You are on page 1of 15

Cyclic Large Strain and Induced Pore Pressure

Models for Saturated Clean Sands


K. Onder Cetin, M.ASCE1; and H. Tolga Bilge, S.M.ASCE2

Abstract: Semiempirical probabilistic models are described to assess cyclic large strain and induced excess pore-water pressure responses of
fully saturated clean sands. For this purpose, available cyclic simple shear and triaxial tests were compiled and studied. The resulting r u versus
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

γ, and γ versus N databases are composed of 101 and 84 cyclic test data, respectively. Key parameters of the proposed ru and γ models are
defined as critical shear strain, relative density, effective confining stress, and equivalent number of loading cycles. Consistent with the
maximum likelihood methodology, model coefficients were estimated by maximizing the likelihood function. For comparison purposes,
the compiled database was again used to evaluate the performance of existing r u models. Both for comparison and calibration purposes,
for each framework, two separate sets of limit-state models were used: model implemented with (1) the original and (2) the updated model
coefficients. The model performances are assessed by simple statistics (i.e., mean and standard deviation) of residuals. It is concluded that
existing models produce inconsistently biased predictions that vary in the range of 2.5 to 70%. The successes of the proposed and existing
models are also assessed for the validation database composed of additional 10 cyclic test results. In addition to (1) repeated improved
predictions, (2) differentiating contractive or dilative cyclic soil responses, and (3) incorporation of strain-dependent modulus degradation
effects, the main advantage of the proposed methodology is the probabilistic nature of model predictions, which enables the incorporation of
the model uncertainty into pore pressure generation predictions. DOI: 10.1061/(ASCE)GT.1943-5606.0000631. © 2012 American Society
of Civil Engineers.
CE Database subject headings: Strain; Pore-water; Water pressure; Triaxial tests; Shear tests; Sand (soil type).
Author keywords: Shear strain; Pore-water pressure; Cyclic response; Cyclic triaxial tests; Cyclic simple shear tests.

Introduction review of these approaches will be subsequently presented, a pre-


vious discussion on some of the critical issues is preferred.
During earthquakes, shear stresses attributable to stress-wave 1. Over the years, the reliability of cyclic testing has improved
propagation induce cyclic shear strains leading to rearrangement and the number of available test data has increased. Thus, ex-
of soil particles and generation of excess pore-water pressure. isting cyclic-testing data-based pore-water pressure generation
Undrained loading-induced elevated pore-water pressure reduces models deserve a revisit.
the effective confining stress-dependent soil stiffness, which in turn 2. Excess pore pressure assessments founded on implemented
triggers the vicious cycle of further strain and excess pore pressure constitutive models generally require a large number of input
accumulation. In the extreme, excess pore-water pressure ap- parameters, which are difficult to be determined under the ab-
proaches to total stress, defining the onset of cyclic soil liquefac- sence of advance laboratory testing and interpretation of their
tion. Understanding this outlined cyclic pore pressure and corollary results. Thus, empirically, or semiempirically based models
straining response of saturated cohesionless soils has been a major have and will continue to establish the state of practice.
concern of geotechnical and earthquake engineering since the very 3. The accuracy and uncertainty with regard to the predictions of
early days of the professions. existing models have not been fully understood and quantified
A number of methodologies, which are grouped under the head- yet, with the exception of a limited number of attempts
ings of (1) stress based (e.g., Seed et al. 1975; Booker et al. 1976), (e.g., Chameau and Clough 1983; Polito et al. 2008).
(2) strain-based (e.g., Martin et al. 1975; Dobry et al. 1985), 4. Despite its numerous advantages, during strain-controlled
(3) energy-based (e.g., Davis and Berrill 1982; Green et al. tests, difficulties in testing specimens beyond shear strain le-
2000), (4) plasticity theory based (e.g., Prevost 1985; Elgamal et al. vels of 1% inevitably diverted researchers to stress-controlled
2003), (5) other (e.g., Oda et al. 2001; Eggzelos 2004), were pro- testing for developing an understanding of coupled pore pres-
posed to assess quantitatively these responses. Although a detailed sure response beyond 1% shear strain levels.
On the basis of these observations, the main motivation of this
1
Professor, Dept. of Civil Engineering, Middle East Technical Univ., study is defined as to (1) assess the validity of existing cyclically
06531, Ankara, Turkey (corresponding author). E-mail: onder@ce.metu induced pore-water pressure models and the accuracy of their pre-
.edu.tr dictions, and (2) develop a practical-to-use and a robust semiem-
2
Graduate Researcher, Dept. of Civil Engineering, Selcuk Univ., Konya, pirical model to assess cyclic straining and induced excess pore
Turkey. pressure responses of fully saturated clean sands. For this purpose,
Note. This manuscript was submitted on December 17, 2009; approved
results of stress-controlled cyclic triaxial (CTXT) and simple shear
on September 7, 2011; published online on September 9, 2011. Discussion
period open until August 1, 2012; separate discussions must be submitted tests (SSTs) were compiled and studied. Following a brief review of
for individual papers. This paper is part of the Journal of Geotechnical and the available literature, data compilation efforts are discussed. As
Geoenvironmental Engineering, Vol. 138, No. 3, March 1, 2012. ©ASCE, part of the literature review, emphasis will be given only to method-
ISSN 1090-0241/2012/3-309–323/$25.00. ologies, which are more widely referred to and used in practice.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 309

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Review of Existing Models the change in volumetric strain attributable to cyclic shearing.
 r and Δεvd are defined as follows:
The variables E
In addition to the proposed model, for comparison purposes, cyclic
pore pressure responses of clean sand specimens were also assessed r ¼ ðσ0v Þ1m
E ð4Þ
by widely used stress-based methods of Seed et al. (1975) and m · k2 · ðσ0v0 Þnm
Polito et al. (2008); strain-based methods of Martin et al. (1975),
Dobry et al. (1985), and Byrne (1991); and energy-based model of c3 · ε2vd
Green et al. (2000). However, before then, a brief summary on the Δεvd ¼ c1 · ðγ  c2 · εvd Þ þ ð5Þ
γ þ c4 · εvd
evolution of these methods will be presented.
where σ0v and σ0v0 , respectively = vertical and initial vertical effec-
Stress-Based Models
tive stresses; γ = induced cyclic shear strain; εvd = accumulated
On the basis of stress-controlled CTXT results on Monterey and volumetric strain; and ci , m, n, k 2 = model coefficients. For loose
Sacramento River sands, Lee and Albaisa (1974) proposed an em- crystal silica sands (DR ≈ 45%), model coefficients, c1 through c4 ,
pirical relationship between excess pore-water pressure ratio (r u ) were recommended to be used as 0.80, 0.79, 0.45, and 0.73, respec-
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

and cycle number ratio (N=N liq ), which was claimed to be insensi- tively. The following correction was proposed for relative density
tive to soil type, relative density (DR ), initial effective confining states other than 45%:
stress (σc;0 ), and number of cycles to liquefaction (N liq ). On the ba-
sis of De Alba et al. (1975) experimental data, Seed et al. (1975) ðΔεvd ÞDR1 ¼ R · ðΔεvd ÞDR ¼45 ð6Þ
developed an empirical, closed-form solution as given in Eq. (1)
     where, on the basis of the findings of Silver and Seed (1971), R = a
1 1 1 N 1=α function of DR for crystal silica sands, presented as follows.
ru ¼ þ sin 2 · 1 ð1Þ
2 π N liq
R ¼ 0:00031 · ð100  DR Þ2 þ 0:062 for 45 < DR < 80% ð7Þ
where α = recommended to be functions of soil properties and test
conditions with an average value of 0.7; and N = number of equiv- Martin et al. (1975) formulated the cyclic shear strain (γ) response,
alent uniform loading cycles. The proposed expression was as presented in Eqs. (8)–(10)
subsequently simplified by Booker et al. (1976). Chameau and τ ·a
Clough (1983) adopted the same functional form in their probabil- γ ¼ pffiffiffiffiffi0 ð8Þ
istically based model. Wang and Kavazanjian (1985) benefitted σv  τ · b
from this approach to predict pore-water pressure response under
εvd
transient loads. Liyanapathirana and Poulos (2002) implemented a ¼ A1  ð9Þ
it in a finite-element-based model. More recently, Polito et al. A2 þ A3 · εvd
(2008) statistically reevaluated Seed et al.’s (1975) methodology
εvd
and concluded that the model coefficient α needs to be estimated b ¼ B1  ð10Þ
as a function of cyclic stress ratio (CSR), fines content (FC), and DR , B2 þ B3 · εvd
given as follows:
where τ = cyclic shear stress; and A1 , A2 , A3 , B1 , B2 , B3 = model
α ¼ 0:01166 · FC þ 0:007397 · DR þ 0:01034 · CSR þ 0:5058 coefficients.
ð2Þ On the basis of Martin et al.’s (1975) findings, Finn et al. (1976,
1977) developed an effective-stress-based constitutive model,
Eq. (2) is suggested to be used for soils with FC < 35%. For clean which has been widely used for ground response and liquefaction
sands (FC ¼ 0), because of the equation’s relative insensitivity to assessments. Subsequently, Byrne (1991) simplified this approach
CSR, DR remains as the major controlling model parameter. For the and recommended a new incremental volumetric strain definition
given coefficients, the value of α is estimated to vary in the range of as given in Eq. (11)
0.73 to 1.14.  
Original and modified versions of Seed et al.’s approach have Δεvd ε
¼ c1 · exp c2 · vd ð11Þ
stayed as relatively simpler and widely used cyclic excess pore γ γ
pressure generation models. However, (1) difficulties in defining
and estimating N liq , (2) lack of effective differentiation between where c1 and c2 are defined as follows:
cyclic liquefaction and mobility responses, and (3) the need to con-
vert transient earthquake excitations to equivalent harmonic cycles c1 ¼ 7;600 · ðDR Þ2:5 ð12Þ
are reported as major limitations.
c2 ¼ 0:4=c1 ð13Þ
Strain-Based Models
Byrne (1991) model was subsequently implemented in dynamic-
Martin et al. (1975) proposed an alternative framework on the basis
coupled stress-flow finite difference software Fast Lagrangian
of the results of strain-controlled cyclic tests performed on dry
Analysis of Continua (FLAC). Both Finn et al. (1976, 1977)
sands. The authors derived semiempirical relationships between
and Byrne (1991) models were based on physical laws, and incor-
the tendencies of volumetric straining and excess pore-water pres-
poration of them into commercially available codes has increased
sure generation of dry and fully saturated sands subjected to strain
their use. However, the requirement of a large number of model
cycles. The following expression was proposed:
parameters is the cost of these models.
Δu ¼ E
 r · Δεvd ð3Þ On another path, Dobry et al. (1982) attempted to link cyclic
shear straining to pore-water pressure generation. The authors de-
where Δu = increase in excess pore-water pressure; E r = veloped a unique relationship among ru , cyclic shear strain ampli-
tangent modulus of the one-dimensional (1D) unloading curve tude, and the number of loading cycles on the basis of undrained,
corresponding to the initial effective vertical stress; and Δεvd is strain-controlled CTXT results. A threshold shear strain level of

310 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


102 %, below which applied shear strains do not trigger significant dεv = incremental vertical strain; σ0h = horizontal effective stress;
excess pore-water pressures, was proposed. Subsequently, on the dεh = incremental radial strain; τ vh = horizontal shear stress acting
basis of the results of Dobry et al. (1982) and the model of Martin on a plane having a vertical normal vector; dεvh = incremental shear
et al. (1975), Dobry et al. (1985) proposed an alternative method strain resulting from τ vh ; τ hv = vertical shear stress acting on a plane
by defining cyclic excess pore-water pressure as a function of un- having a horizontal normal vector; and dεhv = incremental shear
drained unloading-reloading modulus of the sand skeleton and strain resulting from τ hv . For undrained cyclic triaxial and simple
volumetric strain tendency per cycle, which was assumed to be con- shear conditions, respectively, W s in Eq. (16) can be solved numeri-
stant for strain-controlled testing. The proposed methodology was cally as follows:
subsequently modified by Vucetic and Dobry (1986), and the fol-
lowing closed-form solution was proposed: 1 Xn1
Ws ¼ 0 ½ðσ þ σd;i Þ · ðεa;iþ1  εa;i Þ ð17Þ
2σ0 i¼1 d;iþ1
p · f · N c · F · ðγc  γtvp
Þb
r u;N ¼ ð14Þ
1 þ f · N c · F · ðγc  γtvp Þb
1 X n1
Ws ¼ ½ðτ þ τ i Þ · ðγiþ1  γi Þ ð18Þ
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

where γc = cyclic shear strain amplitude; and γtvp = volumetric 2σ00 i¼1 iþ1
threshold shear strain, below which no significant pore-water pres-
sure is generated and is usually between 0.01 and 0.02% for most where n = number of load increments applied to specimen; σd;i and
sands (Dobry et al. 1982). Model coefficient f can be assumed as 1 σd;iþ1 = applied deviator stresses at load increments “i” and “i þ 1,”
or 2, depending on whether pore pressures are induced by one-or respectively; εa;i and εa;iþ1 = axial strains at load increments “i”
two-directional shaking, whereas the rest of coefficients F, p, and b and “i þ 1,” respectively; and τ i , τ iþ1 and γi , γiþ1 = applied
are obtained by laboratory data-based fitting attempts. This model shear stresses and shear strains at load increments “i” and “i þ 1,”
was subsequently utilized by Matasovic and Vucetic (1993), respectively.
Matasovic (2006), and Hashash (2009) in their effective-stress- Green et al. (2000) suggested a procedure for the determination
based nonlinear ground response modeling softwares D-Mod, of PEC from cyclic test data. Subsequently, Polito et al. (2008) ex-
D-Mod_2, Deepsoil, respectively. pressed PEC as a function of DR and FC given as follows:
Recently, strain-based approaches were revisited by Hazirbaba
lnðPECÞ ¼ expðc3 · DR Þ þ c4 for FC < 35% ð19Þ
(2005), and the effects of nonplastic fines on pore pressure gener-
ation were investigated. In addition to the major limitation of not where c3 and c4 = model coefficients and estimated as 0.0139 and
being able to reliably test specimens at larger strain levels (> 1%), 1:021, respectively. The standard deviation of this model was re-
during strain-controlled tests, unique matching of shear strain and ported as 0.6591. Green et al.’s (2000) procedure by adopting Polito
r u responses independent of relative state (i.e., lack of differentiat- et al.’s (2008) expression for the estimation of PEC is used for com-
ing the dilative and contractive responses) and applied number of parison purposes and will be simply referred to as Green Mitchell
strain cycles limit wider use of the strain-controlled testing-based Polito (GMP) model from this point on.
models. In the following sections, these limitations will be dis- Energy-based models do not require the use of a load conversion
cussed further. scheme to estimate equivalent harmonic loading cycles of a
transient earthquake excitation, and can conveniently capture the
Energy-Based Models effects of stress (or strain) histories on actual r u versus γ responses,
A separate leg of research studies has evolved in strain-energy con- adversely require, a priori, either the complete stress-strain history
cepts. These studies attempted to link the change in excess pore or a constitutive model that enables the estimation of induced strain
pressure to the energy dissipated per unit volume of soil during levels under applied stress cycles.
the same cycle of loading. Nemat-Nasser and Shokooh (1979) first
established the governing equations relating energy dissipation to Plasticity Theory-Based Models
volumetric densification of dry sands and pore pressure generation A number of researchers (Prevost 1985; Pastor et al. 1990; Wang
of fully saturated sands. On the basis of their theoretical work, suc- et al. 1990; Ishihara 1993; Fukutake and Ohtsuki 1995; Elgamal
ceeding attempts appeared mostly focusing on the development of et al. 2003; Park and Byrne 2004), benefiting from different
empirical models benefiting from experimental data (e.g., Davis branches of theory of plasticity, developed effective stress-based
and Berrill 1982, 2001; Yamazaki et al. 1985; Law et al. 1990; models capable of predicting cyclic excess pore pressure response.
Hsu 1995; Wang et al. 1997; Green et al. 2000). An extensive sum- These constitutive model-based assessments have been increas-
mary of previous studies, along with a new set of calibration param- ingly implemented in various numerical analyses software. How-
eters, was presented by Green (2001). Additionally, Green et al. ever, for the most conventional projects, adequate laboratory or
(2000) proposed the following model: field test data to evaluate accurately the model input parameters
rffiffiffiffiffiffiffiffiffi are absent. Sourcing from either budget or time constraints, this
Ws reduces the success of the overall numerical schemes, independent
ru ¼ ≤1 ð15Þ
PEC of how precise the constitutive model predictions are. This will stay
as a major obstacle against wider use of these models.
where W s = energy dissipated per unit volume of soil divided by the
initial effective mean confining stress; and pseudoenergy capacity Other Studies
(PEC) = serves the purpose of a calibration parameter. Increments
in W s can be related to stress conditions, and increments in strain by There exist a number of other studies and methods aiming to assess
the excess pore-water pressure response of fully saturated clean
ðσ0v dεv þ 2σ0h dεh þ τ vh dγvh þ τ hv dγhv Þ sands. Because of page limitations, only a few will be referred
dW s ¼ ð16Þ
σ0m0 to. Marcuson et al. (1990) used r u to estimate the reduction of soil
stiffness during earthquake loading. Benefiting from the works of
where dW s = incremental dissipated energy normalized by the Tokimatsu and Yoshimi (1983), Evans (1987), and Hynes (1988),
initial effective mean stress (σ0m0 ); σ0v = vertical effective stress; they attempted to develop a relationship between residual ru value

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 311

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


and factor of safety against liquefaction (FSliq ). Oda et al. (2001) Table 1. Databases Reviewed
linked the increment in pore-water pressure at the end of the first Number of test data
cycle to the number of cycles required to cause a double-amplitude
Number Reference γ versus N ru versus γ
axial strain of 5% in CTXTs, regardless of DR and CSR. Sub-
sequently, Wu et al. (2003) improved this approach on the basis 1 Seed et al. (1975) — 2
of cyclic SST results performed on Monterey sand. They developed 2 Prakash (1981) — 1
a closed-form solution, which relates the number of cycles to 3 Das (1983) — 1
induce 6% double-amplitude shear strain (adopted as the initial 4 Ishihara (1985) — 2
liquefaction criterion by them) and the increment in pore-water
5 Riemer et al. (1994) — 6
pressure at the end of the first cycle. Wu et al. (2003) also reeval-
6 Evans and Rollins (1999) — 2
uated the Lee and Albaisa (1974) approach and concluded that
their upper limit pore-water pressure generation curve does not 7 Ishihara (1996) 1 1
serve as an upper cap for the observed laboratory responses. Sim- 8 Park and Desai (2000) 3 3
ilarly, Eggzelos (2004) developed an empirically based pore-water 9 Ueng et al. (2004) 1 —
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

pressure model by using the r u value at the end of the first loading 10 Vaid and Chern (1983) 2 2
cycle as a model input parameter, and extended this work to assess 11 Arulmoli et al. (1992) 4 4
pore-water pressure generation behavior of gravels. Boulanger et al. 12 Sriskandakumar (2003) 2 2
(1991) focused on the effects of the presence and direction of initial 13 Ghionna and Porcino (2006) 1 1
static shear stresses on pore-water pressure response of clean sands 14 Porcino and Caridi (2007) 1 1
by performing a series of SSTs. Kammerer (2002) performed 15 Boulanger et al. (1991) 13 13
additional bidirectional SSTs and concluded that r u values typically
16 Wu et al. (2003) 55 50
increase more rapidly under multidirectional loading than in
17 Bilge (2005) 16 38
“equivalent” unidirectional tests. However, the complex nature
of the problem made it difficult to propose a quantitative model Total 99 129
addressing the effects of the bidirectional loading and the presence
of initial static shear stresses. More recently, Chang et al. (2007)
revisited previous experimental efforts on pore-water pressure laboratory tests were adopted from Boulanger et al. (1991) database
modeling, and addressed several factors, such as (1) upward for r u versus γ and γ versus N databases, respectively. In their study,
seepage, (2) formation of water film, (3) stratified soil effects, and results of stress-controlled SSTs performed on moist tamped
(4) alteration of ground motion, which are believed to affect cyclic Sacramento River sand were presented. Specimens were consoli-
shearing of soils in situ. Considering these factors, they have high- dated under an effective vertical stress of 207 kPa, and cyclic load-
lighted the significance of monitoring pore-water pressure behavior ing was applied at 0.2 Hz. Fifty out of 101 and 55 out of 84 cyclic
in situ. laboratory tests were adopted from Wu et al. (2003) database for r u
versus γ and γ versus N databases, respectively. The results of
stress-controlled SSTs performed on wet pluviated Monterey num-
Proposed Model ber 0=30 sand were presented in their study. Specimens were first
K 0 -consolidated to an effective vertical stress of 40, 80, or 180 kPa,
Developing a semiempirical model naturally requires fulfilling the and cyclic loading was applied at 0.1 Hz. Similarly, the results of 38
following steps: (1) data compilation, filtering, and consolidation; and 16 stress-controlled CTXTs performed on poorly graded (SP)
(2) defining the variables (descriptors) and the mathematical form Kizilirmak River sand specimens were compiled from Bilge (2005)
(limit state) of the problem (i.e., model development); (3) assessing for r u versus γ and γ versus N databases, respectively. Samples
the model coefficients; and (4) validation and calibration. These were reconstituted in the laboratory by either dry pluviation or
will be discussed next. moist-tamping method depending on the target relative density,
and then they were consolidated isotropically under a confinement
Database Compilation Efforts pressure of 100 kPa. The frequency of cyclic loading was selected
Efforts aiming to develop a semiempirical or empirical model, or to as 1 Hz. Maximum double-amplitude shear strain values were
validate and calibrate existing models require the compilation of estimated by simply multiplying the maximum double-amplitude
high-quality database. For this purpose, as listed in Table 1, test axial strain values by a value of 1.50, which was estimated on
results reported by various researchers were compiled and studied. the basis of the elasticity theory for undrained conditions (i.e.,
Databases listed in 1 through 9 of Table 1 were eliminated because for Poisson’s ratio of 0.5). A summary of test conditions [test type,
of lack of complete documentation of both pore pressure and confining stresses, σ0v;0 or (σ0c;0 ), applied uncorrected test CSR
shear strain responses at every cycle of shearing. As will be sub- values], along with specimen relative density and resulting maxi-
sequently discussed, Vaid and Chern (1983), Arulmoli et al. (1992), mum r u and γ values, is presented in Table 2.
Sriskandakumar (2003), Ghionna and Porcino (2006); and Porcino Distribution of specimens with respect to DR is relatively
and Caridi (2007) databases were reserved for validation purposes. uniform, whereas the resulting databases, more pronounced in
Hence, the proposed models were developed by using Boulanger the r u versus γ database, are concluded to be biased with tests of
et al. (1991), Wu et al. (2003), and Bilge (2005) databases. As pre- initial confining stresses of 100 kPa. This sampling bias will be
sented in Table 2, because of the fact that the proposed γ versus N subsequently addressed and corrected in the assessment of the pro-
model requires the input of maximum cyclic shear strain at the end posed model coefficients.
of the 20th loading cycle, if r u and γ responses are available until For comparison purposes, the compiled database is presented in
the end of the maximum number of loading cycles, N max , equals to the r u versus N=N liq domain along with the upper and lower boun-
or greater than 20, then these test results are included in both r u daries of Seed et al. (1975) and Polito et al. (2008) frameworks as
versus γ, and γ versus N databases. If not, they are only used in given in Fig. 1. For the purpose of achieving a r u value of 1.0 at
the r u versus γ database. Thus, 13 out of 101 and 13 out of 84 cyclic N=N liq ¼ 1:0, N liq is defined as the number of cycles to the first

312 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Table 2. Summary of the Test Results Used for the Development of the Proposed Models
Data source Test type Test number DR (%) σ0v;0 (or σ0c;0 ) (kPa) Cyclic stress ratio N max r u;max γmax (%)
Boulanger et al. (1991) Simple shear test 1 54 207 0.125 20 0.43 0.20
2 54 207 0.160 20 1.00 30.00
3 56 207 0.140 20 0.43 0.40
4 55 207 0.160 20 1.00 4.10
5 56 207 0.150 20 0.35 0.40
6 54 207 0.155 20 0.85 1.80
7 56 207 0.150 20 1.00 12.00
8 56 207 0.150 20 0.95 6.00
9 56 207 0.170 20 0.40 0.20
10 36 207 0.100 20 0.40 0.25
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

11 35 207 0.100 20 0.37 0.35


12 35 207 0.130 20 0.45 0.15
13 35 207 0.110 20 0.45 0.20
Wu et al. (2003) Simple shear test 14 58 33 0.324 9 1.00 25.00
15 55 80 0.166 6 1.00 23.80
16 45 85 0.249 8 1.00 33.00
17 43 81 0.200 11 1.00 34.20
18 56 85 0.339 7 1.00 40.00
19 56 85 0.247 22 1.00 30.80
20 77 77 0.364 14 1.00 15.40
21 79 79 0.406 15 1.00 10.40
22 71 80 0.340 19 1.00 11.50
23 75 85 0.447 17 1.00 14.10
24 50 75 0.220 10 1.00 17.60
25 44 81 0.151 60 1.00 30.90
26 31 82 0.181 9 1.00 30.00
27 61 32 0.236 42 1.00 33.00
28 45 32 0.287 8 1.00 36.00
29 33 39 0.226 9 1.00 35.50
30 33 34 0.171 16 1.00 21.00
31 58 98 0.223 14 1.00 9.60
32 56 40 0.258 19 1.00 17.20
33 66 36 0.240 38 1.00 6.15
34 50 36 0.201 26 1.00 17.10
35 40 43 0.173 26 1.00 13.80
36 67 37 0.351 20 1.00 17.40
37 46 43 0.216 10 1.00 26.00
38 63 42 0.276 32 1.00 31.00
39 60 79 0.233 18 1.00 11.00
40 38 34 0.156 22 1.00 16.90
41 47 45 0.217 16 1.00 28.20
42 63 43 0.407 11 1.00 21.00
43 63 173 0.221 28 1.00 16.30
44 64 182 0.247 20 1.00 16.00
45 65 178 0.297 10 1.00 13.40
46 45 180 0.157 30 1.00 15.00
47 54 182 0.179 26 1.00 12.10
48 85 18 0.355 22 1.00 8.30
49 55 180 0.183 14 1.00 16.40
50 60 182 0.180 70 1.00 27.00
51 84 178 0.355 40 1.00 7.55
52 81 179 0.351 30 1.00 11.90
53 81 178 0.368 30 1.00 8.90
54 83 177 0.435 20 1.00 12.00
55 51 180 0.182 19 1.00 19.00
56 54 180 0.199 7 1.00 16.00

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 313

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Table 2. (Continued.)
Data source Test type Test number DR (%) σ0v;0 (or σ0c;0 ) (kPa) Cyclic stress ratio N max r u;max γmax (%)
57 540 176 0.220 13 1.00 25.00
58 81 178 0.460 19 1.00 14.70
59 78 38 0.440 28 1.00 9.60
60 74 44 0.395 20 1.00 6.70
61 81 83 0.424 6 1.00 7.90
62 49 81 0.312 11 1.00 24.50
63 64 79 0.429 20 1.00 22.50
64 54 95 0.195 20 0.46 0.65
65 59 85 0.166 20 0.31 0.35
66 48 40 0.164 20 0.37 0.20
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

67 40 85 0.099 20 0.17 0.20


68 43 85 0.136 20 0.30 0.30
69 82 80 0.401 20 1.00 7.20
70 80 79 0.292 20 0.66 1.50
71 70 80 0.227 20 0.61 0.70
72 77 81 0.205 20 0.37 0.25
73 50 75 0.220 20 1.00 17.50
74 44 83 0.162 20 0.34 0.15
75 34 80 0.162 20 0.87 6.10
76 38 36 0.141 20 0.28 0.50
77 37 82 0.119 20 0.25 0.10
78 74 40 0.322 20 0.77 1.90
79 63 40 0.371 20 0.94 4.20
80 75 36 0.360 20 0.48 0.50
81 50 40 0.157 20 0.46 0.70
82 78 40 0.224 20 0.37 0.40
83 81 37 0.418 20 0.70 2.30
84 82 35 0.479 20 0.82 4.00
85 49 39 0.173 20 0.52 0.60
86 38 40 0.128 20 0.23 0.20
87 60 181 0.158 20 0.30 0.35
88 78 179 0.273 20 0.83 3.40
89 78 180 0.214 20 0.47 0.50
90 78 81 0.367 20 0.92 5.95
91 81 86 0.371 20 0.90 5.60
Bilge (2005) Cyclic triaxial test 92 75 100 0.390 20 0.74 3.84
93 85 100 0.500 20 0.59 1.61
94 80 100 0.450 20 0.60 1.41
95 82 100 0.550 20 0.74 4.51
96 75 100 0.500 12 1.00 16.60
97 80 100 0.400 20 0.37 0.37
98 66 100 0.580 11 1.00 9.96
99 77 100 0.210 20 0.23 0.20
100 60 100 0.350 14 1.00 6.50
101 69 100 0.200 20 0.09 0.12
102 65 100 0.100 20 0.23 0.08
103 71 100 0.480 17 1.00 25.00
104 75 100 0.250 20 0.45 0.57
105 75 100 0.200 20 0.25 0.41
106 65 100 0.250 20 0.71 4.19
107 65 100 0.550 14 1.00 15.23
108 60 100 0.400 9 1.00 35.84
109 74 100 0.150 20 0.34 1.28
110 35 100 0.200 5 1.00 32.00
111 65 100 0.300 13 1.00 11.01
112 57 100 0.350 11 1.00 19.34

314 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Table 2. (Continued.)
Data source Test type Test number DR (%) σ0v;0 (or σ0c;0 ) (kPa) Cyclic stress ratio N max r u;max γmax (%)
113 75 100 0.140 20 0.15 0.21
114 72 100 0.400 12 1.00 22.27
115 57 100 0.380 4 1.00 24.39
116 64 100 0.420 7 1.00 12.63
117 46 100 0.300 19 1.00 22.01
118 82 100 0.310 20 0.23 0.49
119 64 100 0.350 10 1.00 8.11
120 51 100 0.380 4 1.00 24.11
121 55 100 0.200 17 1.00 20.00
122 60 100 0.500 13 1.00 28.00
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

123 75 100 0.500 11 1.00 19.82


124 59 100 0.240 18 1.00 24.59
125 45 100 0.400 10 1.00 20.26
126 59 100 0.280 8 1.00 21.49
127 85 100 0.250 20 0.14 0.36
128 75 100 0.400 20 0.75 4.35
129 84 100 0.120 20 0.04 0.09

occurrence of r u ¼ 1:0. The upper and lower boundaries of Polito Thus, direct comparisons between strain and stress test result-based
et al. (2008) were obtained by using Eq. (2) with different combi- methods are concluded to be misleading and not fair.
nations of DR (30–85%) and CSR (0.05–0.50). Consistent with the
findings of Wu et al. (2003), these comparisons revealed that Limit-State Models
the proposed boundary curves are not only biased, but also cannot The model for the limit-state function has the general form g ¼ g
encapsulate all the data points, indicating a larger variability in r u (x, Θ), where x is a set of descriptive parameters and Θ is the set of
versus N=N liq domain. unknown model coefficients. Inspired by previous studies and ob-
Dobry (1985) also proposed upper and lower boundaries in the servational trends, it is concluded that the key parameters affecting
r u versus γ domain on the basis of the results of strain-controlled excess pore-water pressure generation response are induced—
cyclic tests, and these boundaries are presented comparatively with cyclic shear strain (γmax ), DR , and σ0v;0 (or σ0m;0 ). As shown in Fig. 3,
the compiled database in Fig. 2. As briefly referred to previously, owing to lack of strong r u versus γmax trends as a function of sample
there exists a major disagreement in r u versus γ responses of stress- preparation technique (i.e., moist tamped, air or wet pluviated), and
and strain-controlled test data because of significantly different with the aim of a less complex model, an input parameter to address
strain paths followed during these cyclic tests. In simpler terms, sample preparation technique was not included in the proposed
Dobry’s boundary curves correspond to a specific case, in which mathematical model. Strong correlation between shear and volu-
10 cycles of shear strains were applied, and at the end of which metric straining (e.g., Ishihara and Yoshimine 1992; Cetin et al.
corollary r u values were estimated. However, during stress- 2009), and volumetric straining and excess pore pressure genera-
controlled cyclic testing, the estimated r u values correspond to the tion encouraged the use of γmax alone, rather than together with
first occurrence (rather than the 10th) of the recorded maximum volumetric strain in the limit-state model. Various functional forms
have been tested, and the following mathematical (functional)
strain levels, and are less dependent on the stress (or strain) histories.

Fig. 1. Comparison of the test results with Seed et al. (1975) and Polito Fig. 2. Comparison of the test results with Dobry (1985) boundary
et al. (2008) boundary curves curves

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 315

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


N 1;60;CS , as recommended by Cetin et al. (2009) and repeated
in Eq. (21).

lnðγmax;20 Þ
 
0:025 · N 1;60;CS þ lnðCSRSS;20;1D;1 atm Þ þ 2:613
¼ ln
0:004 · N 1;60;CS þ 0:001
 1:880
lim : 5 ≤ N 1;60;CS ≤ 40; 0:05 ≤ CSRSS;20;1D;1 atm
≤ 0:60 and 0% ≤ γmax ≤ 50% ð21Þ

2. γmax;N , defined as the maximum double-amplitude cyclic shear


strain at the end of Nth loading cycle, is estimated as a func-
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

tion of γmax;20 , γcritical , N, and DR , as given in Eq. (22).


Fig. 3. Proposed r u versus γmax;N model along with test results used in gγmax;N ðγmax;N ; γmax;20 ; N; γcritical ; DR ; ΘÞ
this study 
¼ lnðγmax;N Þ  ln γmax;20

form, where θi;ru;N represents the set of unknown model coefficients,    θ γmax;20 θ2;γ ·lnðDR Þ 
1 1 N 1;γmax;N ·ð γcritical Þ max;N
was concluded to be improved, not only because of its best fit to · þ arcsin 2 1
2 π 20
available data, but also because of its power to differentiate dilative  θ 
and contractive soil responses. N 3;γmax;N
· þ εlnðγmax;N Þ
20
gru;N ðDR ; σ0v;0 ; γmax;N ; r u;N ; ΘÞ ¼ lnðru;N Þ
lim : N ≤ 20 ð22Þ
 
θ1;ru;N · γmax;N
 ln 1  exp  For the estimation of γmax;N , the use of a magnitude (duration)
θ2;ru;N þ θ3;ru;N · ln σ0v;0  ðDR =100Þ ð20Þ
  scaling factor is not necessary. However, equivalent number of
θ4;ru;N uniform stress cycles of the scenario earthquake at the soil site
· þ1 þ εlnðru;N Þ
1 þ γmax;N needs to be estimated, the estimation of which is discussed
elsewhere (e.g., Liu et al. 2001; Cetin and Bilge 2012).
The first denominator of the exponential term [i.e., θ2;ru;N þ 3. γcritical is a curve-fitting parameter and is defined as the critical
θ3;ru;N · ln σ0v0  ðDR =100Þ] is written in a functional form very sim- shear strain value below which (i.e., γmax;20 < γcritical ) the cyc-
ilar to dilation index (Konrad 1988). Similarly, the multiplier inside lic response can be best described by cyclic mobility type of
θ4;r response with the characteristics of constantly decreasing rate
the exponent term [i.e., ð1þγmax;N
u;N
þ 1Þ] has a form very similar to of shear strain accumulation with increased number of loading
strain-dependent modulus degradation (Hardin and Drnevich cycles. However, if γmax;20 > γcritical , a sudden increase in the
1972). Thus, the selected mathematical form is capable of capturing rate of shear strain accumulation is observed. The variable
and differentiating between the contractive and dilative responses, γcritical is proposed to be estimated as a function of DR and
and can effectively incorporate the effects of strain-dependent σ0v;0 , as given in Eq. (23)
modulus degradation on interrelated excess pore pressure genera-
tion response. This will be illustrated further in the section that gγcritical ðγcritical ; DR ; σ0v;0 ; ΘÞ ¼ lnðγcritical Þ
follows.  
θ1;γcritical · DR
Eq. (20) alone has limited practical use because of difficulties  ln þ θ
ðσ0v;0 Þθ2;γcritical
3;γcritical
associated with the prediction of γmax;N . Even though there exist
some semiempirical methods for the prediction of cyclically in- ð23Þ
duced shear strains as a function of a demand (CSR or FSliq )
and a capacity (DR , SPT-N, or CPT-qc ) term (e.g., Tokimatsu and Eqs. (20) and (22) include a random model correction term
Seed 1984; Ishihara and Yoshimine 1992; Shamoto et al. 1998; (ε) to account for the possibilities that (1) missing descriptive
Cetin et al. 2009), their use is limited to only M w 7.5 earthquakes variables may exist, and (2) the adopted mathematical expres-
(or 20 loading cycles) or requires magnitude (duration) corrections sion may not have the ideal functional form. It is reasonable
applied on cyclic shear stresses, which themselves are another and also convenient to assume that ε follows a normal
controversial issue waiting to be resolved. Not to bring along the distribution with a mean of zero for the aim of producing
uncertainties associated with magnitude corrections, and with the an unbiased model (i.e., one that on the average makes cor-
intention of capturing the cyclic liquefaction and mobility re- rect predictions). The standard deviation of ε, denoted as σε ,
sponses, an alternative procedure is proposed, which eliminate however, is unknown and must be estimated. As will be sub-
the need to use magnitude (duration) corrections. As part of this sequently illustrated, data scatter is observed to be reduced
procedure, by increasing maximum shear strain levels; thus, model
1. γmax;20 , defined as the maximum double-amplitude cyclic uncertainty is preferred to be a function of γmax;N itself. This
shear strain at the end of the 20th loading cycle, is estimated suggests a heteroscedastic σε model as expressed in Eqs. (24)
as a function of CSRSS;20;1D;1 atm (i.e., CSR value correspond- and (25), which are proposed for Eqs. (20) and (22), respec-
ing to a 1D, cyclic SST performed under a confinement tively. The set of unknown coefficients of the models, there-
pressure of 100 kPa for 20 uniform loading cycles) and fore, is Θ ¼ ðθ; σε Þ

316 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


 
1 0:407 · γmax;N
σεlnðru;N Þ ¼ θ4;ru;N
ð24Þ lnðr u;N Þ ¼ ln 1  exp 
ðγmax;N Þ þ θ5;ru;N 0:486 þ 0:025 · ln σ0v;0  ðDR =100Þ
 
0:620
· þ1  σεlnðru;N Þ ð27Þ
1 þ γmax;N
1
σεlnðγmax;N Þ ¼ θ5;γmax;N
ð25Þ
ðγmax;N Þ þ θ6;γmax;N

 
1 1
In forward analysis, uncertainties in the input parameters lnðγmax;N Þ ¼ ln γmax;20 · þ arcsin
(e.g., N 1;60;CS , CSR, DR , and σ0v;0 ) need to be properly con- 2 π
volved along with the model error. In doing so, addressing   1:285·ðγmax;20 Þ0:125·lnðDR Þ 
N γcritical
possible correlations among input parameters is crucial to es- × 2 1
20
timate the true level of accumulated uncertainty. For a more  0:088 
N
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

complete illustration of how these uncertainties can be as- ·  σεlnðγmax;N Þ ð28Þ


sessed, refer to Cetin (2000). 20

Likelihood Function
where γcritical ¼ ðσ 0 Þ0:295  7:65,
0:943·DR
σεlnðru;N Þ ¼ ðγ 1
,
max;N Þ
1:375 þ1:889
For the r u;N versus γmax;N model [Eq. (20)], under the assumptions v;0

σεlnðγmax;N Þ ¼ ðγ 1
and γmax;20 can be estimated by Eq. (21).
that ðDR Þi , ðσ0v;0 Þi , and ðγmax;N Þi values of the ith test are exact max;N Þ
0:626 þ1:557

(i.e., no measurement error), and statistically independent, noting To develop an understanding of the range of model predictions,
that gð…Þ ¼ ^gð…Þ þ εi has the normal distribution with mean ^g Fig. 3 presents ru versus γmax curves for (1) DR ¼ 100% and σ0v;0 ¼
and standard deviation σε , then the likelihood function for j, k, 40 kPa (“dilative”), and (2) DR ¼ 35% and σ0v;0 ¼ 400 kPa (“con-
m, and n number of tests in which σ0v;0 were applied as 40, 80, tractive”), intended to differentiate both dilative and contractive soil
100, and 180 kPa, respectively, can be written for Eq. (20) as a responses. In the same figure, mean and mean 2σεlnðru;N Þ r u curves
function of unknown coefficients as given in Eq. (26). In are also shown corresponding to median DR and σ0v;0 values of the
Eq. (26), φ½: = standard normal probability density function. compiled database, which are estimated as 63% and 99 kPa, respec-
tively. When drawing mean þ2σεlnðru;N Þ curves, an upper limit of 1.0
is applied to estimated r u values. Similarly, in Fig. 4, by using
Y
j
wσ0 ¼40
Y
jþk
wσ 0 ¼80 Eq. (28), model boundary curves for DR equals to 60%, and cor-
Lru;N ðθ; σε Þ ¼ fφ½·g v;0 × fφ½·g v;0

i¼1 i¼jþ1
responding to three different γmax;20 =γcritical values of 0.05, 1, and
10 are overlain on the compiled data, in the γ=γ20 versus N domain.
Y
jþkþm
wσ0 ¼100
Y
jþkþmþn
wσ0 ¼180 As revealed by both of the figures, the proposed models can (1) dif-
× fφ½·g v;0 × fφ½·g v;0 ð26Þ ferentiate both dilative and contractive responses and data trends in
i¼jþkþ1 i¼jþkþmþ1
an unbiased manner, and (2) encapsulate the data pairs with the
estimated model standard deviations, which enables direct assess-
As referred to previously, tests with initial vertical confining ment of uncertainties in model predictions. To further strengthen
stresses of 100 kPa were overrepresented. To correct against this these observations, the differences between natural logarithm of
sampling disparity problem, weighting factors wσ0v;0 ¼40 , wσ0v;0 ¼80 , the recorded (test) and predicted values, referred to as residuals,
wσ0v;0 ¼100 and wσ0v;0 ¼180 were applied, as shown in the likelihood are also shown as a function of γmax;N in Figs. 5 and 6 for r u;N
function, consistent with the available literature (Manski and and γmax;N models, respectively. It is consistently illustrated by both
Lerman 1977; Hsieh et al. 1985; Cetin et al. 2002). In the estima- of the figures that a larger scatter in residuals is observed at
tion of the corrective weighting factors, most attention was given γmax;N < 1%, supporting our choice of γmax;N -dependent σε . This
not to increase the database size or available data information observation is also supported from dynamic soil response point of
(i.e., likelihood information) artificially through weighting factors. view, because r u values have a natural upper cap of 1.0, expected to
Thus, it was preferred simultaneously to downweight the overre-
presented and upweight the underrepresented, which produced
the choice of 1.193, 1.125, 0.49, and 1.193 as corrections factors
of wσ0v;0 ¼40 , wσ0v;0 ¼80 , wσ0v;0 ¼100 and wσ0v;0 ¼180 , respectively. A detailed
discussion on how to assess databases with sampling disparity
problems and weighting factors is available at Manski and Lerman
(1977) and Cetin et al. (2002). A similar methodology is also fol-
lowed for the development of γmax;N versus N model’s [Eq. (22)]
likelihood function. However, because of differences in the data-
base characteristics, weighting factors were estimated and used
as 1.281, 0.791, and 0.928 for σ0v;0 values of 40, 100, and
200 kPa, respectively.

Model Coefficients
Consistent with the maximum likelihood methodology, model co-
efficients are estimated by maximizing the likelihood functions.
The final forms of the resulting proposed models are presented
in Eqs. (27) and (28) along with  one standard deviation of the
Fig. 4. Proposed γmax;N model along with test results used in this study
model error term.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 317

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


be reached at large strains, also suggesting a reduced model uncer-
tainty in the predictions at these large strain responses.
Last, but not least, to check the power of the proposed math-
ematical form (i.e., limit-state functions) and to see whether any
trend as a function of model input variables (descriptors) is hidden
in residuals, residuals versus DR , σ0v;0 and N plots are also prepared,
as presented in Bilge and Cetin (2009), and lack of trend as a func-
tion of any of these input variables confirmed the validity of se-
lected functional forms. Because of almost zero mean residuals
(μresidual;ru ¼ 0:001 and μresidual;γmax;N ¼ 0:085) in both relatively
small strain (γmax;N < 1%) and large strain (γmax;N > 1%) domains,
and their relatively smaller standard deviations (σresidual;ru ¼ 0:224
and σresidual;γmax;N ¼ 0:308), the proposed models are considered to
be relatively more accurate and precise.
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Residuals of the proposed r u;N model


Calibration of and Comparisons with the Existing
Frameworks

For comparison purposes, the compiled database is also used to test


the performance of the stress-based models of Seed et al. (1975)
and Polito et al. (2008); strain-based models of Martin et al.
(1975), Dobry et al. (1985), and Byrne (1991); and energy-based
model of Green et al. (2000). Input parameters for these models are
adopted consistently with the recommendations and definitions
presented in the source documents. Both for comparison and
calibration purposes, for each framework, two alternatives were
followed: model implemented with (1) the original, and (2) the up-
dated model coefficients. The first alternative presents an opportu-
nity to make a judgment regarding which model, in its original
form, is the least biased, and naturally, what should be the average
calibration (correction) factors. As a result of the second assess-
ment (i.e., by comparing the “updated” models), it is possible to
decide which framework has the best limit-state model or func-
Fig. 6. Residuals of the proposed γmax;N model
tional form. In simpler terms, a framework may have a better

Table 3. Summary of Limit-State Functions, Coefficients, and Model Performances


Model parameters
e e
Model limit-state functions θ1 θ2 θ3 θ4 θ5 θ6
n h  io
This study θ ·γ
lnðr u;N Þ ¼ ln 1  exp  θ2 þθ3 ·ln 1σ0 max;N θ4 0:407 0:486 0.025 0.620 1.375 1.889
ðDR =100Þ · 1þγmax;N þ 1
v;0
 εlnðru;N Þ

Seed et al. n h  1=θ io Original 0.7 — — — — —


lnðr u;N Þ ¼ ln þ π1 · arcsin 2 NNliq  1  εlnðru;N Þ
1 1
(1975) 2 Updated 1.215 — — — 0.356 1.225

Polito et al. n h  1=ðθ ·FCþθ ·D þθ ·CSRþθ Þ o Original — 0.007397 0.01034 0.5058 — —


lnðru;N Þ ¼ ln þ π1 · arcsin 2 NNliq  1  εlnðru;N Þ Updated
1 1 2 R 3 4
(2008)a 2 — 0:013 1.197 1.675 0.425 1.250

Martin et al. n h io Original — — — — — —


ðσ0v Þ1θ1 c ·ε2
(1975)b lnðr u;N Þ ¼ ln θ1 ·θ3 ·ðσ0v;0 Þθ2 θ1
· c1 · ðγ  c2 · εvd Þ þ γmax;N3 þcvd4 ·εvd  εlnðru;N Þ Updated 0.80 1.148 0.00011 — 0.258 0.614

Dobry et al. h i Original — — — — — —


θ1 ·θ2 ·N·ðγmax;N θ3 Þθ4
(1985)c lnðru;N Þ ¼ ln 1þθ ·N·ðγ θ Þθ4
 εlnðru;N Þ Updated 1.00 0.206 0.010 1.161 1.448 1.130
2 max;N 3

Byrne  n  o Original — — — — — —
ðσ0v Þ1θ1 εvd
(1991)d lnðr u;N Þ ¼ ln θ1 ·θ3 ·ðσ0v;0 Þθ2 θ1
· c1 · γ · exp c2 · γmax;N  εlnðru;N Þ Updated 0.80 1.229 0.00091 — 0.978 0.869

Green Mitchell nqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffio Original 0.0139 1:021 — — — —


lnðru;N Þ ¼ ln exp½expðθ1 ·DR Þþθ2   εlnðr u;N Þ
Ws
Polito model Updated 0.017 2:016 — — 1.170 1.700
a
FC value is adopted as zero for clean sands, and θ1 was not assessed.
b
Determination of coefficients c1 to c4 is based on Martin et al. (1975).
c
Cyclic loading is applied in 1D and f ¼ 1:0.
d
Determination of coefficients c1 and c2 is based on Byrne (1991).
e
σεlnðru;N Þ ¼ θ5;r
1
.
ðγmax;N Þ
u;N þθ6;ru ;N

318 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Table 4. Summary of Model Performances
γmax;N ≥ 1% 0 < γmax;N < 1% 0 < γmax;N < ∞
Method μresidual σresidual Maxresidual Minresidual μresidual σresidual Maxresidual Minresidual μresidual σresidual
This study 0:014 0.164 1.128 1:221 0.046 0.362 0.608 1:532 0.001 0.224
Seed et al. (1975) 0.525 0.357 1.818 0:923 0.353 0.506 1.690 0:667 0.456 0.431
“Updated” Seed et al. (1975) 0.267 0.225 0.651 1:799 0:196 0.403 0.988 1:032 0.081 0.383
Polito et al. (2008) 0.339 0.244 1.057 1:620 0:065 0.466 1.146 0:876 0.182 0.378
“Updated” Polito et al. (2008) 0.193 0.221 0.798 2:061 0:232 0.451 0.995 1:191 0.041 0.365
“Updated” Martin et al. (1975) 0.060 0.502 2.190 1:311 0.240 0.737 1.868 1:480 0.121 0.585
“Updated” Dobry et al. (1985) 0.052 0.195 1.335 1:886 0.145 0.648 1.223 0:706 0.070 0.357
“Updated” Byrne (1991) 0:129 0.283 1.385 1:842 0:091 0.589 1.667 1:569 0:116 0.412
Green Mitchell Polito model 0.092 0.227 1.136 1:149 0.069 0.475 0.889 0:911 0.080 0.309
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

“Updated” Green Mitchell Polito model 0.016 0.242 2.190 5:747 0:104 0.554 1.868 1:978 0:025 0.335

functional form (limit state), but poorly estimated model coeffi- underpredict r u;N values in both γmax;N < 1% and γmax;N ≥ 1%
cients in its original form may reduce its accuracy. subsets. The extent of this underprediction varies in a very wide
For the “updated” models (models which have the same func- range from 6 to 70%. The “updated” Dobry et al. (1985) model
tional form as with the original models, but with updated model produces approximately 7% lower values, in the average. Updated
coefficients), maximum likelihood approach, as discussed previ- models of GMP (2008), Seed et al. (1975), Polito et al. (2008), and
ously, was used for the estimation of the model coefficients. A “original” Polito et al. (2008) produce smaller predictions in
summary of the functional forms and model coefficients is pre- γmax;N < 1% range, and higher values for the rest. In contrast,
sented in Table 3. The performance of the models is assessed the “updated” Byrne model produces 11% higher predictions,
by simple statistics (i.e., mean and standard deviation) of residuals, in the average. As the mean of their residuals is almost zero
and is presented in Table 4. A smaller absolute mean residual, and their standard deviations are relatively smaller, the proposed
jμresidual;ru j, and σresidual;ru can be simply interpreted as a relatively frameworks are concluded to be unbiased and more precise, as
more accurate and precise model. A positive μresidual;ru indicates compared with the updated models.
that the model predictions of r u are lower than the values recorded
(measured) during cyclic testing (i.e., unconservatively biased).
Estimated μresidual and σresidual values are summarized in Table 4, Validation of the Calibrated Models
for 0 < γmax;N < 1% and γmax;N ≥ 1% subsets, separately. Fig. 7
illustrates the continuous variation of μresidual;ru and μresidual;γmax;N as As the basis of the proposed and “updated” available models,
a function of, γupper limit as defined in Eq. (29): CTXT and SST performed on Sacramento River, Monterey number
Z γ 0=30, and Kizilirmak River sands were used. To eliminate possible
upper limit concerns regarding whether the proposed models are general
μresidual ¼ ðResidualÞ · dγ ð29Þ
0 (global) enough to assess the performance of other types of sands,
from available literature, a total number of 10 independent cyclic
As illustrated by this figure, the proposed frameworks produce tests were compiled and reserved for validation purposes. Test and
consistently unbiased estimates of ru and γmax;N responses in a sample characteristics, and DR , σ0v;0 (or σ0c;0 ), and uncorrected test
wide range of strain levels extending from 0.01 to more than CSR values of the validation set are presented in Table 5. It should
10%. However, as summarized in Table 4, the “original” Seed et al. be noted that this data set was only reserved for validation assess-
(1975), GMP (2008), and Martin et al. (1975) models consistently ments, and was carefully not included in the compiled database
used for the development of both the proposed r u model and
the calibrated existing r u models.
Proposed and existing models with both original and updated
model coefficients were used to simulate each test result in the val-
idation data set. Because of lack of precise stress-strain history data,
GMP model predictions were not presented comparatively for the
validation database. Each input parameter is adopted consistently
as defined and recommended in the source documents. As illus-
trated in Fig. 8 for test V2, r u;N responses estimated by all methods
are compared with the actual response (response recorded during
cyclic testing) according to the key input parameter of each model.
A complete documentation of these comparisons can be found at
Bilge and Cetin (2009). Similarly, the accuracy and precision of the
model predictions are assessed by simple statistics of residuals
(i.e., mean, μresidual;ru and standard deviation, σresidual;ru ), which
are summarized in Table 6 for all models. Fig. 9 presents continu-
ous variation of mean residuals calculated for strain range
γmax;N ≤ γupper limit . As revealed by this figure because of consis-
tently small (almost zero) mean residual μresidual;ru and smaller
Fig. 7. Variation of mean r u and γmax;N residuals with γupper limit
σresidual;ru values, it is concluded that the proposed models again

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 319

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Table 5. Summary of Database Compiled for the Validation of the Proposed Models
Number Test type Sand type DR (%) σ0v0 (or σ0c0 ) (kPa) Cyclic stress ratio Reference
V1 CTXT Ottawa 71.3 200 0.251 Vaid and Chern (1983)
V2 CTXT Ottawa 55.6 200 0.157 Vaid and Chern (1983)
V3 CTXT Gioia Tauro 35 40 0.21 Ghionna and Porcino (2006)
V4 SST Fraser River 44 200 0.10 Sriskandakumar (2003)
V5 SST Fraser River 40 100 0.10 Sriskandakumar (2003)
V6 SST Nevada 40 80 0.093 Arulmoli et al. (1992)
V7 SST Nevada 40 160 0.071 Arulmoli et al. (1992)
V8 SST Nevada 40 80 0.181 Arulmoli et al. (1992)
V9 SST Nevada 40 80 0.134 Arulmoli et al. (1992)
V10 SST Ticino 75 100 0.20 Porcino and Caridi (2007)
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Comparisons of the predicted and observed cyclic responses for the validation set (test V2)

Table 6. Model Performances for the Validation Database


γmax;N ≥ 1% 0 < γmax;N < 1% 0 < γmax;N < ∞
Method μresidual σresidual Maxresidual Minresidual μresidual σresidual Maxresidual Minresidual μresidual σresidual
This study 0.036 0.086 0.280 0:067 0.059 0.107 0.320 0:210 0.045 0.095
Seed et al. (1975) 0.210 0.205 0.392 0:171 0.202 0.227 0.739 0:083 0.206 0.214
“Updated” Seed et al. (1975) 0.089 0.144 0.305 0:680 0:128 0.232 0.285 0:144 0:032 0.209
Polito et al. (2008) 0.111 0.207 0.564 0:487 0:020 0.263 0.648 0:088 0.101 0.224
“Updated” Polito et al. (2008) 0.071 0.128 0.287 0:879 0:121 0.232 0.307 0:093 0:026 0.212
“Updated” Martin et al. (1975) 1:040 0.683 0.000 2:799 2:140 0.313 0.000 2:780 1:300 0.642
“Updated” Dobry et al. (1985) 0.041 0.201 1.459 0:291 0.383 0.507 0.933 0:291 0.172 0.391
“Updated” Byrne (1991) 0:020 0.225 0.697 0:555 0.023 0.805 1.372 1:737 0:003 0.529

produced quite unbiased and the most precise predictions for sands is described. For this purpose, results of SST and CTXT were
the validation database. For validation database, the “updated” studied. The resulting r u versus γ and γ versus N databases are
Byrne model also produced consistently unbiased predictions. composed of 101 and 84 cyclic test data. Distribution of specimens
However, the model’s relatively higher standard deviations (ranks with respect to DR is relatively uniform, whereas the resulting data-
the second highest after the “updated” Martin et al. model) unfav- base is concluded to be biased with tests of initial confining stresses
orably affect the performance of the model (especially the precision of 100 kPa. This sampling bias is subsequently corrected in the
of predictions).
assessment of the proposed model coefficients. Inspired by pre-
vious studies and observational trends from test results, it is con-
Summary and Conclusions cluded that the key parameters affecting excess pore-water pressure
generation response are γmax , DR , and σ0v;0 (or σ0m;0 ). Strong corre-
A semi empirically based probabilistic model to investigate large lation between shear and volumetric straining (e.g., Ishihara and
strain excess pore-water pressure response of fully saturated clean Yoshimine 1992; Cetin et al. 2009), and volumetric straining and

320 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Notation

The following symbols are used in this paper:


CSR = cyclic stress ratio;
CSRSS;20;1D;1 atm = CSR value corresponding to a 1D, 20 uniform
loading cycles SST under a confinement pressure of
100 kPa (¼ 1 atm);
DR = relative density (%);
E r = tangent modulus of the 1D unloading curve
corresponding to the initial effective vertical stress;
FC = fines content;
FSliq = factor of safety against liquefaction;
M w = moment magnitude;
N = number of loading cycles applied in test;
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

N liq = number of cycles to liquefaction;


Fig. 9. Variation of mean r u residuals with γupper limit for the validation
N 1;60;CS = overburden, fines, and the procedure-corrected
database
standard penetration test (SPT) blow counts;
PEC = pseudo-energy capacity;
r u = excess pore-water pressure ratio;
excess pore pressure generation encouraged the use of γmax alone, r u;N = excess pore-water pressure ratio at Nth loading
rather than together with the volumetric strain in the limit-state cycle;
model. Various functional forms have been tested, and the proposed W s = energy dissipated per unit volume of soil divided by
mathematical (functional) form was concluded to be improved, not the initial effective mean confining stress;
only because of its best fit to available data, but also because of its Δu = increase in pore-water pressure;
power to differentiate dilative and contractive soil responses. Con- Δεvd = change in volumetric strain attributable to cyclic
sistent with the maximum likelihood methodology, model coeffi- shearing;
cients are estimated by maximizing the likelihood function. εa = axial strain;
For comparison purposes, the compiled database was again used εvd = accumulated volumetric strain;
to evaluate the performance of the stress-based models of Seed et al. γc = cyclic shear strain;
(1975) and Polito et al. (2008); strain-based models of Martin et al. γcritical = critical shear strain (%);
(1975), Dobry et al. (1985), and Byrne (1991); and energy-based γmax = maximum double-amplitude shear strain (%);
model of Green et al. (2000). Both for comparison and calibration γmax;20 = maximum double-amplitude shear strain at 20th
purposes, for each framework, two separate sets of limit-state mod- loading cycle (%);
els were used: model implemented with (1) the original and (2) the γmax;N = maximum double-amplitude shear strain at Nth
updated model coefficients. The model performances are assessed loading cycle (%);
by simple statistics (i.e., mean and standard deviation) of residuals σ0c = effective confinement stress;
and summarized in Table 4. A smaller absolute mean residual, σ0c;0 = initial effective confinement stress;
σ0m = mean effective stress;
jμresidual;ru j, and σresidual;ru can be simply interpreted as a relatively
σ0m;0 = initial mean effective stress;
less biased and more accurate model. It is concluded that the “origi-
σ0v = vertical effective stress;
nal” Seed et al. (1975) and GMP (2008), and “updated” Martin et al.
σ0v;0 = initial vertical effective stress; and
(1975), and Dobry et al. (1985) model predictions of r u;N are con- τ = cyclic shear stress.
sistently smaller than the test values, the extent of which varies in
the range of 5–70%. In contrast, the “updated” Byrne (1991) model
consistently overpredicts r u;N values. “Updated” models of GMP
(2008), Polito et al. (2008), Seed et al. (1975), and original model References
of Polito et al. (2008) overpredict and underpredict r u;N values for Arulmoli, K., Muraleetharan, K. K., Hosain, M. M., and Fruth, L. S.
shear strain levels less than and greater than 1%, respectively. The (1992). “VELACS laboratory testing program, soil data report.” Rep.
extent of these inconsistent predictions varies in the range from 2 to to the National Science Foundation, The Earth Technology Corpora-
70%. The comparison of measured and predicted responses indi- tion, Irvine, CA, Washington, D.C.
cated that the proposed excess pore-water pressure model produced Bilge, H. T. (2005). “Volumetric and deviatoric strain behavior of cohesion-
the most unbiased response as compared with the updated models. less soils under cyclic loading.” M.Sc. thesis, Middle East Technical
Univ., Ankara, Turkey.
However, to eliminate the concerns regarding whether the proposed
Bilge, H. T., and Cetin, K. O. (2009). “Cyclic large strain and induced pore
models are general (global) enough to assess the performance of pressure response of saturated clean sands.” Rep. No. METU/GTENG
other types of sands, a total number of 10 cyclic tests was studied, 09/12-01, Middle East Technical Univ. Soil Mech. & Foundation Eng.
and DR , σ0v;0 (or σ0c;0 ) values and variation of r u and γmax with num- Research Center, Ankara, Turkey. 〈www.ce.metu.edu.tr/~onder/
ber of stress cycles were assessed. On the basis of repeated im- Publications/reports/T_10.pdf〉.
proved predictions of the proposed models for this control data Booker, J. R., Rahman, M. S., and Seed, H. B. (1976). “GADFLEA—a
set, the proposed procedure is concluded to be a better alternative computer program for the analysis of pore pressure generation and dis-
sipation during cyclic or earthquake loading.” EERC Report No. 76-24,
to existing models. In addition to differentiating contractive and
Univ. of California, Berkeley, CA.
dilative cyclic soil responses and improved model predictions, Boulanger, R. W., Seed, R. B., Chan, C. K., Seed, H. B., and Sousa, J.
the main advantage of the proposed methodology is the probabi- (1991). “Liquefaction behavior of saturated sands under uni-directional
listic nature of the model predictions, which enables incorporation and bi-directional monotonic and cyclic simple shear loading.” Geotech-
of the model uncertainty into pore pressure generation predictions. nical Report No. UCB/GT/91-08, Univ. of California, Berkeley, CA.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 321

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


Byrne, P. M. (1991). “A cyclic shear-volume coupling and pore pressure Proc., John Booker Memorial Symp.—Developments in Theoretical
model for sand.” Proc., 2nd Int. Conf. on Recent Advances in Geomechanics, D. W. Smith and J. P. Carter, eds., Balkema, Rotterdam,
Geotechnical Earthquake and Soil Dynamics, 47–55. Netherlands, 383–390.
Cetin, K. O. (2000). “Reliability-based assessment of seismic soil liquefac- Hardin, B. O., and Drnevich, V. P. (1972). “Shear modulus and damping in
tion initiation hazard.” Ph.D. dissertation, Univ. of California, Berkeley, soils: Design equations and curves.” J. Soil Mech. Found. Div., 98(7),
CA. 667–692.
Cetin, K. O., and Bilge, H. T. (2012) “Performance-based assessment of Hashash, Y. M. A. (2009). “DEEPSOIL V3.5 1-D nonlinear and equivalent
magnitude (duration) scaling factors.” J. Geotech. Geoenviron. Eng., linear wave propagation analysis program for geotechnical seismic site
138(3), 324–334. response analysis of soil deposits.” User Manual and Tutorial, Dept. of
Cetin, K. O., Bilge, H. T., Wu, J., Kammerer, A. M., and Seed, R. B. (2009). Civil and Environmental Engineering, Univ. of Illinois at Urbana-
“Probabilistic models for cyclic straining of saturated clean sands.” Champaign, Urbana, IL.
J. Geotech. Geoenviron. Eng., 135(3), 371–386. Hazirbaba, K. (2005). “Pore pressure generation characteristics of sands
Cetin, K. O., Der Kiureghian, A., and Seed, R. B. (2002). “Probabilistic and silty sands: A strain approach.” Ph.D. dissertation, Univ. of Texas,
models for the initiation of seismic soil liquefaction.” Struct. Saf., Austin, TX.
24(1), 67–82. Hsieh, D., Manski, C. F., and McFadden, D. (1985). “Estimation of
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

Chameau, J. L., and Clough, G. W. (1983). “Probabilistic pore pressure response probabilities from augmented retrospective observations.” J.
analysis for seismic loading.” J. Geotech. Eng. Div., 109(4), 507–524. Am. Stat. Assoc., 80(391), 651–662.
Chang, W., Rathje, E. M., Stokoe II, K. H., and Hazirbaba, K. (2007). “In Hsu, H. L. (1995). “Study on the relationship between shear work and
situ pore-pressure generation behavior of liquefiable sand.” J. Geotech. pore-water pressure for saturated sand in undrained test.” First Int.
Geoenviron. Eng., 133(8), 921–931. Conf. on Earthquake Geotechnical Engineering, K. Ishihara, ed.,
Das, B. M. (1983). Fundamentals of soil dynamics, Elsevier, New York, Vol. 1, Balkema, Rotterdam, Netherlands, 301–307.
414. Hynes, M. E. (1988). “Pore pressure generation characteristics of gravel
Davis, R. O., and Berrill, J. B. (1982). “Energy dissipation and seismic under undrained cyclic loading.” Ph.D. dissertation, Univ. of California,
liquefaction in sands.” Earthquake Eng. Struct. Dyn., 10(1), 59–68. Berkeley, CA.
Davis, R. O., and Berrill, J. B. (2001). “Pore pressure and dissipated energy Ishihara, K. (1985). “Stability of natural deposits during earthquakes.”
in earthquakes—Field verification.” J. Geotech. Geoenviron. Eng., Proc., 11th Int. Conf. Soil Mech. Found. Eng., Taylor & Francis Group,
127(3), 269–274. London, 321–376.
De Alba, P., Chan, C. K., and Seed, H. B. (1975). “Determination of Ishihara, K. (1993). “Liquefaction and flow failure during earthquakes.”
soil liquefaction characteristics by large scale laboratory tests.” EERC Geotechnique, 43(3), 351–415.
Rep. No. 75-14, Univ. of California, Berkeley, CA. Ishihara, K. (1996). “Soil behaviour in earthquake geotechniques.” Oxford
Dobry, R. (1985). “Liquefaction of soils during earthquakes.” Rep. No. Engineering Science Series, Clarendon Press, New York, 360.
CETS-EE-001, National Research Council (NRC), Committee on Ishihara, K., and Yoshimine, M. (1992). “Evaluation of settlements in
Earthquake Engineering, Washington, DC. sand deposits following liquefaction during earthquakes.” Soils Found.,
Dobry, R., Ladd, R. S., Yokel, F. Y., Chung, R. M., and Powell, D. (1982). 32(1), 173–188.
“Prediction of pore-water pressure buildup and liquefaction of sands Kammerer, A. M. (2002). “Undrained response of Monterey 0=30
during earthquakes by the cyclic strain method.” National Bureau of sand under multi-directional cyclic shear loading conditions.” Ph.D.
Standards Building Science Series 138, National Bureau of Standards, dissertation, Univ. of California, Berkeley, CA.
U.S. Dept. of Commerce, Washington, D.C. Konrad, J. M. (1988). “Interpretation of flat plate dilatometer tests in sands
Dobry, R., Pierce, W. G., Dyvik, R., Thomas, G. E., and Ladd, R. S. in terms of the state parameter.” Geotechnique, 38(2), 263–277.
(1985). “Pore pressure model for cyclic straining of sand.” Research Law, K. T., Cao, Y. L., and He, G. N. (1990). “An energy approach for
Rep. 1985-06, Rensselaer Polytechnic Institute, Troy, NY. assessing seismic liquefaction potential.” Can. Geotech. J., 27(3),
Eggzelos, D. N. (2004). “Experimental and theoretical investigation of soil 320–329.
behavior under cyclic loading.” Ph.D. dissertation, National Technical Lee, K., and Albaisa, A. (1974). “Earthquake induced settlements in satu-
Univ. of Athens, Athens, Greece. rated sands.” J. Geotech. Eng. Div., 100(4), 387–405.
Elgamal, A., Yang, Z., Parra, E., and Ragheb, A. (2003). “Modeling of Liu, A. H., Stewart, J. P., Abrahamson, N. A., and Moriwaki, Y. (2001).
cyclic mobility in saturated cohesionless soils.” Int. J. Plast., 19(6), “Equivalent number of uniform stress cycles for soil liquefaction analy-
883–905. sis.” J. Geotech. Geoenviron. Eng., 127(12), 1017–1026.
Evans, M. D. (1987). “Undrained cyclic triaxial testing of gravels: The ef- Liyanapathirana, D. S., and Poulos, H. G. (2002). “A numerical model
fect of membrane compliance.” Ph.D. dissertation, Univ. of California, for dynamic soil liquefaction analysis.” Soil Dyn. Earthquake Eng.,
Berkeley, CA. 22(9–12), 1007–1015.
Evans, M. D., and Rollins, K. M. (1999). “Developments in gravelly soil Manski, C. F., and Lerman, S. R. (1977). “The estimation of choice prob-
liquefaction and dynamic behavior.” Physics and mechanics of soil abilities from choice-based samples.” Econometrica, 45(8), 1977–1988.
liquefaction, P. L. Lade and J. A. Yamamuro, eds., Balkema, Rotterdam, Marcuson, W. F., Hynes, M. E., and Franklin, A. G. (1990). “Evaluation
Netherlands, 91–102. and use of residual strength in seismic safety analysis of embankments.”
Finn, W. D. L., Lee, K., and Martin, G. R. (1977). “An effective stress Earthquake Spectra, 6(3), 529–572.
model for liquefaction.” J. Geotech. Eng. Div., 103(6), 517–533. Martin, G. R., Finn, W. D. L., and Seed, H. B. (1975). “Fundamentals
Finn, W. D. L., Martin, G. R., and Byrne, P. M. (1976). “Seismic response of liquefaction under cyclic loading.” J. Geotech. Eng. Div., 101(5),
and liquefaction of sands.” J. Geotech. Eng. Div., 102(8), 841–856. 423–438.
Fukutake, K., and Ohtsuki, A. (1995). “Three-dimensional liquefaction Matasovic, N. (2006). “D-MOD_2—a computer program for seismic re-
analysis of partially improved ground.” Proc., 1st Int. Conf. on Earth- sponse analysis of horizontally layered soil deposits, earthfill dams,
quake Geotechnical Engineering, Balkema, Brookfield, VT, 863–868. and solid waste landfills.” User’s Manual, GeoMotions, LLC, Lacey,
Ghionna, V. N., and Porcino, D. (2006). “Liquefaction resistance of undis- WA, 20.
turbed and reconstituted samples of a natural coarse sand from Matasovic, N., and Vucetic, M. (1993). “Seismic response of horizontally
undrained cyclic triaxial tests.” J. Geotech. Geoenviron. Eng., 132(2), layered soil deposits.” Rep. No. ENG 93-182, School of Engineering
194–202. and Applied Science, Univ. of California, Berkeley, CA.
Green, R. A. (2001). “Energy-based evaluation and remediation of liquefi- Nemat-Nasser, S., and Shokooh, A. (1979). “A unified approach to densi-
able soils.” Ph.D. dissertation, Civil Engineering, Virginia Polytechnic fication and liquefaction of cohesionless sand in cyclic shearing.” Can.
Institute and State Univ., Blacksburg, VA. Geotech. J., 16(4), 659–678.
Green, R. A., Mitchell, J. K., and Polito, C. P. (2000). “An energy-based Oda, M., Kawamoto, K., Suzuki, K., Fujimori, H., and Sato, M. (2001).
excess pore-water pressure generation model for cohesionless soils.” “Microstructural interpretation on reliquefaction of saturated granular

322 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012

J. Geotech. Geoenviron. Eng. 2012.138:309-323.


soils under cyclic loading.” J. Geotech. Geoenviron. Eng., 127(5), Tokimatsu, K., and Seed, H. B. (1984). “Simplified procedures of the evalu-
416–423. ation of settlements in clean sands.” Rep. No. UCB/GT-84/16, Univ. of
Park, S. S., and Byrne, P. M. (2004). “Practical constitutive model for soil California, Berkeley, CA.
liquefaction.” 9th Int. Symp. on Numerical Models in Geomechanics Tokimatsu, K., and Yoshimi, Y. (1983). “Empirical correlation of soil lique-
(NUMOG IX), Taylor & Francis Group, London, 181–186. faction based on SPT N-value and fines content.” Soils Found., 23(4),
Park, I. J., and Desai, C. S. (2000). “Cyclic behavior of sand using disturbed 56–74.
state concept.” J. Geotech. Geoenviron. Eng., 126(9), 834–846. Ueng, T-S., Sun, C-W., and Chen, C-W. (2004). “Definition of fines and
Pastor, M., Zienkiewicz, O. C., and Chan, A. H. C. (1990). “Generalized liquefaction resistance of Maoluo River sand.” Soil Dyn. Earthquake
plasticity and the modeling of soil behavior.” Int. J. Numer. Anal. Meth- Eng., 24(9–10), 745–750.
ods Geomech., 14(3), 151–190. Vaid, Y. P., and Chern, J. C. (1983). “Effect of static shear on resistance of
Polito, C. P., Green, R. A., and Lee, J. (2008). “Pore pressure generation liquefaction.” Soils Found., 23(1), 47–60.
models for sands and silty soils subjected to cyclic loading.” J. Geotech. Vucetic, M., and Dobry, R. (1986). “Pore pressure build-up and liquefac-
Geoenviron. Eng., 134(10), 1490–1500. tion at level sandy sites during earthquakes.” Research Rep. CE-86-3,
Porcino, D., and Caridi, G. (2007). “Pre- and post-liquefaction response of Dept. of Civil Engineering, Rensselaer Polytechnic Institute, Troy, NY.
sand in cyclic simple shear.” GEO-Denver 2007: New Peaks in Geo-
Wang, G. J., Takemura, J., and Kuwano, J. (1997). “Evaluation of excess
Downloaded from ascelibrary.org by New York University on 05/12/15. Copyright ASCE. For personal use only; all rights reserved.

technics, Dynamic Response and Soil Properties (GSP 160), Denver.


pore-water pressures of intermediate soils due to cyclic loading by
Prakash, S. (1981). Soil dynamics, McGraw-Hill, New York, 419.
energy method.” Computer methods and advances in geomechanics,
Prevost, J. H. (1985). “A simple plasticity theory for frictional cohesionless
J. X. Yuan, ed., Balkema, Rotterdam, Netherlands, 2215–2220.
soils.” Soil Dyn. Earthquake Eng., 4(1), 9–17.
Riemer, M. F., Gookin, W. B., Bray, J. D., and Arango, I. (1994). “Effects Wang, J., and Kavazanjian, E. (1985). “Pore-water pressure development in
of loading frequency and control on the liquefaction behavior of non-uniform cyclic triaxial tests.” John A. Blume Earthquake Engineer-
clean sands.” Geotechnical Eng. Report No. UCB/GT/94-07, Univ. ing Center Rep. No. 73, Technical Dept. of Civil and Environmental
of California, Berkeley, CA. Engineering, Stanford Univ., Stanford, CA.
Seed, H. B., Martin, P. P., and Lysmer, J. (1975). “The generation and dis- Wang, Z. L., Dafalias, Y. F., and Shen, C. K. (1990). “Bounding surface
sipation of pore-water pressures during soil liquefaction.” Geotechnical hypoplasticity model for sand.” J. Eng. Mech., 116(5), 983–1001.
Report No. EERC 75-26, Univ. of California, Berkeley, CA. Wu, J., Seed, R. B., and Pestana, J. M. (2003). “Liquefaction triggering
Shamoto, Y., Zhang, J., and Tokimatsu, K. (1998). “New charts for predict- and post liquefaction deformations of Monterey 0=30 sand under
ing large residual post-liquefaction ground deformation.” Soil Dyn. uni-directional cyclic simple shear loading.” Geotechnical Engineering
Earthquake Eng., 17(7–8), 427–438. Research Rep. No. UCB/GE-2003/01, Univ. of California, Berkeley,
Silver, M. L., and Seed, H. B. (1971). “Volume changes in sands during CA.
cyclic loading.” J. Soil Mech. Found. Div., 97(9), 1171–1182. Yamazaki, F., Towhata, I., and Ishihara, K. (1985). “Numerical model for
Sriskandakumar, S. (2003). “Characterization of cyclic loading response of liquefaction problem under multi-directional shearing on horizontal
air-pluviated Fraser River sand for validation of numerical models.” plane.” Proc., 5th Int. Conf. on Numerical Methods in Geomechanics,
M.Sc. thesis, Univ. of British Columbia, Vancouver, Canada. Vol. 1, Taylor & Francis Group, London, 399–406.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / MARCH 2012 / 323

J. Geotech. Geoenviron. Eng. 2012.138:309-323.

You might also like