Вы находитесь на странице: 1из 17

923

A practical iterative procedure to estimate seismic-induced deformations of shallow rectangular structures


Antonio Bobet, Gabriel Fernandez, Hongbin Huo, and Julio Ramirez

Abstract: An iterative procedure is proposed to estimate seismic-induced distortions of cut-and-cover rectangular structures. The procedure is based on an existing analytical solution for deep rectangular structures subjected to far-field shear stress which assumes elastic behavior of the soil and structure, tied contact at the soilstructure interface, and static loading. The new proposed procedure builds on the analytical solution and approximates dynamic response with a pseudostatic analysis and incorporates soil-stiffness degradation through an iterative scheme where the soil shear modulus is changed in each iteration based on the shear strain of the soil obtained in the previous iteration. The presence of the ground surface and slip at the soilstructure interface are neglected in the method proposed, but their effects are shown to be small and have compensating results when soil nonlinearity is introduced. Predictions obtained from the analytical solution have been verified by a series of numerical tests, which include the response of the Daikai subway station during the 1995 Kobe earthquake in Japan and the Los Angeles Civic Center subway station subjected to the 1994 Northridge earthquake in California. The relative errors in terms of deformation between analytical and numerical results are smaller than 15%. The procedure results in stresses on the structure that compare well with those obtained with the numerical method when there is no slip between soil and structure. If slip is allowed, the analytical solution overpredicts tensile normal stresses and underpredicts compressive normal stresses. Key words: seismic design, soilstructure interaction, underground structures, relative stiffness, shear modulus degradation, cut-and-cover. Resume : On propose une procedure iterative pour estimer les distorsions induites par des seismes dans des coupes et des structures rectangulaires quelles soutiennent. La procedure est basee sur une solution analytique existante pour des structu` res rectangulaires profondes soumises a une contrainte de cisaillement etendue qui suppose un comportement elastique du ` sol et de la structure, un contact serre a linterface sol structure, et un chargement statique. La nouvelle procedure propo see est basee sur la solution analytique et la reponse dynamique approximative avec une analyse pseudo-statique, et incor ` pore une degradation de la rigidite du sol au cours du processus diteration ou le module de cisaillement du sol est change ` a chaque iteration en partant de la deformation en cisaillement du sol obtenue au cours de literation precedente. Dans la ` methode proposee, on neglige la presence de la surface du sol et du glissement a linterface sol structure, mais il est montre que leurs effets sont faibles et ont des resultats compensatoires lorsquon introduit la non linearite du sol. Les pre dictions obtenues par la solution analytique ont ete verifiees par une serie dessais numeriques qui comprenaient la reponse de la station de Daikai durant le seisme de Kobe en 1995, et la station du Centre Civique de Los Angeles durant le seisme de Northridge en 1994. Les erreurs relatives en termes de deformation entre les resultats analytiques et numeriques sont plus petites que 15 %. La procedure resulte en des contraintes sur la structure qui se comparent bien avec celles obtenues avec la methode numerique lorsquil ny a pas de glissement entre le sol et la structure. Si on permet un glissement, la so lution analytique surestime la prediction des contraintes normales de traction et sous estime les predictions de contraintes normales en compression. Mots-cles : conception se`smique, interaction sol structure, structures souterraines, rigidite relative, degradation du module de cisaillement, coupe et couvert. [Traduit par la Redaction]

Introduction
As an integral part of the infrastructure of modern society, underground structures are used for a wide range of applica-

tions, including storage, sewage, water conveyance, and transportation systems such as subways and railways. The support of underground facilities in seismic zones must be designed for static overburden loads and for additional de-

Received 14 November 2006. Accepted 12 February 2008. Published on the NRC Research Press Web site at cgj.nrc.ca on 2 July 2008. A. Bobet1 and J. Ramirez. School of Civil Engineering, Purdue University, West Lafayette, IN 47907, USA. G. Fernandez. School of Civil Engineering, University of Illinois at Urbana, Urbana, IL 61801, USA. H. Huo. Earth Systems, 79-811B Country Club Drive, Indio, CA 92203, USA.
1Corresponding

author (e-mail: bobet@purdue.edu).


doi:10.1139/T08-026
#

Can. Geotech. J. 45: 923938 (2008)

2008 NRC Canada

924

Can. Geotech. J. Vol. 45, 2008

formations imposed by earthquakes. Seismic-induced deformations can be produced by ground failure or shaking. Ground failure includes several types of ground instability, such as direct shear displacements of active faults intersecting the structure, landslides, liquefaction of the surrounding ground, and tectonic uplift and subsidence. Ground shaking refers to vibrations produced by propagating waves through the ground. Shear waves are the most damaging to underground structures because they produce racking or ovalization of the structure (Merritt et al. 1985; Hashash et al. 2001). There are two basic approaches in present seismic design of underground structures. One approach is to carry out dynamic, nonlinear soilstructure interaction analysis using numerical methods. The input motions in these analyses are time histories emulating design response spectra and are applied to the boundaries of a soil island to represent vertically propagating shear waves. The second approach assumes that the seismic ground motions induce a pseudostatic loading on the structure. This approach allows for the development of analytical relationships to evaluate the magnitude of seismic-induced strains in underground structures (Timoshenko and Goodier 1970; Peck et al. 1972; Einstein and Schwartz 1979; Hendron and Fernandez 1983; Merritt et al. 1985; Wang 1993; Penzien and Wu 1998; Penzien 2000; Hashash et al. 2001). Such relationships generally apply to deep tunnels in relatively stiff ground. Cut-and-cover subway structures are generally rectangular shallow structures often placed on soft soils, which exhibit significant changes of stiffness with deformation. A closed-form solution for rectangular tunnels has been developed by Huo et al. (2006) that addresses some of the shortcomings of previous solutions. The development of the solution is based on complex variable theory and conformal mapping techniques and relies on the following assumptions: (i) a deep rectangular structure inside an infinite medium; (ii) plane-strain conditions in any transverse section perpendicular to the longitudinal axis of the structure; (iii) homogeneous and isotropic ground; (iv) elastic response of structure and surrounding ground; (v) pseudo-static analysis. The solution provides the normalized deformation of a rec2 3 2

tangular structure in an infinite elastic medium subjected to a far-field shear stress (Fig. 1). Because of the symmetry of the problem, the stress distribution around the structure is as depicted in Fig. 2a. The shear stress (ti) is constant along the perimeter of the structure, and the normal stress has a linear distribution along the four sides of the rectangle, with a maximum magnitude pi1 along the side of length a and pi2 along the side of height b. Because of moment equilibrium, pi1 a2 pi2 b2 . A positive sign of the normal and shear stresses indicates a distribution, as shown in Fig. 2a. The structure deformations are given by 1 stru G 1  2 Npi2 Mpi2  i L s ff b

where Dstru/Dff is the ratio of the structure racking deformation over the free-field ground deformation; ns is the Poissons ratio of the structure; G is the shear modulus of the ground; b is the height of the structure; M, N, and L are parameters that depend on the magnitude of the stresses applied at the perimeter of the structure and on the stiffness of the ground and the structure;  i is the deformation of the free-standing structure due to a unit shear stress ti; and pi2 is the deformation of the free-standing structure due to the linear normal stress distribution (Fig. 2a) with pi2 1 and pi1 b=a2 . Alternatively, when the magnitudes of ti and pi2 are known and for plane strain, the structure deformation Dstru, which is the difference between the displacements of the top and bottom of the structure (see Fig. 2b), can be estimated as 2 stru 1  2  i  i pi2 pi2 s

The values of  i and pi2 can be obtained analytically or numerically from structural analysis. For the case of a rectangular structure with width a and height b with an interior central column, with member stiffnesses: (EI)w (lateral walls), (EI)s (bottom slabs), and (EI)c (central column), the deformations are as follows (Huo et al. 2006): 3

 i 3

1 4 b 24

2 4  1 5  4 3 2 5 EIc EIs EIw 2EIs 2EIs EIw

1 pi2 b4 4

2 1 3 EIc EIw EIs 2 3 2 3 1 4  1 5  7 5 5 4 45EIc EIs EIw 480EIs 2EIs EIw 1 1  3EIc 6EIw 2EIs

2008 NRC Canada

Bobet et al. Fig. 1. Rectangular structure in an infinite medium.

925

where  = a/b is the aspect ratio of the structure (Fig. 1). If there is no central column, (EI)c = 0, and eq. [3] simplifies to 2 3 1 44  1 5  i b 24 EIs EIw 2 3 4 1 44  1 5 pi2 b 60 EIs EIw The parameters M and N in eq. [1] depend on the Poissons ratio of the ground (n) and the aspect ratio of the structure () (Huo et al. 2006) and are given in Fig. 3. The parameter L is defined as 5 L
a 1  2 Npi2 G F1 s a 1  2 Mpi2  i G F2 s

pends on  = a/b, the ratio of length to height of the structure, which is a measure of the shape of the structure. They also showed that the normalized displacement depends on the relative stiffness between the structure and the surrounding ground, and not on the absolute stiffness of the structure or the ground. There are fundamental difficulties for the application of the analytical solution to practical problems. First, the solution was developed for deep rectangular underground structures, whereas cut-and-cover structures have relatively shallow burial depths. Second, the analytical solution was based on a pseudo-static approach, whereas earthquakes induced dynamic loading. Third, the ground was considered elastic while soils exhibit elastoplastic hysteretic behavior. Lastly, derivation of the solution was done with the assumption of a tied interface between ground and structure whereas in reality slip may occur. All these difficulties are investigated and evaluated in this paper, and an approach is offered to overcome the difficulties. A practical method is proposed where the analytical solution is combined with an iterative scheme to account for the ground shear modulus degradation with cyclic loading (i.e., equivalent linear method). The new method is validated using comparisons with numerical simulations of the seismic responses of the Daikai subway station in Kobe, Japan, and the Civic Center subway station in Los Angeles, Calif.

Description of the method


The first step in the method is to obtain the seismicinduced free-field ground deformations at elevations corresponding to the top and bottom of the structure. These free-field ground deformations can be estimated as Dff = gffb = tffb/G, where Dff is the displacement difference between the top and bottom of the structure; G is the ground shear modulus; b is the height of the structure; and tff and gff are the far-field ground shear stress and shear strain, respectively. The shear modulus, shear stress, and shear strain are those computed at the center of the structure. The far-field shear strain can be obtained from ground response analyses. The computer program SHAKE (Schnabel et al. 1972), is widely employed to estimate the free-field soil deformations for various types of ground and for any given ground motion. The normalized structure deformation can thus be obtained from eq. [1], given the structure shear and normal compliances  i and pi2 , the aspect ratio , the ground shear modulus G, and the Poissons ratio n. The appropriate shear modulus G is the result of the iterative process discussed later in the paper. The analytical formulation is based on the assumption that both the ground and the structure behave elastically. This may be acceptable for a well-designed structure in a relatively low seismic area, where no damage during seismic motions is anticipated and thus the reduction of stiffness of the soil and structure is moderate. Under higher seismic motions, this assumption may not be appropriate. Therefore, in this paper, the change of soil stiffness with structure deformations is introduced in the formulation by an iterative process, where the shear modulus of the soil input into the equations is adjusted depending on the ground deformations obtained in the previous iteration (equivalent linear method).
#

where F1 and F2 can be found from Fig. 4 and are also a function of n and . The stresses along the perimeter of the structure are pi1 6 pi2 2 G ff b

pi2 M i N i L G ff b

The structure deformation can be obtained directly from eq. [1] and the stresses from eq. [6], or alternatively the stresses can be computed from eq. [6] and the deformations from eq. [2]. Huo et al. (2006) provided extensive comparisons between predictions obtained with the analytical solution and results from numerical simulations using finite element methods (FEMs). The comparisons showed that the analytical solution provides results with differences typically within a 2%8% range with respect to numerical methods. Huo et al. (2006) also showed that the normalized structure displacement is independent of the absolute size of the structure; in other words, the normalized displacement de-

2008 NRC Canada

926 Fig. 2. Seismic-induced structure (a) loading and (b) deformations.

Can. Geotech. J. Vol. 45, 2008

Fig. 3. Parameters M and N.

The process ends when the shear modulus used in the last iteration corresponds to the ground deformations (Bobet et al. 2006). If the structure moves into its nonlinear regime, the degradation of the structure stiffness may be significant and thus the constant-stiffness assumption for the structure may introduce large errors (a methodology similar to that proposed for the soil could also be used). Figure 5 shows the flow chart for the iterative procedure, which is summarized in the following steps: (1) Obtain the free-field ground deformation Dff from the shear strain at the center of the structure. The free-field ground deformation can be estimated from ground response analyses, e.g., SHAKE. (2) Compute the structure compliances  i and pi2 and the aspect ratio . This can be done from closed-form solu-

(3)

(4) (5)

(6)

tions such as those of eqs. [3] and [4] or from structural analysis. Compute parameters M, N, and L. Use the value of the ground shear modulus obtained in the previous iteration (step 6). An initial estimate of the ground shear modulus is needed to start the iteration process. Find the structure distortion Dstru from eq. [1], given Dff from step 1. Determine the shear strain g of the ground close to the structure. The soil deformations around the structure are constrained by the presence of the structure. Huo et al. (2005) showed that there is a volume of ground attached to the structure that moves with the structure and determines the structure response. Thus, g = Dstru/b. Update the ground shear modulus on the basis of the
#

2008 NRC Canada

Bobet et al. Fig. 4. Parameters F1 and F2.

927

shear strain estimated. The shear modulus is obtained from the Gg curve of the soil (e.g., from resonant column tests; Seed et al. 1986), given the magnitude of the shear strain found in step 5. (7) Evaluate the difference between the soil strains computed in the current and previous iterations. If the differences are small (e.g., error 3 < 1%), convergence has been reached and the iteration process is completed. The structure deformation is that obtained in step 4. If the differences are larger than the maximum error allowed, a new iteration is attempted with the updated shear modulus of the soil. In the cases presented in this paper, convergence has always been found after a reasonable number of iterations. (8) Compute normal and shear stresses along the perimeter of the structure from eq. [6] using the final, updated, shear modulus of the ground. The procedure assumes that the soil around the structure is homogeneous. A gradual change of soil properties with depth can still be effectively captured with the model because (i) free-field deformations are computed with actual soil properties, and (ii) changes of soil stiffness around the structure are averaged using the stiffness of the soil located at mid-depth of the structure as reference. If significant differences in soil properties are encountered at the site (e.g., a much softer or stiffer soil layer, the fill material between the "

structure and the original soil is much softer or stiffer than the in situ soil, or there are elements that locally affect the soil response such as the remains of temporary support during construction), the method proposed may not give accurate predictions. However, if soil properties can be bracketed within a reasonable range of values, the solution can provide a quick estimate of expected results, which then can be used as a preliminary design or as a first step for more complex (and expensive) numerical methods.

Verification of the method


The analytical solution has been verified with four cases: three different cross sections of the Daikai subway station in Kobe, Japan, subjected to the 1995 Kobe earthquake and the Civic Center subway station of the Los Angeles, California, Metro Red Line subjected to the 1994 Northridge earthquake. The responses of the two stations were calculated using a full dynamic numerical analysis with hysteretic elastoplastic soil model (Huo et al. 2005) and the analytical method proposed. The soil model used in the numerical analysis included an increase in damping and a reduction of the shear modulus of the ground with an increase in strain. Details of the model can be found in Huo et al. (2005). In essence, stresses and strains are related by #

toct

1 toct;c Gmax goct goct;c 1 jgoct goct;c j=n goct;y

2008 NRC Canada

928 Fig. 5. Iterative procedure for computing structure deformations.

Can. Geotech. J. Vol. 45, 2008

with a nonlinear soil model and with those from the new method proposed. Comparisons are provided for three distinct sections of the Daikai subway station and the Los Angeles Civic Center subway station. Daikai subway station The Daikai subway station, a cut-and-cover rectangular structure with central columns, constructed in the subway system in Kobe, Japan, collapsed during the Hyogoken Nambu (Kobe) earthquake of 17 January 1995. Three different cross sections can be distinguished in the station (Fig. 6): (i) section 1, the wide (platform) section that collapsed during the earthquake; (ii) section 2, which includes the running tunnels between stations; and (iii) section 3, which corresponds to the two-story access structure. Section 1 (Fig. 6a) was a rectangular, reinforced concrete box structure, 7.17 m high and 17.0 m wide, with a soil cover of 4.8 m above the top of the structure (Shawky and Maekawa 1996). The central columns had a rectangular cross section of 0.4 m 1.0 m with an axial spacing of 3.5 m in the longitudinal direction. The thickness of the lateral wall was 0.7 m, and the top and bottom slabs were 0.8 and 0.85 m thick, respectively. The running tunnel (Fig. 6b) was also a rectangular, reinforced concrete box structure, 6.36 m high and 9.0 m wide, with a 5.2 m soil cover. The central columns of the running tunnel had cross sections of 0.4 m 0.6 m with axial spacing of 2.5 m in the longitudinal direction. The thickness of the lateral walls and the top slabs was 0.4 m and the bottom slab 0.44 m. The access station, section 3, had two levels (Fig. 6c). The section was 10.12 m high and 26.0 m wide, with an average soil cover of 1.9 m. The central columns were identical to those of section 1. The lateral walls were 0.5 m thick, the top slab was 0.75 m thick, of the bottom slab was 0.85 m thick at the center and 0.55 m at the sides, and the middle slab was 0.35 m thick. At the location of the station, the soil profile consisted of manmade fill close to the surface underlain by a sequence of Holocene clay and sand layers and Pleistocene sand, clay, and gravel layers. The thickness of the fill ranged from 1 to 2 m and was underlain by Holocene clay with a thickness of about 11.5 m. The clay was followed by a Holocene sand to a depth of about 58 m below the surface. Layers of different thickness of Pleistocene sand and clay were encountered between depths of 5 and 17.5 m; a Pleistocene gravel at least several metres thick was found below the above layers. The standard penetration test (SPT) values increased with an increase in depth. For the Holocene deposits, the blow counts generally ranged between 10 and 20, and the lower Pleistocene soil deposits showed larger values, particularly at the center of the station, and increased to 50 or more on the Pleistocene clay and gravel at depths below 1517 m. The groundwater table was located at a depth of 68 m below the surface. Further details can be found in Huo et al. (2005). Extensive modeling of each of the sections and the surrounding soil was carried out by Huo et al. (2005) and Parra-Montesinos et al. (2006) to investigate the load transfer mechanisms between the structures and the soil and to identify the causes for the observed different behavior of the sections during the earthquake. In essence, each of the
#

where toct and goct are the octahedral shear stress and strain, respectively; toct,c and goct,c are the octahedral shear stress and shear strain at the last load reversal, respectively; Gmax is the initial tangent shear modulus and has a magnitude Gmax = ty/gy in which ty and gy are the reference stress and strain, respectively; and n is the scale factor (n = 1 for initial loading; n = 2 for unloading and reloading). The tangent shear modulus can be obtained from eq. [7] as 8 G @  oct 1 Gmax @ oct 1 j oct oct;c j=n oct;y 2

The model has been incorporated into the finite element code ABAQUS (Hibbitt, Karlsson and Sorensen, Inc. 2002). The following sections provide a comparison of structure distortions obtained from the dynamic numerical method

2008 NRC Canada

Bobet et al.

929

Fig. 6. Cross sections of the Daikai station: (a) section 1, collapsed station; (b) section 2, tunnel; and (c) section 3, access. All dimensions are given in millimetres.

2008 NRC Canada

930

Can. Geotech. J. Vol. 45, 2008

three sections of the Daikai subway station was placed inside a large discretization of the soil medium, with the lateral boundaries far enough from the structure to minimize their influence on the structure. The actual acceleration recorded at Port Island was input at the bottom of the discretization (Fig. 7). The soil properties used for the model were obtained from shear wave velocity and SPT measurements at the site (Shawky and Maekawa 1996; Huo et al. 2005). The small-strain shear modulus of the soil increased from Gmax = 80 MPa at the surface to 200 MPa at 58 m below the surface, which was the limit of the discretization of the model. The Poissons ratio was taken as 0.35, and the total unit weight of the soil was 19.6 kN/m3 (Shawky and Maekawa 1996). The structure, with the central columns, was assumed to have elastic behavior throughout the entire analysis or at least prior to failure (this is an acceptable assumption, even for section 1, since it exhibited a brittle failure, as shown in Parra-Montesinos et al. 2006). The concrete of the structure was modeled as an elastic material, with a unit weight of 25 kN/m3, Poissons ratio of 0.15, and Youngs modulus of 24 GPa for the frame (Chuto Industrial Inc. 1997) and 7 GPa for the columns. The Youngs modulus of the central column of the station was obtained by performing a three-dimensional (3D) finite element method (FEM) analysis of the structure with the actual dimensions and spacing of the central columns and matching the stiffness of the 3D structure with that of a two-dimentional (2D) structure where the column was assumed to be a continuous wall; hence, the actual spacing of the column was taken into consideration with the reduced stiffness (7 GPa is approximately the ratio of 24 GPa, the stiffness of the frame, to 3.5 m, the spacing of the columns of section 1; the procedure was also applied to sections 2 and 3). It was assumed that a frictional interface existed between the structure and the surrounding ground which was supposed to follow the Coulomb friction law with a coefficient of friction m of 0.4. Special contact elements were placed between the soil and the structure, which allowed for detachment of the soil from the structure in tension and enforced the Coulomb friction law in compression. Further details about the discretization and simulations are found in Huo et al. (2005). Figure 8 shows the predicted central column distortions for section 1 (results after maximum distortion are theoretical, since the central column should have collapsed beyond this point). Analogous simulations were conducted for sections 2 and 3. Table 1 shows the maximum central column distortion for each section. For the analytical solution, the dimensions of the structure used are those in Fig. 6. It has been found that the analytical solution provides the best results, compared with those of the numerical model, when the dimensions of the structure used in the equations correspond to those of the interior opening (Huo et al. 2006). This behavior is the result of the constraint that the thickness of the structure slabs imposes on the ground behind, and so for practical purposes the effective structure dimensions are those of the interior opening. Thus, a and b were measured from wall to wall and correspond to the interior dimensions of the structure, rather than to the distance between the axes of the horizontal and vertical members. The aspect ratios  = a/b were 2.82 for section 1, 1.48 for section 2, and 2.78 for section 3. The

elastic properties of the structure elements are the same as those used for the numerical analyis, with the Youngs modulus of the central column in each section reduced accordingly to approximate the 3D response. The structure deformations under a unit shear and unit normal stress,  i and pi2 , were obtained numerically from structural analysis and were  i 0:6 and pi2 0:092 for section 1,  i 0:76 and pi2 0:23 for section 2, and  i 0:62 and pi2 0:08 for section 3. The small-strain ground shear modulus at the center of each structure was equal to Gmax = 97 MPa for sections 1 and 2 and 94 MPa for section 3. The free-field soil deformations induced by the Kobe earthquake were obtained from ground response analysis using the program SHAKE 91 (Idriss and Sun 1992), given actual soil properties and input accelerations. The free-field ground deformations obtained from the shear strain of the soil at the center of each cross section were 14, 12, and 19 cm for sections 1, 2, and 3, respectively. The structure deformations for each section were obtained following the proposed iterative method (Fig. 5). The procedure followed to obtain the solution for section 1 is included in Appendix A as an example; the same process was repeated for sections 2 and 3. The shear modulus reduction of the soil with an increase in shear strain was taken from Fig. 9, which also shows the increase of damping with an increase in shear strain (data from Seed et al. 1986; Vucetic and Dobry 1991) and the results of soil stiffness and damping predicted by the soil model from eqs. [7] and [8]. The comparison between the model and experiments is good. The final horizontal structure (column) deformations of sections 1, 2, and 3 were 3.8, 3.5, and 2.1 cm, respectively. The analytical results compare well with the numerical results, namely deformations of 4.0, 3.1, and 2.2 cm, respectively (Table 1). The differences between the analytical and numerical results are 5%, 12.9%, and 4.5%, respectively, well within the degree of accuracy expected given the uncertainties involved in seismic analysis. It is important to point out that the Dstru/Dff ratios obtained (e.g., 3.8/14 = 0.27 for section 1 and 0.29 and 0.11 for sections 2 and 3, respectively) result in structural deformations significantly lower than those in the free-field. Thus, the proposed method is a considerable improvement of current analytical methods, which suggest that underground structures behave as flexible members that follow free-field deformations. Los Angeles Civic Center Metro Red Line station The Metro Red Line is the first modern heavy-rail subway in Los Angeles. The downtown part of the Red Line takes passengers under downtown Los Angeles, from Union station to Civic Center station. The total length of the Metro Red Line is 29 km. Construction started in September 1986, and the downtown part of the Metro Red Line was opened to service on 30 January 1993. The 6.8 magnitude Northridge earthquake in 1994 caused no significant damage to the existing subway tunnels and stations. The Civic Center station is a reinforced concrete structure built by cut-and-cover. The transverse cross section is a rectangular box with a central column, with dimensions somewhat variable along the longitudinal axis of the station.
#

2008 NRC Canada

Bobet et al. Fig. 7. Daikai subway station: input accelerations.

931 Fig. 9. Shear modulus reduction and damping increase with an increase in strain.

Table 1. Comparison of predicted column distortions from numerical and analytical models. Station Daikai, section 1 Daikai, section 2 Daikai, section 3 Los Angeles Civic Center Numerical (cm) 4.0 3.1 2.2 6.8 Analytical (cm) 3.8 3.5 2.1 6.6 Error (%) 5.0 12.9 4.5 2.9

Fig. 8. Daikai subway station, section 1: column distortion.

Figure 10 shows a slightly idealized cross section of the west part of the station. The structure at this location is 19.3 m wide 18.5 m high, with an overburden of 3.17 m. The thickness of the top and bottom slabs is 1.0 and 1.5 m, respectively. The thickness of the lateral walls is 1.2 m. The top one third of the central column has a 0.9 m 0.9 m square cross section, and the bottom two thirds has a rectangular cross section with 0.9 m in the transverse direction and 1.4 m in the longitudinal direction. For simplicity, the cross section of the column is modeled as 0.9 m 1.4 m in the following analyses. The longitudinal axial spacing of the columns is 9.0 m. There are two intermediate slabs in the structure that divide the station into three floors, with thicknesses of 0.56 and 0.5 m, respectively, as shown in

Fig. 10. The concrete of the structure had a unit weight of 22.6 kN/m3. A Youngs modulus Es = 30 000 MPa was used for the lateral walls and slabs and Es = 4700 MPa for the central column to account for the spacing of the columns. The Poissons ratio of the structure was taken as 0.15 (Southern California Rapid Transit District 1986). Soil exploration consisting of borings and SPT tests was performed for the design of the station (Southern California Rapid Transit District 1986). The site at the Civic center station was composed of a weathered silty claystone, known as the Fernando Formation. In the west portion of the station, a shallow layer of alluvial silty clay was found. The silty clay layer extended from the surface down to 7 m depth on top of the claystone layer, the Fernando Formation. The claystone layer can be divided into two sublayers: a shallower soft silty claystone down to 16 m and a deeper stiff silty claystone down to the maximum depth of the borings, 36.6 m (Fig. 10). Based on SPT blowcounts, the maximum small-strain shear modulus of the soil layers was taken as 40, 95, and 177 MPa for the top alluvium silty clay and the soft and stiff claystones of the Fernando Formation, respectively. The Poissons ratio of the soil was taken as 0.35. The strong motions from the Northridge earthquake on 17 January 1994 were used in the analyses. The ground motions recorded at the Glendale Las Palmas earthquake station, approximately 3 km north of the Civic Center station, were used in both the numerical and analytical simulations. Figure 11 shows the recorded horizontal accelerations, which shows a peak value of 0.36g. In the models, the accelerations were arbitrarily imposed at the bottom of the soil exploration, which was at 36.6 m below the surface.
#

2008 NRC Canada

932 Fig. 10. Los Angeles Civic Center subway station. All dimensions are given in millimetres.

Can. Geotech. J. Vol. 45, 2008

Fig. 11. Los Angeles Civic Center subway station: input accelerations.

Modeling and discretization of the Civic Center station were similar to those of the Daikai station. The cross section of the Civic Center station was placed at the center of a soil island 36.6 m deep, with free lateral boundaries far enough from the station such that their presence had no effect on the structure motions. The earthquake accelerations (Fig. 11) were input at the bottom of the discretization. The structure was modeled as elastic and with the dimensions shown in Fig. 10. The soil behavior was approximated with the soil model described previously, with properties obtained from soil exploration. The structureground interface was modeled, similar to that of the Daikai station, as frictional with m = 0.4, using special interface elements. Figure 12 shows the center column deformationtime history. The maximum column distortion is 6.8 cm (Table 1), which corresponds to a column drift of 0.39%, much smaller than that of section 1 of the Daikai station (0.8%). This column drift is below the distortion required for yielding of the columns and is consistent with the observation that no damage occurred in the station during the Northridge earthquake. The first step for the analytical solution was to obtain the free-field deformations at the center of the structure. This was done with SHAKE 91 and with soil properties and input accelerations exactly the same as those used for the numerical simulation. The free-field shear strain produced a struc#

2008 NRC Canada

Bobet et al.

933 Fig. 12. Los Angeles Civic Center subway station: column distortion.

ture deformation of 21 cm, measured from axis to axis between the top and bottom slabs. The aspect ratio of the station was  = 1.05, and the structure compliances due to unit shear and normal loads were  i 1:35 and pi2 0:36, respectively, which were obtained numerically. The maximum shear modulus Gmax of the ground at the Civic Center station site, taken at the middle height of the structure, was 95 MPa. The structure deformations were obtained following the iterative procedure discussed earlier, with shear modulus degradation with shear strain taken from Fig. 9. The end result was a column distortion of 6.6 cm (Table 1) or a column drift of 0.38%. The difference between the analytical and numerical results for the Los Angeles Civic Center station is 2.9%. The differences for sections 1, 2, and 3 of the Daikai station are 2.5%, 12.9%, and 4.5% (Table 1), respectively. Such differences are within the degree of accuracy expected in this type of problem, given the uncertainties involved in the analysis.

Dynamic, depth, and interface effects


The method proposed assumes that the dynamic amplification of stresses associated with a stress wave impinging on the structure is negligible (i.e., a pseudo-static analysis is appropriate). This assumption is correct if the wavelength of peak velocities is at least eight times greater than the width of the opening (Hendron and Fernandez 1983). Under these conditions, the free-field stress gradient across the opening is relatively small, and the seismic loading can be considered a pseudo-static load. In most underground openings, the quasi-static conditions are usually satisfied, and thus the pseudo-static approach followed in the proposed method does not introduce significant errors. The analytical solution was developed with the assumption that the underground structure was far from the surface and there was no slip between the ground and the structure at the interface. However, the depth of cut-and-cover structures is usually small compared with the characteristic size of the structure, and slip at the ground-surface contact is possible. The two factors, depth and interface, have small effects on the structure deformations; when the nonlinear behavior of the soil is included in the analysis, there are compensating results that minimize the effects of the two factors. Wang (1993) showed that the depth of an underground structure has relatively minor effects compared with other factors such as free-field ground movements and relative stiffness. The normalized structure deformation Dstru/Dff slowly decreases as burial depth decreases, with a maximum reduction with respect to a deep structure of about 15% 20%. The authors expanded the analyses done by Wang (1993) and conducted a series of linear elastic numerical analyses on rectangular structures with aspect ratios ranging from  = 1 to 3, with different relative stiffnesses at increasing depths. Even though it was found that the influence of depth on structure deformations depended on the shape of the structure (Huo 2005), the results essentially confirmed previous observations (Wang 1993). For overburden (distance from the surface to the top of the structure) to structure height ratios greater than about 1.5, the deformations are those of a deep structure. Thus, the analytical solution,

which considers a deep tunnel, would initially provide a conservative estimate of the structure deformations. To investigate the significance of friction at the ground structure interface, a series of FEMs were carried out with ABAQUS (Hibbitt, Karlsson and Sorensen, Inc. 2002). A rectangular structure in an infinite medium was considered (Fig. 1). The boundaries were placed at about 1020 times the dimensions of the structure such that the presence of the boundaries had no influence on the response of the structure. The ground was assumed isotropic, homogeneous, and elastic. Special interface elements were placed between the structure and the surrounding ground such that slip occured when the applied shear stress was larger than the shear strength, which was modeled following the Coulomb friction law. A small constant normal stress with magnitude 120 kPa was first applied to the boundaries of the discretization. This was done to prevent the soil from detaching from the structure (i.e., the normal stresses at the interface were compressive). The seismic-induced deformations were modeled by applying a constant shear stress at the far boundaries of the discretization with a magnitude of 50 kPa. Different interfaces were evaluated, with coefficients of friction ranging from 0 (full slip) to very large (no slip, tied interface). Cohesion was always taken as zero. The properties used in the analysis were as follows: Es = 24 000 MPa and ns = 0.15 for the structure, and G = 80 MPa and n = 0.35 for the ground. The thickness of the slabs and walls of the structure was 0.5 m. Figure 13 shows the effects of the frictional interface. The structure distortions, normalized with respect to the distortions obtained with a tied interface (the assumption made with the analytical solution), were plotted as a function of the interface friction and for three different structure aspect ratios, i.e.,  = 1 (square, 4 m 4 m), 2 (4 m 8 m), and 3 (4 m 12 m). The structure deformations increase with an increase in the friction coefficient. The structure deformations with an interface m > 1.2 (friction angle of 508) converge on
#

2008 NRC Canada

934 Fig. 13. Effect of friction coefficient on normalized structure deformation.

Can. Geotech. J. Vol. 45, 2008

those with a tied interface. The structure deformations with a tied interface and assuming elastic soil behavior are 12%, 9%, and 7% greater than those with m = 0.4 for  = 1, 2, and 3, respectively. The structure deformations are the smallest with m = 0.2 and are about 84%88% of the case with a tied interface. In Fig. 13, results with 0 < m < 0.2 have been shaded because openings between the ground and the structure appear in the numerical model owing to the small values of the friction coefficient. The results indicate that an interface with a tied contact yields larger structure deformations than those with a frictional interface. Hence, predictions with the analytical solution would initially lie on the safe side. A similar observation was made for circular tunnels by Peck et al. (1972) who indicated that it is conservative to assume no slip between the ground and the tunnel liner during a seismic event. When the elastoplastic soil behavior is included in the dynamic numerical analyses, the structure deformations are slightly larger with a frictional interface than with a tied interface. The nonlinear numerical analyses for each of the three sections of the Daikai station and the Los Angeles Civic Center station have been repeated exactly as was done originally with a frictional interface but now using a tied interface. Table 2 shows a comparison between structure distortions with a frictional interface and those with a tied interface. A tied interface decreases structure deformations about 5%10% with respect to a frictional interface when the nonlinear soil behavior model is used. This is in contrast with what happens with an elastic soil model where the structure deformations increase 10%15% with a tied interface. The reason for the difference is the interplay that exists between the constraint that the structure imposes on the soil and the soil stiffness degradation with an increase in strain. With an elastic model, the soil stiffness does not change. Since shear stresses from the soil are better transferred to the structure with a tied interface, the structure deformations are larger than those with a frictional interface. However, with a strain-softening soil model, the stiffness of

the soil is reduced with soil deformations. A frictional interface has two opposing effects: (i) it limits the magnitude of the shear stress that can be transmitted to the structure; and (ii) it allows larger deformations to develop within the soil surrounding the structure, since slip can occur at the interface. As the soil deforms more, the stiffness of the soil is reduced, which in turn induces further deformations in the soil with the imposed earthquake motions. This is consistent with the numerical predictions that indicate a reduction of the soil shear modulus G/Gmax to 78% around the structure with a frictional interface compared with a reduction to 85% with a tied interface. Thus, the softer soil will deform more, but the capacity of the soil to transmit deformations to the structure in shear is reduced. The end result is that the soil stiffness degradation phenomenon prevails, and the final outcome from the interrelated structure and soil behavior is larger structure deformations with a frictional interface. The effects and interplay among depth, friction, and shear modulus degradation tend to reduce the differences between the analytical and numerical results. First, neglecting depth effects in the analytical solution would result in overestimating the structure deformation by 15%20%. Second, the analytical solution (tied interface) with elastic soil behavior also results in overestimating the structure deformation by 10%15%. On the other hand, the elastoplastic hysteretic soil model shows that a frictional interface, which is the expected behavior, coupled with shear modulus degradation results in an increase of the structure deformation by about 5%10%. Hence, the net effect of the three factors results in the analytical solution (tied interface) giving predictions within 5%15% of those of the numerical model, which is within the error range shown in Table 1.

Interface stresses
Once the structure deformations are obtained with the iterative procedure proposed, the stresses at the interface between the structure and the soil can be computed with eq. [6] using the soil shear modulus found in the last iteration. As with eq. [1], the inside dimensions of the structure are used for the computations. The magnitudes of the normal and shear stresses obtained with the analytical solution and the numerical method are examined at the peak response of the structure (at about 4 s for the Daikai station and about 10 s for the Los Angeles Civic Center station). The following cases are investigated: tied interface and frictional interface with m = 0.4. Table 3 shows the magnitude of the normal and shear stresses applied to the structure obtained with the numerical and analytical methods for the Daikai and Los Angeles Civic Center stations, with the assumption of a tied interface. Figure 2a provides the shape of the stress distribution around the structure. The differences are acceptable, given the uncertainties involved in this type of analysis, and are within a 15%20% difference range. The normal stresses reported in Table 3 are due only to the dynamic-induced deformations. In other words, the initial normal stresses due to geostatic conditions need to be added to find the final normal stress. Although the similarities between the numerical and analytical results are not unexpected because, after all, structure distortions are not significantly affected by the in#

2008 NRC Canada

Bobet et al. Table 2. Predicted column distortions from the numerical model for frictional and tied interfaces. Station Daikai, section 1 Daikai, section 2 Daikai, section 3 Los Angeles Civic Center Frictional interface (m = 0.4; cm) 4.0 3.1 2.2 6.8 Tied interface (cm) 3.8 3.0 2.1 6.0 Difference (%) 5.0 3.2 4.5 11.8

935

Table 3. Comparison of interface stresses from elastoplastic dynamic numerical and analytical methods for the Daikai and Los Angeles Civic Center stations with a tied interface. Station Daikai, section 1 Daikai, section 1 Daikai, section 2 Daikai, section 2 Daikai, section 3 Daikai, section 3 Los Angeles Civic Center Los Angeles Civic Center Stress (kPa) Normal Shear Normal Shear Normal Shear Normal Shear Numerical pi1 205; pi2 ti = 335 pi1 213; pi2 ti = 290 pi1 167; pi2 ti = 261 pi1 140; pi2 ti = 181 Analytical pi1 164; pi2 ti = 334 pi1 190; pi2 ti = 264 pi1 150; pi2 ti = 298 pi1 160; pi2 ti = 193 Error (%) 19.9; 15.4 0.2 10.8; 4.1 9.0 10.3; 1.7 14.2 14.3; 14.3 6.6

1088 418 1169 154

1287 401 1149 176

terface (Table 2), the results are not realistic because tension develops at the tied interface. In reality, the soil cannot sustain tension, and the shear stresses at the soilstructure contact are limited by the frictional strength of the interface. Figure 14 shows the stress distribution around section 1 of the Daikai station with a frictional interface, where the soil can detach from the structure and slip. There are important differences with the analytical solution: (i) the stress distribution is no longer antisymmetric, and (ii) there is no tension at the interface (the total normal stresses should be obtained from those shown in Fig. 14a plus geostatic stresses). Both effects are the result of allowing slip and preventing tension along the interface. The shear stresses are no longer constant, since their magnitude is limited to the total normal stress times the coefficient of friction, m = 0.4. As a consequence, the analytical solution does not predict well the actual stress distribution around the station. It is interesting to note that a frictional interface redistributes the normal stresses, with a clear shift to compression, as tension is not possible; and the shear stress distribution is no longer constant but proportional to the normal stresses in those areas where there is slip. This redistribution of stresses does not substantially change the structure deformations (Table 2). This is to some extent an expected result, since deformations are mostly controlled by the relative stiffness between the soil and structure while interface and depth effects have relatively minor importance. Similar conclusions were obtained for the Los Angeles Civic Center station. Even though the analytical solution is unable to approximate stresses with a frictional interface, these stresses can be obtained employing numerical methods (e.g., finite element method with frictional interface elements at the soil structure interface) using as input parameter results from the analytical solution. In this case, the far-field shear stresses that are needed for the simulation can be obtained from the analytical procedure proposed, tff = Ggff, where G is the

soil shear modulus obtained at the last iteration, and gff is the far-field shear strain of the soil from SHAKE; the stiffness of the soil that needs to be used in the numerical model is the magnitude of the shear modulus resulting from the last iteration. As an alternative, and for predimensioning of the structural elements, shear forces and bending moments can be estimated, given the stresses resulting from the analytical solution, or by following the recommendation by Wang (1993), where the loads and moments are the result of a horizontal load applied to the top slab of the structure, with a magnitude such that the load produces the structure deformations given by the iterative method.

Summary and conclusions


An iterative scheme is presented in this paper to estimate deformations of cut-and-cover rectangular structures during earthquakes. The method is based on an existing analytical solution for deep rectangular structures in an elastic medium (Huo et al. 2006) and on previous work on the failure of the Daikai subway station in Japan during the 1995 Kobe earthquake (Huo et al. 2005). The previous research confirmed that the relative stiffness between the soil and the structure was the parameter that controlled structure deformations during an earthquake and that the presence of a structure stiffer than the surrounding soil significantly constrained the deformations of the soil adjacent to the structure. In this paper, the existing analytical solution has been modified by including an iterative procedure where the ground shear modulus is reduced based on the deformations of the structure (e.g., equivalent linear method). The iteration scheme is completed when the ground stiffness used as input in the analytical solution matches the stiffness that the ground should have with the deformations predicted by the solution. In the scheme, the deformations of the ground are taken as those of the structure, since the groundstructure
#

2008 NRC Canada

936 Fig. 14. (a) Normal stresses and (b) shear stresses (both in kPa) alonga interface (m = 0.4) of section 1 of the Daikai subway station.

Can. Geotech. J. Vol. 45, 2008

interface is assumed to be tied, and the ground adjacent to the structure controls the deformations of the structure. Four cases were used for verification of the proposed solution: the three sections of the Daikai subway station (sections 1, 2, and 3) and the Los Angeles Civic Center subway station. The four cases were analyzed with a dynamic numerical method with an elastoplastic hysteretic soil model using actual ground motions recorded during the Kobe earthquake of 1995 for the Daikai station and the Northridge earthquake of 1994 for the Los Angeles Civic Center station. The same four cases were also analyzed using the analytical solution with the iterative scheme, with free-field ground deformations obtained at mid-depth of the structure, given the actual ground motions and ground properties at each of the sites. Comparisons between the numerical and analytical results show that the deformations are well predicted by the analytical method, with differences within 5%15%. The fact that the analytical solution matches the numerical results is attributed to three factors that produce compensating effects: depth, interface, and nonlinear soil behavior. Depth and interface effects are not considered in the analytical solution; disregarding these two factors should produce predictions that overestimate the results by about 15%20%. As the ground around the structure is less constrained with a frictional interface, its shear modulus degradation increases, and thus the analytical solution should underpredict structure displacements. Numerical simulations indicate that the errors are of the order of 10%. Hence, the combination of the three factors, namely depth, interface, and shear modulus degradation, results in the analytical sol-

ution giving predictions that should be accurate within an error of 5%15%. The analytical solution provides the magnitude and distribution of normal and shear stresses along the soilstructure interface reasonably well when there is a tied contact. A tied interface imposes the same deformations of the soil and structure at the contact; for a small overburden, tension may be produced at some areas along the interface, as the initial normal stress may be too small compared with the tensile normal stresses induced by the shear-induced loading. For an elastic material or a stiff soil, this behavior may still be possible, but it may not be reasonable for soft soils and in particular for sandy soils. A frictional interface where slip is allowed produces a deviation of the normal stress distribution compared with that assumed in the analytical solution, with a significant increase of compression relative to tension. The end result is that, when a frictional interface is introduced, the analytical solution, which assumes a tied interface, overpredicts tension and underpredicts compression. Independently of interface characteristics, the analytical solution predicts quite well the structure deformations because the primary factor that controls deformations is the relative stiffness between the structure and the ground, whereas the interface has secondary effects. Although the proposed method may be used for predimensioning of the structural elements, accurate prediction of structure stresses requires a numerical model. Given the level of accuracy provided by the analytical solution and the number of cases investigated for its verification, it can be concluded that the proposed solution is a good approximation for a preliminary seismic design of cutand-cover rectangular structures, in particular where deformations need to be estimated. Because the method does not include nonlinear deformations of the structure and is developed with the assumption of a structure more rigid than the surrounding ground, it should only be applied to structures that are not expected to experience significant damage. It may not be appropriate for assessment of deformations or risk for existing structures that are underdesigned or for properly designed structures under seismic ground motion that significantly exceeds design levels. In both these cases, response involves highly nonlinear structural or soil behavior that could well exceed the ability of the proposed approach. However, the level of deformation estimated with the approach can be used to assess if nonlinear structural behavior will occur and thus if a more sophisticated model will be needed. The proposed method, taking into consideration the level of uncertainty that still exists for predicting the response of ground and structure during cycles of loading, should be viewed as a first approximation and not as a replacement for detailed dynamic numerical models. The analytical solution indicates that, in addition to seismic motion input data, the relative stiffness between the structure and the degraded soil is the key parameter that determines the soilstructure response behavior. The analytical solution also provides a quantitative assessment of the influence of the relative stiffness on soilstructure behavior and allows one to determine when efforts should be made to accurately determine the soil and structure stiffnesses. Even though the four cases analyzed in this paper provide confidence in the solution de#

2008 NRC Canada

Bobet et al.

937 jected to far-field shear stresses. Tunnelling and Underground Space Technology, 21: 613625. doi:10.1016/j.tust.2005.12.135. Idriss, I.M., and Sun, J.I. 1992. SHAKE 91: A computer program for conducting equivalent linear seismic response analyses of horizontally layered soil deposits. Users guide. University of California, Davis, Calif. 13 pp. Merritt, J.L., Monsees, J.E., and Hendron, A.J., Jr. 1985. Seismic design of underground structures. In Proceedings of the 1985 Rapid Excavation Tunneling Conference, 1620 June 1985. New York. Edited by C.D. Mann and M.N. Kelley. Society of Mining Engineers of theMining, Metallurgical and Petroleum Engineeers, New York. Vol. 1, pp. 104131. Parra-Montesinos, G.J., Bobet, A., and Ramirez, J. 2006. Evaluation of soilstructure interaction and structural collapse in Daikai subway station during Kobe earthquake. American Concrete Institute Structural Journal, 103: 113122. Peck, R.B., Hendron, A.J., Jr., and Mohraz, B. 1972. State of the art of soft-ground tunneling. In Proceedings of the Rapid Excavation and Tunneling Conference (RETC), Chicago, 1972. Edited by K.S. Lane and L.A. Garfield. ASCE, New York. Vol. 1, pp. 259295. Penzien, J. 2000. Seismically induced raking of tunnel linings. Earthquake Engineering & Structural Dynamics, 29: 683691. doi:10.1002/(SICI)1096-9845(200005)29:5<683::AIDEQE932>3.0.CO;2-1. Penzien, J., and Wu, C.L. 1998. Stresses in linings of bored tunnels. Earthquake Engineering & Structural Dynamics, 27: 283300. doi:10.1002/(SICI)1096-9845(199803)27:3<283::AIDEQE732>3.0.CO;2-T. Schnabel, P.B., Lysmer, J., and Seed, H.B. 1972. SHAKE: A computer program for earthquake response analysis of horizontally layered sites. Report UCB/EERC-72/12. Earthquake Engineering Research Center, University of California, Berkeley, Calif. 102 pp. Seed, H.B., Wong, R.T., Idriss, I.M., and Tokimatsu, K. 1986. Moduli and damping factors for dynamic analyses of cohesionless soils. Journal of the Geotechnical Engineering Division, ASCE, 112: 10161032. Shawky, A., and Maekawa, K. 1996. Collapse mechanism of underground RC structures during Hanshin great earthquake. In Proceedings of the 1st International Conference on Concrete Structures, Cairo, Egypt, 24 January 1996. pp. 117. Southern California Rapid Transit District. 1986. Metro rail project, Civic center station stage I and Tunnels Union station to 5th/ Hill station. Southern California Rapid Transit District, Los Angeles. Timoshenko, S.P., and Goodier, J.N. 1970. Theory of elasticity. McGraw-Hill Publishing Company, New York. Vucetic, M., and Dobry, R. 1991. Effect of soil plasticity on cyclic response. Journal of the Geotechnical Engineering Division, ASCE, 117: 89107. doi:10.1061/(ASCE)0733-9410(1991) 117:1(89). Wang, J.N. 1993. Seismic design of tunnels: a state-of-the-art approach. Monograph 7, Parsons Brinckerhoff Quade & Douglas, Inc., New York.

veloped, additional verifications should be made with other structures and earthquake input motions with a sufficiently large and diverse number of cases. Because of the uncertainty associated with estimation of seismic-induced deformations imposed on underground structures, their design must include a significant level of ductility such that brittle failures are avoided. This is particularly important in cut-and-cover structures that have to resist significant bending moments, and thus their behavior is not that different from that of above-ground structures. Collapse of an underground structure may be caused, similar to that of above-ground structures, in critical elements where the moment capacity and (or) shear capacity is exhausted. This is quite different from deep circular tunnels where the support works mostly in compression. Formation of plastic hinges in the support during earthquake-induced motions may not necessarily cause the collapse of the structure, as the compression capacity of the liner may not be significantly reduced. This provides a level of redundancy that is not available in shallow rectangular structures.

Acknowledgements
The research has been supported in part by the US National Science Foundation, Structural Systems and Hazards Mitigation Program, under grant CMS-0000136. This support is gratefully acknowledged.

References
Hibbitt, Karlsson and Sorensen, Inc. 2002. ABAQUS/explicit users manual, version 6.3. Hibbitt, Karlsson and Sorensen, Inc., Pawtucket, R.I. Bobet, A., Fernandez, G., Ramirez, J., and Huo, H. 2006. A practical method for assessment of seismic-induced deformations of underground structures [CD-ROM]. In GeoCongress 2006: Proceedings of Geotechnical Engineering in the Information Technology Age, Atlanta, Ga., 26 February 1 March 2006. Edited by D.J. DeGroot, J.T. DeLong, J.D. Frost, and L.G. Baise. ASCE, Reston, Va. 6 pp. Chuto Industrial Inc. 1997. Recovery record of the Daikai station after Kobe earthquake. Chuto Industrial Inc., Japan. [In Japanese.] Einstein, H.H., and Schwartz, C.W. 1979. Simplified analysis for tunnel supports. Journal of the Geotechnical Engineering Division, ASCE, 105(GT4): 499518. Hashash, Y.M.A., Hook, J.J., Schmidt, B., and Yao, J.I. 2001. Seismic design and analysis of underground structures. Tunnelling and Underground Space Technology, 16: 247293. doi:10.1016/ S0886-7798(01)00051-7. Hendron, A.J., and Fernandez, G. 1983. Dynamic and static design considerations for underground chambers. In Seismic design of embankments and caverns. Edited by T.R. Howard. ASCE, New York. pp. 157197. Huo, H. 2005. Seismic design and analysis of rectangular underground structures. Ph.D. thesis, Purdue University, West Lafayette, Ind. Huo, H., Bobet, A., Fernandez, G., and Ramrez, J. 2005. Load transfer mechanisms between underground structure and surrounding ground: evaluation of the failure of the Daikai station. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 131: 15221533. doi:10.1061/(ASCE)1090-0241(2005) 131:12(1522). Huo, H., Bobet, A., Fernandez, G., and Ramrez, J. 2006. Analytical solution for deep rectangular underground structures sub-

Appendix A
Calculations of deformations of section 1 of the Daikai subway station during the Kobe earthquake are included as an example. Calculation steps follow the procedure depicted in Fig. 5. (1) Free-field ground deformations are computed from the
#

2008 NRC Canada

938 Table A1. Iterations for section 1 of the Daikai station. Iteration 1st 2nd 3rd 4th 5th 6th 7th 8th 9th 10th 11th 12th 13th 14th 15th 16th 17th 18th 19th 20th 21st gi not applicable 7.31 102 5.64 104 4.98 102 1.05 103 1.45 102 2.95 103 9.24 103 3.69 103 9.09 103 3.98 103 8.15 103 4.76 103 7.74 103 6.06 103 7.24 103 6.40 103 7.07 103 6.73 103 6.90 103 6.83 103 Gi (MPa) 97.00 0.16 30.00 0.30 4.60 0.85 2.90 1.08 2.70 1.17 2.50 1.40 2.30 1.80 2.15 1.90 2.10 2.00 2.05 2.03 2.04 Dstru/Dfree-field 2.8811 0.0222 1.9636 0.0413 0.5720 0.1164 0.3642 0.1456 0.3584 0.1568 0.3350 0.1876 0.3053 0.2390 0.2854 0.2522 0.2788 0.2655 0.2721 0.2695 0.2708 Dstru (cm) 40.34 0.31 27.49 0.58 8.01 1.63 5.10 2.04 5.02 2.20 4.69 2.63 4.27 3.35 4.00 3.53 3.90 3.72 3.81 3.77 3.79 gi+1 7.31 5.64 4.98 1.05 1.45 2.95 9.24 3.69 9.09 3.98 8.50 4.76 7.74 6.06 7.24 6.40 7.07 6.73 6.90 6.83 6.87

Can. Geotech. J. Vol. 45, 2008

102 104 102 103 102 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103

Gi+1 (MPa) 0.16 30.00 0.30 4.60 0.85 2.90 1.08 2.70 1.17 2.50 1.40 2.30 1.80 2.15 1.90 2.10 2.00 2.05 2.03 2.04 2.04

|gi+1 - gi|/max(gi+1, gi) (%) 100 99 99 98 93 80 68 60 59 56 53 42 39 22 16 12 10 5 2 1 0

(2)

(3)

(4) (5) (6)

shear strain at the mid-depth of the structure using the computer code SHAKE 91. Dff = 14 cm. Structure compliances are obtained from numerical analysis of the cross section of the structure by imposing unit shear and normal loads. Thus,  i 0:6 and pi2 0:092.  = 15.6/5.52 = 2.82. Soil and structure Poissons ratios are n = 0.35 and ns = 0.15, respectively. From Fig. 3, M = 3.941 and N = 0.145. From Fig. 4 and eq. [5], L = 0.65. It is assumed in the first iteration that G = Gmax = 97 MPa. Note that L depends on G and changes with iterations. From eq. [1], Dstru/Dff = 2.881. Thus, Dstru = 40.34 cm. g1 = Dstru/b = 0.073. A shear modulus G1 = 0.16 MPa is obtained from the Gg curve of Fig. 9, given the magnitude of the shear strain g1.

(7) A new iteration is attempted with G = G1 = 0.16 MPa. New iterations are performed with updated values of the shear modulus until the error is small. All iterations are included in Table A1. The final structure deformation is 3.8 cm. The number of iterations could be significantly reduced with a better estimate of the final shear modulus in the first iteration. In the example, G = Gmax was taken as the starting value. For strong ground motions, it is unlikely that the soil deformations are going to be that small. In general, a smaller number of iterations may be attained with the initial guess G = 0.5Gmax.

2008 NRC Canada

Вам также может понравиться