Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Engineering Geology 76 (2005) 235 – 261

www.elsevier.com/locate/enggeo

Geotechnical analysis of paleoseismic shaking using


liquefaction features: a major updating
Scott M. Olsona,*, Russell A. Greenb, Stephen F. Obermeierc,d
a
Department of Civil and Environmental Engineering, University of Illinois – Urbana-Champaign,
2230d Newmark Civil Engineering Laboratory, Urbana, IL 61801, USA
b
Department of Civil and Environmental Engineering, University of Michigan, 2372 G.G. Brown Building, Ann Arbor, MI 48109-2125, USA
c
Emeritus, U.S. Geological Survey, Reston, VA, USA
d
EqLiq Consulting, Rockport, IN 47635, USA
Accepted 2 July 2004
Available online 13 September 2004

Abstract

A new methodology is proposed for the geotechnical analysis of strength of paleoseismic shaking using liquefaction effects.
The proposed method provides recommendations for selection of both individual and regionally located test sites, provides
techniques for validation of field data for use in back-analysis, and presents a recently developed energy-based solution to back-
calculate paleoearthquake magnitude and strength of shaking. The proposed method allows investigators to qualitatively assess
the influence of post-earthquake density change and aging. The proposed method also describes how the back-calculations from
individual sites should be integrated into a regional assessment of paleoseismic parameters.
D 2004 Elsevier B.V. All rights reserved.

Keywords: Liquefaction; Paleoliquefaction; Paleoseismicity; Paleoseismic shaking; Geotechnical analysis; Earthquake magnitude; Soil aging;
Magnitude bound

Contents

1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
2. Analytical approaches to paleoliquefaction studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
2.1. Back-calculations using liquefaction evaluation procedures . . . . . . . . . . . . . . . . . . . . . . . . . 237
2.1.1. Cyclic stress method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
2.1.2. Overview of Green and Mitchell’s energy-based method . . . . . . . . . . . . . . . . . . . . . 240
2.1.3. Brief discussion of the cyclic stress and energy-based methods . . . . . . . . . . . . . . . . . . 242

* Corresponding author. Tel.: +1 217 265 7584; fax: +1 217 265 8041.
E-mail addresses: olsons@uiuc.edu (S.M. Olson)8 rugreen@engin.umich.edu (R.A. Green)8 sobermei@yahoo.com (S.F. Obermeier).

0013-7952/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.enggeo.2004.07.008
236 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

2.2. Back-calculations using magnitude-bound method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243


3. Uncertainties in applying the analytical approaches. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
3.1. Factors related to liquefaction susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
3.1.1. Aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
3.1.2. Density changes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
3.2. Factors related to field observations, ground failure mechanism, and field setting. . . . . . . . . . . . . . 247
3.3. Factors related to seismicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
3.4. Validity of in-situ testing techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
4. Recommended approach to paleoliquefaction studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.1. Regional factors affecting paleoliquefaction studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.1.1. Seismological considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.1.2. Geotechnical considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
4.2. Field techniques for performing back-calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.2.1. Bracketing the properties of the source bed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.2.2. Selecting a representative penetration resistance . . . . . . . . . . . . . . . . . . . . . . . . . . 252
4.2.3. Back-calculations using observations in sectional view . . . . . . . . . . . . . . . . . . . . . . . 253
4.2.4. Back-calculations using observations in plan view . . . . . . . . . . . . . . . . . . . . . . . . . 256
5. Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258

1. Introduction liquefaction susceptibility of the source beds (Olson


et al., 2001) and the uncertainty inherent in the
Investigators are increasingly using studies of procedure. Still, this method has been used exten-
paleoliquefaction features to estimate the character- sively for the analysis of paleoliquefaction effects,
istics of ground motions of prehistoric earthquakes, such as those from the 1811–1812 New Madrid
even for events that occurred far back in Holocene earthquakes (e.g., Schneider, 1999) and the 1755
time. Several analytical methods are available for Cape Ann, MA earthquake (Ellis and de Alba,
geotechnical analysis, but only two have been used 1999). Similarly, in the Pacific Northwest, Obermeier
extensively. Major reliance has been placed on one and Dickenson (2000) used the regional pattern of
of these, the cyclic stress method (i.e., the method liquefaction effects in combination with the cyclic
originally devised by Seed and Idriss, 1971 and stress method to estimate the strength of shaking
Whitman, 1971), which, when applied regionally, from the subduction earthquake of 1700 AD.
can be used to estimate the combination of peak Virtually all major paleoseismic studies that have
ground acceleration (pga) and earthquake magnitude attempted to back-calculate the value of pga (peak
(M) required to induce liquefaction at a site. The ground acceleration) at a site have been based either
other principal technique, which we refer to as the directly or indirectly on the cyclic stress method
magnitude-bound method, uses the range of lique- (including those that use the Ishihara, 1985 and the
faction effects (i.e., the most distal site of liquefac- Pond, 1996 energy–stress methods).
tion from the energy center) to estimate the value of In several important studies, paleoseismic inter-
M. Although other methods are available, generally pretations have been based largely on the magnitude-
they are considered less established. bound method. For example, the range of liquefac-
In studies that have been based on the cyclic tion effects predating the 1886 Charleston, SC
stress method, paleoseismic interpretations often are earthquake (M~7.2) was used in combination with
subject to many uncertainties. These uncertainties the sizes of liquefaction features to demonstrate that
include the unknown influence of aging effects on previous events had been at least as strong as that of
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 237

1886 (Obermeier, 1993, 1996; Talwani and uncertainties. Finally, we recommend a regional
Schaeffer, 2001). In the Wabash Valley of Indiana– approach for integrating and interpreting the results
Illinois, Munson and Munson (1996), Pond (1996), from back-calculations of individual sites.
and Obermeier (1998) have used this method to
demonstrate that some of the prehistoric earthquakes
there almost certainly exceeded M 7. In the New 2. Analytical approaches to paleoliquefaction
Madrid Seismic Zone of Missouri–Arkansas, Tuttle studies
(1999) used the sizes and range of liquefaction
features to show that some late Holocene earth- As stated previously, two categories of analytical
quakes probably approached the strongest of the approaches are commonly used in paleoliquefaction
1811–1812 events, of M ~ 7.5 to 8. Still, the studies to estimate seismic parameters. The first
magnitude-bound method is limited by the need for category involves using geotechnical procedures to
calibration from historic earthquakes in the same determine earthquake ground motions required to
tectonic setting. induce liquefaction. The second category is referred
Obermeier et al. (2001) recently summarized the to as the magnitude-bound method and entails the use
geologic and field methods for paleoseismic inter- of empirical correlations relating earthquake magni-
pretations as well as the principal analytical methods tude to the most distal sites of liquefaction. These
that are being used by the scientific and engineering approaches are outlined below.
communities-at-large. They provided an extensive
discussion of the proper conduct for the geologic 2.1. Back-calculations using liquefaction evaluation
portions of a paleoseismic study, which we believe procedures
remains valid. In this paper we attempt to clarify
and update several aspects of back-calculation Numerous liquefaction evaluation procedures
techniques, as well as analysis and interpretation have been proposed in the literature, including
procedures. stress-based, strain-based, and energy-based meth-
First, we briefly describe analytical procedures ods. In nearly all cases the motivation for their
used for back-analysis. Those discussed herein are development was to evaluate the liquefaction poten-
the cyclic stress method, the Green–Mitchell energy- tial of soils at sites subjected to design (i.e., future)
based method (Green, 2001), and the magnitude- earthquake motions. We refer to such use as
bound method. Green (2001) recently developed and bforward analysis.Q However, in addition to forward
verified a new energy-based approach for liquefac- analysis, liquefaction evaluation procedures have
tion analysis, as an alternative to the cyclic stress proven valuable for estimating the magnitude and
method. This method, summarized herein, can be associated peak ground acceleration at sites of
used for paleoseismic analysis to evaluate the values liquefaction for pre-instrumental earthquakes (i.e.,
of pga–M combinations that are required to induce bback-analysisQ).
liquefaction at a site. The method eliminates some We present below brief overviews of two lique-
of the uncertainties inherent to the cyclic stress faction evaluation procedures, as they were developed
method. for performing forward analyses. The first is the
Next, we present recommendations and proce- widely used cyclic stress method, while the second is
dures that reduce uncertainties inherent in the a newly developed energy-based procedure.
collection and application of field data for back-
calculations. We address seismological and geo- 2.1.1. Cyclic stress method
technical issues that affect site selection, as well as The most widely used method for evaluating
field techniques that can be used to bracket potential liquefaction is the stress-based procedure first
source bed properties and assist in selecting a proposed by Seed and Idriss (1971) and Whitman
representative penetration resistance. We recommend (1971). This empirical procedure was originally
methodologies for performing back-calculations and developed using observations of laboratory and field
interpreting results at individual sites so as to reduce data, and has been continually refined by newer
238 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

studies and by the increase in the number of thus avoiding the use of the r d factor. However,
liquefaction case histories (e.g., NRC, 1985; Seed ground response analyses require the selection or
et al., 1985; Youd and Idriss, 1997; Youd et al., synthesis of representative ground motion time
2001; Finn, 2002). The reader is referred to these histories, for which the M and pga are yet unknown
publications for more detailed information on the (i.e., the purpose of performing the paleoliquefaction
cyclic stress method. study is to determine the pga–M combination).
Additionally, we note that a large body of literature
2.1.1.1. Seismic demand. The damage potential of exists regarding the merits and shortcomings of
earthquake ground motions is a function of both the equivalent-linear and non-linear ground response
amplitude and the duration of earthquake-induced analyses and various numerical codes. Accordingly,
motions (i.e., demand), wherein the amplitude of the an investigator should have a solid understanding of
earthquake-induced demand is quantified in terms of these issues prior to using any ground response
cyclic stress ratio (CSR). CSR can be determined for program for paleoseismic analysis.
any desired depth in a soil profile by either performing The duration of ground shaking is typically
a numerical ground response analysis or by using the correlated to earthquake magnitude via magnitude
following equation: scaling factors (MSF). MSF are inversely propor-
tional to the square root of duration of strong
motion (Green and Mitchell, 2003) and are pre-
sh;avg sh;max amax rvo
CSR ¼  0:65  0:65 rd ð1Þ sented in reference to M 7.5 events. Numerous
rvo
V rvo
V g rvoV
correlations for MSF have been proposed, as shown
in Fig. 1 (Seed and Idriss, 1982; Ambraseys, 1988;
where s h,avg = average (equivalent) horizontal seis- Arango, 1996; Andrus and Stokoe, 1997; Youd and
mic shear stress c 0.65s h,max; r vo V = initial vertical Noble, 1997). Youd and Idriss (1997) and Youd et
effective stress; a max = peak ground acceleration al. (2001) describe the development of these
(pga); g = acceleration of gravity; r vo = initial relationships. As may be observed from this figure,
vertical total stress; and r d = dimensionless stress the MSF vary greatly at all magnitudes, and
reduction factor accounting for flexibility of the particularly at magnitudes less than about 6.5.
soil column. At present, no clear consensus exists among the
Eq. (1) constitutes the bsimplified,Q or approx- geotechnical earthquake engineering community as to
imate, procedure to estimate the amplitude of earth- which of the proposed MSF relations is most correct.
quake-induced demand. The r d factor exhibits a wide However, the MSF relations proposed by Seed and
range of values particularly at depths greater than 10 Idriss (1982) and by Andrus and Stokoe (1997) are
m, and the average value of the range is commonly commonly used, expressions for which are given
used in engineering practice (Youd et al., 2001). below.
However, numerous investigators suggest that the r d
factors commonly used with the cyclic stress method 6:5
MSFSeed and Idriss ¼ ð2aÞ
can be far too large at depths less than 10 to 15 m, ð M  1Þ
resulting in cyclic stress ratios that are too large (e.g.,
Nishiyama et al., 1977; Iwasaki et al., 1978; Maugeri  3:3
M
et al., 1989; Hwang, 1993; and Hwang et al., 1995). MSFAndrus and Stokoe ¼ ð2bÞ
For example, analyzing seismic data from the Lotung 7:5
and Hualien, Taiwan Large Scale Seismic Testing
(LSST) downhole array sites, Hwang et al. (1995)
showed that rd at a depth of 6 m can be as much as In addition to MSF, the effect of overburden
25% smaller than the average value recommended by pressure on liquefaction resistance is incorporated in
Youd et al. (2001). the cyclic stress method using the factor K r . K r adjusts
Alternately, values of s h,max with depth can be the liquefaction resistance of a soil to the comparable
determined by numerical ground response analyses, value at one atmosphere effective confining stress. As
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 239

Fig. 1. Magnitude scaling factors proposed by various investigators (adapted from Youd and Noble, 1997). P L is the probability of a liquefaction
occurrence.

discussed in Youd and Idriss (1997), K r is a function quake motions were categorized as liquefied and
of the relative density of the soil, as well as the initial non-liquefied, largely on the basis of the presence or
effective confining stress. Youd et al. (2001) provide absence of surficial liquefaction features (e.g., sand
relationships for K r at various values of relative boils). For each case history, the seismic demand
density. The following expression constitutes the was estimated using Eq. (3) and plotted as a function
bsimplifiedQ approach to estimate seismic demand. of the penetration resistance of the soil. The boundary
giving a reasonable separation of the liquefied and
amax rvo 1 1 non-liquefied points defines the CRR (or capacity
CSRM7:5 ¼ 0:65 rd ð3Þ
g rvo V MSF Kr curve). Fig. 2 presents one such relationship using
SPT-based case histories.
Note that while K r actually applies to liquefac- Following an analogous procedure, Stark and
tion resistance, it may be used to adjust the seismic Olson (1995) developed a CRR boundary using cone
demand, as above. Additionally, it should be noted penetration test (CPT) data. Similarly, Andrus and
that Eq. (3) applies only to free-field, level ground Stokoe (1997) developed a CRR boundary using
sites (i.e., slopes less than 6%). small strain shear wave velocity data.
The factor of safety against liquefaction (FSliq) is
2.1.1.2. Liquefaction resistance (capacity). In the defined as the ratio of liquefaction resistance (or
cyclic stress method the capacity of the soil (or capacity) to seismic demand.
liquefaction resistance) is quantified in terms of
cyclic resistance ratio (CRR). Empirical correlations
capacity
relating CRR to in-situ properties [e.g., SPT (N 1)60, FS liq ¼ ð4Þ
demand
CPT q T1, or shear wave velocity V s1] were
developed through the analysis of earthquake case
histories. Sites containing sandy soils that were Liquefaction is predicted when FSliq is less than or
subjected to known (or reasonably estimated) earth- equal to unity.
240 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

Fig. 2. Cyclic resistance ratio (CRR) curve. The case history data (raw) are from Ambraseys (1988) and were reduced by Green (2001)
according to Youd et al. (2001). Normalized SPT blowcount, (N 1)60cs, is adjusted for fines content of soil according to Youd et al. (2001).

2.1.2. Overview of Green and Mitchell’s energy-based by performing a numerical ground response analysis or
method by using the following bsimplifiedQ expression.
The Green–Mitchell energy-based liquefaction
evaluation procedure is a conceptual and mathemat- DW1
ical unification of the cyclic stress and strain-based NED ¼ d N eqv ð5Þ
rmo
V
methods (e.g., Dobry et al., 1982), with the capacity
curve being empirically derived using field case where DW 1 = the energy dissipated in one equivalent
histories supplemented with laboratory data (Green, cycle of loading; r moV = the initial mean effective
2001). confining stress; and N eqv = the number of equivalent
cycles. Using the principles of visco-elasticity, as
2.1.2.1. Seismic demand. In this procedure the outlined in Green (2001), Eq. (5) can be rewritten as:
seismic demand imposed on the soil is quantified in  2
terms of normalized energy demand (NED), which is 2pDc amax
NED ¼   d 0:65 rvo rd d N eqv
the area bound by the shear stress–strain hysteresis rV d G d G g
mo max Gmax c
loops (DW), divided by the initial mean effective
confining stress (r mo
V ). NED can be determined either ð6Þ
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 241

where the previously undefined terms are: D c = vis-


cous damping ratio at the shear strain c; G max = secant
shear modulus corresponding to c V 104%; and ( G/
G max)c = ratio of secant shear moduli corresponding to
c and c V 104%, respectively.
Because both D and G/G max are functions of shear
strain, the determination of the earthquake-induced
shear strain is central to using Eq. (6). Green (2001)
adopted the iterative approach proposed by Dobry et
al. (1982) to determine strain compatible values of D Fig. 4. Determination of shear modulus and damping ratios from the
and G/G max. Starting with the relation between stress respective degradation curves.
and strain c = s/G, Dobry et al. (1982) derived the
following expression: The remaining unknowns in Eq. (6) are G max and
N eqv. Various correlations have been proposed for
0:65d amax
g d rvod r d G max, one of which is that given in Seed et al. (1986)
c¼   ð7Þ relating G max and (N 1)60:
Gmax d GGmax
c !0:5
 1=3 rmo
V
Using (secant) shear modulus degradation curves Gmax ¼ 440 ðN1 Þ60 Pa1 ð8Þ
Pa2
such as those proposed by Ishibashi and Zhang (1993),
Eq. (7) is solved iteratively as illustrated in Fig. 3. For where Pa1 and Pa2 are atmospheric pressure having
the first iteration, a value of G/G max is assumed and c the same units as G max and r mo
V , respectively. For these
is then computed. In the second iteration the ratio of calculations, Green (2001) estimated r moV using empiri-
G/G max corresponding to the value of c that was cal correlations relating penetration resistance to the
computed in the first iteration is used. The process is ratio of horizontal to vertical effective stress (K o).
repeated until the assumed and computed ratios are However, K o may be assumed reasonably to be 0.5.
within a tolerable error. The shear strains determined Finally, Green (2001) analyzed data from a series of
by this procedure are analogous to those computed ground response analyses and developed a correlation
using btotal stressQ ground response computer pro- relating N eqv to earthquake moment magnitude (M) and
grams (e.g., SHAKE’91). Accordingly, the reduction epicentral distance (R e). Fig. 5 presents the correlation.
of the secant shear modulus of the soil relates only to
the inherent non-linear behavior of the soil at a given
effective confining stress and does not reflect or
include the influence of shaking-induced excess pore
pressures. Once c is determined, the damping and
shear modulus ratios are easily determined from the
respective degradation curves, as illustrated in Fig. 4.

Fig. 3. Iterative solution of Eq. (7) to determine the effective shear Fig. 5. Contour plot of constant N eqv as a function of epicentral
strain g at a given depth in a soil profile. distance (R e) and earthquake magnitude (M) (from Green, 2001).
242 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

2.1.2.2. Liquefaction resistance (capacity). Using an As with the cyclic stress liquefaction evaluation
analogous approach to that used to develop the procedure, liquefaction is predicted where demand
correlation relating CRR and (N 1)60 shown in Fig. equals or exceeds capacity.
2, Green (2001) developed a capacity curve for the
energy-based procedure via the analysis of liquefac- 2.1.3. Brief discussion of the cyclic stress and energy-
tion case histories. Sites containing sandy soils that based methods
were subjected to earthquake motions were catego- The cyclic stress and the Green–Mitchell energy-
rized as liquefied and non-liquefied. For each of the based liquefaction evaluation methods are not com-
case histories, the seismic demand was estimated pletely independent, but rather have commonalities.
using Eq. (6) and plotted as a function of the Fig. 7 illustrates this using a shear stress–strain
normalized penetration resistance of the soil. Fig. 6 hysteretic loop. Assuming the loop shown in this
presents the normalized energy capacity (NEC) figure represents an equivalent cycle, the cyclic stress
boundary that gives a reasonable separation of the method defines the seismic demand in terms of
liquefied and non-liquefied data. CSR = s h,avg /r vo
V (i.e., Eq. (1)). Similarly, the

Fig. 6. Energy-based capacity curve developed using 126 liquefaction field case histories (Green, 2001).
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 243

where R = hypocentral distance in km. This expression


relates to the elastic energy of the seismic waves
arriving at the site, rather than energy dissipated
through friction as a result of the interparticle move-
ment of the sand grains. The Green–Mitchell procedure
uses the latter definition of energy (i.e., dissipated
energy) because it relates directly to the breakdown of
soil structure and therefore is closely related to the
liquefaction phenomenon. Green (2001) presented a
detailed critique of Law et al. (1990) and other energy-
Fig. 7. Graphical illustration of the inter-relationship of the cyclic based procedures.
stress and the Green–Mitchell energy-based liquefaction evaluation
procedures (Green, 2001).
2.2. Back-calculations using magnitude-bound
Green–Mitchell energy-based method defines the method
seismic demand in terms of the area bound by the
hysteretic loop, or DW 1 (i.e., Eq. (5)). The magnitude-bound method allows an inves-
As indicated by Green (2001), one advantage that tigator to estimate the magnitudes of paleoearth-
the Green–Mitchell energy-based method has over the quakes using empirical, regionally dependent
cyclic stress method is that it circumvents the need for correlations relating M to the site-to-source distance
MSF and K r , both of which introduce considerable of the most distal liquefaction feature (R max). Fig. 8
scatter (or uncertainty) into the cyclic stress approach. presents several such correlations for a variety of
Although the Green–Mitchell energy-based proce- geographic and tectonic settings (Kuribayashi and
dure relies on a correlation relating M, site-to-source Tatsuoka, 1975; Youd and Perkins, 1978; Keefer,
distance, and N eqv (Fig. 5), which may be viewed as 1984; Ambraseys, 1988; Carter and Seed, 1988;
being analogous to MSF, the approaches used to derive Papadopoulos and Lefkopolous, 1993; Obermeier et
the correlation shown in Fig. 5 and MSF are al., 1993; Pond and Martin, 1996; Galli, 2000. This
fundamentally different, with the former resulting in figure includes correlations that define the site-to-
considerably less scatter. Green (2001) derived the source distance in terms of epicentral distance (Fig.
correlation in Fig. 5 from numerical ground response 8a) and correlations that define site-to-source dis-
analyses using rock outcrop motions for a range of tance in terms of closest distance to fault or distance
earthquake motions that were recorded at varying site- from energy center (Fig. 8b). Implementing the
to-source distances, with most of the motions being magnitude-bound method for back-analysis consists
from western U.S. earthquakes having shallow focal of two steps. The first step is to determine R max for
depths. For other tectonic settings (e.g., subduction the paleoearthquake. The second is to use a regional
zones), correlations relating M, site-to-source distance, correlation to relate R max to M.
and N eqv can be developed from representative accel- As discussed by Obermeier et al. (2001), back-
eration time histories, either recorded or synthetic. calculations using the magnitude-bound method
Finally, the Green–Mitchell procedure is only one of should be based on regional correlations, as opposed
a plethora of energy-based liquefaction evaluation to correlations derived from worldwide data, because
procedures that have been proposed in literature, with the factors that control the greatest distance from the
the procedure proposed by Law et al. (1990) having earthquake source at which liquefaction occurs are
been used in several paleoliquefaction studies (e.g., regionally dependent. These factors include: (1)
Pond and Martin, 1996). Law et al. (1990) defined the earthquake source characteristics; (2) transmission
seismic demand (T) imposed on the soil using the characteristics, and (3) regional soil liquefaction
following expression. susceptibility. For example, Obermeier et al. (1993)
and Pond (1996) showed that for the central United
101:5M States the moderate liquefaction susceptibility of the
T¼ ð9Þ
R4:3 regional sediments limits the farthest distance to
244 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

Fig. 8. (a) Magnitude-bound curves for varying geographic and tectonic settings. Note that site-to-source distance is quantified in terms of
epicentral distance. (b) Magnitude-bound curves for varying geographic and tectonic settings. Note that the site-to-source distance is quantified
in terms of closest distance to fault, except for the curve by Obermeier et al. and Pond and Martin, which is distance from the energy center
Obermeier et al. (2001). That curve is for the central U.S. and is based on using a value of M ~ 7.6–7.7 for the largest of the 1811–1812 New
Madrid earthquakes.

liquefaction for moderate earthquakes (M b 6.8), as Subsequent sections of this paper expand on their
compared to Ambraseys (1988) worldwide bound. In discussion and the methods used to locate the most
contrast, for large earthquakes (M N 7), the regional distant site of liquefaction.
transmission characteristics of the central and eastern
U.S. (CEUS) extend the farthest distance to liquefac-
tion beyond Ambraseys’ bound. 3. Uncertainties in applying the analytical
Thus, when using the magnitude-bound method for approaches
paleoseismic back-analysis, investigators must con-
sider several issues. These issues include: (1) deter- Despite the apparent simplicity of the above
mination of energy center for paleoearthquakes; (2) approaches to paleoliquefaction back-analysis, there
accounting for regional source and transmission are numerous uncertainties in their application. These
characteristics, as well as liquefaction susceptibility; uncertainties can be categorized as: (1) factors related
and (3) the effort and methods used in the field to liquefaction susceptibility; (2) factors related to field
investigation to locate the most distant site of observations, as well as the ground failure mechanism
liquefaction. Obermeier et al. (2001) present a detailed and the field setting; (3) factors related to seismicity;
discussion of the first two factors noted above. and (4) the validity of in-situ testing techniques.
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 245

3.1. Factors related to liquefaction susceptibility secondary compression effects are virtually complete
within a few hundred years or so (e.g., see laboratory
The occurrence of liquefaction can cause drastic data in Mesri et al., 1990) for these sandy soils.
changes in properties that influence the subsequent Still, with the passage of thousands of years there
liquefaction susceptibility of a deposit. These is a further substantial decrease in liquefaction
changes, discussed in detail by Olson et al. (2001), susceptibility in many field settings (e.g., Youd
relate mainly to changes in properties often associated and Perkins, 1978, Table 2). For example, Youd and
with passage of time (i.e., bagingQ), and to changes in Perkins indicate in their table that river channel
sediment density resulting from liquefaction. We deposits will likely decrease in liquefaction suscept-
summarize below the influence of these two changes ibility from bvery highQ to a value of bhigh,Q
in order to provide an understanding of the field through Holocene time. Similarly, dune deposits
procedures that we propose later. We do not consider will decrease from bhighQ to a value of bmoderate.Q
factors such as grain size and shape because they do As suggested by Terzaghi et al. (1996), we suspect
not change as a result of liquefaction, and presumably that a portion of this increase in liquefaction
their effects are reflected in the field measurement of resistance is in response to preshearing resulting
penetration resistance. from applications of low-level seismic shaking (i.e.,
too low to have caused significant porewater
3.1.1. Aging pressure increase or liquefaction).
Natural and man-made deposits develop a structure The authors are not aware of any field data that
through time that results in soil properties such as prove the effect of preshearing occurs in-situ, but
increased shear strength, modulus, and penetration there are ample laboratory data demonstrating this
resistance (Schmertmann, 1991). This vaguely de- phenomenon (e.g., Finn et al., 1970; Bjerrum, 1973;
fined process is termed baging.Q The effects of aging Lee and Focht, 1975; Seed et al., 1977; Vaid et al.,
are attributed to several sources, primarily mechanical 1989). For example, Seed et al. (1977) used large-
and chemical (i.e., cementation). Mechanical sources scale shaking tests to examine the effect of preshear-
include the minor readjustment of grains resulting ing on the liquefaction resistance of Monterey 0 sand.
from secondary compression and preshearing. Secon- They showed that preshearing under low-level shak-
dary compression is the process of minor grain ing increased the liquefaction resistance by more than
adjustments into a more stable configuration under 40% at 15 cycles of shaking, while the corresponding
constant vertical effective stress (Mesri and God- increase in relative density was only 1% (from 54% to
lewski, 1977; Mesri et al., 1990). We define preshear- 54.7%). Similarly, Vaid et al. (1989) tested Ottawa C-
ing as the minor adjustment of grains into a more 109 sand at an initial relative density of 36% to
stable configuration due to macroscopic shear examine preshearing effects. Vaid et al. showed that
stresses, which can be either transient or sustained the number of shaking cycles (at s/r 1cV = 0.125; where
and either cyclic or monotonic. r 1cV is the major principal effective stress after
Mesri et al. (1990) indicated that freshly deposited consolidation) required to trigger liquefaction
and/or densified clean sands exhibit substantial increased from 14 with no preshearing, to 38 with
increases in penetration resistance during drained 0.1% axial strain due to preshearing, and to 125 with
secondary compression, especially in the first several 0.2% axial strain due to preshearing.
months (i.e., first few log cycles of time following the As these laboratory tests suggest, preshearing
end of primary consolidation). After that time the rate would produce nearly imperceptible settlements at
of increase decreased greatly. Doubling and tripling of the ground surface. However, these slight movements
penetration resistance occasionally was noted during at the particle contacts produce better interlocking
this initial period. We suspect that it is because of this among grains and a significant increase in liquefaction
rapid recovery in penetration resistance that the resistance. Therefore, in the time range of more than a
relations in Fig. 2 (collected within a few years after few hundred years, and where cementation is not
the causative earthquake) are at least approximately present, the net effect of preshearing and secondary
correct for engineering design. And, we believe that compression is most typically an increase in pene-
246 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

tration resistance and correspondingly a decrease in


liquefaction susceptibility.
The contribution of cementation to aging un-
doubtedly increases liquefaction resistance in many
field settings, especially where the level of the
watertable fluctuates. This fluctuation can permit
precipitation of compounds such as carbonates and
the development of oxides through the zone above the
watertable. However, this process is typically impor-
tant only above the lowest level of the watertable.
The actual changes in liquefaction susceptibility
that take place through time can vary greatly from
geographic region to region. For example, in coastal
South Carolina, Martin and Clough (1994) found that
the liquefaction susceptibility of many beach deposits
older than 80,000 years remains high to very high; the Fig. 9. Ratio of post-densification q c value (measured 1 to 30 days
reasons for the persistence of this exceptionally high after densification) to pre-densification q c value for different areas
and densification procedures (from Mesri et al., 1990).
susceptibility are uncertain. Thus, for paleoseismic
analysis, it is apparent that the effects of aging can be lower than the pre-disturbance value (with the value
either large or very small, and must be evaluated on a of one on the ordinate being the pre-disturbance
case-by-case basis. value), depending on the percentage change in void
ratio, De R. (The value De R can be shown to be
3.1.2. Density changes equivalent to the percentage change in relative
The occurrence of liquefaction in a sediment layer density, DD r). Most sites that experienced a large
results in an increase in density due to dissipation of change in relative density (i.e., significant settlements
shaking-induced porewater pressures and reconsoli- due to disturbance) exhibited a post-disturbance
dation (excluding a thin zone immediately below a penetration resistance greater than the pre-disturbance
fine-grained cap). However, this increase in density value. In contrast, all sites shown in the figure that
may or may not result in an increase in liquefaction experienced smaller changes in relative density
resistance. The reason for a potential decrease in exhibited a post-disturbance penetration resistance
liquefaction resistance following liquefaction is the that was smaller than the pre-disturbance value. Olson
destruction of the pre-earthquake (aged) soil structure, et al. (2001) discussed this concept in detail.
which had developed through mechanisms such as While the magnitude of the change in penetration
secondary compression, preshearing, and cementa- resistance resulting from ground modification may
tion. Terzaghi et al. (1996) and Oda et al. (2001) have differ from that resulting from liquefaction, the
suggested that destruction of the aged soil structure by pattern of changes observed by Mesri et al. (1990)
large cyclic shearing and liquefaction may substan- suggests that post-earthquake penetration resistance
tially reduce post-shaking liquefaction resistance. can be higher or lower than the pre-earthquake value,
In the following discussion, we use changes in depending on the magnitude of density change. In
penetration resistance from ground modification pro- turn, the magnitude of density change strongly
grams to illustrate changes in liquefaction resistance depends on the severity of liquefaction and duration
that can occur. Mesri et al. (1990) compiled data that of strong shaking. Whereas an increase in post-
demonstrate the potential changes in penetration earthquake penetration resistance can occur at sites of
resistance from before and after ground modification severe liquefaction, a temporary post-earthquake
(see Fig. 9; Solymar, 1984; Dowding and Hryciw, decrease can occur at sites of marginal or moderate
1986). Fig. 9 illustrates that immediately following liquefaction. However, we believe that it is likely that
modification (i.e., less than 30 days) the post- only minor changes in penetration resistance from
disturbance penetration resistance may be higher or the pre-earthquake values should remain at many
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 247

(and likely most) sites of marginal liquefaction by data presented in Figs. 2 and 6 were determined
the time in situ measurements are made. Our almost exclusively from sediment samples taken from
reasoning is that at sites of marginal liquefaction, borings, in conjunction with judgment. In many cases,
reconsolidation will be completed quickly (on the only the most susceptible strata could confidently
order of hours to days) resulting in secondary have been assumed to liquefy (e.g., Fear and
compression starting shortly after the earthquake. In McRoberts, 1995).
this case, the post-earthquake penetration resistance Still, considering all factors that were used in
will recover very quickly. development of the boundary curves in Figs. 2, 6,
We note that the above discussion applies to clean and 8), we believe that for back-analysis of paleo-
sands. Density changes in silty sands and sandy silts seismic shaking the curves should be used in their
resulting from ground modification or liquefaction are present versions. Only with extensive further research
less predictable (e.g., Green, 2001). can the curves be modified with confidence, because
different factors related to the observed field occur-
3.2. Factors related to field observations, ground rence of liquefaction cause the curves to be shifted in
failure mechanism, and field setting different directions.

The data used to develop Figs. 2, 6, and 8 are 3.3. Factors related to seismicity
exclusively from field observations made in plan
view. Investigators typically have designated sites as At some of the sites corresponding to the data in
bliquefiedQ on the basis of the presence of surficial Figs. 2 and 6, the value of pga was not well
liquefaction features, such as venting of sediment to constrained, even for recent earthquakes (e.g., Bou-
the surface (i.e., sand boils), ground cracking asso- langer et al., 1997). Furthermore, it is widely known
ciated with liquefaction (e.g., lateral spreading), or that the level of shaking can vary greatly in a short
surface settlements. Other evidence of liquefaction distance horizontally (e.g., Idriss and Seed, 1968;
included tilting or settling of overlying structures and Stone et al., 1987; Seed et al., 1990). This factor could
floating of underground structures. lead to substantial errors in the estimated CSR and
The data in Figs. 2, 6, and 8 incorporate all NED for some of the data presented in Figs. 2 and 6,
mechanisms of ground failure–hydraulic fracturing, respectively.
lateral spreading, and surface oscillations. However, In addition to the uncertainty associated with
the ground failure mechanism may control whether or estimating the amplitude of the demand, estimations
not surface manifestations develop for some field of ground motion duration, as represented in terms of
settings and for some intensities of earthquake shaking bequivalent number of cyclesQ (i.e., MSF in the cyclic
(examples are discussed in detail by Obermeier et al., stress method) also may be highly uncertain (e.g., see
2001). For example, liquefaction features at level sites Fig. 1). This is evident from the large scatter in the
(e.g., hydraulic fracturing) may not manifest at the data presented in Seed et al. (1975) relating earth-
surface where the fine-grained cap is relatively thick or quake magnitude to equivalent number of cycles and
where the watertable is located below the top of the in the range of proposed MSF from various studies
stratum that liquefied. Accordingly, it is likely that (e.g., Youd and Idriss, 1997). Additionally, there is
some of the data points in Figs. 2 and 6 that are uncertainty in applying MSF that were developed
designated as sites of bno liquefactionQ did, in fact, from a compilation of ground motions from western
experience liquefaction. Similarly, liquefaction likely U.S. and Japanese earthquakes to other tectonic
occurs beyond the farthest site-to-source distance that settings, such as the CEUS.
is used for the magnitude-bound method (Fig. 8), but
such occurrences are not discernable in plan view. 3.4. Validity of in-situ testing techniques
Another major factor contributing to a non-distinct
transition from sites of liquefaction to those of no The validity of paleoseismic analysis using lique-
liquefaction in Figs. 2 and 6 is caused by the inability faction sites relies on there being measurable differ-
to determine which stratum actually liquefied. The ences in penetration resistance between sites that
248 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

experienced liquefaction and adjacent sites that did not. the sphere of influence can be up to 20 diameters in
However, in some settings such as that in the coastal stiff soils. However, because of the influence of the
plain of South Carolina, there are often little or no length of the sampler in contact with the soil, the
measurable differences in penetration resistance effective diameter controlling the sphere of influence
between sites with liquefaction features and adjacent may be greater than the actual diameter of the
sites with no liquefaction features (Talwani, 2001). For sampling spoon. In other words, in soft soils the
sites such as these, back-analysis techniques using SPT N-value is influenced by a sphere of soil with a
penetration tests have limited validity. diameter ranging anywhere between 11 cm (twice the
We suggest that a possible explanation for the lack actual diameter of the spoon) and 1.8 m (3 the
of measurable difference in penetration resistance at the length of the sampler).
South Carolina sites may be the presence of weak Measurement of shear wave velocity (V s) by cross-
cementation or weak aging effects in the source sand hole or downhole methods is limited by practical
beds. Our rationale is as follows. In-situ penetration considerations of borehole conditions (i.e., casing/
testing methods such as the SPT and CPT induce very grout/soil continuity, stiffness compatibility of grout
large strains in the soil prior to passage of the and soil, borehole verticality, etc.), spacing of measure-
equipment in the ground. As a result, at sites of no ments, and spacing and number of boreholes used for
liquefaction, weak bonds at particle contacts that measurements (in the crosshole method). For example,
remained intact during seismic events may be disrupted as the spacing of boreholes increases in the crosshole
by the large strains that precede the penetrating method, the apparent shear wave velocity in soft soils
equipment. However, as a result of destroying the increases considerably (e.g., ASTM D 4428/D
weak effects, there may be little difference in pene- 4428M). However, the accuracy and resolution of
tration resistance at a site of liquefaction and an crosshole tests is generally constant with depth, with a
adjacent site of no liquefaction. Non-destructive shear highest resolution of about 60 cm. Downhole measure-
wave velocity testing may provide a means to assess the ments (including those made using the seismic cone
presence of cementation at such sites. However, shear penetration test) are subject to considerable interpreta-
wave velocity should not be used for back-analysis at tion, and these interpretations are greatly complicated
these sites because liquefaction resistance relationships by multilayered subsurface profiles (e.g., Lodge El-
using shear wave velocity data were developed using Telbany et al., 1996). For example, using downhole
case history data almost exclusively from sites having methods it may be possible to resolve a 1-m-thick layer
young, uncemented sandy soils (Andrus and Stokoe, that exhibits a 50% shear wave velocity contrast to the
1997). In such locales, back-calculations can only be surrounding layers, but it may not be possible to resolve
made at liquefaction sites (i.e., sites of no liquefaction a 3-m-thick layer with a 20% velocity contrast (King,
cannot be used to provide upper limits), resulting in 2002). Spectral analysis of surface waves (SASW)
greater uncertainty in back-analysis results. eliminates some of the practical issues involved with
Lastly, it is not unusual that marginal liquefaction borehole measurements, but getting reliable results
effects are the result of liquefaction of a thin stratum. depends on procedural details and requires expert
(We provide a definition of marginal liquefaction in operation and interpretation. In addition, SASW
Section 4.) Thus, one must consider whether the tools resolution decreases significantly with depth and has
that are being used for assessment of liquefaction a finest resolution of about 0.3 to 1 m. Suspension
susceptibility, especially the SPT and shear wave logging provides very high resolution (as fine as 20 to
velocity methods, can accurately assess the source bed 30 cm, at least in soil deposits), is performed in a single
properties. The sampling spoon for the SPT typically borehole, and can be conducted to great depths (up to
is about 60 cm long, which means that as much as 60 about 600 m)—offering many advantages over other
cm of soil are in contact with the sampler and directly methods. It should be noted that values of V s measured
influence the SPT value. In addition, Robertson and using suspension logging and the crosshole method
Wride (1997) suggest that a sphere of soil up to two or may differ by up to about 30% over a small depth range
three times the diameter of the SPT sampler can at a single site (Owen and Roblee, 2000). Differences
influence penetration resistance in soft soils, whereas with other methods may be greater. In practice, the
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 249

smallest discernible layer thickness is generally con- At the bplanningQ level, we subsequently discuss
sidered to be about 1 m, and the thinnest layer identified regional factors that affect site selection, data inter-
in the V s-based liquefaction database is 1.2 m (Andrus pretation, and details of back-analysis that must be
and Stokoe, 1997). considered prior to the field conduct of a paleolique-
Clearly the CPT method offers advantages for faction study. We divide these regional factors into
analysis of marginal liquefaction sites, although the seismological and geotechnical issues.
CPT also has limitations for very thin layers, both soft Our recommended approach for collecting field
and stiff (Lunne et al., 1997; Robertson and Wride, data for back-calculations is primarily to use sites of
1997; Vreugdenhil et al., 1994). Robertson and Wride marginal liquefaction that are distributed over a wide
(1997) suggested that the minimum layer thickness for area to the extent possible. A site of marginal
which the CPT can provide an accurate penetration liquefaction typically only has minor liquefaction-
resistance is approximately 7 to 11 cm in soft soils. The induced effects, which can be manifested in a
minimum thickness increases considerably in stiff number of ways depending on the method of
soils, to as much as 70 cm. At marginal liquefaction observation. In plan view (e.g., looking down from
sites, potentially liquefiable soils often are loose to above at a feature on the ground surface), there can
medium dense, suggesting that the minimum identifi- be sparse and/or small sand boils, minor settlements
able layer thickness is somewhere between 7 and 70 (less than several centimeters), minor lateral spread-
cm, perhaps in the 20 to 30 cm range. ing (on the order of centimeters, not tens of
centimeters or meters), or minor ground cracking
that could be associated with liquefaction. In sec-
4. Recommended approach to paleoliquefaction tional view (e.g., looking at the cross-section of a
studies feature exposed in a river bank), there can be
development of small sills along the base of an
In order to address and reduce many of the overlying fine-grained cap or small (i.e., thin) dikes
uncertainties discussed above, we have developed a that may or may not have extended to the ground
new, systematic approach for the conduct of paleo- surface at the time of the earthquake. Sites of more
liquefaction studies. The approach is outlined as severe liquefaction and no liquefaction also should
follows. be used, but the field effort and interpretation should
be focused on sites of marginal liquefaction.
(1) Plan field work. By using the field data, back-calculated combina-
(2) Perform field work. tions of M and pga at individual sites are then
(3) Estimate energy center of the paleoearthquake. compared with appropriate regional acceleration
(4) Use regional magnitude-bound relationship and attenuation models in order to estimate the actual
R max to estimate magnitude. value of earthquake magnitude. This procedure was
(5) Perform back-calculations using liquefaction first used by Pond (1996) for the Wabash Valley of
evaluation procedures at individual sites to Indiana–Illinois to demonstrate that M ~ 7.5 earth-
estimate likely combination of pga and M. quakes had struck the region in mid-Holocene time. In
(6) Integrate back-calculations at individual sites the following sections, we present our recommended
into a regional assessment to better assess the approach, which greatly expands upon and refines the
magnitude of the paleoearthquake. technique first used by Pond (1996).

Obermeier et al. (2001) discussed items 2 through 4.1. Regional factors affecting paleoliquefaction
4 in detail, although additional thoughts regarding the studies
magnitude-bound method are described above. In the
balance of this paper, we focus on the remaining tasks 4.1.1. Seismological considerations
of the proposed approach. In a companion paper Paleoliquefaction studies involve matching the
(Green et al., 2004), we illustrate steps 5 and 6 using back-calculated strength of shaking from individual
case history examples. sites with regional acceleration attenuation models, as
250 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

we noted above. Thus a representative or a suite of so, the slopes of the capacity curves become steep.
representative acceleration attenuation models must Thus a small error in selecting the representative
be selected. The selected attenuation model(s) should (N 1)60 value causes a large error in the back-calculated
be based on appropriate characterization of the value of pga. Therefore, to the extent possible, one
postulated earthquake source and regional transmis- must be very careful in making interpretations in the
sion characteristics. However, at present there is little meizoseismal zone of a large earthquake. In addition,
data on which to base attenuation models for larger very little data exist to substantiate the liquefaction
magnitude earthquakes (i.e., greater than M ~ 6) for resistance of very loose sands (i.e., (N 1)60 less than
the central and eastern U.S. (CEUS). For example, the about 3 or 4). Youd and Idriss (1997) suggested that
2002 National Seismic Hazard Maps (Frankel et al., liquefaction resistance is nearly constant in this very
2002) for the CEUS are based on five attenuation low penetration resistance range, while other lique-
models because each model bbrings novel and faction resistance relationships suggest that liquefac-
important aspects to the problem.Q tion resistance continues to decrease with decreasing
It may be necessary in some seismological settings penetration resistance. This difference results in
to consider the possibility of significant focusing of considerable uncertainty in back-calculations at sites
seismic energy within the bedrock, especially for with very loose sands.
larger magnitude earthquakes (i.e., greater than Another factor that may significantly influence
M ~ 7). For example, it is well known that strike–slip back-calculation results is amplification (or deampli-
faulting can cause the strength of shaking to be much fication) of bedrock motions in alluvial deposits.
higher along the strike of the fault than elsewhere; for Predicting ground motions in very thick alluvium
thrust faulting, the upthrown block will likely have the (e.g., many hundreds of meters) can lead to significant
stronger shaking. Selected attenuation models for errors, as is well known. Furthermore, selecting (and
back-calculation should address these issues. Further- modifying) existing records or generating synthetic
more, locales for paleoliquefaction searches can be time histories for use in ground response analyses
selected on the basis of this information in order to involves considerable uncertainty, particularly for
improve the chances of encountering candidate sites regions like the central U.S. where strong ground
for back-calculation. motions have not been measured for earthquakes
In addition, an investigator must be aware that exceeding about M 5. Experienced seismologists
attenuation models, which typically are based on should participate in the selection and development
small historic earthquakes in the region (at least for of acceleration time histories if a ground response
regions like the CEUS where large earthquakes occur analysis is undertaken. Thus, to the extent possible,
infrequently), show only the largest values of peak site selection and back-calculations should be done so
acceleration that are predicted for larger magnitude as to minimize the unknowns associated with the
earthquakes. Furthermore, the nature of shaking is ground response analysis. Where possible, one
likely to be much more random within the meizo- approach is to conduct geotechnical testing at some
seismal zone, in comparison to beyond. Attenuation sites having thick alluvium, and some at nearby sites
models are unable to account for this random having thin alluvium.
variability at small site-to-source distances. As a result of these seismological and geotechnical
uncertainties, paleoliquefaction sites that are selected
4.1.2. Geotechnical considerations for detailed testing should be located preferentially
Making reliable back-calculations using liquefac- throughout the region of strongest shaking beyond the
tion evaluation procedures in regions of very strong meizoseismal region (as was done by Pond, 1996).
shaking is inherently more uncertain than in regions of Selecting sites located beyond the meizoseismal region
moderate to low shaking because of the nature of can reduce many of the aforementioned uncertainties.
liquefaction susceptibility. For example, very strong Detailed testing at sites located within the meizoseis-
shaking can trigger level-ground liquefaction in mal region should involve testing at many sites. This
relatively dense sands. But, one can see readily from practice will allow an investigator to identify anom-
Figs. 2 and 6 that for (N 1)60 values exceeding 20 or alous back-calculation results.
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 251

4.2. Field techniques for performing the capacity curves in Figs. 2 and 6. This is not to say
back-calculations that only sites of marginal liquefaction lie near that
boundary, or that some sites of marginal liquefaction
Our basic approach for performing back-calcula- do not lie far from the boundary. But this assumption
tions is to use sites of marginal liquefaction to the is a reasonable first approximation for most field
extent possible, because the changes between pre- and settings and conditions. Furthermore, a regional
post-earthquake liquefaction susceptibility and pene- assessment will clarify the validity of this assumption
tration resistance are likely to be small. Again, we for individual sites.
suggest that testing be done at some sites of more For back-analysis, one generally cannot a priori
severe liquefaction as a means to provide reasonable determine which strata actually liquefied, even with
lower bound estimates of seismic parameters (see excellent exposures of possible source strata in
Green et al., 2004). sectional view. One can only observe where fluid-
The recommended approach is designed to bracket ization effects occurred. (Fluidization is the process
the influence of liquefaction on changes in the key whereby sediment is transported by flowing water.
properties of the source bed (i.e., density and aging This flowage is caused by elevated pore-water
effects) from the pre-earthquake condition. Test sites pressure in response to processes such as static
can be observed in either sectional or plan view; artesian conditions or seismic liquefaction. See
however, the method of site selection (i.e., sectional or Obermeier (1996) for a more complete discussion of
plan view) can influence the interpretation of repre- the process.) Typical effects of fluidization include
sentative penetration resistance and the level of development of dikes and sills along the base of an
uncertainty associated with the back-calculation. overlying fine-grained cap, and development of sand
Below, we provide recommendations for the inter- boils when sand is carried to the ground surface.
pretation and selection of a representative penetration However, there are field situations, especially those
resistance for various field settings. In addition, we involving marginal liquefaction, where it is difficult or
provide recommendations for the interpretation of site impossible to visibly determine which strata actually
conditions and back-analysis results for sites inves- liquefied, because the fluidization effects are so
tigated in both sectional and plan view. difficult to detect using the naked eye. Whether
In developing this approach, we presumed that in fluidization occurred within a particular stratum may
most cases the following sequence of events occurs in be discernible only by CT-scan imaging or similar
response to seismic shaking: (a) breakdown of pre- processing on undisturbed samples (e.g., Hurst and
existing aging effects; (b) liquefaction; (c) densifica- Cronin, 2001). But even if small fluidization effects
tion; and then (d) redevelopment of aging effects. Item are visible by such an analysis, it is generally difficult
(d) is time dependent and presents several possibilities, to determine whether those effects formed by pro-
which we qualitatively account for by consideration of cesses active during initial deposition, having nothing
the relative ages of the liquefaction features and those to do with seismic shaking (e.g., see Figs. 33 and 34
of the host sands. The approach applies where a site of in Obermeier, 1996).
marginal liquefaction effects is juxtaposed with a site Generally, the only means for determining whether
having no liquefaction effects, and, in addition, sedi- strata were marginally liquefied is by in-situ testing of
ments at both sites were laid down in response to the possible source beds at side-by-side sites, with one
same sedimentary processes operating essentially at the site having what appears to be marginal liquefaction
same time. That is, the site of liquefaction effects is effects and the other none. This allows the investigator
basically a lateral extension of the site having no to bracket the properties of the source bed in terms of
liquefaction effects. Employing a similar framework, factors and conditions that influence penetration
sites of more severe liquefaction can also be used. resistance and similarly influence liquefaction resist-
ance. Fortunately, in many field settings the beds
4.2.1. Bracketing the properties of the source bed having the highest likelihood of liquefying lie directly
For many field settings, a site of marginal beneath or very close to an overlying fine-grained cap.
liquefaction can reasonably be assumed to lie near This is due mainly to relations between liquefaction
252 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

susceptibility of sediments as well as to the nature of density upon liquefaction are already reflected in the
seismic shaking through an alluvial column. When penetration data. Thus, at liquefaction and no lique-
using observations in sectional view, the large contrast faction sites there is no need to compensate for
in permeability between a cap and underlying sand changes in density from the pre-earthquake condition
strata enhances development of fluidization effects, when using these relationships (and similar CPT- or
often making it relatively easy to select strata that are V s-based relationships) for back-analyses.
candidates for having liquefied. An assessment of the post-earthquake liquefaction
In summary, identifying the source bed that susceptibility at a specific site also requires consid-
liquefied (particularly at sites with only plan view eration of potential aging effects of the source bed.
observations) requires that multiple in-situ tests be Based on our previous discussion of post-earthquake
conducted in proximity to an identified liquefaction aging, we believe that it is reasonable to assume as a
feature, as illustrated in the companion paper (Green first approximation that any changes in penetration
et al., 2004) by means of using actual field data. resistance at sites of marginal liquefaction have been
recovered as a result of post-earthquake aging. Using
4.2.2. Selecting a representative penetration a regional assessment of sites will clarify the validity
resistance of this assumption for individual sites. At sites of
Selecting a representative penetration resistance for more severe liquefaction observed in sectional view,
back-analysis involves two primary issues: (1) or at any site of liquefaction observed in plan view,
accounting for changes in penetration resistance due potential changes in penetration resistance, and thus
to changes in liquefaction susceptibility (i.e., density liquefaction resistance, need to be considered on a
change and aging) at each penetration test location; case-by-case basis using the discussion presented
and (2) selecting a representative penetration resist- herein as guidance.
ance value from multiple penetration tests with proper
regard for the method of observation and the ground 4.2.2.2. Interpreting multiple penetration tests. Se-
failure mechanism. lecting a representative penetration resistance from
multiple in-situ tests at a site involves engineering and
4.2.2.1. Effect of changes in liquefaction susceptibil- geologic judgment. The investigator must consider the
ity. Both the SPT- and CPT-based liquefaction thickness and lateral continuity of the suspected source
resistance relationships (e.g., Figs. 2 and 6) employ bed, the method of observation, and the ground failure
field case history databases that were collected almost mechanism, among others factors. Some of these
exclusively after the causative earthquake, both at factors and considerations are discussed below.
sites of liquefaction and no liquefaction (Youd, 1999; The thickness of the source bed has two primary
Olson and Stark, 1998). These data typically were impacts. First, in-situ tests have limitations in their
collected some months after the earthquake, and ability to detect thin layers, as we discussed pre-
almost always within several years. The penetration viously. Where thin layers are present, the true
resistance data were not modified in any way to material properties may not be measured due to the
correct for the effects of liquefaction or earthquake influence of surrounding layers, whether softer or
shaking. However, it was the pre-earthquake sediment stiffer. Furthermore, the investigator must consider
properties that controlled whether or not liquefaction whether the suspected thin layer was capable of
occurred when the earthquake struck. The difference causing the postulated failure mechanism. For exam-
between the pre-earthquake and post-earthquake ple, a single, thin liquefiable layer may not be capable
liquefaction susceptibility can be significant in many of producing a sufficient volume of water or
cases, but is probably small in many others depending magnitude of excess porewater pressure to trigger
on the magnitude of density change (i.e., the severity hydraulic fracturing (see Green et al., 2004). On the
of liquefaction) and the influence of post-earthquake other hand, the same thin layer may be capable of
aging (Olson et al., 2001). causing lateral spreading.
Because the data used to develop Figs. 2 and 6 In-situ tests must be conducted close enough to the
were collected after the earthquake, any changes in observed feature to provide reasonable assurance that
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 253

the measured properties represent the properties of the penetration resistance values that are too large. For
suspected source bed. For example, the limited lateral example, based on the work of Bartlett and Youd
continuity of sand deposits of small, braided streams (1992), we suggest that sandy soils with (N 1)60
may require that penetration tests be conducted within values greater than about 15 to 20 generally do not
a few meters of the observed feature. produce lateral spreads beyond the meizoseismal
In general, for studies of marginal liquefaction zone.
features conducted in sectional view, we suggest that As alluded to in the previous paragraph, the ground
the minimum penetration resistance of the source failure mechanism can also influence the selection of
material [in terms of (N 1)60 or q T1, etc.] from each a representative penetration resistance value. For
penetration test be used for back-calculation. In example, the areal extent of liquefaction may be
sectional view, the size and extent of marginal much smaller to trigger hydraulic fracturing than to
liquefaction features often are readily evaluated, trigger extensive lateral spreading. In light of this
allowing penetration test locations to be selected and effect, Table 1 provides general guidelines for select-
bmarginal liquefaction/no liquefactionQ designations ing a representative penetration resistance value from
to be assigned. multiple tests.
For studies at any liquefaction site in plan view
and at sites of severe liquefaction in sectional view 4.2.3. Back-calculations using observations in sec-
(particularly sites of lateral spreading), the size and tional view
extent of liquefaction effects at depth generally As discussed previously, the capacity curves (e.g.,
cannot be readily determined from a single or few Figs. 2 and 6) are based on plan view observations.
borings. Thus multiple in-situ tests are generally Thus, the use of sectional view observations for
necessary within a very localized area to evaluate back-calculations must be consistent with the
candidates for a source bed(s). In order to make a original development of the curves. Yet we have
reasonable estimate of the lower limit of the strength noted previously that plan view observations cannot
of shaking at a site, but not an estimate of the detect some sites of marginal liquefaction; however,
extreme lower limit, we suggest using the value of these same effects would have been observed in
penetration resistance corresponding to the highest sectional view. Accordingly, one must assess
minimum liquefaction susceptibility that is common whether the liquefaction feature observed in sec-
to the data from multiple tests in proximity to the tional view would have been observed at the ground
liquefaction feature(s). In other words, we suggest surface, if the field setting had been more optimal
using the highest (or reasonably highest) minimum for showing effects of liquefaction in plan view. For
value of normalized penetration resistance common example, for a site of hydraulic fracturing, the
to multiple tests. To illustrate this method, assume following questions should be asked: bWould the
penetration tests are performed at six individual test liquefaction feature have been observed from the
locations at a site of severe liquefaction along a surface?Q or bWould small sand boils have devel-
streambank exposure several hundred meters in oped, if the cap had been thinner?Q Similarly, where
length. If the individual penetration tests yield the observed paleoliquefaction feature extends con-
minimum (N 1)60 values of 7, 11, 12, 13, 6, and tinuously for a large lateral distance along the base
12, we would suggest using a value of about 12 for of a fine-grained cap: bWould lateral spreading have
the exposure, as a whole, assuming that aging effects likely developed if a free face had been nearby?Q
are negligible. This selected value would also be The scenarios enumerated below describe many
subject to the proviso that the field setting be of the possibilities that might be encountered at sites
relatively uniform along the streambank exposure. of marginal liquefaction. These scenarios qualita-
In a companion paper, Green et al. (2004) provide tively consider the effects of post-earthquake density
additional discussion and field examples of this change and subsequent aging on liquefaction resist-
assessment. ance via penetration test results. To make these
Obviously, an assessment of the engineering and assessments, we consider how the measured pene-
geologic data at the site is required to screen tration resistance corresponds to the penetration
254 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

Table 1
Guidelines for selecting a representative penetration resistance value
Ground Sectional view observations Plan view observations (any severity of liquefaction)
failure of marginal liquefaction and sectional view observations of severe liquefaction
mechanism
Hydraulic Designate individual penetration tests Use highest minimum value of penetration
fracturing as locations of marginal liquefaction or no resistance that is common among multiple penetration
liquefaction based on proximity to observed tests performed in proximity to individual
liquefaction features. Use lowest value of liquefaction features created by hydraulic fracturing.
penetration resistance at each test location.

Lateral Designate penetration tests within the probable Use highest minimum value of penetration
spreading limits of lateral spread as marginal liquefaction. resistance that is common among penetration tests
Designate tests outside these limits as no liquefaction. scattered along the length of the lateral spread
Use highest minimum value of penetration resistance (regardless of their proximity to venting features).
common among multiple tests for each designation. This length can be hundreds of meters at places
subjected to strong earthquake shaking.

Surface Same as for hydraulic fracturing. Penetration tests Use highest minimum value of penetration resistance
oscillations should be performed within a few meters of that is commonly present (and typically near the base of
observed liquefaction feature. the fine-grained cap), and is located within a few tens of
meters of the dikes caused by surface oscillations.

Indeterminate Same as for hydraulic fracturing Use lowest value of penetration resistance that is realistically
mechanism feasible for any of the three candidate mechanisms listed above.

resistance that existed shortly (i.e., a couple of (A) The lowest value of (N 1)60 at a site of
months to a couple of years) after the causative marginal liquefaction [i.e., (N 1)60 (min,
earthquake. This time period is consistent with the marg liq)] is slightly less than the value
original development of the capacity curves, and of (N 1) 60 at an adjacent site of no
thus the penetration resistance representative of this liquefaction [i.e., (N 1)60 (min, no liq)].
time period should be used for back-analysis. This criterion suggests that there is only
Because our proposed techniques are rather elabo- a minor difference in soil properties
rate to describe, Fig. 10 presents the techniques in between the sites of marginal and no
flow chart form. In the following discussion and in liquefaction.
Fig. 10, we generically use (N 1)60 or simply N 1 to (B) The higher values of (N 1)60 at the site
represent any overburden-normalized penetration of no liquefaction increase incrementally.
resistance or in-situ test value. We also refer to In other words, there is only a gradual
pga and seismic demand interchangeably below. increase in (N 1)60 as one moves away
While seismic demand actually encompasses both from the observed marginal liquefaction
pga and M and the following back-calculations feature. This incremental increase further
provide values of both pga and M, a better suggests that there is not a large change
assessment of M should be made by integrating in soil properties between adjacent sites.
the back-calculations from individual sites into a It also suggests that the deposit was laid
regional assessment. down in a consistent manner, with the
only difference being a slight change in
(1) There likely has been no significant increase or sand density.
decrease in the value of (N 1)60 at the sites of (C) There is a large difference between the
marginal and no liquefaction resulting from ages of the liquefaction features and the
earthquake shaking if all the following condi- source deposits that liquefied. This crite-
tions in A, B, and C are met. rion, if (A) and (B) are met, suggests that
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 255

Fig. 10. (a) Flow chart for back-calculation using marginal liquefaction features observed in sectional view where minimum (N 1)60 at marginal
liquefaction site is only slightly less than minimum (N 1)60 at adjacent site of no liquefaction. (b) Flow chart for back-calculation using marginal
liquefaction features observed in sectional view where minimum (N 1)60 at marginal liquefaction site is much less than minimum (N 1)60 at
adjacent site of no liquefaction.

aging effects (if present) should be rela- Thus, using the current (N 1)60 value for back-
tively small. analysis of the marginal liquefaction feature
For the conditions in (1A), (1B), and (1C), the should yield the actual seismic demand required
value of (N 1)60 (min, marg liq) is likely suitable to trigger liquefaction.
to estimate the actual value of pga [i.e., pga (2) If conditions (1A) and (1B) are met, but the
(actual)]. In other words, the current (N 1)60 age of the marginal liquefaction features is
value is likely representative of the (N 1)60 value only slightly less than the age of the source
existing shortly after the causative earthquake. sand, then it is not possible to definitively
256 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

evaluate the influence of aging effects because The first occurs where the age difference
any potential aging effects may be similar and between liquefaction event and the age of the
indistinguishable between the adjacent sites. source sand is at least a few hundred years. In
While some methods are available to estimate this case, the effects of aging (where present)
the increase in (N 1)60 resulting from aging can greatly influence the back-analysis results.
(see Olson et al., 2001), these methods can The current (N 1)60 (min, marg liq) yields an
yield a wide range of bage-correctedQ (N 1)60 estimate of pga (actual) that can range from
values, and none of the methods have been being very reasonable (where aging effects are
verified using field data. As a result, the minor) to somewhat too high (where aging has
current value of (N 1)60 (min, marg liq) may increased the value of (N 1)60 over time) to even
be larger than the value of (N 1)60 existing significantly too low (where there was a
shortly after the causative earthquake due to significant post-earthquake decrease in (N 1)60
the effects of aging. Thus, the current (N 1)60 and aging is not present or has not completely
value is suitable for estimating pga (max recovered the loss in (N 1)60). Alternately, the
possible)—not pga (actual). Furthermore, this minimum value of (N 1)60 at the site of no
value of pga (max possible) may be consid- liquefaction will yield pga (max possible).
erably larger than pga (actual), depending on (5) In the second scenario, for conditions in (4)
the contribution from aging. except that the age difference is small, the
(3) If criterion (1A) is met but (1B) is not, (N 1)60 effects of aging (if present) should be similar at
(min, no liq) is likely to be suitable for the site of marginal and no liquefaction. Thus,
estimating pga (actual) and is almost certain to the current (N 1)60 (min, marg liq) yields
be suitable for estimating pga (max possible). In estimates of pga (actual) that can range from
this case, it is not possible to discern if the being very reasonable (where aging effects are
depositional environment was consistent with- minor) to much too low (where there was a
out other geologic evidence nor is it possible to significant post-earthquake decrease in (N 1)60
definitively evaluate the effects of aging. How- and aging is not present or has not completely
ever, because the site of no liquefaction did not recovered the loss in (N 1)60). Again, the value
experience liquefaction and because it is (and of (N 1)60 (min, no liq) will yield pga (max
likely was at the time of the causative earth- possible).
quake) only slightly more resistant to liquefac-
tion than the adjacent site that did experience Note that for all cases above, an estimate of pga
liquefaction, the current (N 1)60 (min, no liq) (max possible) can be obtained.
likely provides a reasonable estimate of pga
(actual). In any case, the current (N 1)60 (min, no 4.2.4. Back-calculations using observations in plan
liq) almost certainly provides pga (max possi- view
ble). The value of pga (max possible) may be Back-calculations using observations of lique-
significantly larger than pga (actual), but likely faction features made in plan view is inherently
would not be because of the only slight differ- more uncertain than using observations in section
ence in liquefaction resistance between the view because the size and extent of liquefaction
adjacent sites of marginal liquefaction and no effects at depth cannot be determined from plan
liquefaction. view observations. Thus at many field sites
(4) In cases where (N 1)60 (min, marg liq) is much where only plan view observations are made, it
less than (N 1)60 (min, no liq), it is more difficult may not be possible to definitively designate
to qualitatively evaluate the influence of post- individual in-situ tests in proximity to a liquefaction
earthquake densification and aging effects. As a feature as bliquefaction/marginal liquefaction/no
result, there is considerably more uncertainty liquefaction.Q
involved with back-calculations of marginal Still, for purposes of back-calculation, sites of bno
liquefaction sites. Two basic scenarios result. liquefactionQ can be identified as sites having poten-
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 257

tially liquefiable sandy sediments (i.e., below the Additional recommendations for site selection and
watertable, Holocene age, relatively low penetration interpretation using plan view observations are as
resistance, etc.) but lacking visible liquefaction follows.
features or evidence of liquefaction (i.e., sand boils,
ground cracking, settlement, tilted buildings, etc.). (1) For a site of hydraulic fracturing, the cap should
Sites of no liquefaction defined as such will provide a not be highly variable in elevation along the base
reasonable upper limit of the magnitude (i.e., duration because of the tendency for sills to form beneath
of demand). At sites of marginal and severe lique- a sloping base and extend far laterally before
faction where multiple in-situ tests are performed, it is developing into dikes that vent onto the surface
possible to make a reasonable estimate of the lower (e.g., Obermeier et al., 2001). Thus, testing
limit of the strength of shaking (i.e., amplitude of should be avoided along fine-grained channel
demand) at a site using a representative value of fillings.
penetration resistance selected using the guidelines in (2) For a site of lateral spreading, multiple in-situ
Table 1. tests should be performed along the length of the
Appropriate site selection for performing back- lateral spread. In selecting a representative
calculations using observations in plan view also penetration resistance, one must consider the
depends on the ground failure mechanism. For possibility that in some in-situ tests the highest
liquefaction resulting from hydraulic fracturing, the minimum penetration resistance values may be
optimal sites are where the cap is relatively thin from sediments that did not liquefy, but rather
(i.e., preferably no more than a meter or two in failed in horizontal shear.
thickness), and the watertable at the time of the (3) Scroll bars may provide excellent test sites
earthquake was located within or at the base of the because the depositional history and age of the
cap. deposit is virtually identical along individual
Whereas the thickness of the fine-grained cap scrolls.
and depth of the watertable can influence surface (4) In order to determine whether aging might be a
manifestations at sites of hydraulic fracturing, cap significant contributor to the value of (N 1)60 (min,
thickness has little or no influence on development marg liq), in-situ tests should preferably be
of lateral spreading (Ishihara, 1985; Youd and conducted at nearby sites of no liquefaction. Final
Garris, 1995). Still, the use of lateral spreading interpretation of the possible effect of aging on
features suffers from the fact that lateral spreading (N 1)60 should be based on relative ages of the
may not develop at higher values of (N 1)60, even liquefaction features in relation to the age of
where there has been relatively strong shaking. possible source deposits (basically as was done
Thus, the best sites for using lateral spreading above for observations in sectional view).
features are beyond the meizoseismal zone, espe-
cially for very strong earthquakes.
For surface oscillations, it appears that the cap 5. Discussion and conclusions
thickness does not have a major bearing on
development of liquefaction features at the surface There can be considerable uncertainty in the
(Obermeier et al., 2001), but this may not always estimate of peak ground acceleration and earthquake
be the case. We suspect that the role of surface magnitude based on the analysis of a single
oscillations in development of surface manifesta- paleoliquefaction site. This uncertainty can result
tions is largely that of fracturing the cap, thereby from unusual bedrock shaking, undefined conditions
providing a conduit for venting to the surface. affecting the strength of shaking in alluvium (such as
Regardless of the actual role, where this mechanism a clay layer located at depth amplifying the ground
has been operative, back-calculations using sites of motion), in-situ test data that inconclusively define a
marginal liquefaction associated with surface oscil- representative value, and other factors. Therefore, we
lations that adjoin sites of no liquefaction effects recommend that individual back-calculations be
are valid. integrated into a regional assessment. In a compan-
258 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

ion paper (Green et al., 2004), we provide details of paper, Green et al. (2004) further describe the
this regional assessment as well as an example of its proposed back-analysis approach and provide exam-
application. ples of its application.
Besides allowing investigators to identify poten-
tially anomalous back-calculation results, performing a
regional assessment can be used to qualitatively Acknowledgments
evaluate the effects of post-earthquake density change
and aging. For studies using observations in either We thank Gerry Wieczorek of U.S. Geological
sectional or plan views, in-situ testing can be performed Survey for a thoughtful review of this paper.
at many sites, regionally, where the ages of the source
sands vary greatly from the age of the paleoearthquake,
with some source sands being only slightly older (i.e., a References
few hundred years older) and some being considerably
older. If the regional attenuation of pga beyond the Ambraseys, N.N., 1988. Engineering seismology. Earthquake
Engineering and Structural Dynamics 17, 1 – 105.
meizoseismal zone forms a smooth plot, then aging is
Andrus, R.D., Stokoe II, K.H., 1997. Liquefaction resistance based
probably not an important contributor to the values of on shear wave velocity: Proc. of the NCEER Workshop on
(N 1)60 (min, marg liq) used for back-calculations. Such Evaluation of Liquefaction Resistance of Soils. In: Youd, T.L.,
an approach was used, de facto, by Pond (1996) in his Idriss, I.M., (Eds.), Technical Report NCEER-97-0022, State
study of the Wabash Valley earthquakes. A major Univ. of New York at Buffalo, NY, pp. 89–128.
Arango, I., 1996. Magnitude scaling factors for soil liquefaction
problem with such an approach is that there may be a
evaluations. ASCE Journal of Geotechnical Engineering 122
huge expenditure of time and money before the effects (11), 929 – 936.
of changes in density and aging become clear. And, Bartlett, S.F., Youd, T.L., 1992. Empirical analysis of horizontal
even in the final analysis, it will probably be difficult to ground displacement generated by liquefaction-induced lateral
assess which sites are most trustworthy. spread: Technical Report No. NCEER-92-0021, State Univ. of
New York at Buffalo, NY, 114 pp.
Because of uncertainties involving the cyclic stress
Bjerrum, L., 1973. Geotechnical problems involved in foundations
and the Green–Mitchell energy-based liquefaction of structures in the North Sea. Geotechnique 23 (3), 319 – 358.
evaluation procedures and the need for local calibration Boulanger, R.W., Mejia, L.H., Idriss, I.M., 1997. Liquefaction at
of the magnitude-bound method, paleoseismic inter- moss landing during Loma Prieta earthquake. ASCE Journal of
pretations have been based on using multiple, inde- Geotechnical and Geoenvironmental Engineering 123 (5),
453 – 467.
pendent methods in some studies. We strongly
Carter, D.P., Seed, H.B., 1988. Liquefaction potential of sand
advocate this approach wherever possible. Where the deposits under low levels of excitation, Report No. UCB/EERC-
different methods yield the same results there can be 88/11, Earthquake Engineering Research Center, Univ. of
very good confidence in interpretations of the strength California, Berkeley, CA. 119 pp.
of shaking, at least in terms of hazard assessment. Dobry, R., Ladd, R.S., Yokel, F.Y., Chung, R.M., Powell, D.,
1982. Prediction of pore water pressure buildup and liquefac-
Unfortunately the number of field settings where
tion of sands during earthquakes by the cyclic strain method.
completely independent assessments can be made is NBS Building Science Series, vol. 138. U.S. Dept. of
quite limited, at least in the United States, because of Commerce. 152 pp.
the short historic record of only a few hundred years. Dowding, C.H., Hryciw, R.D., 1986. A laboratory study of blast
As a result, the need exists to reduce uncertainties densification of saturated sands. ASCE Journal of Geotechnical
Engineering 112 (2), 187 – 199.
in every phase of paleoliquefaction studies. Our Ellis, C., de Alba, P., 1999. Acceleration distribution and
proposed procedures and recommendations can epicentral location of the 1755 bCape AnnQ earthquake from
potentially reduce uncertainties in site selection, field case histories of ground failure. Seismological Research Letters
data collection, and the back-analysis itself (i.e., 70 (6), 758 – 773.
using both the cyclic stress and the Green–Mitchell Fear, C.E., McRoberts, E.C., 1995. Reconsideration of initiation of
liquefaction in sandy soils. ASCE Journal of Geotechnical
energy-based methods). Furthermore, our proposed
Engineering 121 (3), 249 – 261.
techniques allow, at least in a qualitative sense, Finn, W.D.L., 2002. State of the art for the evaluation of
different uncertainties to be examined for their seismic liquefaction potential. Computers and Geotechnics
relevance to the particular study. In a companion 29, 329 – 341.
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 259

Finn, W.D.L., Bransby, P.L., Pickering, D.J., 1970. Effect of strain Lee, K.L., Focht, J.A., 1975. Liquefaction potential of Ekofisk Tank
history on liquefaction of sands. ASCE Journal of the Soil in North Sea. ASCE Journal of the Geotechnical Engineering
Mechanics and Foundations Division 96 (SM6), 1917 – 1934. Division 100 (GTI), 1 – 18.
Frankel, A.D., Petersen, M.D., Mueller, C.S., Haller, K.M., Lodge El-Telbany, A., Rector, J.W., Mitchell, J.K., Seed, R.B.,
Wheeler, R.L., Leyendecker, E.V., Wesson, R.L., Harmsen, 1996. The accuracy of shear wave velocity measurements taken
S.C., Cramer, C.H., Perkins, D.M., Rukstales, K.S., 2002. using the seismic cone penetrometer. Proc. Symposium on the
Documentation for the 2002 update of the national seismic Application of Geophysics to Engineering and Environmental
hazard maps, U.S. Geological Survey Open-File Report 02-420. Problems. Environmental and Engineering Geophysical Society,
Galli, P., 2000. Empirical relationships between magnitude and Keystone, CO, pp. 159 – 169.
distance for liquefaction. Tectonophysics 324, 169 – 187. Lunne, T., Robertson, P.K., Powell, J.J.M., 1997. Cone Penetration
Green, R.A., 2001. Energy-based evaluation and remediation of Testing in Geotechnical Practice. Blackie Academic and
liquefiable soils: PhD Thesis, Civil Engineering Dept., Virginia Professional. 312 pp.
Tech, Blacksburg, VA. 397 pp. http://scholar.lib.vt.edu/theses/ Martin, J.R., Clough, G.W., 1994. Seismic parameters from
available/etd-08132001-170900/. liquefaction evidence. ASCE Journal of Geotechnical Enginee-
Green, R.A., Mitchell, J.K., 2003. A closer look at Arias intensity- ring 120 (8), 1345 – 1361.
based liquefaction evaluation procedures. Proc., 7th Pacific Maugeri, M., Carrubba, P., Frenna, S.M., 1989. Seismic induced
Conference on Earthquake Engineering, Univ. of Canterbury, shear stress for liquefaction analysis. Proc. 4th Inter. Conf. on
Christchurch, New Zealand, Feb. 13–15, Paper No. 94. 9 p. Soil Dynamics and Earthquake Engineering, Southampton,
Green, R.A., Obermeier, S.F., Olson, S.M., 2004. Engineering England, vol. 1, pp. 1 – 20.
geologic and geotechnical analysis of paleoseismic shaking Mesri, G., Godlewski, P.M., 1977. Time- and stress-compressibility
using liquefaction effects: field examples. Engineering Geology, interrelationship. ASCE Journal of Geotechnical Engineering
this issue; doi:10.1016/j.enggeo.2004.07.026. 103 (5), 417 – 430.
Hurst, A., Cronin, B.T., 2001. The origin of consolidation laminae Mesri, G., Feng, T.W., Benak, J.M., 1990. Postdensification
and dish structures in some deep-water sandstones. Journal of penetration resistance of clean sands. ASCE Journal of Geo-
Sedimentary Research 71 (1), 136 – 143. technical Engineering 116 (7), 1095 – 1115.
Hwang, J.H., 1993. Evaluation of simplified method for liquefaction Munson, P.J., Munson, C.A., 1996. Paleoliquefaction evidence for
analysis: report to Provisional Office of High Speed Rail, recurrent strong earthquakes since 20,000 yr BP in the Wabash
Taiwan. Cited in Hwang et al., 1995. Valley of Indiana: Final Report, submitted to the U.S. Geo-
Hwang, J.H., Chang, C.T., Chen, C.H., 1995. Study on stress logical Survey; Anthropology Dept., Indiana University, Bloo-
reduction factor rd for liquefaction analysis. Proc. 1st Interna- mington, IN, March 1996, 137 pp.
tional Conf. on Earthquake Geotechnical Engineering, Tokyo, Nishiyama, H., Yahagi, K., Nakagawa, S., Wada, K., 1977. Practical
Japan, vol. 1, pp. 617 – 622. method of predicting sand liquefaction. Proc. 9th Internat. Conf.
Idriss, I.M., Seed, H.B., 1968. An analysis of ground on Soil Mechanics and Foundation Engineering, Tokyo, Japan,
motions during the 1957 San Francisco earthquake. vol. 2, pp. 305 – 308.
Bulletin of the Seismological Society of America 58 (6), NRC (National Research Council), 1985. Liquefaction of Soils
2013 – 2032. During Earthquakes: Committee on Earthquake Engineering,
Ishibashi, I., Zhang, X., 1993. Unified dynamic shear moduli and Commission on Engineering and Technical Systems, National
damping ratios of sand and clay. Soils and Foundations 33 (1), Research Council. National Academy Press, Washington, DC.
182 – 191. 240 pp.
Ishihara, K., 1985. Stability of natural deposits during earthquakes. Obermeier, S.F., 1993. Paleoliquefaction features as indicators of
Proc. 11th International Conf. on Soil Mechanics and Founda- potential earthquake activity in the southeastern and central
tion Engineering, San Francisco, CA, vol. 1, pp. 321 – 376. United States. Transportation Research Record No. 1141.
Iwasaki, T., Tatsuoka, F., Yasuda, S., 1978. A practical method for Highway Research Board, Washington, pp. 42 – 52.
assessing soil liquefaction potential based on case studies at Obermeier, S.F., 1996. Use of liquefaction-induced features for
various sites in Japan. Proc. 2nd International. Conf. On paleoseismic analysis—an overview of how seismic liquefaction
Microzonation for Safer Construction and Research Applica- features can be distinguished from other features and how their
tions, San Francisco, CA, vol. 2, pp. 885 – 896. regional distribution and properties can be used to infer the
Keefer, D.K., 1984. Landslides caused by earthquakes. Geological location and strength of Holocene paleo-earthquakes. Enginee-
Society of America Bulletin 95, 406 – 421. ring Geology 44, 1 – 76.
King, T.J., 2002. URS Corporation, St. Louis, MO. Personal Obermeier, S.F., 1998. Liquefaction evidence for strong earthquakes
Communication. of Holocene and latest Pleistocene ages in the states of Indiana
Kuribayashi, E., Tatsuoka, F., 1975. Brief review of liquefaction and Illinois, USA. Engineering Geology 50, 227 – 254.
during earthquakes in Japan. Soils and Foundations 15 (4), Obermeier, S.F., Dickenson, S.E., 2000. Liquefaction evidence for
81 – 92. the strength of ground motions resulting from late Holocene
Law, K.T., Cao, Y.L., He, G.N., 1990. An energy approach for Cascadia subduction earthquakes, with emphasis on the event of
assessing seismic liquefaction potential. Canadian Geotechnical 1700 AD. Bulletin of the Seismological Society of America 90
Journal 27 (3), 320 – 329. (4), 876 – 896.
260 S.M. Olson et al. / Engineering Geology 76 (2005) 235–261

Obermeier, S.F., Martin, J.R., Frankel, A.D, Youd, T.L., Munson, Seed, H.B., Idriss, I.M., 1982. Ground motions and soil liquefaction
P.J., Munson, C.A., Pond, E.C., 1993. Liquefaction evidence for during earthquakes. Monograph Series. Earthquake Engineering
one or more strong Holocene earthquakes in the Wabash Valley Research Institute, Oakland, CA. 134 pp.
of southern Indiana and Illinois. U.S. Geological Survey Seed, H.B., Arango, I., Chan, C.K., 1975. Evaluation of Soil
Professional Paper 1536, 27 pp. Liquefaction Potential During Earthquakes. Report No. EERC
Obermeier, S.F., Pond, E.C., Olson, S.M., with contributions by 75-28. Earthquake Engineering Research Center, University of
Green, R.A., Mitchell, J.K., and Stark, T.D., 2001. Paleolique- California, Berkeley, CA. 113 pp.
faction studies in continental settings—geologic and geotech- Seed, H.B., Mori, K., Chan, C.K., 1977. Influence of seismic
nical factors in interpretations and back-analysis: U.S. history on liquefaction of sands. ASCE Journal of the Geo-
Geological Survey Open-File Report 01-029, 75 pp. http:// technical Engineering Division 102 (GT4), 246 – 270.
pubs.usgs.gov/openfile/of01-029. Seed, H.B., Tokimatsu, K., Harder, L.F., Chung, R.L., 1985.
Oda, M., Kawamoto, K., Suzuki, K., Fujimori, H., Sato, M., 2001. Influence of SPT procedures in soil liquefaction resistance
Microstructural interpretation on reliquefaction of saturated evaluations. ASCE Journal of Geotechnical Engineering 111
granular soils under cyclic loading. ASCE Journal of Geo- (12), 1425 – 1445.
technical and Geoenvironmental Engineering 127 (5), 416 – 423. Seed, H.B., Wong, R.T., Idriss, I.M., Tokimatsu, K., 1986. Moduli
Olson, S.M., Stark, T.D., 1998. CPT based liquefaction resistance of and damping factors for dynamic analysis of cohesionless soils.
sandy soils. Proc. Specialty Conf. On Geotechnical Earthquake ASCE Journal of Geotechnical Engineering 112 (GT11),
Engineering and Soil Dynamics III, ASCE, Aug. 3–6, Seattle, 1016 – 1032.
Washington, vol. 1, pp. 325 – 336. Seed, R.B., Dickenson, S.E., Reimer, M.F., Bray, J.D., Sitar, N.,
Olson, S.M., Obermeier, S.F., Stark, T.D., 2001. Interpretation Mitchell, J.K., Idriss, I.M., Kayen, R.E., Kropp, A., Harder, L.F.,
of penetration resistance for back-analysis at sites of Power, M.S., 1990. Preliminary Report on the Principal Geo-
previous liquefaction. Seismological Research Letters 72 technical Aspects of the October 17, 1989 Loma Prieta earth-
(1), 46 – 59. quake: Report UCB/EERC-90/05. Earthquake Engineering
Owen, W.P., Roblee, C.J., 2000. Borehole velocity logging for Research Center, University of California, Berkeley, CA. 137 pp.
Caltrans’ earthquake engineering program: comparison of field Solymar, Z.V., 1984. Compaction of alluvial sands by deep blasting.
measurements. Proc. 1st Intl. Conf. on the Application of Canadian Geotechnical Journal 21 (2), 305 – 321.
Geophysical Methodologies and NDT to Transportation Facili- Stark, T.D., Olson, S.M., 1995. Liquefaction resistance using CPT
ties and Infrastructure, Geophysics 2000, Federal Highway and field case histories. ASCE Journal of Geotechnical
Administration and Missouri Dept. of Transportation, Dec. 11– Engineering 121 (12), 856 – 869.
15, St. Louis, MO, pp. 1.5.1 – 1.5.13. Stone, W.C., Yokel, F.Y., Celebi, M., Hanks, T., Leyendecker, E.V.,
Papadopoulos, G.A., Lefkopoulos, G., 1993. Magnitude–distance 1987. Engineering aspects of the September 19, 1985 Mexico
relations for liquefaction in soil from earthquakes. Bulletin of earthquake. NBS Building Science Series, vol. 165. National
the Seismological Society of America 83 (3), 925 – 938. Bureau of Standards, Washington, DC. 207 pp.
Pond, E.C., 1996. Seismic parameters for the central United States Talwani, P., 2001. University of South Carolina, Columbia, SC.
based on paleoliquefaction evidence in the Wabash Valley: PhD Personal Communication.
thesis, Virginia Tech, Civil Engineering Dept., Blacksburg, VA. Talwani, P., Schaeffer, W.T., 2001. Recurrence rates of large
583 pp. earthquakes in the South Carolina Coastal Plain based on
Pond, E.C., Martin II, J.R., 1996. Seismic parameters for the central paleoliquefaction data. Journal of Geophysical Research 106
United States based on paleoliquefaction evidence in the (B4), 6621 – 6642.
Wabash Valley: Final Report, submitted to the U.S. Geological Terzaghi, K., Peck, R.B., Mesri, G., 1996. Soil Mechanics in
Survey; Civil Engineering Dept., Virginia Tech, Blacksburg, Engineering Practice (Third ed.). Wiley, New York. 549 pp.
VA. August 1996, 583 pp. Tuttle, M.P., 1999. Late Holocene earthquakes and their
Robertson, P.K., Wride, C.E., 1997. Cyclic liquefaction and its implications for earthquake potential of the New Madrid
evaluation based on the SPT and CPT. In: Youd, T.L., Idriss, Seismic Zone, central United States. PhD thesis, Department
I.M. (Eds.), Proc. of the 1996 NCEER Workshop on of Geology, University of Maryland, College Park, MD.
Evaluation of Liquefaction Resistance of Soils, Technical 250 pp.
Report NCEER-97-0022. State University of New York at Vaid, Y.P., Chung, E.K.F., Kuerbis, R.H., 1989. Preshearing and
Buffalo, NY, pp. 41 – 87. undrained response of sand. Soils and Foundations 29 (4),
Schmertmann, J.H., 1991. The mechanical aging of soils. ASCE 49 – 61.
Journal of Geotechnical Engineering 117 (9), 1288 – 1330. Vreugdenhil, R., Davis, R., Berrill, J., 1994. Interpretation of cone
Schneider, J., 1999. Liquefaction response of soils in mid-America penetration results in multilayered soils. International Journal
evaluated by seismic cone tests. MS thesis, School of Civil and for Numerical and Analytical Methods in Geomechanics 18,
Environmental Engineering, Georgia Institute of Technology, 585 – 599.
Atlanta, GA. 273 pp. Whitman, R.V., 1971. Resistance of soil to liquefaction and
Seed, H.B., Idriss, I.M., 1971. Simplified procedure for evaluating settlement. Soils and Foundations 11 (4), 59 – 68.
liquefaction potential. ASCE Journal of the Soil Mechanics and Youd, T.L., 1999. Brigham Young University, Provo, UT. Written
Foundations Division 97 (9), 1249 – 1273. Communication.
S.M. Olson et al. / Engineering Geology 76 (2005) 235–261 261

Youd, T.L., Garris, C.T., 1995. Liquefaction-induced ground- Youd, T.L., Perkins, D.M., 1978. Mapping liquefaction-induced
surface disruption. ASCE Journal of Geotechnical Engineering ground failure potential. ASCE Journal of Geotechnical
121 (11), 805 – 809. Engineering Division 104 (GT4), 433 – 446.
Youd, T.L., Idriss, I.M. (Eds.), 1997. Proceedings of the NCEER Youd, T.L., Idriss, I.M., Andrus, R.D., Arango, I., Castro, G.,
Workshop on Evaluation of Liquefaction Resistance of Soils. Christian, J.T., Dobry, R., Finn, W.D.L., Harder, L.F., Hynes,
Technical Report NCEER-97-0022, State University of New M.E., Ishihara, K., Koester, J.P., Liao, S.S.C., Marcuson, W.F.,
York at Buffalo, NY. 276 pp. Martin, G.R., Mitchell, J.K., Moriwaki, Y., Power, M.S.,
Youd, T.L., Noble, S.K., 1997. Liquefaction criteria based on Robertson, P.K., Seed, R.B., Stokoe, K.H., 2001. Liquefaction
statistical and probabilistic analysis. In: Youd, T.L., Idriss, resistance of soils–Summary report from the 1996 NCEER and
I.M., (Eds.), Proc. of the NCEER Workshop on Evaluation of 1998 NCEER/NSF workshops on evaluation of liquefaction
Liquefaction Resistance of Soils. Technical Report NCEER 97- resistance of soils. ASCE Journal of Geotechnical and Geo-
0022, State University of New York at Buffalo, NY. pp. 210–216. environmental Engineering 127 (4), 297 – 313.

You might also like