Three-Dimensional Nonlinear Seismic Ground Motion Modeling in Basins

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Physics of the Earth and Planetary Interiors 137 (2003) 81–95

Three-dimensional nonlinear seismic ground


motion modeling in basins
Jifeng Xu a , Jacobo Bielak b,∗ , Omar Ghattas b , Jianlin Wang c
a The Boeing Company, P.O. Box 3707 MC 7L-21, Seattle, WA 98124-2207, USA
b Department of Civil and Environmental Engineering, Carnegie Mellon University, Pittsburgh, PA 15213, USA
c Synopsis, Inc., 700 East Middlefield Rd., Mountain View, CA 94043, USA

Received 16 October 2001; received in revised form 10 May 2002; accepted 31 May 2002

Abstract
In this paper, we report on the development and application of a parallel numerical methodology for simulating large-scale
earthquake-induced ground motion in highly heterogeneous basins whose soil constituents can deform nonlinearly. We target
sedimentary basins with large contrasts in wavelengths for which regular grid methods become inefficient, and overcome the
problem of multiple physical scales by using unstructured finite element triangulations. We illustrate the methodology with
an example of an idealized basin, which contains a deep and a shallow sub-basin. The simulations show significant amplitude
reduction of the ground accelerations due to inelastic soil behavior at sites above the deepest portions of the sub-basins, yet
little shift in frequency. Under the assumption of linear anelastic material behavior, there is a rapid spatial distribution of the
ground acceleration of the basin, which differs markedly from that for a one-dimensional (1D) analysis. This characteristic
three-dimensional nature of the ground motion is preserved for the elastoplastic model. Concerning the ground displacement,
the main qualitative difference between the elastic and inelastic models is the occurrence of significant permanent deformations
in the inelastic case. These residual displacements can have practical implications for the design of long structures such as
bridges and structures with large plan dimensions.
© 2003 Elsevier Science B.V. All rights reserved.
Keywords: Ground motion; Basins; Simulations; Nonlinear soil behavior; Drucker–Prager elastoplastic model; Finite element method

1. Introduction Strong motion seismologists had not until recently


devoted much research to nonlinear phenomena since
Nearly all the models used until recently in seismol- compelling evidence for nonlinear effects in the ob-
ogy for predicting ground motion induced by earth- served motion, other than in liquefied sites, was scarce.
quakes have been based on the assumption of linear In the last decade, however, a number of accelero-
elastic behavior of the soil. On the other hand, for a grams have been recorded during strong earthquakes
number of years nonlinear soil amplification has been that have made it possible to infer nonlinear response.
routinely taken into consideration in geotechnical en- The most common manifestations of inelastic soil
gineering practice (Seed and Idriss, 1969; Finn, 1991). behavior involve the reduction in shear wave velocity
and the increase in soil damping with increasing load
∗ Corresponding author. Tel.: +1-412-268-2958; (Hardin and Drnevich, 1972). Accordingly, the corre-
fax: +1-412-268-7813. sponding nonlinear site effects include the lowering
E-mail address: jbielak@cmu.edu (J. Bielak). of the site amplification factor as the amplitude of the

0031-9201/03/$ – see front matter © 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0031-9201(03)00009-8
82 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

seismic loading increases; in some cases this is accom- studying the ratio of the smoothed Fourier amplitude
panied by the lowering of the resonance frequencies spectra from a soil site and a nearby rock site, and
in the spectra of the recorded ground. Thus, evidence observed pronounced strong motion deamplification
of nonlinear soil response during strong earthquakes effects. The 1994 Northridge earthquake also presents
can be identified directly if weak and strong motion plausible evidence of nonlinear soil response (Bardet
records are simultaneously obtained on the ground and Davis, 1996). Trifunac and Todorovska (1996) es-
surface and at depth in a borehole, or through simula- timated an upper bound on the distance from the earth-
tion, as long as records are available for strong shak- quake source where nonlinear soil response affects
ing. Since these simultaneous records have been rare, peak acceleration. They analyzed the observed strong
nonlinear site effects are oftentimes inferred from motion amplitudes in the San Fernando valley and
ground motion records obtained during strong and found that noticeable reduction in recorded horizon-
weak earthquakes at both rock and soil locations by tal peak accelerations occur at sites with shear wave
filtering out the earthquake source radiation and prop- velocity less than 360 m/s and distance from the fault
agation path effects. Identifying an appropriate refer- less than 15–20 km. By comparing ground motion be-
ence rock site, though essential for this approach, still tween the Northridge earthquake and its aftershocks,
represents a considerable challenge, as rock properties Field et al. (1997) reported that sediment deamplifica-
at the free surface differ from those at the basement. tion was up to a factor of two during the main shock
Nonlinear soil behavior has received increased at- implying significant nonlinearity. More recently,
tention from seismologists in recent years. SMART1 O’Connell (1999) has shown that much of the same
(Strong Motion Accelerograph Array in Taiwan) and data can be explained by linear response and scatter-
SMART2 (Abrahamson et al., 1987) have provided ing of waves in the upper kilometers of the earth’s
reliable records of events available for the study of crust. Strong motion deamplification effects were also
nonlinear site effects in a particular seismic region observed for the aftershocks of the 1983 Coalinga
in Taiwan in which a carefully established downhole earthquake in California by Jarpe et al. (1988).
accelerograph array was deployed. Analyses of these Besides the well-known amplitude deamplifica-
records (Chang et al., 1989; Wen, 1994; Beresnev tion and changes in resonance frequency, Archuleta
et al., 1995) revealed significant nonlinear soil re- (1998) has identified a characteristic spike-like wave-
sponse in both deamplification and reduction of wave form with many sharp peaks in some strong motion
velocities in ground motions with peak ground acceler- accelerograms as evidence of nonlinear response.
ation larger than 0.15 g. In Japan, nonlinear behavior of Archuleta et al. (2000) cite a number of examples
soil sediments was identified at Kuno in the Ashigara where this spiky waveform has been observed. This
valley by using borehole records (Satoh et al., 1995). characteristic high frequency waveform was first no-
At Port Island, the motion at different depths recorded ticed by Porcella (1980) and later thoroughly analyzed
during the 1995 Hyogoken–Nambu by a borehole by Iai et al. (1995).
array demonstrated strong nonlinear features of seis- Simulations of nonlinear earthquake response be-
mic amplification in reclaimed land areas and also gan in the late 1960s. These early studies were con-
in Holocene and Pleistocene soil deposits (Sato ducted for horizontally layered soils and vertically
et al., 1996). The Loma Prieta earthquake of 1989 in incident waves, by either an equivalent linear or a
California presents another case in which nonlinear direct nonlinear method. These are the main meth-
site effects have been detected. Chin and Aki (1991) ods that are still used today in engineering practice.
found a pervasive nonlinear site effect at sediment In the equivalent linear method, the soil response is
sites in the epicentral region by eliminating the influ- evaluated in an iterative manner. First, trial values for
ences of the radiation pattern and topography. They average strain are chosen, then soil properties are de-
concluded that the site amplification factor depends on termined in accordance with the trial values of strain,
the acceleration level and nonlinear effects may appear and finally the response of the model is calculated. If
at levels above 0.1–0.3 g. This is in the range expected the calculated strains differ significantly from the trial
from geotechnical engineering studies. Darragh and values, the cycle is repeated (Idriss and Seed, 1968;
Shakal (1991) estimated the response of a soil site by Schnabel et al., 1972). Many studies have concluded
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 83

that equivalent linear approach cannot reproduce some records observed at the top of a volcanic hill. Results
of the important characteristics of the seismic ground of our basin simulations show that whereas the ground
motion, especially for the cases of strong loadings motion decreases due to soil nonlinearity, the spatial
(Streeter et al., 1974; Finn et al., 1978). The equiva- variation of this motion follows closely that of the lin-
lent linear method overestimates the seismic response ear model, showing clear basin effects. Also, despite
due to pseudo-resonance at periods corresponding to the significant reduction in peak response, there is only
the strain-compatible stiffness used in the final elastic a small change in the dominant frequencies.
iteration analysis. Also, since the method is elastic it
cannot predict the permanent deformations that occur
during an earthquake. 2. Method of analysis
In the direct nonlinear method, the shear modulus
is modified at every time step according to the current The mathematical problem under consideration is
strain, so that the nonlinear stress–strain relationship one of earthquake-induced waves traveling from a
is closely followed. A number of representations have flat-layered halfspace of rock into a basin with het-
been used for the backbone curves and yield rules of erogeneous and possibly nonlinear soils and irregular
soil stress–strain relationships. They include, e.g. lin- geometry. We assume small deformations throughout.
ear (Idriss and Seed, 1968), multilinear (Joyner, 1975), We will use standard finite element techniques to solve
Yu et al., 1993), and one by Archuleta et al. (2000) this problem. Topics of special importance in the de-
based on a modified Masing rule with a provision for velopment of the model, such as the governing equa-
pore pressure (Elgamal, 1991). tions, the constitutive laws governing the materials,
While one-dimensional (1D) simulations can yield the type of loading, and the type of elements are dis-
reasonable estimates of nonlinear effects under ver- cussed briefly next; details can be found in Xu (1998).
tically incident seismic excitation, they cannot rep- The governing equation for the balance of momen-
resent the effects of surface waves and basin effects. tum is given by:
The few 2D and 2.5D studies to date of nonlinear soil
amplification (e.g. Joyner and Chen, 1975; Joyner, ρ ü − div ␴ = f (1)
1975; Elgamal, 1991; Marsh et al., 1995; Zhang and
where ρ is the density of the material, u the displace-
Papageorgiou, 1996) have confirmed the importance
ment vector field, σ the stress tensor field, and f the
of nonlinearity on site response. In particular, a 2.5D
body force, which as will be explained subsequently,
study of the Marina District in San Francisco dur-
is introduced to represent the seismic excitation.
ing the 1989 Loma Prieta earthquake found that the
Standard Galerkin discretization in space by finite
focusing and lateral interferences often observed in
elements produces a system of ordinary differential
studies based on linear soil behavior are still present
equations of the form:
for strong excitation though not as prominently as for
weak excitation (Zhang and Papageorgiou, 1996). The 
M Ü (t) + B T σ dV e = F (t) (2)
use of an equivalent linearization technique meant Ve
e
that no permanent deformations could be detected.
In this study, we describe a finite element methodol- where M is the global mass matrix; U the nodal dis-
ogy for modeling three-dimensional ground motion in placement time-dependent vector; F a time-dependent
basins, including inelastic behavior of the soil, due to vector of applied nodal forces; Ve the element volume;
arbitrary seismic excitation, and illustrate it for an ide- B the element matrix of shape function derivatives;
alized basin. The soil within the basin is idealized as and σ the current stress tensor within each element
a Drucker–Prager elastoplastic material (Drucker and computed by an appropriate procedure from the soil
Prager, 1952). The method is first verified for a sim- constitutive law. This system of equations is nonlin-
ple problem by comparing our results with those from ear because the stress σ depends nonlinearly on the
ABAQUS, a well-tested commercial software pack- current displacement U and the loading history.
age. This is the same software that was used recently The main factor influencing the reliability of numer-
by Sincraian and Oliveira (2001) to model the seismic ical calculations of nonlinear dynamic soil behavior
84 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

is the implementation of the stress–strain law. For this to efficiently resolve multiscale phenomena by tailor-
initial analysis of the ground motion of soil deposits in ing the mesh size to local wavelengths, and the ease
3D basins, we idealize the soil, both clays and sands, with which they handle traction interface and bound-
as a Drucker–Prager material. We selected this mate- ary conditions, and complex geometries, including
rial because: (1) the implementation of its constitutive arbitrary topography. Since the mesh generator in
law is similar to that required for more complex con- Archimedes builds meshes that are made up of tetra-
stitutive laws; (2) it can represent soil dilatancy and its hedra with straight edges, we use 10-node subpara-
parameters can be related to the physical soil proper- metric quadratic tetrahedra for our calculations. The
ties (cohesion and friction angle) in a rather straight- reason we introduce a quadratic approximation of the
forward way; and (3) despite its relative simplicity, it displacement field within each element, rather than the
can lead to reasonable agreement between the results linear approximation we used for our elastic models,
of simulations and observations for problems that do is to be able to represent exactly linear strain fields,
not involve soil liquefaction. This satisfactory perfor- for as it is well-known, piecewise constant strain ap-
mance of the Drucker–Prager model was observed, in proximation leads to poor accuracy for elastoplastic
particular, in a study aimed at predicting the mono- problems.
tonic and cyclic response of pile foundations to axial Two important issues must be considered for solv-
and lateral loads (Trochanis et al., 1991). ing wave propagation problems in infinite domains by
In our implementation of the finite element method the finite element method. One is the need to render
we use lumped mass diagonal matrices for each ele- the domain of computation finite and to limit the oc-
ment; the integral in the second term in (2) is evaluated currence of spurious reflections. This is accomplished
by Gauss numerical integration. As is usually done here by introducing an absorbing boundary condition
in displacement-based finite element formulations, the at the outer boundary of the computational domain.
algorithmic framework for the elastoplastic analysis is We use a simple dashpot approach (Lysmer and
strain driven. Given a prescribed strain increment at Kuhlemeyer, 1969) for this purpose, which consists
a step n + 1, the problem is to update the stress ten- of adding viscous dampers at each boundary node.
sor at each Gauss point at the new time tn+1 given its This gives rise to a diagonal damping matrix with
value at the previous time tn . We use a return mapping non-zero terms associated only with boundary nodes.
algorithm (Simo and Taylor, 1985; Ortiz and Simo, The second point that requires attention is the
1986), which is capable of accommodating arbitrary need to incorporate the seismic excitation into the
yield criteria, flow rules, and hardening laws. model, which for the example we will consider later
In order to treat efficiently large-scale 3D problems consists of an incident plane wave. This is carried
that involve elastoplastic laws, we use an element-by- out by means of a two-step method developed by
element procedure to implement Eq. (2) into our soft- Bielak and his co-workers (Bielak and Christiano,
ware. That is, the product of the mass matrix and 1984; Cremonini et al., 1988; Bielak et al., 2003). In
the corresponding nodal accelerations, and the sec- the first step, one introduces an auxiliary model that
ond term in (2) are evaluated separately for each fi- encompasses the source and a background structure
nite element and then assembled. Notice that no global from which the basin has been removed. The second
matrices need to be stored; only vectors need to be problem models the local site effects. Its input is a set
assembled. of equivalent localized forces derived from the first
We have built a parallel elastoplastic wave prop- step. These forces act only within a single layer of el-
agation simulation code on top of Archimedes, an ements adjacent to the interface between the exterior
environment for solving unstructured mesh finite ele- region and the region of interest.
ment problems on parallel computers (“Archimedes”,
1998), as a generalization of our earlier parallel elas-
tic wave propagation simulation code (Bao et al., 3. Model verification
1998). Archimedes includes 2D and 3D mesh genera-
tors, a mesh partitioner, a parceler and a parallel code Before we can apply the incremental finite ele-
generator. We favor finite elements for their ability ment code just described to a general situation, it is
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 85

necessary to verify it against known solutions from We present here results from one of these tests.
established methods. To check the purely linear ver- The physical problem under study is a solid cube
sion of our new software we compared results for fixed at the bottom surface and subjected to a uniform
simple idealized situations (Bao, 1998) against those pressure over a central portion of its top surface, as
from our linear wave propagation software tool (Bao shown in Fig. 1. The remaining sides are traction-free.
et al., 1998), with satisfactory results. The material is modeled by an elasto-perfectly-plastic
Nonlinear analysis of 3D problems presents some Drucker–Prager law (with zero frictional angle). The
difficulties to both developers and users of finite ele- applied pressure varies in time as a Ricker pulse with
ment technology. Unlike linear problems, in a model a central frequency of 0.4 Hz. The cube is divided
undergoing inelastic deformation, both normal and into smaller cubes and each cube is, in turn, subdi-
shear components of stress undergo changes corre- vided into tetrahedra with straight edges. Nodes are
sponding to changes of either the normal or shear assigned at each vertex and midside, and the solution
component of strain. It is well-known that finite el- for the displacement field within each tetrahedron
ement plasticity solutions can become highly inaccu- is of the form of a complete quadratic polynomial.
rate, especially in the fully plastic range. Inaccuracies This results in the 10-node subparametric quadratic
occur not only due to numerical inaccuracies but also tetrahedral element mentioned earlier. We use an
due to the basic incremental character of the plasticity explicit central difference method to solve the equa-
law. To verify the correctness of our implementation tions of motion (2) and four-point Gauss numerical
and test the accuracy of our numerical procedure, quadrature to evaluate the integral in the second
we developed several small-scale test problems (Xu, term.
1998) and compared the results with those from Synthetic acceleration and displacement seismo-
ABAQUS (from Hibbit, Karlsson, and Sorensen, Inc., grams obtained by our code are shown in Figs. 2
Providence, Rhode Island), a well-tested commercial and 3 for two locations, B and D on the top surface
software package. of the cube. These are labeled plasto-quake, after our

Fig. 1. Test problem for elastoplastic computations with a quadratic tetrahedral mesh. The cube is subjected to a uniform distributed load
on the middle third of its top surface that varies like a Ricker pulse over time.
86 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

Fig. 2. Comparison of dynamic response at B by: (a) ABAQUS/standard with implicit method; (b) plasto-quake with explicit method.
Displacements and accelerations are shown in the direction of the three coordinate axes.

code’s name. The corresponding results obtained with can be parallelized readily. The agreement between
ABAQUS using the exact same elements are also the results of the solutions is very good, both for
shown in these figures, for comparison. Instead of accelerations and displacements. Notice that whereas
an explicit solver, ABAQUS uses and implicit solver the displacement seismograms are fairly smooth, the
in conjunction with the 10-node tetrahedron. Even accelerograms exhibit spikes (high frequency compo-
though our explicit solver requires a smaller time step, nents). Some of these may correspond to the actual
which is dictated by the Courant stability criterion, it solution; others correspond to noise from our numeri-
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 87

Fig. 3. Comparison of dynamic response at D by: (a) ABAQUS/standard with implicit method; (b) plasto-quake with explicit method.
Displacements and accelerations are shown in the direction of the three coordinate axes.

cal approximations, as we expect our model to provide One measure of the numerical error is given by the
accurate solutions only for frequencies up to 1 Hz, traces for the displacement in the y-direction at point
corresponding to the rule of 10 points per wavelength B, which should be identically zero due to symmetry
we used to construct the finite element mesh. We (part of the error arises, of course because the mesh
include these unfiltered results primarily to show the is not completely symmetric about the vertical plane
close agreement between the two solution techniques. through B).
88 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

Table 1
Material properties of basin layers and the surrounding halfspace
Density Poisson’s Shear wave Attenuation Cohesion Friction
(g/cm3 ) ratio velocity (m/s) factor (Q) (kPa) angle (◦ )
Top layer 1.85 0.32 200 25 50 17
Second layer 1.90 0.30 400 50 160 20
Halfspace 2.10 0.28 800 100 – –
Layers obey Drucker–Prager elastoplastic constitutive law; halfspace is linearly elastic. In addition, all three materials have attenuation.

4. Illustrative example function is a Ricker pulse with a central frequency of


0.28 Hz, and the surface free-field peak acceleration
We now examine the effects of 3D nonlinear soil outside the basin is 0.06 g. With the material proper-
behavior on seismic ground motion of an idealized ties and excitation frequencies established, the finite
heterogeneous basin with an irregular basement. The element mesh is tailored to the local wavelengths. The
basin, shown in Fig. 4, consists of a 100 m soil layer shortest wavelength at a given point is determined by
underlain by a stiffer soil of variable depth, which are the local shear wave velocity and the highest frequency
surrounded by an elastic halfspace. The basin has a of the input excitation, which is taken to be 0.8 Hz. For
maximum depth of 1 km and a length of 4 km. The la- future reference, the first resonant frequency of the top
bel I refers to the box that defines the boundary of the layer under linear elastic behavior is 0.5 Hz. The re-
computational domain, where the absorbing bound- sulting mesh, coarsened for visualization purposes and
ary conditions are applied. The effective earthquake partitioned into 64 subdomains, is shown in Fig. 5. It
excitation is applied on the strip of elements between has on the order of 280,000 elements, 400,000 nodes,
II and III, which is the next to the outermost layer and, thus, close to one million degrees of freedom. The
of elements. Five points on the basin’s surface, A, B, simulations were executed on 128 processors of the
C, D, and E, are used to describe the basin response. Cray T3D at the Pittsburgh Supercomputing Center.
The material in the basin is taken to be stiff Lower While the main objective of the present analysis is
Oxford unweathered clay and is assumed to obey the to examine basic differences between the elastic re-
Drucker–Prager elastoplastic constitutive relations. sponse and inelastic response of the basin, we are also
The corresponding cohesion and friction angle are interested in assessing the basin effects on the ground
given in Table 1. This table also lists the density, elas- motion; that is, we ask the question of how differ-
tic and attenuation properties of the soil deposits and ent is the 3D response of the basin from that for a
surrounding halfspace. These properties are assumed one-dimensional (1D) analysis. For this comparison
to remain constant over time, i.e. no degradation is we concentrate only on the elastic response and con-
taken into consideration. The anelastic attenuation is sider a single free-surface location, D, above the deep-
taken to be of the Rayleigh type, i.e. within each finite est part of the basin. The 1D analysis was performed
element the damping matrix is assumed to be a linear for a soil column whose properties are identical to
combination of the mass and initial tangent stiffness those beneath point D, using as input the same incident
matrices. This yields a damping ratio, 1/[2Q(ω)], that S-wave as for the 3D simulation. The main effects of
is inversely and directly proportional to frequency, re- the lateral confinement of the basin on ground motion
spectively. The factors of proportionality are selected are illustrated by Fig. 6, which shows the horizontal
so as to minimize the integral of the squared differ- components of acceleration and displacement at D in
ence between the actual frequency-varying damping the direction of motion x of the incoming wave. The
ratio and the target damping ratio, 1/[2Q(ω)] (Bielak seismograms have been low-pass filtered to 0.8 Hz, in
et al., 1999). keeping with the accuracy of our numerical approx-
The excitation consists of a transient, vertically in- imations. Here and in the subsequent figures we fo-
cident S-wave polarized in the direction of the x-axis cus on the horizontal ground motion because, except
(axis of symmetry through line ABCD). The forcing for ground settlement, this is the one that controls
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 89

Fig. 4. Idealized three-dimensional basin, surrounded by an elastic halfspace. (a) Interface between basin and rock and cross-sections
through plane of symmetry and deepest portion of largest sub-basin. (b) Plan view of basin with contour lines of basement surface; each
step is 77 m. A, B, C, D, and E are observation points, and I, II, III, and IV are various surfaces described in the text.
90 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

Fig. 5. Finite element mesh partitioned for 64 subdomains.

Fig. 6. Acceleration and displacement seismograms at point D from 1D simulation and 3D simulation, in the direction of the x-axis.

the earthquake performance of constructed facilities. creases the maximum amplitude of the response, and
In agreement with previous investigations (e.g. Bielak second, it extends the duration of the strong phase
et al., 1999, Olsen, 2000), the basin affects the 1D re- of the ground motion. For instance, whereas the peak
sponse in two important ways. First, it significantly in- 1D acceleration is 0.10 g, the corresponding 3D value
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 91

Fig. 7. Elastic and elastoplastic displacement and acceleration seismograms in the x-direction at the observation points shown in Fig. 4.

is 0.18 g; also, the duration of shaking is more than the two sub-domains, with a peak value of about 3,
double for 3D than for 1D. This comparison shows and surface Rayleigh waves are prominent. Surface
that neglecting 3D effects can lead to erroneous re- waves are also responsible for the much longer dura-
sults. Accordingly, all subsequent results refer to our tion of the ground motion within the basin than in the
3D model. free-field and with respect to that for the 1D analysis.
Given the peak acceleration of 0.18 g we expect that The convex shape of the basement in the middle re-
the soil will exhibit nonlinear behavior. Fig. 7 shows gion precludes the amplification of motion and longer
the horizontal acceleration and displacements in the duration from developing at C, where a defocusing ef-
x-direction at the five locations shown at the bottom of fect, in contrast to the focusing that takes place at B
Fig. 5. Results are presented both for an elastic model and D, is clearly visible.
in which yielding of the soil material is not allowed, The main effect of the elastoplastic soil behavior on
and for the elastoplastic case. The accelerograms have the accelerations shown in Fig. 7 is to reduce the am-
been low-pass filtered to 0.8 Hz in keeping with the plitude of the maximum response, by about a factor
accuracy of our numerical approximations. This re- of 2 in the deepest regions where the shear strains are
moves spikes such as those present in Figs. 2 and 3. largest. Closer to the basin edge this effect is smaller;
For the elastic model, the amplification of the incom- the reduction at A is only 10%. It is noteworthy that
ing wave is very strong atop the deepest points above contrary to many 1D soil amplification studies, no
92 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

Fig. 8. Residual displacements within the basin: (a) horizontal along x-axis; (b) vertical.

significant reduction is observed in the dominant fre- an average value of the ground acceleration over the
quencies of the response and that the surface wave entire duration of the seismic event. This quantity is
effects remain quite strong for the elastoplastic case. defined as the tensor quantity:
A comparison of displacement synthetics shows that 
π T
the main qualitative difference with the elastic sim- Iij = ai (t)aj (t) dt (3)
ulation of displacements is the occurrence of perma- 2g 0
nent deformations. The distribution of these horizontal where T is the earthquake duration, and ai (t) the com-
and vertical residual displacements within the basin is ponent of ground motion acceleration in the ith direc-
shown in Fig. 8. Permanent vertical displacements are tion. The total horizontal Arias intensity is defined as:
responsible for structure settlements, as observed in
Ih = Ixx + Iyy (4)
Mexico City during the 1985 Michoacán earthquake.
Further, differential ground displacement is responsi- The square root of Ih is shown in Fig. 9 for the elastic
ble for damage to long structures and structures with and elastoplastic cases, respectively. We use the square
large plan dimensions. It is interesting that while the root so that the plotted quantity will vary linearly with
horizontal permanent displacement varies gradually the peak free-field acceleration in the elastic case. Re-
over the basin and the peaks occur in the regions of flecting the difference in the accelerograms, the peak
maximum depth, the displacement in the vertical di- value of the intensity for the elastoplastic case is about
rection, exhibits a rapid spatial variation, and the peak one-half that for the elastic case. The spatial distribu-
values occur near the basin confluences. tion of the two, however, is similar, with maxima oc-
It is usual in engineering applications to use a sin- curring above the deepest points. The Arias intensity
gle parameter to estimate the severity of an earthquake varies considerably (up to a factor of 2.5 for the elas-
at a particular location. Peak acceleration is perhaps tic case and 2 for the elastoplastic one) away from the
the most common one. This value, however, is not edges. The three-dimensionality of the response is ap-
a good predictor of potential damage to most engi- parent, as different points underlain by essentially the
neering structures, as it is related to the response of same soil layers exhibit different responses. Naturally,
structures with very high frequencies, and represents a a 1D analysis would predict identical results at points
value that may occur only at an isolated instant. In the with the same soil characteristics.
following, we use instead another quantity, the accel- To further examine the basin effects on the ground
eration Arias intensity (Arias, 1970), which provides motion, another parameter introduced by Arias (1996)
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 93

Fig. 9. Distribution of Arias acceleration intensity, (Ih )1/2 , within the basin surface, in (m/s)1/2 . Left panel: elastic model; right panel:
elastoplastic model. Notice different scales, but pattern similarity.

to measure the dominant orientation of the horizontal hand, if δ = 1, the horizontal motion occurs only in
ground motion is plotted in Fig. 10. This parameter, a single direction. Clearly, on the axis of symmetry,
denoted by δ, is defined by: which contains the points ABDE, the ground motion
[(Ixx − Iyy )2 + 4Ixy
2 ]1/2 takes place along that axis, since the incident wave is
I1 − I2
δ= = (5) polarized along the x-axis. Despite the simple nature
Ixx + Ixy I1 + I 2 of the incident motion, the ground motion shows a
in which I1 and I2 are the largest and smallest princi- variety of dominant orientations, as characterized by
pal horizontal Arias acceleration intensities. δ varies the wide range of δ. Since for a 1D analysis δ would
between zero and one. If δ = 0, the ground motions in be equal to unity throughout the basin, the variabil-
the x- and y-directions are comparable. On the other ity of this parameter underscores the significance of

Fig. 10. Distribution of Arias acceleration directivity, within the basin surface. This parameter is a measure of the variability of the dominant
orientation of horizontal ground acceleration throughout the basin. δ = 1 indicates motion only along x-axis; δ = 0 denotes comparable
motion in the x- and y-directions. Left panel: elastic model; right panel: elastoplastic model.
94 J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95

the basin effects on ground motion. Notice that for ing Center’s Cray T3D were provided under PSC
this idealized basin, the directivity coefficient for the Grant BCS-960001P. We are grateful for this support.
elastoplastic case exhibits essentially the same pattern Thanks also to Antonio Fernández for help with the
as for the elastic case. preparation of the figures, and to the two reviewers
Rob Graves and Hiroshi Takenaka for their useful
comments and suggestions.
5. Concluding remarks

In summary, results of our simulations show that References


the elastoplastic soil behavior results in an overall
reduction of the ground acceleration throughout the Abrahamson, N.A., Bolt, B.A., Darragh, R.B., Penzien, J., Tsai,
basin. On the other hand, the characteristic 3D spatial Y.B., 1987. The SMART1 accelerograph array, a review
(1980–1987). Earthquake Spectra 3, 263–287.
distribution of the ground acceleration observed for
Archimedes home page, 1998 (http://www.cs.cmu.edu/∼quake/
the elastic basin is also preserved in the elastoplas- Archimedes.html).
tic case. This means that one-dimensional analyses Archuleta, R.J., 1998. Direct observation of nonlinear soil response
usually performed to evaluate nonlinear effects might in acceleration time histories. Seism. Res. Lett. 69, 149.
need to be re-evaluated to account for basin effects. Archuleta, R.J., Bonilla, L.F., Lavallée, D., 2000. Nonlinearity in
In addition, in our nonlinear example the soils experi- observed and computed accelerograms. In: Proceedings of the
12th World Conference on Earthquake Engineering (Paper No.
enced significant residual displacements; these cannot 1934), Auckland, New Zealand.
be modeled by linear of equivalent nonlinear analy- Arias, A., 1970. A measure of earthquake intensity. In: Hansen,
ses. The presence of nonlinearity in ground motion is R.J. (Ed.), Seismic Design for Nuclear Power Plants. MIT Press,
good news in that the amplifying effects of sediments, Cambridge, MA, pp. 438–483.
on average, are apparently not as great as implied by Arias, A., 1996. Local directivity of strong ground motion. In:
Proceedings of the 11th World Conference on Earthquake
weak motion studies. On the negative side, lack of
Engineering (Paper No. 1240), Acapulco, Mexico.
linearity suggests that methods such as the empirical Bao, H., 1998. Finite element simulation of earthquake ground
Green’s function method that make use of recordings motion in realistic basins. Ph.D. Thesis. Carnegie Mellon
of small earthquakes to predict strong ground motion University, Pittsburgh, PA.
at sediment sites might have to be revised to take Bao, H., Bielak, J., Ghattas, O., Kallivokas, L.F., O’Hallaron,
nonlinearity of the surficial soils into consideration. D.R., Shewchuk, J.R., Xu, J., 1998. Large-scale simulation of
elastic wave propagation in heterogeneous media on parallel
Also, spatial dynamic displacements and residual computers. Comput. Methods Appl. Mech. Eng. 152, 85–102.
deformations can have practical implications for the Bardet, J.P., Davis, C., 1996. Engineering observations on ground
design and retrofit of long structures, such as bridges motion at the Van Norman complex after the 1994 Northridge
and structures with large plan dimensions. It should earthquake. Bull. Seism. Soc. Am. 86, 333–349.
be emphasized that the numerical results presented Beresnev, I.A., Wen, K.L., Yeh, Y.T., 1995. Seismological evidence
for nonlinear elastic ground behavior during large earthquakes.
in this paper are based on a single idealized basin, in
Soil Dyn. Earthquake Eng. 14, 103–114.
which the soft first layer is rather deep. Additional Bielak, J., Christiano, P., 1984. On the effective seismic input for
parametric studies and comparisons with observations nonlinear soil–structure interaction systems. Earthquake Eng.
are needed before results can be fully generalized. Struct. Dyn. 14, 103–114.
Bielak, J., Xu, J., Ghattas, O., 1999. Earthquake ground motion and
structural response in alluvial valleys. J. Geotech. Geoenviron.
Eng. 125, 413–423.
Acknowledgements Bielak, J., Loukakis, K., Hisada, Y., Yoshimura, C., 2003. Domain
reduction method for three-dimensional earthquake modeling
This work was supported by the US National Sci- in localized regions. Part I. Theory. Bull. Seism. Soc. Am., in
ence Foundation’s High Performance Computing press.
and Communication program and KDI program, un- Chang, C.Y., Power, M.S., Tang, Y.K., Mok, C.M., 1989. Evidence
of nonlinear soil response during a moderate earthquake. In:
der grants CMS-9318163 and CMS-9980063. The Proceedings of the 12th International Conference on Soil
cognizant program director is Dr. Clifford J. Astill. Mechanics and Foundation Engineering, vol. 3, Rio de Janeiro,
Computing services on the Pittsburgh Supercomput- Brazil, pp. 1927–1930.
J. Xu et al. / Physics of the Earth and Planetary Interiors 137 (2003) 81–95 95

Chin, B.H., Aki, K., 1991. Simultaneous study of the source, path, Olsen, K.B., 2000. Site amplification in the Los Angeles Basin
and site effects on strong ground motion during the 1989 Loma from 3D modeling of ground motion. Bull. Seism. Soc. Am.
Prieta earthquake: a preliminary result on pervasive nonlinear 90, S77–S94.
site effects. Bull. Seism. Soc. Am. 81, 1859–1884. Ortiz, M., Simo, J.C., 1986. An analysis of a new class of
Cremonini, M.G., Christiano, P., Bielak, J., 1988. Implementation integration algorithms for elastoplastic constitutive relations.
of effective seismic input for soil–structure interaction systems. Int. J. Num. Math. Eng. 23, 356–366.
Earthquake Eng. Struct. Dyn. 16, 615–625. Porcella, R.L., 1980. Atypical accelerograms recorded during
Darragh, R.B., Shakal, A.F., 1991. The site response of two rock recent earthquakes. Seismic Engineering Program Report,
and soil station pairs to strong and weak ground motion. Bull. May–August 1980. Geological Survey Circular 854-B, pp. 1–7.
Seism. Soc. Am. 81, 1885–1899. Sato, K., Kokusho, T., Matsumoto, M., Yamada, E., 1996.
Drucker, D.C., Prager, W., 1952. Soil mechanics and plastic Nonlinear seismic response and soil property during strong
analysis or limit design. Q. Appl. Mech. 10, 157–164. motion. Soils Found. Jpn. Geotechnol. Soc. (special issue),
Elgamal, A.-W., 1991. Shear hysteretic elasto-plastic earthquake 41–52.
response of soil systems. Earthquake Eng. Struct. Dyn. 20, Satoh, T., Sato, T., Kawase, H., 1995. Nonlinear behavior of soil
371–387. sediments identified by using borehole records observed at
Field, E.H., Johnson, P.A., Beresnev, I.A., Zeng, Y., 1997. Ashigara valley, Japan. Bull. Seism. Soc. Am. 85, 1821–1834.
Nonlinear ground-motion amplification by sediments during the Schnabel, P., Seed, H.B., Lysmer, J., 1972. Modification of
1994 Northridge earthquake. Nature 390, 599–602. seismograph records for effects of local soil conditions. Bull.
Finn, W.D.L., 1991. Geotechnical engineering aspects of micro- Seism. Soc. Am. 62, 1649–1664.
zonation. In: Proceedings of the 4th International Conference Seed, H.B., Idriss, I.M., 1969. The influence of soil conditions on
on Seismic Zonation, vol. 1, Stanford, CA, pp. 199–259. ground motions during earthquakes. J. Soil Mech. Found. Div.
Finn, W.D.L., Martin, G.R., Lee, M.K.W., 1978. Comparison ASCE 94, 93–137.
of dynamic analysis of saturated sand. In: Proceedings of Simo, J.C., Taylor, R.L., 1985. Consistent tangent operators for
the ASCE Geotechnology Engineering Division Specialty rate-independent plasticity. Comput. Methods Appl. Mech. Eng.
Conference on Earthquake Engineering and Soil Dynamics, 48, 101–118.
Pasadena, CA, pp. 472–491. Sincraian, M.V., Oliveira, C.S., 2001. A 2D sensitivity study of
Hardin, B.O., Drnevich, V.P., 1972. Shear modulus and damping in the dynamic behavior of a volcanic hill in the Azores Islands:
soils: measurement and parameter effects. J. Soil Mech. Found. comparison with 1D and 3D models. Pure Appl. Geophys. 158,
Div. ASCE 8, 473–496. 2431–2450.
Iai, S., Morita, T., Kameoka, T., Matsunaga, Y., Abiko, K., 1995. Streeter, V.L., Wylie, E.R., Richart, R.F., 1974. Soil motion
Response of a dense sand deposit during 1993 Kushiro-Oki computations by characteristics method. J. Geotechnol. Eng.
earthquake. Soils Found. 35, 115–131. Div. ASCE 100, 247–263.
Idriss, I.M., Seed, H.B., 1968. Seismic response of horizontal soil Trifunac, M.D., Todorovska, M.I., 1996. Nonlinear soil response,
layers. J. Soil Mech. Found. Div. ASCE 94, 1003–1031. earthquake—1994 Northridge, California. J. Geotechnol. Eng.
Jarpe, S.P., Cramer, C.H., Tucker, B.E., Shakal, A.F., 1988. A 122, 725–735.
comparison of observations of ground response to weak and Trochanis, A.M., Bielak, J., Christiano, P., 1991. Three-dim-
strong ground motion at Coalinga, California. Bull. Seism. Soc. ensional nonlinear study of piles. J. Geotechnol. Eng. 117,
Am. 78, 421–435. 429–439.
Joyner, W.B., 1975. A method for calculating nonlinear seismic Wen, K.L., 1994. Nonlinear soil response in ground motions.
response in two-dimensions. Bull. Seism. Soc. Am. 65, 1337– Earthquake Eng. Struct. Dyn. 23, 599–608.
1357. Xu, J., 1998. Three-dimensional simulation of wave propagation in
Joyner, W.B., Chen, A.T.F., 1975. Calculation of non-linear ground inelastic media on parallel computers. Ph.D. Thesis. Carnegie
response in earthquakes. Bull. Seism. Soc. Am. 65, 1315–1336. Mellon University, Pittsburgh, PA.
Lysmer, J., Kuhlemeyer, R.L., 1969. Finite dynamic model for Yu, W., Anderson, J.G., Siddharthan, R., 1993. On the
infinite media. J. Eng. Mech. Div. ASCE 95, 859–877. characteristics of nonlinear response. Bull. Seism. Soc. Am.
Marsh, J., Larkin, T.J., Haines, A.J., Benites, R.A., 1995. Compari- 83, 218–224.
son of linear and nonlinear seismic responses of two-dim- Zhang, B., Papageorgiou, A.S., 1996. Simulation of the response
ensional alluvial basins. Bull. Seism. Soc. Am. 85, 874–889. of the Marina District Basin, San Francisco, California, to the
O’Connell, D.R.H., 1999. Replication of apparent nonlinear 1989 Loma Prieta Earthquake. Bull. Seism. Soc. Am. 86, 1382–
seismic response with linear wave propagation models. Science 1400.
283, 2045–2050.

You might also like