computational Mechanics Publications Beskos, D. E., Anagnostopoulos, S. A., Eds., Southampton, U.K., 3, 725773.

480

485

490

495

500

505

510

515

520

525

530

24

J. C. Virella, L. E. Surez, and L. A. Godoy

Veletsos, A. S. and Yang, J. Y. [1977] Earthquake response of liquid storage tanks-Advances in Civil Engineering through Mechanics, Proceedings of the Second Engineering Mechanics Specialty Conference (ASCE), Raleigh, NC, 124. Virella, J. C., Godoy, L. A., and Surez, L. E. [2006a] Fundamental modes of tank-liquid systems 535 under horizontal motions, Engineering Structures 28(10), 14501461. Virella, J. C., Godoy, L. A., and Surez, L. E. [2006b] Dynamic buckling of anchored steel tanks subjected to horizontal earthquake excitation, Journal of Constructional Steel Research 62, 521531. Virella, J. C., Surez, L. E., and Godoy, L. A. [2005] Effect of pre-stress states on the impulsive 540 modes of vibration of cylindrical tank-liquid systems under horizontal motions, Journal of Vibration and Control 11(9), 11951220.

Вам также может понравиться