Vous êtes sur la page 1sur 22

Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

www.elsevier.com/locate/soildyn

Practical applications of a nonlinear approach to analysis


of earthquake-induced liquefaction and deformation of earth structures
Zhi-Liang Wang*, Faiz I. Makdisi, John Egan
Geomatrix Consultants, Inc., 2101 Webster Street, 12th Floor, Oakland, CA 94612, USA

Accepted 11 November 2004

Abstract
Seismic stability, liquefaction, and deformation of earth structures are critical issues in geotechnical earthquake engineering practice. At
present, the equivalent linear approach is considered the ‘state of practice’ in common use. More recently, dynamic analyses incorporating
nonlinear, effective-stress-based soil models have been used more frequently in engineering applications. This paper describes a bounding
surface hypoplasticity model for sand [Wang ZL. Bounding surface hypoplasticity model for granular soils and its applications. PhD
Dissertation for the University of California at Davis, U.M.I. Dissertation Information Service, Order No. 9110679; 1990; Wang ZL, Dafalias
YF, Shen CK. Bounding surface hypoplasticity model for sand. ASCE, J Eng Mech 1990;116(5):983–1001; Wang ZL, Makdisi FI.
Implementing a bounding surface hypoplasticity model for sand into the FLAC program. In: Proceedings of the international symposium on
numerical modeling in geomechanics. Minnesota, USA; 1999. p. 483–90] incorporated into a two-dimensional finite difference analysis
program [Itasca Consulting Group, Inc. FLAC (Fast Lagrangian Analysis of Continua), Version 4. Minneapolis, MN; 2000] to perform
nonlinear, effective-stress analyses of soil structures. The soil properties needed to support such analyses are generally similar to those
currently used for equivalent linear and approximate effective-stress analyses. The advantages of using a nonlinear approach are illustrated
by comparison with results from the equivalent linear approach for a rockfill dam. The earthquake performance of a waterfront slope and an
earth dam were evaluated to demonstrate the model’s ability to simulate pore-pressure generation and liquefaction in cohesionless soils.
q 2005 Elsevier Ltd. All rights reserved.

Keywords: Stability of dams; Earthquake performance; Nonlinearity of soils; Plasticity model; Liquefaction

1. Introduction past earthquakes; (d) by comparing the dynamic-induced


shear stresses to the cyclic strength, the potential for
The state-of-practice for evaluating the seismic stability liquefaction of the embankment and foundation soils is
of earth dams was developed in the early and mid-1970s by estimated; (e) for the zones of the embankment that are
the late Professor Seed and co-workers at the University of determined to have liquefied during the earthquake, a
California at Berkeley [5]. The basic elements of this residual strength is assigned based on the density of the soil;
analysis included the following steps: (a) earthquake ground (f) the stability of the embankment and its foundation is
motions are estimated at bedrock underlying the dam and its evaluated using limit equilibrium stability analyses; if the
soil foundation; (b) the response of the embankment to the embankment is found to be stable, earthquake-induced
base rock excitation is computed to estimate the dynamic permanent displacements are estimated using Newmark-
stresses induced in representative elements of the embank- type deformation analyses. In this procedure, the dynamic
ment; (c) the cyclic strength of the embankment soils is response of an embankment is estimated using an iterative
evaluated using in situ tests and liquefaction resistance equivalent linear approach [6] to model the nonlinear strain-
correlation curves based on observed performance during dependent modulus and damping properties of the soils.
With this approach, however, the computed seismic
response and shear stresses within embankment soils do
* Corresponding author. Fax: C1 510 663 4141. not reflect the effects of induced pore pressure during
E-mail address: zlwang@geomatrix.com (Z.-L. Wang). earthquake shaking. The amount of induced pore pressure
0267-7261/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. and the potential for liquefaction are estimated at the end of
doi:10.1016/j.soildyn.2004.11.032 the specified duration of shaking.
232 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

Nomenclature
a model parameter m state-dependent index
b model parameter M critical stress ratio in pKq space
A constant Mp phase transformation stress ratio in pKq space
c1,c2,c3,c4 constants (nD)ij, (nN)ij unit tensor indicating the direction of the
C(x) modulus degradation function deviatoric plastic strain increment, and devia-
d dilatancy toric unit loading, respectively
d3ij, d3eij , d3pij total, elastic, and plastic deviatoric strain p mean normal stress
increment pc mean normal stress at critical state
d3v, d3ev , d3pv total, elastic, and plastic volumetric strain pm maximum mean normal stress
increment pa atmospheric pressure
d3vd, d3pd shear-induced volumetric strain and plastic q deviatoric stress in triaxial space
strain increment rij, rij stress ratio and image stress ratio, respectively
dp increament of mean stress R, Rf, Rp, Rm, stress ratio invariants for loading, failure,
drkl increament of deviatoric stress ratio tensor phase transformation, and maximum,
dsij increament of deviatoric stress tensor respectively
e, ec void ratio, critical void ratio Rp stress ratio invariants for dilatancy
eij, eeij , epij deviatoric strain, elastic and plastic deviatoric sij deviatoric stress tensor
strains, respectively V(e) a function of void ratio
g gravity w a function controls shear-induced volume
g(q) shape function on pZconst plane changes [1,2]
Gmax elastic shear modulus a model parameter
Go model parameter aij projection center for rij
G secant modulus dij Kronecker delta
hr model parameter 3v, 3vd total and shear-induced volumetric strain
h(x) heavyside step function g engineering shear strain
Hr, Hp plastic shear modulus for drij and dp mechan- ga shear strain amplitude
isms, respectively k model parameter
Ip state pressure index r, r distances from projection center
J second deviatoric stress invariant svo overburden pressure
kr model parameter t shear stress
K elastic bulk modulus tm maximum shear stress
K0 coefficient of lateral earth pressure at rest h stress ratio in triaxial space
Kr, Kp plastic bulk modulus for drij and dp mechan- x accumulated plastic deviatoric strain
isms, respectively

In cases where an earth structure is subjected to severe The procedure uses a nonlinear, bounding surface
ground motions, and where liquefaction may occur soon plasticity constitutive model for sand [1,2], incorporated
during shaking, an alternative approach to account for the into the finite-difference computer program FLAC [4].
effects of buildup of pore pressure and the potential for Because the basis for such an analysis is a cyclic plasticity
liquefaction during earthquake shaking is to use a nonlinear soil model, we will briefly introduce this model before
effective-stress analysis. This paper describes a nonlinear, shifting our focus to applications. Recent improvements to
fully coupled dynamic analysis procedure for estimating the the model include: (1) a newly proposed state parameter, the
potential for buildup of pore pressure, the potential for state pressure index, and its use in defining a dilatancy
liquefaction, and the resulting permanent deformation of curve; (2) simulation of the critical-state behavior of sands;
earth structures. The basic elements for such analyses are: and (3) simulation of post-liquefaction deformation of
(1) an estimated site ground motion; (2) a constitutive model sands. These new formulations are verified by comparing
to simulate soil behavior under conditions of cyclic loading model simulations to laboratory test results.
and liquefaction; (3) a computer program capable of This paper also introduces our current practice of using
performing dynamic analyses that are fully coupled (that this bounding surface plasticity model, as incorporated into
include mechanical aspects and groundwater flow); and (4) the computer program FLAC, to perform nonlinear, fully
relevant laboratory and in situ measurements of soil coupled dynamic analyses. The procedures for using and
properties. calibrating the model also are explained. The soil tests to
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 233

support such analyses are generally similar to those


J
currently used for equivalent linear and simplified effec-
tive-stress analyses. The advantages of using a nonlinear
approach are illustrated by comparing its results with those Line of Phase Failure Surface
from an equivalent linear analysis for a rockfill dam. The Transformation R-Rf=0
earthquake performance of a waterfront slope and an earth R-Rp=0
dam are also presented to demonstrate the model’s ability to
simulate pore-pressure generation and liquefaction in
cohesionless soils. Given the increasing availability of Maximum
Pre-stress Surface
high-speed personal computers, the newly formulated R-Rm=0
analytical methods are becoming more practical and
attractive despite their theoretical complexity. Loading
dσij

2. Bounding surface hypoplasticity model for sand Unloading


dσij

We will briefly introduce the basic features of the model


p
and its capability to simulate sand behavior during
monotonic and cyclic loading under drained and undrained Fig. 1. Surfaces and stress variables in JKp plane.
conditions, focusing on the prediction of liquefaction. Then,
we examine its advantages compared with using existing step function h(pKpm) and the Macauley brackets hdpi
models like the Mohr–Coulomb or hyperbolic stress–strain indicate that the plastic mechanism due to dp operates only
relations. when pZpm and dpO0. In a simplified two-dimensional
version of the model [3], the flat-cap is omitted, because
under undrained conditions mean pressure often decreases
2.1. Basic formulations
and the flat-cap is inactive. Be cautious to use this in a
settlement calculation, where consolidation may occur, in
The general analytical formulation of the bounding
such case pZpm and dpO0, so that the flat-cap is needed.
surface hypoplasticity model for sand [1,2] was written in
The elastic shear and bulk moduli are functions of p
terms of tensors of stress rate and strain rate. The elastic and
plastic strain rates are given as rffiffiffiffiffi
p
Gmax Z Go pa VðeÞ (3)
dsij dpdij pa
d3eij Z C (1)
2Gmax 3K in which
 
p ðnD Þij dij ð2:973 K eÞ2
d3ij Z C pðdrkl ðnN Þkl Þ VðeÞ Z
Hr 3Kr 1 Ce
  rffiffiffiffiffi
rij dij 1 Ce p
C C hðp K pm Þhdpi (2) K Z pa (4)
Hp 3Kp k pa
Gmax and K are the incremental elastic shear and bulk in which Go and k are model parameters that define the
moduli, respectively. Hr and Kr are the plastic shear and elastic moduli, e is void ratio, and pa is atmospheric
bulk moduli, respectively, associated with the tensor of the pressure.
stress ratio increment drij. Similarly, Hp and Kp are the
plastic shear and bulk moduli, respectively, associated with
the mean pressure increment dp. nD is a unit vector in stress
space along the direction of the deviatoric plastic strain
increment, and nN is the associated direction of deviatoric
unit loading. dij is a unit tensor along the p axis. Using J2Z
0.5sijsij and RZJ/p, the failure surface (RKRfg(q)Z0)
and the loading surface (RKRmg(q)Z0), as shown in Figs. 1
and 2, are defined in the stress space JKp with a shape
function g(q) on pZconstant plane (q is the Lode angle). A
flat-cap surface, pKpmZ0, is proposed for simulating
consolidation and rebound. The pm is the maximum value of
p experienced in the past loading history. The heavyside Fig. 2. Surfaces and stress variables in pZconstant plane.
234 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

The plastic modulus Hr for monotonic and cyclic loading 2.2. Comparison with laboratory test results
is
This cyclic nonlinear model has been used to simulate
  m 
Rf r sand and silt behavior under various loading and saturation
Hr Z Gmax hr CðxÞ K1 (5) conditions [1]. Yamada and Ishihara [7] conducted
Rm r
extensive laboratory studies using a true triaxial device.
where RZRm is the maximum pre-stress bounding surface, They studied Fuji River sand under complex stress paths
and RZRf is the failure bounding surface. r and r are including rotational shear, and all these test results were
distances from the projection center (as shown in Fig. 2) to simulated using the model with one set of parameters [1,2].
the current stress and maximum pre-stress surfaces, Towhata and Ishihara [8] performed cyclic torsional shear
respectively. Index mZ ð1C gðqÞÞRm =r is state-dependent tests for Toyoura sand. The model simulation and test
for irregular cyclic loading. For stress-controlled cyclic results (for both the stress–strain relation and effective-
loading, the shear stress-shear strain relations derived based stress path) are compared in Fig. 3. In these comparisons,
on Eqs. (3) and (5) conform to the well-known Masing rules. the cyclic contractive and dilative behavior of the sand is
C(x) is a function of a hardening/softening measure; e.g. the well predicted by the model.
accumulated plastic strain, volumetric and/or deviatoric.
C(x) increases (during hardening) or decreases (during 2.3. Comparison with other nonlinear models
softening) with x.
The plastic volumetric strain rate is controlled by the Although there is a number of plasticity constitutive
plastic bulk modulus, Kr: models for soils published in the literature, but, many of
rffiffiffiffiffi these models have not been used in the industry. In recent
1 Ce p years, a PEER (Pacific Earthquake Engineering Research)
K r Z pa (6)
wk pa supported research project is aimed to shorten the gap
between the research tools and practical applications using a
For plastic modulus Hp and Kp as well as other details of multi-yield surface model [9]. But, in the industry,
the model formulation for cyclic loading, readers are currently, two nonlinear soil models are often used in the
referred to the original publications [1,2]. dynamic analysis of soil structures—the Mohr–Coulomb

Fig. 3. Model Simulation of Cyclic Torsional Test Results [8].


Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 235

(or bilinear model), and the hyperbolic stress–strain   K1


2t t t
formulation [10]. Two parameters characterize these two G=Gmax Z 1 K m lnð1 K Þ C Þ
hr t tm tm (9)
models in the simple shear mode: the soil strength, tm, and
initial shear modulus, Gmax. The hyperbolic stress–strain g Z t=G
relationship is defined by:
in which G is secant modulus and hr is a model parameter.
The modulus reduction curve, G/Gmax, for this model is the
Gmax g first equation in Eq. (9).
tZ (7) For a given soil strength, tm, and shear modulus, Gmax, the
1 C Gtmax g
m
modulus reduction curve (Eq. (9)) is a function of shear
stress, and the function varies with model parameter hr, as
It has an equivalent functional form as shown in Fig. 6. This parameter can be calibrated against a
given soil modulus reduction curve to obtain the best fit to the
laboratory data. The capability of this model to fit specific
g Z t=G G=Gmax Z 1 K t=tm (8) modulus reduction curves enables a good comparison with
results obtained using an equivalent linear model, except at
in which G is the secant modulus and G/Gmax is the large strain ranges (1% or higher), where the equivalent
‘modulus reduction curve.’ linear approach may unreasonably compute stresses that
Modulus reduction curves represent basic soil properties exceed the soil strength. Other ‘modified hyperbolic’ models
used for analyzing the dynamic response of soil structures. have been developed by adding more parameters to Eq. (7) to
Seed and Idriss [6] summarized laboratory tests and gave provide a better fit to published soil modulus reduction
upper- and lower-bound curves for sands, as shown by the curves; however, these analytical equations are restricted to
thin dashed lines in Fig. 4. The top first and second lines relationships between shear stress and shear strain only. For
show modulus reduction curves based on the Mohr– appropriate engineering applications, a soil model should
Coulomb and hyperbolic models, respectively. For a given provide a general relationship between stress increment and
soil strength, tm, and an initial shear modulus, Gmax,, these strain increment tensors (e.g. see Eqs. (1) and (2)).
modulus reduction curves are uniquely defined for each of The original hyperbolic stress–strain relation represents
the two models, as shown in Fig. 6. These curves do not nonlinear elasticity; i.e. volumetric strain only is affected by
adequately match the ranges for sand presented by Seed and mean stress changes, it does not predict the volumetric
Idriss [6]. As commonly recognized, modulus reduction changes induced by variation of shear stress or there is no
curves differ for different soils. Therefore, it is desirable to coupling between deviatoric stress (including shear stress)
provide constitutive models that can better fit the curves and volumetric strain. Several empirical formulations have
derived from the test data. been proposed to compute volumetric strains due to shear
In the bounding surface hypoplasticity model, Eqs. (1) strain changes, such as in Martin et al. [11]:
and (2) can be integrated under monotonic simple shear
loading with pZconstant condition. The analytical solution d3vd Z c1 ðga K c2 3vd Þ C c3 32vd =ðga C c4 3vd Þ (10)
of the model for this loading condition is Based on this empirical equation, the volumetric strain
increment per cycle is a function of the total accumulated
Gmax = 2484 ksf, τm = 2.6 ksf volumetric strain, 3vd, and the amplitude of the shear strain
from Model of Wang (1990) with hr = 0.2
from Hyperbolic Curve
cycle, ga. Although the relation considers the strain
from Mohr-Colomb Model hardening from accumulated volumetric strain, it does not
Range for Sand after Seed & Idriss (1970) simulate the dilative behavior of sand within a cycle.
1.0 In the model of Wang [1,2], the plastic volumetric strain
Normalized Shear Modulus, G/Gmax

0.9 increment (due to shear) is computed using an indepen-


0.8 dently defined plastic bulk modulus—Kr in Eq. (1). The
0.7 ratio between the plastic bulk modulus, Kr, and plastic shear
0.6 modulus, Hr, is identified [12] as:
0.5
rffiffiffi
0.4 d3pv 3 Hr
0.3 p Z
d3d 2 Kr
0.2
rffiffiffi  a    
0.1 3 Gmax hr p h b Mp K h
0.0 Z (11)
0.0001 0.001 0.01 0.1 1 10 2 Kkr pm M h
Effective Shear Strain Amplitude (%)
For a case in which model parameters aZ0 and bZ1, it
Fig. 4. Comparison of modulus reduction curves from nonlinear models. takes the following form
236 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

coupled analysis is the one in which calculation of


displacements and pore pressure is carried out by solving
a coupled system of equations that includes the motion
equation and the diffusion equation. We applied this
approach to the following practical problems in geotechni-
cal earthquake engineering: (1) seismic response analysis
for soft soil sites subjected to very strong earthquake
motions; (2) seismic stability analysis for soil–structure–
foundation interaction involving liquefiable soils; (3)
earthquake-induced deformation analysis of embankment
Fig. 5. Model simulation for volumetric strain under cycle drained shear and earth dams; and (4) estimates of periods for strength
loading. recovery and settlement in post-earthquake and post-
liquefaction conditions.
d3pv
dZ Z AðMp K hÞ (12)
d3pd 3.1. Seismic site response using nonlinear approach
in which A is a constant, and hZq/p is the stress ratio. This
is similar to the dilatancy equation first proposed by Rowe In current practice, the site response to a given
[13] using the energy concept. In plasticity theory, dilatancy earthquake input motion is commonly obtained using an
simply relates volumetric part of plastic strain rate to a scalar equivalent linear approach. For low-level earthquake
measure of its deviatoric part. It can be seen from Eq. (11) or excitations (e.g. input peak acceleration scaled down to
(12), if stress ratio hZq/p is greater than Mp, we have about 0.1g), the nonlinear and equivalent linear approaches
d3pv ! 0, indicating dilative behavior. Fig. 5 is an example of produce essentially the same results (see Fig. 6), when the
computed volumetric strain under drained cyclic loading model was calibrated using the same dynamic soil proper-
showing the contractive and dilative behavior of sand [1]. ties, and similar modulus reduction and damping curves.
When an input motion with a peak acceleration of 0.56g was
applied at a depth of 6.4 m to a site having a soft clay layer
3. Nonlinear, fully coupled dynamic analyses at depth between 2.4 and 6.4 m, the computed ground
motions at the ground surface were different. The equivalent
The above is an effective-stress based, fully nonlinear linear SHAKE analysis predicted a surface motion having
cyclic plasticity model. It has been implemented into the site an amplified peak acceleration of 0.8g, as shown in Fig. 7.
response analysis (with multi-directional shaking) program
5%-Damped
SUMDES [14], and incorporated into the two-dimensional,
Conputed Surface Motion: Equivalent Linear Results (SHAKE)
large-deformation code FLAC [4], to perform fully non-
Computed Surface Motion: Non-linear Results (SUMDES)
linear, fully coupled dynamic response analyses. A fully Spactral Matched Input Motion
4
0.5
5%-Damped
10% input of M8: SHAKE
10% input of M8: Non-linear Results (SUMDES)
0.4
3
Spectral Acceleration (g)
Spectral Acceleration (g)

0.3
2

0.2

1
0.1

0 0
0.01 0.1 1 10 0.01 0.1 1 10
Period(sec) Period (sec)

Fig. 6. Comparison of response spectra at site surface using 10% of input Fig. 7. Comparison of response spectra at site surface and input motion
motion (M8 event). (M8 event).
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 237

300 this site during the earthquake [15]. The predicted


liquefaction in the top 20 m of fill is shown in Fig. 8 in
Effective Mean Pressure p', kPa

at 32m terms of time histories showing the reduction in mean


effective pressure within the sandy layers. The computed
200
surface acceleration time histories match the recordings
reasonably well for both pre- and post-liquefaction periods,
at 16m as shown in Fig. 9.
at 12m
100
at 6m 3.3. Predicting deformation using the nonlinear model

Given an infinite slope surface, the static stresses can be


0
10 20 30 40 expressed by an element of soil having a non-zero initial
Times, Second shear stress (and normal stresses). Under a sinusoidal cyclic
shear stress loading, the initial shear stress renders the cyclic
Fig. 8. Computed time histories of mean effective pressure at Port of Island
during Kobe earthquake. stress–strain loops asymmetric, and shear deformation
accumulates during each successive cycle, as shown in
The nonlinear analysis [14] de-amplified the peak accelera- Fig. 10. If the soil element were saturated, pore water
tion at the surface. In the nonlinear analysis an effective pressure may be generated, resulting in a reduction in
friction angle of 308 was assumed for the soft clay layer. The effective stress. Because the elastic shear modulus of the soil
equivalent linear approach, however, predicts shear stresses (Eq. (3)) is a function of the mean effective pressure, the shear
in the soft clay layer that exceed the assumed shear strength modulus also will decrease. Thus, the accumulated defor-
of the layer. For instance, the ratio between the computed mation will significantly increase as shown by the dashed line
peak shear stress and vertical confining pressure from the in Fig. 10. Incorporating the step-by-step degradation of
equivalent linear analysis reached 0.92 at a depth of 5.8 m. modulus due to pore water pressure generation is often called
‘fully-coupled’ effective-stress analysis.
3.2. Simulation of liquefaction and post-liquefaction
site response 3.4. Post-earthquake and post-liquefaction settlement
and strength recovery
We used records obtained from a downhole array at Port
Island, Japan during the 1991 Kobe earthquake, to compute The seismic stability of an offshore structure foundation
the site response and predict the observed liquefaction at was analyzed for specified design earthquakes [16].

0.8 Recorded at Ground Surface


Acceleration, g

0.4 Computed Using SUMDES (3D)


0.0
–0.4
–0.8
10 15 20 25 30 35
0.6
0.4 at Depth 16m
Acceleration, g

0.2
0.0
–0.2
–0.4
10 15 20 25 30 35
0.4
0.2
Acceleration, g

at Depth 32m
0.0
–0.2
–0.4
–0.6
10 15 20 25 30 35
Time,seconds

Fig. 9. Comparison of recorded and computed acceleration time histories, N00E component at Port of Island.
238 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

Initial Shear Stress 0.2 ksf 4. Practical applications


Initial Normal Stress 1.8 ksf
0.8
Drained Condition In this section, we describe practical applications of the
Undrained Condition
above nonlinear approach to evaluating the seismic stability
0.6
of a rockfill dam, the performance of a waterfront slope
(berth) during the 1989 Loma Prieta, California, earthquake,
0.4 and prediction of liquefaction of the foundation and
Shear Stress, ksf

estimates of permanent deformation of an earth dam.


0.2
4.1. A rockfill dam
0
New Exchequer Dam, located on the Merced River east
of Merced, California, is a 150 m high, concrete-faced,
–0.2
rockfill dam that impounds Lake McClure, providing a gross
storage capacity of one million acre-feet at normal
–0.4 maximum reservoir level. The crest length is 426.7 m.
–0.004 0 0.004 0.008 0.012 The volume of rockfill is four million cubic meters. Both
Shear Strain upstream and downstream slopes are 1.4H to 1.0V. The dam
Fig. 10. Model simulation for sand with initial shear stress under cyclic sits in a deep, relatively narrow canyon eroded in highly
loading. competent, meta-andesite rock, and all rockfill was placed
on a fully stripped foundation. The seismic stability of the
The results of the analysis indicated that a silty sand layer dam was evaluated for potential earthquake ground motions
underlying the site (Layer No. 2 in Fig. 11) would liquefy from a maximum magnitude, Mw 6.5 earthquake on the Bear
except for an area directly beneath the foundation. Effective Mountains fault, about 1/2 km from the site. Estimated
stresses at locations A–C beneath the foundation (Fig. 12) median peak horizontal accelerations at the site exceeded
were computed after earthquake shaking and were found 0.5g. The dam was analyzed using both equivalent linear
to increase with time, due to dissipation of excess pore and nonlinear approaches.
water pressure. This analysis was performed to resolve a Two-dimensional dynamic finite element analyses were
concern that the structure could become unstable if storm performed using the computer program QUAD4M [17] for
loading occurred after an earthquake had softened the the maximum cross-section of the dam (Fig. 13). The
silty sand layer, and before enough time had passed for variation of shear modulus and damping ratio with shear
the earthquake-induced excess pore water pressures to strain is shown in Fig. 14. Details of this analyses are
dissipate. presented in [18]. Dynamic finite element analyses provide

Fig. 11. Grid and boundary conditions for axis-symmetric analysis using FLAC.
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 239

During Earthquake After Earthquake 28 Hours


2.4 2.4
k=0.0001cm/sec for Silt layers

A
2.0 2.0
A

B
1.6 1.6
B
Mean Pressure, ksf

1.2 C 1.2

C
0.8 0.8

0.4 0.4

0.0 0.0
0 4 8 12 16 20 24 28 20000 40000 60000 80000 100000
Time, second Time, second

Fig. 12. Computed times histories of mean pressure during and after earthquake.

the distribution of peak accelerations within the embank- deformation (about 18.3 cm) was for a potential ‘shallow’
ment and seismic coefficient time histories within potential sliding surface within the upper one-fourth of the down-
sliding masses for estimating permanent slope deformation. stream slope.
Slope stability analyses were performed to estimate factors The response of New Exchequer Dam also was evaluated
of safety and yield accelerations, ky, for potential sliding using nonlinear analyses employing the bounding surface
surfaces within the upstream and downstream slopes of the hypoplasticity model incorporated into the finite difference
embankment. Permanent deformations were estimated program FLAC. Six parameters are required in the nonlinear
using the concept of yield acceleration proposed by dynamic response analyses. Three of these parameters (f,
Newmark [19] and modified by Makdisi and Seed [20]. Go, and hr) can be estimated from parameters used in the
The computed peak horizontal acceleration at the crest of equivalent linear analysis. f is the Mohr–Coulomb friction
the dam was about 0.85g, and the greatest amplification angle; Go is related to K2max; and hr is a model parameter
occurred in the upper 12.2 m of the embankment. that characterizes the relationship between shear modulus
The greatest estimated earthquake-induced permanent and shear strain. The model parameter hr for each material

Fig. 13. Maximum section of New Exchequer dam.


240 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

1.0
0.9

Normalized Shear Modulus, G/Gmax


0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0001 0.001 0.01 0.1 1

Upper Bound Sand Curve (Seed & Idriss, 1970) p=8ksf


Non-linear Model at Typical Depths (p= 8 to 39 ksf)

30 18
23
39
Damping Ratio (%)

ksf
20

10
Lower Bound
(Seed & Idriss, 1970)
0
0.0001 0.001 0.01 0.1 1

Effective Shear Strain Amplitude (%)

Fig. 14. Relationships of normalized modules and damping ratio with shear strain.

zone was calibrated to provide modulus and damping Computed peak horizontal displacements at the crest of the
relationships with shear strain similar to those used in the embankment were less than 61 cm in the downstream
equivalent linear analysis (Fig. 14). direction; computed vertical settlements were about
The response of the maximum section of the dam was 36.6 cm. Contours of computed horizontal and vertical
computed using the same input acceleration time history as displacements throughout the embankment at the end of
used for the equivalent linear analysis. For simplicity the earthquake shaking are shown in Fig. 17. The maximum
rock base was assumed rigid. This assumption may have horizontal displacements at the end of earthquake shaking
resulted in higher computed accelerations compared to (within the upper one-fourth of the downstream slope) were
those from the equivalent linear analysis, which treated the about 46 cm. The corresponding maximum vertical settle-
bedrock foundation as an elastic half-space. The nonlinear ments at the crest were about 37 cm. The estimated
analysis provides the time history of acceleration and permanent displacements from the equivalent linear
displacement at any location within the embankment, as analyses were less than 20 cm.
well as contours of horizontal and vertical displacement at
the end of earthquake shaking. Fig. 15 shows the time 4.2. A water front slope
history of computed acceleration at the crest of the dam,
along with the acceleration time history computed from the Following the 1989 Mw7 Loma Prieta earthquake in
equivalent linear analysis. The computed maximum crest California, significant damage was observed at the site of
acceleration from the nonlinear analysis is about 1.1g, the Port of Oakland’s Seventh Street Marine Terminal [21].
compared to 0.85g from the equivalent linear approach. The Damage included a slope failure along part of a dike near the
phases of the two time histories are in reasonable agreement southwest corner of the terminal; damage to battered, pre-
(Fig. 15). stressed concrete piles in the wharf structure; settlement as
Time histories of horizontal and vertical displacements at great as 0.3 m of the landside crane rail; numerous cracks in
the crest of the embankment are presented in Fig. 16. the pavement in the yard area; and movement of some of
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 241

1.0
Equivalent Linear Model

Acceleration (g)
0.5

0.0

–0.5

–1.0
0 4 8 12 16 20 24 28 32 36
1.0
Non-linear Bounding Surface Hypo-plasticity Model
Acceleration (g)

0.5

0.0

–0.5

–1.0
0 4 8 12 16 20 24 28 32 36
Time (seconds)

Fig. 15. Crest acceleration time histories computed from equivalent linear and nonlinear approaches.

the armor stone slope protection on the face of the dike. the perimeter dike are believed to consist primarily of
Other earthquake effects at the site included liquefaction- hydraulically placed dredged sand.
induced sand boils, cracks at the ground surface in unpaved A finite difference model of the wharf embankment
areas near the dike, and lateral and vertical displacement as profile is presented in Fig. 19. The two lateral side boundary
great as 0.3 m between the ground surface and adjacent pile- condition is so called ‘transmitting boundary’ allow wave
supported structures along the dike. The level of ground propagation through the vertical interface. In FLAC, this is
shaking recorded at a ground motion instrument within named as ‘Free Field boundary’.
2 km of the site was about 0.28g. Dynamic model parameters (maximum shear modulus
The observed behavior of a typical profile at Berth 40 and the variation of modulus and damping ratio with shear
without the wharf structure (shown in Fig. 18) during the strain) were developed based on soil properties used in
Loma Prieta earthquake was simulated using the fully equivalent linear analyses. Cyclic strength model par-
coupled, nonlinear effective-stress analysis described in this ameters were calibrated based on in situ standard
paper. penetration test (SPT) data (converted to (N1)60) and the
Construction records of the berth at this site indicate that observed occurrence of liquefaction during past earth-
after soft mud was dredged from beneath the dike area, quakes. Such cyclic strength curves, shown in Fig. 20, are
dredged sand fill was hydraulically placed by bottom-dump based on the liquefaction chart of Seed et al. [22], the
barges up to the elevation of the original mud line. A small magnitude scaling factors reported in Youd and Idriss [23],
rock toe dike was constructed on top of the first sand lift. and the magnitude vs. equivalent number of cycles
After construction of the rock toe dike, a second sand lift relationship of Seed and Idriss [24]. The dashed lines in
was placed to the elevation of the top of the toe dike. Rock Fig. 20 show the cyclic strength curves from the field data;
spoil material was then placed by end-dumping from trucks. the solid curves were predicted based on calibrated model
Compacted sand fill was later placed on top of the rock parameters. Details of the model parameters used in the
spoil. Materials used to create the terminal yard area behind analysis are described in [25]. The input ground motion used
Displacement at Crest ( feet)

2.0
1.5
1.0 Horizontal Displacement Relative to Rock Base
0.5
0.0
–0.5
–1.0
–1.5 Vertical Displacement
–2.0
0 4 8 12 16 20 24 28 32 36
Time (seconds)

Fig. 16. Computed time histories of crest displacements using FLAC and bounding surface hypoplasticity model[1].
242 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

Fig. 17. Computed displacement contours using FLAC and bounding surface hypoplasticity model [1].

in the analysis was based on the recording from the strong Fig. 21 shows that the computed effective mean stress in
motion instrument near the site. The time history recorded at layers A and B dropped to its lowest level after 15 s of
the ground surface was deconvolved to compute an interface shaking, indicating the occurrence of liquefaction. In
motion at the top of the stiff old clay (bay mud) layer contrast, the top layer, C, did not reach full liquefaction
underlying the site at a depth of about 20 m. because the underlying layers A and B had already liquefied
The analysis was performed first for the free-field zone and fully softened, and thus could not transmit the strong
(i.e. a one-dimensional column) representing areas away shaking to the soil above. However, this layer also was
from the crest of the slope. Three layers (A–C in Fig. 19) softened by the significant reduction in effective stress.
located at the bottom, middle, and top of the submerged Fig. 22(a) and (b) show the relationships between the
sand fill were selected to show the computed decrease in shear stress and effective mean stress, as well as shear
mean effective stress, the effective-stress path, and the shear strain computed for layer B, using the bounding surface
stress vs. shear strain variation during earthquake shaking. model.
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 243

–100 –90 –80 –70 –60 –50 –40 –30 –20 –10 0 10 20 30 40 50 60 70 80 90 100
40 40

30 30

20 20

10 Compacted sand fill 10


Rip rap
0 Rock spoil dike: 0

–10 Sand fill Dike fill Rock toe dike: –10


Elevation, feet

–20 –20

–30 Medium dense to dense clayey/silty sand –30

–40 –40
Dense to very dense sand
–50 Medium dense to dense clayey sand –50

–60 –60
Dense to very dense sand
–70 –70
Stiff to very stiff silty clays
–80 –80
–100 –90 –80 –70 –60 –50 –40 –30 –20 –10 0 10 20 30 40 50 60 70 80 90 100
Horizontal Distance, feet

Fig. 18. Typical cross section at Berth 40, Port of Oakland.

Similar responses were computed for zones near the SPT blow counts obtained from field investigations by the
slope (locations D–F in Fig. 19). The effective-stress time US Army Corps of Engineers (Corps), Sacramento District,
histories (Fig. 23) show that the sand fill near the rock toe are shown in Fig. 26. These values and laboratory test
dike (zone F) had liquefied, while the sand fill beneath the results were used to calibrate the model parameters.
crest (zone D) experienced dilative behavior after initial
liquefaction. The rock spoil portion (zone E) did not liquefy. (1) Parameter f is the effective friction angle. For the
The computed time histories of displacement, both alluvium, the pervious and transition zones it was estimated
horizontal and vertical, at the crest are shown in Fig. 24. based on the laboratory data as shown in Fig. 27.
The base displacement is shown as a reference for the crest (2) The Corps of Engineers conducted a geophysical
movements. The computed deformed shape of the slope at investigation where shear wave velocities were measured
the end of earthquake shaking is shown in Fig. 25. The in borings drilled through the upstream and downstream
greatest displacement values (about 0.5 m) were those slopes of the embankment. The measured shear wave
developed at the rock toe below the water surface. The velocities were presented in a ‘normalized’ form in terms of
horizontal displacement at the crest was computed to be Vs1 (see Fig. 26):
about 37 cm. The computed settlement at the crests was  1=4
about 24 cm. These computed deformations are consistent pa
Vs1 Z Vs (13)
with those observed during the Loma Prieta earthquake. svo

Based on Eq. (3), it can be shown that the model parameter


4.3. An earth dam Go is related to Vs1 by the following equation:

Success Dam serves primarily as a flood control and


water conservation project that was constructed between
1958 and 1961. The dam is a rolled-earthfill embankment
44 m high and 1038 m long. It consists of a central
impervious core along with upstream and downstream
pervious shells. The pervious shells of the embankment
were constructed on in situ recent alluvium, alluvial fan
deposits, terrace deposits, and overburden (slope wash)
deposits that may be susceptible to liquefaction during a
seismic event [26]. A typical cross-section of the dam
showing values of normalized shear wave velocities and Fig. 19. Finite difference grid and typical material zones.
244 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

0.6

0.5

Cyclic Stress Ratio, τ/ mo' d=2 d=2.5


0.4
(N1 )60 =20

0.3

d=1

0.2 (N1 )60 =15


(N1 )60 =10

0.1
(N1)60=5

0.0
1 2 3 4 5 6 7 8 9 10 20 30 40 50 60 70 80 90 100
Number of Cycle to Liquefaction, N

Fig. 20. Cyclic strength of sand with 5% (or less) fines based on liquefaction chart [24].
rffiffiffiffiffiffiffi
g svo calibrated using the results of monotonic laboratory tests.
Go Z V2 (14)
gpa VðeÞ s1 p This calculation was made only for the recent alluvium that is
susceptible to liquefaction. Fig. 29 shows the triaxial shear
test results and model simulation using krZ0.55 and an
(3) Parameter hr characterizes the nonlinear relationship estimated phase transformation ratio, Rp, as 0.85 of the
between shear modulus and shear-strain amplitude. It can be failure ratio, Rf, that corresponds to a friction angle of 368.
calibrated against laboratory test results or published (5) Model parameter d controls the development of excess
relationships for different soils. For Success Dam, there are
insufficient cyclic test results to calibrate this parameter. (a) 0.8
Instead, we selected published modulus reduction relation- Computed Effective Stress Path in Layer B
Shear Stress (ksf)

ships. For example, for the cohesionless transition, pervious, 0.4


and alluvium zones, the upper-bound relationship for sand
developed by Seed and Idriss [24] was selected to calibrate 0.0
the model parameters, the calibration procedure involved
searching for a value of parameter hr that provides a best fit to –0.4
the selected modulus reduction curves as shown in Fig. 28.
(4) Parameter kr, which affects the changes in effective stress –0.8
under monotonic (and virgin) loading conditions, was 0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2
Effective Mean Pressure (ksf)
2.0 (b) 0.8
Effective Mean Pressure (ksf)

A At Inland free field Computed Shear Stess Vs. Shear Strain in Layer B
1.6
Shear Stress (ksf)

0.4
B
1.2
C 0.0
0.8
C –0.4
0.4

–0.8
0.0 –0.002 –0.001 0 0.001 0.002 0.003 0.004
0 4 8 12 16 20 24 28 32 36 40
Time (sec) Shear Strain

Fig. 21. Computed mean effective pressure time histories in sand fill layers Fig. 22. Computed mean effective pressure, shear stress and shear and shear
at inland free field. strain in sand fill layer B at inland free field.
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 245

2.0 pore water pressure due to cyclic loading under undrained


Effective Mean Pressure (ksf)

conditions (or shear-induced volumetric change under


1.6 Under Crest and Slope
drained conditions). The number of cycles required to
D cause liquefaction of a soil element was estimated based on
1.2
E the computed effective-stress path (or shear stress vs.
0.8 effective mean pressure), as shown in Fig. 30 for a given
cyclic stress. From the above example, for the selected
0.4 F parameters (krZ0.55 and dZ3) at the specified stress ratio of
0.3, seven cycles are needed to reduce the mean effective
0.0
0 4 8 12 16 20 24 28 32 36 40 pressure to the point of initiating or ‘triggering’ liquefaction.
Time (sec)
The procedure for simulating cyclic loading was repeated for
Fig. 23. Computed mean pressure time histories in sand fill zones under selected locations in the recent alluvium (represented by
crest and slope. mean effective stresses of 4 and 8 kips per square foot (ksf))
using different cyclic stress ratios. The number of cycles to
2.0 trigger liquefaction was determined for each stress ratio.
1.6
Each combination of stress ratio and the corresponding
Computed Displacement, feet

X-displacement at crest number of cycles to cause liquefaction in the simulated test


1.2
was used to construct the model simulation curves shown by
0.8
dashed lines in Fig. 31. For the recent alluvium layer under
0.4
X-displacement at base the upstream slope (represented by a SPT blow count of 10), a
0.0 model parameter dZ1 provided a reasonable fit to the
–0.4 strength curve based on the field data. A similar curve using a
–0.8 model parameter dZ3 was developed for the recent alluvium
Y-displacement at crest
–1.2 layer under the downstream slope, represented by a blow
–1.6 count of 15. Model parameters are presented in Table 1.
–2.0
0 4 8 12 16 20 24 28 32 36 40 A rock input motion provided by the Corps was used in
Time (sec)
the dynamic analysis. The estimated peak acceleration of
Fig. 24. Computed horizontal and vertical displacement time histories at the record is about 0.22g, and the time history represents
crest of slope.

JOB TITLE : (*10^ 2)

FLAC (Version 3.40)


0.750

LEGEND

18-Sep-99 4:14
step 209630
Cons. Time 3.7710E+10
0.250
–1.341E+02 <x< 1.335E+02
–1.626E+02 <y< 1.049E+02

Exaggerated Grid Distortion


Magnification = 5.000E+00
Max Disp = 1.761E+00
–0.250

–0.750

–1.250

Geomatrix Consultants, Inc.


2101 Webster Street, 12 Floor, Oa
–1.000 –0.500 0.000 0.500 1.000
(*10^ 2)

Fig. 25. Computed deformed slope at Berth 40 (displacements shown are magnified five times).
246 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

Fig. 26. Typical cross section of success dam (1 footZ0.3m).

a relatively long duration of shaking (an 80 s record), a gradual softening and resulting in accumulation of shear
simulating a magnitude 8.0 event on the Owens Valley fault. strains (maximum strains are not shown).
The input motion is specified at the base of the old dense The predicted deformed shape of the embankment at the
alluvium layer underlying the dam foundation. Fig. 32 end of earthquake shaking is shown in Fig. 35. This figure
shows both the input acceleration time history and the shows that at the end of earthquake shaking, the computed
computed acceleration time history at the crest. Fig. 32 maximum settlement at the crest of the dam is about 5.5 m.
indicates that the input motion was amplified at the crest of Maximum computed horizontal displacements of points on
the dam for at least the first 25–30 s. As discussed below, the the embankment slopes were 1.5 and 15 m at the upstream
recent alluvium layer liquefied after about 20–30 s of and downstream toes, respectively.
shaking. It may be useful to compare the predicted large
Dynamic shear stresses result in buildup of excess pore deformation for this embankment with the actual perform-
pressures in the zones within the embankment and ance of other similar dams observed in past earthquakes.
foundation that have relatively low blow counts, such as Sheffield Dam was constructed on a silty sand layer with an
the recent alluvium layer in the foundation. The increase in estimated (N1)60 for clean sand of about 8 blows per foot.
pore water pressure results in a corresponding decrease in Sheffield Dam failed near the end of earthquake shaking
the mean effective stress. Fig. 33 shows the time histories of during the 1925 Santa Barbara earthquake [27] as a result of
the mean effective stresses computed for four selected sliding of the entire embankment on the liquefied layer.
locations in the recent alluvium underlying the upstream
slope of the dam. The initial (pre-earthquake) values were 40
those computed from static stress and seepage analysis. The Pervious Zone
B-4B @ elev. 596.4'
lowest allowable mean effective stress was specified at 35 B-5B @ elev. 597.2'
Principal Stress Difference, (σ1'–σ3')/2 (ksf)

B-6B @ elev. 608.5'


14.4 kPa (0.3 ksf) to simulate a state of initial liquefaction. B-7B @ elev. 613.0'
Liquefaction occurred first in the recent alluvium near the 30
B-8B @ elev. 618.6'
B-11B @ elev. 655.6'
upstream and downstream toes of the slope (at about 12– B-12B @ elev. 668.0'
16 s) and then extended toward the centerline of the dame sinφ'=tanψ, c'=a/cosφ B-15B @ elev. 595.5'
25 φ'=37 degrees, c'=0 B-18B @ elev. 651.5'
(at about 28–30 s).
Fig. 34 shows computed shear stresses and shear strains 20
for four zones in the recent alluvium underlying the
upstream shell of the embankment (similar, but much larger ψ=31 degrees
15
shear strains were computed in the alluvium beneath the
downstream side). The left columns on this figure show 10
relations of shear stress vs. mean effective stress. The failure
envelope, corresponding to a friction angle of 368, is also 5
shown in this figure using dashed lines.
The right column in Fig. 34 shows the shear stress-shear 0
a=0
strain relations for zones in the recent alluvium under the 0 5 10 15 20 25 30 35 40
upstream shell. As the effective-stress paths move from the Average Principal Effective Stress, (σ1'+σ3')/2 (ksf)
initial mean stress to minimum mean stress (triggering of
liquefaction), the shear stress and shear strain loops show Fig. 27. Effective Stress Stength Envelope of Previous Zone.
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 247

1.0

Normalized Shear Modulus (G/Gmax) and Damping (β)


0.9

0.8

0.7

0.6

0.5
from Seed and Idriss (1970)
Transition Zone
0.4
Upstream Pervious Shell
0.3 Downstream Pervious Shell
Upstream Recent Alluvium
0.2 Downstream Recent Alluvium
Old Alluvium
0.1

0.0
0.0001 0.001 0.01 0.1 1
Shear Strain, (%)

Fig. 28. Normalized Shear Modulus and Damping Ratios vs. Shear Strain for Sand.

Our estimated deformation for the Success Dam, using the represented by the concept of critical state. Applications
analyses described above, are consistent with the observed of critical-state soil mechanics to the constitutive modeling
performance of Sheffield Dam during and after the 1925 of sand have gained wide acceptance in geotechnical
Santa Barbara earthquake. engineering. To simulate the behavior of sand at the critical
state, the phase transformation stress ratio line originally
5. Recent developments in the constitutive model proposed by Ishihara et al. [29], was revised as a curved
for sand dilatancy line by Wang et al. [12] that passes through the
critical state. This line is defined using a ‘state pressure
Laboratory tests on Toyoura sand [28] show that sand index,’ IpZP/Pc. Here, pc is the mean pressure at the critical
behavior under large shear deformations can be well state corresponding to the current void ratio.

25 25

20 20
sin φ'=3M/(6+M) Line of Phase
φ'=36 degrees Transformation
Deviator Stress, q (ksf)

M=1.46, Failure Line


15 15
1

10 10

5 5

0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 5 10 15 20 25
Axial Strain Mean Effective Stress, p', (σ1'+2σ3')/3' (ksf)

Model Simulations TH-15 specimen 2


TH-14 specimen 1 TH-15 specimen 3
TH-14 specimen 2 TH-15b specimen 1
TH-14 specimen 3 TH-15b specimen 2
TH-15, specimen 1 TH-15b specimen 3

Fig. 29. Comparison of Model Simulations and Triaxial Test Results.


248 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

(a)

Shear Stress, τ (ksf)


0

–1
Effective Stress Path

0 2 4 6
Effective Confining Pressure, P' (ksf)

(b)

1
Shear Stress, τ (ksf)

–1
Shear Stess Vs. Shear Strain

–0.20 –0.15 –0.10 –0.05 0.00 0.05 0.10 0.15 0.20


Shear Strain

Fig. 30. Computed Cyclic Stress-Strain Relations for Stress RatioZ0.30 of Recent Alluvium.

0.6
Cyclic Strength Relations Based on Seed et al. (1985)
(N1)60=20 Model simulation for (N ) =10 using d=1, kr=0.55, from Po'=4ksf
1 60
Model simulation for (N ) =15 using d=3, kr=0.55, from Po'=4ksf
1 60
(N1)60=15 Model simulation for (N ) =10 using d=1, kr=0.55, from Po'=8 ksf
1 60
0.5 Model simulation for (N ) =15 using d=3, kr=0.55, from Po'=8ksf
1 60
Cyclic Stress Ratio, τ/σmo'

0.4

(N1)60=10

0.3

(N1)60=5
0.2

0.1

0.0
1 2 3 4 5 6 7 8 9 10 20 30 40 50
Number of Cycles to Liquefaction, N

Fig. 31. Cyclic Strength Relations and Model Simulations.


Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 249

Table 1
Material zones and model parameters Success Dam

Zone no. Zone description Void ratio f0 Go k hr kr d Rp/Rf b


a
I Impervious core 0.773 18–41 200 0.02 1.663 100 100 1.0 2.0
II Transition 0.685 37 354 0.002 0.241 100 100 1.0 2.0
III Pervious shell (US) 0.392 37 335 0.002 0.308 100 100 1.0 2.0
(DS) 0.392 37 336 0.002 0.443 100 100 1.0 2.0
IV Recent alluvium 0.685 36 260 0.002 0.189 0.55 1.0 0.85 2.0
(US)
(DS) 0.685 36 354 0.002 0.227 0.55 3.0 0.85 2.0
V Old Alluvium 0.620 41 1230 0.002 1.200 100 100 1.0 2.0
a
For clay core at a depth that is equivalent to a strength of CZ0.8 ksf and fZ118.

0.4
Horizontal Acc. (g)

computed motion at crest


0.2

–0.2

–0.4
0.4
Horizontal Acc. (g)

input motion at base


0.2

0.0

–0.2

–0.4
0 10 20 30 40 50 60 70 80 90
Time(sec)

Fig. 32. Horizontal acceleration time histories at base and computed at crest of Success dam.

10
Upstream Recent Alluvium
at zone (20, 2)
8
Effective Mean Pressure (ksf)

6 at zone (16, 2)

4
at zone (12, 2)

at zone (7, 2)

0
0 10 20 30 40 50 60 70 80
Time(sec)

Fig. 33. Computed mean effective pressure time histories beneath upstream shell of Success Dam (1 ksfZ47.9 kpa).
250 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

3 3
at zone (7, 2) at zone (7, 2)
2 2

Shear Stress, τ 1 1

0 0

–1 –1

–2 –2
0 1 2 3 4 5 6 7 8 9 10 –0.04 –0.03 –0.02 –0.01 0
3 3
at zone (12, 2) at zone (12, 2)
2 2
Shear Stress, τ

1 1

0 0

–1 –1

–2 –2
0 1 2 3 4 5 6 7 8 9 10 –0.04 –0.03 –0.02 –0.01 0
3 3
at zone (16, 2) at zone (16, 2)
2 2
Shear Stress, τ

1 1

0 0

–1 –1

–2 –2
0 1 2 3 4 5 6 7 8 9 10 –0.04 –0.03 –0.02 –0.01 0
3 3
at zone (20, 2) at zone (20, 2)
2 2
Shear Stress, τ

1 1

0 0

–1 –1

–2 –2
0 1 2 3 4 5 6 7 8 9 10 –0.04 –0.03 –0.02 –0.01 0
Mean Effective Stress, po Shear Strain, εxy

Fig. 34. Computed stress and strain relations in recent alluvium beneath upstream shell of Success Dam (stress ub ksf, 1 ksfZ47.9 kPa).

The dilatancy line is defined as: consolidated at different confining pressures. The model

simulation using this approach is compared with test results
p
q Z Md p Z Mo C ðM K Mo Þ p (15) on Toyoura sand. Readers are referred to [12] for details.
pc Another desirable improvement to the model of Wang
Using this state-dependent dilatancy line to replace the [1] is the ability to simulate post-liquefaction deformation.
phase transformation line (a straight line in Fig. 1), the model Recent laboratory tests (Kamerrer, 2001, personal com-
can simulate the behavior of Toyoura sand at critical state munication) demonstrate that shear deformation develops
using a unique set of parameters for different void ratios continuously following initial liquefaction, in sands
Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252 251

Boundary of Undeformed Grid (*10^3)


Deformed Grid 0.700

0.500
–5.000 –3.000 -1.000 1.000 3.000 5.000
(*10^2)

Fig. 35. Computed deformed grids of the Success Dam at end of earthquake shaking (1 footZ0.3 m)

subjected to cyclic loading. The effective-stress paths after high-speed personal computers, the above analyses are
liquefaction repeat dilative and contractive behaviors that becoming more practical and useful in current practice.
are similar in each cycle, but the corresponding shear strain
amplitude increases continuously. Wang and Dafalias [30]
proposed a strain-history dependence of plastic shear
modulus and used an intrinsic variable to capture the Acknowledgements
reduction in the plastic shear modulus. The degredation
factor from Eq. (5) is modified as follows The waterfront slope example was part of a study
performed for the Port of Oakland; the authors acknowledge
1 Cx
CðxÞ Z (16) their support. The authors also wish to thank Messrs. Ronn
1 C ax Rose and Matt Allen, of the US Army Coprs of Engineers,
and the intrinsic variable is the accumulated plastic Sacramento District, for their permission to present the
deviatoric strain: results of analyses of Success Dam. L. Scheibel and C.C.
Ð qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Chin of Geomatrix Consultants participated in the projects
xZ 2depij depij (17) described in this paper.

Results of simple shear tests for post-liquefaction


deformation of Nevada sand (Kammemer, 2001) were
simulated using the model and Eq. (16). References
The new development of the model is still limited
because there are new laboratory test results on the [1] Wang ZL. Bounding surface hypoplasticity model for granular soils
uniqueness of critical-state line, so that we have not apply and its applications. PhD Dissertation for the University of California
the modification to our practical applications. at Davis, U.M.I. Dissertation Information Service, Order No.
9110679; 1990.
[2] Wang ZL, Dafalias YF, Shen CK. Bounding surface hypoplasticity
model for sand. ASCE, J Eng Mech 1990;116(5):983–1001.
6. Conclusions [3] Wang ZL, Makdisi FI. Implementing a bounding surface hypoplas-
ticity model for sand into the FLAC program. In: Proceedings of the
international symposium on numerical modeling in geomechanics.
A nonlinear effective-stress approach is developed and Minnesota, USA; 1999. p. 483–90.
applied to the seismic stability and deformation analyses of [4] Itasca Consulting Group, Inc. FLAC (Fast Lagrangian Analysis of
earth structures that may experience strong earthquake Continua), Version 4. Minneapolis, MN; 2000.
shaking or induced liquefaction. The key element to [5] Seed HB, Lee KL, Idriss IM, Makdisi FI. The slides in the San
Fernando dams during the earthquake of February 9, 1971. J Geotech
performing such analysis is a constitutive model capable
Eng Div, ASCE 1975;101(GT7):651–88.
of simulating soil behavior under cyclic loading conditions. [6] Seed HB, Idriss IM. Soil moduli and damping factors for dynamic
The model’s capability is demonstrated by comparing its response analyses. Earthquake Engineering Research Center, Univer-
simulation with laboratory test results. sity of California, Berkeley, EERC79-10 1970.
The advantages of using the soil model [1,2] to simulate [7] Yamada Y, Ishihara K. Undrained deformation characteristics of sand
nonlinear behavior and pore-pressure generation and in multi-directional shear. Soils Found 1983;23(1):61–79.
[8] Towhata I, Ishihara K. Undrained strength of sand undergoing cyclic
liquefaction in sandy materials are illustrated using results rotation of principal stress axes. Soils Found 1985;25(2):135–47.
for the seismic deformation and liquefaction analysis of [9] Yang Z, Elgamal A, Parra E. Computational model for cyclic mobility
practical applications that include a rockfill dam, a water and associated shear deformation. J Geotech Geoenviron Eng 2003;
front slope and an earth dam. 129(12):1119–27.
Because the soil properties needed to support such [10] Konder RL, Zelasko JS. A hyperbolic stress–strain formulation for
sands. Proceedings, second Pan American conference on soil
analyses are generally similar to those currently used in mechanics and foundation engineering; 1963. p. 289–324.
performing equivalent linear and simplified effective-stress [11] Martin GR, Finn WD, Seed HB. Fundamentals of liquefaction under
analyses, and considering the increasing availability of cyclic loading. J Geotech Div ASCE 1975;101(GT5):423–38.
252 Z.-L. Wang et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 231–252

[12] Wang ZL, Dafalias YF, Li XS, Makdisi FI. State pressure index for Juran I, editors. Proceedings, grouting, soil improvement and
modeling sand behavior. J Geotech Geoenviron Eng 2002;128(6): geosynthetics, vol. 2. ASCE, Geotechnical Special Publication No.
511–9. 30; 1992. p. 867–78.
[13] Rowe PW. The stress–dilatancy relation for static equilibrium of an [22] Seed HB, Tokimatsu K, Harder LF, Chung R. Influence of SPT
assembly of particles in contact. Proc R Soc Lond; Ser A 1962;269: procedures in soil liquefaction resistance evaluations. J Geotech Eng
500–27. Div ASCE 1985;111(12):1425–45.
[14] Li XS, Wang ZL, Shen CK. SUMDES: a nonlinear procedure for [23] Youd TL, Idriss IM. Summary reports. In: Proceedings of NCEER
response analysis of horizontally layered sites subjected to multi- workshop on evaluation of liquefaction resistance of soils. NCEER-
directional earthquake loading. Davis: Department of civil engineer- 0022; 1997.
ing, University of California; 1992. [24] Seed HB, Idriss IM. Ground motion and soil liquefaction during
[15] Wang ZL, Chang CY, Mok CM. Evaluation of a site response using earthquakes. Monograph series, Earthquake Engineering Research
downhole array data from a liquefied site. Third international Institute. Berkeley, CA: University of California; 1982.
conference on recent advances in geotechnical earthquake engineer- [25] Wang ZL, Egan J, Scheibel L, Makdisi FI. Simulation of earthquake
ing and soil dynamics, San Diego, CA, No. 430 2001. performance of a waterfront slope using fully coupled effective stress
[16] Geomatrix Consultants, Inc. Seismic response, liquefaction, and approach. FLAC and numerical modeling in geomechanics. Proceed-
deformation analysis for platform site WHP-1. Bohai Bay, Offshore ings of the second international FLAC symposium. Lyon, France;
China: Phillips China, Inc.; 2001. 2001.
[17] Hudson M, Idriss IM, Beikae M. User’s manual for QUAD4M: a [26] Allen MG, Nickell JS, Sherer SG. Geotechnical characterization of
computer program to evaluate the seismic response of soil structures success dam for a dam safety earthquake engineering evaluation.
using finite element procedures and incorporating a compliant base. Fourth international conference on recent advances in geotechnical
Davis: Center for geotechnical modeling, University of California; earthquake engineering and soil dynamics and symposium in honor of
1994. Professor W.D. Liam Finn. San Diego; CA 2001.
[18] Makdisi FI, Wang ZL, Edwards WD. Seismic stability of new [27] US Army Corps of Engineers. Report on investigation of failure of
exchequer dam and gated spillway. Twentieth annual USCOLD Sheffield Dam. Santa Barbara, CA; 1949.
lecture series. Seattle, Washington; 2000. p. 437–58. [28] Verdugo R, Ishiharam K. The steady state of sandy soils. Soils Found
[19] Newmark NM. Effects of earthquakes on dams and embankments. 1996;36(2):81–91.
Geotechnique 1965;15(2):139–60. [29] Ishihara K, Tatsuoka F, Yasuda S. Undrained deformation and
[20] Makdisi FI, Seed HB. Simplified procedure for estimating dam and liquefaction of sand under cyclic stress. Soils Found 1975;15(1):
embankment earthquake-induced deformation. J Geotech Eng Div 29–44.
ASCE 1978;104(GT7):849–67. [30] Wang ZL, Dafalias YF. Simulation of post-liquefaction deformation
[21] Egan JA, Hayden RF, Scheibel LL, Otus M, Serventi GM. Seismic of sand. Paper No. 183. ASCE 15th engineering mechanics
repair at Seventh Street Marine Terminal. In: Borden RH, Holtz RD, conference. New York: Columbia University; 2002.

Vous aimerez peut-être aussi