You are on page 1of 12

Return Period of Soil Liquefaction

Steven L. Kramer, M.ASCE1; and Roy T. Mayfield2

Abstract: The paper describes a performance-based approach to the evaluation of liquefaction potential, and shows how it can be used
to account for the entire range of potential ground shaking. The result is a direct estimate of the return period of liquefaction, rather than
a factor of safety or probability of liquefaction conditional upon ground shaking with some specified return period. As such, the
performance-based approach can be considered to produce a more complete and consistent indication of the actual likelihood of lique-
faction at a given location than conventional procedures. In this paper, the performance-based procedure is introduced and used to
compare likelihoods of the initiation of liquefaction at identical sites located in areas of different seismicity. The results indicate that the
likelihood of liquefaction depends on the position and slope of the peak acceleration hazard curve, and on the distribution of earthquake
magnitudes contributing to the ground motion hazard. The results also show that the consistent use of conventional procedures for the
evaluation of liquefaction potential produces inconsistent actual likelihoods of liquefaction.
DOI: 10.1061/共ASCE兲1090-0241共2007兲133:7共802兲
CE Database subject headings: Earthquakes; Liquefaction; Sand; Penetration tests; Hazards.

Introduction probability of liquefaction conditional upon ground shaking with


some specified return period. As such, the performance-based
Liquefaction of soil has been a topic of considerable interest to approach can be considered to produce a more complete and
geotechnical engineers since its devastating effects were widely consistent indication of the likelihood of liquefaction at a given
observed following 1964 earthquakes in Niigata, Japan and location than conventional procedures. In this paper, the
Alaska. Since that time, a great deal of research on soil liquefac- performance-based procedure is introduced and then used to com-
tion has been completed in many countries that are exposed to pare the actual likelihoods of liquefaction at identical sites located
this important seismic hazard. This work has resulted in the de- in areas of different seismicity; the results show that the consis-
velopment of useful empirical procedures that allow the determin- tent use of conventional procedures for evaluation of liquefaction
potential produces inconsistent actual likelihoods of liquefaction.
istic and probabilistic evaluation of liquefaction potential for a
specified level of ground shaking.
In practice, the level of ground shaking is usually obtained
from the results of a probabilistic seismic hazard analysis Liquefaction Potential
共PSHA兲; although that ground shaking model is determined
Liquefaction potential is generally evaluated by comparing con-
probabilistically, a single level of ground shaking is selected and
sistent measures of earthquake loading and liquefaction resis-
used within the liquefaction potential evaluation. In reality,
tance. It has become common to base the comparison on cyclic
though, a given site may be subjected to a wide range of ground
shear stress amplitude, usually normalized by initial vertical ef-
shaking levels ranging from low levels that occur relatively fre-
fective stress and expressed in the form of a cyclic stress ratio,
quently to very high levels that occur only rarely, each with dif-
CSR, for loading and a cyclic resistance ratio, CRR, for resis-
ferent potential for triggering liquefaction.
tance. The potential for liquefaction is then described in terms of
This paper shows how the entire range of potential ground
a factor of safety against liquefaction, FSL = CRR/ CSR.
shaking can be considered in a fully probabilistic liquefaction
potential evaluation using a performance-based earthquake engi-
neering 共PBEE兲 framework. The result is a direct estimate of the Characterization of Earthquake Loading
return period of liquefaction, rather than a factor of safety or The CSR is most commonly evaluated using the “simplified
method” first described by Seed and Idriss 共1971兲, which can be
1
Professor, Dept. of Civil and Environmental Engineering, Univ. of expressed as
Washington, Seattle, WA 98195-2700. E-mail: kramer@u.washington.edu
2
Consulting Engineer, Kirkland, WA 98034; formerly, Graduate Re- amax ␴vo rd
search Assistant, Dept. of Civil and Environmental Engineering, Univ. of CSR = 0.65 · · 共1兲
Washington. E-mail: roy@mayfield.name g ␴⬘vo MSF
Note. Discussion open until December 1, 2007. Separate discussions where amax = peak ground surface acceleration; g = acceleration of
must be submitted for individual papers. To extend the closing date by
gravity 共in same units as amax兲; ␴vo = initial vertical total stress;
one month, a written request must be filed with the ASCE Managing
Editor. The manuscript for this paper was submitted for review and pos- ␴⬘vo = initial vertical effective stress; rd = depth reduction factor;
sible publication on April 4, 2006; approved on July 20, 2006. This paper and MSF= magnitude scaling factor, which is a function of earth-
is part of the Journal of Geotechnical and Geoenvironmental Engineer- quake magnitude. The depth reduction factor accounts for com-
ing, Vol. 133, No. 7, July 1, 2007. ©ASCE, ISSN 1090-0241/2007/7- pliance of a typical soil profile, and the MSF acts as a proxy for
802–813/$25.00. the number of significant cycles, which is related to the ground

802 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Fig. 1. 共a兲 Deterministic cyclic resistance curves proposed by Youd et al. 2001, ASCE; 共b兲 cyclic resistance curves of constant probability of
liquefaction with measurement/estimation errors by Cetin et al. 2004, ASCE.

motion duration. It should be noted that two pieces of loading Research 1997兲. The liquefaction evaluation procedure described
information—amax and earthquake magnitude—are required for by Youd et al. 共2001兲 will be referred to hereafter as the NCEER
estimation of the CSR. procedure. The NCEER procedure has been shown to produce
reasonable predictions of liquefaction potential 共i.e., few cases of
Characterization of Liquefaction Resistance nonprediction for sites at which liquefaction was observed兲 in
past earthquakes, and is widely used in contemporary geotechni-
The CRR is generally obtained by correlation to in situ test re- cal engineering practice. For the purposes of this paper, a conven-
sults, usually standard penetration 共SPT兲, cone penetration 共CPT兲, tionally liquefaction-resistant site will be considered to be one for
or shear wave velocity 共Vs兲 tests. Of these, the SPT has been most which FSL 艌 1.2 for a 475-year ground motion using the NCEER
commonly used and will be used in the remainder of this paper. A
procedure. This standard is consistent with that recommended by
number of SPT-based procedures for deterministic 共Seed and
Martin and Lew 共1999兲, for example, and is considered represen-
Idriss 1971; Seed et al. 1985; Youd et al. 2001; Idriss and Bou-
tative of those commonly used in current practice.
langer 2004兲 and probabilistic 共Liao et al. 1988; Toprak et al.
1999; Youd and Noble 1997; Juang and Jiang 2000; Cetin et al.
Probabilistic Approach
2004兲 estimation of liquefaction resistance have been proposed.
Recently, a detailed review and careful reinterpretation of lique-
Deterministic Approach faction case histories 共Cetin 2000; Cetin et al. 2004兲 was used to
Fig. 1共a兲 illustrates the widely used liquefaction resistance curves develop new probabilistic procedures for the evaluation of lique-
recommended by Youd et al. 共2001兲, which are based on discus- faction potential. The probabilistic implementation of the Cetin et
sions at a National Center for Earthquake Engineering Research al. 共2004兲 procedure produces a probability of liquefaction, PL,
共NCEER兲 workshop 共National Center for Earthquake Engineering which can be expressed as


PL = ⌽ −
共N1兲60共1 + ␪1FC兲 − ␪2 ln CSReq − ␪3 ln M w − ␪4 ln共␴⬘vo/pa兲 + ␪5FC + ␪6
␴␧
册 共2兲

where ⌽ = standard normal cumulative distribution function; faction includes both loading terms 共again, peak acceleration, as
共N1兲60 = corrected SPT resistance; FC= fines content 共in percent兲; reflected in the CSR, and magnitude兲 and resistance terms 共SPT
CSReq = cyclic stress ratio 关Eq. 共1兲 without MSF兴; M w = moment resistance, FC, and vertical effective stress兲. Mean values of the
magnitude; ␴⬘vo = initial vertical effective stress, pa is atmospheric model coefficients are presented for two conditions in Table 1—a
pressure in same units as ␴⬘vo; ␴␧ = measure of the estimated model case in which the uncertainty includes parameter measurement/
and parameter uncertainty; and ␪1–␪6 are model coefficients ob- estimation errors and a case in which the effects of measurement/
tained by regression. As Eq. 共2兲 shows, the probability of lique- estimation errors have been removed. The former would corre-

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 803


Table 1. Cetin et al. 共2004兲 Model Coefficients with and without Measurement/Estimation Errors 共Adapted from Cetin et al. 2002兲
Case Measurement/estimation errors ␪1 ␪2 ␪3 ␪4 ␪5 ␪6 ␴␧
I Included 0.004 13.79 29.06 3.82 0.06 15.25 4.21
II Removed 0.004 13.32 29.53 3.70 0.05 16.85 2.70

spond to uncertainties that exist for a site investigated with a found that the average effective stress for their critical layers were
normal level of detail and the latter to a “perfect” investigation at lower effective stresses 共⬃0.65 atm兲 instead of the standard 1
关i.e., no uncertainty in any of the variables on the right-side of atm, and made allowances for those differences. Also, the basic
Eq. 共2兲兴. Fig. 1共b兲 shows contours of equal PL for conditions in shapes of the cyclic resistance curves are different—the Cetin
which measurement/estimation errors are included; the et al. 共2004兲 curves 共Fig. 3兲 have a smoothly changing curvature
measurement/estimation errors have only a slight influence on the
while the Youd et al. 共2001兲 curve 共Fig. 1兲 is nearly linear at
model coefficients but a significant effect on the uncertainty term,
intermediate SPT resistances 关共N1兲60 ⬇ 10– 22兴 with higher curva-
␴ ␧.
Direct comparison of the procedures described by Youd et al. tures at lower and higher SPT resistances. An approximate com-
共2001兲 and Cetin et al. 共2004兲 is difficult because various aspects parison of the two methods can be made by substituting CRR for
of the procedures are different. For example, Cetin et al. 共2004兲 CSReq in Eq. 共2兲 and then rearranging the equation in the form

CRR = exp 冋 共N1兲60共1 + ␪1FC兲 − ␪3 ln M w − ␪4 ln共␴⬘vo/pa兲 + ␪5FC + ␪6 + ␴␧⌽−1共PL兲


␪2
册 共3兲

where ⌽−1 = inverse standard normal cumulative distribution Table 1兲 shows equivalence of FSL when a value of PL ⬇ 0.6 is
function. The resulting value of CRR can then be used in the used. Cetin et al. 共2004兲 suggest the use of a deterministic curve
common expression for FSL. Arango et al. 共2004兲 used this for- equivalent to that given by Eq. 共3兲 with PL = 0.15, which would
mulation without measurement/estimation errors 共Case II in Table produce a more conservative result than the NCEER procedure.
1 and found that the Cetin et al. 共2004兲 and NCEER procedures The differences between the two procedures are most pronounced
yielded similar values of FSL for a site in San Francisco when a at high CRR values; the NCEER procedure contains an implicit
value of PL ⬇ 0.65 was used in Eq. 共3兲. A similar exercise for a assumption of 共N1兲60 = 30 as an upper bound to liquefaction sus-
site in Seattle with measurement/estimation errors 共Case I in ceptibility while Cetin et al. 共2004兲, whose database contained

Fig. 2. Magnitude and distance deaggregation of 475-year peak acceleration hazard for site in Seattle

804 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Fig. 3. Distributions of magnitude contributing to peak rock outcrop acceleration for different return periods in Seattle: 共a兲 TR = 108 years; 共b兲
TR = 224 years; 共c兲 TR = 475 years; 共d兲 TR = 975 years; 共e兲 TR = 2,475 years; and 共f兲 TR = 4,975 years

considerably more cases at high CSR levels, indicate that lique- analyzed by the U.S. Geological Survey 共USGS兲 共http://
faction is possible 共albeit with limited potential effects兲 at 共N1兲60 eqhazmaps.usgs.gov兲 are shown in Fig. 3. The decreasing signifi-
values above 30. cance of lower magnitude earthquakes for longer return periods,
evident in Fig. 3, is a characteristic shared by many other
locations.
Seismic Hazard Analysis

Ground shaking levels used in seismic design and hazard evalu- Performance-Based Liquefaction Potential
ations are generally determined by means of seismic hazard Evaluation
analyses. Deterministic seismic hazard analyses are used most
often for special structures or for estimation of upper bound In practice, liquefaction potential is usually evaluated using de-
ground shaking levels. In the majority of cases, however, ground terministic CRR curves, a single ground motion hazard level, for
shaking levels are determined by probabilistic seismic hazard example, for ground motions with a 475-year return period, and a
analyses. single earthquake magnitude, usually the mean or mode. In con-
Probabilistic seismic hazard analyses consider the potential trast, the performance-based approach incorporates probabilistic
levels of ground shaking from all combinations of magnitude and CRR curves and contributions from all hazard levels and all
distance for all known sources capable of producing significant earthquake magnitudes.
shaking at a site of interest. The distributions of magnitude and The roots of performance-based liquefaction assessment are in
distance, and of ground shaking level conditional upon magnitude the method of seismic risk analysis introduced by Cornell 共1968兲.
and distance, are combined in a way that allows estimation of the The first known application of this approach to liquefaction as-
mean annual rate at which a particular level of ground shaking sessment was presented by Yegian and Whitman 共1978兲, although
will be exceeded. The mean annual rate of exceeding a ground earthquake loading was described as a combination of earthquake
motion parameter value, y *, is usually expressed as ␭y*; the recip- magnitude and source-to-site distance rather than peak accelera-
rocal of the mean annual rate of exceedance is commonly referred tion and magnitude. Atkinson et al. 共1984兲 developed a procedure
to as the return period. The results of a PSHA are typically pre- for the estimation of the annual probability of liquefaction using
sented in the form of a seismic hazard curve, which graphically linearized approximations of the CRR curves of Seed and Idriss
illustrates the relationship between ␭y* and y *. 共1983兲 in a deterministic manner. Marrone et al. 共2003兲 described
The ground motion level associated with a particular return liquefaction assessment methods that incorporate probabilistic
period is therefore influenced by contributions from a number of CRR curves and the full range of magnitudes and peak accelera-
different magnitudes, distances, and conditional exceedance prob- tions in a manner similar to the PBEE framework described
ability levels 共usually expressed in terms of a parameter, ␧, de- herein. Hwang et al. 共2005兲 described a Monte Carlo simulation-
fined as the number of standard deviations by which ln y * exceeds based approach that produces similar results.
the natural logarithm of the median value of y for a given M and PBEE is generally formulated in a probabilistic framework to
R兲. The relative contributions of each M-R pair to ␭y* can be evaluate the risk associated with earthquake shaking at a particu-
quantified by means of a deaggregation analysis 共McGuire, 1995兲; lar site. The risk can be expressed in terms of economic loss,
the deaggregated contributions of magnitude and distance are fre- fatalities, or other measures. The Pacific Earthquake Engineering
quently illustrated in diagrams such as that shown in Fig. 2. Be- Research 共PEER兲 has developed a probabilistic framework for
cause both peak acceleration and magnitude are required for PBEE 共Cornell and Krawinkler 2000; Krawinkler 2002; Deierlein
cyclic stress-based evaluations of liquefaction potential, the mar- et al. 2003兲 that computes risk as a function of ground shaking
ginal distribution of magnitude can be obtained by summing the through the use of several intermediate variables. The ground
contributions of each distance and ␧ value for each magnitude; motion is characterized by an intensity measure, IM, which could
magnitude distributions for six return periods at a site in Seattle be any one of a number of ground motion parameters 共e.g., amax,

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 805


Arias intensity, etc.兲. The effects of the IM on a system of interest
are expressed in terms used primarily by engineers in the form of
engineering demand parameters, or EDPs 共e.g., excess pore pres-
sure, FSL, etc.兲. The physical effects associated with the EDPs
共e.g. settlement, lateral displacement, etc.兲 are expressed in terms
of damage measures, or DMs. Finally, the risk associated with the
DM is expressed in a form that is useful to decisionmakers by
means of decision variables, DV 共e.g. repair cost, downtime,
etc.兲. The mean annual rate of exceedance of various DV levels,
␭DV, can be expressed in terms of the other variables as

NDM NEDP NIM

␭dv = 兺 兺 兺 P关DV ⬎ dv兩DM = dmk兴


k=1 j=1 i=1

⫻P关DM = dmk兩EDP = edp j兴P关EDP = edp j兩IM = imi兴⌬␭imi


共4兲

where P关a 兩 b兴 describes the conditional probability of a given b;


and NDM, NEDP, and NIM = number of increments of DM, EDP, and
IM, respectively. Extending this approach to consider epistemic Fig. 4. Seismic hazard curves for Seattle deaggregated on the basis
uncertainty in IM, although not pursued in this paper, is straight- of magnitude. Total hazard curve is equal to sum of hazard curves for
forward. By integrating over the entire hazard curve 共approxi- all magnitudes.
mated by the summation over i = 1, NIM兲, the performance-based
approach includes contributions from all return periods, not just
the return periods mandated by various codes or regulations.
nonexceedance of FSL* = 1.0, i.e., TR,L = 1 / ⌳FS* =1.0.
For a liquefiable site, the geotechnical engineer’s initial con- L
tribution to this process for evaluating liquefaction hazards comes The PEER framework assumes IM sufficiency, i.e., that the
primarily in the evaluation of P关EDP兩 IM兴. Representing the EDP intensity measure is a scalar that provides all of the information
by FSL and combining the probabilistic evaluation of FSL with the required to predict the EDP. This sufficiency, however, does not
results of a seismic hazard analysis allows the mean annual rate of exist for cyclic stress-based liquefaction potential evaluation pro-
nonexceedance of a selected factor of safety, FSL* , to be computed cedures as evidenced by the long-recognized need for a MSF.
as Therefore, FSL depends on more than just peak acceleration as an
intensity measure, and calculation of the mean annual rate of
exceeding some factor of safety against liquefaction, FSL* , can be
NIM modified as
⌳FS* =
L

i=1
P关FSL ⬍ FSL* 兩IMi兴⌬␭IMi
共5兲
N M Namax

The value of ⌳FS* should be interpreted as the mean annual rate


⌳FS* =
L
兺 兺 P关FSL ⬍ FSL* 兩amax ,m j兴⌬␭a
j=1 i=1
i maxi,m j
共6兲
L
共or inverse of the return period兲 at which the actual factor of
safety will be less than FSL* . Note that ⌳FS* increases with increas- where NM and Namax = number of magnitude and peak acceleration
L
ing FSL* since weaker motions producing higher factors of safety increments into which “hazard space” is subdivided; and
occur more frequently than stronger motions that produce lower ⌬␭amax ,m j = incremental mean annual rate of exceedance for inten-
i
factors of safety. The mean annual rate of factor of safety nonex- sity measure, amaxi, and magnitude, m j. The values of ␭amax,m can
ceedance is used because nonexceedance of a particular factor of be visualized as a series of seismic hazard curves distributed with
safety represents an undesirable condition, just as exceedance of respect to magnitude according to the results of a deaggregation
an intensity measure does; because lower case lambda is com- analysis 共Fig. 3兲; therefore, their summation 共over magnitude兲
monly used to represent mean annual rate of exceedance, an yields the total seismic hazard curve for the site 共Fig. 4兲. The
upper case lambda is used here to represent mean annual rate of conditional probability term in Eq. 共6兲 can be calculated using the
nonexceedance. Since liquefaction is expected to occur when probabilistic model of Cetin et al. 共2004兲, as described in Eq. 共2兲,
CRR⬍ CSR 共i.e., when FSL* ⬍ 1.0兲, the return period of liquefac- with CSR= CSReq,i · FSL* 共with CSReq,i computed from amax,i兲 and
tion corresponds to the reciprocal of the mean annual rate of M w = m j, i.e.


P关FSL ⬍ FSL* 兩amaxi,m j兴 = ⌽ −
共N1兲60共1 + ␪1FC兲 − ␪2 ln共CSReq,i · FSL* 兲 − ␪3 ln m j − ␪4 ln共␴⬘vo/pa兲 + ␪5FC + ␪6
␴␧
册 共7兲

806 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Another way of characterizing liquefaction potential is in P关N ⬍ Nreq兴 = P关FSL ⬍ 1.0兴. The PBEE approach can then be ap-
terms of the liquefaction resistance required to produce a desired plied in such a way as to produce a mean annual rate of exceed-
*
level of performance. For example, the SPT value required to ance for Nreq
resist liquefaction, Nreq, can be determined at each depth of inter-
est. The difference between the actual SPT resistance and the N M Namax
required SPT resistance would provide an indication of how much
soil improvement might be required to bring a particular site to an
␭ N* =
req
兺 兺 P关Nreq ⬎ Nreq
j=1 i=1
*
兩amax ,m j兴⌬␭a
i maxi,m j
共8兲

acceptable factor of safety against liquefaction. Given that lique-


faction would occur when N ⬍ Nreq, or when FSL ⬍ 1.0, then where

P关Nreq ⬎ Nreq
*
兩amaxi,m j兴 = ⌽ −冋 *
Nreq共1 + ␪1FC兲 − ␪2 ln CSReq,i − ␪3 ln m j − ␪4 ln共␴⬘vo/pa兲 + ␪5FC + ␪6
␴␧
册 共9兲

*
The value of Nreq can be interpreted as the SPT resistance re- ing. The wide range of smoothly increasing SPT resistance, while
quired to produce the desired performance level for shaking with perhaps unlikely to be realized in a natural depositional environ-
a return period of 1 / ␭N* . ment, is useful for illustrating the main points of this paper.
req

Locations
Comparison of Conventional and
Performance-Based Approaches In order to illustrate the effects of different seismic environments
on liquefaction potential, the hypothetical site was assumed to be
Conventional procedures provide a means for evaluating the liq- located in each of the ten United States cities listed in Table 2. For
uefaction potential of a soil deposit for a given level of loading. each location, the local seismicity was characterized by the proba-
When applied consistently to different sites in the same seismic bilistic seismic hazard analyses available from the USGS 共using
environment, they provide a consistent indication of the likeli- the 2002 interactive deaggregation link with listed latitudes and
hood of liquefaction 共expressed in terms of FSL or PL兲 at those longitudes兲. In addition to being spread across the United States,
sites. The degree to which they provide a consistent indication of these locations represent a wide range of seismic environments;
liquefaction likelihood when applied to sites in different seismic the total seismic hazard curves for each of the locations are shown
environments, however, has not been established. That issue is in Fig. 6. The seismicity levels vary widely—475-year peak ac-
addressed in the remainder of this paper. celeration values range from 0.12g 共Butte兲 to 0.66g 共Eureka兲.
Two of the locations 共Charleston and Memphis兲 are in areas of
low recent seismicity with very large historical earthquakes, three
Idealized Site 共Seattle, Portland, and Eureka兲 are in areas subject to large-
Potentially liquefiable sites around the world have different like- magnitude subduction earthquakes, and two 共San Francisco and
lihoods of liquefaction due to differences in site conditions San Jose兲 are in close proximity 共⬃60 km兲 in a very active envi-
共which strongly affect liquefaction resistance兲 and local seismic ronment.
environments 共which strongly affect loading兲. The effects of seis-
mic environment can be isolated by considering the liquefaction Conventional Liquefaction Potential Analyses
potential of a single soil profile placed at different locations.
Fig. 5 shows the subsurface conditions for an idealized, hypo- Two sets of conventional deterministic analyses were performed
thetical site with corrected SPT resistances that range from rela- to illustrate the different degrees of liquefaction potential of the
tively low 关共N1兲60 = 10兴 to moderately high 关共N1兲60 = 30兴. Using the
cyclic stress-based approach, the upper portion of the saturated Table 2. Peak Ground Surface 共Quaternary Alluvium兲 Acceleration
sand would be expected to liquefy under moderately strong shak- Hazard Information for Ten U.S. Cities
Latitude Longitude 475-year 2,475-year
Location 共N兲 共W兲 amax amax
Butte, MT 46.003 112.533 0.120 0.225
Charleston, SC 32.776 79.931 0.189 0.734
Eureka, CA 40.802 124.162 0.658 1.023
Memphis, TN 35.149 90.048 0.214 0.655
Portland, OR 45.523 122.675 0.204 0.398
Salt Lake City, UT 40.755 111.898 0.298 0.679
San Francisco, CA 37.775 122.418 0.468 0.675
San Jose, CA 37.339 121.893 0.449 0.618
Santa Monica, CA 34.015 118.492 0.432 0.710
Fig. 5. Subsurface profile for idealized site Seattle, WA 47.530 122.300 0.332 0.620

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 807


amax,surface
Fa = = exp关− 0.15 − 0.13 ln amax,rock兴 共10兲
amax,rock
The amplification factor was applied deterministically so the un-
certainty in peak ground surface acceleration is controlled by the
uncertainties in the attenuation relationships used in the USGS
PSHAs. The uncertainties in peak ground surface accelerations
for soil sites are usually equal to or somewhat lower than those
for rock sites 共e.g., Toro et al. 1997; Stewart et al. 2003兲.
The results of the first set of analyses are shown in Fig. 7. Fig.
7共a兲 shows the variation of FSL with depth for the hypothetical
soil profile at each location. The results are, as expected, consis-
tent with the seismic hazard curves—the locations with the high-
est 475-year amax values have the lowest factors of safety against
liquefaction. Fig. 7共b兲 expresses the results of the conventional
det
analyses in a different way—in terms of Nreq , the SPT resistance
Fig. 6. USGS total seismic hazard curves for quaternary alluvium required to produce a performance level of FSL = 1.2 with the
conditions at different site locations 475-year ground motion parameters for each location. The 共N1兲60
values for the hypothetical soil profile are also shown in Fig. 7共b兲,
det
and can be seen to exceed the Nreq values at all locations/depths
for which FSL ⬎ 1.2. It should be noted that Nreq det
艋 30 for all cases
since the NCEER procedure implies zero liquefaction potential
hypothetical soil profile at the different site locations. The first set
共infinite FSL兲 for 共N1兲60 ⬎ 30.
of analyses was performed using the NCEER procedure with
The results of the second set of analyses are shown in Fig. 8,
475-year peak ground accelerations and magnitude scaling factors det det
both in terms of FSL and Nreq . The FSL and Nreq values are gen-
computed using the mean magnitude from the 475-year deaggre-
erally quite similar to those from the first set of analyses, except
gation of peak ground acceleration. The second set of analyses
that required SPT resistances are slightly in excess of 30 共as al-
was performed using Eq. 共2兲 with PL = 0.6 to produce a determin- lowed by the NCEER-C procedure兲 for the most seismically ac-
istic approximation to the NCEER procedure; these analyses will tive locations in the second set. The similarity of these values
be referred to hereafter as NCEER-C analyses. It should be noted confirms the approximation of the NCEER procedure by the
that, although applied deterministically in this paper, the NCEER-C procedure.
NCEER-C approximation to the NCEER procedure used here is
not equivalent to the deterministic procedure recommended by
Cetin et al. 共2004兲. In all analyses, the peak ground surface ac- Performance-Based Liquefaction Potential Analyses
celerations were computed from the peak rock outcrop accelera- The performance-based approach, which allows consideration of
tions obtained from the USGS 2002 interactive deaggregations all ground motion levels and fully probabilistic computation of
using a quaternary alluvium amplification factor 共Stewart et al. liquefaction hazard curves, was applied to each of the site loca-
2003兲 tions. Fig. 9 illustrates the results of the performance-based analy-

Fig. 7. Profiles of: 共a兲 factor of safety against liquefaction; 共b兲 required SPT resistance obtained using NCEER deterministic procedure with
475-year ground motions

808 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Fig. 8. Profiles of: 共a兲 factor of safety against liquefaction; 共b兲 required SPT resistance obtained using NCEER-C deterministic procedure for
475-year ground motions

ses for an element of soil near the center of the saturated zone 共at application of conventional procedures for evaluation of liquefac-
a depth of 6 m, at which 共N1兲60 = 18 for the hypothetical soil pro- tion potential. For each site location, the process is as follows:
file兲. Fig. 9共a兲 shows factor of safety hazard curves, and Fig. 9共b兲 1. At the depth of interest, determine the SPT resistance re-
PB
shows hazard curves for Nreq , the SPT resistance required to resist quired to produce a factor of safety of 1.2 using the conven-
liquefaction. Note that the SPT resistances shown in Fig. 9共b兲 are tional approach 关from either Figs. 7共b兲 or 8共b兲兴. At that SPT
those at which liquefaction would actually be expected to occur, resistance, the soils at that depth would have an equal lique-
rather than the values at which FSL would be as low as 1.2 共cor- faction potential 共i.e., FSL = 1.2 with a 475-year ground mo-
responding to a conventionally liquefaction resistant soil as de- tion兲 at all site locations as evaluated using the conventional
fined previously兲, which were plotted in Figs. 7 and 8. Therefore, approach.
the mean annual rates of exceedance in Fig. 9 are equal at each 2. Determine the mean annual rate of exceedance for the SPT
PB
site location for FSL = 1.0 and Nreq = 18. resistance from Step 1 using results of the type shown in Fig.
9共b兲 for each depth of interest. Since Fig. 9共b兲 shows the SPT
Equivalent Return Periods resistance for FSL = 1.0, this is the mean annual rate of lique-
The results of the conventional deterministic analyses shown in faction for soils with this SPT resistance at the depth of
Figs. 7 and 8 can be combined with the results of the interest.
performance-based analyses shown in Fig. 9 to evaluate the return 3. Compute the return period as the reciprocal of the mean an-
periods of liquefaction produced in different areas by consistent nual rate of exceedance.

Fig. 9. Seismic hazard curves for 6-m depth: 共a兲 factor of safety against liquefaction, FSL for 共N1兲60 = 18; 共b兲 required SPT resistance, Nreq
PB
, for
FSL = 1.0

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 809


Fig. 10. Profiles of return period of liquefaction for sites with equal liquefaction potential as evaluated by 共a兲 NCEER procedure; 共b兲 NCEER-C
procedure using 475-year ground motion parameters

4. Repeat Steps 1–3 for each depth of interest. Table 3 shows the return periods of liquefaction at a depth of
This process was applied to all site locations in Table 2 to evalu- 6 m 共the values are approximately equal to the averages over the
ate the return period for liquefaction as a function of depth for depth of the saturated zone for each site location兲 for conditions
each location; the calculations were performed using 475-year that would be judged as having equal liquefaction potential using
ground motions and again using 2,475-year ground motions. conventional procedures. Using both the NCEER and NCEER-C
Fig. 10 shows the results of this process for both sets of con- procedures, the actual return periods of liquefaction can be seen
ventional analyses. It is obvious from Fig. 10 that consistent ap- to vary significantly from one location to another, particularly for
plication of the conventional procedure produces inconsistent the case in which the conventional procedure was used with
return periods, and therefore different actual likelihoods of lique- 2,475-year motions.
faction, at the different site locations. Examination of the return The actual return periods depend on the seismic hazard curves
period curves shows that they are nearly vertical at depths greater and deaggregated magnitude distributions, and are different for
than about 4 m, indicating that the deterministic procedures are the NCEER and NCEER-C procedures. Using the NCEER pro-
relatively unbiased with respect to SPT resistance. The greater cedure with 475-year motions, the actual return periods of lique-
verticality of the curves based on the NCEER-C analyses results faction range from as short as 236 years 关for Eureka, which is
from the consistency of the shapes of those curves and the con- affected by the 共N1兲60 = 30 limit implied by the NCEER proce-
stant PL curves given by Eq. 共2兲, which were used in the dure兴 to 565 years 共Memphis兲; the corresponding 50-year prob-
performance-based analyses. Differences between the shapes of ability of liquefaction 共under the Poisson assumption兲 in Eureka
the NCEER curve 关Fig. 1共a兲兴 and the curves 关Fig. 1共b兲兴, particu- would be more than double that in Memphis. The return periods
larly for sites subjected to very strong shaking 共hence, very high computed using the NCEER-C procedure with 475-year motions
CSRs兲 such as San Francisco and Eureka, contribute to depth- are more consistent, but still range from 341 years 共San Jose兲 to
dependent return periods for the NCEER results. about 570 years 共Charleston and Memphis兲.
If deterministic liquefaction potential evaluations are based on
Table 3. Liquefaction Return Periods for 6 m Depth in Idealized Site at 2,475-year ground motions using the NCEER procedure, the im-
Different Site Locations Based on Conventional Liquefaction Potential plied limit of 共N1兲60 = 30 produces highly inconsistent actual re-
Evaluation Using 475-year and 2,475-year Motions turn periods—the 50-year probability of liquefaction in Eureka
would be more than six times that in Portland. All but two of the
475-year motions 2,475-year motions
ten locations would require 共N1兲30 = 30 according to that proce-
Location NCEER NCEER-C NCEER NCEER-C dure and, as illustrated in Fig. 9共b兲, the return periods for
PB
Nreq = 30 vary widely in the different seismic environments. The
Butte, MT 348 418 1,592 2,304
variations are smaller but still quite significant using the
Charleston, SC 532 571 1,433a 2,725
NCEER-C procedure.
Eureka, CA 236a 483 236a 1,590
Differences in regional seismicity can produce significant dif-
Memphis, TN 565 575 1,277a 2,532
ferences in ground motions at different return periods. Leyen-
Portland, OR 376 422 1,675 1,508 decker et al. 共2000兲 showed that short-period 共0.2 s兲 spectral
Salt Lake City, UT 552 543 1,316a 2,674 acceleration, for example, increased by about 50% when going
San Francisco, CA 355a 503 355a 1,736 from return periods of 475 years to 2,475 years in Los Angeles
San Jose, CA 360 341 532a 1,021 and San Francisco but by 200–500% or more in other areas of the
Santa Monica, CA 483 457 794a 1,901 country. The position and slope of the peak acceleration hazard
Seattle, WA 448 427 1,280a 2,155 curve clearly affect the return period of liquefaction. However,
a
Upper limit of 共N1兲60 = 30 implied by NCEER procedure reached. the regional differences in magnitude distribution 共i.e., the rela-

810 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Table 4. Required Penetration Resistances for 475-year Liquefaction Hazard in Element of Soil at 6 m Depth in Hypothetical Site at Different Site
Locations, and Ratios of Equivalent Factors of Safety
NCEER NCEER-C
NCEER NCEER-C 475-year det
CRR共Nreq 兲 det
CRR共Nreq 兲
det det PB
Location Nreq Nreq Nreq PB
CRR共Nreq 兲 PB
CRR共Nreq 兲
Butte, MT 5.6 6.6 7.3 0.88 0.95
Charleston, SC 14.5 15.9 12.1 1.19 1.32
Eureka, CA 30.0 34.9 34.8 0.0a 1.01
Memphis, TN 19.1 19.4 15.7 1.28 1.31
Portland, OR 17.0 17.9 18.8 0.88 0.94
Salt Lake City, UT 22.7 22.5 21.1 1.12 1.11
San Francisco, CA 30.0 31.8 31.5 0.0a 1.02
San Jose, CA 28.5 28.3 29.6 0.92 0.91
Santa Monica, CA 27.6 27.3 27.5 1.01 0.99
Seattle, WA 23.8 23.5 24.2 0.97 0.95
a
Zero values caused by upper limit of 共N1兲60 = 30 implied by NCEER procedure.

tive contribution to peak acceleration hazard from each magni- 14.3 for a 475-year return period of liquefaction 共performance-
tude兲 also contribute to differences in return period; San based procedure兲. For Portland, the relative unconservatism in the
Francisco and San Jose have substantially different return periods conventional approach means that the required SPT resistance
for liquefaction despite the similarity of their hazard curves be- increases slightly from 17.0 共deterministic procedure兲 to 17.5
cause the relative contributions of large magnitude earthquakes on 共performance-based procedure兲.
the San Andreas fault are higher for San Francisco than for San Because cyclic resistance ratio varies nonlinearly with SPT
Jose. resistance, it is also useful to consider the difference between the
deterministic and performance-based approaches from a factor of
safety standpoint. Since FSL is proportional to CRR, the ratio of
Conditions for Consistent Liquefaction Potential the CRR values corresponding to the SPT resistances in Table 4
can be thought of as factor of safety ratios that describe the
The differences in actual liquefaction return periods produced by “extra” liquefaction resistance required by the deterministic crite-
conventional liquefaction potential evaluations make it difficult to rion 共FSL = 1.2 for 475-year ground motion兲 relative to that re-
establish uniform and consistent procedures for conventional quired by the performance-based criterion 共475-year return period
evaluation of seismic hazards such as liquefaction potential. The for liquefaction兲. These values, shown in Table 4, range from 0.88
performance-based approach, however, provides a framework in to 1.28 for the NCEER procedure and from 0.91 to 1.32 for the
which design and evaluation can be based on a specified return NCEER-C procedure; higher values are associated with locations
period for liquefaction rather than on the basis of a factor of that are “penalized” by conventional deterministic procedures.
safety or probability of liquefaction for a ground motion with a The zero values shown for Eureka and San Francisco result from
specified return period. It is useful to compare the differences in the fact that NreqPB
⬎ 30 for those highly active areas; the corre-
possible requirements for acceptable liquefaction resistance pro- sponding CRR from the NCEER procedure is infinite. In Mem-
duced by the conventional and performance-based approaches. phis and Charleston, for example, the deterministic criterion
For the purposes of this paper, two alternative criteria will be would result in a factor of safety some 30% higher than that
considered: associated with an actual return period of 475 years. At several
1. The previously described conventional criterion of a mini- other locations, the same deterministic criterion would result in
mum FSL = 1.2 with 475-year ground motions; and factors of safety lower than those associated with the same actual
2. A performance-based criterion of a 475-year return period liquefaction hazard. In effect, the conventional procedure results
for liquefaction, i.e., TR共FSL = 1.0兲 = 475 years. in an owner in Memphis designing for an equivalent factor of
The 475-year return period used in the second criterion is in- safety that is about 45% higher than an owner in Portland.
tended as an example; a suitable specific return period for an
actual performance-based liquefaction criterion would need to be
identified and endorsed by a group of experienced professionals. Summary and Conclusions
The SPT resistances required to satisfy the first criterion at a
depth of 6 m, computed using the NCEER and NCEER-C proce- The evaluation of liquefaction potential involves comparison of
dures, are listed in Table 4, as well as the SPT resistances required consistent measures of loading and resistance. In contemporary
to satisfy the second criterion at the same depth. For those loca- geotechnical engineering practice, such comparisons are com-
tions at which the return periods for liquefaction in Table 3 were monly made using cyclic shear stresses expressed in terms of
less than 475 years, the SPT resistances required for liquefaction normalized cyclic stress and cyclic resistance ratios. The CSR is
with an actual return period of 475 years are increased, and vice usually estimated using a simplified procedure in which the level
versa for locations at which the Table 2 return periods were of ground shaking is related to peak ground surface acceleration
greater than 475 years. For a location like Memphis, the relative and earthquake magnitude. Criteria by which liquefaction resis-
conservatism in the conventional approach means that the re- tance is judged to be adequate are usually expressed in terms of a
quired SPT resistance of 19.1 for FSL = 1.2 with the 475-year single level of ground shaking.
motion 共NCEER procedure兲 is reduced to an SPT resistance of For a given soil profile at a given location, liquefaction can be

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 811


caused by a range of ground shaking levels—from strong ground References
motions that occur relatively rarely to weaker motions that occur
more frequently. Performance-based procedures allow consider- Arango, I., Ostadan, F., Cameron, J., Wu, C. L., and Chang, C. Y. 共2004兲.
ation of all levels of ground motion in the evaluation of liquefac- “Liquefaction probability of the BART Transbay Tube backfill.”
tion potential. By integrating a probabilistic liquefaction evalua- Proc., 11th Int. Conf. on Soil Dynamics and Earthquake Engineering
tion procedure with the results of a PSHA, this paper presented a and 3rd Int. Conf. on Earthquake Geotechnical Engineering, Berke-
methodology for performance-based evaluation of liquefaction ley, Calif., I, 456–462.
potential. The methodology was used to illustrate differences be- Atkinson, G. M., Finn, W. D. L., and Charlwood, R. G. 共1984兲. “Simple
tween performance-based and conventional evaluation of lique- computation of liquefaction probability for seismic hazard applica-
faction potential. These analyses led to the following conclusions: tions.” Earthquake Spectra, 1共1兲, 107–123.
1. The actual potential for liquefaction, considering all levels of Cetin, K. O. 共2000兲. “Reliability-based assessment of seismic soil lique-
faction initiation hazard.” Ph.D. dissertation, Univ. of California,
ground motion, is influenced by the position and slope of the
Berkeley.
peak acceleration hazard curve and by the distributions of
Cetin, K. O., et al. 共2004兲. “Standard penetration test-based probabilistic
earthquake magnitude that contribute to peak acceleration and deterministic assessment of seismic soil liquefaction potential.” J.
hazard at different return periods. Geotech. Geoenviron. Eng., 130共12兲, 1314–1340.
2. Consistent application of conventional procedures for evalu- Cetin, K. O., Der Kiureghian, A., and Seed, R. B. 共2002兲. “Probabilistic
ation of liquefaction potential 共i.e., based on a single ground models for the initiation of soil liquefaction.” Struct. Safety 24, 67–
motion level兲 to sites in different seismic environments can 82.
produce highly inconsistent estimates of actual liquefaction Cornell, C. A. 共1968兲. “Engineering seismic risk analysis.” Bull. Seismol.
hazards. Soc. Am. 58共5兲, 1583–1606.
3. Criteria based on conventional procedures for evaluation of Cornell, C. A., and Krawinkler, H. 共2000兲. “Progress and challenges in
liquefaction potential can produce significantly different liq- seismic performance assessment.” PEER News, April 1–3, PEER,
uefaction hazards even for sites in relatively close proximity Berkeley, Calif.
Deierlein, G. G., Krawinkler, H., and Cornell, C. A. 共2003兲. “A frame-
to each other. For the locations considered in this paper, such
work for performance-based earthquake engineering.” Proc., 2003
criteria were generally more strict 共i.e., resulted in longer
Pacific Conf. on Earthquake Engineering, Wellington, New Zealand.
return periods, hence lower probabilities, of liquefaction兲 for
Idriss, I. M., and Boulanger, R. W. 共2004兲. “Semiempirical procedures for
locations with flatter peak acceleration hazard curves than for
evaluating liquefaction potential during earthquakes.” Proc., 11th Int.
locations with steeper hazard curves, and for locations at
Conf. on Soil Dynamics and Earthquake Engineering and 3rd Int.
which large magnitude earthquakes contributed a relatively Conf. on Earthquake Geotechnical Engineering, Berkeley, Calif., I,
large proportion of the total hazard. 32–56.
4. Criteria that would produce more uniform liquefaction haz- Hwang, J. H., Chen, C. H., and Juang, C. H. 共2005兲. “Liquefaction hazard
ards at locations in all seismic environments could be devel- analysis: A fully probabilistic method.” Proc., of the Sessions of the
oped by specifying a standard return period for liquefaction. Geo-Frontiers 2005 Congress, Earthquake Engineering and Soil Dy-
Performance-based procedures such as the one described in namics, R. W. Boulanger et al., eds, ASCE, Reston, Va., Paper No. 22.
this paper could be used to evaluate individual sites with Juang, C. H., and Jiang, T. 共2000兲. “Assessing probabilistic methods for
respect to such criteria. liquefaction potential evaluation.” Soil dynamics and liquefaction
The performance-based methodology described in this paper 2000, R. Y. S. Pak and J. Yamamuro, eds., Geotechnical Special Pub-
makes use of a recently developed procedure for estimation of the lication, 107, ASCE, New York, 148–162.
probability of liquefaction. While this procedure is very well Krawinkler, H. 共2002兲. “A general approach to seismic performance as-
suited for implementation into the performance-based methodol- sessment.” Proc., Int. Conf. on Advances and New Challenges in
ogy, other probabilistic liquefaction procedures could also be Earthquake Engineering Research, ICANCEER, Hong Kong.
used. The methodology, which deals with the initiation of lique- Leyendecker, E. V., Hunt, R. J., Frankel, A. D., and Rukstales, K. R.
faction, can also be extended to estimate return periods for vari- 共2000兲. “Development of maximum considered earthquake ground
ous effects of liquefaction such as lateral spreading displacement motion maps.” Earthquake Spectra, 16共1兲, 21–40.
or ground surface settlement; research on these issues is under- Liao, S. S. C., Veneziano, D., and Whitman, R. V. 共1988兲. “Regression
models for evaluating liquefaction probability.” J. Geotech. Engrg.,
way.
114共4兲, 389–411.
Marrone, J., Ostadan, F., Youngs, R., and Litehiser, J. 共2003兲. “Probabi-
listic liquefaction hazard evaluation: Method and application.” Proc.,
Acknowledgments 17th Int. Conf. Structural Mechanics in Reactor Technology (SMiRT
17), Prague, Czech Republic, Paper No. M02-1.
The research described in this paper was funded by the Washing- Martin, G. R., and Lew, M., eds. 共1999兲. “Recommended procedures for
ton State Department of Transportation; the support of Tony Allen implementation of DMG Special Publication 117—Guidelines for
and Kim Willoughby is gratefully acknowledged. A portion of the analyzing and mitigating liquefaction hazards in California.” Southern
work was completed while the senior writer was on sabbatical California Earthquake Center, Los Angeles.
leave at the International Centre for Geohazards at the Norwegian McGuire, R. K. 共1995兲. “Probabilistic seismic hazard analysis and design
earthquakes: Closing the loop.” Bull. Seismol. Soc. Am., 85共5兲, 1275–
Geotechnical Institute, where he benefited greatly from discus-
1284.
sions with Dr. Farrokh Nadim. The performance-based approach
National Center for Earthquake Engineering Research. 共1997兲. Proc.,
described in the paper was motivated by the senior writer’s in-
NCEER Workshop on Evaluation of Liquefaction Resistance of Soils,
volvement with PEER, and particularly by discussions with Pro- T. L. Youd and I. M. Idriss eds., Technical Rep. No. NCEER, 97-022,
fessor C. Allin Cornell. The writers are grateful to Dr. Donald G. NCEER, Buffalo, N.Y.
Anderson, Professor Stephen E. Dickenson, Dr. Robert M. Pyke, Seed, H. B., and Idriss, I. M. 共1971兲. “Simplified procedure for evaluating
Professor Jonathan P. Stewart, and Steven G. Vick for their con- soil liquefaction potential.” J. Soil Mech. and Found. Div. 97, 1249–
structive comments. 1273.

812 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007


Seed, H. B and Idriss, I. M 共1983兲. Ground motions and soil liquefaction Toro, G. R., Abrahamson, N. A., and Schneider, J. F. 共1997兲. “Model of
during earthquakes, Earthquake Engineering Research Institute, Ber- strong ground motions from earthquakes in central and eastern North
keley, Calif., 134. America: Best estimates and uncertainties.” Seismol. Res. Lett., 68共1兲,
Seed, H. B., Tokimatsu, K., Harder, L. F., and Chung, R. M. 共1985兲. 41–57.
“Influence of SPT procedures in soil liquefaction resistance evalua- Yegian, M. Y., and Whitman, R. V. 共1978兲. “Risk analysis for ground
tions.” J. Geotech. Engrg., 111共12兲, 1425–1445. failure by liquefaction.” J. Geotech. Engrg. Div., 104, 921–938.
Stewart, J. P., Liu, A. H., and Choi, Y. 共2003兲. “Amplification factors for
Youd, T. L., et al. 共2001兲. “Liquefaction resistance of soils: Summary
spectral acceleration in tectonically active regions.” Bull. Seismol.
report from the 1996 NCEER and 1998 NCEER/NSF workshops on
Soc. Am., 93共1兲, 332–352.
Toprak, S., Holzer, T. L., Bennett, M. J., and Tinsley, J. C. III. 共1999兲. evaluation of liquefaction resistance of soils.” J. Geotech. Geoenvi-
“CPT- and SPT-based probabilistic assessment of liquefaction.” Proc., ron. Eng., 127, 817–833.
Youd, T. L., and Noble, S. K. 共1997兲. “Liquefaction criteria based on
7th U.S.-Japan Workshop on Earthquake Resistant Design of Lifeline
Facilities and Countermeasures Against Liquefaction, Multidisci- statistical and probabilistic analyses.” Proc., NCEER Workshop on
plinary Center for Earthquake Engineering Research, Buffalo, N.Y., Evaluation of Liquefaction Resistance of Soils, National Center for
69–86. Earthquake Engineering Research, Buffalo, N.Y., 201–205.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / JULY 2007 / 813

You might also like