Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

FLOW LIQUEFACTION SIMULATION USING A

COMBINED EFFECTIVE STRESS – TOTAL STRESS


MODEL
Ernest Naesgaard and Peter M. Byrne
Department of Civil Engineering – University of British Columbia, Vancouver,
B.C., Canada

ABSTRACT
Pore water redistribution and soil mixing can give post-liquefaction strengths much lower than those from undrained
laboratory tests. Fully coupled effective stress numerical analyses have given important insight into the process but
further development and calibration is needed. In the meantime a combined effective stress – total stress analysis is
recommended. In the combined procedure the effective stress analysis is used to determine zones of liquefaction and
deformations that occur during strong shaking. Following strong shaking liquefied elements are changed to a total
stress model with a specified residual strength that is back-calculated from case histories.

RÉSUMÉ
La redistribution des pressions d’eau interstitielles et le mélange de sols granuleux donnera résistances après-
liquéfaction beaucoup plus basses que celles des essais de laboratoire non drainés. Des analyses numériques
entièrement couplées d’effort effectif ont donné un éclairage important dans le procès mais autres développements et
calibrations sont nécessaires. En attendant, une analyse combinée d’effort effectif – effort totale est recommandée.
Dans cette procédure combinée l’analyse d’effort effectif est utilisée pour déterminer des zones de liquéfaction et
déformations qui occurrent pendant la secousse. Après la secousse les éléments liquéfiés sont changés en modèle
total d’effort avec une résistance résiduelle spécifiée qui est retro-calculée à partir des antécédents.

1 INTRODUCTION generally does not occur, even for loose sands on


relatively steep slopes subjected to strong shaking.
Numerous flow and bearing failures have occurred during
or following strong earthquake shaking when liquefaction Fully coupled effective stress numerical analyses
is triggered (Ishihara 1984, Kokusho 2003, Hamada procedures developed at the University of British
1992, Seed et al. 1987, Poulos et al. 1985). These flow Columbia have been used to emulate field and centrifuge
liquefaction failures are deemed to occur when the static test case histories and have given important insight into
driving stress exceeds the soil shear strength. Shear liquefied strength. The difficulties with the fully coupled
strength of soil following triggering of liquefaction has effective stress analyses are that (1) the results are
been called residual or liquefied strength, and has dependent on stratigraphy details that often are not well
commonly been assumed to be an undrained strength known and often on a scale smaller than the size of the
parameter. Seed (1987) back-calculated these strengths smallest practical element in the numerical model, (2)
from case histories while others (Poulos et al.1985) localization occurs below the low permeability barriers
attempted to determine them from laboratory tests. and the behaviour becomes element size dependent
Recent work (Vaid & Eliadorani 1998, Yoshida & Finn (Naesgaard et al. 2005, Yang & Elgamal 2002), (3) mixing
2000, Kokusho 1999, Kokusho 2003, Kulasingam et al. of layered soils often occurs during liquefaction and
2004, Sento et al. 2004, Seid-Karbasi & Byrne 2004, quantification of the effects is difficult and (4) the effects
Naesgaard et al. 2005; Malvick 2005, and Naesgaard et of out-of-plane shaking are not considered in the current
al. 2006) has shown that the undrained assumption is two-dimensional analyses. Perhaps the coupled effective
often not correct and may be unconservative. They have stress analyses alone will be able to fulfill design
demonstrated that pore water flow and pressure requirements with further development and calibration;
redistribution which occurs during and following however, it is currently recommended that an alternative
earthquake shaking may result in relatively thin zones total-stress analysis should also be carried out in
with very high void ratio, or in the extreme, water inter- conjunction with the coupled effective stress analysis to
layers, below low permeability layers. These high void overcome the shortcomings discussed above. In the
ratio/water inter-layer zones have very low to near zero proposed effective-stress/total-stress approach the
shear strength, much lower than that obtained from residual strength that is back-calculated from case
undrained laboratory element tests. Without a low histories (similar to those originally proposed by H. Seed
permeability barrier to retard the escape of groundwater, and colleagues (Seed 1987)) is introduced into the
and/or some form of soil mixing (Naesgaard & Byrne process. In this manner the empirical experience used in
2005, Vasquez-Herrera & Dobry 1989), flow liquefaction current design practice is included in the process.
2 SOIL LIQUEFACTION AND PORE WATER between time of earthquake shaking and time of flow
REDISTRIBUTION failure. There are numerous case histories where soil
liquefaction occurred during earthquake shaking but
When typical loose sand, relative density Dr = 40%, is related flow failure did not occur until some time after end
tested in drained (dry) cyclic simple shear it is initially of shaking. The classical examples are the Lower San
contractive on loading and unloading. However on Fernando Dam where the upstream face of the dam failed
loading to large strains the stress state exceeds the 20 to 30s after end of earthquake shaking (Seed, 1973,
constant volume friction angle (Φcv) and the soil becomes Seed, 1987) and in Niigata where eyewitnesses reported
dilative. Dense sand behaves in a similar manner except that the girders of the Showa Ohashi Bridge fell a few
the dilative response is more pronounced. Both dense minutes after the earthquake motion had ceased
and loose sands are contractive on unloading. The net (Hamada, 1992).
result of cyclic loading is generally a reduction in sample
volume. Numerical analyses have shown that the blockage effect
on flow caused by the presence of the barrier causes the
If the pores are filled with water that is prevented from highest rate of expansion to occur directly beneath the
escaping (undrained condition) then pore pressures will barrier. This is important as this can cause localization
increase when the soil skeleton attempts to contract, and and formation of a thin very weak layer at the interface.
decrease when the soil skeleton attempts to expand The expansion is believed to be mainly due to the influx of
(dilate). With repeated cycles the stress path may reach water rather than from shear induced dilation. If the zone
the zero effective stress / zero shear strength origin and directly under the barrier is thought of as a very thin soil
true liquefaction occurs. However, with continued element, with plane boundaries, then an extremely small
monotonic shearing to large strains the soil will attempt to influx of pore water will cause a large volumetric
dilate and gain strength. The residual strength will not be expansive strain under the barrier and may result in the
reached until, (i) the pore-water cavitates and thus allows element going to the critical state at zero effective stress.
the sample to increase in volume and reach the steady Further inflow will cause localization and formation of a
state, or (ii) the high mean effective stress generated by water film with zero effective stress and zero shear
dilation suppresses the dilation and the soil reaches its strength. Further shearing does not induce dilation. The
critical state strength, or (iii) the sand grains crush and strength will remain zero until the water film drains.
the soil reaches a critical state of the crushed material.
The strength of the sand reached in (i), (ii) or (iii) is With a static shear bias, the expanding zone under the
generally much higher than the commonly accepted barrier will attempt to fail prior to reaching the critical state
‘undrained” residual strengths back-calculated from case and zero strength. This will result in a large shear strain
histories and is likely higher than the drained strength. that is a function of inflow volume and soil density
(Boulanger and Truman 1996, Sento et al. 2004).
If, in lieu of undrained loading, a small inflow of water However, with continued inflow the layer eventually
(expansion) happens then shear can occur without reaches the critical state at a shear stress corresponding
strength gain. Upon reaching the critical state further to the static bias and any further inflow will lead to a flow
inflow will cause total loss of strength (Vaid & Eliadorani slide.
1998, Boulanger & Truman 1996, Sento et al. 2004).
This process of void redistribution was originally proposed In real soils the barrier boundary will not be perfectly
by Whitman (1985). plane, infinitely thin or of infinite lateral extent but will
have undulations, varying normal total stresses, finite
Natural and many man-made soils are often layered and grain sizes, varying permeability, etc. (Naesgaard et al.
have variations in grain-size and permeability. 2005). Earthquake duration and slope geometry may also
Earthquake shaking and related liquefaction will induce affect the residual strength. These items will result in the
pore water gradients and flow that will result in liquefied (residual) shear strength along the interface
contraction of the soil skeleton with outflow of pore water varying both in time and space with an average value that
in some areas and expansion with inflow of pore water in is greater than zero. This ‘average value’ is the residual
other areas. Inflow gives soil strengths considerably lower shear strength that is believed to be reflected in the low
than that obtained assuming undrained conditions. When values back-calculated from case histories.
there is a low permeability layer or barrier, the upward
migrating pore water gets trapped beneath the barrier and
forms a layer of local expansion with low effective stress. 3 SOIL MIXING
At the limit, an actual water interlayer will develop. In
embankments and water-edge slopes, liquefaction within Another procedure that may result in high void ratios,
the center portion of embankments or back from a water- excess pore water and liquefaction is soil mixing
edge slope will induce pore pressures equal to the local (Naesgaard & Byrne 2005, Vasquez-Herrera & Dobry
overburden pressure. When this high pressure water 1989). Natural water-pluviated soils are often layered.
migrates laterally toward the edges of the embankment or When layers of coarse and fine grained soils are mixed
toward the water-edge slope (location of lower together in an undrained or near-undrained environment
overburden pressure), expansion and potential for than the resulting post-liquefaction shear strength of the
strength reduction will also occur. Since water flow and mixed soil is much lower than the shear strength of the
pressure redistribution takes time there is often a delay individual layers. This is illustrated in Fig. 1 and 2.
During the soil liquefaction and post-liquefaction shear Back-calculating post-liquefaction ‘residual’ strength from
process there are numerous occasions during which case histories and treating it as an “undrained” soil
extensive turbulence and mixing of particles occurs. This strength property that is a function of penetration
has the potential for creating zones with post-liquefaction resistance or soil density is a gross simplification with a
strengths well below those that would be obtained from large scatter in available data (Fig. 3). However at this
undrained laboratory tests on the individual non-mixed time this may still be the best strength data available.
layers.

Figure 1. ‘A’ is prior to mixing, ‘B’ & ‘C’ are two


alternative mixed configurations. ‘A’ has
significantly higher residual strength than ‘B’ & ‘C’.
From Naesgaard & Byrne (2005).

4 RESIDUAL STRENGTHS FROM BACK-ANALYSIS


OF FIELD CASE HISTORIES Figure 3. Residual strength ratio (Sr/σ'vo) of liquefied
soil versus equivalent-clean-sand- corrected SPT (N1)60,
Field experience during past earthquakes indicates that
for σ'vo  400 kPa. From Idriss & Boulanger (2007).
residual strengths can be much lower than values
obtained from undrained laboratory tests on undisturbed
samples. This is postulated to be due to the presence of 5 PROPOSED COMBINED EFFECTIVE STRESS –
low permeability barriers and soil mixing. Based on back- TOTAL STRESS APPROACH
analysis of field case histories, Seed and Harder (1990),
Olson and Stark (2002), Idriss and Boulanger (2007) and Fully coupled effective stress numerical analyses can
others have proposed residual shear strength (Sr) or model the triggering of liquefaction, the cyclic mobility
strength ratio (Sr/p') for liquefied soil as a function of deformations that occur during shaking, and emulate the
either SPT (N1)60 or normalized CPT tip resistance qc1. pore pressure migration and trapping of water under low
The recent procedure by Idriss and Boulanger (2007) permeability barriers as observed in field and centrifuge
includes different relationships for cases with and without test case histories (Byrne et al. 2006). One of the
low permeability barriers (Fig.3) and has been used in the difficulties with the fully coupled effective stress numerical
example analysis in this paper. modelling is that localization occurs below the low
permeability barriers and the behaviour becomes element
size dependent (Yang & Elgamal 2002, Seid-Karbasi &
Byrne 2004, Naesgaard et al. 2005). Using the very
small elements necessary to correctly simulate the
behaviour is not practical due to time and memory
constraints. Large elements can be made to behave like
small elements by curtailing dilation in the large element
at a void ratio that is considerably less than that
calculated from critical state theory (Naesgaard et al.
2005). However, it is difficult to know if the coupled
effective stress model is correctly capturing the behaviour
and further work in calibrating the model is required. As a
tool for use in current design practice it is proposed to
augment the coupled effective stress analyses with a total
stress analysis which uses liquefied (residual) strengths
back-calculated from case-histories.

The proposed numerical process is as follows:


Figure 2. Post-liquefaction strengths of individual
sandy SILT and silty SAND layers are much higher (1) The model is built and brought to static equilibrium.
than the post-liquefacton strengths of a mixture of the (2) A dynamic coupled effective stress analysis is
two soils. From Naesgaard & Byrne (2005). carried out to the end of strong earthquake shaking.
During the analyses zones where liquefaction has plastic effective stress model with the mechanical
been triggered (pore pressure ratio Ru (u/σ'vo) > 0.70 behaviour of the sand skeleton and pore water flow fully
to 1.0) are tracked. Following end of strong coupled ((Beaty & Byrne 1998; Byrne et al. 2004). The
earthquake shaking a data file is saved and model includes a yield surface related to the developed
analyses are continued as two separate cases (3a & friction angle, non-associative flow rule, and definitions for
3b). loading, unloading, and hardening. Key elastic and
(3a) In one of the analysis branches the coupled plastic parameters used are adjusted so as to give a
effective stress analysis is continued to allow further good match with simple shear laboratory tests as the
pore pressure redistribution to occur and to bring loading path of this test, including rotation of principal
insight into the related localization and potential stress axes, closely approximates that which occurs
post-liquefaction flow. during earthquake loading. The UBCSAND constitutive
(3b) In the other analysis branch, elements within the model is run within the finite difference program FLAC 5.0
post-shaking model that have liquefied during the (ITASCA 2005). In the FLAC program dynamic analyses
first phase are changed to a total stress constitutive are carried out in the time domain with full coupling
model with a ‘residual’ strength and the analysis is between groundwater flow and mechanical loading.
continued until static equilibrium is reached. The
proposed post-liquefaction residual strength is that The UBCSAND model has been calibrated against simple
back-calculated from case histories. shear laboratory tests, centrifuge tests with and without
(4) The results from both the coupled effective-stress impermeable silt barriers (Yang et al. 2004, Phillips et al.
analysis and the combined effective stress/total- 2005, Seid-Karbasi et al. 2005, Byrne & Park 2005) and
stress analysis should be considered for design. the empirical liquefaction triggering charts (Idriss &
Boulanger 2007). The model is able to emulate both the
A flow chart illustrating the procedure is given in Fig. 4. observed drained behaviour of loose sand soils
(Contraction when sheared below the Φcv and dilative
above the Φcv) and the build-up of pore pressure and soil
5.1 Coupled Effective Stress Analysis liquefaction that occurs in undrained simple shear tests.
The model is also able to emulate the behaviour of a flow
A coupled effective stress constitutive model (UBCSAND) failure when a low permeability barrier was present and
and analyses procedure for modelling earthquake no failure when the barrier was absent or when drains
shaking and liquefaction has been developed at the were installed through the barrier (Byrne et al. 2006,
University of British Columbia. UBCSAND is an elasto- Naesgaard et al. 2005) and able to emulate the post-
shaking failure of the Lower San Fernando Dam
(Naesgaard et al. 2006). Atigh & Byrne, 2004, showed
that UBCSAND could emulate the behaviour of the triaxial
tests with fluid inflow carried out by Vaid & Eliadorani,
1998.

As mentioned, pore pressure redistribution in soil profiles


with low permeability barriers can cause localization
below the barrier and the analyses results become
element size dependent. Having very small elements in
the model is not practical due to analysis time constraints.
To counter this effect, large elements are made to behave
like smaller elements by setting dilation to zero or near
zero in elements that are expanding due to water inflow.
This change is made at the end-of-strong shaking so the
calibrated shear-induced liquefaction triggering behaviour
is not affected.

5.2 Post-liquefaction Total Stress Analysis

The pore pressure ratio Ru is monitored during the


coupled effective stress analysis and the maximum value
in each element is retained. At the end of strong shaking,
elements with maximum Ru greater than a value around
0.7 to 1.0 are deemed to have liquefied and are changed
to a Mohr Coulomb constitutive model with a friction and
dilation angle of zero, and cohesion set equal to the
residual strength Sr. The Sr is that back-calculated from
case histories as proposed by Seed & Harder (1990),
Figure 4. Flow chart showing combined coupled Olson & Stark 2002, and Idriss & Boulanger 2007. The
effective stress / total stress analyses procedure. shear modulus, bulk modulus, density, porosity and
permeability in the liquefied elements are set equal to the
values present at the end of strong shaking (end of phase In the actual dam the spacing of the low permeability
2) in the initial coupled effective stress analysis. layers is much finer than that in the numerical model.
Limitations on smallest element size are necessary for
analysis time to be in a practical range and therefore the
6 LOWER SAN FERNANDO DAM - EXAMPLE behaviour of finely spaced layers must be emulated with
ANALYSIS larger elements by introducing a dilation cut-off in any
elements expanding beyond their original volume.
In 1971 the Lower San Fernando Dam (LSFD) was
shaken by a large earthquake with a peak velocity pulse 6.1 Coupled Effective Stress Analysis of LSFD –
of around 0.6m/s and peak ground acceleration of Phases 2 & 3a
approximately 0.5 g (Seed 1973; Seed et al. 1989, Castro
et al. 1989, Castro 1995). Approximately 20 to 30 s after In the example LSFD analysis the coupled effective
the end of the earthquake shaking the upstream face of stress analysis was carried out using the UBCSAND
the dam catastrophically failed leaving only 1.5m of free- constitutive model within the program FLAC as described
board and putting a large population at risk. Extensive in section 5.1 for all saturated sand elements while a total
investigation and analyses of the dam were conducted stress Mohr Coulomb model was used for the clay/silt
following the event. From the studies it was concluded core portion of the dam.
that the hydraulic fill soil within lower and central portions
of the dam had liquefied and overlying portions of the Initially the whole model was brought to static equilibrium
dam had flowed out riding on the liquefied soil (Fig. 5). using a Mohr Coulomb model and drained stiffness
parameters (phase 1). Then prior to earthquake shaking
In the current example the liquefiable portions of hydraulic the sand elements were changed to the UBCSAND model
fill soils were given (N1)60 values of 12 while the clayey while portions of the clay/silt core were left with a Mohr
core of the dam was given undrained shear strength of Coulomb model but given undrained stiffness parameters.
18% of the initial vertical overburden pressure (Fig.6). The model is then brought to equilibrium in dynamic
The UBCSAND constitutive model was used for the mode and all displacements reset to zero (end of phase
potentially liquefiable sandy soil portions of the dam while 1). Following this the dynamic analysis with earthquake
a Mohr Coulomb model was used for the clayey core and motion input at the base was run until excessive grid
portions of the dam above the water table. The effective distortion stopped the analysis (≈ 40 s). This constitutes
stress analyses were fully-coupled (mechanical - pore phase 2 and 3a in Fig. 4. A file was saved at after end of
water flow) and run within the program FLAC. strong shaking (15 s) for use in the phase 3b Total Stress
alternate analysis.

Figure 5. Lower San Fernando Dam which failed into the reservoir approximately 30s after end of earthquake
shaking. Black areas are soil that was mixed and deemed to have liquefied. Lower figure is reconstructed profile
which shows location of liquefied soil before failure. (Seed et al. 1973).
Figure 6. Grid, locations of low permeability barriers, (N1)60) and shear strength input parameters used in Low San
Fernando numerical example.
recommended by Idriss and Boulanger (2007) (Fig. 3).
The dynamic analysis was then continued until stopped
6.2 Post-Liquefaction Total Stress Analysis of LSFD – by excessive grid distortion (≈ 18 s).
Phase 3b
6.3 Results of LSFD example analysis
Elements which had liquefied (Ru)max > 0.7) during the
strong shaking portion of the analysis (Phase 2) where During strong shaking (phase 2) the upstream shell
changed to a Mohr Coulomb model with zero friction moved approximately 3 m upstream and the downstream
angle, zero dilation, and a cohesion set equal to a shell moved approximately 1.5 m downstream.
residual strength (Sr) calculated using the initial pre- Liquefaction occurred both within the upstream and
shaking effective stress (p') and the Sr/p' ratio downstream portions of the dam as shown in Fig. 7.
Strong shaking ended at approximately 10 s (Fig. 8).

Continuing the coupled effective stress analysis (Phase


3a) resulted in very little movement within the dam until
approximately 32 s at which time the upstream shell
proceeded to fail in a similar manner to that reported for
the actual dam (Fig. 8). Fig. 9 shows displacement and
Figure 7. Profile showing zones deemed to have velocity of the dam at 40 s at which time the program
liquefied at end of strong shaking (Ru > 0.7). stopped due to excessive distortion of some of the
elements. The downstream face did not move
given important insight into flow liquefaction behaviour.
However, there is uncertainty in the procedures and a
combined effective stress – total stress analysis carried
out in addition to the fully coupled analysis is
recommended. In the combined procedure the effective
stress analysis is used to determine zones of liquefaction
and deformations that occur during strong shaking.
Following strong shaking liquefied elements are changed
to a total stress model with a specified residual strength.
The proposed residual strength is that back-calculated
from case histories similar to those originally proposed by
H. Seed and colleagues. In this manner the empirical
experience used in current design practice is included in
the analytical process.

An example back-analysis of the Lower San Fernando


Dam demonstrates the proposed analysis procedure.
Both the coupled effective stress analysis and the total
stress alternative indicate a flow failure of the upstream
shell into the reservoir but no flow failure of the
downstream slope as was the case in the 1971 failure.
Figure 8. Displacement time histories of point near The coupled effective stress analysis also indicated that
upstream toe of dam for both 3a coupled effective the failure would occur approximately 25 s after the end
stress analysis (ESA) and 3b total stress analysis of strong shaking - similar to what was actually observed.
(TSA) alternative. Both analyses stopped due to
excessive element distortion.
ACKNOWLEDGEMENTS
significantly following end of strong shaking as was the
case of the actual dam.
The authors acknowledge the support from the British
Columbia Ministry of Transportation - UBC Professional
In the Phase 3b total stress alternative the dam
Partnership program, the Canadian Council of
proceeded to fail once the liquefied elements where
Professional Engineers, Trow Associates Inc., and the
changed to a Mohr Coulomb model with residual strength
National Scientific and Engineering Research Council
Sr as given by the Idriss & Boulanger (2007) relationship.
Strategic Liquefaction Grant No. NSERC 246394. The
This is not unexpected as the Lower San Fernando dam
writers would like to acknowledge the contribution of
was one of the key case histories used in deriving the
Mahmood Seid Karbasi, Sung Sik Park, Michael Beaty, in
relationship. Again the dam only failed in the upstream
development of the UBCSAND model and analysis
direction as was the case with the actual dam.
procedures and Pascale Rouse for abstract translation.

7 CONCLUSIONS REFERENCES

Fully coupled effective stress numerical analyses have Atigh, E. and Byrne, P.M. 2004. Liquefaction flow of
emulated field and centrifuge test case histories and submarine slopes under partially undrained
conditions: an effective stress approach, Canadian
Geotech. J., V. 41, pp. 154-165.
Beaty, M.H. 2001. A Synthesized Approach for Estimating
Liquefaction-induced Displacements of Geotechnical
Structures, Ph.D. Thesis, University of British
Columbia, Dept. of Civil Eng.
Beaty, M., and Byrne, P.M. 1998. An effective stress
model for predicting liquefaction behaviour of sand,
Geotechnical Earthquake Engineering and Soil
Dynamics III. P. Dakoulas, M. Yegian, and R Holtz
(eds.), ASCE, Geotechnical Special Publication 75
(1), pp. 766-777.
Boulanger, R. W., and Truman, S. P. 1996. Void
redistribution in sand under post-earthquake loading,
Can. Geotech. J., 33, 829-834.
Byrne, P.M., Park, S.S., Beaty, M., Sharp, M., Gonzalex,
Figure 9. (a) Horizontal velocity (m/s) and (b) L., Abdoun, T. 2004. Numerical modeling of dynamic
horizontal displacement (m) at end of analyses (40s) centrifuge tests, 13th World Conf. on Earthquake
for coupled effective stress analysis. Eng., Vancouver, B.C., paper 3387.
Byrne, P.M., Naesgaard, E., and Seid-Karbasi, M. 2006. Phillips, R.,, Tu, M., and Coulter, S., 2004. C-CORE.
Analysis and design of earth structures to resist Earthquake Induced Damage Mitigation from Soil
th
seismic soil liquefaction, 59 Canadian Geotechnical Liquefaction. Data report – Centrifuge Test CT5”. For
Conf., Vancouver. University of British Columbia. C-CORE Report R-04-
Castro, G., Keller, T.O., and Boynton, S.S. 1989. Re- 068-145, December.
evaluation of the Lower San Fernando Dam, Rpt. 1 Poulos, S.J., Castro, G., and France, W., 1985.
Vol. 1 & 2, US Army Engineer Waterways Experiment Liquefaction evaluation procedure, J. Geotechnical
Station, Contract Rpt. GL-89-2. Engineering Div., ASCE, Vol. 111, No.6, pp. 772-792.
Castro, G. 1995. Empirical Methods in Liquefaction Seed, H.B., Lee, K.L., Idriss, I.M., and Makdisi, F., 1973.
Evaluation, Primer Ciclo de Conferencias Analysis of the Slides in the San Fernando Dams
Internationales Leonardo Zeevaert, Mexico City. during the Earthquake of Feb. 9, 1971, Earthquake
Hamada, M. 1992. Large ground deformations and their Engineering Research Center, Rpt. EERC 73-2, June.
effects on lifelines: 1964 Niigata Earthquake, Proc., Seed, H.B., 1987. Design problems in soil liquefaction, J.
Lifeline Performance during Past Earthquakes, V. 1: Geotechnical Eng. Div., ASCE, Vol. 113, No.8, 827-
Japanese Case Studies Tech. Rep NCEER-92-0001, 845.
Buffalo, N.Y., 3/1-3/123. Seed, H.B., Seed, R.B., Harder, L.F., and Jong, H.L.,
Idriss, I.M., and Boulanger, R.W. 2006. Semi-empirical 1989. Re-evaluation of the Lower San Fernando
procedures for evaluating liquefaction potential during Dam, Rpt. 2, US Army Engineer Waterways
earthquakes. Soil Dynamics and Earthquake Experiment Station, Contract Rpt. GL-89-2.
Engineering 26, 115-130. Seed, R.B., and Harder, L.F., Jr., 1990. SPT-based
Idriss, I.M., and Boulanger, R.W. 2007. SPT- and CPT- analysis of cyclic pore pressure generation and
based relationships for the residual shear strength of undrained residual strength., In Proc. of the H.B. Seed
th
liquefied soils, Proc. 4 Int. Conf. on Earthquake Memorial Symposium, Bi-Tech Publishing Ltd., Vol. 2,
Geotechnical Engineering, Thessaloniki, Greece, pp. 351–376.
June. Seid-Karbasi, M., and Byrne, P.M., 2004. Liquefaction,
Ishihara, K. 1984. Post-earthquake failure of a tailings lateral spreading and flow slides, Proc. 57th Canadian
dam due to liquefaction of the pond deposit, Proc. Geotechnical Conf., Session 2c, pp. 23-30.
Inter. Conf., Case Histories in Geotechnical Eng., Seid-Karbasi, M. Byrne, P.M., Naesgaard, E., Park, S.S.,
Rolla, Missouri, V. 3, pp. 1129-1143. Wijewickreme, D., and Phillips, R., 2005. Response of
ITASCA, 2005. FLAC - Fast langrangian analysis of Sloping Ground with Liquefiable Materials during an
continua, Version 5.0 ITASCA Consulting Group Inc., Earthquake: a class A prediction, 11th IACMAC Conf.,
Minneapolis, Minnesota. Italy.
Kokusho, T. 2003. Current state of research on flow Sento, N., Kazama, M., Uzuoka, R., Ohmura, and
failure considering void redistribution in Liquefied Ishimaru, M., 2004. Possibility of Postliquefaction
deposits, Soil Dynamics and Earthquake Engineering Flow Failure due to Seepage, J. of Geotech.
23, 585-603. Geoenviron. Eng., ASCE, Vol. 130, No. 7, July 1,er 1,
Kokusho, T., 1999. Water film in liquefied sand and its pp. 707-716.
effect on lateral spread, J. Geotech. Geoenviron. Vaid, Y.P., and Eliadorani, A., 1998. Instability and
Eng. 125(10), pp. 817–826. liquefaction of granular soils under undrained and
Kulasingam, R., Malvick, E.J., Boulanger, R.W., and partially drained states, Canadian Geotech. J. Vol.
Kutter, B.L., 2004. Strength loss and localization of silt 35, pp. 1053–1062.
interlayers in slopes of liquefied sand, J. of Geotech. Yang, D., Naesgaard, E., Byrne, P.M. Adalier, K., and
Geoenviron. Eng., ASCE, Vol. 130, No. 11, November Abdoun, T. (2004). “Numerical Model Verification and
1, pp. 1192-1202. Calibration of George Massey Tunnel Using
Malvick, E. J., 2005. Void Redistribution-Induced Shear Centrifuge Models.” Canadian Geot. J., 41(5), 921-
Localization and Deformation in Slope, Ph.D. Thesis, 942.
University of California, Davis. Vasquez-Herrera, A. and Dobry, R., 1989. Re-evaluation
Naesgaard, E., and Byrne, P. M., 2005. Flow Liquefaction of the Lower San Fernando Dam, Rpt. 3, US Army
due to Mixing of Layered Deposits, Proc. of Geot. Engineer Waterways Experiment Station, Contract
Earthquake Eng. Satellite Conf., TC4 Committee, Rpt. GL-89-2.
ISSMGE, Osaka, Japan, Sept. Whitman, R.V 1985. On liquefaction. In: Proceedings,
Naesgaard, E., Byrne, P. M., Seid-Karbasi, M., and Park, 11th International Conference on Soil Mechanics and
S. S., 2005. Modeling flow Liquefaction, its mitigation, Foundation Engineering. San Francisco, CA. A.A.
and comparison with centrifuge tests, Proc. of Geot. Balkema, pp 1923-1926.
Earthquake Eng. Satellite Conf., TC4 Committee, Yang, Z., and Elgamal, A., 2002. Influence of permeability
ISSMGE, Osaka, Japan, Sept. on liquefaction-induced shear deformation, Journal of
Naesgaard, E., Byrne, P.M., and Seid-Karbasi, M. 2006. Engineering Mechanics, ASCE, Vol. 128, No. 7, July,
Modeling flow liquefaction and pore water pp. 720-729.
th
redistribution mechanisms, 8 NCEE, San Francisco, Yoshida, N., and Finn, W.D.L., 2000. Simulation of
April liquefaction beneath an impermeable surface layer,
Olson, S. M. and Stark, T. D., 2002. Liquefied strength Soil Dyn. Earthquake Eng. 19, pp. 333-338.
ratio from liquefaction flow failure case histories,
Canadian Geot. J., Vol 39, pp. 629–647.

You might also like