SlideShare una empresa de Scribd logo
1 de 18
Descargar para leer sin conexión
Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
Dynamics of nonlinear porous media with applications
to soil liquefaction
Radu Popescua
, Jean H. Prevostb,, George Deodatisc
, Pradipta Chakraborttya
a
Faculty of Engineering and Applied Science, Memorial University, St. John’s, Nfld, Canada A1B 3X5
b
Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ 08544, USA
c
Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, NY 10027, USA
Abstract
This paper provides a description of the extension of Biot’s theory for dynamic behavior of saturated porous media into the nonlinear
regime that was introduced by the second author in 1980. It also provides a finite element implementation of this extension and two
numerical applications involving the seismic behavior of saturated soil deposits. In the first numerical application, the dynamic
interaction between liquefying soil and a structure sitting on the ground surface is examined, with emphasis on the interplay between the
seismic loading rate and the (evolving) characteristic frequency of the soil–structure system. The attenuation of seismic energy as the
seismic waves pass through softened soil is also discussed. The second numerical application involves the seismically induced liquefaction
of stochastically spatially variable soils. It is explained why more pore-water pressure is generated in a heterogeneous soil than in a
corresponding uniform soil. Comparisons are also provided with experimental centrifuge data.
r 2006 Elsevier Ltd. All rights reserved.
1. Introduction
Soils consist of an assemblage of particles with different
sizes and shapes which form a skeleton whose voids are
filled with water and air or gas. The word ‘‘soil’’, therefore,
implies a mixture of assorted mineral grains with various
fluids. Hence, soil in general must be looked at as a one-
phase (dry soil), two-phase (saturated soil), or multi-phase
(partially saturated soil) material whose state is to be
described by the stresses and displacements (velocities)
within each phase. There are still uncertainties on how to
deal analytically with partially saturated soils, and this is
currently an area of active research. Attention is restricted
in the following to fully saturated soils. The stresses carried
by the soil skeleton are conventionally called ‘‘effective
stresses’’ in the soil mechanics literature [1], and those in
the fluid phase are called ‘‘pore-fluid pressures’’.
In a saturated soil, when free drainage conditions
prevail, the steady-state pore-fluid pressures depend only
on the hydraulic conditions and are independent of the soil
skeleton response to external loads. Therefore, in that case,
a single-phase continuum description of soil behavior is
certainly adequate. Similarly, a single-phase description of
soil behavior is also adequate when no drainage (i.e., no
flow) conditions prevail. However, in intermediate cases in
which some flow can take place, there is an interaction
between the skeleton strains and the pore-fluid flow. The
solution of these problems requires that soil behavior be
analyzed by incorporating the effects of the transient flow
of the pore-fluid through the voids, and therefore requires
that a two-phase continuum formulation be available for
saturated porous media. Such a theory was first developed
by Biot [2–4] for an elastic porous medium. Applications of
Biot’s theory to finite element analysis of saturated porous
media have been reported by numerous authors (see, e.g.,
Refs. [5,6] for a summary of the wide range of existing
finite element formulations). Mesgouez et al. [7] present a
review of work on applications of Biot’s theory in transient
wave propagation in saturated porous media. In most
formulations, the soil skeleton has either elastic or visco-
elastic behavior.
An extension of Biot’s theory into the nonlinear inelastic
range is presented by Prévost [8], where soil is viewed as a
ARTICLE IN PRESS
www.elsevier.com/locate/soildyn
0267-7261/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.soildyn.2006.01.015
Corresponding author. Tel.: +1 609 258 5424; fax: +1 609 258 2760.
E-mail address: prevost@princeton.edu (J.H. Prevost).
multi-phase medium and the modern theories of mixtures
developed by Green and Naghdi [9] and Eringen and
Ingram [10] are used. General mixture results can be shown
through formal linearization of the field and constitutive
equations, to reduce Biot linear poroelastic model (see, e.g.,
Ref. [11]). The resulting equations are summarized in
Section 2.
The first part of this paper presents briefly Biot’s basic
theory of dynamics of saturated porous media, its
extension into the nonlinear regime [8], and its implemen-
tation in the finite element code DYNAFLOW [12]. Two
numerical examples are presented next, illustrating this
extension of Biot’s theory into the nonlinear dynamic
behavior of saturated soils. Both examples are presented in
relation to centrifuge experiments of structures on homo-
geneous or non-homogeneous saturated soil deposits
subjected to seismic loads. The first numerical example
provides results of a study aimed at better understanding
the loading and failure mechanisms of structures on
liquefiable soil, while the second example examines the
mechanisms of excess pore-water pressure (EPWP) build-
up and liquefaction of natural soil deposits exhibiting
stochastic spatial variability of their properties. The first
example is an original study extending the results of
previous research on earth dams (e.g., [13]) to structures on
liquefiable soil deposits. The second example is also an
original study that continues and expands on previous
work of the authors on seismically induced liquefaction of
randomly heterogeneous soil [14–17]. Based on detailed
analysis of both experimental and numerical results, it
provides an interpretation for a particularly interesting and
important behavior detected in previous work, namely that
more pore-water pressure is generated by seismic loads in a
heterogeneous soil than in an equivalent uniform soil. The
numerical calculations presented in both examples are
validated based on results of recent centrifuge experiments
[18,19]. Besides demonstrating the capabilities of the
nonlinear extension of Biot’s theory implemented in
DYNAFLOW to handle relatively complex phenomena
(such as dynamic soil–structure interaction and seismic
behavior of randomly variable soils), these studies provide
new insight into the mechanisms of seismic behavior of
soil–structure systems and interesting practical results for
geotechnical design.
2. Dynamics of porous media
2.1. Basic theory
During deformation, the solid particles, which form the
soil skeleton, undergo irreversible motions such as slips at
grain boundaries, creations of voids by particles coming
out of a packed configuration, and combinations of such
irreversible motions. When the particulate nature and the
microscopic origin of the phenomena involved are not
explicitly sought, phenomenological equations then pro-
vide an adequate description of the behavior of the various
phases which form the soil medium. In multi-phase
theories, the conceptual model is, thus, one in which each
phase (or constituent) enters through its averaged proper-
ties obtained as if the particles were smeared out in space.
In other words, the particles nature of the constituents is
described in terms of phenomenological laws as the
particles behave collectively as a continuum. Soil is, thus,
viewed herein as consisting of a solid skeleton interacting
with the pore fluids.
Balance of linear momentum equations for each
constituent then can be written simply as (see, e.g., Ref. [8])
rS
aS
¼ r  s0S
 ð1  nw
Þ  rpw
 x  ðvS
 vw
Þ þ rS
b, (1)
rw dvw
dt
¼ rw
ðvS
 vw
Þ  rvw
 nw
rpw
þ x  ðvS
 vw
Þ þ rw
b, ð2Þ
where the motion of the solid phase is used as the reference
motion. In Eqs. (1) and (2), s0s
is the solid (effective) stress,
as
the solid acceleration, vs
(vw
) the solid (fluid) velocity, b
the body force (per unit mass), pw
the pore-fluid pressure,
rS
¼ ð1  nw
Þrs and rw
¼ nw
rw with rs the solid mass
density, rw the fluid mass density, and nw
the porosity;
x ¼ nw2
gwk1
where k is the hydraulic conductivity (L/T)
and gw ¼ rwg the fluid unit weight, g ¼ ðjjbjjÞ acceleration
of gravity.
In the case of a compressible pore fluid, the pore-fluid
pressure is determined from the computed velocities
through time integration of the following equation:
dpw
dt
¼ ðlw
=nw
Þ r  ð1  nw
ÞvS
þ r  ðnw
vw
Þ
 
, (3)
where lw
is the fluid bulk modulus. In the case of an
incompressible pore fluid, the pore-fluid pressure is
determined from the computed velocities through the
following equation:
pw
¼ ðlw
=nw
Þ r  ð1  nw
ÞvS
þ r  ðnw
vw
Þ
 
, (4)
where lw
is a penalty parameter used to enforce the
incompressibility constraint [20,21].
2.2. Weak form/semi-discrete finite element equations
The weak formulation associated with the initial
boundary value problem is obtained by proceeding along
standard lines (see, e.g., Refs. [22,23]). The associated
matrix problem is obtained by discretizing the domain
occupied by the porous medium into non-overlapping finite
elements. Each element is defined by nodal points at which
shape functions are prescribed. In general, two sets of
shape functions are required for the solid displacement and
the fluid velocity field, respectively. However, attention in
the following is restricted to low-order (i.e., four-node
plane and eight-node brick) finite elements which are the
most efficient in nonlinear analysis, and thus the same
shape functions are used for both the solid and the fluid.
The shape functions for the solid displacement and fluid
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 649
velocity associated with node A are denoted by NA
in the
following. They satisfy the relation NA
ðxB
Þ ¼ dAB in which
xB
denotes the position vector of node B and dAB ¼ Kro-
necker delta. The Galerkin counterpart of the weak
formulation is expressed in terms of the shape functions
and gives rise to the following system of equation:
Ms
0
0 Mw
 #
as
aw
( )
þ
Z Z
ZT
Z
 #
vs
vw
( )
þ
ns
nw
( )
¼
fs ext
fw ext
( )
, ð5Þ
where Ma
, aa
, va
, na
, and fa ext
represent the (generalized)
mass matrix, acceleration, velocity, and force vectors,
respectively (a ¼ s; w for the solid and fluid phases,
respectively). The element contributions to node A from
node B for directions i and j to the matrices appearing in
Eq. (5) are defined in the following. The terms are
integrated over the spatial domain occupied by the
element, Oe
. For computational simplicity, a diagonal
‘‘lumped mass’’ matrix is used:
mAB
ij
 a
¼ dijdAB
Z
Oe
ra
NA
dO. (6)
The momentum transfer terms give rise to the damping
matrix Z:
ZAB
ij
 
¼
Z
Oe
NA
xijNB
dO. (7)
The external force ðfa
Þext
(i.e., body force, surface
traction) is computed as follows:
ðfA
i Þa ext
¼
Z
Oe
ra
biNA
dO þ boundary terms: (8)
The internal stress forces na
are computed as follows:
ðnA
i Þs
¼
Z
Oe
ðs0s
ij  ns
pwdijÞNA
;j dO (9)
and
ðnA
i Þw
¼
Z
Oe
rw
NA
ðvw
j  vs
j Þ dO 
Z
Oe
NA
;i nw
pw dO, (10)
where ns
¼ 1  nw
.
2.3. Time integration
Time integration of the semi-discrete finite element
equations is accomplished by a finite difference time-
stepping algorithm. In general, implicit and explicit
integration procedures are available. Explicit procedures
are computationally the simplest since they do not require
(for a diagonal mass matrix) equation solving to advance
the solution. However, stability restricts the size of the
allowable time step. On the other hand, unconditional
stability can usually be achieved in implicit procedures,
which require solution of a system of equations at each
time step. For the problem at hand, a purely explicit
procedure is not usually appropriate because of the
unreasonably stringent time-step restriction resulting from
the presence of the very stiff fluid in the mixture (even for
highly nonlinear solid material models). The methods used
combine the attractive features of explicit and implicit
integration (see, e.g., Refs. [24,25]), and fall under the
category of ‘‘split operator methods’’. Different portions of
the system of equations are treated implicitly and explicitly,
reducing the system of equations to be solved. The specific
split to be made is obviously problem dependent, and the
appropriate implicit/explicit splits for the problem at hand
are detailed in Refs. [20,21].
2.4. Nonlinear iterations
At each step the resulting nonlinear algebraic problem is
solved by Newton–Raphson type iterations. Quasi-Newton
procedures with line searches are typically used.
3. Modeling seismic-induced soil liquefaction
Liquefaction (as defined by Castro and Poulos [26]) is a
phenomenon wherein a saturated sand subjected to
monotonic or cyclic shear loads looses a large percentage
of its shear resistance and flows in a manner resembling a
liquid. Marcuson [27] defines liquefaction as the transfor-
mation of a granular material from a solid to a liquefied
state as a consequence of increased pore-water pressure
and reduced effective stress. This phenomenon occurs most
readily in loose to medium dense granular soils that have a
tendency to compact when sheared. In saturated soils,
pore-water pressure drainage may be prevented due to the
presence of silty or clayey seam inclusions, or may not have
time to occur due to rapid loading—such as in the case of
seismic loads. In this situation, the tendency to compact is
translated into an increase in pore-water pressure. This
leads to a reduction in effective stress, and a corresponding
decrease of the frictional shear strength. If the EPWP
generated at a certain location in a purely frictional soil
(e.g., sand) reaches the initial value of the effective vertical
stress, then, theoretically, all shear strength is lost at that
location and the soil liquefies and behaves like a viscous
fluid. Liquefaction-induced large ground deformations are
a leading cause of disasters during earthquakes. EPWPs are
induced also in moderately dense to dense granular
materials subjected to cyclic loads, but due to their
tendency to dilate during shear, the softening is only
temporary leading to increased cyclic shear strains, but not
to major strength loss and large ground deformations [28].
This phenomenon is known as cyclic mobility.
Two important aspects need to be addressed in a
rigorous seismic analysis of saturated soils involving
EPWP build-up: (1) coupled analysis: solid and fluid
coupled field equations have to be used in a step-by-step
(time domain) dynamic analysis to correctly capture the
inertial and dissipative coupling terms; and (2) soil
constitutive model: accurate simulation of dynamically
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
650
induced EPWP build-up and continuous softening of the
material requires soil models able to reproduce the
experimentally observed nonlinear hysteretic behavior and
shear stress-induced anisotropic effects, and to reflect the
strong dependency of plastic dilatancy on effective stress
ratio. Advanced plasticity models, such as multi-yield or
bounding surface plasticity, in combination with kinematic
hardening rules, can offer a material representation of
considerable power and flexibility (e.g., [29]). As such soil
constitutive models usually require a relatively large number
of parameters, a well-defined methodology for calibrating
those parameters based on results of standard in situ and/or
laboratory soil tests is also desirable.
The first aspect is addressed in the DYNAFLOW code
by the extension of Biot’s theory into the nonlinear regime,
discussed in Section 2. Nonlinear dynamic constitutive
behavior of saturated soil under partially drained condi-
tions is modeled using a kinematic hardening, multi-yield
constitutive model based on a simple plasticity theory [30].
The yield function is described in the principal stress space
by a set of nested rounded Mohr–Coulomb yield surfaces.
A non-associative plastic flow rule is used for the
dilatational component of the plastic deformation. The
model has been tailored (1) to retain the extreme versatility
and accuracy of the simple multi-surface J2 theory in
describing observed shear nonlinear hysteretic behavior
and shear stress-induced anisotropic effects, and (2) to
reflect the strong dependency of the shear-induced dila-
tancy on the effective stress ratio. Accurate simulation of
shear-induced plastic dilation and of hysteretic effects
under cyclic loading, together with full coupling between
solid and fluid equations, allows capturing the build-up
and dissipation of pore-water pressures and modeling the
gradual softening and hardening of soil materials. The
required constitutive model parameters can be derived
from the results of conventional laboratory (e.g., triaxial,
simple shear [31]) or in situ (e.g., standard penetration,
cone penetration, wave velocity [32]) soil tests. The multi-
yield plasticity soil constitutive model, its implementation
algorithm, and the methodology for estimating the
constitutive model parameters have been repeatedly
validated in the past for soil liquefaction computations,
based on both centrifuge experimental results (e.g.,
[31,33,34]) and full-scale measurements (e.g., [35]).
Practical applications of the extension of Biot’s theory
into the nonlinear regime, including its implementation in
DYNAFLOW, are presented in the remaining of this paper
through results of two studies involving simple structures
on homogeneous or non-homogeneous saturated soil
deposits subjected to seismic loads.
4. Example 1: tower structures on liquefiable soil
Any soil–structure system has its own characteristic
frequency, which depends on material properties, geome-
try, degree of saturation of the soil, and contact conditions
at the soil–foundation interface. This characteristic fre-
quency may decrease during dynamic excitation due to
degradation of the soil effective shear modulus as a result
of pore pressure build-up and/or large shear strains. Any
mechanical system is more sensitive to dynamic loading as
its characteristic frequency becomes closer to the frequency
range corresponding to the maximum spectral values of the
excitation. Consequently, both the frequency content of
seismic excitation and the evolution of structural frequency
characteristics can make a significant difference in the
dynamic response of soil–structure systems.
Several centrifugal studies of seismic behavior of tower
structures on soil foundation are reported, among which:
Morris [36] and Weissman and Prevost [37]—tower struc-
tures on dry soil, Madabhushi and Schofield [18]—tower
structures on saturated soil. Several tests with different input
seismic acceleration have been performed to study the effect
of characteristic frequency. Madabhushi and Schofield [18]
concluded that, when the predominant frequency of the
input seismic motion was lower than the characteristic
frequency of the system, the characteristic frequency of the
system decreased during shaking to a value close to the
predominant input seismic frequency due to build up of
excess pore pressure and subsequent degradation of soil
stiffness. Popescu [13] performed nonlinear dynamic finite
element analyses of an embankment dam on liquefiable
foundation and concluded that the interplay between the
frequency content of the seismic motion, the vibration
characteristics of the structure and the possible evolution of
those characteristics during shaking had significant influence
on the predicted dynamic response.
In this example, the behavior of two different tower
structures (with different characteristic frequencies but
with the same mass) resting on a homogeneous liquefiable
soil deposit is investigated for a range of intensities of the
seismic input motion. The saturated soil is modeled as a
two-phase porous medium, and the finite element calcula-
tions are conducted in terms of effective stress, using fully
coupled solid–fluid equations. The soil and structures
analyzed in this study correspond to those used by
Madabhushi and Schofield [18] in a series of centrifuge
experiments. The earthquake motion used here spans a
wide range of seismic intensities and is different from that
used in the centrifuge experiments.
The main purpose of this study is to investigate in detail
the changes in soil strength induced by EPWP build-up,
leading to changes in the vibration characteristics of a
soil–structure system during and after the earthquake. It is
shown how those changes and their interplay with the
frequency content of the seismic excitation affect: (1) the
dynamic excitation of the structure and (2) the amount of
seismic energy transmitted to the system and, eventually,
the structural response.
4.1. Finite element model and analysis method
Two different structures are analyzed. Tower 1 with
a characteristic frequency of 4.8 Hz, corresponding to a
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 651
two-story building, and Tower 2 with a characteristic
frequency of 1.4 Hz, corresponding to a seven-story
building. The dynamic characteristics of both structures
are assumed to be dominated by their first mode of
vibration, and the structures are idealized as single degree
of freedom systems. The plane strain assumption is used in
the finite element analyses, and the foundation is idealized
as a 10 m wide continuous strip footing. The factor of
safety for bearing capacity under static conditions and
assuming local bearing capacity failure is about 20 for both
structures. In the numerical analyses, the structure material
is idealized as linear elastic. The soil, assumed saturated
below the foundation (ground water table at 1 m depth), is
idealized as a two-phase porous material and discretized
into four-node quadrilateral continuum two-phase porous
elements with four degrees of freedom per node (two for
the solid phase and two for the fluid phase displacements).
The soil constitutive parameters are listed in Table 1.
Use of a rigid box in an experimental earthquake
simulation induces a series of spurious seismic waves that
are generated by the motion of each end wall of the box
and are subsequently reflected back by the opposite wall.
This phenomenon, which may induce a behavior in the
model different from the real field, is partly attenuated by
placing a certain amount of duxseal (a relatively soft
material) at each end wall of the box to create absorbing
boundaries. It is assumed that those phenomena (wave
reflection due to simulation of rigid end walls and
subsequent attenuation when modeling the presence of
duxseal) are reproduced by the numerical model. Duxseal
has been used in the centrifuge experiments to prevent
seismic wave reflection from the lateral boundaries of the
rigid centrifuge box and thus helping the soil deposit to
behave like a semi-infinite medium [18]. This material is
considered also in the numerical analyses, and idealized as
linear elastic. The material properties assumed for duxseal
and structure are listed in Table 2. The finite element mesh
is shown in Fig. 1.
The seismic acceleration is applied at the base and lateral
boundaries of the analysis domain, same as in the
centrifuge experiment performed in a rigid box. The base
input accelerations are generated using a procedure for
simulation of non-stationary stochastic processes [38]
capable of generating seismic ground motion time histories
that are compatible with prescribed response spectra and
have a prescribed modulating envelope for amplitude
variation. Two hundred acceleration time histories have
been generated to be compatible with the response spectra
type-1 and type-3 recommended by the Uniform Building
Code [39] and corresponding to different soil conditions.
All 100 generated acceleration time histories in each group
are compatible to the corresponding response spectrum,
ARTICLE IN PRESS
Table 1
Parameters of the multi-yield plasticity model, and values used for the saturated soil in Example 1
Constitutive parameter Symbol Value Type
Mass density—solid rs
2710 kg/m3
State parameters
Porosity nw
0.43
Hydraulic conductivity k 0.0002 m/s
Low strain elastic shear modulus G0 20 MPa Low strain elastic parameters
Poisson’s ratio n 0.33
Power exponent n 0.5
Friction angle at failure f 381 Yield and failure parameters
Maximum deviatoric strain (comp/ext) edev
max
0.08/0.06
Coefficient of lateral stress k0 0.426
Stress–strain curve coefficient a 0.21
Dilation angle c 321 Dilation parameters
Dilation parameter Xpp 0.2
Table 2
Constitutive parameters for linear elastic materials used in Examples 1 and 2
Constitutive parameter Duxseal
(Examples 1 and 2)
Tower structures
(Example 1)
Structurea
(Example 2)
ESB boxb
Dural Rubber
Mass density (kg/m3
) 1650 3530 3200 3953 3953
Young’s modulus (MPa) 8 73,000 73,000 73,000 158
Poisson’s ratio 0.46 0.33 0.33 0.33 0.45
a
Mass calculated to obtain a pressure on soil equivalent to the one in the centrifuge experiments.
b
Mass density for both materials and Young’s modulus of rubber have been fitted to obtain equivalent vibration and deformability characteristics with
those of the ESB box used in the centrifuge experiments.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
652
and consequently have similar frequency content. They
also have similar duration of the strong motion. Aside
from the above common characteristics, each individual
acceleration time history is different from the others in the
same group (e.g., the peak values occur at different time
instants in each sample), as each time history constitutes a
different sample of the target non-stationary stochastic
process. The accelerations in each group have been scaled
to a range of Arias intensities [40], IA ¼ 0:1922:1 m=s,
corresponding to a range of peak ground accelerations,
PGA ¼ 0:120:4 g. The duration of seismic motion con-
sidered here is about 15 s and the total analysis time is 40 s.
4.2. Analysis results
Contours of EPWP ratio with respect to the initial
effective vertical stress calculated for one case (Tower 1
subjected to type-3 acceleration sample #51, with IA ¼
1:12 m=s and PGA ¼ 0:25 g) are shown in Color Plate 1 (at
the end of the paper). The soil in the free field is predicted
to liquefy, but, due to increased overburden pressure, the
soil below the structure does not liquefy. There is, however,
significant pore-water pressure build-up and subsequent
softening of the soil in this area, leading to large
settlements and rotations of the structure. The numerical
model also reproduces gradual dissipation of pore-water
pressure after the end of the earthquake (Color Plate 1c).
Color Plate 2 illustrates the predicted failure mechanism
for the case presented in Color Plate 1. At time T ¼ 5 s, the
average EPWP ratio in the area below the structure is
about 75% (Color Plate 2b), and the corresponding
reduction in effective stress leads to a factor of safety for
local bearing capacity failure equal to 1. The failure
mechanism illustrated by the maximum shear strain
contours in Color Plate 2a is characteristic for a local
bearing capacity failure, consisting of an active wedge
below the structure, clearly shown by the shear strain
contours, and fading transition and passive wedges. The
failure wedges are not symmetric, due to the presence of
lateral inertial forces.
Due to softening of soil below the structure, the seismic
motion is strongly attenuated, especially for large Arias
intensities (IA), and the acceleration at the base of the
structure has significantly less specific energy compared to
the input. In Fig. 2, it is shown how at low seismic intensity
(sample #1 with IA ¼ 0:19 m=s) the seismic energy is
entirely transmitted to the structure (Figs. 2a and b), while
for high seismic intensities (sample #90 with IA ¼ 1:84 m=s,
where the soil liquefies), the input seismic energy is largely
attenuated by the soil (Fig. 2c), especially for the high
frequency components of the base motion (Fig. 2d). For
the second case, it can be noted from Fig. 2c how the
acceleration amplitudes at the base of the structure are
similar to those of the input motion for the first 2–3 s of
shaking, when the pore-water pressures in the soil are still
small. After that, as the soil starts to soften due to build-up
ARTICLE IN PRESS
Fig. 1. Seismic analysis of a tower structure: finite element mesh.
-1
0
1
m/s2
0 5 10 15 20
-3
-2
-1
0
1
2
3 input
base of struct.
m/s2
Time (sec)
(c)
(a)
(b)
(d)
10-1
100
101
10-2
100
Sample # 1
Sample # 1
10-1
100
101
10-2
100
Period (sec)
Response
Spectrum(m/s
2
)
input
base of struct.
Sample # 90
Sample # 90
Fig. 2. Comparison between acceleration time histories (a, c) and response spectra (b, d) used as input at the base of the mesh and computed at the base of
the structure for two samples of type-1 acceleration used for Tower 1 (corresponding to two different Arias intensity values).
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 653
of EPWP, the maximum accelerations transmitted to the
structure base become significantly lower than the ones
input at the base of the analysis domain. It can be noted
from the computed acceleration response spectrum shown
in Fig. 2d that the high frequency components of the
seismic acceleration are predominantly attenuated, while
the energy of the low frequency components that are closer
to the natural frequency of the liquefied soil deposit is not
very much affected by the seismic wave propagation
through soil. This is also evident from the low frequency
sinusoidal shape exhibited by the computed acceleration
time history at the base of the structure (red line in the
electronic version and darker line in the paper version in
Fig. 2c). The shape of this computed acceleration time
history is similar to accelerations recorded at shallow depth
in a landfill that liquefied during the Hyogoken-Nambu
Earthquake (e.g., [41]).
It is mentioned that the computed reduction in accel-
eration amplitude at the base of the structure for large
Arias intensities of the input seismic motion should not be
misinterpreted as a safer situation for the case in Fig. 2c
(sample #90 with IA ¼ 1:84 m=s) than for the case in
Fig. 2a (sample #1 with IA ¼ 0:19 m=s). Even if structure
accelerations are in general smaller for sample #90 than for
sample #1, the structure shaken by sample #90 is predicted
to fail due to bearing capacity failure of the liquefied soil,
while the structure shaken by sample #1 is predicted to
be safe.
The effects of earthquake intensity and loading rate on
the settlement and rotation of the rigid foundation for the
two types of structures are presented in Fig. 3 in terms of
fragility curves, which are an illustrative and practical way
of expressing the probability of exceeding a certain degree
of structural response as a function of load intensity.
Fragility curves have been used extensively in earthquake
engineering to describe the seismic vulnerability of
structures as a function of the severity of the seismic event
(e.g., [42,43]). The procedure used here for establishing the
fragility curves for foundation settlements and rotation
follows that proposed by Shinozuka et al. [42]. After
selecting a specific response threshold, all earthquake
intensities IA for which the response exceeds this threshold
are assigned to the unity probability level, while all the
other intensities are assigned to the zero probability level.
Then, a shifted lognormal distribution function is fitted to
those points using the maximum likelihood method.
Fragility curves can include effects of multiple sources of
uncertainty related to material resistance or load char-
acteristics. The curves used in this example reflect only
uncertainty in the seismic ground motion.
The curves presented here express the probability of
exceeding certain thresholds in the response, namely
foundation settlement of 20 cm in Fig. 3a and base
rotations of 1/300 in Fig. 3b. Those results represent
earthquake effects in terms of rigid motion of the structure.
From both figures it can be concluded that, for the same
level of seismic intensity, type-3 acceleration has stronger
effect on both settlements and rotations than type-1
acceleration. This is due to a change in vibration
characteristics of the soil to fundamental frequencies close
to the dominant frequency values of type-3 acceleration.
The same behavior has also been observed by Madabhushi
and Schofield [18] in the centrifuge experiments. The initial
characteristic periods of the soil–structure system (before
the earthquake) are T1 ¼ 0:6 s, for Tower 1, and
T2 ¼ 0:8 s, for Tower 2. During the earthquake, due to
significant increase in pore-water pressure and subsequent
soil softening, this characteristic period goes up. For most
cases analyzed here, when the soil liquefies in the free field
and the EPWP increases significantly below structure, the
characteristic period of the soil–structure system for both
Towers 1 and 2 goes up to about 1.5 s. Consequently, the
seismic energy input at the base of the soil layer is
delivered, for most of the earthquake duration, to a system
that has a characteristic period of about 1.5 s. As shown in
Fig. 4, this value is closer to the range of maximum spectral
amplitudes of type-3 acceleration than to those of type-1,
and this results in significantly more energy delivered at the
resonance frequency to the soil–structure system when
acted by the type-3 input than when acted by type-1 input.
ARTICLE IN PRESS
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
Arias Intensity (m/s)
Prob.
that
rotation
exceeds
1/300
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
Arias Intensity (m/s)
Prob.
that
settlement
exceeds
20cm
Tower1 - Type 1
Tower2 - Type 1
Tower1 - Type 3
Tower2 - Type 3
Tower1 - Type 1
Tower2 - Type 1
Tower1 - Type 3
Tower2 - Type 3
(a) (b)
Fig. 3. Comparison between analysis results in terms of fragility curves: (a) settlements and (b) rotations (type-1 and type-3 refer to the corresponding
UBC response spectra, and Towers 1 and 2 refer to the two structures considered).
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
654
For example, for the case illustrated in Fig. 4, more than
two times more energy is delivered at the resonance
frequency by the type-3 input acceleration than by the
type-1 input. It is mentioned that when computing the
characteristic period before the earthquake and at various
time instants during the dynamic analysis, the coupling
effects in Eqs. (1) and (2) have been neglected.
It can be noticed from Fig. 3 that, for type-1 input
motion, the fragility curves calculated for Tower 1 are
located to the left of the ones calculated for Tower 2. This
means that a lower value of seismic intensity corresponds
to a given probability of seismic damage for Tower 1 than
for Tower 2. Apparently Tower 2, more flexible, can
attenuate some part of the seismic energy by structural
deflections. It is also interesting that, when subjected
to seismic motion type-3, where the degree of soil
liquefaction is higher than for type-1 input, the computed
responses of both structures are similar, as both structures
behave as rigid blocks compared to the deformability of
liquefied soil.
5. Example 2: effects of stochastic soil variability on
liquefaction
5.1. Results from previous studies
Natural soil properties vary from one point to another,
even within so-called ‘‘uniform’’ soil layers. In addition to
inducing uncertainty in the computed response, natural
spatial variability of soil properties within geologically
distinct layers affects the mechanical behavior of geotech-
nical structures. For example, for phenomena involving the
presence of a failure surface (such as encountered in
landslides or in bearing capacity failures) the actual failure
surface can deviate from its theoretical position to pass
selectively through weaker soil zones and thus the average
mobilized strength is reduced when compared to that of a
corresponding uniform soil (e.g., [44,45]). For the case of
seismically induced soil liquefaction, it was shown (e.g.,
[14]) that a larger amount of EPWP is generated in a
heterogeneous soil than in the corresponding uniform soil
having geomechanical properties equal to the average
properties of the variable soil.
Experimental evidence of the effects of soil heterogeneity
on liquefaction resistance is presented in Refs. [46,47].
Budiman et al. [46] performed a series of undrained cyclic
triaxial tests on sand specimens containing up to 25%
gravel inclusions. They found that the liquefaction
resistance of sand with inclusions was lower than that of
uniform sand. Moreover, the reduction in liquefaction
resistance was more pronounced for samples with a higher
content of gravel. Konrad and Dubeau [47] conducted
undrained cyclic triaxial tests on fine Ottawa sand and on
silica silt. The study consisted of three series of tests: the
first two series were conducted to characterize the cyclic
strength of the sand and of the silt, respectively. The third
series was performed on layered samples, consisting of one
horizontal silt layer sandwiched between two sand layers.
The study results revealed that layering induced a much
lower cyclic resistance than that developed in either of the
materials in uniform samples.
Pioneering work in numerical analysis of the effects of
soil heterogeneity on liquefaction resistance was presented
by Ohtomo and Shinozuka [48]. Later on, Popescu [32] and
Popescu et al. [14,15] conducted a systematic study of those
effects using the multi-yield surface constitutive model
implemented in DYNAFLOW. A Monte-Carlo simulation
(MCS) method, combining digital generation of non-
Gaussian stochastic vector fields representing the spatial
distribution of various soil properties over the analysis
domain with dynamic effective stress finite element
analyses, was used to quantify the effects of soil hetero-
geneity on the amount and pattern of EPWP build-up in a
saturated soil deposit subjected to seismic loads. It was
shown that, for the same average values of soil parameters,
more EPWP build-up was predicted in stochastic analyses,
accounting for soil heterogeneity, than in so-called
‘‘deterministic analyses’’, assuming uniform soil properties.
Characteristic percentiles of soil strength were proposed for
use in deterministic analyses, resulting in a response similar
to that predicted by more expensive stochastic analyses.
Effects of the degree of soil variability and of the seismic
loading rate were also investigated. Fenton and Vanmarcke
[49] addressed the three-dimensional (3D) aspect of
liquefaction of randomly variable soils through an analysis
of liquefaction potential at the Wildlife Site, Imperial
Valley, California. The soil properties were modeled as a
3D random field; however, because of limited computa-
tional resources, the liquefaction analysis was carried out
in 1D vertical columns, without any coupling in the
horizontal plane. Consequently, the study could only
address the pointwise initiation of liquefaction, and did
ARTICLE IN PRESS
10-1
100
101
10-1
100
101
Period - T (sec)
Spectral
acc.
-Sa
(m/s
2
)
UBC-Type 1
UBC-Type 3
2.1m/s2
1.0m/s2
T1
=1.5s
Fig. 4. Comparison of the spectral acceleration values of the two types of
spectra used for the input motion at the characteristic period of the
soil–structure system during shaking. This characteristic period increases
to the value of about T1 ¼ 1:5 s from the initial values of 0.6 and 0.8 s.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 655
not include the effects of 3D pore-water pressure redis-
tribution before and after initial liquefaction. Another
analysis of the same site was presented by Elkateb et al.
[50]. A 3D stochastic analysis was performed, but the
liquefaction potential was assessed separately at each point
in space using an empirical method (e.g., [51,52]). Although
this approach could not capture any interaction between
zones of soil with different liquefaction strengths and
different amounts of pore-water pressure build-up, it
provided some very interesting conclusions on the effect
of loose soil zones on the probability of liquefaction
failure. It was found that liquefaction assessments using
mean values of soil properties (e.g., from in situ CPT
results) are not able to capture the effect of loose soil
pockets within the soil mass and consequently are on the
unsafe side. Popescu et al. [17] performed MCSs including
fully coupled 3D dynamic finite element analyses to
investigate the effects of true 3D soil variability on
liquefaction of soil deposits. They concluded that 2D plane
strain stochastic analyses can provide sufficiently accurate
information for liquefaction strength assessment of hor-
izontally layered soil deposits, while liquefaction-induced
differential settlements can only be captured accurately by
a 3D analysis, with correct simulation of pore-water
migration. Koutsourelakis et al. [53] and Popescu et al.
[16] analyzed the effects of spatial variability on soil
liquefaction for a wide range of earthquake intensities
using again MCSs, and presented fragility curves for simple
structures on liquefiable soil.
To illustrate the effects of soil heterogeneity on
seismically induced EPWP build-up, some of the results
obtained in Ref. [14] for a loose to medium dense soil
deposit, with geomechanical properties and spatial
variability characteristics estimated based on a series of
piezocone test results from a hydraulic fill deposit
are reproduced in Fig. 5. The results in Fig. 5b show
ARTICLE IN PRESS
20 one - phase finite
480 two - phase finite elements
for porous media (4 dof/node)
with dimensions: 3m x 0.5m
Finite elements
analysis set-up
free field boundary
conditions
elements (2 dof/node)
12.0m
1.6m
Water table
Vertical scale is two
times larger than
horizontal scale
60.0m
(a)
(b)
(c)
Excess pore pressure ratio predited using six samples functions of a stochastic vector field with
cross-correlation structure and probability distribution functions estimated from piezocone test results
Grey scale range
for the excess pore
pressure ratio (u/σv0)
0.90...1.00
0.70...0.90
0.50...0.70
0.00...0.50
Sample function # 1 Sample function # 2 Sample function # 3
Sample function # 6
Sample function # 5
Sample function # 4
Excess pore pressure ratio predicted using deterministic input soil parameters
Deterministic - average 50 - percentile 60 - percentile
90 - percentile
80 - percentile
70 - percentile
(u/σv0
)max = 0.44
´
´
Fig. 5. Monte-Carlo simulations of seismically induced liquefaction in a saturated soil deposit, accounting for natural variability of the soil properties
(after [14]): (a) finite element analysis setup; (b) contours of EPWP ratio for six sample functions used in the Monte-Carlo simulations; (c) contours of
EPWP ratio using deterministic (i.e., uniform) soil parameters and various percentiles of soil strength.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
656
computed contours of EPWP ratio with respect to the
initial effective vertical stress for six sample functions
of a stochastic field representing six possible realizations of
soil properties over the analysis domain (see Ref. [14] for
more details on the soil properties). Soil liquefaction
(EPWP ratio larger than approximately 0.9) was predicted
for most sample functions shown in Fig. 5b. Analysis of an
assumed uniform soil deposit, with strength characteristics
corresponding to the average strength of the soil samples
used in MCSs, resulted in no soil liquefaction (the
maximum predicted EPWP ratio was 0.44). It can be
concluded from these results that both the pattern and
the amount of dynamically induced EPWP build-up
are strongly affected by the spatial variability of soil
properties. For the same average values of the soil
parameters, more EPWP build-up was predicted in the
stochastic analysis (MCS), accounting for spatial varia-
bility, than in the deterministic analysis, assuming uniform
soil properties.
Popescu et al. [14] suggested that in order to predict
more accurate values of EPWP build-up in a deterministic
analysis assuming uniform soil properties, one has to use a
modified soil strength which is lower than the average
strength of the variable soil. The results presented in
Fig. 5c illustrate the idea of a ‘‘characteristic percentile’’
of soil strength for dynamic liquefaction analysis of
randomly heterogeneous soils. This is a percentile of
the field-recorded soil strength values which is lower
than the average value. The characteristic percentile is
selected in such a way that, when used in a deterministic
dynamic analysis assuming uniform soil strength, it
results in a computed EPWP ratio equivalent to that
obtained from the stochastic analysis accounting for
soil spatial variability. For example, for the soil properties
and earthquake intensity considered in Ref. [14],
the resulting characteristic percentile was somewhere
between 70% and 80% (Fig. 5c). A comparison between
the results presented in Figs. 5b and c also shows clearly
the differences in the predicted liquefaction pattern of
natural heterogeneous soil vs. that of assumed uniform
soil.
The effects of soil spatial variability on structures
founded on liquefiable soil deposits are illustrated in
Fig. 6 (from Ref. [16]). The structure is the same as Tower
2 presented in the previous example, founded on
either heterogeneous soil, or on corresponding uniform
soil, with the same strength as the average strength of
the variable soil. It is mentioned that the soil properties
for uniform soil, as well as the lateral boundary conditions,
are different in the study presented in Fig. 6 from those
used in Example 1. Fragility curves expressing the
exceedance probability of two thresholds in the response,
namely foundation settlements larger than 20 cm and
base rotations larger than 1/150, are plotted in Figs. 6a
and b, respectively. As both the spatial distribution of
soil properties and the input seismic motion vary
randomly from one sample to another, the curves shown
in Fig. 6 reflect uncertainties in both spatial distribution
of soil strength and in the seismic ground motion.
The fragility curves are plotted for two types of seismic
input (Types 1 and 3 discussed in Example 1), and for
two assumptions related to soil properties: (1) variable
soil, with properties described by a random field with
correlation distances yH ¼ 8 m in horizontal direction
and yV ¼ 2 m in vertical direction, and a coefficient
of variation (CV) ¼ 0.5 (see Ref. [16] for more details)
and (2) corresponding uniform soil. For all cases
analyzed, the fragility curves calculated for variable
soil are shifted to the left as compared to the ones
for uniform soil, indicating that, for a given exceedance
probability, a significantly lower seismic intensity is
needed when soil variability is accounted for than for an
assumed uniform soil. For example, from Fig. 6a, for
type-1 input acceleration, a 50% exceedance probability
of 20 cm foundation settlement is predicted at Arias
intensity of the seismic motion IA ¼ 1:7 m=s for uniform
soil, and at IA ¼ 1:0 m=s when the soil variability is
accounted for.
ARTICLE IN PRESS
Fig. 6. Comparison between analysis results for variable soil and
equivalent uniform soil, for Tower 2 in Example 1 (after [16]): (a)
settlements and (b) rotations. The foundation conditions considered in this
analysis are different from those discussed in Example 1.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 657
5.2. Mechanisms of liquefaction of heterogeneous soils
It was postulated by Popescu et al. [14] that presence of
loose pockets would lead to earlier initiation of EPWP and
local liquefaction than in a corresponding uniform soil.
After that, the pressure gradient between loose and dense
zones would lead to water migration, softening, and
liquefaction of the dense sand. This assumption that
presence of loose pockets is responsible for lower liquefac-
tion resistance was theoretically verified by two types of
numerical results. First, for a given average value of the soil
strength and a given earthquake intensity, the resulting
amount of EPWP build-up in a variable soil was increasing
with the assumed degree of variability of soil properties,
expressed by the CV, which also controls the amount
of loose pockets in the soil mass (e.g., [15]). Second, for
the same CV of spatial variation, more EPWP was
predicted when the soil strength fluctuations followed
a symmetrical probability distribution function (PDF)—
e.g., truncated Gaussian—than for a positively skewed
PDF—e.g., lognormal [54]. For the same mean value,
a symmetrical PDF has a fatter left tail (indicating a
larger amount of loose pockets) than a positively
skewed PDF.
Several practical guidelines for assessment of liquefac-
tion potential and for design of structures on liquefiable
soil with spatially varying properties resulted from
the numerical studies discussed before. DYNAFLOW
and the multi-yield plasticity constitutive model have
been repeatedly validated for liquefaction assessment
based on centrifuge experimental results and field observa-
tions. All those validation studies were based on determi-
nistic analyses, assuming uniform soil properties in
horizontal direction. To validate the geotechnical design
recommendations for liquefaction of heterogeneous soil
mentioned above, the capability of the numerical model
to capture the real behavior of spatially variable soil
also needs to be verified based on experimental data. Such
a verification is performed as part of this numerical
example.
There are very few experimental studies dealing with
liquefaction of spatially variable soil. The two studies
mentioned in Section 5.1 were based on undrained cyclic
triaxial tests for samples made of different soils (Budiman
et al. [46] used sand and gravel, and Konrad and Dubeau
[47] used sand and silt). A detailed analysis of the behavior
of non-homogeneous soils was performed by Chakrabortty
et al. [55] using DYNAFLOW. They reproduced the cyclic
undrained triaxial tests on sandwiched samples made of
dense sand and silt layers described in Ref. [47], and
analyzed the mechanism by which a sample made of two
different soils liquefies faster than each of the soils tested
separately in uniform samples. The explanation, resulting
from a detailed analysis of the numerical results, was that
water was squeezed from the more deformable silt layer
and injected into the neighboring sand, leading to
liquefaction of the dense sand.
Regarding liquefaction mechanisms of soil deposits
involving the same material, but with spatially variable
strength, Ghosh and Madabhushi [19] performed a series
of centrifuge experiments to analyze the effects of localized
loose patches in a dense sand deposit subjected to seismic
loads. They observed that EPWP is first generated in the
loose sand patches, and then the water migrates into the
neighboring dense sand, reduces the effective stress and
softens the dense soil that can liquefy. As discussed before,
it is believed that a similar phenomenon is responsible for
lower liquefaction resistance of heterogeneous soil than
that of the corresponding uniform soil (e.g., [14]). There-
fore, this phenomenon is analyzed in details in the
following numerical example, based on the backanalysis
of Ghosh and Madabhushi’s centrifuge experiments,
and on a study for a hypothetical chess-board like
heterogeneous soil. This is an original study aimed
at: (1) improving understanding of the mechanisms
of liquefaction of spatially variable sandy soil deposits
and (2) verifying if the nonlinear Biot model implemented
in DYNAFLOW is able to capture experimentally
observed phenomena including water migration from loose
to dense sand zones, with subsequent softening and
liquefaction of the entire soil mass.
A detailed description of the centrifuge tests analyzed in
this example is presented in Ref. [19]. Most of the
experimental results are posted on the WWW [56]. The
general layout of the tests discussed here—structure on
dense sand deposit with included horizontal loose sand
layer (test BG4) and included loose sand patch (test
BG5)—is sketched in Fig. 7a, along with transducer
locations and the finite element mesh used in the back-
analyses. The location of the loose sand layer in test BG4 is
indicated in Fig. 7a by two thick horizontal lines, and the
loose sand patch in test BG5 is shown by the orange
rectangular area below the structure in the electronic
version (darker area in the paper version). The centrifuge
container used in the experiments was an equivalent shear
beam (ESB) box, ideally matching the stiffness of a soil
column. The vibration and deformability characteristics of
the Cambridge centrifuge ESB box are presented in details
in Ref. [57]. Those characteristics have been reproduced for
plane strain analysis in the finite element model used here.
A duxseal padding placed between soil and the lateral walls
of the box in the centrifuge model was also included in the
numerical model, as shown in Fig. 7a. Duxseal properties,
as well as the properties of the ESB box materials (dural
and rubber), all modeled as linear-elastic materials, are
listed in Table 2. The base input accelerations used in the
two tests [56] are presented in Figs. 7b and c.
The soil properties at two relative densities used in the
centrifuge experiments (Dr ¼ 85% for the dense sand and
Dr ¼ 45% for the loose sand) have been estimated based
on information provided in Ref. [19] and on the back-
analysis of test BG4. Those soil properties have been
subsequently checked in a backanalysis of test BG5. A
comparison of numerical results obtained in this study with
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
658
ARTICLE IN PRESS
Fig. 8. Recorded (after [19]) and computed excess pore-water pressure time histories at three locations: below the structure (P2 and P5) and in the free field
(P6) (transducer locations are shown in Fig. 7). The calculations are done using the acceleration time histories shown in Figs. 7b and c for tests BG4 and
BG5, respectively. The figure is plotted at model scale.
Fig. 7. Finite element simulation of tests BG4 and BG5: (a) finite element mesh, (b) input acceleration for test BG4, and (c) input acceleration for
test BG5.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 659
centrifuge experimental records in terms of EPWP at three
locations presented in Fig. 8 shows acceptable match. The
soil constitutive parameters are listed in Table 3.
To analyze in detail the mechanism of water migra-
tion between loose and dense sand observed in the
experiments, and to have a better basis for comparison,
both centrifuge tests have been numerically simulated
with DYNAFLOW using the same base input acceleration,
presented in Fig. 7c. Some results in terms of EPWP
ratio with respect to the initial effective vertical stress
are presented in Fig. 9. Evolution of EPWP ratio at
neighboring locations in test BG4 below the structure
and in the free field are presented in Figs. 9a and b,
respectively. In those figures, P4 and P7 are located in
the loose sand layer, while P5 and P6 are located in dense
sand. While some pore-water pressure dissipation is
predicted to occur after the earthquake in the loose
sand, increase in pore-water pressure after the end of
ARTICLE IN PRESS
Table 3
Parameters of the multi-yield plasticity model used for the saturated soil in Example 2
Constitutive parameter Symbol Dr ¼ 45% Dr ¼ 65% Dr ¼ 85%
Mass density—solid (kg/m3
) rs
2650 2650 2650
Porosity nw
0.45 0.425 0.40
Hydraulic conductivity (m/s) k 1.35  104
1.1  104
8  105
Low strain elastic shear modulus (MPa) G0 12 16 20
Poisson’s ratio n 0.25 0.25 0.25
Power exponent n 0.5 0.5 0.5
Friction angle at failure (deg.) f 39 42.5 45
Maximum deviatoric strain (comp/ext) edev
max
0.07/0.04 0.07/0.04 0.07/0.04
Coefficient of lateral stress k0 0.7 0.7 0.7
Stress–strain curve coefficient a 0.182 0.182 0.182
Dilation angle (deg.) c 32 32 32
Dilation parameter Xpp 0.05 0.02 0.006
(c)
(a)
(b) (d)
P7
P6
0 10 20 30 40 50
0
0.2
0.4
0.6
0.8
1
Test BG4
Time (sec)
EPWP
Ratio
0 10 20 30 40 50
0
0.2
0.4
0.6
0.8
1
Test BG5
Time (sec)
EPWP
Ratio
0 10 20 30 40 50
0
0.2
0.4
0.6
0.8
1
Test BG4
Time (sec)
EPWP
Ratio
P4
P5
P4
P5
Fig. 9. Computed excess pore-water pressure (EPWP) ratios at pairs of transducers, located one in dense sand and one in loose sand: (a) test BG4, below
structure; (b) test BG4, free field; (c) test BG5, below structure; (d) transducer locations. The calculations are done using the acceleration time history in
Fig. 7c as input motion for both tests.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
660
shaking is simulated at the locations in dense sand, due
to migration of water from loose sand. The same trend
is observed in test BG5—Fig. 9c shows computed
EPWP ratios at two neighboring locations below
the structure (P4 in the loose sand patch and P5 in dense
sand).
Fig. 10 compares computed EPWP ratios at the same
location (transducer P6 in dense sand in the free field)
for tests BG4 and BG5. An increase in residual EPWP
ratio from about 0.2 in test BG5 to about 0.4 in test
BG4 is attributed to water migration from the loose
sand, present in the free field close to location P6 in
test BG4, but not present in the free field in test BG5.
The effects of water migration with gradual increase in
EPWP in test BG4 as compared to test BG5 can be
observed starting at time T ¼ 20 s. Simulation of water
migration from loose to dense sand by the numerical model
is also illustrated in Fig. 11 by computed time histories
of water displacements (relative to the solid phase) at
two locations at the border between loose and dense sand
in test BG5.
To quantify the effects of alternating loose and dense soil
patches, as encountered in natural soil deposits, a
hypothetical soil–structure system (Fig. 12) is analyzed
next. The soil is made of a chess-board like pattern of
loose (Dr ¼ 45%) and dense (Dr ¼ 85%) sand zones.
It is mentioned that natural soils exhibit a smooth
(gradual) variation in relative density between loose
and dense zones, while the example analyzed here
assumes sudden variation in soil properties. The reason
for selecting this chess-board like layout is that sudden
changes in soil properties may amplify the effects of
spatial variability, leading to easier result analysis and
understanding of the underlying mechanisms. To avoid
any boundary effects, free field boundary conditions
are imposed in the following calculations (corresponding
solid and fluid nodal displacements for all pairs of
nodes located at the same elevations on the lateral
boundaries of the mesh are identical). The input motion
used in this example corresponds to the acceleration
time history in Fig. 7c divided by a factor of 2 to simulate
a small to moderate intensity earthquake. For compa-
rison, three other analyses are performed for a structure
ARTICLE IN PRESS
0 10 20 30 40 50
-0.1
0
0.1
0.2
0.3
0.4
0.5
Time (sec)
EPWP
Ratio
Test BG4-P6
Test BG5-P6
Fig. 10. Comparison of computed excess pore-water pressure (EPWP)
ratios in the free field (dense sand) for tests BG4 and BG5. The
calculations are done using the acceleration time history in Fig. 7c as input
motion for both tests.
0 10 20 30 40 50
-5
0
5
10
15
Time (sec)
Relative
water
movement
[water-solid]
(mm)
Vertical displacements at Node A
Horizontal displacements at Node B
Fig. 11. Computed relative movement of water with respect to the solid
phase at the interface between loose and dense sand in test BG5. Node
locations are shown in Fig. 9d. Positive displacements are upward
(vertical) and left to right (horizontal).
Fig. 12. Finite element mesh and distribution of loose and dense soil patches for a hypothetical soil–structure system analyzed in Example 2.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 661
founded on an assumed uniform soil deposit, for three
different relative densities: dense (Dr ¼ 85%) and loose
(Dr ¼ 45%), corresponding to the patches in the chess-
board example, and medium dense (Dr ¼ 65%), correspon-
ding to the average relative density of the heterogeneous
soil.
Some results in terms of EPWP ratio time histories
computed in the free field are presented in Fig. 13. Fig. 13a
shows the evolution of the EPWP ratio in two adjacent
elements of the chess-board variable soil, one located
in a loose soil patch (element 1—shown in Fig. 12) and
the other in a dense one (element 2). The effects of
water migration are evident here: the dense sand in the
chess-board deposit reaches about 95% EPWP ratio,
while at the same location in a uniform dense sand deposit
the maximum EPWP ratio is about 25% (compare
Figs. 13a and c). Figs. 13b and c show computed EPWP
ratios in the free field for the chess-board soil deposit and
for the other three cases with assumed uniform soil. It is
apparent from those results that the excess pressures
calculated in the heterogeneous soil (at both loose and
dense sand locations) are closer to the ones for uniform
loose sand than to those calculated in the uniform medium
dense sand. Computed structure settlements shown in
Fig. 13d also indicate a significantly softer behavior of the
heterogeneous soil than that of the corresponding uniform
soil (i.e., Dr ¼ 65%). Color Plates 3 and 4 present contours
of EPWP ratio at time T ¼ 20 s and of maximum shear
strains at the end of analysis, respectively. From both color
plates it is obvious that the heterogeneous soil response
(Color Plates 3a and 4a) is much closer to the response of
the uniform loose sand (Color Plates 3c and 4c) than to
that of the uniform medium dense sand (Color Plates 3b
and 4b).
6. Summary and conclusions
The theory of dynamics of saturated porous media, its
extension into the nonlinear regime, and its implementa-
tion in the DYNAFLOW code have been presented, along
with numerical applications illustrating the capabilities of
the numerical model to deal with complex phenomena,
such as seismic behavior of structures on liquefiable soil
deposits. The numerical results are validated through
comparisons with centrifuge experimental data.
The frequency content of seismic motion has a major
role in predicted structural response. In the first numerical
example, it is shown how the characteristic frequency of the
soil–structure system is reduced to values close to the
dominant frequency range of type-3 inputs, due to soil
softening after build-up of pore pressures. Consequently,
considerably more damage is predicted for this type of
seismic motion. The results of this study also point out to
important attenuation of the earthquake motion when
ARTICLE IN PRESS
0 10 20 30 40 50
0.25
0.2
0.15
0.1
0.05
0
Time (sec)
Settlement
(m)
Dense sand
Medium dense sand
Chess-board
Loose sand
0 10 20 30 40 50
-0.2
0
0.2
0.4
0.6
0.8
1
Element 1
Time (sec)
EPWP
Ratio
Loose sand
Chess-board
Medium dense sand
Dense sand
(b)
(a)
(c) (d)
Settlement at base of structure
0 10 20 30 40 50
-0.2
0
0.2
0.4
0.6
0.8
1
Time (sec)
EPWP
Ratio
Element 2
(Dense sand)
Element 1
(Loose sand)
0 10 20 30 40 50
-0.2
0
0.2
0.4
0.6
0.8
1
Element 2
Time (sec)
EPWP
Ratio
Medium dense sand
Chess-board
Loose sand
Dense sand
Fig. 13. (a) EPWP ratio time histories for the chess-board type heterogeneous soil computed at two adjacent locations in the free field (one in the loose
soil—element 1, and the other in the dense soil—element 2); (b) EPWP ratio in the free field in element 1; (c) EPWP ratio in the free field in element 2; (d)
computed settlements of the structure. The four curves in (b)–(d) correspond to results from four different analyses defined in the text. Element locations
are shown in Fig. 12.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
662
seismic waves pass through soft soil after considerable
increase in pore-water pressure. The reduction in specific
seismic energy was found to range between zero, for
small earthquake intensities, to about 75%, for large
earthquake intensities (IA about 2 m/s and PGA about
0.4 g). As expected, the numerical model predicts higher
attenuation for the high frequency components of the base
motion.
In the second numerical example, it was mentioned that
soil heterogeneity leads to more EPWP build-up than
that predicted for corresponding uniform soil and, for
structures founded on liquefiable soils, to reduction in
average mobilized soil strength due to deviation of
the failure surface to pass through looser soil zones.
Those effects result in larger predicted damage when
accounting for soil spatial variability, as compared to
assuming a corresponding uniform soil. The mechanism
of liquefaction of heterogeneous soils was also studied in
this example through backanalysis of centrifuge tests
including loose soil zones in a dense sand deposit. From
the experimental results, it was concluded that the
pore pressure gradient created from the initial liquef-
action of loose zones leads to migration of water in the
adjacent dense zones and makes the dense soil liable to
liquefaction. Several numerical examples were presented
that confirmed the capabilities of the numerical model to
reproduce those experimentally observed phenomena
and validated previous research results and design recom-
mendations for liquefaction of stochastically spatially
variable soils.
ARTICLE IN PRESS
Plate 4. (Example 2) Computed maximum shear strain contours and deformed meshes at the end of analysis (time T ¼ 50 s) for four different soil
structure systems, defined as: (a) chess-board distributed loose and dense patches; (b) uniform medium dense soil (Dr ¼ 65%); (c) uniform loose soil
(Dr ¼ 45%); (d) uniform dense soil (Dr ¼ 85%).
Plate 3. (Example 2) Computed EPWP ratio contours at time T ¼ 20 s for four different soil structure systems, defined as: (a) chess-board distributed
loose and dense patches; (b) uniform medium dense soil (Dr ¼ 65%); (c) uniform loose soil (Dr ¼ 45%); (d) uniform dense soil (Dr ¼ 85%).
Plate 1. (Example 1) EPWP ratio contours for Tower 1, seismic motion
type-3, sample #51 (IA ¼ 1:12 m=s). Deformation magnification
factor ¼ 5.
Plate 2. (Example 1) Tower 1, seismic motion type-3, sample #51 at time
T ¼ 5 s: (a) contours of maximum shear strains and deformed mesh and
(b) contours of EPWP ratio.
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 663
Acknowledgments
The financial support provided by NSERC under
Research Grant No. RG203795-02 and by NSF under
Grant No. 0075998 is gratefully acknowledged. The
authors are also indebted to Dr. B. Ghosh for providing
some of the results of centrifuge experiments analyzed in
Example 2.
References
[1] Terzaghi K. Theoretical soil mechanics. New York: Wiley; 1943.
[2] Biot MA. Theory of elasticity and consolidation for a porous
anisotropic solid. J Appl Phys 1955;26:182–5.
[3] Biot MA. Theory of propagation of elastic waves in a fluid-saturated
porous solid, I. Low frequency range. J Acoust Soc Am 1956;28:
168–91.
[4] Biot MA, Willis DG. The elastic coefficients of a theory of
consolidation. J Appl Mech ASME 1957;29:594–601.
[5] Simon BR, Wu JSS, Zienkiewicz OC, Paul DK. Evaluation of u–w
and u–p finite element methods for the dynamic response of saturated
porous media using one-dimensional models. Int J Numer Anal Meth
Geomech 1986(10):461–82.
[6] Simon BR, Wu JSS, Zienkiewicz OC. Evaluation of higher order,
mixed and hermitean finite element procedures for dynamic analysis
of saturated porous media using one-dimensional models. Int J
Numer Anal Meth Geomech 1986(10):483–99.
[7] Mesgouez A, Lefeuve-Mesgouez G, Chambarel A. Transient
mechanical wave propagation in semi-infinite porous media using a
finite element approach. Soil Dyn Earthquake Eng 2005;25(6):
421–30.
[8] Prévost JH. Mechanics of continuous porous media. Int J Eng Sci
1980;18(5):787–800.
[9] Green AE, Naghdi PM. A dynamical theory of interacting continua.
Int J Eng Sci 1965;3:231–41.
[10] Eringen AC, Ingram JD. A continuum theory of chemically reacting
media I. Int J Eng Sci 1965;3:197–213.
[11] Bowen RM. Compressible porous media models by use of the theory
of mixtures. Int J Eng Sci 1982;20(6):697–735.
[12] Prevost JH. DYNAFLOW—a nonlinear transient finite element
analysis program, Version 02. Technical Report, Department of Civil
and Environmental Engineering, Princeton University, Princeton,
NJ, 2002. /http://www.princeton.edu/dynaflow/S.
[13] Popescu R. Influence of seismic loading rate on soil liquefaction. Can
Geotech J 2002;39:331–44.
[14] Popescu R, Prevost JH, Deodatis G. Effects of spatial variability on
soil liquefaction: some design recommendations. Geotechnique
1997;47(5):1019–36.
[15] Popescu R, Prevost JH, Deodatis G. Characteristic percentile of soil
strength for dynamic analyses. Geotech Earthquake Eng Soil Dyn,
Geotech Spec Publ 75, ASCE 1998:1461–71.
[16] Popescu R, Chakrabortty P, Prevost JH. Fragility curves for tower
structures on stochastically variable soil. In: Proceedings of the
international conference on structural safety and reliability,
ICOSSAR05, Rome, Italy, June 2005 [on CD-ROM].
[17] Popescu R, Prevost JH, Deodatis G. 3D effects in seismic liquefaction
of stochastically variable soil deposits. Geotechnique 2005;55(1):
21–31.
[18] Madabhushi SPG, Schofield AN. Centrifuge modeling of tower
structures on saturated sands subjected to earthquake perturbations.
Geotechnique 1993;43(4):555–65.
[19] Ghosh B, Madabhushi SPG. Effects of localized soil inhomogeneity
in modifying seismic soil–structure interaction. In: Proceedings of the
16th ASCE engineering mechanics conference, Seattle, WA, July
2003.
[20] Prévost JH. Nonlinear transient phenomena in saturated porous
media. Comput Meth Appl Mech Eng 1982;30:3–18.
[21] Prévost JH. Wave propagation in fluid-saturated porous media: an
efficient finite element procedure. Soil Dyn Earthquake Eng
1985;4(4):183–202.
[22] Bathe KJ, Wilson EL. Numerical methods in finite element analysis.
Englewood Cliffs, NJ: Prentice-Hall; 1976.
[23] Zienkiewicz OC. The finite element method. 3rd ed. London,
England: McGraw-Hill; 1977.
[24] Hughes TJR, Liu WK. Implicit–explicit finite elements in nonlinear
transient analysis. Comput Meth Appl Mech Eng ASME 1978;45:
371–4.
[25] Hughes TJR, Liu WK. Implicit–explicit finite elements in nonlinear
transient analysis: implementation and numerical examples. J Appl
Mech ASME 1978;45:395–8.
[26] Castro G, Poulos J. Factors affecting liquefaction and cyclic mobility.
J Geotech Eng Div ASCE 1977;103(GT6):501–16.
[27] Marcuson III WF. Definition of terms related to liquefaction.
J Geotech Eng Div ASCE 1978;104(9):1197–200.
[28] Youd TL, et al. Liquefaction resistance of soils: summary report from
the 1996 NCEER and 1998 NCEER/NSF workshops on evaluation
of liquefaction resistance of soils. J Geotech Geoenviron Eng ASCE
2001;127(10):817–33.
[29] Prevost JH, Popescu R. Constitutive relations for soil materials.
Electron J Geotech Eng 1996;1 http://www.ejge.com/1996/Ppr9609/
Ppr9609.htm.
[30] Prevost JH. A simple plasticity theory for frictional cohesionless soils.
Soil Dyn Earthquake Eng 1985;4(1):9–17.
[31] Popescu R, Prevost JH. Centrifuge validation of a numerical model for
dynamic soil liquefaction. Soil Dyn Earthquake Eng 1993;12(2):73–90.
[32] Popescu R. Stochastic variability of soil properties: data analysis,
digital simulation, effects on system behavior. PhD thesis, Princeton
University, Princeton, NJ, 1995.
[33] Popescu R, Prevost JH. Comparison between VELACS numerical
‘class A’ predictions and centrifuge experimental soil test results. Soil
Dyn Earthquake Eng 1995;14(2):79–92.
[34] Jafari-Mehrabadi A, Popescu R. Mitigative measures against soil
liquefaction: numerical modeling and centrifuge experiments. Can
Geotech J, in review.
[35] Keanne CM, Prevost JH. An analysis of earthquake data observed at
the Wildlife Liquefaction Array Site, Imperial County, California. In:
Proceedings of the second US–Japan workshop on liquefaction, large
ground deformations and effects on lifelines, New York, 1989.
p. 39–53.
[36] Morris DV. The centrifugal modeling of dynamic soil–structure
interaction and earthquake behavior. PhD thesis, University of
Cambridge, 1979.
[37] Weissman K, Prevost JH. Centrifugal modeling of dynamic
soil–structure interaction. Technical Report NCEER-89-0040, Prin-
ceton University, 1989.
[38] Deodatis G. Non-stationary stochastic vector processes: seismic
ground motion applications. Probab Eng Mech 1996;11:149–68.
[39] Uniform Building Code. International conference of building officials
(ICBO), Whittier, CA, vol. 2, 1994.
[40] Arias A. A measure of earthquake intensity. In: Hansen RJ, editor.
Seismic design for nuclear power plants. Cambridge, MA: MIT Press;
1970. p. 438–83.
[41] Iwasaki Y, Tai M. Strong motion records at Kobe Port Island. Soils
and Foundations, Special Issue on Geotechnical Aspects of the
January 17 1995 Hyogoken-Nambu Earthquake, January 1996.
p. 29–40.
[42] Shinozuka M, Feng MQ, Lee J, Naganuma T. Statistical analysis of
fragility curves. J Eng Mech ASCE 2000;126(12):1224–31.
[43] Deodatis G, Saxena V, Shinozuka M. Effect of spatial variability of
ground motion on bridge fragility curves. In: Proceedings of the
eighth ASCE specialty conference on probabilistic mechanics
and structural reliability, University of Notre Dame, IL, 2000
[on CD-ROM].
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665
664
[44] Focht JA, Focht III JA. Factor of safety and reliability in
geotechnical engineering (discussion). J Geotech Geoenviron Eng
ASCE 2001;127(8):704–6.
[45] Popescu R, Deodatis G, Nobahar A. Effects of soil heterogeneity on
bearing capacity—qualitative assessment. Probab Eng Mech 2005;20:
324–41.
[46] Budiman JS, Mohammadi J, Bandi S. Effect of large inclusions on
liquefaction of sand. In: Proceedings of the conference of geotechni-
cal engineering division. Geotechnical Special Publication 56, ASCE;
1995. p. 48–63.
[47] Konrad J-M, Dubeau S. Cyclic strength of stratified soil samples. In:
Proceedings of the 55th Canadian geotechnical conference: ground and
water: theory to practice, Niagara Falls, Ont., October 2002. p. 89–94.
[48] Ohtomo K, Shinozuka M. Two-dimensional spatial severity of
liquefaction. In: Proceedings of the eighth Japan earthquake
engineering symposium, Tokyo, 1990.
[49] Fenton GA, Vanmarcke EH. Spatial variation in liquefaction risk.
Geotechnique 1998;48(6):819–31.
[50] Elkateb T, Chalaturnyk R, Robertson PK. Simplified geostatistical
analysis of earthquake-induced ground response at the Wildlife Site,
California, USA. Can Geotech J 2003;40:16–35.
[51] Seed HB, Idriss IM. Simplified procedure for evaluating soil
liquefaction potential. J Geotech Eng Div ASCE 1971;97(9):1249–73.
[52] Robertson PK, Wride CE. Evaluating cyclic liquefaction potential
using the cone penetration test. Can Geotech J 1998;35(3):442–59.
[53] Koutsourelakis S, Prevost JH, Deodatis G. Risk assessment of an
interacting structure–soil system due to liquefaction. Earthquake Eng
Struct Dyn 2002;31:851–79.
[54] Popescu R, Prevost JH, Deodatis G. Influence of spatial variability of
soil properties on seismically induced soil liquefaction. In: Proceed-
ings of the uncertainty in the geologic environment: from theory to
practice. Madison, WI: ASCE; 1996. p. 1098–112.
[55] Chakrabortty P, Jafari-Mehrabadi A, Popescu R. Effects of low
permeability soil layers on seismic stability of submarine slopes. In:
Proceedings of the 57th Canadian geotechnical conference, Quebec
City, PQ, 2004.
[56] Geotechnical and Environmental Research Group. Behaviour of
rigid foundations on layered soil, Report #331, Cambridge Uni-
versity, 2004. /http://www-civ.eng.cam.ac.uk/geotech_new/S.
[57] Zeng X, Schofield AN. Design and performance of an equivalent
shear beam container for earthquake centrifuge modeling. Geotech-
nique 1996;46(1):83–102.
ARTICLE IN PRESS
R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 665

Más contenido relacionado

Similar a 23. The effect of plastic fines on the pore pressure generation characteristics.pdf

Molecular Fractal Surfaces Analysis ICSE4-ffvi
Molecular Fractal Surfaces Analysis ICSE4-ffviMolecular Fractal Surfaces Analysis ICSE4-ffvi
Molecular Fractal Surfaces Analysis ICSE4-ffvifferrieu
 
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...IJERA Editor
 
Annals of Limnology and Oceanography
Annals of Limnology and OceanographyAnnals of Limnology and Oceanography
Annals of Limnology and Oceanographypeertechzpublication
 
Consolidation settlement with sand drains – analytical and numerical approaches
Consolidation settlement with sand drains – analytical and numerical approachesConsolidation settlement with sand drains – analytical and numerical approaches
Consolidation settlement with sand drains – analytical and numerical approachesUmed Paliwal
 
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...coreconferences
 
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...P singh
 
case study on terzaghi’s theory
case study on terzaghi’s theorycase study on terzaghi’s theory
case study on terzaghi’s theoryAbhishek Mangukiya
 
Ess 16-3959-2012
Ess 16-3959-2012Ess 16-3959-2012
Ess 16-3959-2012Simone88888
 
Structure soil-structure interaction literature review
Structure soil-structure interaction literature reviewStructure soil-structure interaction literature review
Structure soil-structure interaction literature reviewÁngel J. Alicea, EIT
 
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...P singh
 
Technical NoteSoil Compressibility Models for a Wide Stres.docx
Technical NoteSoil Compressibility Models for a Wide Stres.docxTechnical NoteSoil Compressibility Models for a Wide Stres.docx
Technical NoteSoil Compressibility Models for a Wide Stres.docxssuserf9c51d
 
Finite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateFinite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateeSAT Publishing House
 
Finite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateFinite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateeSAT Journals
 
Analysis pore pressuregen
Analysis pore pressuregenAnalysis pore pressuregen
Analysis pore pressuregengefyra-rion
 
1. Studies of Rainfall-induced Slope Failure.pdf
1. Studies of Rainfall-induced Slope Failure.pdf1. Studies of Rainfall-induced Slope Failure.pdf
1. Studies of Rainfall-induced Slope Failure.pdfSaktiIdrus1
 
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...Anastasia Dollari
 
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...Satyajit Mojumder
 

Similar a 23. The effect of plastic fines on the pore pressure generation characteristics.pdf (20)

Molecular Fractal Surfaces Analysis ICSE4-ffvi
Molecular Fractal Surfaces Analysis ICSE4-ffviMolecular Fractal Surfaces Analysis ICSE4-ffvi
Molecular Fractal Surfaces Analysis ICSE4-ffvi
 
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...
Magneto-Thermosolutal Convection in Maxwellian Dusty Viscoelastic Fluid throu...
 
Annals of Limnology and Oceanography
Annals of Limnology and OceanographyAnnals of Limnology and Oceanography
Annals of Limnology and Oceanography
 
Bm4201421428
Bm4201421428Bm4201421428
Bm4201421428
 
Consolidation settlement with sand drains – analytical and numerical approaches
Consolidation settlement with sand drains – analytical and numerical approachesConsolidation settlement with sand drains – analytical and numerical approaches
Consolidation settlement with sand drains – analytical and numerical approaches
 
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...
The Influence of Pore Air Pressure on Slope Stability Under Various Rainfall ...
 
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...
FIELD AND THEORETICAL ANALYSIS OF ACCELERATED CONSOLIDATION USING VERTICAL DR...
 
case study on terzaghi’s theory
case study on terzaghi’s theorycase study on terzaghi’s theory
case study on terzaghi’s theory
 
Ess 16-3959-2012
Ess 16-3959-2012Ess 16-3959-2012
Ess 16-3959-2012
 
Structure soil-structure interaction literature review
Structure soil-structure interaction literature reviewStructure soil-structure interaction literature review
Structure soil-structure interaction literature review
 
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...
TRANSIENT ANALYSIS ON 3D DROPLET OVER HORIZONTAL SURFACE UNDER SHEAR FLOW WIT...
 
Technical NoteSoil Compressibility Models for a Wide Stres.docx
Technical NoteSoil Compressibility Models for a Wide Stres.docxTechnical NoteSoil Compressibility Models for a Wide Stres.docx
Technical NoteSoil Compressibility Models for a Wide Stres.docx
 
Finite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateFinite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plate
 
Finite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plateFinite element analysis of a floating rectangular plate
Finite element analysis of a floating rectangular plate
 
Analysis pore pressuregen
Analysis pore pressuregenAnalysis pore pressuregen
Analysis pore pressuregen
 
1. Studies of Rainfall-induced Slope Failure.pdf
1. Studies of Rainfall-induced Slope Failure.pdf1. Studies of Rainfall-induced Slope Failure.pdf
1. Studies of Rainfall-induced Slope Failure.pdf
 
Sullivan
SullivanSullivan
Sullivan
 
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...
Nonlinear Darcy flow dynamics during ganglia stranding and mobilization in he...
 
Final-Thesis-present.pptx
Final-Thesis-present.pptxFinal-Thesis-present.pptx
Final-Thesis-present.pptx
 
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...
Physical review e volume 76 issue 3 2007 [doi 10.1103%2 fphysreve.76.036303] ...
 

Más de PinakRay2

21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf
21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf
21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdfPinakRay2
 
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...PinakRay2
 
25. Centrifuge validation of a numerical model for.pdf
25. Centrifuge validation of a numerical model for.pdf25. Centrifuge validation of a numerical model for.pdf
25. Centrifuge validation of a numerical model for.pdfPinakRay2
 
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdfPinakRay2
 
22. The effect of plastic fines on the pore pressure generation characteristi...
22. The effect of plastic fines on the pore pressure generation characteristi...22. The effect of plastic fines on the pore pressure generation characteristi...
22. The effect of plastic fines on the pore pressure generation characteristi...PinakRay2
 
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdfPinakRay2
 
17. A critical state interpretation for the cyclic liquefaction.pdf
17. A critical state interpretation for the cyclic liquefaction.pdf17. A critical state interpretation for the cyclic liquefaction.pdf
17. A critical state interpretation for the cyclic liquefaction.pdfPinakRay2
 

Más de PinakRay2 (7)

21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf
21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf
21. Cyclic and post-cyclic monotonic behavior of Adapazari soils.pdf
 
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...
24. Evaluation of temperature and freeze–thaw effects on excess pore pressure...
 
25. Centrifuge validation of a numerical model for.pdf
25. Centrifuge validation of a numerical model for.pdf25. Centrifuge validation of a numerical model for.pdf
25. Centrifuge validation of a numerical model for.pdf
 
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf
20. Evaluation of static and dynamic properties of sand–fines mixtures.pdf
 
22. The effect of plastic fines on the pore pressure generation characteristi...
22. The effect of plastic fines on the pore pressure generation characteristi...22. The effect of plastic fines on the pore pressure generation characteristi...
22. The effect of plastic fines on the pore pressure generation characteristi...
 
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf
14. Mechanism of Liquefaction Response in Sand–Silt Dynamic.pdf
 
17. A critical state interpretation for the cyclic liquefaction.pdf
17. A critical state interpretation for the cyclic liquefaction.pdf17. A critical state interpretation for the cyclic liquefaction.pdf
17. A critical state interpretation for the cyclic liquefaction.pdf
 

Último

XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXssuser89054b
 
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...ranjana rawat
 
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...ranjana rawat
 
Generative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTGenerative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTbhaskargani46
 
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLPVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLManishPatel169454
 
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...SUHANI PANDEY
 
Double rodded leveling 1 pdf activity 01
Double rodded leveling 1 pdf activity 01Double rodded leveling 1 pdf activity 01
Double rodded leveling 1 pdf activity 01KreezheaRecto
 
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Booking
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 BookingVIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Booking
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Bookingdharasingh5698
 
KubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlyKubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlysanyuktamishra911
 
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756dollysharma2066
 
chapter 5.pptx: drainage and irrigation engineering
chapter 5.pptx: drainage and irrigation engineeringchapter 5.pptx: drainage and irrigation engineering
chapter 5.pptx: drainage and irrigation engineeringmulugeta48
 
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...Christo Ananth
 
Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)simmis5
 
Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . pptDineshKumar4165
 
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Stand
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night StandCall Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Stand
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Standamitlee9823
 
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete Record
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete RecordCCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete Record
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete RecordAsst.prof M.Gokilavani
 
data_management_and _data_science_cheat_sheet.pdf
data_management_and _data_science_cheat_sheet.pdfdata_management_and _data_science_cheat_sheet.pdf
data_management_and _data_science_cheat_sheet.pdfJiananWang21
 

Último (20)

XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
 
Call Now ≽ 9953056974 ≼🔝 Call Girls In New Ashok Nagar ≼🔝 Delhi door step de...
Call Now ≽ 9953056974 ≼🔝 Call Girls In New Ashok Nagar  ≼🔝 Delhi door step de...Call Now ≽ 9953056974 ≼🔝 Call Girls In New Ashok Nagar  ≼🔝 Delhi door step de...
Call Now ≽ 9953056974 ≼🔝 Call Girls In New Ashok Nagar ≼🔝 Delhi door step de...
 
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
The Most Attractive Pune Call Girls Manchar 8250192130 Will You Miss This Cha...
 
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...
The Most Attractive Pune Call Girls Budhwar Peth 8250192130 Will You Miss Thi...
 
Generative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTGenerative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPT
 
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELLPVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
PVC VS. FIBERGLASS (FRP) GRAVITY SEWER - UNI BELL
 
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
VIP Model Call Girls Kothrud ( Pune ) Call ON 8005736733 Starting From 5K to ...
 
Double rodded leveling 1 pdf activity 01
Double rodded leveling 1 pdf activity 01Double rodded leveling 1 pdf activity 01
Double rodded leveling 1 pdf activity 01
 
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Booking
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 BookingVIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Booking
VIP Call Girls Ankleshwar 7001035870 Whatsapp Number, 24/07 Booking
 
KubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghlyKubeKraft presentation @CloudNativeHooghly
KubeKraft presentation @CloudNativeHooghly
 
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756
FULL ENJOY Call Girls In Mahipalpur Delhi Contact Us 8377877756
 
chapter 5.pptx: drainage and irrigation engineering
chapter 5.pptx: drainage and irrigation engineeringchapter 5.pptx: drainage and irrigation engineering
chapter 5.pptx: drainage and irrigation engineering
 
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...
Call for Papers - Educational Administration: Theory and Practice, E-ISSN: 21...
 
Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)Java Programming :Event Handling(Types of Events)
Java Programming :Event Handling(Types of Events)
 
Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . ppt
 
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
(INDIRA) Call Girl Aurangabad Call Now 8617697112 Aurangabad Escorts 24x7
 
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Stand
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night StandCall Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Stand
Call Girls In Bangalore ☎ 7737669865 🥵 Book Your One night Stand
 
Roadmap to Membership of RICS - Pathways and Routes
Roadmap to Membership of RICS - Pathways and RoutesRoadmap to Membership of RICS - Pathways and Routes
Roadmap to Membership of RICS - Pathways and Routes
 
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete Record
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete RecordCCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete Record
CCS335 _ Neural Networks and Deep Learning Laboratory_Lab Complete Record
 
data_management_and _data_science_cheat_sheet.pdf
data_management_and _data_science_cheat_sheet.pdfdata_management_and _data_science_cheat_sheet.pdf
data_management_and _data_science_cheat_sheet.pdf
 

23. The effect of plastic fines on the pore pressure generation characteristics.pdf

  • 1. Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 Dynamics of nonlinear porous media with applications to soil liquefaction Radu Popescua , Jean H. Prevostb,, George Deodatisc , Pradipta Chakraborttya a Faculty of Engineering and Applied Science, Memorial University, St. John’s, Nfld, Canada A1B 3X5 b Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ 08544, USA c Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, NY 10027, USA Abstract This paper provides a description of the extension of Biot’s theory for dynamic behavior of saturated porous media into the nonlinear regime that was introduced by the second author in 1980. It also provides a finite element implementation of this extension and two numerical applications involving the seismic behavior of saturated soil deposits. In the first numerical application, the dynamic interaction between liquefying soil and a structure sitting on the ground surface is examined, with emphasis on the interplay between the seismic loading rate and the (evolving) characteristic frequency of the soil–structure system. The attenuation of seismic energy as the seismic waves pass through softened soil is also discussed. The second numerical application involves the seismically induced liquefaction of stochastically spatially variable soils. It is explained why more pore-water pressure is generated in a heterogeneous soil than in a corresponding uniform soil. Comparisons are also provided with experimental centrifuge data. r 2006 Elsevier Ltd. All rights reserved. 1. Introduction Soils consist of an assemblage of particles with different sizes and shapes which form a skeleton whose voids are filled with water and air or gas. The word ‘‘soil’’, therefore, implies a mixture of assorted mineral grains with various fluids. Hence, soil in general must be looked at as a one- phase (dry soil), two-phase (saturated soil), or multi-phase (partially saturated soil) material whose state is to be described by the stresses and displacements (velocities) within each phase. There are still uncertainties on how to deal analytically with partially saturated soils, and this is currently an area of active research. Attention is restricted in the following to fully saturated soils. The stresses carried by the soil skeleton are conventionally called ‘‘effective stresses’’ in the soil mechanics literature [1], and those in the fluid phase are called ‘‘pore-fluid pressures’’. In a saturated soil, when free drainage conditions prevail, the steady-state pore-fluid pressures depend only on the hydraulic conditions and are independent of the soil skeleton response to external loads. Therefore, in that case, a single-phase continuum description of soil behavior is certainly adequate. Similarly, a single-phase description of soil behavior is also adequate when no drainage (i.e., no flow) conditions prevail. However, in intermediate cases in which some flow can take place, there is an interaction between the skeleton strains and the pore-fluid flow. The solution of these problems requires that soil behavior be analyzed by incorporating the effects of the transient flow of the pore-fluid through the voids, and therefore requires that a two-phase continuum formulation be available for saturated porous media. Such a theory was first developed by Biot [2–4] for an elastic porous medium. Applications of Biot’s theory to finite element analysis of saturated porous media have been reported by numerous authors (see, e.g., Refs. [5,6] for a summary of the wide range of existing finite element formulations). Mesgouez et al. [7] present a review of work on applications of Biot’s theory in transient wave propagation in saturated porous media. In most formulations, the soil skeleton has either elastic or visco- elastic behavior. An extension of Biot’s theory into the nonlinear inelastic range is presented by Prévost [8], where soil is viewed as a ARTICLE IN PRESS www.elsevier.com/locate/soildyn 0267-7261/$ - see front matter r 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.soildyn.2006.01.015 Corresponding author. Tel.: +1 609 258 5424; fax: +1 609 258 2760. E-mail address: prevost@princeton.edu (J.H. Prevost).
  • 2. multi-phase medium and the modern theories of mixtures developed by Green and Naghdi [9] and Eringen and Ingram [10] are used. General mixture results can be shown through formal linearization of the field and constitutive equations, to reduce Biot linear poroelastic model (see, e.g., Ref. [11]). The resulting equations are summarized in Section 2. The first part of this paper presents briefly Biot’s basic theory of dynamics of saturated porous media, its extension into the nonlinear regime [8], and its implemen- tation in the finite element code DYNAFLOW [12]. Two numerical examples are presented next, illustrating this extension of Biot’s theory into the nonlinear dynamic behavior of saturated soils. Both examples are presented in relation to centrifuge experiments of structures on homo- geneous or non-homogeneous saturated soil deposits subjected to seismic loads. The first numerical example provides results of a study aimed at better understanding the loading and failure mechanisms of structures on liquefiable soil, while the second example examines the mechanisms of excess pore-water pressure (EPWP) build- up and liquefaction of natural soil deposits exhibiting stochastic spatial variability of their properties. The first example is an original study extending the results of previous research on earth dams (e.g., [13]) to structures on liquefiable soil deposits. The second example is also an original study that continues and expands on previous work of the authors on seismically induced liquefaction of randomly heterogeneous soil [14–17]. Based on detailed analysis of both experimental and numerical results, it provides an interpretation for a particularly interesting and important behavior detected in previous work, namely that more pore-water pressure is generated by seismic loads in a heterogeneous soil than in an equivalent uniform soil. The numerical calculations presented in both examples are validated based on results of recent centrifuge experiments [18,19]. Besides demonstrating the capabilities of the nonlinear extension of Biot’s theory implemented in DYNAFLOW to handle relatively complex phenomena (such as dynamic soil–structure interaction and seismic behavior of randomly variable soils), these studies provide new insight into the mechanisms of seismic behavior of soil–structure systems and interesting practical results for geotechnical design. 2. Dynamics of porous media 2.1. Basic theory During deformation, the solid particles, which form the soil skeleton, undergo irreversible motions such as slips at grain boundaries, creations of voids by particles coming out of a packed configuration, and combinations of such irreversible motions. When the particulate nature and the microscopic origin of the phenomena involved are not explicitly sought, phenomenological equations then pro- vide an adequate description of the behavior of the various phases which form the soil medium. In multi-phase theories, the conceptual model is, thus, one in which each phase (or constituent) enters through its averaged proper- ties obtained as if the particles were smeared out in space. In other words, the particles nature of the constituents is described in terms of phenomenological laws as the particles behave collectively as a continuum. Soil is, thus, viewed herein as consisting of a solid skeleton interacting with the pore fluids. Balance of linear momentum equations for each constituent then can be written simply as (see, e.g., Ref. [8]) rS aS ¼ r s0S ð1 nw Þ rpw x ðvS vw Þ þ rS b, (1) rw dvw dt ¼ rw ðvS vw Þ rvw nw rpw þ x ðvS vw Þ þ rw b, ð2Þ where the motion of the solid phase is used as the reference motion. In Eqs. (1) and (2), s0s is the solid (effective) stress, as the solid acceleration, vs (vw ) the solid (fluid) velocity, b the body force (per unit mass), pw the pore-fluid pressure, rS ¼ ð1 nw Þrs and rw ¼ nw rw with rs the solid mass density, rw the fluid mass density, and nw the porosity; x ¼ nw2 gwk1 where k is the hydraulic conductivity (L/T) and gw ¼ rwg the fluid unit weight, g ¼ ðjjbjjÞ acceleration of gravity. In the case of a compressible pore fluid, the pore-fluid pressure is determined from the computed velocities through time integration of the following equation: dpw dt ¼ ðlw =nw Þ r ð1 nw ÞvS þ r ðnw vw Þ , (3) where lw is the fluid bulk modulus. In the case of an incompressible pore fluid, the pore-fluid pressure is determined from the computed velocities through the following equation: pw ¼ ðlw =nw Þ r ð1 nw ÞvS þ r ðnw vw Þ , (4) where lw is a penalty parameter used to enforce the incompressibility constraint [20,21]. 2.2. Weak form/semi-discrete finite element equations The weak formulation associated with the initial boundary value problem is obtained by proceeding along standard lines (see, e.g., Refs. [22,23]). The associated matrix problem is obtained by discretizing the domain occupied by the porous medium into non-overlapping finite elements. Each element is defined by nodal points at which shape functions are prescribed. In general, two sets of shape functions are required for the solid displacement and the fluid velocity field, respectively. However, attention in the following is restricted to low-order (i.e., four-node plane and eight-node brick) finite elements which are the most efficient in nonlinear analysis, and thus the same shape functions are used for both the solid and the fluid. The shape functions for the solid displacement and fluid ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 649
  • 3. velocity associated with node A are denoted by NA in the following. They satisfy the relation NA ðxB Þ ¼ dAB in which xB denotes the position vector of node B and dAB ¼ Kro- necker delta. The Galerkin counterpart of the weak formulation is expressed in terms of the shape functions and gives rise to the following system of equation: Ms 0 0 Mw # as aw ( ) þ Z Z ZT Z # vs vw ( ) þ ns nw ( ) ¼ fs ext fw ext ( ) , ð5Þ where Ma , aa , va , na , and fa ext represent the (generalized) mass matrix, acceleration, velocity, and force vectors, respectively (a ¼ s; w for the solid and fluid phases, respectively). The element contributions to node A from node B for directions i and j to the matrices appearing in Eq. (5) are defined in the following. The terms are integrated over the spatial domain occupied by the element, Oe . For computational simplicity, a diagonal ‘‘lumped mass’’ matrix is used: mAB ij a ¼ dijdAB Z Oe ra NA dO. (6) The momentum transfer terms give rise to the damping matrix Z: ZAB ij ¼ Z Oe NA xijNB dO. (7) The external force ðfa Þext (i.e., body force, surface traction) is computed as follows: ðfA i Þa ext ¼ Z Oe ra biNA dO þ boundary terms: (8) The internal stress forces na are computed as follows: ðnA i Þs ¼ Z Oe ðs0s ij ns pwdijÞNA ;j dO (9) and ðnA i Þw ¼ Z Oe rw NA ðvw j vs j Þ dO Z Oe NA ;i nw pw dO, (10) where ns ¼ 1 nw . 2.3. Time integration Time integration of the semi-discrete finite element equations is accomplished by a finite difference time- stepping algorithm. In general, implicit and explicit integration procedures are available. Explicit procedures are computationally the simplest since they do not require (for a diagonal mass matrix) equation solving to advance the solution. However, stability restricts the size of the allowable time step. On the other hand, unconditional stability can usually be achieved in implicit procedures, which require solution of a system of equations at each time step. For the problem at hand, a purely explicit procedure is not usually appropriate because of the unreasonably stringent time-step restriction resulting from the presence of the very stiff fluid in the mixture (even for highly nonlinear solid material models). The methods used combine the attractive features of explicit and implicit integration (see, e.g., Refs. [24,25]), and fall under the category of ‘‘split operator methods’’. Different portions of the system of equations are treated implicitly and explicitly, reducing the system of equations to be solved. The specific split to be made is obviously problem dependent, and the appropriate implicit/explicit splits for the problem at hand are detailed in Refs. [20,21]. 2.4. Nonlinear iterations At each step the resulting nonlinear algebraic problem is solved by Newton–Raphson type iterations. Quasi-Newton procedures with line searches are typically used. 3. Modeling seismic-induced soil liquefaction Liquefaction (as defined by Castro and Poulos [26]) is a phenomenon wherein a saturated sand subjected to monotonic or cyclic shear loads looses a large percentage of its shear resistance and flows in a manner resembling a liquid. Marcuson [27] defines liquefaction as the transfor- mation of a granular material from a solid to a liquefied state as a consequence of increased pore-water pressure and reduced effective stress. This phenomenon occurs most readily in loose to medium dense granular soils that have a tendency to compact when sheared. In saturated soils, pore-water pressure drainage may be prevented due to the presence of silty or clayey seam inclusions, or may not have time to occur due to rapid loading—such as in the case of seismic loads. In this situation, the tendency to compact is translated into an increase in pore-water pressure. This leads to a reduction in effective stress, and a corresponding decrease of the frictional shear strength. If the EPWP generated at a certain location in a purely frictional soil (e.g., sand) reaches the initial value of the effective vertical stress, then, theoretically, all shear strength is lost at that location and the soil liquefies and behaves like a viscous fluid. Liquefaction-induced large ground deformations are a leading cause of disasters during earthquakes. EPWPs are induced also in moderately dense to dense granular materials subjected to cyclic loads, but due to their tendency to dilate during shear, the softening is only temporary leading to increased cyclic shear strains, but not to major strength loss and large ground deformations [28]. This phenomenon is known as cyclic mobility. Two important aspects need to be addressed in a rigorous seismic analysis of saturated soils involving EPWP build-up: (1) coupled analysis: solid and fluid coupled field equations have to be used in a step-by-step (time domain) dynamic analysis to correctly capture the inertial and dissipative coupling terms; and (2) soil constitutive model: accurate simulation of dynamically ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 650
  • 4. induced EPWP build-up and continuous softening of the material requires soil models able to reproduce the experimentally observed nonlinear hysteretic behavior and shear stress-induced anisotropic effects, and to reflect the strong dependency of plastic dilatancy on effective stress ratio. Advanced plasticity models, such as multi-yield or bounding surface plasticity, in combination with kinematic hardening rules, can offer a material representation of considerable power and flexibility (e.g., [29]). As such soil constitutive models usually require a relatively large number of parameters, a well-defined methodology for calibrating those parameters based on results of standard in situ and/or laboratory soil tests is also desirable. The first aspect is addressed in the DYNAFLOW code by the extension of Biot’s theory into the nonlinear regime, discussed in Section 2. Nonlinear dynamic constitutive behavior of saturated soil under partially drained condi- tions is modeled using a kinematic hardening, multi-yield constitutive model based on a simple plasticity theory [30]. The yield function is described in the principal stress space by a set of nested rounded Mohr–Coulomb yield surfaces. A non-associative plastic flow rule is used for the dilatational component of the plastic deformation. The model has been tailored (1) to retain the extreme versatility and accuracy of the simple multi-surface J2 theory in describing observed shear nonlinear hysteretic behavior and shear stress-induced anisotropic effects, and (2) to reflect the strong dependency of the shear-induced dila- tancy on the effective stress ratio. Accurate simulation of shear-induced plastic dilation and of hysteretic effects under cyclic loading, together with full coupling between solid and fluid equations, allows capturing the build-up and dissipation of pore-water pressures and modeling the gradual softening and hardening of soil materials. The required constitutive model parameters can be derived from the results of conventional laboratory (e.g., triaxial, simple shear [31]) or in situ (e.g., standard penetration, cone penetration, wave velocity [32]) soil tests. The multi- yield plasticity soil constitutive model, its implementation algorithm, and the methodology for estimating the constitutive model parameters have been repeatedly validated in the past for soil liquefaction computations, based on both centrifuge experimental results (e.g., [31,33,34]) and full-scale measurements (e.g., [35]). Practical applications of the extension of Biot’s theory into the nonlinear regime, including its implementation in DYNAFLOW, are presented in the remaining of this paper through results of two studies involving simple structures on homogeneous or non-homogeneous saturated soil deposits subjected to seismic loads. 4. Example 1: tower structures on liquefiable soil Any soil–structure system has its own characteristic frequency, which depends on material properties, geome- try, degree of saturation of the soil, and contact conditions at the soil–foundation interface. This characteristic fre- quency may decrease during dynamic excitation due to degradation of the soil effective shear modulus as a result of pore pressure build-up and/or large shear strains. Any mechanical system is more sensitive to dynamic loading as its characteristic frequency becomes closer to the frequency range corresponding to the maximum spectral values of the excitation. Consequently, both the frequency content of seismic excitation and the evolution of structural frequency characteristics can make a significant difference in the dynamic response of soil–structure systems. Several centrifugal studies of seismic behavior of tower structures on soil foundation are reported, among which: Morris [36] and Weissman and Prevost [37]—tower struc- tures on dry soil, Madabhushi and Schofield [18]—tower structures on saturated soil. Several tests with different input seismic acceleration have been performed to study the effect of characteristic frequency. Madabhushi and Schofield [18] concluded that, when the predominant frequency of the input seismic motion was lower than the characteristic frequency of the system, the characteristic frequency of the system decreased during shaking to a value close to the predominant input seismic frequency due to build up of excess pore pressure and subsequent degradation of soil stiffness. Popescu [13] performed nonlinear dynamic finite element analyses of an embankment dam on liquefiable foundation and concluded that the interplay between the frequency content of the seismic motion, the vibration characteristics of the structure and the possible evolution of those characteristics during shaking had significant influence on the predicted dynamic response. In this example, the behavior of two different tower structures (with different characteristic frequencies but with the same mass) resting on a homogeneous liquefiable soil deposit is investigated for a range of intensities of the seismic input motion. The saturated soil is modeled as a two-phase porous medium, and the finite element calcula- tions are conducted in terms of effective stress, using fully coupled solid–fluid equations. The soil and structures analyzed in this study correspond to those used by Madabhushi and Schofield [18] in a series of centrifuge experiments. The earthquake motion used here spans a wide range of seismic intensities and is different from that used in the centrifuge experiments. The main purpose of this study is to investigate in detail the changes in soil strength induced by EPWP build-up, leading to changes in the vibration characteristics of a soil–structure system during and after the earthquake. It is shown how those changes and their interplay with the frequency content of the seismic excitation affect: (1) the dynamic excitation of the structure and (2) the amount of seismic energy transmitted to the system and, eventually, the structural response. 4.1. Finite element model and analysis method Two different structures are analyzed. Tower 1 with a characteristic frequency of 4.8 Hz, corresponding to a ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 651
  • 5. two-story building, and Tower 2 with a characteristic frequency of 1.4 Hz, corresponding to a seven-story building. The dynamic characteristics of both structures are assumed to be dominated by their first mode of vibration, and the structures are idealized as single degree of freedom systems. The plane strain assumption is used in the finite element analyses, and the foundation is idealized as a 10 m wide continuous strip footing. The factor of safety for bearing capacity under static conditions and assuming local bearing capacity failure is about 20 for both structures. In the numerical analyses, the structure material is idealized as linear elastic. The soil, assumed saturated below the foundation (ground water table at 1 m depth), is idealized as a two-phase porous material and discretized into four-node quadrilateral continuum two-phase porous elements with four degrees of freedom per node (two for the solid phase and two for the fluid phase displacements). The soil constitutive parameters are listed in Table 1. Use of a rigid box in an experimental earthquake simulation induces a series of spurious seismic waves that are generated by the motion of each end wall of the box and are subsequently reflected back by the opposite wall. This phenomenon, which may induce a behavior in the model different from the real field, is partly attenuated by placing a certain amount of duxseal (a relatively soft material) at each end wall of the box to create absorbing boundaries. It is assumed that those phenomena (wave reflection due to simulation of rigid end walls and subsequent attenuation when modeling the presence of duxseal) are reproduced by the numerical model. Duxseal has been used in the centrifuge experiments to prevent seismic wave reflection from the lateral boundaries of the rigid centrifuge box and thus helping the soil deposit to behave like a semi-infinite medium [18]. This material is considered also in the numerical analyses, and idealized as linear elastic. The material properties assumed for duxseal and structure are listed in Table 2. The finite element mesh is shown in Fig. 1. The seismic acceleration is applied at the base and lateral boundaries of the analysis domain, same as in the centrifuge experiment performed in a rigid box. The base input accelerations are generated using a procedure for simulation of non-stationary stochastic processes [38] capable of generating seismic ground motion time histories that are compatible with prescribed response spectra and have a prescribed modulating envelope for amplitude variation. Two hundred acceleration time histories have been generated to be compatible with the response spectra type-1 and type-3 recommended by the Uniform Building Code [39] and corresponding to different soil conditions. All 100 generated acceleration time histories in each group are compatible to the corresponding response spectrum, ARTICLE IN PRESS Table 1 Parameters of the multi-yield plasticity model, and values used for the saturated soil in Example 1 Constitutive parameter Symbol Value Type Mass density—solid rs 2710 kg/m3 State parameters Porosity nw 0.43 Hydraulic conductivity k 0.0002 m/s Low strain elastic shear modulus G0 20 MPa Low strain elastic parameters Poisson’s ratio n 0.33 Power exponent n 0.5 Friction angle at failure f 381 Yield and failure parameters Maximum deviatoric strain (comp/ext) edev max 0.08/0.06 Coefficient of lateral stress k0 0.426 Stress–strain curve coefficient a 0.21 Dilation angle c 321 Dilation parameters Dilation parameter Xpp 0.2 Table 2 Constitutive parameters for linear elastic materials used in Examples 1 and 2 Constitutive parameter Duxseal (Examples 1 and 2) Tower structures (Example 1) Structurea (Example 2) ESB boxb Dural Rubber Mass density (kg/m3 ) 1650 3530 3200 3953 3953 Young’s modulus (MPa) 8 73,000 73,000 73,000 158 Poisson’s ratio 0.46 0.33 0.33 0.33 0.45 a Mass calculated to obtain a pressure on soil equivalent to the one in the centrifuge experiments. b Mass density for both materials and Young’s modulus of rubber have been fitted to obtain equivalent vibration and deformability characteristics with those of the ESB box used in the centrifuge experiments. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 652
  • 6. and consequently have similar frequency content. They also have similar duration of the strong motion. Aside from the above common characteristics, each individual acceleration time history is different from the others in the same group (e.g., the peak values occur at different time instants in each sample), as each time history constitutes a different sample of the target non-stationary stochastic process. The accelerations in each group have been scaled to a range of Arias intensities [40], IA ¼ 0:1922:1 m=s, corresponding to a range of peak ground accelerations, PGA ¼ 0:120:4 g. The duration of seismic motion con- sidered here is about 15 s and the total analysis time is 40 s. 4.2. Analysis results Contours of EPWP ratio with respect to the initial effective vertical stress calculated for one case (Tower 1 subjected to type-3 acceleration sample #51, with IA ¼ 1:12 m=s and PGA ¼ 0:25 g) are shown in Color Plate 1 (at the end of the paper). The soil in the free field is predicted to liquefy, but, due to increased overburden pressure, the soil below the structure does not liquefy. There is, however, significant pore-water pressure build-up and subsequent softening of the soil in this area, leading to large settlements and rotations of the structure. The numerical model also reproduces gradual dissipation of pore-water pressure after the end of the earthquake (Color Plate 1c). Color Plate 2 illustrates the predicted failure mechanism for the case presented in Color Plate 1. At time T ¼ 5 s, the average EPWP ratio in the area below the structure is about 75% (Color Plate 2b), and the corresponding reduction in effective stress leads to a factor of safety for local bearing capacity failure equal to 1. The failure mechanism illustrated by the maximum shear strain contours in Color Plate 2a is characteristic for a local bearing capacity failure, consisting of an active wedge below the structure, clearly shown by the shear strain contours, and fading transition and passive wedges. The failure wedges are not symmetric, due to the presence of lateral inertial forces. Due to softening of soil below the structure, the seismic motion is strongly attenuated, especially for large Arias intensities (IA), and the acceleration at the base of the structure has significantly less specific energy compared to the input. In Fig. 2, it is shown how at low seismic intensity (sample #1 with IA ¼ 0:19 m=s) the seismic energy is entirely transmitted to the structure (Figs. 2a and b), while for high seismic intensities (sample #90 with IA ¼ 1:84 m=s, where the soil liquefies), the input seismic energy is largely attenuated by the soil (Fig. 2c), especially for the high frequency components of the base motion (Fig. 2d). For the second case, it can be noted from Fig. 2c how the acceleration amplitudes at the base of the structure are similar to those of the input motion for the first 2–3 s of shaking, when the pore-water pressures in the soil are still small. After that, as the soil starts to soften due to build-up ARTICLE IN PRESS Fig. 1. Seismic analysis of a tower structure: finite element mesh. -1 0 1 m/s2 0 5 10 15 20 -3 -2 -1 0 1 2 3 input base of struct. m/s2 Time (sec) (c) (a) (b) (d) 10-1 100 101 10-2 100 Sample # 1 Sample # 1 10-1 100 101 10-2 100 Period (sec) Response Spectrum(m/s 2 ) input base of struct. Sample # 90 Sample # 90 Fig. 2. Comparison between acceleration time histories (a, c) and response spectra (b, d) used as input at the base of the mesh and computed at the base of the structure for two samples of type-1 acceleration used for Tower 1 (corresponding to two different Arias intensity values). R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 653
  • 7. of EPWP, the maximum accelerations transmitted to the structure base become significantly lower than the ones input at the base of the analysis domain. It can be noted from the computed acceleration response spectrum shown in Fig. 2d that the high frequency components of the seismic acceleration are predominantly attenuated, while the energy of the low frequency components that are closer to the natural frequency of the liquefied soil deposit is not very much affected by the seismic wave propagation through soil. This is also evident from the low frequency sinusoidal shape exhibited by the computed acceleration time history at the base of the structure (red line in the electronic version and darker line in the paper version in Fig. 2c). The shape of this computed acceleration time history is similar to accelerations recorded at shallow depth in a landfill that liquefied during the Hyogoken-Nambu Earthquake (e.g., [41]). It is mentioned that the computed reduction in accel- eration amplitude at the base of the structure for large Arias intensities of the input seismic motion should not be misinterpreted as a safer situation for the case in Fig. 2c (sample #90 with IA ¼ 1:84 m=s) than for the case in Fig. 2a (sample #1 with IA ¼ 0:19 m=s). Even if structure accelerations are in general smaller for sample #90 than for sample #1, the structure shaken by sample #90 is predicted to fail due to bearing capacity failure of the liquefied soil, while the structure shaken by sample #1 is predicted to be safe. The effects of earthquake intensity and loading rate on the settlement and rotation of the rigid foundation for the two types of structures are presented in Fig. 3 in terms of fragility curves, which are an illustrative and practical way of expressing the probability of exceeding a certain degree of structural response as a function of load intensity. Fragility curves have been used extensively in earthquake engineering to describe the seismic vulnerability of structures as a function of the severity of the seismic event (e.g., [42,43]). The procedure used here for establishing the fragility curves for foundation settlements and rotation follows that proposed by Shinozuka et al. [42]. After selecting a specific response threshold, all earthquake intensities IA for which the response exceeds this threshold are assigned to the unity probability level, while all the other intensities are assigned to the zero probability level. Then, a shifted lognormal distribution function is fitted to those points using the maximum likelihood method. Fragility curves can include effects of multiple sources of uncertainty related to material resistance or load char- acteristics. The curves used in this example reflect only uncertainty in the seismic ground motion. The curves presented here express the probability of exceeding certain thresholds in the response, namely foundation settlement of 20 cm in Fig. 3a and base rotations of 1/300 in Fig. 3b. Those results represent earthquake effects in terms of rigid motion of the structure. From both figures it can be concluded that, for the same level of seismic intensity, type-3 acceleration has stronger effect on both settlements and rotations than type-1 acceleration. This is due to a change in vibration characteristics of the soil to fundamental frequencies close to the dominant frequency values of type-3 acceleration. The same behavior has also been observed by Madabhushi and Schofield [18] in the centrifuge experiments. The initial characteristic periods of the soil–structure system (before the earthquake) are T1 ¼ 0:6 s, for Tower 1, and T2 ¼ 0:8 s, for Tower 2. During the earthquake, due to significant increase in pore-water pressure and subsequent soil softening, this characteristic period goes up. For most cases analyzed here, when the soil liquefies in the free field and the EPWP increases significantly below structure, the characteristic period of the soil–structure system for both Towers 1 and 2 goes up to about 1.5 s. Consequently, the seismic energy input at the base of the soil layer is delivered, for most of the earthquake duration, to a system that has a characteristic period of about 1.5 s. As shown in Fig. 4, this value is closer to the range of maximum spectral amplitudes of type-3 acceleration than to those of type-1, and this results in significantly more energy delivered at the resonance frequency to the soil–structure system when acted by the type-3 input than when acted by type-1 input. ARTICLE IN PRESS 0 0.5 1 1.5 2 0 0.2 0.4 0.6 0.8 1 Arias Intensity (m/s) Prob. that rotation exceeds 1/300 0 0.5 1 1.5 2 0 0.2 0.4 0.6 0.8 1 Arias Intensity (m/s) Prob. that settlement exceeds 20cm Tower1 - Type 1 Tower2 - Type 1 Tower1 - Type 3 Tower2 - Type 3 Tower1 - Type 1 Tower2 - Type 1 Tower1 - Type 3 Tower2 - Type 3 (a) (b) Fig. 3. Comparison between analysis results in terms of fragility curves: (a) settlements and (b) rotations (type-1 and type-3 refer to the corresponding UBC response spectra, and Towers 1 and 2 refer to the two structures considered). R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 654
  • 8. For example, for the case illustrated in Fig. 4, more than two times more energy is delivered at the resonance frequency by the type-3 input acceleration than by the type-1 input. It is mentioned that when computing the characteristic period before the earthquake and at various time instants during the dynamic analysis, the coupling effects in Eqs. (1) and (2) have been neglected. It can be noticed from Fig. 3 that, for type-1 input motion, the fragility curves calculated for Tower 1 are located to the left of the ones calculated for Tower 2. This means that a lower value of seismic intensity corresponds to a given probability of seismic damage for Tower 1 than for Tower 2. Apparently Tower 2, more flexible, can attenuate some part of the seismic energy by structural deflections. It is also interesting that, when subjected to seismic motion type-3, where the degree of soil liquefaction is higher than for type-1 input, the computed responses of both structures are similar, as both structures behave as rigid blocks compared to the deformability of liquefied soil. 5. Example 2: effects of stochastic soil variability on liquefaction 5.1. Results from previous studies Natural soil properties vary from one point to another, even within so-called ‘‘uniform’’ soil layers. In addition to inducing uncertainty in the computed response, natural spatial variability of soil properties within geologically distinct layers affects the mechanical behavior of geotech- nical structures. For example, for phenomena involving the presence of a failure surface (such as encountered in landslides or in bearing capacity failures) the actual failure surface can deviate from its theoretical position to pass selectively through weaker soil zones and thus the average mobilized strength is reduced when compared to that of a corresponding uniform soil (e.g., [44,45]). For the case of seismically induced soil liquefaction, it was shown (e.g., [14]) that a larger amount of EPWP is generated in a heterogeneous soil than in the corresponding uniform soil having geomechanical properties equal to the average properties of the variable soil. Experimental evidence of the effects of soil heterogeneity on liquefaction resistance is presented in Refs. [46,47]. Budiman et al. [46] performed a series of undrained cyclic triaxial tests on sand specimens containing up to 25% gravel inclusions. They found that the liquefaction resistance of sand with inclusions was lower than that of uniform sand. Moreover, the reduction in liquefaction resistance was more pronounced for samples with a higher content of gravel. Konrad and Dubeau [47] conducted undrained cyclic triaxial tests on fine Ottawa sand and on silica silt. The study consisted of three series of tests: the first two series were conducted to characterize the cyclic strength of the sand and of the silt, respectively. The third series was performed on layered samples, consisting of one horizontal silt layer sandwiched between two sand layers. The study results revealed that layering induced a much lower cyclic resistance than that developed in either of the materials in uniform samples. Pioneering work in numerical analysis of the effects of soil heterogeneity on liquefaction resistance was presented by Ohtomo and Shinozuka [48]. Later on, Popescu [32] and Popescu et al. [14,15] conducted a systematic study of those effects using the multi-yield surface constitutive model implemented in DYNAFLOW. A Monte-Carlo simulation (MCS) method, combining digital generation of non- Gaussian stochastic vector fields representing the spatial distribution of various soil properties over the analysis domain with dynamic effective stress finite element analyses, was used to quantify the effects of soil hetero- geneity on the amount and pattern of EPWP build-up in a saturated soil deposit subjected to seismic loads. It was shown that, for the same average values of soil parameters, more EPWP build-up was predicted in stochastic analyses, accounting for soil heterogeneity, than in so-called ‘‘deterministic analyses’’, assuming uniform soil properties. Characteristic percentiles of soil strength were proposed for use in deterministic analyses, resulting in a response similar to that predicted by more expensive stochastic analyses. Effects of the degree of soil variability and of the seismic loading rate were also investigated. Fenton and Vanmarcke [49] addressed the three-dimensional (3D) aspect of liquefaction of randomly variable soils through an analysis of liquefaction potential at the Wildlife Site, Imperial Valley, California. The soil properties were modeled as a 3D random field; however, because of limited computa- tional resources, the liquefaction analysis was carried out in 1D vertical columns, without any coupling in the horizontal plane. Consequently, the study could only address the pointwise initiation of liquefaction, and did ARTICLE IN PRESS 10-1 100 101 10-1 100 101 Period - T (sec) Spectral acc. -Sa (m/s 2 ) UBC-Type 1 UBC-Type 3 2.1m/s2 1.0m/s2 T1 =1.5s Fig. 4. Comparison of the spectral acceleration values of the two types of spectra used for the input motion at the characteristic period of the soil–structure system during shaking. This characteristic period increases to the value of about T1 ¼ 1:5 s from the initial values of 0.6 and 0.8 s. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 655
  • 9. not include the effects of 3D pore-water pressure redis- tribution before and after initial liquefaction. Another analysis of the same site was presented by Elkateb et al. [50]. A 3D stochastic analysis was performed, but the liquefaction potential was assessed separately at each point in space using an empirical method (e.g., [51,52]). Although this approach could not capture any interaction between zones of soil with different liquefaction strengths and different amounts of pore-water pressure build-up, it provided some very interesting conclusions on the effect of loose soil zones on the probability of liquefaction failure. It was found that liquefaction assessments using mean values of soil properties (e.g., from in situ CPT results) are not able to capture the effect of loose soil pockets within the soil mass and consequently are on the unsafe side. Popescu et al. [17] performed MCSs including fully coupled 3D dynamic finite element analyses to investigate the effects of true 3D soil variability on liquefaction of soil deposits. They concluded that 2D plane strain stochastic analyses can provide sufficiently accurate information for liquefaction strength assessment of hor- izontally layered soil deposits, while liquefaction-induced differential settlements can only be captured accurately by a 3D analysis, with correct simulation of pore-water migration. Koutsourelakis et al. [53] and Popescu et al. [16] analyzed the effects of spatial variability on soil liquefaction for a wide range of earthquake intensities using again MCSs, and presented fragility curves for simple structures on liquefiable soil. To illustrate the effects of soil heterogeneity on seismically induced EPWP build-up, some of the results obtained in Ref. [14] for a loose to medium dense soil deposit, with geomechanical properties and spatial variability characteristics estimated based on a series of piezocone test results from a hydraulic fill deposit are reproduced in Fig. 5. The results in Fig. 5b show ARTICLE IN PRESS 20 one - phase finite 480 two - phase finite elements for porous media (4 dof/node) with dimensions: 3m x 0.5m Finite elements analysis set-up free field boundary conditions elements (2 dof/node) 12.0m 1.6m Water table Vertical scale is two times larger than horizontal scale 60.0m (a) (b) (c) Excess pore pressure ratio predited using six samples functions of a stochastic vector field with cross-correlation structure and probability distribution functions estimated from piezocone test results Grey scale range for the excess pore pressure ratio (u/σv0) 0.90...1.00 0.70...0.90 0.50...0.70 0.00...0.50 Sample function # 1 Sample function # 2 Sample function # 3 Sample function # 6 Sample function # 5 Sample function # 4 Excess pore pressure ratio predicted using deterministic input soil parameters Deterministic - average 50 - percentile 60 - percentile 90 - percentile 80 - percentile 70 - percentile (u/σv0 )max = 0.44 ´ ´ Fig. 5. Monte-Carlo simulations of seismically induced liquefaction in a saturated soil deposit, accounting for natural variability of the soil properties (after [14]): (a) finite element analysis setup; (b) contours of EPWP ratio for six sample functions used in the Monte-Carlo simulations; (c) contours of EPWP ratio using deterministic (i.e., uniform) soil parameters and various percentiles of soil strength. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 656
  • 10. computed contours of EPWP ratio with respect to the initial effective vertical stress for six sample functions of a stochastic field representing six possible realizations of soil properties over the analysis domain (see Ref. [14] for more details on the soil properties). Soil liquefaction (EPWP ratio larger than approximately 0.9) was predicted for most sample functions shown in Fig. 5b. Analysis of an assumed uniform soil deposit, with strength characteristics corresponding to the average strength of the soil samples used in MCSs, resulted in no soil liquefaction (the maximum predicted EPWP ratio was 0.44). It can be concluded from these results that both the pattern and the amount of dynamically induced EPWP build-up are strongly affected by the spatial variability of soil properties. For the same average values of the soil parameters, more EPWP build-up was predicted in the stochastic analysis (MCS), accounting for spatial varia- bility, than in the deterministic analysis, assuming uniform soil properties. Popescu et al. [14] suggested that in order to predict more accurate values of EPWP build-up in a deterministic analysis assuming uniform soil properties, one has to use a modified soil strength which is lower than the average strength of the variable soil. The results presented in Fig. 5c illustrate the idea of a ‘‘characteristic percentile’’ of soil strength for dynamic liquefaction analysis of randomly heterogeneous soils. This is a percentile of the field-recorded soil strength values which is lower than the average value. The characteristic percentile is selected in such a way that, when used in a deterministic dynamic analysis assuming uniform soil strength, it results in a computed EPWP ratio equivalent to that obtained from the stochastic analysis accounting for soil spatial variability. For example, for the soil properties and earthquake intensity considered in Ref. [14], the resulting characteristic percentile was somewhere between 70% and 80% (Fig. 5c). A comparison between the results presented in Figs. 5b and c also shows clearly the differences in the predicted liquefaction pattern of natural heterogeneous soil vs. that of assumed uniform soil. The effects of soil spatial variability on structures founded on liquefiable soil deposits are illustrated in Fig. 6 (from Ref. [16]). The structure is the same as Tower 2 presented in the previous example, founded on either heterogeneous soil, or on corresponding uniform soil, with the same strength as the average strength of the variable soil. It is mentioned that the soil properties for uniform soil, as well as the lateral boundary conditions, are different in the study presented in Fig. 6 from those used in Example 1. Fragility curves expressing the exceedance probability of two thresholds in the response, namely foundation settlements larger than 20 cm and base rotations larger than 1/150, are plotted in Figs. 6a and b, respectively. As both the spatial distribution of soil properties and the input seismic motion vary randomly from one sample to another, the curves shown in Fig. 6 reflect uncertainties in both spatial distribution of soil strength and in the seismic ground motion. The fragility curves are plotted for two types of seismic input (Types 1 and 3 discussed in Example 1), and for two assumptions related to soil properties: (1) variable soil, with properties described by a random field with correlation distances yH ¼ 8 m in horizontal direction and yV ¼ 2 m in vertical direction, and a coefficient of variation (CV) ¼ 0.5 (see Ref. [16] for more details) and (2) corresponding uniform soil. For all cases analyzed, the fragility curves calculated for variable soil are shifted to the left as compared to the ones for uniform soil, indicating that, for a given exceedance probability, a significantly lower seismic intensity is needed when soil variability is accounted for than for an assumed uniform soil. For example, from Fig. 6a, for type-1 input acceleration, a 50% exceedance probability of 20 cm foundation settlement is predicted at Arias intensity of the seismic motion IA ¼ 1:7 m=s for uniform soil, and at IA ¼ 1:0 m=s when the soil variability is accounted for. ARTICLE IN PRESS Fig. 6. Comparison between analysis results for variable soil and equivalent uniform soil, for Tower 2 in Example 1 (after [16]): (a) settlements and (b) rotations. The foundation conditions considered in this analysis are different from those discussed in Example 1. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 657
  • 11. 5.2. Mechanisms of liquefaction of heterogeneous soils It was postulated by Popescu et al. [14] that presence of loose pockets would lead to earlier initiation of EPWP and local liquefaction than in a corresponding uniform soil. After that, the pressure gradient between loose and dense zones would lead to water migration, softening, and liquefaction of the dense sand. This assumption that presence of loose pockets is responsible for lower liquefac- tion resistance was theoretically verified by two types of numerical results. First, for a given average value of the soil strength and a given earthquake intensity, the resulting amount of EPWP build-up in a variable soil was increasing with the assumed degree of variability of soil properties, expressed by the CV, which also controls the amount of loose pockets in the soil mass (e.g., [15]). Second, for the same CV of spatial variation, more EPWP was predicted when the soil strength fluctuations followed a symmetrical probability distribution function (PDF)— e.g., truncated Gaussian—than for a positively skewed PDF—e.g., lognormal [54]. For the same mean value, a symmetrical PDF has a fatter left tail (indicating a larger amount of loose pockets) than a positively skewed PDF. Several practical guidelines for assessment of liquefac- tion potential and for design of structures on liquefiable soil with spatially varying properties resulted from the numerical studies discussed before. DYNAFLOW and the multi-yield plasticity constitutive model have been repeatedly validated for liquefaction assessment based on centrifuge experimental results and field observa- tions. All those validation studies were based on determi- nistic analyses, assuming uniform soil properties in horizontal direction. To validate the geotechnical design recommendations for liquefaction of heterogeneous soil mentioned above, the capability of the numerical model to capture the real behavior of spatially variable soil also needs to be verified based on experimental data. Such a verification is performed as part of this numerical example. There are very few experimental studies dealing with liquefaction of spatially variable soil. The two studies mentioned in Section 5.1 were based on undrained cyclic triaxial tests for samples made of different soils (Budiman et al. [46] used sand and gravel, and Konrad and Dubeau [47] used sand and silt). A detailed analysis of the behavior of non-homogeneous soils was performed by Chakrabortty et al. [55] using DYNAFLOW. They reproduced the cyclic undrained triaxial tests on sandwiched samples made of dense sand and silt layers described in Ref. [47], and analyzed the mechanism by which a sample made of two different soils liquefies faster than each of the soils tested separately in uniform samples. The explanation, resulting from a detailed analysis of the numerical results, was that water was squeezed from the more deformable silt layer and injected into the neighboring sand, leading to liquefaction of the dense sand. Regarding liquefaction mechanisms of soil deposits involving the same material, but with spatially variable strength, Ghosh and Madabhushi [19] performed a series of centrifuge experiments to analyze the effects of localized loose patches in a dense sand deposit subjected to seismic loads. They observed that EPWP is first generated in the loose sand patches, and then the water migrates into the neighboring dense sand, reduces the effective stress and softens the dense soil that can liquefy. As discussed before, it is believed that a similar phenomenon is responsible for lower liquefaction resistance of heterogeneous soil than that of the corresponding uniform soil (e.g., [14]). There- fore, this phenomenon is analyzed in details in the following numerical example, based on the backanalysis of Ghosh and Madabhushi’s centrifuge experiments, and on a study for a hypothetical chess-board like heterogeneous soil. This is an original study aimed at: (1) improving understanding of the mechanisms of liquefaction of spatially variable sandy soil deposits and (2) verifying if the nonlinear Biot model implemented in DYNAFLOW is able to capture experimentally observed phenomena including water migration from loose to dense sand zones, with subsequent softening and liquefaction of the entire soil mass. A detailed description of the centrifuge tests analyzed in this example is presented in Ref. [19]. Most of the experimental results are posted on the WWW [56]. The general layout of the tests discussed here—structure on dense sand deposit with included horizontal loose sand layer (test BG4) and included loose sand patch (test BG5)—is sketched in Fig. 7a, along with transducer locations and the finite element mesh used in the back- analyses. The location of the loose sand layer in test BG4 is indicated in Fig. 7a by two thick horizontal lines, and the loose sand patch in test BG5 is shown by the orange rectangular area below the structure in the electronic version (darker area in the paper version). The centrifuge container used in the experiments was an equivalent shear beam (ESB) box, ideally matching the stiffness of a soil column. The vibration and deformability characteristics of the Cambridge centrifuge ESB box are presented in details in Ref. [57]. Those characteristics have been reproduced for plane strain analysis in the finite element model used here. A duxseal padding placed between soil and the lateral walls of the box in the centrifuge model was also included in the numerical model, as shown in Fig. 7a. Duxseal properties, as well as the properties of the ESB box materials (dural and rubber), all modeled as linear-elastic materials, are listed in Table 2. The base input accelerations used in the two tests [56] are presented in Figs. 7b and c. The soil properties at two relative densities used in the centrifuge experiments (Dr ¼ 85% for the dense sand and Dr ¼ 45% for the loose sand) have been estimated based on information provided in Ref. [19] and on the back- analysis of test BG4. Those soil properties have been subsequently checked in a backanalysis of test BG5. A comparison of numerical results obtained in this study with ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 658
  • 12. ARTICLE IN PRESS Fig. 8. Recorded (after [19]) and computed excess pore-water pressure time histories at three locations: below the structure (P2 and P5) and in the free field (P6) (transducer locations are shown in Fig. 7). The calculations are done using the acceleration time histories shown in Figs. 7b and c for tests BG4 and BG5, respectively. The figure is plotted at model scale. Fig. 7. Finite element simulation of tests BG4 and BG5: (a) finite element mesh, (b) input acceleration for test BG4, and (c) input acceleration for test BG5. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 659
  • 13. centrifuge experimental records in terms of EPWP at three locations presented in Fig. 8 shows acceptable match. The soil constitutive parameters are listed in Table 3. To analyze in detail the mechanism of water migra- tion between loose and dense sand observed in the experiments, and to have a better basis for comparison, both centrifuge tests have been numerically simulated with DYNAFLOW using the same base input acceleration, presented in Fig. 7c. Some results in terms of EPWP ratio with respect to the initial effective vertical stress are presented in Fig. 9. Evolution of EPWP ratio at neighboring locations in test BG4 below the structure and in the free field are presented in Figs. 9a and b, respectively. In those figures, P4 and P7 are located in the loose sand layer, while P5 and P6 are located in dense sand. While some pore-water pressure dissipation is predicted to occur after the earthquake in the loose sand, increase in pore-water pressure after the end of ARTICLE IN PRESS Table 3 Parameters of the multi-yield plasticity model used for the saturated soil in Example 2 Constitutive parameter Symbol Dr ¼ 45% Dr ¼ 65% Dr ¼ 85% Mass density—solid (kg/m3 ) rs 2650 2650 2650 Porosity nw 0.45 0.425 0.40 Hydraulic conductivity (m/s) k 1.35 104 1.1 104 8 105 Low strain elastic shear modulus (MPa) G0 12 16 20 Poisson’s ratio n 0.25 0.25 0.25 Power exponent n 0.5 0.5 0.5 Friction angle at failure (deg.) f 39 42.5 45 Maximum deviatoric strain (comp/ext) edev max 0.07/0.04 0.07/0.04 0.07/0.04 Coefficient of lateral stress k0 0.7 0.7 0.7 Stress–strain curve coefficient a 0.182 0.182 0.182 Dilation angle (deg.) c 32 32 32 Dilation parameter Xpp 0.05 0.02 0.006 (c) (a) (b) (d) P7 P6 0 10 20 30 40 50 0 0.2 0.4 0.6 0.8 1 Test BG4 Time (sec) EPWP Ratio 0 10 20 30 40 50 0 0.2 0.4 0.6 0.8 1 Test BG5 Time (sec) EPWP Ratio 0 10 20 30 40 50 0 0.2 0.4 0.6 0.8 1 Test BG4 Time (sec) EPWP Ratio P4 P5 P4 P5 Fig. 9. Computed excess pore-water pressure (EPWP) ratios at pairs of transducers, located one in dense sand and one in loose sand: (a) test BG4, below structure; (b) test BG4, free field; (c) test BG5, below structure; (d) transducer locations. The calculations are done using the acceleration time history in Fig. 7c as input motion for both tests. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 660
  • 14. shaking is simulated at the locations in dense sand, due to migration of water from loose sand. The same trend is observed in test BG5—Fig. 9c shows computed EPWP ratios at two neighboring locations below the structure (P4 in the loose sand patch and P5 in dense sand). Fig. 10 compares computed EPWP ratios at the same location (transducer P6 in dense sand in the free field) for tests BG4 and BG5. An increase in residual EPWP ratio from about 0.2 in test BG5 to about 0.4 in test BG4 is attributed to water migration from the loose sand, present in the free field close to location P6 in test BG4, but not present in the free field in test BG5. The effects of water migration with gradual increase in EPWP in test BG4 as compared to test BG5 can be observed starting at time T ¼ 20 s. Simulation of water migration from loose to dense sand by the numerical model is also illustrated in Fig. 11 by computed time histories of water displacements (relative to the solid phase) at two locations at the border between loose and dense sand in test BG5. To quantify the effects of alternating loose and dense soil patches, as encountered in natural soil deposits, a hypothetical soil–structure system (Fig. 12) is analyzed next. The soil is made of a chess-board like pattern of loose (Dr ¼ 45%) and dense (Dr ¼ 85%) sand zones. It is mentioned that natural soils exhibit a smooth (gradual) variation in relative density between loose and dense zones, while the example analyzed here assumes sudden variation in soil properties. The reason for selecting this chess-board like layout is that sudden changes in soil properties may amplify the effects of spatial variability, leading to easier result analysis and understanding of the underlying mechanisms. To avoid any boundary effects, free field boundary conditions are imposed in the following calculations (corresponding solid and fluid nodal displacements for all pairs of nodes located at the same elevations on the lateral boundaries of the mesh are identical). The input motion used in this example corresponds to the acceleration time history in Fig. 7c divided by a factor of 2 to simulate a small to moderate intensity earthquake. For compa- rison, three other analyses are performed for a structure ARTICLE IN PRESS 0 10 20 30 40 50 -0.1 0 0.1 0.2 0.3 0.4 0.5 Time (sec) EPWP Ratio Test BG4-P6 Test BG5-P6 Fig. 10. Comparison of computed excess pore-water pressure (EPWP) ratios in the free field (dense sand) for tests BG4 and BG5. The calculations are done using the acceleration time history in Fig. 7c as input motion for both tests. 0 10 20 30 40 50 -5 0 5 10 15 Time (sec) Relative water movement [water-solid] (mm) Vertical displacements at Node A Horizontal displacements at Node B Fig. 11. Computed relative movement of water with respect to the solid phase at the interface between loose and dense sand in test BG5. Node locations are shown in Fig. 9d. Positive displacements are upward (vertical) and left to right (horizontal). Fig. 12. Finite element mesh and distribution of loose and dense soil patches for a hypothetical soil–structure system analyzed in Example 2. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 661
  • 15. founded on an assumed uniform soil deposit, for three different relative densities: dense (Dr ¼ 85%) and loose (Dr ¼ 45%), corresponding to the patches in the chess- board example, and medium dense (Dr ¼ 65%), correspon- ding to the average relative density of the heterogeneous soil. Some results in terms of EPWP ratio time histories computed in the free field are presented in Fig. 13. Fig. 13a shows the evolution of the EPWP ratio in two adjacent elements of the chess-board variable soil, one located in a loose soil patch (element 1—shown in Fig. 12) and the other in a dense one (element 2). The effects of water migration are evident here: the dense sand in the chess-board deposit reaches about 95% EPWP ratio, while at the same location in a uniform dense sand deposit the maximum EPWP ratio is about 25% (compare Figs. 13a and c). Figs. 13b and c show computed EPWP ratios in the free field for the chess-board soil deposit and for the other three cases with assumed uniform soil. It is apparent from those results that the excess pressures calculated in the heterogeneous soil (at both loose and dense sand locations) are closer to the ones for uniform loose sand than to those calculated in the uniform medium dense sand. Computed structure settlements shown in Fig. 13d also indicate a significantly softer behavior of the heterogeneous soil than that of the corresponding uniform soil (i.e., Dr ¼ 65%). Color Plates 3 and 4 present contours of EPWP ratio at time T ¼ 20 s and of maximum shear strains at the end of analysis, respectively. From both color plates it is obvious that the heterogeneous soil response (Color Plates 3a and 4a) is much closer to the response of the uniform loose sand (Color Plates 3c and 4c) than to that of the uniform medium dense sand (Color Plates 3b and 4b). 6. Summary and conclusions The theory of dynamics of saturated porous media, its extension into the nonlinear regime, and its implementa- tion in the DYNAFLOW code have been presented, along with numerical applications illustrating the capabilities of the numerical model to deal with complex phenomena, such as seismic behavior of structures on liquefiable soil deposits. The numerical results are validated through comparisons with centrifuge experimental data. The frequency content of seismic motion has a major role in predicted structural response. In the first numerical example, it is shown how the characteristic frequency of the soil–structure system is reduced to values close to the dominant frequency range of type-3 inputs, due to soil softening after build-up of pore pressures. Consequently, considerably more damage is predicted for this type of seismic motion. The results of this study also point out to important attenuation of the earthquake motion when ARTICLE IN PRESS 0 10 20 30 40 50 0.25 0.2 0.15 0.1 0.05 0 Time (sec) Settlement (m) Dense sand Medium dense sand Chess-board Loose sand 0 10 20 30 40 50 -0.2 0 0.2 0.4 0.6 0.8 1 Element 1 Time (sec) EPWP Ratio Loose sand Chess-board Medium dense sand Dense sand (b) (a) (c) (d) Settlement at base of structure 0 10 20 30 40 50 -0.2 0 0.2 0.4 0.6 0.8 1 Time (sec) EPWP Ratio Element 2 (Dense sand) Element 1 (Loose sand) 0 10 20 30 40 50 -0.2 0 0.2 0.4 0.6 0.8 1 Element 2 Time (sec) EPWP Ratio Medium dense sand Chess-board Loose sand Dense sand Fig. 13. (a) EPWP ratio time histories for the chess-board type heterogeneous soil computed at two adjacent locations in the free field (one in the loose soil—element 1, and the other in the dense soil—element 2); (b) EPWP ratio in the free field in element 1; (c) EPWP ratio in the free field in element 2; (d) computed settlements of the structure. The four curves in (b)–(d) correspond to results from four different analyses defined in the text. Element locations are shown in Fig. 12. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 662
  • 16. seismic waves pass through soft soil after considerable increase in pore-water pressure. The reduction in specific seismic energy was found to range between zero, for small earthquake intensities, to about 75%, for large earthquake intensities (IA about 2 m/s and PGA about 0.4 g). As expected, the numerical model predicts higher attenuation for the high frequency components of the base motion. In the second numerical example, it was mentioned that soil heterogeneity leads to more EPWP build-up than that predicted for corresponding uniform soil and, for structures founded on liquefiable soils, to reduction in average mobilized soil strength due to deviation of the failure surface to pass through looser soil zones. Those effects result in larger predicted damage when accounting for soil spatial variability, as compared to assuming a corresponding uniform soil. The mechanism of liquefaction of heterogeneous soils was also studied in this example through backanalysis of centrifuge tests including loose soil zones in a dense sand deposit. From the experimental results, it was concluded that the pore pressure gradient created from the initial liquef- action of loose zones leads to migration of water in the adjacent dense zones and makes the dense soil liable to liquefaction. Several numerical examples were presented that confirmed the capabilities of the numerical model to reproduce those experimentally observed phenomena and validated previous research results and design recom- mendations for liquefaction of stochastically spatially variable soils. ARTICLE IN PRESS Plate 4. (Example 2) Computed maximum shear strain contours and deformed meshes at the end of analysis (time T ¼ 50 s) for four different soil structure systems, defined as: (a) chess-board distributed loose and dense patches; (b) uniform medium dense soil (Dr ¼ 65%); (c) uniform loose soil (Dr ¼ 45%); (d) uniform dense soil (Dr ¼ 85%). Plate 3. (Example 2) Computed EPWP ratio contours at time T ¼ 20 s for four different soil structure systems, defined as: (a) chess-board distributed loose and dense patches; (b) uniform medium dense soil (Dr ¼ 65%); (c) uniform loose soil (Dr ¼ 45%); (d) uniform dense soil (Dr ¼ 85%). Plate 1. (Example 1) EPWP ratio contours for Tower 1, seismic motion type-3, sample #51 (IA ¼ 1:12 m=s). Deformation magnification factor ¼ 5. Plate 2. (Example 1) Tower 1, seismic motion type-3, sample #51 at time T ¼ 5 s: (a) contours of maximum shear strains and deformed mesh and (b) contours of EPWP ratio. R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 663
  • 17. Acknowledgments The financial support provided by NSERC under Research Grant No. RG203795-02 and by NSF under Grant No. 0075998 is gratefully acknowledged. The authors are also indebted to Dr. B. Ghosh for providing some of the results of centrifuge experiments analyzed in Example 2. References [1] Terzaghi K. Theoretical soil mechanics. New York: Wiley; 1943. [2] Biot MA. Theory of elasticity and consolidation for a porous anisotropic solid. J Appl Phys 1955;26:182–5. [3] Biot MA. Theory of propagation of elastic waves in a fluid-saturated porous solid, I. Low frequency range. J Acoust Soc Am 1956;28: 168–91. [4] Biot MA, Willis DG. The elastic coefficients of a theory of consolidation. J Appl Mech ASME 1957;29:594–601. [5] Simon BR, Wu JSS, Zienkiewicz OC, Paul DK. Evaluation of u–w and u–p finite element methods for the dynamic response of saturated porous media using one-dimensional models. Int J Numer Anal Meth Geomech 1986(10):461–82. [6] Simon BR, Wu JSS, Zienkiewicz OC. Evaluation of higher order, mixed and hermitean finite element procedures for dynamic analysis of saturated porous media using one-dimensional models. Int J Numer Anal Meth Geomech 1986(10):483–99. [7] Mesgouez A, Lefeuve-Mesgouez G, Chambarel A. Transient mechanical wave propagation in semi-infinite porous media using a finite element approach. Soil Dyn Earthquake Eng 2005;25(6): 421–30. [8] Prévost JH. Mechanics of continuous porous media. Int J Eng Sci 1980;18(5):787–800. [9] Green AE, Naghdi PM. A dynamical theory of interacting continua. Int J Eng Sci 1965;3:231–41. [10] Eringen AC, Ingram JD. A continuum theory of chemically reacting media I. Int J Eng Sci 1965;3:197–213. [11] Bowen RM. Compressible porous media models by use of the theory of mixtures. Int J Eng Sci 1982;20(6):697–735. [12] Prevost JH. DYNAFLOW—a nonlinear transient finite element analysis program, Version 02. Technical Report, Department of Civil and Environmental Engineering, Princeton University, Princeton, NJ, 2002. /http://www.princeton.edu/dynaflow/S. [13] Popescu R. Influence of seismic loading rate on soil liquefaction. Can Geotech J 2002;39:331–44. [14] Popescu R, Prevost JH, Deodatis G. Effects of spatial variability on soil liquefaction: some design recommendations. Geotechnique 1997;47(5):1019–36. [15] Popescu R, Prevost JH, Deodatis G. Characteristic percentile of soil strength for dynamic analyses. Geotech Earthquake Eng Soil Dyn, Geotech Spec Publ 75, ASCE 1998:1461–71. [16] Popescu R, Chakrabortty P, Prevost JH. Fragility curves for tower structures on stochastically variable soil. In: Proceedings of the international conference on structural safety and reliability, ICOSSAR05, Rome, Italy, June 2005 [on CD-ROM]. [17] Popescu R, Prevost JH, Deodatis G. 3D effects in seismic liquefaction of stochastically variable soil deposits. Geotechnique 2005;55(1): 21–31. [18] Madabhushi SPG, Schofield AN. Centrifuge modeling of tower structures on saturated sands subjected to earthquake perturbations. Geotechnique 1993;43(4):555–65. [19] Ghosh B, Madabhushi SPG. Effects of localized soil inhomogeneity in modifying seismic soil–structure interaction. In: Proceedings of the 16th ASCE engineering mechanics conference, Seattle, WA, July 2003. [20] Prévost JH. Nonlinear transient phenomena in saturated porous media. Comput Meth Appl Mech Eng 1982;30:3–18. [21] Prévost JH. Wave propagation in fluid-saturated porous media: an efficient finite element procedure. Soil Dyn Earthquake Eng 1985;4(4):183–202. [22] Bathe KJ, Wilson EL. Numerical methods in finite element analysis. Englewood Cliffs, NJ: Prentice-Hall; 1976. [23] Zienkiewicz OC. The finite element method. 3rd ed. London, England: McGraw-Hill; 1977. [24] Hughes TJR, Liu WK. Implicit–explicit finite elements in nonlinear transient analysis. Comput Meth Appl Mech Eng ASME 1978;45: 371–4. [25] Hughes TJR, Liu WK. Implicit–explicit finite elements in nonlinear transient analysis: implementation and numerical examples. J Appl Mech ASME 1978;45:395–8. [26] Castro G, Poulos J. Factors affecting liquefaction and cyclic mobility. J Geotech Eng Div ASCE 1977;103(GT6):501–16. [27] Marcuson III WF. Definition of terms related to liquefaction. J Geotech Eng Div ASCE 1978;104(9):1197–200. [28] Youd TL, et al. Liquefaction resistance of soils: summary report from the 1996 NCEER and 1998 NCEER/NSF workshops on evaluation of liquefaction resistance of soils. J Geotech Geoenviron Eng ASCE 2001;127(10):817–33. [29] Prevost JH, Popescu R. Constitutive relations for soil materials. Electron J Geotech Eng 1996;1 http://www.ejge.com/1996/Ppr9609/ Ppr9609.htm. [30] Prevost JH. A simple plasticity theory for frictional cohesionless soils. Soil Dyn Earthquake Eng 1985;4(1):9–17. [31] Popescu R, Prevost JH. Centrifuge validation of a numerical model for dynamic soil liquefaction. Soil Dyn Earthquake Eng 1993;12(2):73–90. [32] Popescu R. Stochastic variability of soil properties: data analysis, digital simulation, effects on system behavior. PhD thesis, Princeton University, Princeton, NJ, 1995. [33] Popescu R, Prevost JH. Comparison between VELACS numerical ‘class A’ predictions and centrifuge experimental soil test results. Soil Dyn Earthquake Eng 1995;14(2):79–92. [34] Jafari-Mehrabadi A, Popescu R. Mitigative measures against soil liquefaction: numerical modeling and centrifuge experiments. Can Geotech J, in review. [35] Keanne CM, Prevost JH. An analysis of earthquake data observed at the Wildlife Liquefaction Array Site, Imperial County, California. In: Proceedings of the second US–Japan workshop on liquefaction, large ground deformations and effects on lifelines, New York, 1989. p. 39–53. [36] Morris DV. The centrifugal modeling of dynamic soil–structure interaction and earthquake behavior. PhD thesis, University of Cambridge, 1979. [37] Weissman K, Prevost JH. Centrifugal modeling of dynamic soil–structure interaction. Technical Report NCEER-89-0040, Prin- ceton University, 1989. [38] Deodatis G. Non-stationary stochastic vector processes: seismic ground motion applications. Probab Eng Mech 1996;11:149–68. [39] Uniform Building Code. International conference of building officials (ICBO), Whittier, CA, vol. 2, 1994. [40] Arias A. A measure of earthquake intensity. In: Hansen RJ, editor. Seismic design for nuclear power plants. Cambridge, MA: MIT Press; 1970. p. 438–83. [41] Iwasaki Y, Tai M. Strong motion records at Kobe Port Island. Soils and Foundations, Special Issue on Geotechnical Aspects of the January 17 1995 Hyogoken-Nambu Earthquake, January 1996. p. 29–40. [42] Shinozuka M, Feng MQ, Lee J, Naganuma T. Statistical analysis of fragility curves. J Eng Mech ASCE 2000;126(12):1224–31. [43] Deodatis G, Saxena V, Shinozuka M. Effect of spatial variability of ground motion on bridge fragility curves. In: Proceedings of the eighth ASCE specialty conference on probabilistic mechanics and structural reliability, University of Notre Dame, IL, 2000 [on CD-ROM]. ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 664
  • 18. [44] Focht JA, Focht III JA. Factor of safety and reliability in geotechnical engineering (discussion). J Geotech Geoenviron Eng ASCE 2001;127(8):704–6. [45] Popescu R, Deodatis G, Nobahar A. Effects of soil heterogeneity on bearing capacity—qualitative assessment. Probab Eng Mech 2005;20: 324–41. [46] Budiman JS, Mohammadi J, Bandi S. Effect of large inclusions on liquefaction of sand. In: Proceedings of the conference of geotechni- cal engineering division. Geotechnical Special Publication 56, ASCE; 1995. p. 48–63. [47] Konrad J-M, Dubeau S. Cyclic strength of stratified soil samples. In: Proceedings of the 55th Canadian geotechnical conference: ground and water: theory to practice, Niagara Falls, Ont., October 2002. p. 89–94. [48] Ohtomo K, Shinozuka M. Two-dimensional spatial severity of liquefaction. In: Proceedings of the eighth Japan earthquake engineering symposium, Tokyo, 1990. [49] Fenton GA, Vanmarcke EH. Spatial variation in liquefaction risk. Geotechnique 1998;48(6):819–31. [50] Elkateb T, Chalaturnyk R, Robertson PK. Simplified geostatistical analysis of earthquake-induced ground response at the Wildlife Site, California, USA. Can Geotech J 2003;40:16–35. [51] Seed HB, Idriss IM. Simplified procedure for evaluating soil liquefaction potential. J Geotech Eng Div ASCE 1971;97(9):1249–73. [52] Robertson PK, Wride CE. Evaluating cyclic liquefaction potential using the cone penetration test. Can Geotech J 1998;35(3):442–59. [53] Koutsourelakis S, Prevost JH, Deodatis G. Risk assessment of an interacting structure–soil system due to liquefaction. Earthquake Eng Struct Dyn 2002;31:851–79. [54] Popescu R, Prevost JH, Deodatis G. Influence of spatial variability of soil properties on seismically induced soil liquefaction. In: Proceed- ings of the uncertainty in the geologic environment: from theory to practice. Madison, WI: ASCE; 1996. p. 1098–112. [55] Chakrabortty P, Jafari-Mehrabadi A, Popescu R. Effects of low permeability soil layers on seismic stability of submarine slopes. In: Proceedings of the 57th Canadian geotechnical conference, Quebec City, PQ, 2004. [56] Geotechnical and Environmental Research Group. Behaviour of rigid foundations on layered soil, Report #331, Cambridge Uni- versity, 2004. /http://www-civ.eng.cam.ac.uk/geotech_new/S. [57] Zeng X, Schofield AN. Design and performance of an equivalent shear beam container for earthquake centrifuge modeling. Geotech- nique 1996;46(1):83–102. ARTICLE IN PRESS R. Popescu et al. / Soil Dynamics and Earthquake Engineering 26 (2006) 648–665 665