Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Bounding surface plasticity model for the seismic liquefaction


analysis of geostructures
Konstantinos I. Andrianopoulos a,, Achilleas G. Papadimitriou b, George D. Bouckovalas a
a
Department of Geotechnical Engineering, School of Civil Engineering, National Technical University of Athens, Greece
b
Department of Civil Engineering, University of Thessaly, Greece

a r t i c l e in fo abstract

Article history: This paper presents the constitutive relations and the simulative potential of a new plasticity model
Received 16 October 2009 developed mainly for the seismic liquefaction analysis of geostructures. The model incorporates the
Received in revised form framework of critical state soil mechanics, while it relies on bounding surface plasticity with a vanished
30 March 2010
elastic region to simulate the non-linear soil response. Key constitutive ingredients of the new model
Accepted 1 April 2010
are: (a) the inter-dependence of the critical state, the bounding and the dilatancy (open cone) surfaces
on the basis of the state parameter c, (b) a (Ramberg–Osgood type) non-linear hysteretic formulation
for the ‘‘elastic’’ strain rate, (c) a discontinuously relocatable stress projection center related to the
‘‘last’’ load reversal point, which is used for mapping the current stress point on model surfaces and as a
reference point for introducing non-linearity in the ‘‘elastic’’ strain rate and finally (d) an empirical
index of the directional effect of sand fabric evolution during shearing, which scales the plastic
modulus. In addition, the paper outlines the calibration procedure for the model constants, and exhibits
its accuracy on the basis of a large number of laboratory element tests on Nevada sand. More
importantly, the paper explores the potential of the new model by presenting simulations of the
VELACS centrifuge tests of Models No 1 and 12, which refer to the free-field liquefaction response of
Nevada sand and the seismic response of a rigid foundation on the same sand, respectively. These
simulations show that the new model can be used successfully for the analysis of widely different
boundary value problems involving earthquake soil liquefaction, with the same set of model constants
calibrated on the basis of laboratory element tests.
& 2010 Elsevier Ltd. All rights reserved.

1. Introduction go beyond elastoplasticity, like the hypoplasticity (e.g. [24]) and


hyperplasticity models (e.g. [25]).
Over the years, a number of constitutive models have been Although pertinent models present the foregoing differences in
developed to simulate the cyclic loading of non-cohesive soils, their fundamental characteristics, the most recent ones usually share
with emphasis on cyclic mobility and/or flow liquefaction, which constitutive ingredients that have offered improved simulations.
may be divided into several categories depending on their Specifically, since the mechanical response of sand is characterized by
fundamental characteristics, such as multi-surface models (e.g. infinite normal consolidation (NC-) lines pending on the initial
[1–4]), two-surface models (e.g. [5–10]), bounding surface models conditions of void ratio and confining pressure, the critical state
(e.g. [11–14]) and generalized plasticity models (e.g. [15–18]). constitutive models that retained the unique NCL idealization from
Furthermore, there are models that build on classical elastoplas- clay response had difficulties in producing quantitatively accurate
ticity by complementing it with key constitutive concepts, like the simulations for all initial conditions. Hence, the introduction of the
multi-laminate concept (e.g. [19]), an endochronic densification state parameter c of Been and Jefferies [26] in constitutive equations,
law (e.g. [20]), the middle surface concept (e.g. [21]), the either implicitly (by Jefferies [27] for monotonic loading) or explicitly
disturbed state concept (e.g. [22]) and the multiple-mechanism (by Manzari and Dafalias [6] for both monotonic and cyclic response),
concept (e.g. [23]). In addition to the above, there are models that allowed the use of a single set of model constants for successful
simulations for any initial void ratio or confining pressure. This
concept has been implemented in many constitutive models there-
 Correspondence to: Mourgkanas 6, Maroussi, 15126 Athens, Greece.
after, regardless of their fundamental characteristics (e.g. the
bounding surface model of Li et al. [13], the two-surface models of
Tel.: +30 210 8069604; fax: + 30 210 6128460.
E-mail addresses: kandrian@tee.gr (K.I. Andrianopoulos), apapad@civ.uth.gr Papadimitriou and Bouckovalas [9], Dafalias and Manzari [28] and the
(A.G. Papadimitriou), gbouck@central.ntua.gr (G.D. Bouckovalas). generalized plasticity model of Ling and Yang [18]).

0267-7261/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.soildyn.2010.04.001
896 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

In addition, various functions of integrals of strain histories simulations for these two boundary value problems were
during the shearing phase have appeared in the literature as obtained after successful implementation of the new model to
scalar multipliers of the plastic modulus (e.g. in the generalized the explicit finite difference code FLAC [35], as detailed in
plasticity model of Pastor et al. [15]). Such a plastic modulus Andrianopoulos [36] and Andrianopoulos et al. [37,38].
multiplier was first explicitly related to evolving sand fabric by In the following, the constitutive equations of the new model
Papadimitriou [29] and Papadimitriou et al. [30] in their are always presented with reference to the antecedent model, so
integrated approach to account, not only for the densifying effect that the reader can clearly identify the similarities and differ-
related to contractive phases of shearing, but also for the opposite entiations between the two models. In terms of notation, tensors
effects related to dilation. In parallel, focusing merely on sand are written in bold face characters so that they can be easily
fabric evolution due to the dilation, Dafalias and Manzari [31] distinguished from scalars, while all presented stress quantities
proposed a scalar multiplier of the dilatancy aiming primarily on are effective. Symbol ‘‘:’’ denotes the double inner product of two
accurately simulating the cyclic mobility phase of shearing. This second-order tensors.
constitutive concept related to evolving sand fabric has continued
to appear in the literature, in different forms of scalar multipliers
of either the plastic modulus (e.g. in generalized plasticity models 2. Constitutive equations
[17,18], in two-surface models [8,9]), or of the dilatancy (e.g. in
multi-surface models [3,4], in two-surface models [28]), or as an 2.1. Model platform
endochronic densification law [20].
Given the foregoing recent advances, most pertinent constitu- As mentioned above, the proposed model is an improvement
tive models have been able to simulate qualitatively the of the antecedent two-surface model of Papadimitriou et al. [8]
hysteretic cyclic response of sands in small to medium cyclic and Papadimitriou and Bouckovalas [9], which was in turn based
shear strains (see thresholds of cyclic shear strain amplitudes in on the model of Manzari and Dafalias [6]. Therefore, its basic
[32]) and have demonstrated the potential to reproduce the equations need not be presented in full detail and are therefore
excess pore pressure buildup in undrained loading under large only outlined in Table 1. In this manner, the paper focuses on the
cyclic shear strains, as well as the well-known ‘‘butterfly’’ shaped alterations and enhancements presented in paragraphs 2.2
loops in the effective stress path that are related to cyclic through 2.6.
mobility. Nevertheless, to our knowledge, quantitative accuracy In particular, the antecedent model had adopted a four (open-
of the response for all cyclic strain levels with the same set of cone) surface formulation (yield, critical state, bounding, dila-
values of model constants has only been demonstrated by tancy) and the inter-dependency of the last three surfaces on the
Papadimitriou et al. [8] and Papadimitriou and Bouckovalas [9], basis of the state parameter c, from Manzari and Dafalias [6]. In
who adopted the constitutive two-surface model platform of the new model, the ‘‘small’’ conical, purely kinematically hard-
Manzari and Dafalias [6], and introduced two key constitutive ening, yield surface has been reduced to a straight line, and
elements: (a) a Ramberg–Osgood type non-linear hysteretic therefore appears as a point on the p-plane of the deviatoric stress
formulation of the ‘‘elastic’’ moduli, that governs shear modulus ratio r space instead of a constant radius circle. This choice was
degradation and hysteretic damping increase for small to medium not made to increase its predictive accuracy (the elastic region
cyclic shear strains, and (b) a scalar multiplier of the plastic was ‘‘small’’ in the antecedent model), but to reduce the
modulus, that builds on the proposals of [29,30] and [31] numerical problems created by its existence during its imple-
mentioned above, and governs the response at medium to large mentation to FLAC. More specifically, in this way, it was made
cyclic shear strains. possible to satisfactorily overcome a number of stability related
The new model presented herein is an improvement of the problems that significantly increase the required computational
foregoing (antecedent) two-surface model of Papadimitriou et al. effort with the explicit stress integration scheme (Sloan [39],
[8] and Papadimitriou and Bouckovalas [9]. As such, it retains the Sloan et al. [40]) used, namely (a) the estimation of a stress path
constitutive merits described above, while it introduces certain crossing point of the yield surface (when going from an elastic to a
alterations that enhance its accuracy, but mainly facilitate its plastic state), and (b) the drift correction resulting from the weak
implementation to numerical codes. Namely, the new model enforcement of the consistency condition, and the subsequent
reduces the ‘‘small’’ conical yield surface to a straight line necessary sub-stepping in the stress integration scheme. The need
emanating from the stress origin (vanished elastic region: a for adopting a zero elastic range was further underlined by the
concept first proposed by Dafalias [33], Dafalias and Popov [34] mixed discretization algorithm [41] adopted in FLAC in order to
for metals), and thus reduces its numerical complexity. In turn, overcome the problem of hourglassing, which may occur with
this change dictates revision of key constitutive ingredients, i.e. constant-strain finite difference quadrilaterals. In more detail,
the definition of the loading direction and the mapping rule, since with the foregoing computer code, the continuum is divided into a
these were based on a yield surface that no longer exists, as well finite difference mesh composed of quadrilateral elements (or
as a revision of the interpolation rule. ‘‘zones’’ in FLAC terminology). Each element is automatically
In particular, this paper presents the constitutive relations of subdivided into two overlaid sets of constant-strain triangular
the new model, outlines the calibration procedure and includes a sub-zones, and stress integration is performed separately for each
comparison of simulations to data from the same large number of of the four sub-zones of the element. However, the isotropic
element tests on Nevada sand (capturing the full behavioral effective stress (p) and strain components (ep) are taken to be
pattern of cyclic loading) that were also used for the verification uniform over the whole quadrilateral element and equal to their
of the antecedent model. Furthermore, unlike the antecedent average value over the four triangular sub-zones, while the
model whose accuracy was ascertained only at the element test deviatoric components of stress s and strain e are treated
level, the accuracy of the new model is explored for two (2) separately for each triangular sub-zone. This averaging procedure
widely different boundary value problems involving earthquake- of p and ep is inherent in FLAC and is performed after the end of
induced liquefaction, namely the VELACS centrifuge tests of each applied strain increment e_ , leading this way to a significant
Models No 1 and 12, which refer to the free-field response of drift error and thus making the enforcement of the consistency
liquefied Nevada sand and the seismic response of a rigid condition at each solution step even more cumbersome numeri-
foundation on the same sand, respectively. The numerical cally. The overall computational cost becomes disproportionally
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 897

Table 1
Equations of constitutive model platform.

Description Form of equation Constants References/notes

Deviatoric stress ratio on bounding surface in triaxial Mcb ¼ Mcc þ kbc /cS ð1Þ Mcc ,kbc Manzari and Dafalias [6], with o 4
compression, as a function of the state parameter c being Macauley brackets, where
(Been and Jefferies [26]) o c 4 ¼0.5(|c|+ c)
Deviatoric stress ratio on dilatancy surface in triaxial Mcd ¼ Mcc þ kdc c ð2Þ kdc Manzari and Dafalias [6]
compression
Critical State Line location in [e, p] – definition of c c ¼ eecs ¼ eGcs þ l lnðpÞ ð3Þ Gcs, l Manzari and Dafalias [6]

c,b,d
Multiaxial generalization of model surfaces on the basis My ¼ gðy,cÞMcc,b,d ð4Þ c ¼ Mec =Mcc
of the Lode angle y and c
 
Shape of model surfaces on the p-plane of the 4c ð1 þ cÞ þ ð1cÞcosð3yÞ Papadimitriou and Bouckovalas [9]
gðy,cÞ ¼  ð5Þ
deviatoric stress ratio space ð1þ cÞð1cÞcosð3yÞ 2

pffiffiffi
Estimation of Lode angle y, on the basis of the conjugate 3 ðrbi : rbi : rbi Þ Li [14]
b cosð3yÞ ¼ 3=2 ð6Þ
(image) deviatoric stress ratio tensor ri on the 2
ð0:5ðrbi : rbi Þ
bounding surface
!  
‘‘Elastic’’ strain rate estimation, on the basis of the e_ ep s_ p_ Kt, Gt are the bulk and shear ‘‘elastic’’
e_ e ¼ e_ e þ I¼ þ I ð7Þ
effective stress increment r_ ¼ s_ þ pI
_ 3 2Gt Kt moduli (see 2.4)

!
Plastic strain rate e_ pp L (see Eq. (9)), R (see Eq. (11))
e_ p ¼ e_ p þ I ¼ /LSR ð8Þ
3

Loading index, in terms of the effective stress r_ L:r_ Kp is the plastic modulus (see 2.5)
L¼ ð9Þ
increment Kp

Loading direction n:r n is a unit deviatoric stress ratio


L ¼ n I ð10Þ
3 tensor (see 2.2)

Plastic strain rate direction (non-associated flow rule) D D is the dilatancy function (see
R ¼ nþ I ð11Þ
3 section 2.3)

Estimation of effective stress r_ increment, on the basis r_ ¼ 2Gt e_ þ Kt e_ p I/LSð2Gt n þ Kt DIÞ ð12Þ On the basis of Eqs. (7)–(8) and (11)
of the applied strain increment e_ ¼ e_ þ ðe_ p =3ÞI
Loading index, in terms of the applied strain increment L ¼ 2Gt n:eðn _ : rÞKt e_ p The sign of L is used for definition of
ð13Þ
Kp þ 2Gt ðn : rÞKt D load reversal (see 2.6)

large given the fact that FLAC is an explicit code and thus [28] and [43], or more elaborate critical state line equations in Eq.
necessarily uses small time steps irrespective of the constitutive (3), like the 3-constant equation of [13]. However, the proposed
model at hand. Based on everything mentioned above it becomes model retains the original form of the equations for simplicity and
obvious, that the choice of a vanished elastic region is practically a consistency with the antecedent model.
‘‘Gordian knot’’ type of solution of all aforementioned numerical The multiaxial generalization of these surfaces is shown
problems, without loss of simulative accuracy. graphically in Fig. 1, that presents their shape in the p-plane
Hence, the proposed model has three open cone-type surfaces (perpendicular to the hydrostatic p axis) of the deviatoric stress-
with apex at the origin of stress space: (i) the critical state surface ratio r space, where r ¼s/p, with s¼ r  pI being the deviatoric
at which deviatoric deformation develops for constant stresses stress tensor (r and I are the effective stress and the identity
and no volume change (constant void ratio e), (ii) the bounding second-order tensors). As shown in this figure, the multiaxial
surface which locates the peak deviatoric stress ratio states and generalization is based on a Lode angle y-dependency of all
(iii) the dilatancy surface which dictates the sign of the plastic surfaces, expressed via Eqs. (4) and (5), where Myc,b,d define the
volumetric strain increment during loading. It is noted that when apertures of the three surfaces, respectively, along radii related to
the stress point is within the dilatancy surface the sand behavior Lode angle y (see Fig. 1), on the basis of their respective apertures
is contractive, while the opposite occurs outside the dilatancy for triaxial compression Mcc,b,d , and model constant c ¼ Mec =Mcc
surface, and hence this surface is practically the stress space ( o1). As depicted in Fig. 1, the Lode angle y is defined in terms of
generalization of the phase transformation line of Ishihara et al. the conjugate (image) deviatoric stress ratio tensor rbi on the
[42]. Furthermore, it should be clarified that the proposed and the bounding surface according to Eq. (6), whose estimation
antecedent models adopt the concept of Manzari and Dafalias [6] procedure (mapping rule) will be fully explained in paragraph 2.2.
who allowed the stress point to marginally cross the bounding The general form of constitutive equations is that of classical
surface in order to simulate strain softening, and in this sense, the elasto-plasticity, i.e. there is a decomposition of the strain rate e_
term ‘bounding’ surface for the proposed model is used in the into an elastic e_ e and a plastic e_ p component. Considering that the
broader sense. The shapes of these model surfaces for triaxial former strain rate is not fully reversible, and that there is no
compression are uniquely defined by the deviatoric stress ratios elastic region in this model, the term ‘‘elastic’’ is presented in
Mcc , Mcb and Mcd (collectively Mcc,b,d ), which are interrelated via the quotes throughout the paper. In particular, the ‘‘elastic’’ and
state parameter c, via Eqs. (1) through (3) in Table 1. Alternatives plastic strain rates are estimated as per Eqs. (7) and (8) in Table 1,
to the forms, but not to the concepts, of these equations have been which also includes all equations that formulate the constitutive
proposed in the literature. For example, smoother exponential model platform (Eq. (9) through (13)). What remains to be defined
functions of c for Mcb and Mcd can be used in Eqs. (1)–(2), as per are the forms given to the unit deviatoric stress ratio n (on the
898 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

basis of the mapping rule, see Section 2.2), the dilatancy function rc,i and rd,i were always estimated by adding the quantity mn to
D (on the basis of the flow rule, see Section 2.3), the non-linear the deviatoric back-stress ratios abi, aci and adi, respectively.
hysteretic ‘‘elastic’’ moduli Kt and Gt (see Section 2.4), the plastic Based on the above, the importance of n was crucial for the
modulus Kp (on the basis of the interpolation rule, see Section 2.5) whole constitutive platform. In addition, basing its estimation on
and the triggering mechanism of load reversal (see Section 2.6). the gradient of the yield surface L provided for stability, since the n
was not dependent on the currently applied loading increment. On
2.2. Mapping rule the contrary, the n was made to depend on the recent loading
history, as this is expressed via the location of the yield surface a
The existence of a ‘‘small’’ conical, purely kinematically relative to the stress point r. Moreover, adopting a circular shape
hardening yield surface in the antecedent model provided the for the yield surface in the p-plane retained the definition of L
means for defining the loading direction L as the gradient to the (and of n) simple, as opposed to adopting a more realistic yield
yield surface. Given that the yield surface had a circular shape in surface shape with third stress invariant dependency (e.g. as in the
the p-plane of the deviatoric stress ratio r space in that model (of two-surface model of [7]) having a small overall gain in accuracy.
radius m, as proposed by Manzari and Dafalias [6]), the deviatoric Reducing the ‘‘small’’ conical, purely kinematically hardening
part of the loading direction L, i.e. n in Eq. (10), was defined along yield surface to a straight line originating from the stress origin
the r–a direction, where a was the kinematically hardening created the need for explicitly defining the n direction and also for
deviatoric back-stress ratio that located the center of the yield introducing a new mapping rule of the current deviatoric stress
surface in the foregoing p-plane (Fig. 2a). The corresponding ratio r on the model surfaces. Furthermore, in order to maintain
direction of n was then used for defining the deviatoric part of the the good performance of the antecedent two-surface model in its
plastic strain rate direction R (Eq. (11)) and also for mapping the zero-elastic range variant presented herein, the two constitutive
location the current stress point on model surfaces. Strictly attributes would have to remain interrelated. In addition, the
speaking, the bounding, critical and dilatancy surfaces in the two- definition of the unit-deviatoric loading direction n should remain
surface model were defined in terms of deviatoric back-stress stable, dependent on recent loading history and simple, as in the
ratios ab,i, ac,i and ad,i, respectively, which mapped the location of two-surface model.
a and not of r (see Fig. 2a). Nevertheless, since r¼ a + mn, the Thus, it was first chosen to base the definition of n on the
corresponding conjugate (image) deviatoric stress ratio points rb,i, location of the conjugate (image) deviatoric stress ratio point on
the bounding surface rbi, and define the n direction as

r bi
n ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffi ð14Þ
r bi : r bi

In the sequel, attempting to find an appropriate mapping rule,


Andrianopoulos et al. [44] studied various schemes (Figs. 1,
2b, 2c) that could potentially be used for cyclic loading conditions,
and compared its predictions with those of a reference two-
surface model similar to [6] and whose mapping rule is depicted
in Fig. 2a. Based on their conclusions, locating rbi by using the
e
elastic prediction of the deviatoric stress ratio rate r_ (Fig. 2b) is
simple and reasonable, but fails in the stability criterion, since
small loading perturbations (common in numerical analyses) will
continuously relocate the rbi. Alternatively, a mapping rule was
considered (Fig. 2c) that is radial on the p-plane of the deviatoric
stress ratio space and, hence the rbi is collinear with r or –r,
depending on whether the r : r_ e scalar product is positive
Fig. 1. Model surfaces and adopted mapping rule in the p-plane of the deviatoric (loading) or negative (reverse loading), respectively. This rule is
stress ratio space, based on a relocatable projection center (based on [12]). simple, introduces the useful concept of reversal of loading, which

Fig. 2. Model surfaces and various alternative mapping rules in the p-plane of the deviatoric stress ratio space: (a) based on the yield surface location [6,9], (b) based on the
elastic prediction of deviatoric stress ratio rate, (c) radial, but taking into account the elastic prediction of deviatoric stress ratio rate.
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 899

discontinuously relocates the rbi point, nevertheless it yields p-plane of the deviatoric stress ratio space is equal to n. Therefore,
non-realistic predictions for non-proportional loadings, since the the deviatoric component of R is not related to the direction of the
rbi continuously fails to be in the direction of loading, with the gradient to the bounding surface at the location of the conjugate
exception of proportional paths like the triaxial, for which it yields (image) point rbi, but merely to the direction of the rbi point itself
identical results with the two-surface formulation. (Eq. (14), Fig. 1). As shown by Dafalias and Manzari [28], this
Building on the foregoing study, the proposed model finally simplification leads to less accurate predictions for monotonic
adopts the mapping rule of Fig. 1 that was first proposed by Wang loading. Nevertheless this form of R is hereby considered accurate
et al. [12], but with respect to the memory loading surface and not enough for simulations of cyclic loading. Furthermore, by
the bounding surface related to the peak stress ratios as performed comparing Eqs. (10) and (11) it is deduced that the new model
here. According to this rule, the estimation of rbi presupposes the adopts a non-associative flow rule, since the dilatancy function D
definition of a relocatable projection centre rref defined as the is not related to the volumetric parameter n:r of the L. This
deviatoric stress-ratio state where the last load reversal took place because the dilatancy function D is defined explicitly as [28]
(see Section 2.6 for triggering of load reversal). Once the reference " sffiffiffiffiffiffiffiffiffiffiffiffiffi#
state rref has been defined, the conjugate (image) point on the d /dd S
D ¼ Ao d 2 ð18Þ
bounding surface rbi is found as the crossing point of the bounding ddmax
surface and the line originating from r and projecting along the (r–
Note that the D in the antecedent model [9] did not include the
rref) direction. During the implementation of this model in FLAC, the
term in brackets, which was added by Dafalias et al. [28] for better
Pegasus procedure of Dowel and Jarratt [45] was adopted for
simulations at high stress ratios. Furthermore, note that the sign
locating rbi, a procedure that was also used by Sloan et al. [40] to
of D is solely dependent on the sign of dd, thus ensuring negative
locate the stress path-yield surface intersection point. This proce-
plastic volumetric strain rates e_ pp o0 for negative values of dd, i.e.
dure is unconditionally convergent in four or five iterations (for an
when the stress point is outside the dilatancy surface.
error tolerance of 10  5 in terms of r values), does not require the use
of derivatives and is detailed in Andrianopoulos et al. [37].
This mapping rule is obviously simple, stable (since small 2.4. Non-linear hysteretic ‘‘elastic’’ moduli
perturbations of the loading do not affect the rbi location, unless
they trigger the load reversal criterion) and depends on the recent Based on what was mentioned in the introduction and Section
loading history (as this is portrayed by the reference state rref). 2.1, the ‘‘elastic’’ strain rate of the proposed model is estimated on
Hence, it retains the fundamental characteristics of the mapping the basis of non-linear hysteretic moduli. In particular, following
rule of the two-surface model, with the drawback of having a isotropic elasticity, the tangential ‘‘elastic’’ bulk modulus Kt is
discontinuously relocatable projection centre rref, as opposed to a related to the tangential ‘‘elastic’’ shear modulus Gt via a constant
kinematically hardening deviatoric back-stress ratio a. elastic Poisson’s ratio n (a model constant). In turn, based on [9],
Having defined rbi (and thus n, as per Eq. (14)) enables the the non-linear hysteretic form of Gt is given the following
depiction of rci and rdi, i.e. the conjugate (image) point on the generalized Ramberg–Osgood type of relation
critical state and the dilatancy surfaces, respectively, as the points  rffiffiffiffiffi 
Gmax Bpa p 1
on these surfaces that corresponds to the same Lode angle y (see Gt ¼ ¼ 2
ð19Þ
T 0:3 þ 0:7e pa T
Eq. (6) and Fig. 1). Analytically, the conjugate (image) points on
the three surfaces (collectively rc.b,d
i ) are given by 
0 ffi1
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffi ! 1 1=2ðrr ref Þ:ðrr ref ÞA
2 c,b,d T ¼ 1þ 1 @   ð20Þ
a1 Gref
rc,b,d
i ¼ M n ð15Þ a max
1 g
3 y pref 1

The scalar distances db and dd (collectively db,d), of the current where B, a1, g1 are model constants, pa ¼98.1 kPa is the atmo-
state of stress from the bounding and the dilatancy surfaces spheric pressure, T is the scalar variable which introduces
respectively, are defined on the p-plane of the deviatoric stress- tangential shear modulus degradation (see illustration in Fig. 3),
ratio space and along the direction of n (see Fig. 1), thereby while pref and Gref
max correspond to the values of p and Gmax (using
retaining the philosophy of the two-surface model. In particular, Eq. (19) for T¼1) at the last load reversal state (where r ¼rref).
these distances are thus estimated as Eqs. (19) and (20) ensure that the ever-current value Gt is
assumed to decrease smoothly from its maximum value Gmax, as a
db,d ¼ ðrb,d
i rÞ : n ð16Þ function of the distance of the current state (tensor r) from the
Note that distances db,d may become positive or negative, last load reversal (rref) on the p-plane of the deviatoric stress-ratio
depending on whether r is within or outside the respective space (see Fig. 1). Compared to the form of these equations in [9],
surface. Besides the foregoing distances, of interest for the Eq. (19) has no longer a maximum value for T equal to (2/a1  1), a
estimation of the plastic modulus and the dilatancy (see Sections fact leading to a smoother transition in the cyclic response from
2.5 and 2.3) are also dbmax and ddmax (collectively db,d
max ), which medium to large cyclic shear strains (see Fig. 3, for comparison
correspond to the ever-current maximum possible values of db,d with the formulation in [9]). In addition, for monotonic loading
along the direction of n, and are given by the shear modulus degradation introduced by T is not required,
rffiffiffi and therefore Gt requires the calibration only of B (see details in
2 b,d
db,d
max ¼ ðM þ Myb,d
þp
Þ ð17Þ Appendix). Therefore, the proposed model requires 2 less
3 y
constants for use in monotonic loading.
Note that the definitions of dbmax and ddmax above follow the It needs to be stressed that the very large majority of elasto-
proposals of Manzari and Dafalias [6] and Dafalias et al. [46], plastic models aiming at cyclic loading of sands in the literature
respectively. depend on the plastic strain rate for modeling soil-nonlinearity
and usually use hypoelastic moduli for the elastic strain rate
2.3. Flow rule component. Such an option works seemingly well for models
aiming at flow liquefaction and/or cyclic mobility, since the
With respect to the flow rule of the new model, Eq. (11) shows pertinent response is characterized by large cyclic shear strains
that the component of the plastic strain rate direction R on the for which the elastic component is a small percentage of the total
900 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

Function hb introduces the highly non-linear interpolation rule


of the new model according to

jdb j3
hb ¼ ho ð22Þ
odbmax jdb j 4

with ho being a model constant, and distances db and dbmax being


estimated by Eqs. (16) and (17), respectively. The foregoing
equation induces a much more non-linear interpolation rule (due
to the exponent 3 in the nominator) in comparison to the
antecedent model [9], a fact that has proved necessary after
introducing the vanished elastic region (e.g. following load
reversal far from failure, a vanished elastic region model without
the exponent 3 would give much larger plastic strains than what
data show and what a two-surface model, like [9], would give due
to the existence of the ‘‘small’’ elastic region).
Function hf is the second key constitutive ingredient providing
enhanced accuracy for the simulation of cyclic response that was
first proposed, in a form similar to the current one, by
Papadimitriou and Bouckovalas [9]. Specifically, this function acts
as an empirical macroscopic indicator of the effect of sand fabric
evolution during shearing on the magnitude of the plastic strain
rate. In more detail, the original proposal [29,30,8] took into
account experimental observations that loading within the
dilatancy surface re-arranged sand particle contacts rendering
the material ‘‘stiffer’’ (e.g. [47]), while loading in dilation (outside
Fig. 3. Exemplary pure shear stress–strain relation [t–g] according to the
proposed Ramberg–Osgood type shear formulation. Comparison with formulation the dilatancy surface) led to much ‘‘softer’’ response upon load
in Papadimitriou and Bouckovalas [9]. reversal (e.g. [42]). Hence, the hf composed of a non-directional
ratio of integrals of the plastic volumetric strain rate, namely of
strain. Nevertheless, if one uses such models for predicting the the whole loading history in the numerator (simulating the
response under smaller cyclic shear strains amplitudes (say ‘‘stiffening’’ response) and of the history of loading in dilation
smaller than 0.05%), he may find that the experimentally between successive load reversals in the denominator (depicting
measured soil non-linearity is under-predicted. An example of the foregoing ‘‘softening’’ response). In parallel, following micro-
such under-prediction is demonstrated by Papadimitriou and mechanical observations of sand particle contact normal orienta-
Bouckovalas [9]. tion distribution (e.g. [48]), Dafalias and Manzari [31,28] proposed
To this extent, it is also worth mentioning that, since the a purely deviatoric fabric-dilatancy tensor that evolves only upon
foregoing publication, pertinent constitutive models have ap- dilative phases of loading. This fabric-dilatancy tensor was used to
peared in the literature with non-linear hysteretic forms of the define a scalar multiplier of the dilatancy function D that leads to
‘‘elastic’’ moduli that are directly based, or merely inspired by the enhanced contractive response upon load reversal following
foregoing proposal. For instance, Loukidis and Salgado [10] dilation. Finally, Papadimitriou and Bouckovalas [9] re-formulated
incorporate the Ramberg–Osgood type formulation unaltered in their non-directional function hf to account for the directivity of
their two-surface plasticity model, while Lopez-Querol and sand fabric evolution during dilation [31,28], by introducing an
Blazquez [20] use it in a simplified manner in their endochronic evolving fabric-dilatancy tensor F ¼f + (fp/3)I, where f is the
liquefaction model. deviatoric component of F and fp ¼trace(F) ¼F:I, which enters
the form of hf as

2.5. Plastic modulus 1 þ /F : IS2 1 þ/fp S2


hf ¼ ¼ ð23Þ
1þ /F : nS 1 þ /f : nS
The basis of the constitutive response for large cyclic shear
with components fp and f having independent rates of evolution
strains is the form given to the plastic modulus Kp (Eqs. (9) and
(13)) and the flow rule (Eqs. (11)–(13)). Specifically, the plastic f_ p ¼ Ne_ pp ð24Þ
modulus Kp is given as [9]

Kp ¼ phb hf db ð21Þ f_ ¼ N oe_ pp 4½Fm n þf ð25Þ

where hb and hf are dimensionless positive functions to be defined with Eq. (25) being adopted from [31,28]. Note that the term
in the sequel, while the sign of distance db from the bounding ‘‘fabric-dilatancy’’ used above for tensor F describes the fact it
surface (Eq. (16)) controls the sign of Kp. Specifically, when db is evolves as a function of the (plastic) volumetric strain rate
negative, the Kp becomes negative too, but the loading index L (Eqs. (24) and (25)) and that it is not actually a true fabric tensor,
remains positive thus simulating post-peak shear strain softening since it does not quantify any physically measurable property
response of dilative soils (with respect to the deviatoric stress of the sand fabric. Positive quantity Fm in Eq. (25) corresponds to
ratio and not deviatoric stress), an attribute of the formulation the ever-current maximum norm allowed for tensor f, while N
originating from Manzari and Dafalias [6]. Note that allowing the corresponds to the intensity of sand fabric evolution for both
stress state to go outside the bounding surface is an original the volumetric and the deviatoric components in Eqs. (24) and
constitutive attribute, which facilitates the simulation of strain (25), respectively. In particular, the foregoing quantities are
softening, as opposed to other relatively more complicated efforts given by
in the literature (e.g. via the disturbed state concept of Park and
Desai [22]). Fm ¼ 4maxjfp j2 ð26Þ
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 901

 
pa the direction of n. In such a case, the increase of the denominator
N ¼ No /co S ð27Þ
s1o of hf leads to larger plastic strains (due to a decrease of Kp), which
under undrained conditions provides the well-known intense
where No is a model constant, while co and s1o correspond to the contraction following dilation at small p values, that may
values of the state parameter and the major principal stress at the eventually lead to liquefaction or cyclic mobility (see Fig. 4b).
beginning of cyclic loading (e.g. at the end of consolidation), both Further details on the operation of hf may be found in
quantities being used for appropriately scaling the intensity N of Papadimitriou and Bouckovalas [9].
sand fabric evolution according to the prevailing initial condi-
tions. Note that based on Eq. (24), scalar fp may become negative
(e.g. after a lengthy dilative phase) and this is the reason for 2.6. Triggering of load reversal
taking the maximum of |fp| for quantifying the norm of f in Eq.
(26). Furthermore, note that due to the exponent 2 in Eq. (26), the Both the two-surface and the new model require the definition
maximum value of the denominator of hf becomes of the same of a relocatable load reversal point that affects the values of the
order of magnitude as the maximum value ever retained by its ‘‘elastic’’ and plastic strain rates. Specifically, the two-surface
numerator. Based on Eq. (27) parameter N becomes non-zero only model [8,9] had different triggering mechanisms of load reversal
for initially dilative states (co o0), implying that the effect of sand for the ‘‘elastic’’ and the plastic strain rates. Nevertheless, given
fabric evolution needs to be simulated for such states only. In the vanished elastic region of the new model, it was considered
extension, the more dilative the initial state, the more important more appropriate to hereby adopt a unique triggering mechanism
should the effect of sand fabric evolution be, and therefore N of load reversal for both strain rates. Thus, in the proposed model,
increases with o–co 4. Finally, the addition of the s1o term in load reversal takes place when the scalar loading index L takes a
Eq. (27), instead of a similar po term as in [8], depicts the negative value, a triggering criterion that has been used success-
additional effect of initial stress anisotropy on the intensity of fully in the literature (e.g. [12,15]). When this occurs, the load
sand fabric evolution. reversal point (reference state) is automatically relocated, giving
Compared to the form of Eqs. (23)–(27) in Papadimitriou and updated values to rref, pref and Gref
max .
Bouckovalas [9], Eq. (26) allows for a 4 times larger Fm value, thus The update in rref affects the location of the conjugate (image)
providing potential for more intense contractive response upon point on the bounding surface rbi (Fig. 1), which in turn affects the
unloading following dilation. This was found necessary after trial- direction of n (Eq. (14)), the cornerstone of the plastic strain rate.
and-error simulations of cyclic mobility tests for compensating In addition, load reversal affects the ‘‘elastic’’ strain rate, since the
the effect of the use of the new more non-linear interpolation rule ‘‘elastic’’ moduli Gt and Kt reduce as a function of the |r–rref|
(Eq. (22)). In addition, the term (pa/s1o) in Eq. (27) no longer distance, pref and Gref
max (Eqs. (19) and (20)). Note that L remains
possesses an exponent, and in this way the proposed model has negative only momentarily, since there is no yield surface and the
one model constant less than the antecedent model [9], without response at every incremental step is elasto-plastic. Thus,
essential loss of accuracy in simulating cyclic loading. immediately after relocating rref, the direction of n changes and
Focusing on the operation of scalar hf, Fig. 4 presents the the scalar loading index L becomes non-negative again, leading to
simulation of an undrained cyclic simple shear test leading to the estimation of plastic strains at every incremental step, as per
cyclic mobility. Observe in Fig. 4 that for load paths within the Eq. (8).
dilatancy surface, scalar hf gradually increases due to its Note that, upon load reversal, both the ‘‘elastic’’ and the plastic
numerator ensuring that the history of cyclic shearing leads to strain rates take values much smaller than those in the preceding
stiffening response. In parallel, Eq. (25) makes the deviatoric time step, thus accurately simulating the stiffer soil response
component f evolve only upon dilation (due to the Macaulay upon load reversal. More importantly, due to the adopted
brackets), and this in the opposite sense to tensor [Fmn+ f] due to interpolation rule (Eq. (22)) the plastic strain rate remains non-
the negative sign in front of N. In this way, the of:n4 term in the zero upon load reversal, as opposed to other models which retain
denominator of hf becomes non-zero, only upon load reversal in memory a last reference point and introduce a Kp-N
following a dilative shearing phase, and this due to the change in condition upon load reversal (e.g. [12,14,28]). It is believed that

Fig. 4. Simulation of an undrained cyclic simple shear test: (a) shear stress t—normal effective sv relation and (b) evolution of hf during the shearing.
902 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

the selected form of the interpolation rule reduces the unrealistic The new model has been implemented in FLAC [35], using the
(instability) effects of numerical perturbations causing false user-defined-model (UDM) capability and therefore the remain-
triggering of the load reversal criterion in boundary value der of this paper presents its accuracy in both element test
problems, as compared to models that introduce a Kp-N simulations, as well as different boundary value problems, using
condition upon load reversal in favor of enhanced accuracy in the foregoing numerical algorithm. Note that the accuracy of the
the simulations of (fully controlled) cyclic loading element tests. antecedent two-surface model [8,9] has only been documented at
It should be clarified that the mapping and interpolation rules, the element test level, since it was never implemented to a
as well as the related load reversal criterion which introduces a numerical code.
discontinuously moving rref point, have been adopted with In short, with the foregoing 2D computer code, the equations
seismic ground response analyses in mind. For non-proportional of motion are integrated using the explicit central difference
non-monotonic paths that are not necessarily related to seismic integration rule. Darcy’s law is invoked for fluid flow in a porous
shaking (e.g. ratcheting, rotational shear) the constitutive model solid, while the incremental formulation of coupled deformation-
has not been fully tested. However, Wang et al. [12] showed that diffusion processes provides the numerical representations for the
the adopted mapping rule (in relation to a memory loading linear quasi-static Biot theory. Stress integration of the constitu-
surface and not the bounding surface as performed here) may tive relations at each zone for a given strain increment e_ is
provide successful simulations, at least for certain rotational shear performed via a second order modified forward Euler method
paths. Still, it is believed that successful simulations of various with automatic error control [39,40], after appropriate manipula-
non-proportional paths may require some additional constitutive tion to account for the complexities of the proposed model.
alterations (e.g. the incorporation of inherent anisotropy, as per Integration is accomplished in one or more sub-increments (or
Dafalias et al. [46], and/or an additional loading mechanism, as substeps), in order to maintain the local truncation error at each
per Li and Dafalias [43]), that are not necessary in a two-surface step below a desired tolerance level STOL set at this study equal to
model like the antecedent one. Moreover, false triggering of the 10  3. Further details about the stress integration algorithm used
load reversal criterion may potentially occur in cases that the in the implementation, as well as its accuracy and efficiency for
stress path approaches the bounding surface at a very shallow boundary value problems involving earthquake-induced liquefac-
angle, especially because the n is not necessarily normal to the tion are presented in Andrianopoulos [36] and Andrianopoulos
bounding surface (see Fig. 1). Nevertheless, such paths are not et al. [37,38].
common in the boundary value problems in question, while Note that all numerical simulations (at the element and
defining the n as collinear to the normal to the rbi point would system level) hereby presented have been performed with the
make the constitutive equations much more complex (see [43]). set of values of the model constants of Table 2.

3. Model performance 3.2. Element response

3.1. General For completeness, the presentation initiates from monotonic


testing. Hence, Fig. 5 presents an exemplary comparison of
The new model requires the calibration of 13 dimensionless simulations to data from monotonic drained triaxial
and positive constants for cyclic loading, and only 11 for compression tests under p ¼80 kPa ( ¼constant) conditions and
monotonic (setting a1 ¼1.0 makes the response independent of different initial relative densities Dr of Nevada sand [49]. Fig. 6
g1 as well, see Eq. (19)). The full list is presented in Table 2, along does the same for monotonic undrained triaxial compression tests
with their values for the Nevada sand database [49]). This table for Dr ¼60% and different initial p values for the same sand. It
also includes their range for typical non-cohesive soils based on becomes obvious that the proposed model offers reasonable
the literature and the authors’ experience. The first 10 constants simulation of the drained and undrained monotonic response
of the list may be directly estimated on the basis of the specified under various initial stresses and densities with a single set of
monotonic and cyclic element tests, while the remaining 3 model constants, that of Table 2 (note the use of B ¼180 and
constants require trial-and-error simulations of element tests. a1 ¼1.0).
Due to paper length limitations, the calibration process is merely Yet, the scope of the new model is primarily the simulation of
outlined in Appendix and the reader is referred to [8] for further seismic ground response and the first step for this purpose is
details. successful simulation of the cyclic response at the element level.

Table 2
Model constants: physical meaning, range for typical sands and values for Nevada sand.

# Physical meaning Typical Range Value

Mcc Deviatoric stress ratio at critical state in triaxial compression (TC) 1.20–1.37 1.25
c Ratio of deviatoric stress ratios at critical state in triaxial extension (TE) over TC 0.68–0.80 0.72
Gcs Void ratio at critical state for p ¼ 1 kPa 0.77–1.03 0.910
l Slope of critical state line in the [e-lnp] space 0.01–0.03 0.022
B Elastic shear modulus constant 550–950 600a
v Elastic Poisson’s ratio 0.10–0.40 0.33
kbc Effect of c on peak deviatoric stress ratio in TC 0.5–4.0 1.45
kdc Effect of c on dilatancy deviatoric stress ratio in TC 0.1–3.0 0.30
g1 Reference cyclic shear strain for non-linearity of ‘‘elastic’’ shear modulus 0.0065%–0.025% 0.025%a
a1 Non-linearity of ‘‘elastic’’ shear modulus 0.45–0.85 0.6a
Ao Dilatancy constant 0.5–1.5 0.8
ho Plastic modulus constant 3,000–100,000 15,000
No Fabric evolution constant 30,000–80,000 40,000

a
for monotonic loading of Nevada sand: B¼ 180, a1 ¼ 1.0 (that renders the value of g1 irrelevant due to T¼ 1 in Eq. (19)).
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 903

Fig. 5. Comparison of simulations to data from monotonic drained triaxial tests on


Nevada sand (p ¼80 kPa, constant) with Dr ¼40% (CIDC 40-100) and Dr ¼ 60% (CIDC
60-75).

Fig. 7. Summary comparison of simulations to data from resonant column tests on


Nevada sand [49], in terms of: (a) the maximum shear modulus Gmax, (b) the
secant shear modulus G/Gmax degradation and (c) the hysteretic damping x
increase curves, with cyclic shear strain gcyc.

In particular, the comparison is made in terms of the maximum


shear modulus Gmax and the degradation of the secant shear
modulus G/Gmax and the increase of the hysteretic damping x with
the amplitude of the cyclic shear strain. Different symbols denote
different relative densities (40% and 60%), while the confining
pressure in these tests ranged from 40 to 160 kPa. The black
curves denote the model simulations, while the related gray
curves of Vucetic and Dobry [50] for plasticity index PI ¼0% are
also added to this figure for comparison purposes.
Focusing on the model accuracy under large cyclic shear strain
amplitudes, Fig. 8 presents a one-to-one comparison of
simulations to data from a typical cyclic undrained simple shear
test on Nevada sand at Dr ¼40% and initial effective stress
svo ¼80 kPa [49]. In order to ascertain whether such a
Fig. 6. Comparison of simulations to data from monotonic undrained triaxial tests satisfactory accuracy is obtained for all initial conditions, Fig. 9
on Nevada sand (Dr ¼60%) with po ¼80 kPa (CIUC 6011) and po ¼160 kPa (CIUC
compares the liquefaction resistance curves from all cyclic simple
6004).
shear tests on Nevada sand [49] with the respective simulations
and pertinent curves from the literature [51]. Furthermore, model
Therefore, Fig. 7 presents the overall comparison of simulations to accuracy has been also ascertained in terms of the rate of excess
data for cyclic shearing under small and medium cyclic shear pore pressure Du buildup in the cycles that precede initial
strain amplitudes measured in the resonant column device ([49]). liquefaction. An example of such an overall comparison is
904 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

Fig. 8. Comparison of simulation to data for a typical cyclic undrained simple shear test on Nevada sand with Dr ¼40% [49].

Fig. 10. Summary comparison of simulations to data from all cyclic undrained
simple shear tests on Nevada sand [49], in terms of the excess pore pressure after
the first load cycle Du1.

less for cyclic loading, while for monotonic loading the number of
model constants reduces by three.

3.3. Seismic response of horizontal liquefiable layer

Results from centrifuge Model tests No 1 and No 2 of the


VELACS project have been used to test the ability of the new
Fig. 9. Summary comparison of liquefaction curves from simulations to data from
model in simulating the one-dimensional (1D) seismic response
all cyclic undrained simple shear tests on Nevada sand [49], as well as established
curves from the literature [51]. of a liquefiable soil layer under level and mildly sloping ground
conditions, respectively. This paper presents comparisons of
simulations to data from centrifuge Model tests No 1, while
presented in Fig. 10, where the excess pore pressure after the first documentation regarding the validation run of centrifuge Model
load cycle Du1 from all cyclic simple shear tests is compared to test No 2 can be found in Andrianopoulos et al. [37].
the respective values from the simulations. The emphasis given The test arrangement and the instrumentation for the fore-
on the first load cycle is backed by the fact that experimental data going model test are shown in Fig. 11. In prototype scale, the
show that the rate of Du buildup decreases after the first (most experiment refers to a 10 m deep Nevada Sand layer of
significant) cycle and increases again only on approaching initial approximately 40% relative density [49], with a permeability
liquefaction (e.g. see typical rate of Du buildup according to Seed coefficient of 2.1  10  5 m/s. The water surface was 1 m above the
and Booker [52]). ground surface. The centrifuge model was built inside a laminar
Overall, a comparison of the proposed model performance box, which was tested at 50g centrifugal acceleration. The
with that of the antecedent model does not show significant excitation applied at the base of the laminar box was essentially
improvement at the element level. Yet, the benefit becomes sinusoidal with a peak horizontal acceleration amplitude of
apparent by the fact that the new model has one model constant 0.235g, a strong motion duration of approximately 11 s and a
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 905

Fig. 11. Schematic illustration of testing configuration of VELACS centrifuge Model No.1 test.

Fig. 12. Time history of base horizontal acceleration for VELACS centrifuge Model
Fig. 13. Variation of permeability used in the numerical simulation of the VELACS
No 1 test [49].
centrifuge Model No 1 test (based on Arulanandan and Sybico [54]).

predominant frequency of 2 Hz (Fig. 12) in prototype scale. The thick free water layer. Flow was allowed only at the top boundary,
response of the level ground was monitored along the axis as well while the bottom and the lateral boundaries were assumed
as closer to the box boundaries with eight (8) accelerometers, impermeable.
eight (8) pore pressure transducers and six (6) LVDTs measuring Initially static equilibrium was achieved, in order to create the
displacements. geostatic stress field, and then the dynamic motion was applied.
The finite-difference grid used for the numerical analysis of the The initial static equilibrium was achieved by assuming elastic
prototype sand layer consisted of 1.0 m  1.0 m square elements behavior for the soil corresponding to a Ko value equal to 0.5
(‘zones’ in FLAC terminology). The actual relative density of the (elastic Poisson’s ratio equal to 0.33), since no further details were
sand was Dr ¼45%, somewhat larger than the Dr ¼40% originally given in the factual report [49]. A coupled dynamic analysis was
specified [49]. Thus, the sand was assumed uniform with an initial performed, taking into account the interaction between the
void ratio of 0.724, corresponding to the former value of Dr. The mechanical behavior and the diffusion process. The permeability
bottom nodes of this grid were set to follow the prescribed input coefficient of the sand was scaled appropriately in order to take
motion in both directions. In particular, in the horizontal direction into account: (a) the faster diffusion rate occurring at centrifugal
the bottom nodes followed the (digitally available) VELACS Ng accelerations, since water was used as pore fluid in this
horizontal acceleration time-history shown in Fig. 12. The experiment (kprototype ¼Nkmodel) and (b) the significant evolution
parasitic vertical acceleration during the experiments was minor of sand fabric during the process of soil liquefaction, which
and consequently was not taken into account in the numerical temporarily increases the permeability coefficient. Namely, the
simulation. The lateral boundaries were free to move in horizontal permeability variation shown in Fig. 13 was used, where the
and vertical direction, while tied to one-another in order to initial permeability increases rapidly as the soil approaches initial
enforce the same horizontal and vertical displacements of the two liquefaction (to2 s), while it decreases more smoothly thereafter.
boundaries, thus simulating the laminar box device used in the The permeability at the point of initial liquefaction when the soil
centrifuge test. A uniform normal stress and related pore particles lose full contact with each other and water flow paths
pressures were applied at the top boundary, to simulate the 1 m are greater, reaches 6.35 times the initial value, while the mean
906 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

increase throughout shaking reaches the value of 1.67. This recording and the simulation depict a liquefaction-induced
procedure was also used by Manzari and Arulanandan [53] to deamplification of the acceleration at the ground surface. More
simulate VELACS Model 1, incorporating in this manner the importantly, after t¼ 5 s the extensive liquefaction of the ground
findings of Arulanandan and Sybico [54], who investigated the nullifies the acceleration at the ground surface, as deduced by
change of sand fabric during liquefaction using electrical both the recording and the simulation. At the mid-depth of the
properties to measure changes in pore shape factor and tortuosity. sand layer, this deamplification is less intense in both simulation
Fig. 14 compares the time histories of the excess pore pressure and recording, and especially the latter. The differences observed
ratio ru ¼ Du/svo from the centrifuge test recordings to their at this depth may be attributed to the relatively lower recorded
numerically simulated counterparts, at various depths ( 1.45, versus simulated excess pore pressures from the depth of  5.0 m
2.6, 5.0,  7.5 m) of the sand layer, always along the axis of and below.
the centrifuge model. Both recordings and simulations show This centrifuge test also showed significant free-field settle-
initial liquefaction (ru ¼1.0) in the upper 5–7 m. In particular, very ments (of the order of 20 cm), which developed mostly during
high values of ru develop from the first seconds of the shaking in shaking and not in the post-shaking excess pore pressure
the upper 2.5 m, while deeper locations develop similarly high dissipation phase. Qualitatively, this type of settlement accumu-
values of ru at later stages of the shaking. Moreover, the rates of lation was attributed by Arulanandan and Sybico [54] to the sand
excess pore pressure buildup and dissipation are satisfactorily fabric changes during initial liquefaction, where the increase of
simulated, with the possible exception of the first 2 loading cycles permeability (see Fig. 13) and the consequent faster dissipation
during which the model shows more intense buildup than the tend to counterbalance the buildup of excess pore pressures due
data. to shaking, while simultaneous sand grain sedimentation in-
Fig. 15 compares the time histories of ground acceleration (in creases surface settlements. The numerical simulation also
percent of g) from the centrifuge test recordings to their showed development of settlement during shaking, but its total
numerically simulated counterparts, at the ground surface (AH3) amount was much smaller (4 cm). The inability of numerical
and at the mid-depth (AH5 at a depth 5.0 m) of the sand layer analyses to predict this amount of free-field settlement
and along the axis of the centrifuge model. The agreement is was demonstrated in the VELACS project, where, apart from Lacy
satisfactory. It is especially noteworthy that the both the [55] and Manzari and Arulanandan [53], most of numerical

Fig. 14. Comparison of data to simulations for the time history of the excess pore pressure ratio Du/svo developed at various depths along the axis of the model of the
VELACS centrifuge Model No 1 test.

Fig. 15. Comparison of data to simulations for the time history of the horizontal acceleration developed at various depths along the axis of the model of the VELACS
centrifuge Model No 1 test.
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 907

simulations ranged between 1 and 7 cm. In the present study, PPT3) also located under the foundation, one (1) pore pressure
this disagreement can be attributed to: (a) the inability of the transducer (PPT4) placed in the free field, as well as, one (1) LVDT
model to predict plastic deformations for paths under (nearly) monitoring the structural settlement. The locations of the fore-
constant values of the deviatoric stress ratio r, (b) the inability of going instruments are approximately depicted in Fig. 16.
the model to simulate the sand fabric changes during initial A series of centrifuge tests were performed for Model No 12 at
liquefaction, as these were highlighted above by Arulanandan and three different universities, namely at Princeton University
Sybico [54], besides the macroscopic change in permeability. (PRNU, [56]), at the Rensselaer Polytechnic Institute (RPI, [57])
However, as depicted in paragraph 3.4 that follows, the foregoing and the University of California at Davis (UCD, [58]). However,
under-prediction of free-field settlements, did not affect the some of the performed tests had deviations from the original
accuracy in simulating foundation settlements, for which the specifications, especially at the base input motion, whose target
mechanism of loading and settlement accumulation is quite excitation consisted of ten (10), more or less, uniform cycles of
different. purely horizontal acceleration with 0.25g amplitude and 2 Hz
frequency. To explore these deviations, Table 3 presents a list of
the successfully executed centrifuge tests of Model No 12 with the
3.4. Seismic response of shallow foundations on liquefiable soil values of the peak acceleration amax (maximum absolute value
during strong shaking), the average acceleration aave (average
The ability of the proposed model to simulate soil-structure value of positive and negative peaks of all significant cycles during
interaction in a liquefiable soil regime is validated via comparison strong shaking), the duration of strong motion td and the values of
of simulations to measurements from the VELACS centrifuge the residual structural settlement, as measured from the various
Model test No 12 which simulates the response of shallow tests, while Fig. 17 presents the applied acceleration of the test at
foundations on liquefiable soils. The test arrangement and the RPI, that was similar to the target excitation (that was never
instrumentation for the foregoing test are shown in Fig. 16. In attained experimentally).
prototype scale, the experiment refers to a 6 m deep Nevada sand The finite difference mesh used for the numerical analysis is
layer with 60% relative density [49] and permeability coefficient also presented in Fig. 16. As depicted in this figure, the zone size
of 2.1  10  5 m/s, overlaid by a 1 m thick, low permeability varied from 0.75 m  0.5 m at the vicinity of the structure to
(5  10  9 m/s), Bonnie silt layer, with the free water surface being 1.5 m  1.0 m far from its location, and the total number of zones
1 m above the silt surface. A 4 m high and 3 m wide rigid square was 280. The input acceleration used in the simulations was
structure is seated 0.5 m below the silt-sand interface (and in the similar to the target excitation (see Fig. 17 for comparison with
plan center of the model container), inducing an average bearing the RPI test acceleration), and was applied both at the bottom and
pressure of 150 kPa. the lateral boundaries in order to simulate the rigid box device.
The centrifuge model was built inside a rigid box with an The vertical acceleration during the experiments was minimal
inside plan area of 282 mm  128 mm and tested at Ng¼100g and was not taken into account during the numerical simulation.
centrifugal acceleration. In general, the response of the ground Bonnie silt layer was simulated by using a simplified version of
and the structure was monitored with three (3) accelerometers the proposed model, which includes only the Ramberg–Osgood
(AccB, AccC and AccD), located at different depths under the type stress–strain relation (Eqs. (19) and (20)) appropriately
foundation, three (3) pore pressure transducers (PPT1, PPT2 and calibrated to simulate the non-linear hysteretic response of a fine

Fig. 16. Schematic illustration of testing configuration of VELACS centrifuge Model Fig. 17. Comparison of numerically applied and measured time history of base
No 12 test, including the mesh and the applied boundary conditions of the horizontal acceleration for VELACS centrifuge Model No 12 test (data from test that had
numerical simulation. input base motion close to the target excitation and was performed at RPI).

Table 3
Measured structural settlements and pertinent numerical estimates, on the basis of the characteristics of the applied base excitation for the VELACS Model No 12.

Source # amax(g) aave(g) td(sec) Settlement (cm)

Princeton University (PRNU) 1 0.30 0.26 6.65 27


3 0.40 0.34 6.70 47
5 0.21 0.18 5.40 22
6 0.24 0.21 5.50 21
University of California at Davis (UCD) 2 0.37 0.31 7.50 18
3 0.25 0.21 6.40 9
Rensselaer Polytechnic Institute (RPI) 1 0.26 0.21 5.70 13
Numerical 1 0.24 0.21 5.75 10.8
2 0.32 0.30 7.25 15.2
908 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

soil with plasticity index PI¼ 15% (B ¼500, e¼0.70, a1 ¼0.64,


g1 ¼0.027%, n ¼0.33). This type of simulation of Bonnie silt is
simplistic, but preferable over using any elastic-perfectly plastic
model available in FLAC, given the inability of such models to
accurately simulate the shear modulus degradation and hysteretic
damping over a wide range of cyclic shear strains. Moreover, this
simplification is not expected to affect the structure settlement
results, since: (a) the structure was founded on the sand and not
the silt layer, (b) the observed mode of failure was mainly
‘punching through’ and not ‘general failure’ and (c) the thickness
of the silt layer is small compared to that of the sand.
The structure was simulated as a rigid block (elastic elements
of very high stiffness) of appropriate unit weight to portray the
desired contact pressure. Since the numerical analysis assumed a
2D plane strain behavior, the actual bearing pressure of the
structure in the 3D model (150 kPa) was reduced to its 2D plane
strain equivalent (95 kPa), as suggested by Popescu and Prevost
[59]. No interface elements were used between the structure and
the rest of the grid. Furthermore, uniform normal stress and
related pore pressures were applied at the top soil boundary, to
simulate the 1 m thick free water layer.
Static equilibrium was achieved by ‘‘building’’ the structure in
multiple steps, and before applying the dynamic excitation. The
initial static equilibrium was achieved by assuming elastic
behavior for the soil corresponding to Ko value equal to 0.5
(elastic Poisson’s ratio equal to 0.33), since no further details were
Fig. 18. Comparison of data to simulations for the time history of the horizontal
given in the factual report [49]. As in the previous case of the level acceleration developed at various depths below the structure of the VELACS
ground response, a coupled dynamic analysis was performed, centrifuge Model No 12 test (data from RPI test with input motion shown in
taking into account the interaction between the mechanical Fig. 13).
behavior and the diffusion process, while the permeability
coefficient was scaled appropriately in order to take into account
the faster diffusion rate of the pore water occurring at centrifugal where high excess pore pressures leading to liquefaction are
Ng accelerations (kprototype ¼N kmodel). Opposite to the previous observed from the initial stages of shaking. On the other hand,
case, no further adjustment of permeability was made for the transducers PP1, PP2 and PP3 in the vicinity of the structure show
evolution of sand fabric during initial liquefaction (as per [53,54]), no liquefaction. Overall, the comparison between numerical and
since only a small portion of the sand layer liquefied. It is the test results is satisfactory, given the fact that most of the
authors’ belief that such type of permeability adjustment should measured excess pore pressure ratios presented has been
potentially be applied only when a great portion of a sand layer smoothed in this figure (as per [60]). The only consistent
liquefies and only when the driving stresses do not produce deviation between numerical and experimental results observed
significant dilation spikes during liquefaction (e.g. due to a static in Fig. 19 is the rate of excess pore pressure dissipation after the
shear stress offset (in gently sloping ground), or due to rocking end of the strong motion (t ¼6.5 s), with the measured rates
effects when a tall superstructure is present). (especially those from the RPI test) being higher that the
Fig. 18 presents a comparison between the simulated seismic simulated ones. One possible explanation for this observed
accelerations at two discrete depths below the structure and the difference is the aforementioned approximate simulation of the
respective recordings from the test at RPI. Observe that the real base input after 6.5 s of shaking.
numerical results plot reasonably close to the centrifuge data, at In addition, Fig. 20 compares the time history of structural
least during the strong motion phase (up to 6 s). After the strong settlement from the numerical analysis to those measured in the
motion phase, the simulations depict smaller accelerations at tests with input motions close to the target excitation, namely
both depths compared to the measurements, a discrepancy that is PRNU (#5, #6), UCD (#3) and the RPI test. Observe that both
attributed to the fact that the real input base motion continues measured and simulated structural settlements are accumulated
with reduced amplitude even after 6 seconds of shaking, contrary only during strong shaking, with the simulation leading to a
to its numerically applied counterpart, which simulates the target residual value of settlement equal to 10.8 cm, which is close to the
excitation (see Fig. 17 for comparison of base motions). lower limit of the three presented tests (ranging from 9 to 22 cm).
Comparing the seismic amplifications at the two locations it is However, mainly due to deviations of the input motion in the
noted, that the numerical analysis depicts a slight amplification of various tests performed by the three universities (ranging from
the input motion (0.36g for ACC-C and 0.33g for ACC-B), while 0.21 to 0.37g), the structural settlements measured during the
similar response is measured in the centrifuge test (0.32g for both whole series of tests ranged between 9 and 47 cm, with the higher
ACC-C and ACC-B). values corresponding to the higher input intensities (see Table 3).
Fig. 19 compares predicted and recorded time histories of the In order to investigate whether the proposed model reproduces
excess pore pressure ratios ru ¼ Du/svo at the locations of the pore the aforementioned increase of measured structural settlements
pressure transducers. Note that this figure includes data only from due to higher intensity motions, a second numerical analysis was
the tests that had input base motions similar to the target performed with intensity comparable to the higher intensities
excitation and were performed in all three universities, namely (equal to 0.32g, as compared to base case of 0.25g, see Table 2).
PRNU (#5, #6), UCD (#3) and the RPI test. In particular, the The results of this investigation are presented in Fig. 21, where it
location of pore pressure transducer PP4 is representative of the is deduced that the numerically estimated structural settlements
conditions away from the structure, i.e. the free-field conditions, (dashed trend line) compare quite well with the measured ones at
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 909

Fig. 19. Comparison of data to simulations for the time history of the excess pore pressure ratio Du/svo developed at various depths below the structure (PPT1, PPT2, PP3),
as well as the free-field (PPT4) of the VELACS centrifuge Model No 12 test (data from tests RPI #1, PRNU #5 and #6, UCD #3).

solid trend line), a fact that is attributed to problems related to the


measuring device used (see [60] for details).

4. Summary and conclusions

This paper presents the constitutive relations and the


simulative potential of a new bounding surface plasticity model
developed mainly for the seismic liquefaction analysis of
geostructures. The new model builds on the two-surface model
of Papadimitriou and Bouckovalas [9] and introduces specific
constitutive alterations aimed at potentially enhancing its
accuracy, but mainly facilitating its implementation to numerical
Fig. 20. Comparison of data to simulations for the time history of the structural
settlement of the VELACS centrifuge Model No 12 test (data from tests: RPI #1, codes. In particular, the ‘‘small’’ conical yield surface of the two-
PRNU #5, UCD #3). surface model is reduced to a straight line emanating from the
stress space origin, a change that dictated a number of revisions,
the most important being the definition of the loading direction
and the forms of the mapping and interpolation rules.
The new model proves able to simulate successfully the cyclic
response of Nevada sand in element tests under small, medium and
large cyclic shear strain amplitudes, irrespective of the initial
conditions, with a single set of values of the 13 model constants
(one less than the two-surface model). This is attributed to three
(3) basic constitutive ingredients: the dependence of model
surfaces on the state parameter c (first proposed by Manzari and
Dafalias [6]), the Ramberg–Osgood type formulation of the ‘‘elastic’’
moduli that governs the response under small to medium cyclic
shear strains and the sand fabric evolution scalar multiplier hf of
the plastic modulus that ensures accuracy in the simulations under
medium to large cyclic shear strains, exactly as in the two-surface
model. Moreover, the new model shows reasonable accuracy in the
simulation of the monotonic response of the same sand.
Unlike the two-surface model, the new bounding surface
Fig. 21. Effect of peak acceleration of the input base motion on the residual
plasticity model has been implemented in a numerical code (FLAC),
structural settlement, based on a compilation of simulations and data from the thus enabling further verification for different boundary value
various centrifuge tests from Model No 12 (see Table 2). problems involving earthquake-induced liquefaction. In particular,
this paper presents the verification against two VELACS centrifuge
experiments, related to the free-field liquefaction response of a
RPI and U.C. Davis under various levels of shaking (lower solid sand layer (Model No 1), and the seismic response of a rigid
trend line). On the contrary, the tests performed at Princeton foundation on a sand layer (Model No 12), whereas Andrianopou-
University gave consistently higher values of settlement (upper los et al. [37] do the same for the laterally spreading response of a
910 K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911

liquefied sand layer (Model No 2). In all presented simulations, dilatancy, i.e. the ratio of the plastic volumetric strain increment
transient and permanent ground motions, as well as excess pore over its deviatoric counterpart, or, with adequate accuracy, the
pressures in the foundation soil and in the free field were ratio of the respective total strain increments that are directly
successfully simulated using the same set of values for the model measured. By depicting the deviatoric stress ratio at phase
constants calibrated on the basis of laboratory element tests. This is transformation (where D¼0 at high stress ratios), one may
credible evidence that the new model is a fairly reliable numerical proceed to the calibration of Ao on the basis of Eq. (28) focusing
tool for performance-based design in the hands of practitioners, as on the area of intermediate to high values of the deviatoric stress
well as an indispensable weapon for researchers in gaining ratio, for better accuracy. If step (f) is performed first, then the
valuable insight to the complex mechanisms which control the estimation of Ao may be based on the ratio of the plastic strain
liquefied ground response and soil-structure interaction. increments, by subtracting the pertinent elastic strain compo-
nents from the measured strain increments, but the gain in
accuracy is usually small.
Acknowledgements
Part B: Calibration of 3 remaining constants, using trial-and-error
The research presented herein was funded by the General runs, as follows:
Secretariat for Research and Technology (G.G.E.T.) of Greece, through
research project EPAN-DP23 (‘‘X-SOILS’’) and by the Basic Research (f) hypoelastic shear modulus constant, B: Its calibration requires
Fund ‘‘Lefkippos’’ of N.T.U.A. (Grant no. 65/1506). These contributions comparison of shear modulus values from simple shear
are gratefully acknowledged. The authors would also like to sincerely loading paths initiating from a Ko ¼1 condition and not
thank the two anonymous reviewers, whose constructive comments exceeding cyclic shear strains of 5  10  6 to pertinent
greatly enhanced the quality of the manuscript. small-strain measurements (e.g. from resonant column tests,
from shear wave velocity measurements via bender elements
5. Appendix Calibration of model constants (lab) or geophysical testing (insitu)). The need for trial-and-
error runs for B stems from the fact that due to the vanished
Given the resemblance of the proposed with the antecedent elastic region, the model exhibits elasto-plastic response even
model, their calibration process need not be presented in full for small cyclic shear strains. Therefore a direct estimation of
detail. Further details may be found in Papadimitriou et al. [8] in B on the basis of Eq. (19) with T¼ 1 from data may only serve
which constants are similarly denoted, with the exception of No as a lower limit for B, unlike in the antecedent two-surface
that appeared as Ho. model in which it gave the final value without any need for
trial-and-error runs. As in [8], for monotonic loading the B
Part A: Calibration of 10 directly ratable constants: value must be reduced (usually by 2 to 4 times) and is
calibrated by fitting Eq. (19) with T¼1 to the initial stages of
(a) Critical state constants, Gcs, l, Mcc and c: These are estimated on the deviatoric stress strain relation of monotonic triaxial tests.
the basis of monotonic element tests (usually triaxial compres- (g) plastic modulus constant, ho: Its value may be estimated by
sion and extension) that have reached the critical state. fitting both the effective stress path and the stress-strain
(b) Elastic Poisson’s ratio, v: This constant has usually small relation of monotonic tests (e.g. triaxial, simple shear), or the
variability in natural sands (e.g. 0.10 to 0.40) and may be first cycle of large-strain cyclic tests. If both drained and
estimated by small-strain measurements (e.g. cross hole in undrained tests are available, then these trial-and-error runs
non-saturated soils), or from the initial stages of 1D unloading may also be used for indirectly calibrating constants for which
with lateral stress measurements (e.g. in a triaxial cell). relevant data are not usually available (e.g. elastic Poisson’s
(c) State parameter dependence constants, kbc and kdc : Their ratio v, as in the case of Nevada sand [49]). The foregoing trial-
calibration requires monotonic compression tests (at least and-error simulations are not affected by the exact value of
two) that initiate from different initial conditions and are the remaining constant No, since the fabric evolution scalar hf
based on correlating the deviatoric stress ratios at the peak does not change considerably during such paths.
and at the phase transformation, respectively, to the values of (h) fabric evolution intensity constant, No: The calibration of No is
c at which they are attained. based on fitting large-strain cyclic shear tests and preferably
(d) Shear modulus degradation constants, g1 and a1: As shown in cyclic undrained simple shear tests if the end goal is the
[8], their values may be estimated from resonant column numerical simulation of liquefaction related phenomena. In
tests, by fitting the Ramberg–Osgood relation to the G/Gmax particular, the value of No is pinpointed by fitting the rate of
versus cyclic shear strain data. For this purpose, the value of excess pore pressure buildup during cyclic loading, as well as
g1 should be correlated to the volumetric threshold shear the number of cycles to reach initial liquefaction. Obviously,
strain a~ tv, which has generally small variability and for non- the more cyclic tests are used for fine-tuning No the better, but
plastic soils ranges from 0.0065% to 0.025% [32]. Then, the in lack of relevant data one may resort to the literature. For
value of a1 practically corresponds to the value of G/Gmax that example, the rate of excess pore pressure buildup in element
corresponds to the selected a~ 1 value of cyclic shear strain on tests may be modeled according to Seed and Booker [58], while
the basis of the resonant column data. In lack of pertinent the number of cycles to liquefaction may be estimated on the
data for the sand at hand, the literature may be successfully basis of liquefaction resistance curves, preferably from shaking
used (e.g. Vucetic and Dobry [50], Ishibashi and Zhang [61]), table (e.g. [51]) or cyclic simple shear tests (e.g. [62]).
or a value ranging between 0.55 and 0.75 is usually
appropriate [29]. For monotonic loading, the a1 is always set
to 1.0, which makes Eq. (19) purely hypoelastic (T¼1) and
independent of g1. References
(e) Dilatancy constant, Ao: Calibration of this constant requires good
quality (small sampling interval) stress-dilatancy data, preferably [1] Prevost JH. Plasticity theory for soil stress-strain behaviour. Journal of
Engineering Mechanics, ASCE 1978;104(5):1177–94.
from drained triaxial tests. In such a case, one should plot the [2] Cubrinovski M, Ishihara K. State concept and modified elastoplasticity for
ever-current value of the deviatoric stress ratio versus the sand modelling. Soils Foundations 1998;38(4):213–25.
K.I. Andrianopoulos et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 895–911 911

[3] Yang Z, Elgamal A, Parra E. Computational model for cyclic mobility and [34] Dafalias YF, Popov EP. A model of nonlinearly hardening materials for
associated shear deformation. Journal of Geotechnical and Geoenvironmental complex loadings. Acta Mechanica 1975;21(3):173–92.
Engineering, ASCE 2003;129(12):1119–27. [35] Itasca. Fast Lagrangian Analysis of Continua. Minneapolis, Minnesota: Itasca
[4] Yang Z, Elgamal A. Multi-surface cyclic plasticity sand model with Lode angle Consulting Group Inc.,; 2005.
effect. Geotechnical and Geological Engineering 2008;26:335–48. [36] Andrianopoulos KI. Numerical modeling of static and dynamic behavior of
[5] Poorooshasb HB, Pietruszczak S. A generalized flow theory for sand. Soils and elastoplastic soils, Doctorate Thesis, 2006; Department of Geotechnical
Foundations 1986;26(2):1–15. Engineering, School of Civil Engineering, National Technical University of
[6] Manzari MT, Dafalias YF. A critical state two-surface plasticity model for Athens (in Greek).
sands. Geotechnique 1997;47(2):255–72. [37] Andrianopoulos KI, Papadimitriou AG, Bouckovalas GD. Explicit integration of
[7] Gajo A, Wood DM. A kinematic hardening constitutive model for sands: the bounding surface model for analysis of earthquake soil liquefaction.
multiaxial formulation. International Journal for Numerical and Analytical International Journal for Numerical and Analytical Methods in Geomechanics,
Methods in Geomechanics 1999;23:925–65. 2009 doi:10.1002/nag.875.
[8] Papadimitriou AG, Bouckovalas GD, Dafalias YF. Plasticity model for sand [38] Andrianopoulos KI, Papadimitriou AG, Bouckovalas GD. Use of a new
under small and large cyclic strains. Journal of Geotechnical and Geoenviron- bounding surface model for the analysis of earthquake-induced liquefaction
mental Engineering ASCE 2001;127(11):973–83. phenomena. In: Proceedings of the Fourth International Conference on
[9] Papadimitriou AG, Bouckovalas GD. Plasticity model for sand under small and Earthquake Geotechnical Engineering; paper No. 1443, 2007.
large cyclic strains: a multiaxial formulation. Soil Dynamics and Earthquake [39] Sloan SW. Substepping schemes for the numerical integration of elastoplastic
Engineering 2002;22:191–204. stress – strain relations. International Journal of Numerical Methods on
[10] Loukidis D, Salgado R. Modeling sand response using two-surface plasticity. Engineering, 1987;24:893–911.
Computers & Geotechnics 2009;36(1–2):166–86. [40] Sloan SW, Abbo AJ, Sheng D. Refined explicit integration of elastoplastic models
[11] Hachiguchi K, Ueno M. Elastoplastic constitutive laws of granular materials, with automatic error control. Engineering Computations 2001;18(1/2):121–54.
Proceedings, Constitutive Equations of Soils. Ninth International Conference [41] Martin J, Cundall PA. Mixed discretisation procedure for accurate solution of
of Soil Mechanics and Foundation Engineering 1977;Sp. Session 9:73–82. plasticity problems. International Journal for Numerical and Analytical
[12] Wang Z-L, Dafalias YF, Shen CK. Bounding surface hypoplasticity model for Methods in Geomechanics 1982;6:129–39.
sand. Journal of Engineering Mechanics, ASCE 1990;116(5):983–1001. [42] Ishihara K, Tatsuoka F, Yasuda S. Undrained deformation and liquefaction of
[13] Li XS, Dafalias YF, Wang ZL. State-dependent dilatancy in critical-state sand under cyclic stresses. Soils and Foundations 1975;15(1):29–44.
constitutive modelling of sand. Canadian Geotechnical Journal 1999;36(4): [43] Li XS, Dafalias YF. A constitutive framework for anisotropic sand including
599–611. non-proportional loading. Geotechnique 2004;54(1):41–55.
[14] Li XS. A sand model with state-dependent dilatancy. Geotechnique 2002;52(3): [44] Andrianopoulos KI, Papadimitriou AG, Bouckovalas GD. Bounding surface
173–86. models of sands: pitfalls of mapping rules for cyclic loading. In: Proceedings
[15] Pastor M, Zienkiewicz OC, Chan AHC. Generalized plasticity and the modeling of the 11th international conference of IACMAG, Torino, June 19–24, 2005,
of soil behaviour. International Journal for Numerical and Analytical Methods Vol. 1, p. 241–248.
in Geomechanics 1990;14(3):151–90. [45] Dowell M, Jarratt P. The Pegasus method for computing the root of an
[16] Iai S, Matsunaga Y, Kameoka T. Strain space plasticity model for cyclic equation. BIT 1972;12:503–8.
mobility. Soils and Foundations 1992;32(2):1–15. [46] Dafalias YF, Papadimitriou AG, Li XS. Sand plasticity model accounting for
[17] Ling HI, Liu H. Pressure-level dependency and densification behavior of sand inherent fabric anisotropy. Journal of Engineering Mechanics, ASCE
through generalized plasticity model. Journal of Engineering Mechanics, ASCE 2004;130(11):1319–33.
2003;129(8):851–60. [47] Ladd CC, Foott R, Ishihara K, Schlosser F, Poulos HG. Stress-deformation and
[18] Ling HI, Yang S. Unified sand model based on the critical state and strength characteristics, State-of-the-art Report. In: Proceedings of the ninth
generalized plasticity. Journal of Engineering Mechanics 2006;132(12): international conference on soil mechanics and foundation engineering,
1380–91. Tokyo, 1977. 2: p. 421–494.
[19] Park S-S, Byrne PM, Practical constitutive model for soil liquefaction. In: [48] Nemat-Nasser S, Tobita Y. Influence of fabric on liquefaction and densifica-
Proceedings of the ninth international symposium on numerical models in tion potential of cohesionless sand. Mechanics of Materials 1982;1:43–62.
geomechanics (NUMOG IX), 2004. p. 181–186. [49] Arulmoli K, Muraleetharan KK, Hossain MM, Fruth LS. VELACS verification of
[20] Lopez-Querol S, Blazquez R. Liquefaction and cyclic mobility model for liquefaction analyses by centrifuge studies -Laboratory Testing Program –
saturated granular media. International Journal for Numerical and Analytical Soil Data Report, Research Report, The Earth. Technology Corporation, 1992.
Methods in Geomechanics 2006;30:413–39. [50] Vucetic M, Dobry R. Effect of soil plasticity on cyclic response. Journal of
[21] Yang Y, Muraleetharan KK, Yu HS. A middle surface concept (MSC) model for Geotechnical Engineering, ASCE 1991;117(1):89–107.
saturated sands in general stress space. International Journal for Numerical [51] DeAlba P, Seed HB, Chan CK. Sand liquefaction in large-scale simple shear
and Analytical Methods in Geomechanics 2006;30:389–412. tests. Journal of the Geotechnical Engineering Division 1976;102(9):909–27.
[22] Park I-J, Desai CS. Cyclic behavior and liquefaction of sand using disturbed [52] Seed HB, Booker JR. Stabilization of potentially liquefiable sand deposits
state concept. Journal of Geotechnical and Geoenvironmental Engineering, using gravel drains. Journal of the Geotechnical Engineering Division, ASCE
ASCE 2000;126(9):834–46. 1977;103(7):755–68.
[23] Aubry D, Hujeux J-C, Lassoudiere F, Meimon Y. A double memory model with [53] Manzari MT, Arulanandan K. Numerical predictions for Model No. 1. In:
multiple mechanisms for cyclic behaviour. In: Proceedings of the interna- Proceedings of the international conference on verification of numerical
tional symposium on numerical models, Balkema, 1982. p. 3–13. procedures for the analysis of soil liquefaction problems, Davis, CA, 17–20
[24] Bauer E, Wu W. A hypoplastic model for granular soils under cyclic loading. October, 1994. 1: p. 179–185.
In: Kolymbas D, editor. Modern Approaches to Plasticity. Elsevier; 1993. p. [54] Arulanandan K, Sybico J. Post liquefaction settlement of sands, Proceedings,
247–58. Wroth Memorial Symposium. England: Oxford University; 1992.
[25] Houlsby GT, Mortara G. A Continuous Hyperplasticity Model for Sands under [55] Lacy SJ. Numerical prediction for Model No. 1. In: Proceedings of the international
Cyclic Loading. In: Proceedings of the International Conference on Cyclic conference on verification of numerical procedures for the analysis of soil
Behaviour of Soils and Liquefaction Phenomena, Bochum, Germany, 31 liquefaction problems, Davis, CA, 17–20 October, 1994. 1: p. 153–168.
March–2 April 2004, 21–26, Balkema. [56] Krstelj I, Prevost JH. Experimental results of Model No. 12. In: Proceedings of
[26] Been K, Jefferies MG. A state parameter for sands. Geotechnique 1985;35(2): the international conference on verification of numerical procedures for the
99–112. analysis of soil liquefaction problems, Davis, CA, 17–20 October, 1994. 1:
[27] Jefferies MG. Nor-Sand: a simple critical state model for sand. Geotechnique p. 1007–1017.
1993;43(1):91–103. [57] Carnevalle R, Elgamal A. Experimental results of RPI centrifuge Model No. 12.
[28] Dafalias YF, Manzari MT. Simple plasticity sand model accounting for fabric In: Proceedings of the international conference on verification of numerical
change effects. Journal of Engineering Mechanics, ASCE 2004;130(6):622–34. procedures for the analysis of soil liquefaction problems, Davis, CA, 17–20
[29] Papadimitriou AG. Elastoplastic modelling of monotonic and dynamic October, 1994. 1: p. 1019–1026.
behavior of soils, Doctorate Thesis, Department of Geotechnical Engineering, [58] Farrel T, Kutter B. Experimental results of Model No. 12. In: Proceedings of the
Faculty of Civil Engineering, National Technical University of Athens, June, international conference on verification of numerical procedures for the analysis
1999, (in Greek). of soil liquefaction problems, Davis, CA, 17–20 October, 1994. 1: p. 1027–1034.
[30] Papadimitriou AG, Bouckovalas GD, Dafalias YF. Use of elasto-plasticity to [59] Popescu R, Prevost JH. Numerical class ‘‘A’’ predictions for Model Nos 1, 2, 3,
simulate cyclic sand behavior. In: Proceedings of the second international 4a, 4b, 6, 7, 11 & 12, Proceedings, International Conference on Verification of
conference on earthquake geotechnical engineering, Lisbon, 1999. 1: p. Numerical Procedures for the Analysis of Soil Liquefaction Problems, Davis,
125–130. CA, 17-20 October, 1994, 2: 1105–1207.
[31] Dafalias YF, Manzari MT. Modeling of fabric effect on the cyclic loading [60] Prevost JH, Krstelj I, Popescu R. Overview of experimental results for
response of granular soils, In: Proceedings, 13th ASCE Engineering Mechanics centrifuge model No. 12. In: Proceedings of the international conference on
Specialty Conference, 1999, Baltimore (in CD-ROM). verification of numerical procedures for the analysis of soil liquefaction
[32] Vucetic M. Cyclic threshold shear strains in soils. Journal of Geotechnical and problems, Davis, CA, 17–20 October, 1994. 2: p. 1619–1634.
Geoenvironmental Engineering, ASCE 1994;120(12):2208–28. [61] Ishibashi I, Zhang X. Unified dynamic shear moduli and damping ratios of
[33] Dafalias YF. On cyclic and anisotropic plasticity: (i) a general model including sand and clay. Soils and Foundations 1993;33(1):182–91.
material behavior under stress reversals, (ii) anisotropic hardening for [62] Troncoso JH. Failure risks of abandoned tailing dams. In: Proceedings of the
initially orthotropic materials, PhD Thesis, 1975, University of California, international symposium on safety and rehabilitation of tailing dams, ICOLD,
Berkeley. Paris, 1990. p. 82–89.

You might also like