Vous êtes sur la page 1sur 19

873

Seismic liquefaction, lateral spreading, and flow


slides: a numerical investigation into void
redistribution
Mahmood Seid-Karbasi and Peter M. Byrne

Abstract: Experience from past earthquakes indicates that seismically induced large lateral spreads and flow slides in
alluvial sand deposits have taken place in coastal and river areas in many parts of the world. The ground slope in these
slides was often not very steep, gentler than a few percent. Recent research indicates that the presence of low-permeability
silt or clay sublayers within the sand deposits is responsible for this behaviour. Such layers form a barrier to upward
flow of water associated with earthquake-generated pore pressures. This causes an accumulation of pore water at the
base of the layers, resulting in greatly reduced strength and possible slope instability. This paper uses an effective
stress coupled stress-flow dynamic analyses procedure to demonstrate the effects of a low-permeability barrier layer on
ground deformations from an earthquake event. The analyses show that an expansion zone develops at the base of barrier
layers in stratified soil deposits under seismic loading which can greatly reduce shear strength and result in large
deformations and flow failure. Without such a layer or layers, the slope may undergo significant displacements, but not
a flow slide. Slopes with a barrier layer can be stabilized by drains.
Key words: liquefaction, lateral spreads, stratification, flow failure, dynamic analysis, UBCSAND model, drain.
Résumé : L’expérience antérieure des tremblements de terre indique que de grands étalements latéraux et des coulées
dans les dépôts de sable alluvionnaires se produisent dans des régions côtières et de rivières dans plusieurs parties du
monde. La pente du terrain dans ces glissements n’est souvent pas très abrupte, plus douce que quelques pour cent. La
recherche récente indique que la présence de sous-couches de silt ou d’argile de faible perméabilité dans les dépôts de
sable est responsable de ce comportement. De tels couches forment une barrière pour l’écoulement de l’eau vers le
haut associé avec les pressions interstitielles générées par le tremblement de terre. Ceci cause une accumulation de pression
interstitielle à la base des couches, et résulte en une résistance grandement diminuée et une instabilité potentielle des
pentes. Cet article utilise une procédure d’analyses en contrainte effective couplée à la dynamique contrainte-écoulement
pour montrer les effets d’une couche-barrière de faible perméabilité sur la déformation causée par un événement de
tremblement de terre. Les analyses montrent qu’une zone d’expansion se développe à la base des couches-barrières dans
les dépôts de sol stratifié sous un chargement séismique qui peut réduire grandement la résistance au cisaillement et
résulter en de grandes déformations et en une rupture par écoulement. Sans une telle couche ou de telles couches, la
pente peut subir des déplacements appréciables, mais pas un glissement par écoulement. Les pentes avec des couches-
barrières peuvent être stabilisées par drains.
Mots-clés : liquéfaction, étalements latéraux, stratification, rupture par écoulement, analyse dynamique, modèle
UBCSAND, drain.

[Traduit par la Rédaction] Seid-Karbasi and Byrne 890

Introduction (Kokusho 2003). Submarine slides have been seismically trig-


gered in many regions as reported by Scott and Zuckerman
Experience from past earthquakes indicates that lateral (1972) and Hamada (1992). More interestingly, lateral spreads
spreads and flow slides have taken place in liquefied ground in or flow slides have occurred not only during earthquake shak-
coastal and river areas in many regions of the world, including ing, but also after earthquake shaking. These large movements
Alaska, Japan (Niigata), and Turkey. Movements may exceed are mainly driven by gravity, although they are initially trig-
several metres, even in gentle slopes of less than a few percent gered to liquefy by seismic stresses.
In the Hakusan District in Niigata City, Japan, an area
Received 3 November 2004. Accepted 29 January 2007. 250 m × 150 m bordering the Shinano River moved toward the
Published on the NRC Research Press Web site at cgj.nrc.ca
on 14 August 2007.
river during the 1964 Niigata earthquake (Kawakami and
Asada 1966). Girders of the Showa Ohashi Bridge fell down
M. Seid-Karbasi1,2 and P.M. Byrne. Department of Civil due to flow failure of the liquefied riverbed, which moved the
Engineering, University of British Columbia, 6250 Applied supporting pile foundations. Eyewitnesses reported that the
Science Lane, Vancouver, BC V6T 1Z4, Canada. girders began to fall a few minutes after the earthquake motion
1
Corresponding author (e-mail: mseidkarbasi@golder.com). had ceased (Hamada 1992). At the lower San Fernando dam, a
2
Present address: Golder Associates Ltd., 500-4260 Still hydraulic fill in California, flow failure due to liquefaction in
Creek Drive, Burnaby, BC V5C 6C6, Canada. the upstream slope occurred about a minute after the strong
Can. Geotech. J. 44: 873–890 (2007) doi:10.1139/T07-027 © 2007 NRC Canada
874 Can. Geotech. J. Vol. 44, 2007

Fig. 1. Cross section of a sand volcano formed on the surface of prise many sublayers as a result of the sedimentation pro-
the disposal pond (modified from Ishihara 1984). cess. A number of researchers have examined the effect of
layering on post-liquefaction sliding, including Scott and
Zuckerman (1972), Huishan and Taiping (1984), Liu and
Qiao (1984), Elgamal et al. (1989), Adalier and Elgamal
(1992), Fiegel and Kutter (1992), Kokusho (1999, 2000),
Kulasingam et al. (2001, 2004), Malvick et al. (2002, 2005,
2006), Yang and Elgamal (2002), Kulasingam (2003), Seid-
Karbasi and Byrne (2004a), Malvick (2005), Byrne et al.
(2006), and Seid-Karbasi3. Based on physical model tests
and site investigations, Kokusho (1999) and Kokusho and
Kojima (2002) concluded that failure can be caused by the
formation of a water film at the base of a sublayer leading to
a zone of essentially zero strength. The water-film effect is
associated with upward flow of water arising from liquefac-
tion. When such flow is impeded by a lower permeability
layer, it results in a void redistribution and an accumulation
of pore water that can lead to the formation of a water film
shaking in the 1971 earthquake (Seed 1987). Mochikoshi tail- or a thin water-rich zone.
ings dams in Japan failed as a result of the 1978 Izu-Ohshim- Kulasingam et al. (2004) using centrifuge testing demon-
Kinkai earthquake due to liquefaction-induced flow slides, strated that water films are an extreme case of void redistri-
causing release of the tailings (Ishihara 1984). One dam failed bution but are not a requisite for large ground deformation to
during the shaking, and a second dam failed 24 h later. Figure 1 occur. Under postseismic conditions, the void expansion
shows a sand boil formed in the Mochikoshi disposal pond may cause the strength to drop significantly and become
comprising sandy silt and silt layers. More recently, Harder and smaller than the driving stress, thereby resulting in large
Stewart (1996) reported that in the 1994 Northridge, California, ground deformations. The permeability contrast to cause
earthquake, liquefaction-induced flow failure of the 24 m high such an effect can be associated with silt and (or) clay layers
Tapo Canyon tailings dam occurred about 10 min after the sandwiched within sand and (or) gravel layers. Layering and
main shock. Yoshida et al. (2005) reported large lateral associated void ratio change during and or after an earth-
displacements in the order of several metres that occurred in quake can explain why the steady-state strength approach
near-level ground conditions in the 1983 Nihonkai–Chubu based on an undrained condition (Poulos et al. 1985) can
earthquake, with the movements obtained by comparing aerial lead to significantly higher values of residual strength than
photographs taken before and after the earthquake. those estimated from back analysis of field case studies
Flow failures in liquefied soil deposits have also occurred (e.g., Seed and Harder 1990; Stark and Mesri 1992; Olson
in submarine slopes triggered by earthquakes. The slopes in and Stark 2002; among others). This has already been
these slides were gentle, normally less than 5° and some- pointed out by a number of researchers (Seed 1987; Castro
times less than 1° (Hampton and Lee 1996). During the et al. 1989; Byrne and Beaty 1997; Kokusho 2003; Seid-
1964 Alaska earthquake, the port cities of Valdez and Karbasi and Byrne 2004a; Byrne et al. 2006).
Seward suffered great loss of life and property from large- Our understanding of the behaviour of liquefiable soils
scale submarine slides (Coulter and Migliaccio 1966; Lemke has increased dramatically in the past 30 years because of
1967). In the 1999 Kocaeli earthquake, Turkey, widespread (i) observations from field case histories, (ii) extensive labo-
large lateral spreads are reported by Kokusho (2003). Coastal ratory testing of soil elements under monotonic and cyclic
areas along Izmit Bay suffered submarine slides triggered by loading, (iii) model testing of earth structures under simu-
the earthquake. The 1995 Aegion earthquake of magnitude lated earthquake loading, and (iv) development of numerical
(Ms) 6.2 caused intense ground failures in Eratini, Greece modeling procedures.
(Bouckovalas et al. 1999). Geotechnical exploration revealed Laboratory model testing suggests that slopes comprised
that the soil profile at the failure sites is characterized by a of clean sands and having relative densities, Dr, as loose as
continuous interchange of silty sand and clay layers. The 20% are unlikely to suffer a flow slide (Kulasingam et al.
ground inclination in these cases was less than 7°. During 2004). These slopes can be readily triggered to liquefy and
the 1980 Mammoth lakes earthquake in California, a 2 km × may undergo large displacements, but their strengths are
20 km area of seafloor, consisting of sand and mud on a generally adequate to prevent a flow slide. However, if the
slope of 0.25°, suffered a flow slide. There was evidence of sands contain low-permeability silt layers that impede drain-
liquefaction in the form of sand boils on the seafloor as age, expansion of the sand skeleton beneath the silt occurs
described by Field et al. (1982). and can lead to the formation of a water film and complete
Although lateral flow failures have been reported in past loss of strength and a flow slide.
earthquakes causing damage to structures (e.g., Berrill et al. Traditionally, drainage conditions in laboratory investiga-
1997), the mechanism leading to large lateral displacements tions to characterize liquefiable soils are considered as either
is still poorly understood. Sand and silt deposits often com- undrained or (fully) drained, but these conditions do not rep-
3
Effects of void redistribution on liquefaction-induced ground deformations in earthquakes: a numerical investigation. Ph.D. dissertation. In
preparation.

© 2007 NRC Canada


Seid-Karbasi and Byrne 875

resent the real situation for stratified deposits in the field be- Fig. 2. Partially drained instability of loose Fraser River sand
cause during and after shaking water migrates from zones (modified from Vaid and Eliadorani 1998): (a) stress paths;
with higher excess pressure towards zones with lower excess ′ , minor
(b) strain paths. Drc, relative density at consolidation; σ3c
pressure. principal stress at consolidation.
In this paper a dynamic coupled stress-flow analysis pro-
cedure is utilized to study the mechanisms involved in large
deformations leading to lateral spreads and flow failures in
gentle slopes in liquefiable ground. Based on such analyses,
implications for design of mitigation methods to resist seismic
loading are examined.

Sand liquefaction and failure


Seismic liquefaction refers to a sudden loss in stiffness
and strength of soil due to cyclic loading effects of an earth-
quake. The loss arises from a tendency for soil to contract
under cyclic loading, and if such contraction is prevented or
curtailed by the presence of water in the pores that cannot
escape, it leads to a rise in pore-water pressure and a resulting
drop in effective stress (Silver and Seed 1971; Martin et al.
1975). If the effective stress drops to zero (100% pore-water
pressure rise), the strength and stiffness also drop to zero
and the soil behaves as a heavy liquid. Unless the soil is
very loose (e.g., Dr ≤ 20%), however, it will dilate and regain
some stiffness and strength as it strains. If this strength is
sufficient, it will prevent a flow slide from occurring but
may still result in excessive displacements commonly re-
ferred to as lateral spreading.
The excess pore water generated by seismic loading gen-
erally drains upwards, and if no barrier layers are present,
dissipation results in a decrease in void ratio and an increase
in strength and stability. The potential for lateral spreading
and flow slides can greatly increase if low-permeability layers, In these tests, expansive volumetric strains (ε v ) were
e.g., silt layers within a soil deposit, impede drainage, form- imposed by injection of water into the samples at a constant
ing a barrier to flow that can result in an expansion of the rate dε v / dε1 = − 0.4, where ε1 is the axial strain. The samples
sand skeleton (void redistribution) and accumulation of pore were stable under the initial stress state. The stress paths fol-
water at the base of the layers. lowed during injection indicate a reduction in effective
The majority of the laboratory element tests conducted to stresses at a constant shear stress. At each initial confining
assess liquefaction resistance of sands have been carried out stress, the stress changes leading to instability occurred with
under undrained conditions. Pore-pressure rise in these ele- little change in shear stress and void ratio and at very small
ments occurs under applied shear stress, causing a reduction axial strains of the order of 0.1%. Positive pore pressures con-
in effective stresses and strength. In recent experimental tinued to develop even beyond the phase-transformation line.
studies, Vaid and Eliadorani (1998) and Eliadorani (2000) dem- This occurs because the rate of imposed expansive volumetric
onstrated that a small net flow of water into an element strain is greater than the tendency for the skeleton to expand.
(injection) causing it to expand results in additional pore- Chu (1991), Bobei and Lo (2003), Sento et al. (2004), and
pressure generation and further reduction in strength. They Yoshimine et al. (2006) have reported similar results for sands
examined this phenomenon by injecting or removing small and silty sand materials. As a result, soil elements can liquefy
volumes of water from the sample as it was being sheared due to expansive volumetric strains that cannot be predicted
and referred to this as a partially drained condition. from analyses based on the results of undrained tests.
The results of inflow tests on Fraser River sand indicate a To investigate the effects of low-permeability layers on
potential for triggering liquefaction at constant shear stress. ground deformations due to earthquakes, it is necessary to
Small imposed expansive volumetric strains resulted in effec- predict the generation, redistribution, and dissipation of
tive stress reduction and flow failure of samples of sand con- excess pore pressures during and after earthquake shaking.
solidated to an initial state of stress corresponding to A fundamental approach requires a dynamic coupled stress-
Rc = σ1c′ / σ′3c = 2, as shown in Fig. 2a, where Rc is the flow analysis. In such an analysis, the volumetric strains are
consolidation effective stress ratio, and σ′1c and σ′3c are the controlled by the compressibility of the pore fluid and flow
major and minor principle effective stresses, respectively. Vaid of water through the soil elements. To predict the instability
and Eliadorani (1998) defined this condition as instability that and liquefaction flow, an effective stress approach based on
occurs when a soil element subjected to small effective stress an elastic–plastic constitutive model (UBCSAND) was used.
perturbation cannot sustain the current stress state and results in The model is calibrated against laboratory element data and
runaway deformation as seen in Fig. 2b, or liquefaction flow. centrifuge data and is described in the next section.

© 2007 NRC Canada


876 Can. Geotech. J. Vol. 44, 2007

Stress–strain model for sand Fig. 3. Moving of yield loci with stress ratio and plastic strain
increment vectors. φcv , friction angle at constant volume; εpv,
The UBCSAND constitutive model is based on the elastic– plastic volumetric strain.
plastic stress–strain model proposed by Byrne et al. (1995) and
has been further developed by Beaty and Byrne (1998) and
Puebla (1999). The model has been successfully used in ana-
lyzing the CANLEX liquefaction embankments (Puebla et al.
1997) and predicting the failure of the Mochikoshi tailings dam
(Seid-Karbasi and Byrne 2004b). It has been used to examine
dynamic centrifuge test data (e.g., Byrne et al. 2004; Seid-
Karbasi et al. 2005) and also partially saturation effects (Seid-
Karbasi and Byrne 2006). It is an incremental elastic–plastic
model in which the yield loci are lines of constant stress ratio
(η = τ / σ′, where η is the shear stress ratio, τ is the shear stress,
and σ′ is the effective normal stress on the plane of maximum
shear stress). The flow rule relating the plastic strain increment
directions is nonassociated (see Fig. 3) and leads to a plastic
potential defined in terms of the dilation angle.
Fig. 4. Plastic shear strain increment and shear modulus.
Elastic properties
The elastic component of response is assumed to be iso-
tropic and specified by a shear modulus, G e , and a bulk
modulus, B e , as follows:
ne
 σ′ 
[1] G e
= K Ge Pa  

 Pa 

[2] Be = α G e

where K Ge is an elastic shear modulus number, Pa is the


atmospheric pressure, σ′ = ( σx′ + σ′y ) / 2, n e is an elastic expo-
nent approximately equal to 0.5, and α ( = 2 (1 + ν) / 3 (1 − 2ν))
depends on Poisson’s ratio, ν, and ranges from 2/3 to 4/3.

Plastic properties
The plastic shear strain increment d γ p is related to the
stress ratio, dη, where η = τ / σ′, as shown in Fig. 4 and can
be expressed as follows:
whereas no compaction is predicted at stress ratios corre-
dη sponding to ϕ cv (plastic potential vector is vertical). For
[3] dγ p =
G p / σ′ stress ratios greater than ϕ cv , shear-induced plastic expan-
sion or dilation is predicted (where plastic potential vectors
where G p is the plastic shear modulus and is given by a slope to the left).
hyperbolic function as follows: This simple flow rule is in close agreement with the char-
2
acteristic behaviour of sand observed in drained laboratory
 η   element testing. The response of sand is controlled by the
= G ip 1 −  Rf 
p
[4] G  skeleton behaviour outlined previously. The presence of a
 η f   fluid (air–water mix) in the pores of the sand acts as a volu-
where G ip is the plastic shear modulus at a low stress ratio metric constraint on the skeleton if drainage is fully or par-
level (η = 0); η f is the stress ratio at failure and equals sin tially curtailed. It is this constraint that causes the pore
ϕ f , where ϕ f is the peak friction angle; and Rf is the failure pressure rise that can lead to liquefaction. Provided the skel-
ratio. G ip in turn is related to G e and the relative density of eton or drained behaviour is appropriately modeled under
the sand. The associated increment of plastic volumetric monotonic and cyclic loading conditions and the stiffness of
strain, dε pv , is related to the increment of plastic shear strain, the pore fluid and drainage are accounted for, the liquefac-
d γ p , through the flow rule as follows: tion response can be predicted.
This model was incorporated in the commercially avail-
[5] dε v = dγ p (sin ϕ cv − η ) able computer code FLAC (fast lagrangian analysis of con-
tinua), version 4.0 (Itasca Consulting Group Inc. 2005). This
where ϕ cv is the friction angle at constant volume or phase program models the soil mass as a collection of grid zones
transformation. It can be seen from Fig. 3 that significant or elements and solves the coupled stress flow problem using
shear-induced plastic compaction is occurring (where plastic an explicit time stepping approach. The program has a number
potential vectors slope to the right) at low stress ratios, of built-in stress–strain models including an elastic–plastic
© 2007 NRC Canada
Seid-Karbasi and Byrne 877

Fig. 5. Comparison of predicted and measured response for Fraser Fig. 6. Predictions of element response in undrained and partially
River sand: (a) stress–strain; (b) excess pore pressure ratio, Ru, drained (inflow) triaxial tests for Fraser River sand: (a) stress–
versus number of cycles; (c) CSR versus number of cycles (tests strain; (b) volumetric strains; (c) stress paths (modified from
data from Sriskandakumar 2004). Atigh and Byrne 2004).

calibrated to reproduce the National Center for Earthquake


Engineering Research (NCEER) 1997 (Youd et al. 2001)
triggering chart, which in turn is based on field experience
during past earthquakes and is expressed in terms of the
standard penetration test resistance value, ( N 1 ) 60 . The model
properties to obtain such an agreement are therefore ex-
pressed in terms of (N1)60.

Mohr–Coulomb model, and UBCSAND is a variation of this Model simulation of laboratory element tests
model in which friction and dilation angles are varied to
incorporate the yield loci and flow rule described previ- The model was applied to simulate cyclic simple shear
ously. Pore fluid stiffness and Darcy hydraulic flow are ba- tests under undrained conditions. Figure 5 shows model pre-
sic to the FLAC program, so only the skeleton stress–strain dictions along with test results on Fraser River sand. The
relation is needed to simulate liquefaction, and that is what test had an initial vertical consolidation stress σ′v = 100 kPa
the UBCSAND model does. Drainage conditions are built and Dr = 40%. The results in terms of stress–strain and
into FLAC, and drained, undrained, or coupled stress flow excess pore pressure ratio, R u , compare reasonably well with
conditions are specified by the user. the laboratory data. A comparison of model predictions with
The key elastic and plastic parameters can be expressed in tests results in terms of required number of cycles to trigger
terms of relative density, Dr , or normalized standard pene- liquefaction for different cyclic stress ratios, CSR, is shown
tration test values, ( N 1 ) 60 . Initial estimates of these parame- in Fig. 5c and shows good agreement.
ters have been approximated from published data and model The model was also used to predict the effect of both the
calibrations. The response of sand elements under mono- undrained condition and partial drainage as observed in
tonic and cyclic loading can then be predicted and the results triaxial monotonic tests. The partial drainage involved inject-
compared with laboratory data. In this way, the model can ing the sample with water to expand its volume as it was
be made to match the observed response over the range of sheared. The injection causes a drastic reduction in strength.
relative density or (N 1)60 values. The model has also been In the numerical model, the same volumetric expansion was
© 2007 NRC Canada
878 Can. Geotech. J. Vol. 44, 2007

Fig. 7. Ground conditions used in this study: (a) case I, uniform profile without low-permeability sublayer; (b) case II, profile with
low-permeability sublayer; (c) case III, profile treated with drain column.

(a) Groundwater table (b) Material: (c)


(m)
10 Silt barrier
Loose sand
9 Drain column

5
Element:
(1,13)
4 (1,10)
(1,5)
3 (1,3)

-1
Firm impervious ground
-1 0 1 2 3 4 (m)

Table 1. Properties of the materials used in the analyses.


Dry density, ρd Porosity, UBCSAND Permeability,
Material (×1000 kg/m3) n (N1)60 k (m/s)
Loose soil, Dr = 40% 1.50 0.448 6.2 8.81×10–4
Silt barrier 1.50 0.448 — 8.81×10–7

applied and the results are shown in Figs. 6a, 6b, and 6c in Fig. 8. Acceleration–time history for base input harmonic motion.
terms of stress–strain, volumetric strain versus mean stress,
and stress path, respectively (model prediction with the solid
line). The predictions are in remarkably good agreement
with the measured data.
The previous simulations illustrate that the model can gen-
erate the appropriate pore pressures and stress–strain response
to undrained loading and can account for the effect of volu-
metric expansion caused by inflow of water into an element.

Soil profile used in the analyses


The soil profile used in this study is a 10 m thick deposit
representing a submerged ground condition as shown in
Fig. 7a. It comprises a loose sand deposit resting on an UBCSAND ( N 1 ) 60 value. The low-permeability silt layer
impermeable rigid foundation. The effect of a low- barrier is simulated with a Mohr–Coulomb model having a
permeability layer within the loose sand at a depth of 4 m is friction angle ϕ = 30° and permeability k one thousand times
examined (see Fig. 7b). Fraser River sand with a relative lower than that of the loose sand layer. Its stiffness in terms
density of Dr = 40% is considered to represent the loose of bulk modulus and shear modulus was modeled as 1.0 ×
sand. Materials properties are listed in Table 1, in which ρ d , 104 and 0.5 × 104 kPa, respectively. It is not considered to
n, and k are the material dry density, porosity, and perme- generate excess pore pressure.
ability, respectively. The UBCSAND model was applied to Input base motion in terms of an acceleration–time history
the loose sand layer with a corresponding equivalent is shown in Fig. 8. It is a harmonic (sinusoid) excitation
© 2007 NRC Canada
Seid-Karbasi and Byrne 879

Fig. 9. Excess pore pressure ratio Ru versus time at selected points (elements (1, 13), (1, 10), and (1, 3), see Fig. 7) with increasing
depth: (a) case I; (b) case II.

applied at the base of the soil layer. It ramps up to 2.5 m/s2 layer was located at a depth of 4 m and had a permeability
within 1 s and dies out in 2 s, lasting for 7 s in total. 1000 times lower than that of the loose sand. The predicted
Analyses were conducted for three cases: (I) sloping behaviour for these cases in terms of excess pore pressure,
ground without a low-permeability layer, (II) sloping ground surface lateral displacement, and deformation pattern are
with a low-permeability layer, and (III) sloping ground with presented and compared in this section.
a low-permeability layer treated with a drain column. Figure 9 shows the time histories of excess pore pressure
ratio, Ru, for selected depths for the two cases (for positions
Analyses and results of the points refer to Fig. 7). The predicted patterns of
excess pore pressure in case I (see Fig. 9a) indicate that
To model the free-field condition, a mesh with 9 × 22 essentially high excess pore pressures build up within the
zones as illustrated in Fig. 7 was used. Material types are soil profile during the strong shaking. Small dilation spikes
recognized with different permeability values as shown in are predicted when R u ≈ 1 and are less pronounced in the
the figure. The nodes on the left and right boundaries were upper parts as a result of upward inflow. It may be seen that
linked to force the soil column to deform as a shear beam. excess pore pressures dissipate somewhat more rapidly at
The earthquake motion was applied as a time history of depth, e.g., Ru at 10 s is 55% and 75% for elements (1, 3)
acceleration at the base of the mesh. and (1, 13), respectively. A similar trend has been observed
in centrifuge tests conducted for level and sloping ground
Ground behaviour in cases I and II (Taboada and Dobry 1993a, 1993b, 1998).
A uniform sloping ground condition with 1° inclination The corresponding Ru time histories for case II at different
without a low-permeability layer was analyzed as a bench- depths are shown in Fig. 9b. Again, pore pressures increase
mark condition (case I), and then the same soil profile with a very rapidly; however, beneath the barrier layer they remain
low-permeability layer (case II) was analyzed. The barrier high (Ru ≈ 100%) after the end of shaking, and dissipation
© 2007 NRC Canada
880 Can. Geotech. J. Vol. 44, 2007

Fig. 10. Deformation pattern of soil profile: (a) without barrier, case I (with maximum lateral displacement of 0.95 m after 14 s);
(b) with barrier (darker area), case II (with maximum lateral displacement of 1.75 m after 30 s).

Fig. 11. Surface lateral displacement (X-dis) versus time for pro- 22 cycles of harmonic motion with peak ground acceleration
files with and without a barrier. (PGA) = 0.23g at 2 Hz using viscose pore fluid. The soil used
in their tests was Nevada sand 120, deposited at Dr = 45%.
The deformation pattern after 30 s for case II is shown in
Fig. 10b. Comparing the pattern and magnitude of lateral
displacements for these cases (with and without a barrier
layer shown in Figs. 10a and 10b), it can be seen that the
magnitude has increased from 0.95 to 1.75 m and the pattern
is quite different, with a large slippage occurring at the base
of the barrier layer.
Time histories of horizontal displacement of the top sur-
face for the two cases (with and without a barrier layer) are
compared in Fig. 11, which indicates that for case I, without
the barrier, displacements occur during shaking and cease
shortly after the end of shaking. It can be seen that the sur-
occurs at greater depths. This indicates that water flows face displacements are much larger when a barrier layer is
from the greater depths towards the layer beneath the bar- present. They are larger during shaking and continue to
rier, causing higher excess pore pressures to last for a signif- increase after shaking ceases.
icantly longer time compared with the case without a Figure 12 shows a profile of volumetric strain beneath
barrier. The injected flow causes an expansion of the layer the barrier for case II and indicates that the lower two
beneath the barrier to occur at essentially zero effective thirds of the soil profile contracts while expansion occurs
stress and leads to large deformation. in the upper one third of the profile, with the highest rate
Figure 10 shows the predicted deformed meshes for the of expansion directly beneath the barrier. Therefore, un-
two cases. As seen from Fig. 10a, distortion in case I is pro- drained conditions do not exist, especially locally within
nounced at the base and tapers off towards the surface, the soil deposit beneath the barrier. This also suggests that
resulting in a maximum displacement of 0.95 m at the top undisturbed samples taken prior to the earthquake will not
surface. This pattern compares well with dynamic centrifuge be representative of conditions during and shortly after the
data reported by Sharp et al. (2003), who modeled a 10 m earthquake due to void redistribution resulting from the up-
uniform liquefiable layer with a 5.2° inclination shaken with ward flow of water.

© 2007 NRC Canada


Seid-Karbasi and Byrne 881

Fig. 12. Profile of volumetric strain beneath the barrier layer (ii) as a 0.1 m soil profile in a 100g field (model scale) fol-
(after 30 s). lowing appropriate conversion laws for modeling described
by Schofield (1981) and Kutter (1995). It is noted that the
mesh size used in the latter way of modeling was one-
hundredth that used in the former.
The base input motion and mechanical properties of materials
are the same as those used previously. The displacement pat-
terns were identical for both models when examined in
prototype scale and are the same as those shown in Fig. 10b
(the profiles of volumetric strain beneath the barrier layer for
both models were also identical and the same as those
depicted in Fig. 12). This suggests that for a soil profile
comprising a barrier layer, contraction at the lower parts and
expansion at the upper parts are characteristic behaviours
that occur due to pore pressure migration and result in void
redistribution regardless of layer size. These findings indi-
cate that the phenomenon of pore-water redistribution can be
captured in centrifuge tests and that the actual physical size
of the layers is not important. What matters from a numerical
modeling point of view is the size of the mesh in relation to
the size of the liquefiable layer.
Further studies were undertaken to analyze the problem
under the same conditions but using meshes of decreased
size for the element beneath the barrier. The results showed
that a decrease in element size results in greater volumetric
expansion for the very first element at the base of the barrier
layer. Figure 13a shows maximum volumetric strain of the
element beneath the barrier versus the element thickness
normalized with respect to soil layer thickness beneath the
barrier. The figure suggests that void redistribution leads to a
The increasing rate of expansion predicted in Fig. 12 as very thin, water-rich zone at the base of the barrier. If
the barrier is approached suggests that very high strains may enough water flows into the soil element, it can expand until
be occurring in the zone at or near the boundary. This is the steady or critical state at zero effective stress is reached.
referred to as strain localization. This corresponds approximately to the maximum void ratio
Similar analyses were also carried out for the same soil state. At this state the skeleton can undergo no further
profile including a barrier layer with zero inclination (level expansion, and this is simulated by setting the dilation angle
ground). In this case, no significant lateral displacements are to zero. Additional inflow will result in the formation of a
predicted because of zero static driving force or shear stress water film at the interface and zero shear strength.
bias for level ground conditions.
The predicted increasing trend in volumetric strain of the
barrier base element with a decrease in its size (shown in
Flow failure Fig. 13a) suggests that volumetric strain of the base element
The results from the previous section revealed that delayed becomes infinity as the size of the base element approaches
large deformations in liquefiable gently sloping grounds can zero. This implies that for a liquefiable layer with a sublayer
occur when a sublayer with low permeability is present. barrier, flow failure occurs regardless of other involved
However, two questions arise in this regard. (1) In view of factors, i.e., liquefied soil layer thickness, density, shaking
the predicted localization in the element beneath the barrier, level and duration, and ground slope.
to what extent is the predicted deformation pattern (i.e., con- Figure 13b shows time histories of volumetric strain for
traction in the lower section and expansion in the upper sec- the base element with various element thickness ratios
tion) related to the model configuration (mesh size effects)? (ETRs) for the analyzed sloping ground using the
(2) What are the requirements for a flow failure? UBCSAND model. It can be seen that the amount of volu-
The rest of this section discusses the results of analyses metric expansion in the base element increases with time
carried out to answer these questions. and with a decrease in element thickness. More detailed
The predicted strain localization implies that the com- analyses indicate that the volumetric expansion does not go
puted results can be mesh-size dependent, as also noted by to infinity at the boundary. Further analyses to examine flow
Yang and Elgamal (2002) and Uzuoka et al. (2003). To and expansion issues were carried out and are presented as
investigate scale effects on the predicted characteristic follows.
behaviour of liquefiable grounds with a sublayer barrier, a Figure 14 shows the analyses results in terms of the iso-
separate series of analyses was conducted for the same soil chrones of Y-Flow, the vertical specific discharge velocity
profile with a barrier but modeled in two different ways: (flow rate per unit area) at different time intervals for the
(i) as a 10 m soil profile in a 1g field (prototype scale), and mesh shown in Fig. 7b using the UBCSAND model. The

© 2007 NRC Canada


882 Can. Geotech. J. Vol. 44, 2007

Fig. 13. Volumetric strain of the base element with various thicknesses: (a) maximum volumetric strain versus normalized element
thickness; (b) time histories of volumetric strain.

figure indicates that water flows upward from the lower forms within the soil mass. In the next section an approach
parts that have the highest excess pore pressures towards the for numerical modeling of flow failures is presented.
upper low-permeability layer where the excess pore pressure
is lowest. Maximum flow occurs near the mid-height of the Approach to modeling localized flow failure
soil layer with a descending trend as it approaches the The effect of mesh size can be approximately accounted
boundaries both top and bottom. Flows initially increase for in the analysis by specifying a dilation cut-off on skeleton
quickly with time and then drop off as pore pressures dissi- expansion that is related to the initial relative density and
pate. Figure 15 shows the time history of volumetric strain mesh size. Denser material would have more dilation capacity
rate for the base element of the mesh shown in Fig. 10b. It to reach its maximum void ratio, at which point dilation
can be seen that the rate of expansion tends to essentially would be set to zero. For a coarser mesh, less expansion is
zero some time after shaking ceases. Similar results were predicted for the element and the amount of expansion
obtained for the meshes with smaller base zone thicknesses. required to trigger dilation cut-off would be less.
Figure 16 shows the maximum expansion rate versus nor- The mesh size should also be related to the particle size.
malized zone thickness. It clearly shows no indication of an Roscoe (1970) noted that a shear band may occur within a
infinite value for the expansion rate as the grid size thin zone with a thickness 10–20 times that of the mean
approaches zero. The results indicate that, although the grain size of the soil, D50 . Similar finding have been reported
expansion increases with a reduction in element thickness, from further experimental investigations (e.g., Alshibli and
its maximum value is finite. Sture 1999; Desrues and Viggiani 2004; Lu et al. 2004) and
From a practical point of view, localized flow failure occurs numerical investigations (e.g., district element method,
in a thin zone of soil when a shear band of limited thickness DEM) by a number of researchers. A reasonable lower limit
© 2007 NRC Canada
Seid-Karbasi and Byrne 883

Fig. 14. Vertical specific discharge (Y-Flow) isochrones at differ- Fig. 16. Maximum expansion rate versus base element thickness
ent times. ratio. Isochrones at different time intervals for UBCSAND model
with 0.5 m element height.

(2002) and Olson and Stark (2003). The threshold volumetric


strain, ( ε v ) T , can be calculated as follows:
e max − e min
[6] ( ε v ) T = Dr
1 + ei

where Dr is the soil relative density, e max is the maximum


void ratio, e min is the minimum void ratio, and e i is the initial
void ratio. The numerator of eq. [6] should be e ss − e min
(where e ss is the steady-state void ratio at zero effective
Fig. 15. Time history of volumetric strain rate of base element. stress), but it can be approximated by e max − e min for the sake
of simplicity. A similar approach was adopted by Tsukamoto
et al. (2004) and Sawada et al. (2006) to estimate potential
maximum liquefaction-induced settlement. In Table 2, the
potential maximum expansion is also provided for a typical
medium-dense state (Dr = 50%). Figure 18 shows the upper
and lower bounds for potential maximum expansion for a
complete range of initial Dr calculated based on eq. [6]. A
range of relative densities from 25% to 80% is of interest
from a practical point of view. For Fraser River sand at Dr
of 40%, (ε v )T = 6.9%.
For the case involved in Fig. 17 with Fraser River sand,
the predicted maximum volumetric strain in the shear band
zone was 5.2%. Since this is less than the threshold value,
i.e., ( ε v ) T = 6.9%, a flow slide is not predicted. If the expan-
sion limit is reached, the soil element deforms in a steady-
on mesh size could be based on the size of the shear band. state condition with essentially zero shear strength. Not all
For a soil profile comprising Fraser River sand of Dr = 40% elements along a potential slide surface will necessarily
with D50 = 0.3 mm, a shear band could form within a zone reach this state, and the low residual strengths back-
3–6 mm in height. In Fig. 17 the data shown in Fig. 13a are calculated from field experience are a reflection of the aver-
plotted in semilogarithmic format with the initial portion age strength on the failure surface.
extrapolated to reach a value corresponding to an element When the element thickness is chosen equal to the shear
thickness of 3 mm (ETR = 3/6000 = 0.0005). The maximum band thickness, there is no correction for element size. It is
volumetric strain predicted for that ETR is about 5.2%. If generally not practical to make the element height this small,
the shear band zone can expand by 5.2% before reaching its and a correction for element size is then needed. To examine
critical state, then it can retain some strength and prevent a this approach, the sloping-ground problem was analyzed using
flow slide. A key question then is as follows: What is the the mesh shown in Fig. 7b. This mesh has an ETR of 0.042
maximum volumetric expansion (threshold strain) that an compared with a localized failure or shear band zone ratio of
element can tolerate before a water film forms? 0.0005. Based on Fig. 13, the predicted volumetric strain in the
Potential maximum volumetric (threshold) strain can be element beneath the barrier would be 0.79% compared with
estimated based on data from various soils available in the 5.10% in the shear band. To account for this, the strain in the
literature. Table 2 lists a number of soils with their minimum element must be modified to ( ε v )* = ε v (5.10/0.79) = 6.4ε v ,
and maximum void ratios compiled by Cubrinovski and Ishihara where 6.4 is a mesh correction factor.
© 2007 NRC Canada
884 Can. Geotech. J. Vol. 44, 2007

Fig. 17. Estimate of maximum volumetric strain, ( εv )max , for a base element with 10D50 (0.003 m) thickness for Fraser River sand
using extended trend of ( εv )max with element thickness ratio.

Table 2. Minimum and maximum void ratio and potential maximum expansion of sands at Dr = 50%.
Void ratio
Expansion potential,
Soil emin emax (εv)50 (%) Reference
Fraser River sand 0.596 0.900 8.8 Vaid and Eliadorani 1998
Cambria sand 0.538 0.767 7.0 Lade et al. 1998
Nevada sand 50/80 0.581 0.858 8.0 Lade et al. 1998
Nevada sand 80/200 0.617 0.940 9.1 Lade et al. 1998
Nevada fines 0.754 1.178 10.8 Lade et al. 1998
Ottawa sand 50/200 0.550 0.805 7.6 Lade and Yamamuro 1997
Ottawa sand F-95 0.580 0.865 8.3 Lade and Yamamuro 1997
Host sand A2 0.60 0.98 10.6 Thevanayagam 1998
Toyoura sand 0.616 0.988 10.3 Zlatovic 1994
Ottawa sand 0.48 0.78 9.2 Salgado et al. 2000
Ottawa sand C-109 0.50 0.83 9.9 Pitman et al. 1994
Quiou sand 0.78 1.20 10.6 Pestana and Whittle 1995
Mine tailings sand 0.69 1.06 9.9 Vaid et al. 1985
Brasted sand 0.48 0.79 9.5 Cornforth 1974
Dune sand 0.54 0.91 10.7 Konrad 1990
Well-rounded silica sand 0.67 1.06 10.5 Konrad 1990
Nerlerk 0%–2% 0.62 0.94 9.0 Sladen et al. 1985
Tottri sand 0.938 1.008 10.2 Takeshita et al. 1995
Monterey No. 9 sand 0.53 0.86 9.7 Riemer et al. 1990
Massey Tunnel sand 0.712 1.102 10.2 Konrad and Pouliot 1997
Quebec sand 0.54 0.79 7.5 Konrad 1998
Sand B 0.50 0.84 10.2 Castro 1969
Sand C 0.66 0.99 9.0 Castro 1969
Sand A 1.23 1.88 12.7 Castro 1969

© 2007 NRC Canada


Seid-Karbasi and Byrne 885

Fig. 18. Maximum expansion potential versus initial relative density Fig. 20. Time histories of surface lateral velocity (X-vel) for two
for sands (based on Table 2 data and eq. [6]). events.

1.0° inclination treated with gravel drains penetrating


Fig. 19. Time histories of surface lateral displacement (X-dis) for through the low-permeability sublayer down to the loose
two events. sand layer at the bottom (case III).
For the study conducted in plane strain conditions, the
drains have been accounted for with a soil column having a
permeability 100 times greater than that of the surrounding
liquefiable soils. In a real prototype problem, the three-
dimensional effects of the drain columns pattern can be
treated in plane strain analysis using an equivalent drain area
approach (e.g., Indraratna and Redana 1997, 2000). Fig-
ure 21a shows the model (case III) with a drain. Figure 21b
shows distributions of maximum excess pore pressure ratio,
Rumax , within the model along with flow vectors after 3.5 s
of shaking. It is seen that Rumax increases with an increase in
distance from the drain column. The drain exhibits some
excess pore pressure during shaking. The maximum pre-
dicted lateral displacement in this case was negligible, even
compared with that of the model without a low-permeability
The analysis showed that the ( ε v )* so computed was less layer (case I). This demonstrates that the drain column is a
than the threshold value of 6.9%, and the results were promising measure to mitigate liquefaction-induced defor-
identical to those predicted in Figs. 11 and 12. When the mations.
duration of base motion was increased from 7 s to 12 s, the A comparison of predicted time histories of excess pore
computed ( ε v )* was greater than 6.9%, at which time dila- pressure buildup for a point at mid-depth of loose sand for
tion was suppressed and a flow slide was predicted. cases I, II, and III is shown in Fig. 22, which indicates that
Figure 19 shows the surface displacement – time histories drains can reduce generated excess pore pressures and signi-
for both cases. It can be seen that surface displacement for ficantly speed up the dissipation of excess pore pressure.
12 s motion is ever increasing, whereas displacements stop The effect of drains in decreasing excess pore pressure during
some time after the end of shaking in the 7 s case. Ground shaking is also reported by Chang et al. (2004) and Brennan
surface velocity is a good indicator of flow failure. Figure 20 and Madabhushi (2005) in field model tests and centrifuge
shows ground surface velocity versus time for the two tests, respectively. Figure 23 shows excess pore pressure ratios
events. For the 7 s event, the surface velocity decreases and measured in field tests within a liquefiable soil for a deposit
then becomes zero after the end of shaking, whereas for the without and with vertical drains subjected to a harmonic
12 s event it is essentially constant over that period. This re- excitation. These data show that the excess pore pressure
veals that flow failure takes place in the second case when build-up rate is significantly slower with drains and confirms
the dilation threshold is reached. the effectiveness of drains as a remediation measure.

Countermeasure for flow failure Practical implications


Drains have the potential to nullify the barrier effect and It was demonstrated that the applied dynamic procedure
curtail or prevent the occurrence of lateral spread or flow can reproduce the seismic response of liquefiable soil when
slides in the event of an earthquake. They can facilitate dis- a low-permeability sublayer is present. The presence of such
sipation of excess pore pressure and mitigate the impedance a sublayer impedes upward drainage that can lead to the for-
effects of low-permeability layers. This is examined for the mation of water films having essentially zero shear strength
same soil profile as that used in the previous analyses with and result in large ground deformation and flow failure. The

© 2007 NRC Canada


886 Can. Geotech. J. Vol. 44, 2007

Fig. 21. Treated model, case III: (a) central vertical drain; (b) Rumax and flow vectors at 3.5 s.

k (m/s) R umax
8.81 x 10-7 0.65
8.81 x 10-4 0.70
8.81 x 10-2 0.75
0.80
0.85
0.90
0.95
1.00
Vertical drain

(a) (b)

analyses also show that drains can be very effective in alle- sands contain low-permeability layers that impede drainage,
viating the destructive effects of sublayer barriers. a void expansion resulting in a water-rich thin zone or water
Important aspects for design are as follows: film may cause a near complete loss of strength of the soil
(1) Appropriate site investigation techniques should be applied directly beneath that layer and result in a flow slide. The low
to detect low-permeability, thin layers within sand and residual strengths based on back analyses of field case histo-
(or) gravel layers. ries and reported by Seed and Harder (1990) are likely a result
(2) If a low-permeability sublayer is present and significant of void expansion related to the presence of barrier layers.
liquefaction is predicted for the design earthquake, either A numerical approach is used in this paper that captures
low residual shear strengths consistent with field back element sand behaviour in monotonic and cyclic loading under
analysis should be used or drainage–densification should different drainage conditions; undrained and inflow were
be considered. The design of drains–densification can utilized to study the effects of low-permeability layers on
be optimized from a coupled stress-flow dynamic analysis. the sloping ground response during earthquake loading. The
(3) The perfect undisturbed sample obtained prior to an following conclusions are made based on the analysis results:
earthquake will not be representative of conditions during (1) The computed results show that for a homogeneous
and shortly after the earthquake due to expansion result- sand layer that is triggered to liquefy, upward flow
ing from the upward flow of water in stratified deposits. resulting from excess pore pressure causes dissipation
Dissipation of excess pore pressure some time after the and a reduction in void ratio within the sand layer. How-
earthquake will reconsolidate soil elements beneath the ever, if a low-permeability layer(s) impedes drainage,
barrier such that the perfect sample taken some time then contraction occurs at the base of the sand layer but
after the earthquake will also not be representative of expansion at the top where flow is impeded, and this
the critical conditions. expansion is responsible for flow failures. The occur-
rence of contraction and expansion zones, respectively,
Conclusions at the lower and upper zones due to pore pressure redis-
tribution is a characteristic behaviour of a liquefiable
Many failures of earth structures and submarine slides deposit with sublayer barriers.
have been reported during past earthquakes worldwide. A (2) Void redistribution results in a thin, water-rich zone at
number of civil structures and soil deposits in coastal or the base of a barrier layer that could ultimately form a
river areas have suffered large deformations during past water film when enough water is available for injection.
earthquakes as a result of soil liquefaction. The deformations The thickness of this shear band zone is related to particle
may exceed several metres, even in gentle slopes of less than diameter; if and when the predicted expansion exceeds
a few percent. Deformations occur not only during but also the threshold expansion, a water film will form, and this
after earthquake shaking. is simulated in the analyses by setting the dilation equal
Clean, loose (e.g., Dr ≥ 20%) sands are unlikely to suffer to zero.
a flow slide. Although they can be triggered to liquefy and (3) It is generally not practical to have the mesh size as
undergo large strains and displacements, their undrained small as the shear band thickness, in which case a cor-
strengths are generally adequate for stability. However, if the rection to the computed expansion is required.
© 2007 NRC Canada
Seid-Karbasi and Byrne 887

Fig. 22. Predicted time history of Ru at mid-depth of loose sand: (4) Most of the relative movements occur at the base of the
(a) without barrier layer; (b) with barrier layer; (c) with barrier low-permeability sublayer. Maximum displacements occur
layer treated with drain (Seid-Karbasi3). at or near the surface above the barrier layer.
(5) A large part of the deformations may occur some time
after shaking has ceased, depending on the time needed
for migration of water from zones with higher excess
pore pressures.
(6) Installation of vertical drains that penetrate the barrier
layers can mitigate the destructive effects of low-
permeability layers. This has been demonstrated in these
analyses and has also been observed from field model
tests.
(7) The design of these drains can be assessed from
dynamic coupled flow – effective stress analyses using
appropriate modeling parameters and design input
motions. The dimensions and location of remediation
measures can be optimized from dynamic analyses.

Acknowledgments
The authors acknowledge the financial support from BC
Hydro through the Professional Partnership program and the
support of the Natural Sciences and Engineering Research
Council of Canada (NSERC) through Strategic Liquefaction
Grant NSERC 246394 and NSERC COSTA Grant 03608-
CG068625. The authors are also grateful to Professor
D.L. Anderson and to Ernest Naesgaard for helpful discus-
sions during the course of this research. In addition, the au-
thors express their appreciation to the reviewers for their
constructive comments, which led to many changes from the
original manuscript.

References
Adalier, K., and Elgamal, A.W. 1992. Post-liquefaction behavior of
soil systems. Report, Department of Civil Engineering,
Rensselaer Polytechnic Institute, Troy, N.Y.
Alshibli, K., and Sture, S. 1999. Sand shear band thickness mea-
surements by digital imaging techniques. Computing in Civil
Engineering, ASCE, 13: 103–109.
Atigh, E., and Byrne, P.M. 2004. Liquefaction flow of submarine
slopes under partially undrained conditions: an effective stress
Fig. 23. Measured Ru in field liquefaction test with and without approach. Canadian Geotechnical Journal, 41(1): 154–165.
a drain (data from Chang et al. 2004). Beaty, M.H., and Byrne, P.M. 1998. An effective stress model for
predicting liquefaction behavior of sand. In Geotechnical Earth-
quake Engineering and Soil Dynamics III: Proceedings of a
Specialty Conference, Seattle, Wash., 3–6 August 1998. Edited
by P. Dakoulas, M. Yegian, and R.D. Holtz. ASCE Geotechnical
Special Publication 75, Vol. 1, pp. 766–777.
Berrill, S.A., Christensen, S.A., Keenan, R.J., Okada, W., and
Pettinga, J.R. 1997. Lateral-spreading loads on a piled bridge
foundation. In Proceedings of the 14th International Conference
on Soil Mechanics and Foundation Engineering, Discussion
Session, Seismic Behavior of Ground and Geotechnical Struc-
tures, Hamburg, Germany, 6–12 September 1997. A.A.
Balkema, Rotterdam. pp. 173–183.
Bobei, D.C., and Lo, S.-C. 2003. Strain path influence on the
behavior of sand with fines. In Soil and Rock America 2003:
Proceedings of the 12th Panamerican Conference on Soil
Mechanics and Geotechnical Engineering, Cambridge, Mass., 22–
26 June 3003. Edited by P.J. Culligan, H.H. Einstein, and A.J.
Whittle. Verlag Gluckauf GMBH, Essen, Germany. pp. 583–588.
© 2007 NRC Canada
888 Can. Geotech. J. Vol. 44, 2007

Bouckovalas, G.D., Gazetas, G., and Papadimitriou, A.G. 1999. Fiegel, G.L., and Kutter, B.L. 1992. Liquefaction mechanism for
Geotechnical aspects of the 1995 Aegion (Greece) earthquake. layered soil. Journal of Geotechnical Engineering, ASCE, 120:
In Proceedings of the 2nd International Conference on Earth- 737–755.
quake Geotechnical Engineering, Lisbon, 21–25 June. Edited by Field, M.E., Gardner, J.V., Jennings, A.E., and Edwards, B.D.
P. Seco e Pinto. A.A. Balkema, Rotterdam. pp. 739–748. 1982. Earthquake induced sediment failures on a 0.258 slope,
Brennan, A.J., and Madabhushi, S.P. 2005. Liquefaction and drainage Klamath River Delta. Journal of California Geology, 10: 542–546.
in stratified soil. Journal of Geotechnical and Geoenvironmental Hamada, M. 1992. Large ground deformations and their effects on
Engineering, ASCE, 131: 876–885. lifelines: 1964 Niigata Earthquake. In Case studies of liquefac-
Byrne, P.M., and Beaty, M.H. 1997. Post-liquefaction shear tion and lifeline performance during past earthquakes. Vol. 1.
strength of granular soils: theoretical/conceptual issues. In Pro- Japanese case studies. Edited by M. Hamada and T.D.
ceedings of the Workshop on Post-Liquefaction Shear Strength O’Rourke. NCEER Technical Report NCEER-92-0001, National
of Granular Soils, Urbana–Champion, Ill., 17–18 April 1997. Center for Earthquake Engineering Research (NCEER), Buffalo,
pp. 16–45. N.Y. pp. 3/1–3/123.
Byrne, P.M., Roy, D., Campanella, R.G., and Hughes, J. 1995. Hampton, M.A., and Lee, H.J. 1996. Submarine landslides. Reviews
Predicting liquefaction response of granular soils from pressure- of Geophysics, 34: 33–59.
meter tests. In Static and Dynamic Properties of Gravelly Soils: Harder, L., and Stewart, J. 1996. Failure of Tapo Canyon tailings
Proceedings of Sessions of the ASCE National Convention, San dam. Journal of Performance of Constructed Facilities, ASCE,
Diego, Calif., 23–24 October 1995. Edited by M.D. Evans and 10(3): 109–114.
R.J. Fragaszy. ASCE Geotechnical Special Publication 56, Huishan, L., and Taiping, Q. 1984. Liquefaction potential of sat-
pp. 122–135. urated sand deposits underlying foundation of structure. In
Byrne, P.M., Park, S., Beaty, M., Sharp, M., Gonzalez, L., and Proceedings of the 8th World Conference on Earthquake En-
Abdoun, T. 2004. Numerical modeling of liquefaction and com- gineering, San Francisco, Calif., 21–28 July 1984. Prentice–
parison with centrifuge tests. Canadian Geotechnical Journal, Hall Inc., Englewood Cliffs, N.J. Vol. 3, pp. 199–206.
41(1): 193–211. Indraratna, B., and Redana, I.W. 1997. Plane strain modeling
Byrne, P.M., Naesgaard, E., and Seid-Karbasi, M. 2006. Analysis smear effects associated with vertical drains. Journal of
and design of earth structures to resist seismic soil liquefaction. Geotechnical Engineering, ASCE, 119: 1321–1329.
In Proceedings of the 59th Canadian Geotechnical Engineering Indraratna, B., and Redana, I.W. 2000. Numerical modeling of
Conference, Hardy Lecture, Vancouver, B.C., 1–4 October 2006. vertical drains with smear and well resistance installed in soft
Canadian Geotechnical Society, Alliston, Ont. pp. 1–24. clay. Canadian Geotechnical Journal, 37(1): 132–145.
Castro, G. 1969. Liquefaction of sands. Ph.D. thesis, Harvard Uni- Ishihara, K. 1984. Post-earthquake failure of a tailings dam due to
versity, Cambridge, Mass. liquefaction of the pond deposit. In Proceedings of the Interna-
Castro, G., Keller, T.O., and Boynton, S.S. 1989. Re-evaluation of tional Conference of Case Histories in Geotechnical Engi-
the Lower San Fernando Dam. Contract Report GL-89-2, US Army neering, University of Missouri-Rolla, Rolla, Mo., May 1984.
Waterways Experiment Station, Vicksburg, Miss. Vols. 1 and 2. Vol. 3, pp. 1129–1143.
Chang, W.J., Rathje, E., Stokoe, K.H., II, and Cox, B.R. 2004. Itasca Consulting Group, Inc. 2005. Fast lagrangian analysis of
Direct evaluation of effectiveness of prefabricated vertical drain continua (FLAC), version 5, user’s guide. Itasca Consulting
in liquefiable sand. Soil Dynamics and Earthquake Engineering Group, Inc., Minneapolis, Minn.
Journal, 24: 723–731. Kawakami, F., and Asada, A. 1966. Damage to the ground and
Chu, J. 1991. Strain softening behavior of granular soils under earth-structures by the Niigata earthquake of June 16, 1964.
strain path testing. Ph.D. thesis, University of New South Wales, Soils and Foundations, 1: 14–30.
Kensington, NSW, Australia. Kokusho, T. 1999. Water film in liquefied sand and its effect on
Cornforth, D.H. 1974. One-dimensional consolidation curves of a lateral spread. Journal of. Geotechnical and Geoenvironmental
medium sand. Géotechnique, 24: 678–683. Engineering, ASCE, 125: 817–826.
Coulter, H.W., and Migliaccio, R.R. 1966. Effects of the earth- Kokusho, T. 2000. Mechanism for water film generation and lat-
quake of March 27, 1964 at Valdez, Alaska. US Geological eral flow in liquefied sand layer. Soils and Foundations, 40:
Survey, Professional Paper 542-C. 36 pp. 99–111.
Cubrinovski, M., and Ishihara, K. 2002. Maximum and minimum Kokusho, T. 2003. Current state of research on flow failure consid-
void ratio characteristics of sand. Soils and Foundations, 42: 65–78. ering void redistribution in liquefied deposits. Journal of Soil
Desrues, J., and Viggiani, G. 2004. Strain localization in sand: an Dynamics and Earthquake Engineering, 23: 585–603.
overview of the experimental results obtained in Grenoble using Kokusho, T., and Kojima, T. 2002. Mechanism for post-
stereophotogrammetry. International Journal for Numerical and liquefaction water film generation in layered sand. Journal of
Analytical Methods in Geomechanics, 28: 279–321. Geotechnical and Geoenvironmental Engineering, ASCE, 128:
Elgamal, A.W., Dobry, R., and Adalier, K. 1989. Small scale 129–137.
shaking table tests of saturated layered sand–silt deposits. In Konrad, J.-M. 1990. Minimum undrained strength of two sands.
Proceedings of the 2nd US–Japan Workshop on Soil Liquefac- Journal of Geotechnical Engineering, ASCE, 116: 932–947.
tion, Large Ground Deformation, and Their Effects on Life- Konrad, J.-M. 1998. Sand state from cone penetration tests. Géo-
lines, Grand Island and Ithaca, N.Y., 26–29 September 1989. technique, 48: 201–215.
Edited by T.D. O’Rourke and M. Hamada. NCEER Technical Konrad, J.-M., and Pouliot, N. 1997. Ultimate state of reconstituted
Report NCEER-89-0032, National Center for Earthquake Engi- and intact samples of deltaic sand. Canadian Geotechnical Journal,
neering Research (NCEER), Buffalo, N.Y. pp. 233–245. 34(5): 737–748.
Eliadorani, A.A. 2000. The response of sands under partially Kulasingam, R. 2003. Effects of void redistribution on
drained states with emphasis on liquefaction. Ph.D. thesis, Civil liquefaction-induced deformations. Ph.D. thesis, Civil and En-
Engineering Department, The University of British Columbia, vironmental Engineering Department, University of California,
Vancouver, B.C. Davis, Calif.

© 2007 NRC Canada


Seid-Karbasi and Byrne 889

Kulasingam, R., Malvick, E.J., Boulanger, R.W., and Kutter, B.L. Pitman, T.D., Robertson, P.K., and Sego, D.C. 1994. Influence of
2001. Void redistribution and localization of shear strains in fines on the collapse of loose sands. Canadian Geotechnical
model sand slopes with silt seams. Report on first year activities. Journal, 31(5): 728–739.
In Proceedings of the US–Japan Cooperative Research in Urban Poulos, S.J., Castro, G., and France, J.W. 1985. Liquefaction evalu-
Earthquake Disaster Mitigation Workshop, Seattle, Wash., 16– ation procedure. Journal of Geotechnical Engineering., ASCE,
18 August 2001. MCEER, N.Y. pp. 117–128. 111: 772–792.
Kulasingam, R., Malvick, E.J., Boulanger, R.W., and Kutter, B.L. Puebla, H. 1999. A constitutive model for sand analysis of the
2004. Strength loss and localization at silt interlayers in slopes CANLEX embankment. Ph.D. thesis, Civil Engineering Depart-
of liquefied sand. Journal of Geotechnical and Geo- ment, University of British Columbia, Vancouver, B.C.
environmental Engineering, ASCE, 130(11): 1192–1202. Puebla, H., Byrne, P.M., and Phillips, R. 1997. Analysis of
Kutter, B.L. 1995. Recent advances in centrifuge modeling of seis- CANLEX liquefaction embankments: prototype and centrifuge
mic shaking. In Proceedings of the 3rd International Conference models. Canadian Geotechnical Journal, 34(5): 641–654.
of Recent Advances in Geotechnical Earthquake Engineering and Riemer, M.F., Seed, R.B., Nicholson, P.G., and Jong, H.L. 1990.
Soil Dynamics, St. Louis, Mo., 2–7 April 1995. University of Steady state testing of loose sands: limiting minimum density.
Missouri-Rolla, Rolla. Vol. 2, pp. 927–942. Journal of Geotechnical Engineering, ASCE, 116: 332–337.
Lade, P.V., and Yamamuro, J.A. 1997. Effects of nonplastic fines Roscoe, K.H. 1970. The influence of strains in soil mechanics.
on static liquefaction of sands. Canadian Geotechnical Journal, Géotechnique, 20: 129–170.
34(6): 918–928. Salgado, R., Bandini, P., and Karim, A. 2000. Shear strength and
Lade, P.V., Liggio, C.D., and Yamamuro, J.A. 1998. Effects of stiffness of silty sand. Journal of Geotechnical and Geo-
non-plastic fines on minimum and maximum void ratios of sand. environmental Engineering, ASCE, 126: 451–462.
ASTM Journal of Geotechnical Testing, 21: 336–347. Sawada, S., Tsukamoto, Y., and Ishihara, K. 2006. Residual defor-
Lemke, R.W. 1967. Effects of the earthquake of March 27, 1964 at mation characteristics of partially saturated sandy soils sub-
Seward, Alaska. US Geological Survey, Professional Paper 542-E. jected to seismic excitation. Journal of Soil Dynamics and
Liu, H., and Qiao, T. 1984. Liquefaction potential of saturated sand Earthquake Engineering, 26: 175–182.
deposits underlaying foundation of structure. In Proceedings of Schofield, A.N. 1981. Dynamics and earthquake geotechnical
the 8th World Conference on Earthquake Engineering, San Fran- centrifuge modeling. In Proceedings of the International Confer-
cisco, Calif., 21–28 July 1984. Prentice–Hall Inc., Englewood ence on Recent Advances in Geotechnical Earthquake Engi-
Cliffs, N.J. pp. 199–206. neering and Soil Dynamics, St. Louis, Mo., 26 April – 2 May
Lu, X., Shuyun, W., Yihua, W., and Cui, P. 2004. An approximate 1981. Edited by S. Prakash. University of Missouri-Rolla, Rolla.
method for evaluating the shear band thickness in saturated Vol. 3, pp. 1081–1100.
sand. International Journal of Numerical and Analytical Methods Scott, R.F., and Zuckerman, K.A. 1972. Sandblows and liquefac-
in Geomechanics, 28: 1533–1541. tion. In The Great Alaskan earthquake of 1964. Engineering
Malvick, E.J. 2005. Void redistribution-induced shear localization Publication 1606, National Academy of Sciences, Washington,
and deformation in slopes. Ph.D. dissertation, University of Cal- D.C., pp. 170–189.
ifornia at Davis, Davis, Calif. 285 pp. Seed, H.B. 1987. Design problems in soil liquefaction. Journal of
Malvick, E.J., Kulasingam, R., Kutter, B.L., and Boulanger, Geotechnical Engineering, ASCE, 113: 827–845.
R.W. 2002. Void redistribution and localized shear strains in Seed, R.B., and Harder, L.F. 1990. SPT-based analysis of cyclic
slopes during liquefaction. In Physical Modeling in pore pressure generation and undrained residual strength. In
Geotechnics, ICPMG‘02: Proceedings of the International Proceedings of the H.B. Seed Memorial Symposium, May 1990.
Conference, St. John’s, Nfld., 10–12 July 2002. Edited by R. Edited by J.M. Duncan. BiTech Publishing Ltd., Richmond,
Phillips, P.J. Guo, and R. Popescu. A.A. Balkema, Rotterdam. B.C. Vol. 2, pp. 351–376.
pp. 495–500. Seid-Karbasi, M., and Byrne, P.M. 2004a. Liquefaction, lateral
Malvick, E.J., Kutter, B.L., Boulanger, R.W., Kabasawa, K., and spreading and flow slides. In Proceedings of the 57th Canadian
Kokusho, T. 2005. Void redistribution research with 1-g and Geotechnical Engineering Conference, Québec City, Que., 24–
centrifuge modeling. In Proceedings of the 16th International 27 October 2004. Canadian Geotechnical Society, Alliston, Ont.
Conference on Soil Mechanics and Geotechnical Engineering p. G13.529.
(ICSMGE), Osaka, 12–16 September 2005. A.A. Balkema, Rot- Seid-Karbasi, M., and Byrne, P.M. 2004b. Embankment dams and
terdam. Vol. 4, pp. 2543–2546. earthquakes. International Journal on Hydropower and Dams,
Malvick, E.J., Kutter, B.L., Boulanger, R.W., and Kulasingam, R. 11(2): 96–102.
2006. Shear localization due to liquefaction-induced void redis- Seid-Karbasi, M., and Byrne, M.P. 2006. Effects of partial satura-
tribution in a layered infinite slope. Journal of Geotechnical and tion on liquefiable ground response. In GeoCongress 2006:
Geoenvironmental Engineering, ASCE, 132: 1293–1303. Geotechnical Engineering in the Information Technology Age,
Martin, G.R., Finn, W.D.L., and Seed, H.B. 1975. Fundamentals of Atlanta, Ga., 26 February – 2 March 2006. Edited by D.J.
liquefaction under cyclic loading. Journal of the Geotechnical DeGroot, J.T. DeJong, D. Frost, and L.G. Baise. ASCE, Reston,
Engineering Division, ASCE, 101: 423–438. Va. Paper 11803.
Olson, S.M., and Stark, T.D. 2002. Liquefied strength ratio from Seid-Karbasi, M., Byrne, P.M., Naesgaard, E., Park, S.,
liquefaction flow failure case histories. Canadian Geotechnical Wijewickreme, D., and Phillips, R. 2005. Response of sloping
Journal, 39(3): 629–647. ground with liquefiable materials during earthquake: a class A
Olson, S.M., and Stark, T.D. 2003. Use of laboratory data to confirm prediction. In International Association of Computer Methods
yield and liquefied strength ratio concepts. Canadian and Advances in Geomechanics, IACMAG: Proceedings of the
Geotechnical Journal, 40(6): 1164–1184. 11th International Conference, Turin, Italy, 19–21 June 2005.
Pestana, J.M., and Whittle, A.J. 1995. Compression model for Edited by G. Barla and M. Barla. Patron Editore, Bologna, Italy.
cohesionless soils. Géotechnique, 45: 611–631. Vol. 3, pp. 313–320.

© 2007 NRC Canada


890 Can. Geotech. J. Vol. 44, 2007

Sento, N., Kazama, M., Uzuoka, R., Ohmur, H., and Ishimaru, M. Tsukamoto, Y., Ishihara, K., and Sawada, S. 2004. Settlement of
2004. Possibility of post liquefaction flow failure due to seepage. silty sand deposits following liquefaction during earthquakes.
Journal of Geotechnical and Geoenvironmental Engineering, Soils and Foundations, 44: 135–148.
ASCE, 130: 707–716. Uzuoka, R., Sento, N., and Kazama, M. 2003. Numerical analysis
Sharp, M., Dobry, R., and Abdoun, T. 2003. Liquefaction centri- on liquefaction-induced progressive deformation with pore water
fuge modeling of sands of different permeability. Journal of pressure migration. In Proceedings of the 3rd International Sym-
Geotechnical and Geoenvironmental Engineering, ASCE, 129: posium on Deformation Characteristics of Geomechanics, Lyon,
1083–1091. 22–24 September 2003. Edited by H. Di Benedetto, T. Doanh,
Silver, M.L., and Seed, H.B. 1971. Volume change in sands during H. Geoffroy, and C. Sauzét. A.A. Balkema, Rotterdam.
cyclic loading. Journal of the Soil Mechanics and Foundations pp. 1095–1101.
Division, ASCE, 97: 1171–1182. Vaid, Y.P., and Eliadorani, A. 1998. Instability and liquefaction of
Sladen, J.A., D’Hollander, R.D., Krahn, J., and Mitchell, D.E. granular soils under undrained and partially drained states.
1985. Back analysis of the Nerlerk berm liquefaction slides. Canadian Geotechnical Journal, 35(6): 1053–1062.
Canadian Geotechnical Journal, 22: 579–588. Vaid, Y.P., Chern, J.C., and Tumi, H. 1985. Confining pressure,
Sriskandakumar, S. 2004. Cyclic loading response of Fraser River grain angularity, and liquefaction. Journal of Geotechnical Engi-
Sand for numerical models simulating centrifuge tests. M.A.Sc. neering, ASCE, 111: 1229–1235.
thesis, Civil Engineering Department, University of British Yang, Z., and Elgamal, A. 2002. Influence of permeability on
Columbia, Vancouver, B.C. liquefaction-induced shear deformation. Journal of Engineering
Stark, T.D., and Mesri, G. 1992. Undrained shear strength of lique- Mechanics, ASCE, 128: 720–729.
fied sands for stability analysis. Journal of Geotechnical Engi- Yoshida, N., Yasuda, S., and Ohya, Y. 2005. Two criteria for
neering, ASCE, 118: 1727–1747. liquefaction-induced flow. In Proceedings of the 16th Interna-
Taboada, V., and Dobry, R. 1993a. Experimental results of model tional Conference on Soil Mechanics and Geotechnical Engi-
No. 1 at RPI. In Verification of Numerical Procedures for the neering (ICSMGE), Performance Based Design in Earthquake
Analysis of Soil Liquefaction Problems (VELACS): Proceedings Geotechnical Engineering: Concepts and Research Satellite
of the International Conference, Davis, Calif., 17–20 October Conference, Osaka, Japan, 10–15 September 2005. A.A. Balkema,
1993. Edited by K. Arulandan, R.F. Scott, and X. Zeng. A.A. Rotterdam. pp. 109–116.
Balkema, Rotterdam. Vol. 1, pp. 3–17. Yoshimine, M., Nishizaki, H., Amano, K., and Hosono, Y. 2006.
Taboada, V., and Dobry, R. 1993b. Experimental results of model Flow deformation of liquefied sand under constant shear load
No. 2 at RPI. In Verification of Numerical Procedures for the and its application to analysis of flow slide of infinite slope.
Analysis of Soil Liquefaction Problems (VELACS): Proceedings Soil Dynamics and Earthquake Engineering Journal, 26: 253–
of the International Conference, Davis, Calif., 17–20 October 264.
1993. Edited by K. Arulandan, R.F. Scott, and X. Zeng. A.A. Youd, T.L., Idriss, I.M., Andrus, R., Arango, I., Castro, G., Christian,
Balkema, Rotterdam. pp. 277–294. J., Dobry, J., Finn, L., Harder, L., Jr., Hynes, H.M., Ishihara, K.,
Taboada, V., and Dobry, R. 1998. Centrifuge modeling of earth- Koester, J., Liao, S.S., Marcuson, W.F., III, Martin, G., Mitchell,
quake-induced lateral spreading in sand. Journal of Geotechnical J.K., Moriwaki, Y., Power, M.S., Robertson, P.K., Seed, R.B.,
Engineering, ASCE, 124: 195–206. and Stokoe, K.H., II. 2001. Liquefaction resistance of soils:
Takeshita, S., Takeishi, M., and Tamada, K. 1995. Static liquefaction summary report from the 1996 NCEER and 1998 NCEER/NSF
of sands and its liquefaction index. In Proceedings of the 1st Workshops on Evaluation of Liquefaction Resistance of Soils.
International Conference on Earthquake Geotechnical Engi- Journal of Geotechnical and Geoenvironmental Engineering,
neering, Tokyo, Japan, 14–16 November 1995. Edited by K. ASCE, 127: 817–833.
Ishihara. A.A. Balkema, Rotterdam. Vol. 1, pp. 177–182. Zlatovic, S. 1994. Residual strength of silty soils. D.Eng. thesis,
Thevanayagam, S. 1998. Effect of fines and confining stress on University of Tokyo, Tokyo.
undrained shear strength of silty sands. Journal of Geotechnical
and Geoenvironmental Engineering, ASCE, 124: 479–491.

© 2007 NRC Canada

Vous aimerez peut-être aussi