Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

DSC Model for Soil and Interface Including Liquefaction

and Prediction of Centrifuge Test


Shashank K. Pradhan1 and Chandra S. Desai, F.ASCE2
Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Realistic predictions of dynamic soil–structure interaction problems require appropriate constitutive models for the character-
ization of soils and interfaces. This paper presents a unified model based on the disturbed state concept 共DSC兲. The parameters for the
models for the Nevada sand, and sand–metal interface are obtained based on available triaxial test data on the sand and interfaces. The
predicted stress–strain–pore water pressure behavior for the sand using the DSC model is compared with the test data. In addition, a finite
element procedure with the DSC model, based on the generalized Biot’s theory, is used to predict the measured responses for a pile
共aluminum兲 sand foundation problem obtained by using the centrifuge test. The predictions compared very well with measured pore water
pressures. The DSC model is used to identify microstructural instability leading to liquefaction. A procedure is proposed to apply the
proposed method for analysis and design for dynamic response and liquefaction.
DOI: 10.1061/共ASCE兲1090-0241共2006兲132:2共214兲
CE Database subject headings: Constitutive models; Liquefaction; Sand; Interfaces; Parameters; Validation; Piles; Centrifuge;
Predictions.

Introduction faction. The parameters for the DSC model for the sand and in-
terfaces are determined by using available laboratory tests.
Interaction between a superstructure and foundation depends on The DSC model is implemented in a nonlinear finite element
the behavior of soil supporting the foundation and interfaces procedure based on the generalized Biot’s theory. The predictions
between structural materials and soils. Hence, it is necessary to are compared with measured responses for a pile–sand foundation
characterize the behavior of soil and interfaces based on appro- by using centrifuge tests.
priate tests. One of the objectives is to implement such realistic
models in conventional and computer procedures such as the
finite element method. The latter can be used for solving complex Constitutive Model
problems involving dynamic loading, nonlinear material behavior
and presence of water. A constitutive model should be such that it allows for different
Piles and pile foundations represent one of the oldest types of significant factors that affect the behavior of soils and interfaces.
foundation system, and have been in use since prehistoric times. The basic idea of the DSC was introduced by Desai 共1974兲 to
Response of pile foundations to earthquake loading can be chal- characterize softening behavior, and later generalized to model
lenging. Studies on piles during earthquakes are usually difficult; the behavior of wide range of materials 共Desai 2001兲. The basis of
therefore, recourse is made to alternative studies, e.g., centrifuge DSC is that when a material is subjected to external excitation or
testing, instrumented pile tests, shake table tests, etc. loading, microstructural changes take place in the material. Ini-
tially, the material is in the relative intact 共RI兲 state. As the dis-
turbance increases, the material transforms from RI state, through
a process of microstructural adjustment, to the fully adjusted 共FA兲
Scope
state or critical state. The disturbance due to the self-adjustment
may cause anisotropy, microcracking leading to fracture, and dis-
A constitutive model based on the disturbed state concept 共DSC兲 turbance leading to softening and liquefaction. The disturbance,
is used for characterizing the cyclic behavior of sands and inter- D, allows for the effect of interaction and coupling between the
faces. It allows for elastic, plastic and creep strains, disturbance materials in the RI and FA states. As a result, it provides for
共softening or stiffening兲 and microstructural instability or lique- improved characterization compared to other models such as clas-
sical damage 共Kachanov 1986兲, damage with crack interaction
1
Lecturer, Applied Mechanics Dept., Univ. of Baroda, India. 共Bazant 1994兲, and conventional plasticity models 共Desai 2001兲.
2
Regents’ Professor, Dept. of Civil Engineering and Engineering In the DSC used herein, the RI behavior is characterized by
Mechanics, The Univ. of Arizona, Tucson, AZ 85721. using the hierarchical single surface 共HISS兲 plasticity model and
Note. Discussion open until July 1, 2006. Separate discussions must the FA behavior by using the critical state concept 共Roscoe et al.
be submitted for individual papers. To extend the closing date by one
1958; Desai 2001兲. The microstructural instability or liquefaction
month, a written request must be filed with the ASCE Managing Editor.
The manuscript for this paper was submitted for review and possible is identified by locating the critical disturbance during deforma-
publication on September 23, 2003; approved on May 9, 2005. This paper tion. Details of the DSC model are given in 关Desai and Ma
is part of the Journal of Geotechnical and Geoenvironmental Engineer- 共1992兲; Desai 共2000b, 2001兲; Desai et al. 共1998兲; Park and Desai
ing, Vol. 132, No. 2, February 1, 2006. ©ASCE, ISSN 1090-0241/2006/ 共2000兲. A brief description of the HISS and FA models, distur-
2-214–222/$25.00. bance and liquefaction identification are given in the following.

214 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


ec = ec0 − ␭ ln 冉 冊
Jc1
3pa
共3a兲

or

Jc1 = 3pa exp 冉 冊


ec0 − ec

共3b兲

where ec0⫽critical void ratio when Jc1 = 3pa and ␭⫽material con-
stant. At the critical state, the shear stress then can be expressed
as
冑J2D
c
= m̄Jc1 共4兲
Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

where m̄⫽slope of critical state line. The values of m̄, ␭, and ec0
Fig. 1. Yield surface in J1 − 冑J2D plane are obtained from triaxial tests on Nevada sand with different
confining pressures 共Arulmoli et al. 1992兲.

HISS Plasticity Model


Disturbance
The yield function for the HISS–␦0 model is given as
Once RI and FA states are defined, the next step is to formulate
J2D the disturbance function, D. There are a number of ways to define
F= − F bF s = 0 共1兲 D, which are explained in Desai 共2001兲; here, the disturbance, D,
p2a is defined as
where ␴i − ␴a
共5兲
冋 册 冋 册
D=
J1 + 3R n
J1 + 3R 2 ␴i − ␴c
Fb = − ␣ +␥
pa pa where the superscripts i, a, and c denote the relative intact, ob-
served, and fully adjusted state and ␴ denotes the measure of
冑27 J3D stress such as shear stress, ␶, or the second invariant of deviatoric
Fs = 共1 − ␤Sr兲−0.5, Sr = 1.5
stress, 冑J2D. At the beginning of loading, the material can be
2 J2D assumed to be entirely in the RI state and so the disturbance D is
The term Fb is related to the shape of the yield surface in zero. As the loading increases, the deformation increases and con-
J1 − 冑J2D space, Fig. 1 共with Sr⫽constant兲; and Fs is related to the sequently the disturbance D increases. When D approaches unity,
shape of the yield surface in the octahedral space 共J1⫽constant兲. the material is in ultimate or critical state. The following function
Sr⫽stress ratio; J1, J2D, and J3D⫽invariants of stress tensor ␴ij; is used to represent disturbance D:
pa⫽atmospheric pressure constant; ␥ and ␤⫽materials param- Z
D = Du共1 − e−A ␰D兲 共6兲
eters associated with the ultimate behavior; n⫽phase change pa-
rameter; and R denotes tensile strength and is related to cohesive where A, Z, and Du⫽material parameters and ␰D⫽trajectory of or
strength. The hardening function, ␣ is defined in terms of accu- accumulated deviatoric plastic strains defined as
mulated plastic strain 共trajectory兲 as
d␰D = 冑dEijpdEijp 共7a兲
h1 and
␣= 共2a兲
␰ h2
1
where h1 and h2⫽material parameters and ␰⫽trajectory of or ac- dEijp = d␧ijp − ␦ijd␧kk
p
共7b兲
3
cumulated plastic strain expressed as
where EijP⫽tensor of deviatoric plastic strain and d␧kk
p
⫽plastic

␰= 冕冑 d␧ijpd␧ijp 共2b兲
volumetric strain.

where d␧ijp⫽tensor of incremental plastic strains. Incremental Equations

Based on the RI and FA behavior and disturbance, the following


FA Model incremental constitutive equations are derived:
d␴a = 共1 − D兲Cid␧i + DCcd␧c + dD共␴c − ␴i兲 共8兲
The fully adjusted state of the material is an asymptotic or ulti-
mate state, which is modeled using the critical state concept where C ⫽constitutive matrix for the RI behavior, for which the
i

共Roscoe et al. 1958兲. At the critical state, shear deformations can HISS plasticity model, Eq. 共1兲, is used; Cc⫽constitutive matrix
continue without further changes in the volume of the material. for FA behavior, for which the critical state, Eq. 共3兲 and 共4兲 are
For geologic materials, the final void ratio ec of the material at- used; ␴ and ␧⫽stress and strain vectors; and d denotes increment.
tained in a shear test is related to the hydrostatic stress Jc1 at the Finite Element Method: The previous DSC model, Eq. 共8兲, is
critical state as implemented in the finite element procedure based on the gener-

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006 / 215

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Typical plot for pressure versus number of cycles: Fig. 3. Determination of critical disturbance, Dc, for typical test:
␴3 = 40 kPa ␴3 = 80 kPa

alized Biot’s theory. The finite element equations are derived as


共Desai 2000a; Shao and Desai 2000; Desai 2001兲 tion 共degradation兲 in effective mean pressure during cyclic load-

册再 冎 冋 册再 冎
ing, and can be identified by using the following steps:

冋 M uu M uw
M wu M ww


+
O O
O Cww


1. Obtain cyclic stress–strain data 共from triaxial tests兲 with dif-
ferent confining pressures.
2. Plot time 共seconds兲 versus D 共disturbance兲.


冤 冥
3. Locate value of the critical disturbance Dc based on the time

再冎
Nu,i␴dV at which liquefaction has occurred. Fig. 3 shows determina-
v fu tion of critical disturbance for typical test with confining


+ = 共9兲
fw pressure of 80 kPa, for the Nevada sand.
Nw,i pdV 4. Calculate average value of critical disturbance Dc which
v indicates a state of instability that can identify initial
liquefaction.
where M⫽mass matrices; C̄⫽damping matrix; U and
Based on this procedure, the values of critical disturbance for the
W⫽vectors of 共solid兲 displacements and relative displacement
three cyclic tests 共for later Nevada sand兲 were found and average
between solids and fluids, respectively; Nu and Nw⫽interpolation
value of critical disturbance was found as 0.86.
functions corresponding to U and W, respectively; ␴⫽stress vec-
Generalized Models: A number of other generalized models
tor; p⫽pore water pressure; fu and fw⫽load vectors; and the sub-
have been proposed, e.g., Mroz et al. 共1978兲, Pestana and Whittle
script comma, in Nu,, denotes derivative. Eq. 共9兲 is solved by
共1999兲, Pastor et al. 共2000兲, and Elgamal et al. 共2002兲. These
using a time integration scheme 共Newmark兲, and provide for dis-
models can account for various factors beyond those in previously
placements, strains, stresses, pore water pressures and disturbance
available 共plasticity兲 models. They are based on continuum plas-
at various nodes and elements at different times.
ticity, and introduce modifications to simulate behavioral features
such as degradation 共softening兲. However, very often, a deform-
ing material involves discontinuities due to microcracking result-
Liquefaction
ing from microstructural modifications. The continuum plasticity
models may not account for such discontinuous deformations. On
In order to predict liquefaction, the disturbance function D is
the other hand, the DSC model accounts for the changing micro-
defined with respect to the observed, intact and critical effective
structure resulting in degradation or strengthening, and also dis-
mean stress J1 / 3 = ␴⬘ from cyclic triaxial tests:
continuous deformations. It also allows intrinsically for the defi-
共␴⬘兲i − 共␴⬘兲a nition of instability or liquefaction based on the internal response
D= 共10兲 共disturbance兲 of a deforming material. Furthermore, introduction
共␴⬘兲i − 共␴⬘兲c
of disturbance with continuum models 共elasticity or plasticity兲 for
The first term 共␴⬘兲i is calculated from the relative intact state, RI behavior, may not involve any more effort than the modifica-
which is projected smoothly from the peak of the first cycle. The tion of continuum plasticity models. Other models such as
second term 共␴⬘兲a is calculated at the peak of each cycle from the MIT-S1 共Pestana and Whittle 1999; Elgamal 2002兲 based on the
test data for each test. The third term 共␴⬘兲c is calculated for each bounding surface plasticity may also involve a greater number of
test as parameters compared to those in the DSC model for comparable

共␴⬘兲c = Jc1 = 3pa exp 冉 冊


ec0 − ec

共11兲
capabilities 共Desai 2001兲.

Fig. 2 shows a typical plot of effective mean stress 共␴⬘兲 versus DSC Parameters
number of cycles 共N兲 from the test data for confining pressure,
␴3 = 40 kPa, for Nevada sand used in the centrifuge test, described To determine the parameters, Nevada sand test results were used
later. from soil data by The Earth Technology Corporation, Huntington
The critical disturbance due to microstructural instabilities can Beach, Calif., for VELACS 共Arulmoli et al. 1992兲. For determin-
be defined to identify liquefaction. It can be related to the reduc- ing parameters, three undrained monotonic triaxial tests, and three

216 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Table 1. Parameters for Nevada Sand and Sand–Aluminum Interface
Nevada sand
aluminum
Group Subgroup Parameters Nevada sand interface
1 Elastic E 40,848.8 kPa 14.6 MPa
parameters ␯ 0.316 0.384
Plasticity ␥ 0.0675 0.246
parameters ␤ 0.0 0.000
3R 0.0 0.0
n 4.1 3.350
h1 0.1245 0.620
Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

h2 0.0725 0.570
2 Critical state m̄ 0.22 0.304
parameters ␭ 0.02 0.0278
ec0 0.712 0.791
3 Disturbance Du 0.99 0.990
parameters Z 0.411 1.195
A 5.02 0.595

undrained cyclic triaxial tests with confining pressures, ␴3, of 40,


80, and 160 kPa at 60% relative density were used. Details of
Fig. 4. Details of mesh and boundary conditions
evaluating these parameters from the test data are given in
Pradhan and Desai 共2002兲. The parameter values are presented in
Table 1. boundary conditions is shown in Fig. 4. One eight-noded element
and axisymmetric idealization was used to model the sand
sample. The mesh was subjected to increments of confining pres-
DSC Model for Interfaces sure and then increments of deviatoric stress.
Typical comparisons between the predicted results and test
Interface test data for the aluminum–Nevada sand were not avail- data only for the independent test are shown in Figs. 5 and 6. The
able for the prediction from the centrifuge tests, described subse- degradation of effective stress shows good agreement, with a dif-
quently. Hence, a neural network procedure was used 共Pradhan
and Desai 2002兲. Here two sets of data were used for the training
of the neural network. Set 1 consisted of the Ottawa sand param-
eters using the triaxial testing 共Park and Desai 2000兲 and the
Ottawa sand–steel interface using the cyclic multidegree of free-
dom 共CYMDOF兲 device 共Alanazy 1996; Desai and Rigby 1997兲.
The second set consisted of the parameters for marine clay at
Sabine, Tex., from triaxial and multiaxial testing 共Katti and Desai
1995兲 and marine clay–steel interface using the CYMDOF 共Shao
and Desai 2000兲. It was assumed that the steel and aluminum
behave approximately similar in interface with the soil. Also, the
Nevada sand gradation curve falls between those of the Ottawa
sand and the Sabine clay. The parameters for the Nevada sand
found from the available triaxial data 共Arulmoli et al. 1992兲 were
used in the neural network, which provided estimates for Nevada
sand–aluminum interface parameters. They are given in Table 1
and were used for the interfaces in the pile test using the centri-
fuge device, described later.

Validations

Laboratory Tests
The model parameters were used to back predict laboratory tests
共Dr = 60%兲 that were employed to find parameters and also an
independent test 共Dr = 40%兲 whose results were not used to find
parameters 共Pradhan and Desai 2002兲. The predictions were ob-
tained using the DSC-DYN2D program, Desai 共2000a兲. The finite
element mesh used for back prediction of laboratory tests with Fig. 5. Effective mean stress versus time, ␴3 = 160 kPa, Dr = 40%

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006 / 217

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Disturbance D versus ␰D, ␴3 = 160 kPa, Dr = 40%

ference in the peaks after about four cycles. The predicted pore
water pressures are in very good agreement with the test data,
except that the numerical prediction shows more detailed varia-
tions. The disturbance from predictions and test data correlates
very well.

Centrifuge Test
Prediction of the behavior of pile foundations in liquefiable sands
under earthquake loading can be a challenging problem. The best
way to understand such a problem and develop a design and
analysis procedure is comparison with full-scale field data, which
can be expensive. Hence, centrifuge studies represent a useful
development in studying fundamental mechanisms of soil-pile
structure interaction. The centrifuge test data used herein were
performed at the National Geotechnical Centrifuge at University
of California, Davis, Calif. 共Kutter et al. 1991, 1994; Wilson et al.
1997a,b,c兲. It has a radius of 9.0 m and is equipped with large
shaking table. It has a maximum model mass of about 2,500 kg,
and available bucket area of 4.0 m2 and a maximum centrifugal
acceleration of 50 g. Details of the centrifuge can be found in
Wilson et al. 共2000兲.

Fig. 8. Finite element mesh for pile: 共a兲 mesh for total domain and
共b兲 details of mesh around pile

Five containers of soil–structure systems were tested at a cen-


trifugal acceleration of 30 g. Full details for each test can be
found in Wilson et al. 共1997a,b兲. In this study, earthquake event J
in the model, referred to as CSP3, was simulated. The soil profile
Fig. 7. Schematic of instrumented pile in centrifuge test CSP3 关data in this container consisted of two horizontal layers. The upper 9.3
adapted from Wilson et al. 共2000兲兴 m layer was medium–dense Nevada sand 共Dr = 55% 兲 and lower

218 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Base motion acceleration 关data adapted from Wilson et al.


共2000兲兴

11.4 m thick layer was dense Nevada sand 共Dr = 80%兲, Fig. 7.
Details of the structure and pile systems can be found in Wilson et
al. 共1997a,b,c兲. Foundation models included single pile founda-
tions, a four pile group, and a nine pile group, with superstructure
mass typically being 500 kN per each supporting pile.
In this analysis, a single prototype pile near the middle, Fig. 8,
made of aluminum pipe with 0.67 m diameter, 72 mm wall thick-
ness, and 20.6 and 16.8 m embedment depth was used. Linear
elastic properties were used for the pile: E = 70.0 GPa and
␯ = 0.33. The DSC model was used for the sand and interface; the
parameters are given in Table 1.
It would be desirable to use a three-dimensional finite element
共FE兲 for the problem; however, such procedures were not readily
available. In their two-dimensional 共2D兲 analysis, Johnson et al.
共1998兲 and Iai 共1998兲 have used “springs” with the capacity equal
to the expected value of the ultimate lateral traction that the soil
exerts against the pile, to connect the pile to the 2D mesh. This is
assumed to allow for motions between the pile and soil, and
three-dimensional effects. It is not clear how the behavior in the
third direction is accounted for, and whether the prepeak and
postpeak responses of the zone between the soil and the pile are
included in this model.
Fig. 10. Comparisons for pore water pressures for Element 139 near
In this work, the 共thin兲 zone of soil around the pile, which is
pile
usually weaker than the soil, is treated as an interface with finite
dimensions that allows the relative motions between the pile and
soil and includes prepeak, peak and postpeak behavior between ␴⬘h = K j␴⬘v 共12b兲
the soil and pile.
In the present analysis, as an approximation, the single pile
K0 = ␯/共1 − ␯兲 共12c兲
was modeled by using two-dimensional FE procedure 共Desai
2000a兲, with plane strain idealization. The plane strain idealiza-
tion for such a pile is considered to be more appropriate than the p = ␥ wh 共12d兲
axisymmetric idealization, partly because the latter allows mainly where ␴⬘v and ␴h⬘⫽effective vertical and horizontal stresses at
axial behavior. Similar plane strain idealizations for piles have depth, h; ␥s⫽submerged unit weight of soil; ␥w⫽unit weight of
been used previously 共Desai et al. 1974; Anandarajah 1992; Ellis water; p⫽initial pore water pressure; and K0⫽coefficient of earth
1997; Fujii et al. 1998兲. pressure at rest.
Fig. 8共a兲 shows the FE mesh, which consisted of 302, 4-noded
elements with 342 nodes. Detailed mesh around the pile is shown
in Fig. 8共b兲. Boundary BC was restrained in the y direction, and Comparison of Results
AB and CD were free to move in both x and y directions. The
base of the mesh was subjected to the acceleration–time history Two finite element analyses were performed, with and without
shown in Fig. 9 共Wilson et al. 2000兲. interfaces. The predictions and observed pore water pressure were
Initial in situ stresses and pore pressures at centers of all ele- compared for a number of elements: 1, 5, 8, 9, 12, and 139.
ments were introduced by using the following expressions: Typical results for Element 139 near the pile, Element 9 away
from the pile but at the same level as Element 139, and Element
␴⬘v = ␥sh 共12a兲 5 away from the pile but at a lower level are presented in Figs.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006 / 219

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Comparisons of pore water pressures for Element 9, away


from pile on same level as 139 Fig. 12. Comparisons of pore water pressures for Element 5, away
from pile at lower level

10–12, respectively. It can be seen from these figures that the


model with the interface provides improved correlation with the 共small兲 strength after the Port Island, Kobe, Japan earthquake.
test results. Pore water pressure, with interface, in Element 139 Thus, in the DSC, the time to liquefaction 共Dc ⬇ 0.86兲 is lower
reaches the initial vertical effective stress, ␴⬘v, after about 9 s than that at complete loss of strength, Fig. 3. Also, it is believed
indicating liquefaction, while that without interface does not in- that the microstructural instability or liquefaction can occur be-
dicate liquefaction. Further, the predictions for both with and fore the complete collapse of the microstructure. Further investi-
without interface, show similar values and trends for elements gation regarding this aspect is desirable.
away from the pile. This can be mainly because the influence of
the relative motions at interface diminishes away from the pile.
Table 2 compares the time when the disturbance reaches its
Table 2. Times to Liquefaction from Disturbance and Conventional
critical value, Dc = 0.86, with the conventional procedure in which
Methods
liquefaction occurs when the excess pore water pressure, Uw,
reaches the initial effective pressure 共␴⬘v兲, for typical elements. Element Uw = ␴⬘v 共s兲 D = Dc 共s兲
The times to liquefaction correlate well. However, those from the 143 1.74 1.23
DSC model are lower than those from the conventional proce- 130 1.83 1.29
dure, implying that the microstructural instability can occur ear- 104 2.55 1.92
lier than the time to liquefaction obtained from the conventional
78 3.36 2.67
procedure. The conventional procedure implies that the soil does
52 3.81 2.94
not possess any strength when liquefaction occurs. In fact, it has
26 9.03 8.22
been reported, e.g., by Desai 共2000b兲, that the soil retained a

220 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Disturbance versus time in Elements 126 and 139

Fig. 13 shows typical variation of disturbance with time in the


interface Element 139 and in the adjacent Element 126. These
figures indicate that the liquefaction for this problem can occur in
the interface element earlier than in the surrounding elements.
The top bucket ring at which the displacements were measured
was not included in the pile mesh. Hence, comparisons for dis-
placements were not available. Also, because the 2D plane strain
idealization does not incorporate the moment of inertia of pile, the
bending moments were not available for comparison with the
recorded bending moments. Hence, the predicted accelerations
from the analysis were compared with the experimental values
near the pile top; the experimental data, Fig. 14, may have been
affected by use of different computer programs. Fig. 14 shows
comparison between predicted accelerations near the top of the
pile from the analysis that includes the interface, and the mea-
sured values. The peak values differ by about 12.5%; overall, the
correlation is considered to be satisfactory.
Figs. 15共a and b兲 show disturbance contours after typical
times, 2 and 20 s. They indicate that critical disturbance
Dc = 0.86 and greater, signifying liquefaction, grows gradually
from the top to a significant extent at 20 s.

Fig. 15. Contours of disturbance at various times: 共a兲 2 s and 共b兲


20 s

Conclusions

The disturbed state concept is used to characterize the behavior of


soils 共sands兲 and interfaces between structural materials 共piles兲
and sands. The parameters for the DSC model are determined
based on the available triaxial tests for Nevada sand. Those for
interface are obtained based on tests for interfaces between Ot-
tawa sand and steel, marine clay and steel, and the triaxial test
Fig. 14. Comparisons between predicted and experimental data for the Nevada sand, through a neural network procedure. A
acceleration near the pile top finite element procedure with the DSC model based on the gen-

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006 / 221

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222


eralized Biot’s theory for saturated materials is used to validate cyclic mobility and post-liquefaction site response.” Soil Dyn. Earth-
the model with test data for the sand. quake Eng., 22共4兲, 259–271.
The procedure is also applied to predict the behavior of pile– Ellis, E. A. 共1997兲. “Soil-structure interaction for full-height piled bridge
Nevada sand by using measurements from the centrifuge tests at abatements constructed on soft clay.” PhD thesis, Cambridge Univ.,
Cambridge, U.K.
the University of California, Davis, California. The correlations
Fujii, S., Cubrinovski, M., Tokimatsu, K., and Hayashi, T. 共1998兲.
between predictions and measurements in terms of pore water
“Analyses of damaged and undamaged pile foundations in liquefied
pressures are considered to be highly satisfactory, and those in
soils during the 1995 Kobe earthquake.” Geotechnical Special Publi-
terms of accelerations near the top of the pile are considered to be
cation 75, Proc., Geotech. Earthquake Eng. and Soil Dynamics III,
satisfactory. The FE results are used to analyze the liquefaction ASCE, Reston, Va., 1187–1198.
behavior in the sand and interface. Iai, S. 共1998兲. “Seismic analysis and performance of retaining structures.”
Based on the results reported herein, the DSC model can Geotechnical Special Publication 75, Proc., Geotech. Earthquake
provide satisfactory and improved predictions for the be-
Downloaded from ascelibrary.org by California State University- Fresno on 10/03/19. Copyright ASCE. For personal use only; all rights reserved.

Eng. And Soil Dynamics III, ASCE, Reston, Va., Vol. 2, 1020–1044.
havior of dynamic soil–structure interaction problems including Johnson, R. K., Riffenburgh, R., Hodali, R., Moriwaki, Y., and Tan, P.
liquefaction. 共1998兲. “Analysis and design of a container terminal wharf at the Port
of Long Beach.” Proc., Ports 98, ASCE, Reston, Va., 436–444.
Kachanov, L. M. 共1986兲. Introduction of continuum damage mechanics,
Martinus Nijholft, Dordrecht, The Netherlands.
Acknowledgments Katti, D. R., and Desai, C. S. 共1995兲. “Modeling and testing of cohesive
soil using disturbed-state concept.” J. Eng. Mech., 121共5兲, 648–658.
The research reported here was partly supported by Grant Nos. Kutter, B. L., Idriss, I. M., Kohnke, T., Lakeland, J., Li, X. S., Sluis, W.,
Zeng, X., Tauscher, R., Goto, Y., and Kubodera, I. 共1994兲. “Design of
CMS 9115316 and CMS-9732811 from the National Science
large earthquake simulator at UC Davis.” Proc., Int. Conf. Centrifuge
Foundation, Washington, D.C.
94, Singapore, 169–175.
Kutter, B. L., Li, X. S., Sluis, W., and Cheney, J. A. 共1991兲. “Perfor-
mance and instrumentation of the large centrifuge at Davis.” Proc.,
Conf. Centrifuge 91, Boulder, Colo. 19–26.
References Mroz, Z., Norris, V. A., and Zienkiewicz, O. C. 共1978兲. “An anisotropic
hardening model for soils and its application to cyclic loading.” Int. J.
Alanazy, A. S. 共1996兲. “Testing and modeling of sand-steel interfaces Numer. Analyt. Meth. Geomech., 2共3兲, 208–221.
under static and cyclic loading.” PhD dissertation, Dept. of Civil Park, I. J., and Desai, C. S. 共2000兲. “Cyclic behavior and liquefaction
Engineering and Engineering Mechanics, The Univ. of Arizona, using disturbed state concept.” J. Geotech. Geoenviron. Eng., 126共9兲,
Tucson, Ariz. 834–846.
Anandarajah, A. 共1992兲. “Fully coupled analysis of a single pile founded Pastor, M., Chan, A. H. C., and Zienkiewicz, O. C. 共2000兲. “Constitutive
in liquefiable sands,” Report, Dept. of Civil Engineering, The Johns and numerical modeling of liquefaction.” Modeling in geomechanics,
Hopkins Univ., Baltimore, Md. Chap. 5, M. Zaman, G. Gioda, and J. Booker, eds., Wiley, Chichester,
Arulmoli, K., Muraleetharan, K. K., Hossain, M. M., and Fruth, L. S. U.K.
共1992兲. “VELACS: Verification of liquefaction analysis by centrifuge Pestana, J. M., and Whittle, A. J. 共1999兲. “Formulation of a unified con-
studies.” Soil data report, The Earth Technology Corporation, Irvine, stitutive model for clays and sands.” Int. J. Numer. Analyt. Meth.
Calif. Geomech., 23共12兲, 1215–1243.
Bazant, Z. P. 共1994兲. “Nonlocal damage theory based on micromechanics Pradhan, S. K., and Desai, C. S. 共2002兲. “Dynamic soil-structure interac-
of crack interactions.” J. Eng. Mech., 120共3兲, 593–617. tion using disturbed state concept and artificial neural networks for
Desai, C. S. 共1974兲. “A consistent finite element technique for work- parameter evaluation.” NSF Rep., Dept. of Civil Engineering and
softening behavior.” Proc., Int. Conf. on Computational Methods in Engineering Mechanics, The Univ. of Arizona, Tucson, Ariz.
Nonlinear Mechanics, J. T. Oden, et al., eds., Univ. of Texas, Austin, Roscoe, K. H., Schofield, A., and Wroth, C. P. 共1958兲. “On yielding of
Tex. soils.” Geotechnique, 8, 22–53.
Desai, C. S. 共2000a兲. “DSC-DYN2D, computer code for dynamic and Shao, C., and Desai, C. S. 共2000兲. “Implementation of DSC model and
static analysis: Dry and saturated materials.” C. S. Desai, Reps. Part I, application for analysis of field pile tests under cyclic loading.” Int. J.
II and III, Tucson, Ariz. Numer. Analyt. Meth. Geomech., 24共6兲, 601–624.
Desai, C. S. 共2000b兲. “Evaluation of liquefaction using disturbed state Wilson, D. W., Boulanger, R. W., and Kutter, B. L. 共2000兲. “Observed
and energy approaches.” J. Geotech. Geoenviron. Eng., 126 共7兲, seismic lateral resistance of liquefying sand.” J. Geotech. Geoenviron.
618–631. Eng., 126共10兲, 898–906.
Desai, C. S. 共2001兲. Mechanics of materials and interfaces: The disturbed Wilson, D. W., Boulanger, R. W., and Kutter, B. L. 共1997a兲. “Soil-pile-
state concept, CRC Press, Boca Raton, Fla. superstructure interaction at soft or liquefiable soil sites–Centrifuge
Desai, C. S., Johnson, L. D., and Hargett, C. M. 共1974兲. “Analysis of data report for CSP1.” Rep. No. UCD/CGMDR-97/02, Center forGeo-
pile-supported gravity lock.” J. Geotech. Eng. Div., Am. Soc. Civ. technical Modeling, Dept. of Civil and Environmental Engineering,
Eng., 100共9兲, 1009–1029. Univ. of California, Davis, Calif.
Desai, C. S., and Ma, Y. 共1992兲. “Modelling of joints and interfaces using Wilson, D. W., Boulanger, R. W., and Kutter, B. L. 共1997b兲. “Soil-pile-
the disturbed state concept.” Int. J. Numer. Analyt. Meth. Geomech., superstructure interaction at soft or liquefiable soil sites–Centrifuge
16共9兲, 623–653. data report for CSP5.” Rep. No. UCD/CGMDR-97/06, Center
Desai, C. S., Park, I. J., and Shao, C. 共1998兲. “Fundamental yet simplified for Geotechnical Modeling, Dept. of Civil and Environmental Engi-
model for liquefaction instability.” Int. J. Numer. Analyt. Meth. Geo- neering, Univ. of California, Davis, Calif.
mech., 22共9兲, 721–748. Wilson, D. W., Boulanger, R. W., and Kutter, B. L. 共1997c兲. “Soil-pile-
Desai, C. S., and Rigby, D. B. 共1997兲. “Cyclic interface and joint shear superstructure interaction at soft or liquefiable soil sites–Centrifuge
device including pore pressure effects.” J. Geotech. Geoenviron. Eng., data report for CSP3.” Rep. No. UCD/CGMDR-97/04, Center for
123共6兲, 568–579. Geotechnical Modeling, Dept. of Civil and Environmental Engineer-
Elgamal, A., Yang, Z., and Parra, E. 共2002兲. “Computational modeling of ing, Univ. of California, Davis, Calif.

222 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / FEBRUARY 2006

J. Geotech. Geoenviron. Eng., 2006, 132(2): 214-222

You might also like