Vous êtes sur la page 1sur 24

442

Evaluating cyclic liquefaction potential using the


cone penetration test

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

P.K. Robertson and C.E. (Fear) Wride

Abstract: Soil liquefaction is a major concern for structures constructed with or on sandy soils. This paper describes
the phenomena of soil liquefaction, reviews suitable definitions, and provides an update on methods to evaluate cyclic
liquefaction using the cone penetration test (CPT). A method is described to estimate grain characteristics directly from
the CPT and to incorporate this into one of the methods for evaluating resistance to cyclic loading. A worked example
is also provided, illustrating how the continuous nature of the CPT can provide a good evaluation of cyclic liquefaction
potential, on an overall profile basis. This paper forms part of the final submission by the authors to the proceedings
of the 1996 National Center for Earthquake Engineering Research workshop on evaluation of liquefaction resistance of
soils.
Key words: cyclic liquefaction, sandy soils, cone penetration test.
Rsum : La liqufaction des sols est un risque majeur pour les structures faites en sable ou fondes dessus. Cet
article dcrit les phnomnes de liqufaction des sols, fait le tour des dfinitions sy rapportant et fournit une mise
jour des mthodes permettant dvaluer la liqufaction cyclique partir de lessai de pntration au cne (CPT). On
dcrit une mthode qui permet destimer les caractristiques granulaires partir du CPT et dincorporer directement les
rsultats dans une des mthodes dvaluation de la rsistance au chargement cyclique. Un exemple avec solution est
aussi prsent pour illustrer comment la nature continue du CPT peut donner une bonne ide du potentiel de
liqufaction cyclique sur la base dun profil densemble. Cet article fait partie de la contribution finale des auteurs aux
comptes-rendus du sminaire NCEER tenu en 1996 sur lvaluation de la rsistance des sols la liqufaction.
Mots cls : liqufaction cyclique, sols sableux, CPT
[Traduit par la Rdaction]

Robertson and Wride

Soil liquefaction is a major concern for structures constructed with or on saturated sandy soils. The phenomenon
of soil liquefaction has been recognized for many years.
Terzaghi and Peck (1967) referred to spontaneous liquefaction to describe the sudden loss of strength of very loose
sands that caused flow slides due to a slight disturbance.
Mogami and Kubo (1953) also used the term liquefaction to
describe a similar phenomenon observed during earthquakes.
The Niigata earthquake in 1964 is certainly the event that
focused world attention on the phenomenon of soil liquefaction. Since 1964, much work has been carried out to explain
and understand soil liquefaction. The progress of work on
soil liquefaction has been described in detail in a series of
state-of-the-art papers, such as those by Yoshimi et al.
(1977), Seed (1979), Finn (1981), Ishihara (1993), and Robertson and Fear (1995). The major earthquakes of Niigata in
1964 and Kobe in 1995 have illustrated the significance and
extent of damage that can be caused by soil liquefaction.
Liquefaction was the cause of much of the damage to the
port facilities in Kobe in 1995. Soil liquefaction is also a

Received April 7, 1997. Accepted March 5, 1998.


P.K. Robertson and C.E. (Fear) Wride. Geotechnical
Group, University of Alberta, Edmonton, AB T6G 2G7,
Canada. e-mail: pkrobertson@civil.ualberta.ca
Can. Geotech. J. 35: 442459 (1998)

459

major design problem for large sand structures such as mine


tailings impoundments and earth dams.
The state-of-the-art paper by Robertson and Fear (1995)
provided a detailed description and review of soil liquefaction and its evaluation. In January 1996, the National Center
for Earthquake Engineering Research (NCEER) in the
United States arranged a workshop in Salt Lake City, Utah,
to discuss recent advances in the evaluation of cyclic liquefaction. This paper forms part of the authors final presentation (Robertson and Wride 1998) to the proceedings of that
workshop (Youd and Idriss 1998). The objective of this paper is to provide an update on the evaluation of cyclic liquefaction using the cone penetration test (CPT). Several
phenomena are described as soil liquefaction. In an effort to
clarify the different phenomena, the mechanisms will be
briefly described and definitions for soil liquefaction will be
reviewed.

Before describing methods to evaluate liquefaction potential, it is important to first define the terms used to explain
the phenomena of soil liquefaction.
Figure 1 shows a summary of the behaviour of a granular
soil loaded in undrained monotonic triaxial compression. In
void ratio (e) and mean normal effective stress (p) space, a
soil with an initial void ratio higher than the ultimate state
line (USL) will strain soften (SS) at large strains, eventually
1998 NRC Canada

Robertson and Wride

443

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 1. Schematic of undrained monotonic behaviour of sand in triaxial compression (after Robertson 1994). LSS, limited
strain-softening response; qST, static gravitational shear stress; Su, ultimate undrained shear strength; SH, strain-hardening response; SS,
strain-softening response; US, ultimate state.

reaching an ultimate condition often referred to as critical or


steady state. The term ultimate state (US) is used here, after
Poorooshasb and Consoli (1991). However, a soil with an
initial void ratio lower than the USL will strain harden (SH)
at large strains towards its ultimate state. It is possible to
have a soil with an initial void ratio higher than but close to
the USL. For this soil state, the response can show limited
strain softening (LSS) to a quasi steady state (QSS) (Ishihara
1993), but eventually, at large strains, the response strain
hardens to the ultimate state. For some sands, very large
strains are required to reach the ultimate state, and in some

cases conventional triaxial equipment may not reach these


large strains (axial strain, a > 20%).
During cyclic undrained loading (e.g., earthquake loading), almost all saturated cohesionless soils develop positive
pore pressures due to the contractive response of the soil at
small strains. If there is shear stress reversal, the effective
stress state can progress to the point of essentially zero effective stress, as illustrated in Fig. 2. For shear stress reversal to occur, ground conditions must be generally level or
gently sloping; however, shear stress reversal can occur in
steeply sloping ground if the slope is of limited height
1998 NRC Canada

444

Can. Geotech. J. Vol. 35, 1998

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 2. Schematic of undrained cyclic behaviour of sand illustrating cyclic liquefaction (after Robertson 1994). qcy, cycling shear stress.

(Pando and Robertson 1995). When a soil element reaches


the condition of essentially zero effective stress, the soil has
very little stiffness and large deformations can occur during
cyclic loading. However, when cyclic loading stops, the deformations essentially stop, except for those due to local
pore-pressure redistribution. If there is no shear stress reversal, such as in steeply sloping ground subjected to moderate
cyclic loading, the stress state may not reach zero effective
stress. As a result, only cyclic mobility with limited deformations will occur, provided that the initial void ratio of the
sand is below the USL and the large strain response is strain
hardening (i.e., the material is not susceptible to a catastrophic flow slide). However, shear stress reversal in the
level ground area beyond the toe of a slope may lead to
overall failure of the slope due to softening of the soil in the
toe region.
Based on the above description of soil behaviour in undrained shear, Robertson and Fear (1995), building on earlier work by Robertson (1994), proposed specific definitions
of soil liquefaction which distinguished between flow liquefaction (strain-softening behaviour; see Fig. 1) from cyclic
softening. Cyclic softening was further divided into cyclic
liquefaction (see Fig. 2) and cyclic mobility. A full description of these definitions is given by Robertson and Wride
(1998) in the NCEER report (Youd and Idriss 1998). Figure 3 presents a flow chart (after Robertson 1994) for the
evaluation of liquefaction according to these definitions. The
first step is to evaluate the material characteristics in terms

of a strain-softening or strain-hardening response. If the soil


is strain softening, flow liquefaction is possible if the soil
can be triggered to collapse and if the gravitational shear
stresses are larger than the ultimate or minimum strength.
The trigger mechanism can be either monotonic or cyclic.
Whether a slope or soil structure will fail and slide will depend on the amount of strain-softening soil relative to
strain-hardening soil within the structure, the brittleness of
the strain-softening soil, and the geometry of the ground.
The resulting deformations of a soil structure with both
strain-softening and strain-hardening soils will depend on
many factors, such as distribution of soils, ground geometry,
amount and type of trigger mechanism, brittleness of the
strain-softening soil, and drainage conditions. Soils that are
only temporarily strain softening (i.e., experience a minimum strength before dilating to US) are not as dangerous as
very loose soils that can strain soften directly to ultimate
state. Examples of flow liquefaction failures are Fort Peck
Dam (Casagrande 1965), Aberfan flowslide (Bishop 1973),
Zealand flowslide (Koppejan et al. 1948), and the Stava tailings dam. In general, flow liquefaction failures are not common; however, when they occur, they take place rapidly
with little warning and are usually catastrophic. Hence, the
design against flow liquefaction should be carried out cautiously.
If the soil is strain hardening, flow liquefaction will generally not occur. However, cyclic softening can occur due to
cyclic undrained loading, such as earthquake loading. The
1998 NRC Canada

Robertson and Wride

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 3. Suggested flow chart for evaluation of soil liquefaction


(after Robertson 1994).

amount and extent of deformations during cyclic loading


will depend on the density of the soil, the magnitude and duration of the cyclic loading, and the extent to which shear
stress reversal occurs. If extensive shear stress reversal occurs, it is possible for the effective stresses to reach zero
and, hence, cyclic liquefaction can take place. When the
condition of essentially zero effective stress is achieved,
large deformations can result. If cyclic loading continues,
deformations can progressively increase. If shear stress reversal does not take place, it is generally not possible to
reach the condition of zero effective stress and deformations
will be smaller, i.e., cyclic mobility will occur. Examples of
cyclic softening were common in the major earthquakes in
Niigata in 1964 and Kobe in 1995 and manifested in the
form of sand boils, damaged lifelines (pipelines, etc.), lateral
spreads, slumping of small embankments, settlements, and
ground-surface cracks. If cyclic liquefaction occurs and
drainage paths are restricted due to overlying less permeable
layers, the sand immediately beneath the less permeable soil
can loosen due to pore-water redistribution, resulting in possible subsequent flow liquefaction, given the right geometry.
Both flow liquefaction and cyclic liquefaction can cause
very large deformations. Hence, it can be very difficult to
clearly identify the correct phenomenon based on observed
deformations following earthquake loading. Earthquake-induced flow liquefaction movements tend to occur
after the cyclic loading ceases due to the progressive nature
of the load redistribution. However, if the soil is sufficiently
loose and the static shear stresses are sufficiently large, the
earthquake loading may trigger essentially spontaneous liquefaction within the first few cycles of loading. Also, if the

445

soil is sufficiently loose, the ultimate undrained strength


may be close to zero with an associated effective confining
stress very close to zero (Ishihara 1993). Cyclic liquefaction
movements, on the other hand, tend to occur during the cyclic loading, since it is the inertial forces that drive the phenomenon. The post-earthquake diagnosis can be further
complicated by the possibility of pore-water redistribution
after the cyclic loading, resulting in a change in soil density
and possibly the subsequent triggering of flow liquefaction.
Identifying the type of phenomenon after earthquake loading
is difficult and, ideally, requires instrumentation during and
after cyclic loading together with comprehensive site characterization.
The most common form of soil liquefaction observed in
the field has been cyclic softening due to earthquake loading. Much of the existing research work on soil liquefaction
has been related to cyclic softening, primarily cyclic liquefaction. Cyclic liquefaction generally applies to level or
gently sloping ground where shear stress reversal occurs
during earthquake loading. This paper is concerned primarily with cyclic liquefaction due to earthquake loading and its
evaluation using results of the CPT.

Much of the early work related to earthquake-induced soil


liquefaction resulted from laboratory testing of reconstituted
samples subjected to cyclic loading by means of cyclic
triaxial, cyclic simple shear, or cyclic torsional tests. The
outcome of these studies generally confirmed that the resistance to cyclic loading is influenced primarily by the state of
the soil (i.e., void ratio, effective confining stresses, and soil
structure) and the intensity and duration of the cyclic loading (i.e., cyclic shear stress and number of cycles), as well
as the grain characteristics of the soil. Soil structure incorporates features such as fabric, age, and cementation. Grain
characteristics incorporate features such as grain-size distribution, grain shape, and mineralogy.
Resistance to cyclic loading is usually represented in
terms of a cyclic stress ratio that causes cyclic liquefaction,
termed cyclic resistance ratio (CRR). The point of liquefaction in a cyclic laboratory test is typically defined as the
time at which the sample achieves a strain level of either 5%
double-amplitude axial strain in a cyclic triaxial test or
34% double-amplitude shear strain in a cyclic simple shear
test. For cyclic simple shear tests, CRR is the ratio of the cyclic shear stress to cause cyclic liquefaction to the initial
vertical effective stress, i.e., (CRR)ss = cyc/vo. For cyclic
triaxial tests, CRR is the ratio of the maximum cyclic shear
stress to cause cyclic liquefaction to the initial effective confining stress, i.e., (CRR)tx = dc/23c . The two tests impose
different loading conditions and the CRR values are not
equivalent. Cyclic simple shear tests are generally considered to be better than cyclic triaxial tests at closely representing earthquake loading for level ground conditions.
However, experience has shown that the (CRR)ss can be estimated quite well from (CRR)tx, and correction factors have
been developed (Ishihara 1993). The CRR is typically taken
at about 15 cycles of uniform loading to represent an equivalent earthquake loading of magnitude (M) 7.5, i.e., CRR7.5.
1998 NRC Canada

446

Can. Geotech. J. Vol. 35, 1998

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 4. Graphical representation of liquefaction criteria for silts


and clays from studies by Seed et al. (1973) and Wang (1979) in
China (after Marcuson et al. 1990): < 15% finer than 0.005 mm,
liquid limit (LL) < 35%, and water content > 0.9 liquid limit.

The CRR for any other size earthquake can be estimated


using the following equation:
[1]

CRR =( CRR7.5 ) (MSF)

where MSF is the magnitude scaling factor (recommended


values are provided in the report by NCEER (Youd and
Idriss 1998), which summarizes the results of the 1996
NCEER workshop).
When a soil is fine grained or contains some fines, some
cohesion or adhesion can develop between the fine particles
making the soil more resistant at essentially zero effective
confining stress. Consequently, a greater resistance to cyclic
liquefaction is generally exhibited by sandy soils containing
some fines. However, this tendency depends on the nature of
the fines contained in the sand (Ishihara 1993). Laboratory
testing has shown that one of the most important index properties influencing CRR is the plasticity index of the fines
contained in the sand (Ishihara and Koseki 1989). Ishihara
(1993) showed that the (CRR)tx appears to increase with increasing plasticity index. Studies in China (Wang 1979) suggest that the potential for cyclic liquefaction in silts and
clays is controlled by grain size, liquid limit, and water content. The interpretation of this criterion as given by
Marcuson et al. (1990) and shown in Fig. 4 can be useful;
however, it is important to note that it is based on limited
data and should be used with caution. Figure 4 suggests that
when a soil has a liquid limit less than 35% combined with a
water content greater than 90% of the liquid limit, it is unclear if the soil can experience cyclic liquefaction, and the
soil should be tested to clarify the expected response to undrained cyclic loading.
Although void ratio (relative density) has been recognized
as a dominant factor influencing the CRR of sands, studies
by Ladd (1974), Mulilis et al. (1977), and Tatsuoka et al.
(1986) have clearly shown that sample preparation (i.e., soil
fabric) also plays an important role. This is consistent with
the results of monotonic tests at small to intermediate strain
levels. Hence, if results are to be directly applied with any
confidence, it is important to conduct cyclic laboratory tests
on reconstituted samples with a structure similar to that in

situ. Unfortunately, it is very difficult to determine the in


situ fabric of natural sands below the water table. As a result, there is often some uncertainty in the evaluation of
CRR based on laboratory testing of reconstituted samples,
although, as suggested by Tokimatsu and Hosaka (1986), either the small strain shear modulus or shear wave velocity
measurements could be used to improve the value of laboratory testing on reconstituted samples of sand. Therefore,
there has been increasing interest in testing high-quality undisturbed samples of sandy soils under conditions representative of those in situ. Yoshimi et al. (1989) showed that
aging and fabric had a significant influence on the CRR of
clean sand from Niigata. Yoshimi et al. (1994) also showed
that sand samples obtained using conventional high-quality
fixed piston samplers produced different CRR values than
those of undisturbed samples obtained using in situ ground
freezing. Dense sand samples showed a decrease in CRR
and loose sand samples showed an increase in CRR when
obtained using a piston sampler, as compared with the results of testing in situ frozen samples. The difference in
CRR became more pronounced as the density of the sand increased.
Based on the above observations, for high-risk projects
when evaluation of the potential for soil liquefaction due to
earthquake loading is very important, consideration should
be given to a limited amount of appropriate laboratory tests
on high-quality undisturbed samples. Recently, in situ
ground freezing has been used to obtain undisturbed samples
of sandy soils (Yoshimi et al. 1978, 1989, 1994; Sego et al.
1994; Hofmann et al. 1995; Hofmann 1997). Cyclic simple
shear tests are generally the most appropriate tests, although
cyclic triaxial tests can also give reasonable results.

The above comments have shown that testing high-quality


undisturbed samples will give better results than testing poor
quality samples. However, obtaining high-quality undisturbed samples of saturated sandy soils is very difficult and
expensive and can only be carried out for large projects for
which the consequences of liquefaction may result in large
costs. Therefore, there will always be a need for simple, economic procedures for estimating the CRR of sandy soils,
particularly for low-risk projects and the initial screening
stages of high-risk projects. Currently, the most popular simple method for estimating CRR makes use of penetration resistance from the standard penetration test (SPT), although,
more recently, the CPT has become very popular because of
its greater repeatability and the continuous nature of its profile.
The late Professor H.B. Seed and his coworkers developed a comprehensive SPT-based approach to estimate the
potential for cyclic softening due to earthquake loading. The
approach requires an estimate of the cyclic stress ratio
(CSR) profile caused by a design earthquake. This is usually
done based on a probability of occurrence for a given earthquake. A site-specific seismicity analysis can be carried out
to determine the design CSR profile with depth. A simplified
method to estimate CSR was also developed by Seed and
Idriss (1971) based on the maximum ground surface acceler 1998 NRC Canada

Robertson and Wride

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 5. Comparison between three CPT-based charts for


estimating cyclic resistance ratio (CRR) for clean sands (after
Ishihara 1993). D50, average grain size; FC, fines content.

447

tails of the SPT to avoid or at least minimize the effects of


some of the major factors, the most important of which is
the energy delivered to the SPT sampler. A full description
of the recommended modifications to the simplified SPT
method to estimate cyclic liquefaction is given by Robertson
and Wride (1998) in the NCEER report (Youd and Idriss
1998).
Cone penetration test (CPT)
Because of the inherent difficulties and poor repeatability
associated with the SPT, several correlations have been proposed to estimate CRR for clean sands and silty sands using
corrected CPT penetration resistance (e.g., Robertson and
Campanella 1985; Seed and de Alba 1986; Olsen 1988;
Olsen and Malone 1988; Shibata and Teparaska 1988;
Mitchell and Tseng 1990; Olsen and Koester 1995; Suzuki et
al. 1995a, 1995b; Stark and Olson 1995; Robertson and Fear
1995).
Although cone penetration resistance is often just corrected for overburden stress (resulting in the term qc1), truly
normalized (i.e., dimensionless) cone penetration resistance
corrected for overburden stress (qc1N) can be given by
[3]

ation (amax) at the site. This simplified approach can be summarized as follows:
[2]

CSR =

a
av
= 065
. max vo rd
g
vo

vo

where av is the average cyclic shear stress; amax is the maximum horizontal acceleration at the ground surface; g =
9.81 m/s2 is the acceleration due to gravity; vo and vo are
the total and effective vertical overburden stresses, respectively; and rd is a stress-reduction factor which is dependent
on depth. Details about the rd factor are summarized by
Robertson and Wride (1998), based on the recommendations
by Seed and Idriss (1971) and Liao and Whitman (1986).
The CSR profile from the earthquake can be compared to
the estimated CRR profile for the soil deposit, adjusted to
the same magnitude using eq. [1]. At any depth, if CSR is
greater than CRR, cyclic softening (liquefaction) is possible.
This approach is the most commonly used technique in most
parts of the world for estimating soil liquefaction due to
earthquake loading.
The approach based on the SPT has many problems, primarily due to the inconsistent nature of the SPT. The main
factors affecting the SPT have been reviewed (e.g., Seed et
al. 1985; Skempton 1986; Robertson et al. 1983) and are
summarized by Robertson and Wride (1998). It is highly
recommended that the engineer become familiar with the de-

q
q c1N = c
Pa2

CQ = q c1

Pa2

where qc is the measured cone tip penetration resistance; CQ


= (Pa/vo)n is a correction for overburden stress; the exponent n is typically equal to 0.5; Pa is a reference pressure in
the same units as vo (i.e., Pa = 100 kPa if vo is in kPa);
and Pa2 is a reference pressure in the same units as qc (i.e.,
Pa2 = 0.1 MPa if qc is in MPa). A maximum value of CQ = 2
is generally applied to CPT data at shallow depths. The normalized cone penetration resistance, qc1N, is dimensionless.
Robertson and Campanella (1985) developed a chart for
estimating CRR from corrected CPT penetration resistance
(qc1) based on the Seed et al. (1985) SPT chart and
SPTCPT conversions. Other similar CPT-based charts were
also developed by Seed and de Alba (1986), Shibata and
Teparaska (1988), and Mitchell and Tseng (1990). A comparison between three of these CPT charts is shown in
Fig. 5. In recent years, there has been an increase in available field performance data, especially for the CPT (Ishihara
1993; Kayen et al. 1992; Stark and Olson 1995; Suzuki et al.
1995b). The recent field performance data have shown that
the existing CPT-based correlations to estimate CRR are
generally good for clean sands. The recent field performance
data show that the correlation between CRR and qc1N by
Robertson and Campanella (1985) for clean sands provides a
reasonable estimate of CRR. Based on discussions at the
1996 NCEER workshop, the curve by Robertson and
Campanella (1985) has been adjusted slightly at the lower
end to be more consistent with the SPT curve. The resulting
recommended CPT correlation for clean sand is shown in
Fig. 6. Included in Fig. 6 are suggested curves of limiting
shear strain, similar to those suggested by Seed et al. (1985)
for the SPT. Occurrence of liquefaction is based on level
ground observations of surface manifestations of cyclic liquefaction. For loose sand (i.e., qc1N < 75) this could involve
large deformations resulting from a condition of essentially
zero effective stress being reached. For denser sand (i.e.,
1998 NRC Canada

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

448

Can. Geotech. J. Vol. 35, 1998

Fig. 6. Recommended cyclic resistance ratio (CRR) for clean


sands under level ground conditions based on CPT. l, limiting
shear strain.

Fig. 7. Summary of variation of cyclic resistance ratio (CRR)


with fines content based on CPT field performance data (after
Stark and Olson 1995).

qc1N > 75) this could involve the development of large pore
pressures, but the effective stress may not fully reduce to
zero and deformations may not be as large as those in loose
sands. Hence, the consequences of liquefaction will vary depending on the soil density as well as the size and duration
of loading. An approximate equation for the clean sand CPT
curve shown in Fig. 6 is given later in this paper (eq. [8]).
Based on data from 180 sites, Stark and Olson (1995) also
developed a set of correlations between CRR and cone tip
resistance for various sandy soils based on fines content and
mean grain size, as shown in Fig. 7, in terms of qc1N. The
CPT combined database is now larger than the original
SPT-based database proposed by Seed et al. (1985).
The field observation data used to compile the CPT database are apparently based on the following conditions, similar in nature to those for the SPT-based data: Holocene age,
clean sand deposits; level or gently sloping ground; magnitude M = 7.5 earthquakes; depth range from 1 to 15 m (345
ft) (84% is for depths < 10 m (30 ft)); and representative average CPT qc values for the layer that was considered to
have experienced cyclic liquefaction.
Caution should be exercised when extrapolating the CPT
correlation to conditions outside of the above range. An important feature to recognize is that the correlation appears to
be based on average values for the inferred liquefied layers.
However, the correlation is often applied to all measured
CPT values, which include low values below the average, as
well as low values as the cone moves through soil layer interfaces. Therefore, the correlation can be conservative in
variable deposits where a small part of the CPT data could
indicate possible liquefaction. Although some of the recorded case histories show liquefaction below the suggested
curve in Fig. 6, the data are based on average values and,

hence, the authors consider the suggested curve to be consistent with field observations. It is important to note that the
simplified approach based on either the SPT or the CPT has
many uncertainties. The correlations are empirical and there
is some uncertainty over the degree of conservatism in the
correlations as a result of the methods used to select representative values of penetration resistance within the layers
assumed to have liquefied. A detailed review of the CPT
data, similar to those carried out by Liao and Whitman
(1986) and Fear and McRoberts (1995) on SPT data, would
be required to investigate the degree of conservatism contained in Figs. 6 and 7. The correlations are also sensitive to
the amount and plasticity of the fines within the sand.
For the same CRR, SPT or CPT penetration resistance in
silty sands is smaller because of the greater compressibility
and decreased permeability of silty sands. Therefore, one
reason for the continued use of the SPT has been the need to
obtain a soil sample to determine the fines content of the
soil. However, this has been offset by the poor repeatability
of SPT data. With the increasing interest in the CPT due to
its greater repeatability, several researchers (e.g., Robertson
and Campanella 1985; Olsen 1988; Olsen and Malone 1988;
Olsen and Koester 1995; Suzuki et al. 1995a, 1995b; Stark
and Olson 1995; Robertson and Fear 1995) have developed
a variety of approaches for evaluating cyclic liquefaction potential using CPT results. It is now possible to estimate grain
characteristics such as apparent fines content and grain size
from CPT data and incorporate this directly into the evaluation of liquefaction potential. Robertson and Fear (1995)
recommended an average correction, which was dependent
on apparent fines content, but not on penetration resistance.
This paper provides modifications to and an update of the
CPT approach suggested by Robertson and Fear (1995). The
proposed equation to obtain the equivalent clean sand nor 1998 NRC Canada

Robertson and Wride


Fig. 8. Normalized CPT soil behaviour type chart, as proposed
by Robertson (1990). Soil types: 1, sensitive, fine grained; 2,
peats; 3, silty clay to clay; 4, clayey silt to silty clay; 5, silty
sand to sandy silt; 6, clean sand to silty sand; 7, gravelly sand to
dense sand; 8, very stiff sand to clayey sand (heavily
overconsolidated or cemented); 9, very stiff, fine grained
(heavily overconsolidated or cemented). OCR, overconsolidation
ratio; , friction angle.

449

Based on extensive field data and experience, it is possible to estimate grain characteristics directly from CPT results using the soil behaviour type chart shown in Fig. 8.
The boundaries between soil behaviour type zones 27 can
be approximated as concentric circles (Jefferies and Davies
1993). The radius of each circle can then be used as a soil
behaviour type index. Using the CPT chart by Robertson
(1990), the soil behaviour type index, Ic, can be defined as
follows:

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

[5]

I c = [(3.47 - Q) 2 + (log F + 122


. ) 2 ]0.5
n

malized CPT penetration resistance, (qc1N)cs, is a function of


both the measured penetration resistance, qc1N, and the grain
characteristics of the soil, as follows:
[4]

(q c1N ) cs = Kc q c1N

where Kc is a correction factor that is a function of the grain


characteristics of the soil, as described later in this paper.
Grain characteristics from the CPT
In recent years, charts have been developed to estimate
soil type from CPT data (Olsen and Malone 1988; Olsen and
Koester 1995; Robertson and Campanella 1988; Robertson
1990). Experience has shown that the CPT friction ratio (ratio of the CPT sleeve friction to the cone tip resistance) increases with increasing fines content and soil plasticity.
Hence, grain characteristics such as apparent fines content of
sandy soils can be estimated directly from CPT data using
any of these soil behaviour charts, such as that by Robertson
(1990) shown in Fig. 8. As a result, the measured penetration resistance can be corrected to an equivalent clean sand
value. The addition of pore-pressure data can also provide
valuable additional guidance in estimating fines content.
Robertson et al. (1992) suggested a method for estimating
fines content based on the rate of pore-pressure dissipation
(t50) during a pause in the CPT.

q vo Pa

where Q = c
P

a2

vo
is
the
normalized
CPT
penetration
resistance
(dimensionless); the exponent n is typically equal to 1.0; F =
[fs/(qc vo)]100 is the normalized friction ratio, in percent;
fs is the CPT sleeve friction stress; vo and vo are the total
and effective overburden stresses, respectively; Pa is a reference pressure in the same units as vo (i.e., Pa = 100 kPa if
vo is in kPa); and Pa2 is a reference pressure in the same
units as qc and vo (i.e., Pa2 = 0.1 MPa if qc and vo are
in MPa).
The soil behaviour type chart by Robertson (1990) uses a
normalized cone penetration resistance (Q) based on a simple linear stress exponent of n = 1.0 (see above), whereas
the chart recommended here for estimating CRR (see Fig. 6)
is essentially based on a normalized cone penetration resistance (qc1N) based on a stress exponent n = 0.5 (see eq. [3]).
Olsen and Malone (1988) correctly suggested a normalization where the stress exponent (n) varies from around 0.5 in
sands to 1.0 in clays. However, this normalization for soil
type is somewhat complex and iterative.
The Robertson (1990) procedure using n = 1.0 is recommended for soil classification in clay type soils when Ic >
2.6. However, in sandy soils when Ic 2.6, it is recommended that data being plotted on the Robertson chart be
modified by using n = 0.5. Hence, the recommended procedure is to first use n = 1.0 to calculate Q and, therefore, an
initial value of Ic for CPT data. If Ic > 2.6, the data should be
plotted directly on the Robertson chart (and assume qc1N =
Q). However, if Ic 2.6, the exponent to calculate Q should
be changed to n = 0.5 (i.e., essentially calculate qc1N using
eq. [3], since vo < < qc) and Ic should be recalculated based
on qc1N and F. If the recalculated Ic remains less than 2.6, the
data should be plotted on the Robertson chart using qc1N
based on n = 0.5. If, however, Ic iterates above and below a
value of 2.6, depending which value of n is used, a value of
n = 0.75 should be selected to calculate qc1N (using eq. [3])
and plot data on the Robertson chart. Note that if the in situ
effective overburden stresses are in the order of 50150 kPa,
the choice of normalization has little effect on the calculated
normalized penetration resistance.
The boundaries of soil behaviour type are given in terms
of the index, Ic, as shown in Table 1. The soil behaviour type
index does not apply to zones 1, 8, or 9. Along the normally
consolidated region in Fig. 8, soil behaviour type index increases with increasing apparent fines content and soil plasticity, and the following simplified relationship is suggested:
1998 NRC Canada

450

Can. Geotech. J. Vol. 35, 1998

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Table 1. Boundaries of soil behaviour type (after Robertson 1990).


Soil behaviour type index, Ic

Zone

Soil behaviour type (see Fig. 8)

Ic < 1.31
1.31 < Ic
2.05 < Ic
2.60 < Ic
2.95 < Ic
Ic > 3.60

7
6
5
4
3
2

Gravelly sand to dense sand


Sands: clean sand to silty sand
Sand mixtures: silty sand to sandy silt
Silt mixtures: clayey silt to silty clay
Clays: silty clay to clay
Organic soils: peats

<
<
<
<

2.05
2.60
2.95
3.60

Fig. 9. Variation of CPT soil behaviour type index (Ic) with apparent fines content in or close to the normally consolidated zone of the
soil behaviour chart by Robertson (1990). PI, plasticity index.

[6a] if I c < 126


. apparent fines content FC (%) = 0
[6b] if 126
. I c 35
.
apparent fines content FC (%) = 175
. I c3.25 3.7
[6c] if I c > 35
. apparent fines content FC (%) = 100
The range of potential correlations is illustrated in Fig. 9,
which shows the variation of soil behaviour type index (Ic)
with apparent fines content and the effect of the degree of
plasticity of the fines. The recommended general relationship given in eq. [6] is also shown in Fig. 9. Note that this
equation is slightly modified from the original work by Robertson and Fear (1995) to increase the prediction of apparent
FC for a given value of Ic.
The proposed correlation between CPT soil behaviour index (Ic) and apparent fines content is approximate, since the
CPT responds to many other factors affecting soil behaviour,
such as soil plasticity, mineralogy, sensitivity, and stress history. However, for small projects, the above correlation provides a useful guide. Caution must be taken in applying

eq. [6] to sands that plot in the region defined by 1.64 < Ic <
2.36 and F < 0.5% in Fig. 8 so as not to confuse very loose
clean sands with denser sands containing fines. In this zone,
it is suggested that the apparent fines content is set equal to
5%, such that no correction will be applied to the measured
CPT tip resistance when the CPT data plot in this zone. To
evaluate the correlation shown in Fig. 9 it is important to
show the complete soil profile (CPT and samples, e.g., see
Fig. 13), since comparing soil samples with an adjacent CPT
at the same elevation can be misleading due to soil stratigraphic changes and soil heterogeneity.
Based on the above method for estimating grain characteristics directly from the CPT using the soil behaviour index
(Ic), the recommended relationship between Ic and the correction factor Kc is shown in Fig. 10 and given by the following equations:
[7a]

if I c 164
. ,

Kc = 10
.

[7b]

. ,
if I c 164

Kc = 0.403I c4 + 5581
. I c3 2163
. I c2
+33.75I c 1788
.
1998 NRC Canada

Robertson and Wride

451

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 10. Recommended grain characteristic correction to obtain clean sand equivalent CPT penetration resistance in sandy soils.

The proposed correction factor, Kc, is approximate, since


the CPT responds to many factors, such as soil plasticity,
fines content, mineralogy, soil sensitivity, and stress history.
However, for small projects or for initial screening on larger
projects, the above correlation provides a useful guide. Caution must be taken in applying the relationship to sands that
plot in the region defined by 1.64 < Ic < 2.36 and F 0.5%
so as not to confuse very loose clean sands with sands containing fines. In this zone, it is suggested that the correction
factor Kc be set to a value of 1.0 (i.e., assume that the sand
is a clean sand).
Note that the relationship between the recommended correction factor, Kc, and soil behaviour type index, Ic, is shown
in Fig. 10 as a broken line beyond an Ic of 2.6, which corresponds to an approximate apparent fines content of 35%.
Soils with Ic > 2.6 fall into the clayey silt, silty clay, and
clay regions of the CPT soil behaviour chart (i.e., zones 3
and 4). When the CPT indicates soils in these regions (Ic >
2.6), samples should be obtained and evaluated using criteria
such as those shown in Fig. 4. It is reasonable to assume, in
general, that soils with Ic > 2.6 are nonliquefiable and that
the correction Kc could be large. Soils that fall in the lower
left region of the CPT soil behaviour chart (Fig. 8), defined
by Ic > 2.6 and F 1.0%, can be very sensitive and, hence,
possibly susceptible to both cyclic and (or) flow liquefaction. Soils in this region should be evaluated using criteria
such as those shown in Fig. 4 combined with additional testing.

Figure 11 shows the resulting equivalent CRR curves for


Ic values of 1.64, 2.07, and 2.59 which represent approximate apparent fines contents of 5, 15, and 35%, respectively.
Influence of thin layers
A problem associated with the interpretation of penetration tests in interbedded soils occurs when thin sand layers
are embedded in softer deposits. Theoretical as well as laboratory studies show that the cone resistance is influenced by
the soil ahead of and behind the penetrating cone. The cone
will start to sense a change in soil type before it reaches the
new soil and will continue to sense the original soil even
when it has entered a new soil. As a result, the CPT will not
always measure the correct mechanical properties in thinly
interbedded soils. The distance over which the cone tip
senses an interface increases with increasing soil stiffness. In
soft soils, the diameter of the sphere of influence can be as
small as two to three cone diameters, whereas in stiff soils
the sphere of influence can be up to 20 cone diameters.
Hence, the cone resistance can fully respond (i.e., reach full
value within the layer) in thin soft layers better than in thin
stiff layers. Therefore care should be taken when interpreting cone resistance in thin sand layers located within soft
clay or silt deposits. Based on a simplified elastic solution,
Vreugdenhil et al. (1994) have provided some insight as to
how to correct cone data in thin layers. Vreugdenhil et al.
have shown that the error in the measured cone resistance
within thin stiff layers is a function of the thickness of the
1998 NRC Canada

452
Fig. 11. CPT base curves for various values of soil behaviour
index, Ic (corresponding to various apparent fines contents, as
indicated).

Can. Geotech. J. Vol. 35, 1998

(qc1N)cs, directly from the measured CPT data. Then, using


the equivalent clean sand normalized penetration resistance
(qc1N)cs, the CRR (for M = 7.5) can be estimated using the
following simplified equation (which approximates the clean
sand curve recommended in Fig. 6):
[8a]
3

if

50 (q c1N ) cs < 160,

(q )
CRR = 93 c1N cs + 0.08
1000

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

[8b]
3

if

layer as well as the stiffness of the layer relative to that of


the surrounding softer soil. The relative stiffness of the layers is reflected by the change in cone resistance from the
soft surrounding soil to the stiff soil in the layer.
Vreugdenhil et al. validated the model with laboratory and
field data.
Based on the work by Vreugdenhil et al. (1994), Robertson and Fear (1995) suggested a conservative correction factor for cone resistance. The corrections apply only to thin
sand layers embedded in thick, fine-grained layers. A full
description of the proposed correction factor is given by
Robertson and Wride (1998) in the NCEER report (Youd
and Idriss 1998). Thin sand layers embedded in soft clay deposits are often incorrectly classified as silty sands based on
the CPT soil behaviour type charts. Hence, a slightly improved classification can be achieved if the cone resistance
is first corrected for layer thickness before applying the classification charts.
Cyclic resistance from the CPT
In an earlier section, a method was suggested for estimating apparent fines content directly from CPT results, using
eq. [6]. Following the traditional SPT approach, the estimated apparent fines content could be used to estimate the
correction necessary to obtain the clean sand equivalent penetration resistance. However, since other grain characteristics also influence the measured CPT penetration resistance,
it is recommended that the necessary correction be estimated
from the soil behaviour type index, as described above.
Hence, eqs. [4], [5], and [7] can be combined to estimate the
equivalent clean sand normalized penetration resistance,

(q c1N ) cs < 50,

(q )
CRR = 0833
. c1N cs + 0.05
1000

In summary, eqs. [4][7] and [8] can be combined to provide an integrated method for evaluating the cyclic resistance (M = 7.5) of saturated sandy soils based on the CPT. If
thin layers are present, corrections to the measured tip resistance in each thin layer may be appropriate. The CPT-based
method is an alternative to the SPT or shear wave velocity
(Vs) based in situ methods; however, using more than one
method is useful in providing independent evaluations of
liquefaction potential. The proposed integrated CPT method
is summarized in Fig. 12 in the form of a flow chart. The
flow chart clearly shows the step-by-step process involved
in using the proposed integrated method based on the CPT
for evaluating CRR and indicates the recommended equations for each step of the process.
Although the proposed approach provides a method to
correct CPT results for grain characteristics, it is not intended to remove the need for selected sampling. If the user
has no previous CPT experience in the particular geologic
region, it is important to take carefully selected samples to
evaluate the CPT soil behaviour type classification.
Site-specific modifications are recommended, where possible. However, if extensive CPT experience exists within the
geologic region, the modified CPT method with no samples
can be expected to provide an excellent guide to evaluate
liquefaction potential.
Application of the proposed CPT method
An example of this proposed modified CPT-based method
is shown in Fig. 13 for the Moss Landing site that suffered
cyclic liquefaction during the 1989 Loma Prieta earthquake
in California (Boulanger et al. 1995, 1997). The measured
cone resistance is normalized and corrected for overburden
stress to qc1N and F and the soil behaviour type index (Ic) is
calculated. The final continuous profile of CRR at N = 15
cycles (M = 7.5) is calculated from the equivalent clean sand
values of qc1N (i.e., (qc1N)cs = Kcqc1N) and eq. [8]. Included
in Fig. 13 are measured fines content values obtained from
adjacent SPT samples. A reasonable comparison is seen between the estimated apparent fines contents and the measured fines contents. Note that for Ic > 2.6 (i.e., FC > 35%;
see eq. [6]), the soil is considered to be nonliquefiable; however, this should be checked using other criteria (e.g.,
Marcuson et al. 1990) (see Fig. 4). The estimated zones of
soil that are predicted to experience cyclic liquefaction are
very similar to those observed and reported by Boulanger et
al. (1995, 1997).
1998 NRC Canada

Robertson and Wride

453

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 12. Flow chart illustrating the application of the integrated CPT method of evaluating cyclic resistance ratio (CRR) in sandy soils.
Vs, shear wave velocity.

The predicted zones of liquefaction are slightly conservative compared to the observed ground response. For example, when the CPT passes from a sand into a clay (e.g., at a
depth of 10.5 m), the cone resistance decreases in the transi-

tion zone and the method predicts a very small zone of possible liquefaction. However, given the continuous nature of
the CPT and the high frequency of data (typically every
20 mm), these erroneous predictions are easy to identify. In
1998 NRC Canada

Can. Geotech. J. Vol. 35, 1998

1998 NRC Canada

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

454

Fig. 13. Application of the integrated CPT method for estimating cyclic resistance ratio (CRR) to the State Beach site at Moss Landing which suffered cyclic liquefaction
during the 1989 Loma Prieta earthquake in California.

Robertson and Wride

455

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 14. Comparison of estimating CRR from the CPT. (a) CRR = 0.05 to 0.3, after Olsen and Koester (1995). c, stress exponent.
(b) CRR = 0.15 and 0.25, after Suzuki et al. (1995b). v , vertical effective stress.

general, given the complexity of the problem, the proposed


methodology provides a good overall prediction of the potential for liquefaction. Since the methodology should be applied to low-risk projects and in the initial screening stages
of high-risk projects, it is preferred that the method produces, in general, slightly conservative results.
Comparison with other CPT methods
Olsen (1988), Olsen and Koester (1995), and Suzuki et al.
(1995b) have also suggested integrated methods to estimate
the CRR of sandy soils directly from CPT results with the
correlations presented in the form of soil behaviour charts.
The Olsen and Koester method is based on SPTCPT conversions plus some laboratory-based CRR data. The method
by Suzuki et al. is based on limited field observations. The
methods by Olsen and Koester and Suzuki et al. are shown
in Fig. 14. The Olsen and Koester method uses a variable
normalization technique, which requires an iterative process

to determine the normalization. The method by Suzuki et al.


uses the qc1N normalization suggested in this paper (eq. [3]
with n = 0.5). The Olsen and Koester method is very sensitive to small variations in measured friction ratio and the
user is not able to adjust the correlations based on
site-specific experience. The friction sleeve measurement for
the CPT can vary somewhat depending on specific CPT
equipment and tolerance details between the cone and the
sleeve and, hence, can be subject to some uncertainty. The
method proposed in this paper is based on field observations
and is essentially similar to those of Olsen and Koester and
Suzuki et al.; however, the method described here is slightly
more conservative and the process has been broken down
into its individual components.
Built into each of the CPT methods for estimating CRR is
the step of correcting the measured cone tip resistance to a
clean sand equivalent value. It is the size of this correction
that results in the largest differences between predicted val 1998 NRC Canada

456

Can. Geotech. J. Vol. 35, 1998

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 15. Approximate comparison of various methods for correcting CPT tip resistance to clean sand equivalent values, based on soil
behaviour type.

ues of CRR from the various methods. Figure 15 provides


an approximate comparison between the methods by Stark
and Olson (1995), Olsen and Koester (1995), and Suzuki et
al. (1995b), in terms of Kc from soil behaviour type index,
Ic. Also superimposed in Fig. 15 is the recommended relationship, based on the equations given in this paper. The
comparisons are approximate because the different authors
use different normalizations when correcting CPT tip resistance for plotting data on a soil classification chart. However, for effective overburden stresses in the order of
100 kPa (-1 tsf (tsf= ton-force per square foot)), all of the
normalization methods should give similar values of normalized penetration resistance for the same value of measured
penetration resistance. The boundaries between different soil
behaviour types are also indicated in Fig. 15, as a guide to
the type of soil in which corrections of certain magnitudes
are suggested by the various methods.
The methods by Suzuki et al. (1995b) and Olsen and
Koester (1995) appear to be consistent with each other, indicating that Kc increases with increasing CRR. Very large
corrections result especially in soils of high Ic. The method
by Stark and Olson (1995) indicates that Kc decreases with
increasing CRR for a given Ic and does not fit in with the
trend of the combined Suzuki et al. and Olsen and Koester
lines. The magnitudes of the corrections suggested by Stark

and Olson are generally smaller than those of the other


methods.
Figure 15 indicates that the recommended correction is
generally more conservative than the corrections proposed
by the other authors. The other methods generally predict
higher values of Kc and suggest that corrections should be
applied beginning at lower values of Ic, particularly for
higher values of CRR. Note that the recommended relationship between Kc and Ic is shown in Fig. 15 as a broken line
beyond Ic = 2.6, which corresponds to an apparent FC of
35% (eq. [6]). This shows that the integrated CPT method
for evaluating CRR, as outlined here, does not apply to soils
that would be classified as clayey silt, silty clay, or clay. As
explained earlier, when interpretation of the CPT indicates
that these types of soils are present, samples should be obtained and evaluated using other criteria, such as those given
in Fig. 4 (Marcuson et al. 1990). It is logical that in
nonliquefiable clay soils, the equivalent correction factor,
Kc, could be very large for Ic > 2.6.

For low-risk, small-scale projects, the potential for cyclic


liquefaction can be estimated using penetration tests such as
the CPT. The CPT is generally more repeatable than the SPT
1998 NRC Canada

Robertson and Wride

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Fig. 16. Summary of liquefaction potential on soil classification


chart by Robertson (1990). Zone A, cyclic liquefaction possible,
depending on size and duration of cyclic loading; zone B,
liquefaction unlikely, check other criteria; zone C, flow
liquefaction and (or) cyclic liquefaction possible, depending on
soil plasticity and sensitivity as well as size and duration of
cyclic loading.

and is the preferred test, where possible. The CPT provides


continuous profiles of penetration resistance, which are useful for identifying soil stratigraphy and for providing continuous profiles of estimated cyclic resistance ratio (CRR).
Corrections are required for both the SPT and CPT for grain
characteristics, such as fines content and plasticity. For the
CPT, these corrections are best expressed as a function of
soil behaviour type index, Ic, which is affected by a variety
of grain characteristics.
For medium- to high-risk projects, the CPT can be useful
for providing a preliminary estimate of liquefaction potential
in sandy soils. For higher risk projects, and in cases where
there is no previous CPT experience in the geologic region,
it is also preferred practice to drill sufficient boreholes adjacent to CPT soundings to verify various soil types encountered and to perform index testing on disturbed samples. A
procedure has been described to correct the measured cone
resistance for grain characteristics based on the CPT soil behaviour type index, Ic. The corrections are approximate,
since the CPT responds to many factors affecting soil behaviour. Expressing the corrections in terms of soil behaviour
index is the preferred method of incorporating the effects of
various grain characteristics, in addition to fines content.
When possible, it is recommended that the corrections be
evaluated and modified to suit a specific site and project.
However, for small-scale, low-risk projects and in the initial
screening process for higher risk projects, the suggested gen-

457

eral corrections provide a useful guide. Correcting CPT results in thin sand layers embedded in softer fine-grained deposits may also be appropriate. The CPT is generally limited
to sandy soils with limited gravel contents. In soils with high
gravel contents, penetration may be limited.
A summary of the CPT method is shown in Fig. 16, which
identifies the zones in which soils are susceptible to cyclic
liquefaction (primarily zone A). In general, soils with Ic >
2.6 and F > 1.0% (zone B) are likely nonliquefiable. Soils
that plot in the lower left portion of the chart (zone C; Ic >
2.6 and F < 1.0%) may be susceptible to cyclic and (or) flow
liquefaction due to the sensitive nature of these soils. Soils
in this region should be evaluated using other criteria. Caution should also be exercised when extrapolating the suggested CPT correlations to conditions outside of the range
from which the field performance data were obtained. An
important feature to recognize is that the correlations appear
to be based on average values for the inferred liquefied layers. However, the correlations are often applied to all measured CPT values, which include low values below the
average for a given sand deposit and at the interface between different soil layers. Hence, the correlations could be
conservative in variable stratified deposits where a small
part of the penetration data could indicate possible liquefaction.
As mentioned earlier, it is clearly useful to evaluate CRR
using more than one method. For example, the seismic CPT
can provide a useful technique for independently evaluating
liquefaction potential, since it measures both the usual CPT
parameters and shear wave velocities within the same borehole. The CPT provides detailed profiles of cone tip resistance, but the penetration resistance is sensitive to grain
characteristics, such as fines content and soil mineralogy,
and hence corrections are required. The seismic part of the
CPT provides a shear wave velocity profile typically averaged over 1 m intervals and, therefore, contains less detail
than the cone tip resistance profile. However, shear wave velocity is less influenced by grain characteristics and few or
no corrections are required (Robertson et al. 1992; Andrus
and Stokoe 1998). Shear wave velocity should be measured
with care to provide the most accurate results possible, since
the estimated CRR is sensitive to small changes in shear
wave velocity. There should be consistency in the liquefaction evaluation using either method. If the two methods provide different predictions of CRR profiles, samples should
be obtained to evaluate the grain characteristics of the soil.
A final comment to be made is that a key advantage of the
integrated CPT method described here is that the algorithms
can easily be incorporated into a spreadsheet. As illustrated
by the Moss Landing example presented here, the result is a
straightforward method for analyzing entire CPT profiles in
a continuous manner. This provides an useful tool for the engineer to review the potential for cyclic liquefaction across a
site using engineering judgement.

The authors appreciate the contributions of the members


of the 1996 NCEER Workshop on Evaluation of Liquefaction Resistance of Soils (T.L. Youd, Chair) which was held
in Salt Lake City, Utah. Particular appreciation is extended
1998 NRC Canada

458

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

to the following individuals: W.D.L. Finn, I.M. Idriss, J.


Koester, S. Liao, W.F. Marcuson III, J.K. Mitchell, R.
Olsen, R. Seed, and T.L. Youd.

Andrus, R.D., and Stokoe, K.H. 1998. Guidelines for evaluation of


liquefaction resistance using shear wave velocity. In Proceedings of the 1996 NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, Salt Lake City, Utah. Edited by T.L.
Youd and I.M. Idriss. Report No. NCEER-97-0022.
Bishop, A.W. 1973. The stability of tips and spoil heaps. Quarterly
Journal of Engineering Geology, 6: 335376.
Boulanger, R.W., Idriss, I.M., and Mejia, L.H. 1995. Investigation
and evaluation of liquefaction related ground displacements at
Moss Landing during the 1989 Loma Prieta earthquake. Center
for Geotechnical Modeling, Department of Civil and Environmental Engineering, College of Engineering, University of California at Davis, Report UCD/CGM-95/02.
Boulanger, R.W., Mejia, L.H., and Idriss, I.M. 1997. Liquefaction
at Moss Landing during Loma Prieta earthquake. Journal of
Geotechnical and Geoenvironmental Engineering, 123(5):
453467.
Casagrande, A. 1965. The role of the calculated risk in earthwork and foundation engineering. The Terzaghi Lecture. Journal
of the Soil Mechanics and Foundations Division, ASCE, 91(4):
140.
Fear, C.E., and McRoberts, E.C. 1995. Reconsideration of initiation of liquefaction in sandy soils. Journal of Geotechnical Engineering, ASCE, 121(3): 249261.
Finn, W.D.L. 1981. Liquefaction potential: developments since
1976. In Proceedings of the 1st International Conference on Recent Advances in Geotechnical Earthquake Engineering and Soil
Dynamics, St. Louis, Mo. Vol. 2. Edited by S. Prakash. University of Missouri-Rolla. May. pp. 655681.
Hofmann, B.A. 1997. In-situ ground freezing to obtain undisturbed
samples of loose sand for liquefaction assessment. Ph.D. thesis,
University of Alberta, Edmonton, Alta.
Hofmann, B.A., Sego, D.C., and Robertson, P.K. 1995. Undisturbed sampling of a deep loose sand deposit using ground
freezing. In Proceedings of the 47th Canadian Geotechnical
Conference, Halifax, N.S. CGS. Sept. pp. 287296.
Ishihara, K. 1993. Liquefaction and flow failure during earthquakes. The 33rd Rankine Lecture. Gotechnique, 43(3):
351415.
Ishihara, K., and Koseki, J. 1989. Cyclic shear strength of
fines-containing sands. In Earthquake Geotechnical Engineering, Discussion Session on Influence of Local Conditions
on Seismic Response, Proceedings of the 12th International
Conference on Soil Mechanics and Foundation Engineering, Rio
de Janeiro. A.A. Balkema, Amsterdam. pp. 101106.
Jefferies, M.G., and Davies, M.P. 1993. Use of CPTu to estimate
equivalent SPT N60. Geotechnical Testing Journal, 16(4):
458467.
Kayen, R.E., Mitchell, J.K., Lodge, A., Seed, R.B., Nishio, S., and
Coutinho, R. 1992. Evaluation of SPT-, CPT-, and shear
wave-based methods for liquefaction potential assessment using
Loma Prieta data. In Proceedings of the 4th JapanU.S. Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures for Soil Liquefaction. Edited by M. Hamada
and T.D. ORourke. Technical Report NCEER-94-0019, Vol. 1,
pp. 177204.
Koppejan, A.W., Van Wamelen, B.M., and Weinberg, L.J.H. 1948.
Coastal landslides in the Dutch province of Zealand. In Proceed-

Can. Geotech. J. Vol. 35, 1998


ings of the 2nd International Conference on Soil Mechanics and
Foundation Engineering, Rotterdam, Holland. pp. 8996.
Ladd, R.S. 1974. Specimen preparation and liquefaction of sands.
Journal of the Geotechnical Engineering Division, ASCE,
110(GT10): 11801184.
Liao, S.S.C., and Whitman, R.V. 1986. A catalog of liquefaction
and non-liquefaction occurrences during earthquakes. Research
Report, Department of Civil Engineering, Massachusetts Institute of Technology, Cambridge, Mass.
Marcuson, W.F., III, Hynes, M.E., and Franklin, A.G. 1990. Evaluation and use of residual strength in seismic safety analysis of
embankments. Earthquake Spectra, 6(3): 529572.
Mitchell, J.K., and Tseng, D.-J. 1990. Assessment of liquefaction
potential by cone penetration resistance. In Proceedings of the
H. Bolton Seed Memorial Symposium, Berkeley, Calif. Edited
by J.M. Duncan. Vol. 2. Bitech Publishers. pp. 335350.
Mogami, T., and Kubo, K. 1953. The behaviour of soil during vibration. In Proceedings of the 3rd International Conference on
Soil Mechanics and Foundation Engineering, Vol. 1.
pp. 152153.
Mulilis, J.P., Seed, H.B., Chan, C.K., Mitchell, J.K., and
Arulanandan, K. 1977. Effects of sample preparation on sand
liquefaction. Journal of the Geotechnical Engineering Division,
ASCE, 103(GT2): 99108.
Olsen, R.S. 1988. Using the CPT for dynamic response characterization. In Proceedings of the Earthquake Engineering and Soil
Dynamics II Conference, American Society of Civil Engineers,
New York. ASCE. pp. 111117.
Olsen, R.S., and Koester, J.P. 1995. Prediction of liquefaction resistance using the CPT. In Proceedings of the International Symposium on Cone Penetration Testing, CPT95, Linkoping,
Sweden, Vol. 2. SGS. Oct. pp. 251256.
Olsen, R.S., and Malone, P.G. 1988. Soil classification and site
characterization using the cone penetrometer test. In Penetration
testing 1988. De Ruiter Balkema, Rotterdam. ISOPT-1, Vol. 2.
pp. 887893.
Pando, M., and Robertson, P.K. 1995. Evaluation of shear stress reversal due to earthquake loading for sloping ground. In Proceedings of the 48th Canadian Geotechnical Conference, Vancouver,
B.C. Vol. 2. CGS. Sept. pp. 955962.
Poorooshasb, H.B., and Consoli, N.C. 1991. The ultimate state. In
Proceedings of the 9th Pan-American Conference.
pp. 10831090.
Robertson, P.K. 1990. Soil classification using the cone penetration
test. Canadian Geotechnical Journal, 27(1): 151158.
Robertson, P.K. 1994. Suggested terminology for liquefaction. In
Proceedings of the 47th Canadian Geotechnical Conference,
Halifax, N.S. CGS. Sept. pp. 277286.
Robertson, P.K., and Campanella, R.G. 1985. Liquefaction potential of sands using the cone penetration test. Journal of
Geotechnical Engineering, ASCE, 22(3): 298307.
Robertson, P.K., and Campanella, R.G. 1988. Design manual for
use of CPT and CPTu. Pennsylvania Department of Transportation.
Robertson, P.K., and Fear, C.E. 1995. Liquefaction of sands and its
evaluation. Keynote lecture. In IS Tokyo 95, Proceedings of the
1st International Conference on Earthquake Geotechnical Engineering. Nov. Edited by K. Ishihara. A.A. Balkema, Amsterdam.
Robertson, P.K., and Wride (ne Fear), C.E. 1998. Cyclic liquefaction and its evaluation based on the SPT and CPT. In Proceedings of the 1996 NCEER Workshop on Evaluation of
Liquefaction Resistance of Soils. Edited by T.L. Youd and I.M.
Idriss. NCEER-97-0022.
1998 NRC Canada

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

Robertson and Wride


Robertson, P.K., Campanella, R.G., and Wightman, A. 1983.
SPTCPT correlations. Journal of Geotechnical Engineering,
ASCE, 109: 14491459.
Robertson, P.K., Woeller, D.J., and Finn, W.D.L. 1992. Seismic
cone penetration test for evaluating liquefaction potential under
cyclic loading. Canadian Geotechnical Journal, 29: 686695.
Seed, H.B. 1979. Soil liquefaction and cyclic mobility evaluation
for level ground during earthquakes. Journal of the Geotechnical
Engineering Division, ASCE, 105(GT2): 201255.
Seed, H.B., and de Alba, P. 1986. Use of SPT and CPT tests for
evaluating the liquefaction resistance of sands. In Use of in situ
tests in geotechnical engineering. Edited by S.P. Clemence.
American Society of Civil Engineers, Geotechnical Special Publication 6, pp. 281302.
Seed, H.B., and Idriss, I.M. 1971. Simplified procedure for evaluating soil liquefaction potential. Journal of the Soil Mechanics
and Foundations Division, ASCE, 97(SM9): 12491273.
Seed, H.B., Tokimatsu, K., Harder, L.F., and Chung, R. 1985. Influence of SPT procedures in soil liquefaction resistance evaluations. Journal of Geotechnical Engineering, ASCE, 111(12):
14251445.
Sego, D.C., Robertson, P.K., Sasitharan, S., Kilpatrick, B.L., and
Pillai, V.S. 1994. Ground freezing and sampling of foundation
soils at Duncan Dam. Canadian Geotechnical Journal, 31(6):
939950.
Shibata, T., and Teparaska, W. 1988. Evaluation of liquefaction potentials of soils using cone penetration tests. Soils and Foundations, 28(2): 4960.
Skempton, A.W. 1986. Standard penetration test procedures and
the effects in sands of overburden pressure, relative density, particle size, aging and overconsolidation. Gotechnique, 36(3):
425447.
Stark, T.D., and Olson, S.M. 1995. Liquefaction resistance using
CPT and field case histories. Journal of Geotechnical Engineering, ASCE, 121(12): 856869.
Suzuki, Y., Tokimatsu, K., Taye, Y., and Kubota, Y. 1995a. Correlation between CPT data and dynamic properties of in situ
frozen samples. In Proceedings of the 3rd International Conference on Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics, St. Louis, Mo. Vol. 1. Edited by S.
Prakash. University of Missouri-Rolla. pp. 210215.

459
Suzuki, Y., Tokimatsu, K., Koyamada, K., Taya, Y., and Kubota, Y.
1995b. Field correlation of soil liquefaction based on CPT data.
In Proceedings of the International Symposium on Cone Penetration Testing, CPT 95, Linkoping, Sweden. Vol. 2. SGS. Oct.
pp. 583588.
Tatsuoka, F., Ochi, K., Fujii, S., and Okamoto, M. 1986. Cyclic undrained triaxial and torsional shear strength of sands for different sample preparation methods. Soils and Foundations, 26(3):
2341.
Terzaghi, K., and Peck, R.B. 1967. Soil mechanics in engineering
practice. 2nd ed. John Wiley & Sons, Inc., New York.
Tokimatsu, K., and Hosaka, Y. 1986. Effects of sample disturbance
on dynamic properties of sand. Soils and Foundations, 26(1):
5364.
Vreugdenhil, R., Davis, R., and Berrill, J. 1994. Interpretation of
cone penetration results in multilayered soils. International Journal for Numerical Methods in Geomechanics, 18: 585599.
Wang, W. 1979. Some findings in soil liquefaction. Water Conservancy and Hydroelectric Power Scientific Research Institute,
Beijing, China.
Yoshimi, Y., Richart, F.E., Prakash, S., Balkan, D.D., and Ilyichev,
Y.L. 1977. Soil dynamics and its application to foundation engineering. In Proceedings of the 9th International Conference on
Soil Mechanics and Foundation Engineering, Tokyo, Vol. 2,
pp. 605650.
Yoshimi, Y., Hatanaka, M., and Oh-Oka, H. 1978. Undisturbed
sampling of saturated sands by freezing. Soils and Foundations,
18(3): 105111.
Yoshimi, Y., Tokimatsu, K., and Hosaka, Y. 1989. Evaluation of
liquefaction resistance of clean sands based on high-quality undisturbed samples. Soils and Foundations, 29(1): 93104.
Yoshimi, Y., Tokimatsu, K., and Ohara, J. 1994. In situ liquefaction
resistance of clean sands over a wide density range.
Gotechnique, 44(3): 479494.
Youd, T.L., and Idriss, I.M. (Editors). 1998. Proceedings of the National Center for Earthquake Engineering Research (NCEER)
Workshop on Evaluation of Liquefaction Resistance of Soils,
Salt Lake City, Utah, January 1996. NCEER-97-0022.

1998 NRC Canada

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

This article has been cited by:


1. B.W. Maurer, R.A. Green, M. Cubrinovski, B.A. Bradley. 2015. Fines-content effects on liquefaction hazard evaluation for
infrastructure in Christchurch, New Zealand. Soil Dynamics and Earthquake Engineering 76, 58-68. [CrossRef]
2. Christopher R. McGann, Brendon A. Bradley, Merrick L. Taylor, Liam M. Wotherspoon, Misko Cubrinovski. 2015. Applicability
of existing empirical shear wave velocity correlations to seismic cone penetration test data in Christchurch New Zealand. Soil
Dynamics and Earthquake Engineering 75, 76-86. [CrossRef]
3. Christopher R. McGann, Brendon A. Bradley, Merrick L. Taylor, Liam M. Wotherspoon, Misko Cubrinovski. 2015. Development
of an empirical correlation for predicting shear wave velocity of Christchurch soils from cone penetration test data. Soil Dynamics
and Earthquake Engineering 75, 66-75. [CrossRef]
4. Marco Chini, Matteo Albano, Michele Saroli, Luca Pulvirenti, Marco Moro, Christian Bignami, Emanuela Falcucci, Stefano Gori,
Giuseppe Modoni, Nazzareno Pierdicca, Salvatore Stramondo. 2015. Coseismic liquefaction phenomenon analysis by COSMOSkyMed: 2012 Emilia (Italy) earthquake. International Journal of Applied Earth Observation and Geoinformation 39, 65-78.
[CrossRef]
5. E. Suba Duman, S. B. Ikizler, Z. Angin. 2015. Evaluation of soil liquefaction potential index based on SPT data in the Erzincan,
Eastern Turkey. Arabian Journal of Geosciences 8, 5269-5283. [CrossRef]
6. Sara Khoshnevisan, Hsein Juang, Yan-Guo Zhou, Wenping Gong. 2015. Probabilistic assessment of liquefaction-induced lateral
spreads using CPT Focusing on the 20102011 Canterbury earthquake sequence. Engineering Geology 192, 113-128. [CrossRef]
7. Nurhan Ecemis, Hasan Emre Demirci, Mustafa Karaman. 2015. Influence of consolidation properties on the cyclic re-liquefaction
potential of sands. Bulletin of Earthquake Engineering 13, 1655-1673. [CrossRef]
8. Guojun Cai, Songyu Liu, Anand J. Puppala, Liyuan Tong. 2015. Identification of Soil Strata Based on General Regression Neural
Network Model From CPTU Data. Marine Georesources & Geotechnology 33, 229-238. [CrossRef]
9. Xinhua Xue, Xingguo Yang. 2015. Seismic liquefaction potential assessed by support vector machines approaches. Bulletin of
Engineering Geology and the Environment . [CrossRef]
10. Abbas Abbaszadeh Shahri, Alireza Malehmir, Christopher Juhlin. 2015. Soil classification analysis based on piezocone penetration
test data A case study from a quick-clay landslide site in southwestern Sweden. Engineering Geology 189, 32-47. [CrossRef]
11. G. Papathanassiou, A. Mantovani, G. Tarabusi, D. Rapti, R. Caputo. 2015. Assessment of liquefaction potential for two liquefaction
prone areas considering the May 20, 2012 Emilia (Italy) earthquake. Engineering Geology 189, 1-16. [CrossRef]
12. Liam M. Wotherspoon, Rolando P. Orense, Russell A. Green, Brendon A. Bradley, Brady R. Cox, Clinton M. Wood. 2015.
Assessment of liquefaction evaluation procedures and severity index frameworks at Christchurch strong motion stations. Soil
Dynamics and Earthquake Engineering . [CrossRef]
13. J.J. Lees, R.H. Ballagh, R.P. Orense, S. van Ballegooy. 2015. CPT-based analysis of liquefaction and re-liquefaction following
the Canterbury earthquake sequence. Soil Dynamics and Earthquake Engineering . [CrossRef]
14. Yusuf Erzin, Nurhan Ecemis. 2015. The use of neural networks for CPT-based liquefaction screening. Bulletin of Engineering
Geology and the Environment 74, 103-116. [CrossRef]
15. Ricardo Dobry, Tarek Abdoun. 2015. 3rd Ishihara Lecture: An investigation into why liquefaction charts work: A necessary step
toward integrating the states of art and practice. Soil Dynamics and Earthquake Engineering 68, 40-56. [CrossRef]
16. Reid David. 2015. Estimating slope of critical state line from cone penetration test an update. Canadian Geotechnical Journal
52:1, 46-57. [Abstract] [Full Text] [PDF] [PDF Plus]
17. Claudio di Prisco, Luca Mancinelli, Letizia Zanelotti, Federico Pisan. 2015. Numerical stability analysis of submerged slopes
subject to rapid sedimentation processes. Continuum Mechanics and Thermodynamics 27, 157-172. [CrossRef]
18. Shpresa Gashi, Neritan Shkodrani. 2015. Seismic Soil Liquefaction for Deterministic and Probabilistic Approach Based on in
<i>Situ</i> Test (CPTU) Data. World Journal of Engineering and Technology 03, 41-49. [CrossRef]
19. Efran Ovando Shelley, Vanessa Mussio, Miguel Rodrguez, Jos G. Acosta Chang. 2015. Evaluation of soil liquefaction from
surface analysis. Geofsica Internacional 54, 95-109. [CrossRef]
20. Abolfazl Eslami. 2015. Investigation of explosive compaction (EC) for liquefaction mitigation using CPT records. Bulletin of
Earthquake Engineering . [CrossRef]
21. Abolfazl Eslami, Arash Pirouzi, J. R. Omer, Mahdi Shakeran. 2015. CPT-Based Evaluation of Blast Densification (BD)
Performance in Loose Deposits with Settlement and Resistance Considerations. Geotechnical and Geological Engineering .
[CrossRef]

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

22. Esra Subasi Duman, Sabriye Banu Ikizler, Zekai Angin, Gokhan Demir. 2014. Assessment of liquefaction potential of the Erzincan,
Eastern Turkey. Geomechanics and Engineering 7, 589-612. [CrossRef]
23. Hamid Masaeli, Faramarz Khoshnoudian, Mohammad Hadikhan Tehrani. 2014. Rocking isolation of nonductile moderately tall
buildings subjected to bidirectional near-fault ground motions. Engineering Structures 80, 298-315. [CrossRef]
24. Pijush Samui, J. Karthikeyan. 2014. The Use of a Relevance Vector Machine in Predicting Liquefaction Potential. Indian
Geotechnical Journal 44, 458-467. [CrossRef]
25. D. Neelima Satyam, K. S. Rao. 2014. Liquefaction Hazard Assessment Using SPT and VS for Two Cities in India. Indian
Geotechnical Journal 44, 468-479. [CrossRef]
26. Ikram Guettaya, Mohamed Ridha El Ouni. 2014. In situ-based assessment of soil liquefaction potential-Case study of an earth
dam in Tunisia. Frontiers of Structural and Civil Engineering 8, 456-461. [CrossRef]
27. Filippo Santucci de Magistris, Giovanni Lanzano, Giovanni Forte, Giovanni Fabbrocino. 2014. A peak acceleration threshold for
soil liquefaction: lessons learned from the 2012 Emilia earthquake (Italy). Natural Hazards 74, 1069-1094. [CrossRef]
28. Nurhan Ecemis, Mustafa Karaman. 2014. Influence of non-/low plastic fines on cone penetration and liquefaction resistance.
Engineering Geology 181, 48-57. [CrossRef]
29. Himmet Karaman, Turan Erden. 2014. Net earthquake hazard and elements at risk (NEaR) map creation for city of Istanbul via
spatial multi-criteria decision analysis. Natural Hazards 73, 685-709. [CrossRef]
30. E. Subas Duman, S. B. Ikizler. 2014. Assessment of liquefaction potential of Erzincan Province and its vicinity, Turkey. Natural
Hazards 73, 1863-1887. [CrossRef]
31. Bart Rogiers, Thomas Vienken, Matej Gedeon, Okke Batelaan, Dirk Mallants, Marijke Huysmans, Alain Dassargues. 2014.
Multi-scale aquifer characterization and groundwater flow model parameterization using direct push technologies. Environmental
Earth Sciences 72, 1303-1324. [CrossRef]
32. Firouzianbandpey S., Griffiths D.V., Ibsen L.B., Andersen L.V.. 2014. Spatial correlation length of normalized cone data in sand:
case study in the north of Denmark. Canadian Geotechnical Journal 51:8, 844-857. [Abstract] [Full Text] [PDF] [PDF Plus]
33. Youssef M. A. Hashash, Byungmin Kim, Scott M. Olson, Sungwoo MoonGeotechnical Issues and Site Response in the Central
U.S 71-90. [CrossRef]
34. Dahl Karina R., DeJong Jason T., Boulanger Ross W., Pyke Robert, Wahl Douglas. 2014. Characterization of an alluvial silt
and clay deposit for monotonic, cyclic, and post-cyclic behavior. Canadian Geotechnical Journal 51:4, 432-440. [Abstract] [Full
Text] [PDF] [PDF Plus]
35. B. Vivek, Prishati Raychowdhury. 2014. Probabilistic and spatial liquefaction analysis using CPT data: a case study for Alameda
County site. Natural Hazards 71, 1715-1732. [CrossRef]
36. Xinhua Xue, Xingguo Yang. 2014. Seismic liquefaction potential assessed by fuzzy comprehensive evaluation method. Natural
Hazards 71, 2101-2112. [CrossRef]
37. Pradyut Kumar Muduli, Sarat Kumar Das. 2014. CPT-based Seismic Liquefaction Potential Evaluation Using Multi-gene Genetic
Programming Approach. Indian Geotechnical Journal 44, 86-93. [CrossRef]
38. Guojun Cai, Anand J. Puppala, Songyu Liu. 2014. Characterization on the correlation between shear wave velocity and piezocone
tip resistance of Jiangsu clays. Engineering Geology 171, 96-103. [CrossRef]
39. Kamal Mohamed Hafez Ismail Ibrahim. 2014. Liquefaction analysis of alluvial soil deposits in Bedsa south west of Cairo. Ain
Shams Engineering Journal . [CrossRef]
40. Gi-Chun Kang, Jae-Won Chung, J. David Rogers. 2014. Re-calibrating the thresholds for the classification of liquefaction potential
index based on the 2004 Niigata-ken Chuetsu earthquake. Engineering Geology 169, 30-40. [CrossRef]
41. Pradyut Kumar Muduli, Sarat Kumar Das, Subhamoy Bhattacharya. 2014. CPT-based probabilistic evaluation of seismic soil
liquefaction potential using multi-gene genetic programming. Georisk: Assessment and Management of Risk for Engineered Systems
and Geohazards 8, 14-28. [CrossRef]
42. Qiang Wang, LiYuan Tong. 2014. Determination Permeability Coefficient from Piezocone. Advances in Materials Science and
Engineering 2014, 1-7. [CrossRef]
43. Ali Akbar Firoozi, Mohd Raihan Taha, S. M. Mir Moammad Hosseini, Ali Asghar Firoozi. 2014. Examination of the Behavior
of Gravity Quay Wall against Liquefaction under the Effect of Wall Width and Soil Improvement. The Scientific World Journal
2014, 1-11. [CrossRef]
44. References 969-982. [CrossRef]

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

45. Juang C. Hsein, Ching Jianye, Wang Lei, Khoshnevisan Sara, Ku Chih-Sheng. 2013. Simplified procedure for estimation of
liquefaction-induced settlement and site-specific probabilistic settlement exceedance curve using cone penetration test (CPT).
Canadian Geotechnical Journal 50:10, 1055-1066. [Abstract] [Full Text] [PDF] [PDF Plus]
46. Sayed M. Ahmed, Sherif W. Agaiby, Ahmed H. Abdel-Rahman. 2013. A unified CPTSPT correlation for non-crushable and
crushable cohesionless soils. Ain Shams Engineering Journal . [CrossRef]
47. Laura Tonni, Paolo Simonini. 2013. Shear wave velocity as function of cone penetration test measurements in sand and silt
mixtures. Engineering Geology 163, 55-67. [CrossRef]
48. Ye Lu, Yong Tan, Guoming Lin. 2013. Characterization of thick varved-clayey-silt deposits along the Delaware River by field and
laboratory tests. Environmental Earth Sciences 69, 1845-1860. [CrossRef]
49. V. K. Singh, S. G. Chung. 2013. Determination of Deformation Modulus for Lower Sands in Nakdong River Delta. Marine
Georesources & Geotechnology 31, 290-307. [CrossRef]
50. Ertan Bol. 2013. The influence of pore pressure gradients in soil classification during piezocone penetration test. Engineering
Geology 157, 69-78. [CrossRef]
51. V. K. Singh, S. G. Chung. 2013. Shear Strength Evaluation of Lower Sand in Nakdong River Delta. Marine Georesources &
Geotechnology 31, 107-124. [CrossRef]
52. Li-Ping Huang. 2013. Seismic risk study of a low-seismicity region dominated by large-magnitude and distant earthquakes.
Journal of Asian Earth Sciences 64, 77-85. [CrossRef]
53. Akhter M. Hossain, Ronald D. Andrus, William M. Camp. 2013. Correcting Liquefaction Resistance of Unsaturated Soil Using
Wave Velocity. Journal of Geotechnical and Geoenvironmental Engineering 139, 277-287. [CrossRef]
54. J. KARTHIKEYAN, PIJUSH SAMUI. 2013. Application of statistical learning algorithms for prediction of liquefaction
susceptibility of soil based on shear wave velocity. Geomatics, Natural Hazards and Risk 1-19. [CrossRef]
55. Mohiuddin Ali KhanSeismic Response of Structures to Liquefaction 57-80. [CrossRef]
56. Lalita G. Oka, Mandar M. Dewoolkar, Scott M. Olson. 2012. Liquefaction assessment of cohesionless soils in the vicinity of large
embankments. Soil Dynamics and Earthquake Engineering 43, 33-44. [CrossRef]
57. A. Flora, S. Lirer, F. Silvestri. 2012. Undrained cyclic resistance of undisturbed gravelly soils. Soil Dynamics and Earthquake
Engineering 43, 366-379. [CrossRef]
58. Guojun Cai, Songyu Liu, Anand J. Puppala. 2012. Liquefaction assessments using seismic piezocone penetration (SCPTU) test
investigations in Tangshan region in China. Soil Dynamics and Earthquake Engineering 41, 141-150. [CrossRef]
59. Gopal S. P. Madabhushi, Stuart K. Haigh. 2012. How Well Do We Understand Earthquake Induced Liquefaction?. Indian
Geotechnical Journal 42, 150-160. [CrossRef]
60. K.S. Vipin, T.G. Sitharam. 2012. A performance-based framework for assessing liquefaction potential based on CPT data. Georisk:
Assessment and Management of Risk for Engineered Systems and Geohazards 6, 177-187. [CrossRef]
61. Min-Woo Seo, Scott M. Olson, Chang-Guk Sun, Myoung-Hak Oh. 2012. Evaluation of Liquefaction Potential Index Along
Western Coast of South Korea Using SPT and CPT. Marine Georesources & Geotechnology 30, 234-260. [CrossRef]
62. Ye Lu, Yong Tan. 2012. Examination of loose saturated sands impacted by a heavy tamper. Environmental Earth Sciences 66,
1557-1567. [CrossRef]
63. H. W. Huang, J. Zhang, L. M. Zhang. 2012. Bayesian network for characterizing model uncertainty of liquefaction potential
evaluation models. KSCE Journal of Civil Engineering 16, 714-722. [CrossRef]
64. Africa M. Geremew, Ernest K. Yanful. 2012. Laboratory Investigation of the Resistance of Tailings and Natural Sediments to
Cyclic Loading. Geotechnical and Geological Engineering 30, 431-447. [CrossRef]
65. Khaled H. Charif, Shadi NajjarComparative Study of Shear Modulus in Calcareous Sand and Sabkha Soils 2697-2706. [CrossRef]
66. VienkenThomas, LevenCarsten, DietrichPeter. 2012. Use of CPT and other direct push methods for (hydro-) stratigraphic
aquifer characterization a field study. Canadian Geotechnical Journal 49:2, 197-206. [Abstract] [Full Text] [PDF] [PDF Plus]
67. M H BAGHERIPOUR, I SHOOSHPASHA, M AFZALIRAD. 2012. A genetic algorithm approach for assessing soil liquefaction
potential based on reliability method. Journal of Earth System Science 121, 45-62. [CrossRef]
68. KuChih-Sheng, JuangC. Hsein, ChangChi-Wen, ChingJianye. 2012. Probabilistic version of the Robertson and Wride method for
liquefaction evaluation: development and application. Canadian Geotechnical Journal 49:1, 27-44. [Abstract] [Full Text] [PDF]
[PDF Plus]
69. Gordon Tung-Chin Kung, Der-Her Lee, Pai-Hsiang Tsai. 2011. Examination of DMT-based methods for evaluating the
liquefaction potential of soils. Journal of Zhejiang University SCIENCE A 12, 807-817. [CrossRef]

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

70. P. K. Robertson. 2011. Closure to Evaluation of Flow Liquefaction and Liquefied Strength Using the Cone Penetration Test
by P. K. Robertson. Journal of Geotechnical and Geoenvironmental Engineering 137, 984-986. [CrossRef]
71. E. Gyri, L. Tth, Z. Grczer, T. Katona. 2011. Liquefaction and post-liquefaction settlement assessment A probabilistic
approach. Acta Geodaetica et Geophysica Hungarica 46, 347-369. [CrossRef]
72. Guojun Cai, Songyu Liu, Anand J. Puppala. 2011. Comparison of CPT charts for soil classification using PCPT data: Example
from clay deposits in Jiangsu Province, China. Engineering Geology 121, 89-96. [CrossRef]
73. K. S. Vipin, T. G. Sitharam. 2011. Evaluation of liquefaction return period based on local site classes: performance based logic
tree approach. International Journal of Geotechnical Engineering 5, 245-254. [CrossRef]
74. Chih-Sheng Ku, C. Hsein Juang, Jianye Ching, Der-Her LeeLiquefaction Probability by Probabilistic Version of Robertson and
Wride Model 247-254. [CrossRef]
75. T. Heidari, R. D. Andrus, S. M. J. MoyseyCharacterizing the Liquefaction Potential of the Pleistocene-Age Wando Formation
in the Charleston Area, South Carolina 510-517. [CrossRef]
76. D. A. Saftner, R. A. Green, P.E., R. D. Hryciw, M. F. Fadden, A. DaCostaCharacterizing Spatial Variability of Cone Penetration
Testing through Geostatistical Evaluation 428-435. [CrossRef]
77. TonniLaura, GottardiGuido. 2011. Analysis and interpretation of piezocone data on the silty soils of the Venetian lagoon (Treporti
test site). Canadian Geotechnical Journal 48:4, 616-633. [Abstract] [Full Text] [PDF] [PDF Plus]
78. KarrayMourad, LefebvreGuy, EthierYannic, BigrasAnnick. 2011. Influence of particle size on the correlation between shear wave
velocity and cone tip resistance. Canadian Geotechnical Journal 48:4, 599-615. [Abstract] [Full Text] [PDF] [PDF Plus]
79. C. Schmelzbach, J. Tronicke, P. Dietrich. 2011. Three-dimensional hydrostratigraphic models from ground-penetrating radar
and direct-push data. Journal of Hydrology 398, 235-245. [CrossRef]
80. Mohammad Rezania, Asaad Faramarzi, Akbar A. Javadi. 2011. An evolutionary based approach for assessment of earthquakeinduced soil liquefaction and lateral displacement. Engineering Applications of Artificial Intelligence 24, 142-153. [CrossRef]
81. Asl Kurtulu. 2011. Pipeline Vulnerability of Adapazar during the 1999 Kocaeli, Turkey, Earthquake. Earthquake Spectra 27,
45-66. [CrossRef]
82. Pijush Samui, Dookie Kim, T.G. Sitharam. 2011. Support vector machine for evaluating seismic-liquefaction potential using shear
wave velocity. Journal of Applied Geophysics 73, 8-15. [CrossRef]
83. Edgar G.DiazE.G. Diaz, FernandoRodrguez-RoaF. Rodrguez-Roa. 2010. Design load of rigid footings on sand. Canadian
Geotechnical Journal 47:8, 872-884. [Abstract] [Full Text] [PDF] [PDF Plus]
84. Ya-Fen Lee, Yun-Yao Chi, C. Hsein Juang, Der-Her Lee. 2010. Annual probability and return period of soil liquefaction in
Yuanlin, Taiwan attributed to Chelungpu Fault and Changhua Fault. Engineering Geology 114, 343-353. [CrossRef]
85. C. HseinJuangC.H. Juang, Chang-YuOuC.-Y. Ou, Chih-ChiehLuC.-C. Lu, ZheLuoZ. Luo. 2010. Probabilistic framework for
assessing liquefaction hazard at a given site in a specified exposure time using standard penetration testing. Canadian Geotechnical
Journal 47:6, 674-687. [Abstract] [Full Text] [PDF] [PDF Plus]
86. Tahereh Heidari, Ronald D. Andrus. 2010. Mapping liquefaction potential of aged soil deposits in Mount Pleasant, South Carolina.
Engineering Geology 112, 1-12. [CrossRef]
87. Mohammad Rezania, Akbar A. Javadi, Orazio Giustolisi. 2010. Evaluation of liquefaction potential based on CPT results using
evolutionary polynomial regression. Computers and Geotechnics 37, 82-92. [CrossRef]
88. P. K.RobertsonP.K. Robertson. 2009. Interpretation of cone penetration tests a unified approach. Canadian Geotechnical Journal
46:11, 1337-1355. [Abstract] [Full Text] [PDF] [PDF Plus]
89. A. Marache, D. Breysse, C. Piette, P. Thierry. 2009. Geotechnical modeling at the city scale using statistical and geostatistical
tools: The Pessac case (France). Engineering Geology 107, 67-76. [CrossRef]
90. M.S. Ravi Sharma, K. Ilamparuthi. 2009. Interpretation of electric piezocone data of Chennai coast. Ocean Engineering 36, 511-520.
[CrossRef]
91. Pai-Hsiang Tsai, Der-Her Lee, Gordon Tung-Chin Kung, C. Hsein Juang. 2009. Simplified DMT-based methods for evaluating
liquefaction resistance of soils. Engineering Geology 103, 13-22. [CrossRef]
92. P.BreulP. Breul, Y.HaddaniY. Haddani, R.GourvsR. Gourvs. 2008. On site characterization and air content evaluation of coastal
soils by image analysis to estimate liquefaction risk. Canadian Geotechnical Journal 45:12, 1723-1732. [Abstract] [Full Text]
[PDF] [PDF Plus]
93. C. Hsein Juang, Chia-Nan Liu, Chien-Hsun Chen, Jin-Hung Hwang, Chih-Chieh Lu. 2008. Calibration of liquefaction potential
index: A re-visit focusing on a new CPTU model. Engineering Geology 102, 19-30. [CrossRef]

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

94. Dawn A.ShuttleD.A. Shuttle, JohnCunningJ. Cunning. 2008. Reply to the discussion by Robertson on Liquefaction potential
of silts from CPTuAppears in Canadian Geotechnical Journal, 45: 140141.. Canadian Geotechnical Journal 45:1, 142-145.
[Citation] [Full Text] [PDF] [PDF Plus]
95. S. Feyza Cinicioglu, Ilknur Bozbey, Sadik Oztoprak, M. Kubilay Kelesoglu. 2007. An integrated earthquake damage assessment
methodology and its application for two districts in Istanbul, Turkey. Engineering Geology 94, 145-165. [CrossRef]
96. Anthony T.C. Goh, S.H. Goh. 2007. Support vector machines: Their use in geotechnical engineering as illustrated using seismic
liquefaction data. Computers and Geotechnics 34, 410-421. [CrossRef]
97. Jennifer A. Lenz, Laurie G. Baise. 2007. Spatial variability of liquefaction potential in regional mapping using CPT and SPT
data. Soil Dynamics and Earthquake Engineering 27, 690-702. [CrossRef]
98. C. Hsein Juang, David Kun Li. 2007. Assessment of liquefaction hazards in Charleston quadrangle, South Carolina. Engineering
Geology 92, 59-72. [CrossRef]
99. Adel M. Hanna, Derin Ural, Gokhan Saygili. 2007. Evaluation of liquefaction potential of soil deposits using artificial neural
networks. Engineering Computations 24, 5-16. [CrossRef]
100. Dawn A Shuttle, John Cunning. 2007. Liquefaction potential of silts from CPTu. Canadian Geotechnical Journal 44:1, 1-19.
[Abstract] [PDF] [PDF Plus]
101. Measurement of Properties 127-158. [CrossRef]
102. Radu Popescu, Jean H. Prevost, George Deodatis, Pradipta Chakrabortty. 2006. Dynamics of nonlinear porous media with
applications to soil liquefaction. Soil Dynamics and Earthquake Engineering 26, 648-665. [CrossRef]
103. I.M. Idriss, R.W. Boulanger. 2006. Semi-empirical procedures for evaluating liquefaction potential during earthquakes. Soil
Dynamics and Earthquake Engineering 26, 115-130. [CrossRef]
104. T. Wichtmann, A. Niemunis, Th. Triantafyllidis, M. Poblete. 2005. Correlation of cyclic preloading with the liquefaction
resistance. Soil Dynamics and Earthquake Engineering 25, 923-932. [CrossRef]
105. D.S. Liyanapathirana, H.G. Poulos. 2004. Assessment of soil liquefaction incorporating earthquake characteristics. Soil Dynamics
and Earthquake Engineering 24, 867-875. [CrossRef]
106. Ellen M Rathje, Ismail Karatas, Stephen G Wright, Jeff Bachhuber. 2004. Coastal failures during the 1999 Kocaeli earthquake
in Turkey. Soil Dynamics and Earthquake Engineering 24, 699-712. [CrossRef]
107. C Hsein Juang, Susan Hui Yang, Haiming Yuan, Eng Hui Khor. 2004. Characterization of the uncertainty of the Robertson and
Wride model for liquefaction potential evaluation. Soil Dynamics and Earthquake Engineering 24, 771-780. [CrossRef]
108. Chih-Sheng Ku, Der-Her Lee, Jian-Hong Wu. 2004. Evaluation of soil liquefaction in the Chi-Chi, Taiwan earthquake using
CPT. Soil Dynamics and Earthquake Engineering 24, 659-673. [CrossRef]
109. Ping-Sien Lin, Chi-Wen Chang, Wen-Jong Chang. 2004. Characterization of liquefaction resistance in gravelly soil: large hammer
penetration test and shear wave velocity approach. Soil Dynamics and Earthquake Engineering 24, 675-687. [CrossRef]
110. Ronald D Andrus, Paramananthan Piratheepan, Brian S Ellis, Jianfeng Zhang, C Hsein Juang. 2004. Comparing liquefaction
evaluation methods using penetration-VS relationships. Soil Dynamics and Earthquake Engineering 24, 713-721. [CrossRef]
111. G. Zhang, P. K. Robertson, R. W. I. Brachman. 2004. Estimating Liquefaction-Induced Lateral Displacements Using the Standard
Penetration Test or Cone Penetration Test. Journal of Geotechnical and Geoenvironmental Engineering 130, 861-871. [CrossRef]
112. Jonathan D. Bray, Rodolfo B. Sancio, Turan Durgunoglu, Akin Onalp, T. Leslie Youd, Jonathan P. Stewart, Raymond B. Seed,
Onder K. Cetin, Ertan Bol, M. B. Baturay, C. Christensen, T. Karadayilar. 2004. Subsurface Characterization at Ground Failure
Sites in Adapazari, Turkey. Journal of Geotechnical and Geoenvironmental Engineering 130, 673-685. [CrossRef]
113. Bin-Lin Chu, Sung-Chi Hsu, Yi-Ming Chang. 2004. Ground behavior and liquefaction analyses in central Taiwan-Wufeng.
Engineering Geology 71, 119-139. [CrossRef]
114. Der-Her Lee, Chih-Sheng Ku, Haiming Yuan. 2004. A study of the liquefaction risk potential at Yuanlin, Taiwan. Engineering
Geology 71, 97-117. [CrossRef]
115. Haiming Yuan, Susan Hui Yang, Ronald D Andrus, C Hsein Juang. 2004. Liquefaction-induced ground failure: a study of the
Chi-Chi earthquake cases. Engineering Geology 71, 141-155. [CrossRef]
116. Ross W. Boulanger. 2003. High Overburden Stress Effects in Liquefaction Analyses. Journal of Geotechnical and Geoenvironmental
Engineering 129, 1071-1082. [CrossRef]
117. J. A. H. Carraro, P. Bandini, R. Salgado. 2003. Liquefaction Resistance of Clean and Nonplastic Silty Sands Based on Cone
Penetration Resistance. Journal of Geotechnical and Geoenvironmental Engineering 129, 965-976. [CrossRef]

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15


For personal use only.

118. M.H. Baziar, N. Nilipour. 2003. Evaluation of liquefaction potential using neural-networks and CPT results. Soil Dynamics and
Earthquake Engineering 23, 631-636. [CrossRef]
119. Tamer Elkateb, Rick Chalaturnyk, Peter K Robertson. 2003. Simplified geostatistical analysis of earthquake-induced ground
response at the Wildlife Site, California, U.S.A. Canadian Geotechnical Journal 40:1, 16-35. [Abstract] [PDF] [PDF Plus]
120. Jean-Pierre Bardet71 Advances in analysis of soil liquefaction during earthquakes 1175-1201. [CrossRef]
121. C. Hsein Juang, Haiming Yuan, Der-Her Lee, Ping-Sien Lin. 2003. Simplified Cone Penetration Test-based Method for
Evaluating Liquefaction Resistance of Soils. Journal of Geotechnical and Geoenvironmental Engineering 129, 66-80. [CrossRef]
122. G Zhang, P K Robertson, R W.I Brachman. 2002. Estimating liquefaction-induced ground settlements from CPT for level
ground. Canadian Geotechnical Journal 39:5, 1168-1180. [Abstract] [PDF] [PDF Plus]
123. T. Liao, P.W. Mayne, M.P. Tuttle, E.S. Schweig, R.B. Van Arsdale. 2002. CPT site characterization for seismic hazards in the
New Madrid seismic zone. Soil Dynamics and Earthquake Engineering 22, 943-950. [CrossRef]
124. W.D.Liam Finn. 2002. State of the art for the evaluation of seismic liquefaction potential. Computers and Geotechnics 29, 329-341.
[CrossRef]
125. C. Hsein Juang, Tao Jiang, Ronald D. Andrus. 2002. Assessing Probability-based Methods for Liquefaction Potential Evaluation.
Journal of Geotechnical and Geoenvironmental Engineering 128, 580-589. [CrossRef]
126. C. Hsein Juang, Haiming Yuan, Der-Her Lee, Chih-Sheng Ku. 2002. Assessing CPT-based methods for liquefaction evaluation
with emphasis on the cases from the Chi-Chi, Taiwan, earthquake. Soil Dynamics and Earthquake Engineering 22, 241-258.
[CrossRef]
127. Der-Her Lee, C Hsein Juang, Chi-Sheng Ku. 2001. Liquefaction performance of soils at the site of a partially completed ground
improvement project during the 1999 Chi-Chi earthquake in Taiwan. Canadian Geotechnical Journal 38:6, 1241-1253. [Abstract]
[PDF] [PDF Plus]
128. James A. Schneider, Paul W. Mayne, Glenn J. Rix. 2001. Geotechnical site characterization in the greater Memphis area using
cone penetration tests. Engineering Geology 62, 169-184. [CrossRef]
129. M Yoshimine, P K Robertson, C E (Fear) Wride. 2001. Undrained shear strength of clean sands to trigger flow liquefaction:
Reply. Canadian Geotechnical Journal 38:3, 654-657. [Citation] [PDF] [PDF Plus]
130. C Hsein Juang, Caroline J Chen, Tao Jiang, Ronald D Andrus. 2000. Risk-based liquefaction potential evaluation using standard
penetration tests. Canadian Geotechnical Journal 37:6, 1195-1208. [Abstract] [PDF] [PDF Plus]
131. J LH Grozic, P K Robertson, N R Morgenstern. 2000. Cyclic liquefaction of loose gassy sand. Canadian Geotechnical Journal
37:4, 843-856. [Abstract] [PDF] [PDF Plus]
132. Ronald D. Andrus, Kenneth H. StokoeLiquefaction Evaluation of Densified Sand at Treasure Island, California 264-278.
[CrossRef]
133. P K Robertson, C E (Fear) Wride, B R List, U Atukorala, K W Biggar, P M Byrne, R G Campanella, D C Cathro, D H Chan,
K Czajewski, WDL Finn, W H Gu, Y Hammamji, B A Hofmann, J A Howie, J Hughes, A S Imrie, J -M Konrad, A Kpper,
T Law, E RF Lord, P A Monahan, N R Morgenstern, R Phillips, R Pich, H D Plewes, D Scott, D C Sego, J Sobkowicz, R
A Stewart, B D Watts, D J Woeller, T L Youd, Z Zavodni. 2000. The CANLEX project: summary and conclusions. Canadian
Geotechnical Journal 37:3, 563-591. [Abstract] [PDF] [PDF Plus]
134. C E (Fear) Wride, E C McRoberts, P K Robertson. 1999. Reconsideration of case histories for estimating undrained shear strength
in sandy soils. Canadian Geotechnical Journal 36:5, 907-933. [Abstract] [PDF] [PDF Plus]
135. M Yoshimine, P K Robertson, C E (Fear) Wride. 1999. Undrained shear strength of clean sands to trigger flow liquefaction.
Canadian Geotechnical Journal 36:5, 891-906. [Abstract] [PDF] [PDF Plus]

Vous aimerez peut-être aussi