Vous êtes sur la page 1sur 24

442

Evaluating cyclic liquefaction potential using the


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

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.

Résumé : La liquéfaction des sols est un risque majeur pour les structures faites en sable ou fondées dessus. Cet
article décrit les phénomènes de liquéfaction des sols, fait le tour des définitions s’y rapportant et fournit une mise à
jour des méthodes permettant d’évaluer la liquéfaction cyclique à partir de l’essai de pénétration au cône (CPT). On
décrit une méthode qui permet d’estimer les caractéristiques granulaires à partir du CPT et d’incorporer directement les
résultats dans une des méthodes d’évaluation de la résistance au chargement cyclique. Un exemple avec solution est
aussi présenté pour illustrer comment la nature continue du CPT peut donner une bonne idée du potentiel de
For personal use only.

liquéfaction cyclique sur la base d’un profil d’ensemble. Cet article fait partie de la contribution finale des auteurs aux
comptes-rendus du séminaire NCEER tenu en 1996 sur l’évaluation de la résistance des sols à la liquéfaction.

Mots clés : liquéfaction cyclique, sols sableux, CPT

[Traduit par la Rédaction] Robertson and Wride 459

major design problem for large sand structures such as mine


tailings impoundments and earth dams.
Soil liquefaction is a major concern for structures con- The state-of-the-art paper by Robertson and Fear (1995)
structed with or on saturated sandy soils. The phenomenon provided a detailed description and review of soil liquefac-
of soil liquefaction has been recognized for many years. tion and its evaluation. In January 1996, the National Center
Terzaghi and Peck (1967) referred to “spontaneous liquefac- for Earthquake Engineering Research (NCEER) in the
tion” to describe the sudden loss of strength of very loose United States arranged a workshop in Salt Lake City, Utah,
sands that caused flow slides due to a slight disturbance. to discuss recent advances in the evaluation of cyclic lique-
Mogami and Kubo (1953) also used the term liquefaction to faction. This paper forms part of the authors’ final presenta-
describe a similar phenomenon observed during earthquakes. tion (Robertson and Wride 1998) to the proceedings of that
The Niigata earthquake in 1964 is certainly the event that workshop (Youd and Idriss 1998). The objective of this pa-
focused world attention on the phenomenon of soil liquefac- per is to provide an update on the evaluation of cyclic lique-
tion. Since 1964, much work has been carried out to explain faction using the cone penetration test (CPT). Several
and understand soil liquefaction. The progress of work on phenomena are described as soil liquefaction. In an effort to
soil liquefaction has been described in detail in a series of clarify the different phenomena, the mechanisms will be
state-of-the-art papers, such as those by Yoshimi et al. briefly described and definitions for soil liquefaction will be
(1977), Seed (1979), Finn (1981), Ishihara (1993), and Rob- reviewed.
ertson 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 Before describing methods to evaluate liquefaction poten-
tial, 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
Received April 7, 1997. Accepted March 5, 1998. soil loaded in undrained monotonic triaxial compression. In
P.K. Robertson and C.E. (Fear) Wride. Geotechnical void ratio (e) and mean normal effective stress (pʹ) space, a
Group, University of Alberta, Edmonton, AB T6G 2G7, soil with an initial void ratio higher than the ultimate state
Canada. e-mail: pkrobertson@civil.ualberta.ca line (USL) will strain soften (SS) at large strains, eventually

Can. Geotech. J. 35: 442–459 (1998) © 1998 NRC Canada


Robertson and Wride 443

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

reaching an ultimate condition often referred to as critical or cases conventional triaxial equipment may not reach these
steady state. The term ultimate state (US) is used here, after large strains (axial strain, εa > 20%).
Poorooshasb and Consoli (1991). However, a soil with an During cyclic undrained loading (e.g., earthquake load-
initial void ratio lower than the USL will strain harden (SH) ing), almost all saturated cohesionless soils develop positive
at large strains towards its ultimate state. It is possible to pore pressures due to the contractive response of the soil at
have a soil with an initial void ratio higher than but close to small strains. If there is shear stress reversal, the effective
the USL. For this soil state, the response can show limited stress state can progress to the point of essentially zero ef-
strain softening (LSS) to a quasi steady state (QSS) (Ishihara fective stress, as illustrated in Fig. 2. For shear stress rever-
1993), but eventually, at large strains, the response strain sal to occur, ground conditions must be generally level or
hardens to the ultimate state. For some sands, very large gently sloping; however, shear stress reversal can occur in
strains are required to reach the ultimate state, and in some steeply sloping ground if the slope is of limited height

© 1998 NRC Canada


444 Can. Geotech. J. Vol. 35, 1998

Fig. 2. Schematic of undrained cyclic behaviour of sand illustrating cyclic liquefaction (after Robertson 1994). qcy, cycling shear stress.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
For personal use only.

(Pando and Robertson 1995). When a soil element reaches of a strain-softening or strain-hardening response. If the soil
the condition of essentially zero effective stress, the soil has is strain softening, flow liquefaction is possible if the soil
very little stiffness and large deformations can occur during can be triggered to collapse and if the gravitational shear
cyclic loading. However, when cyclic loading stops, the de- stresses are larger than the ultimate or minimum strength.
formations essentially stop, except for those due to local The trigger mechanism can be either monotonic or cyclic.
pore-pressure redistribution. If there is no shear stress rever- Whether a slope or soil structure will fail and slide will de-
sal, such as in steeply sloping ground subjected to moderate pend on the amount of strain-softening soil relative to
cyclic loading, the stress state may not reach zero effective strain-hardening soil within the structure, the brittleness of
stress. As a result, only cyclic mobility with limited defor- the strain-softening soil, and the geometry of the ground.
mations will occur, provided that the initial void ratio of the The resulting deformations of a soil structure with both
sand is below the USL and the large strain response is strain strain-softening and strain-hardening soils will depend on
hardening (i.e., the material is not susceptible to a cata- many factors, such as distribution of soils, ground geometry,
strophic flow slide). However, shear stress reversal in the amount and type of trigger mechanism, brittleness of the
level ground area beyond the toe of a slope may lead to strain-softening soil, and drainage conditions. Soils that are
overall failure of the slope due to softening of the soil in the only temporarily strain softening (i.e., experience a mini-
toe region. mum strength before dilating to US) are not as dangerous as
Based on the above description of soil behaviour in un- very loose soils that can strain soften directly to ultimate
drained shear, Robertson and Fear (1995), building on ear- state. Examples of flow liquefaction failures are Fort Peck
lier work by Robertson (1994), proposed specific definitions Dam (Casagrande 1965), Aberfan flowslide (Bishop 1973),
of soil liquefaction which distinguished between flow lique- Zealand flowslide (Koppejan et al. 1948), and the Stava tail-
faction (strain-softening behaviour; see Fig. 1) from cyclic ings dam. In general, flow liquefaction failures are not com-
softening. Cyclic softening was further divided into cyclic mon; however, when they occur, they take place rapidly
liquefaction (see Fig. 2) and cyclic mobility. A full descrip- with little warning and are usually catastrophic. Hence, the
tion of these definitions is given by Robertson and Wride design against flow liquefaction should be carried out cau-
(1998) in the NCEER report (Youd and Idriss 1998). Fig- tiously.
ure 3 presents a flow chart (after Robertson 1994) for the If the soil is strain hardening, flow liquefaction will gen-
evaluation of liquefaction according to these definitions. The erally not occur. However, cyclic softening can occur due to
first step is to evaluate the material characteristics in terms cyclic undrained loading, such as earthquake loading. The

© 1998 NRC Canada


Robertson and Wride 445

Fig. 3. Suggested flow chart for evaluation of soil liquefaction soil is sufficiently loose, the ultimate undrained strength
(after Robertson 1994). 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 cy-
clic loading, since it is the inertial forces that drive the phe-
nomenon. 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 charac-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

terization.
The most common form of soil liquefaction observed in
the field has been cyclic softening due to earthquake load-
ing. Much of the existing research work on soil liquefaction
has been related to cyclic softening, primarily cyclic lique-
faction. Cyclic liquefaction generally applies to level or
gently sloping ground where shear stress reversal occurs
during earthquake loading. This paper is concerned primar-
ily with cyclic liquefaction due to earthquake loading and its
evaluation using results of the CPT.
For personal use only.

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
amount and extent of deformations during cyclic loading outcome of these studies generally confirmed that the resis-
will depend on the density of the soil, the magnitude and du- tance to cyclic loading is influenced primarily by the state of
ration of the cyclic loading, and the extent to which shear the soil (i.e., void ratio, effective confining stresses, and soil
stress reversal occurs. If extensive shear stress reversal oc- structure) and the intensity and duration of the cyclic load-
curs, it is possible for the effective stresses to reach zero ing (i.e., cyclic shear stress and number of cycles), as well
and, hence, cyclic liquefaction can take place. When the as the grain characteristics of the soil. Soil structure incorpo-
condition of essentially zero effective stress is achieved, rates features such as fabric, age, and cementation. Grain
large deformations can result. If cyclic loading continues, characteristics incorporate features such as grain-size distri-
deformations can progressively increase. If shear stress re- bution, grain shape, and mineralogy.
versal does not take place, it is generally not possible to Resistance to cyclic loading is usually represented in
reach the condition of zero effective stress and deformations terms of a cyclic stress ratio that causes cyclic liquefaction,
will be smaller, i.e., cyclic mobility will occur. Examples of termed cyclic resistance ratio (CRR). The point of “liquefac-
cyclic softening were common in the major earthquakes in tion” in a cyclic laboratory test is typically defined as the
Niigata in 1964 and Kobe in 1995 and manifested in the time at which the sample achieves a strain level of either 5%
form of sand boils, damaged lifelines (pipelines, etc.), lateral double-amplitude axial strain in a cyclic triaxial test or
spreads, slumping of small embankments, settlements, and 3–4% double-amplitude shear strain in a cyclic simple shear
ground-surface cracks. If cyclic liquefaction occurs and test. For cyclic simple shear tests, CRR is the ratio of the cy-
drainage paths are restricted due to overlying less permeable clic shear stress to cause cyclic liquefaction to the initial
layers, the sand immediately beneath the less permeable soil vertical effective stress, i.e., (CRR)ss = τcyc/σʹvo. For cyclic
can loosen due to pore-water redistribution, resulting in pos- triaxial tests, CRR is the ratio of the maximum cyclic shear
sible subsequent flow liquefaction, given the right geometry. stress to cause cyclic liquefaction to the initial effective con-
Both flow liquefaction and cyclic liquefaction can cause fining stress, i.e., (CRR)tx = σdc/2σʹ3c . The two tests impose
very large deformations. Hence, it can be very difficult to different loading conditions and the CRR values are not
clearly identify the correct phenomenon based on observed equivalent. Cyclic simple shear tests are generally consid-
deformations following earthquake loading. Earth- ered to be better than cyclic triaxial tests at closely repre-
quake-induced flow liquefaction movements tend to occur senting earthquake loading for level ground conditions.
after the cyclic loading ceases due to the progressive nature However, experience has shown that the (CRR)ss can be esti-
of the load redistribution. However, if the soil is sufficiently mated quite well from (CRR)tx, and correction factors have
loose and the static shear stresses are sufficiently large, the been developed (Ishihara 1993). The CRR is typically taken
earthquake loading may trigger essentially spontaneous liq- at about 15 cycles of uniform loading to represent an equiva-
uefaction within the first few cycles of loading. Also, if the lent earthquake loading of magnitude (M) 7.5, i.e., CRR7.5.

© 1998 NRC Canada


446 Can. Geotech. J. Vol. 35, 1998

Fig. 4. Graphical representation of liquefaction criteria for silts situ. Unfortunately, it is very difficult to determine the in
and clays from studies by Seed et al. (1973) and Wang (1979) in situ fabric of natural sands below the water table. As a re-
China (after Marcuson et al. 1990): < 15% finer than 0.005 mm, sult, there is often some uncertainty in the evaluation of
liquid limit (LL) < 35%, and water content > 0.9 × liquid limit. CRR based on laboratory testing of reconstituted samples,
although, as suggested by Tokimatsu and Hosaka (1986), ei-
ther the small strain shear modulus or shear wave velocity
measurements could be used to improve the value of labora-
tory testing on reconstituted samples of sand. Therefore,
there has been increasing interest in testing high-quality un-
disturbed samples of sandy soils under conditions represen-
tative of those in situ. Yoshimi et al. (1989) showed that
aging and fabric had a significant influence on the CRR of
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

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 re-
sults of testing in situ frozen samples. The difference in
CRR became more pronounced as the density of the sand in-
The CRR for any other size earthquake can be estimated creased.
using the following equation: Based on the above observations, for high-risk projects
when evaluation of the potential for soil liquefaction due to
[1] CRR =( CRR7.5 ) (MSF) earthquake loading is very important, consideration should
where MSF is the magnitude scaling factor (recommended be given to a limited amount of appropriate laboratory tests
For personal use only.

values are provided in the report by NCEER (Youd and on high-quality undisturbed samples. Recently, in situ
Idriss 1998), which summarizes the results of the 1996 ground freezing has been used to obtain undisturbed samples
NCEER workshop). of sandy soils (Yoshimi et al. 1978, 1989, 1994; Sego et al.
When a soil is fine grained or contains some fines, some 1994; Hofmann et al. 1995; Hofmann 1997). Cyclic simple
cohesion or adhesion can develop between the fine particles shear tests are generally the most appropriate tests, although
making the soil more resistant at essentially zero effective cyclic triaxial tests can also give reasonable results.
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 prop- The above comments have shown that testing high-quality
erties influencing CRR is the plasticity index of the fines undisturbed samples will give better results than testing poor
contained in the sand (Ishihara and Koseki 1989). Ishihara quality samples. However, obtaining high-quality undis-
(1993) showed that the (CRR)tx appears to increase with in- turbed samples of saturated sandy soils is very difficult and
creasing plasticity index. Studies in China (Wang 1979) sug- expensive and can only be carried out for large projects for
gest that the potential for cyclic liquefaction in silts and which the consequences of liquefaction may result in large
clays is controlled by grain size, liquid limit, and water con- costs. Therefore, there will always be a need for simple, eco-
tent. The interpretation of this criterion as given by nomic procedures for estimating the CRR of sandy soils,
Marcuson et al. (1990) and shown in Fig. 4 can be useful; particularly for low-risk projects and the initial screening
however, it is important to note that it is based on limited stages of high-risk projects. Currently, the most popular sim-
data and should be used with caution. Figure 4 suggests that ple method for estimating CRR makes use of penetration re-
when a soil has a liquid limit less than 35% combined with a sistance from the standard penetration test (SPT), although,
water content greater than 90% of the liquid limit, it is un- more recently, the CPT has become very popular because of
clear if the soil can experience cyclic liquefaction, and the its greater repeatability and the continuous nature of its pro-
soil should be tested to clarify the expected response to un- file.
drained cyclic loading. The late Professor H.B. Seed and his coworkers devel-
Although void ratio (relative density) has been recognized oped a comprehensive SPT-based approach to estimate the
as a dominant factor influencing the CRR of sands, studies potential for cyclic softening due to earthquake loading. The
by Ladd (1974), Mulilis et al. (1977), and Tatsuoka et al. approach requires an estimate of the cyclic stress ratio
(1986) have clearly shown that sample preparation (i.e., soil (CSR) profile caused by a design earthquake. This is usually
fabric) also plays an important role. This is consistent with done based on a probability of occurrence for a given earth-
the results of monotonic tests at small to intermediate strain quake. A site-specific seismicity analysis can be carried out
levels. Hence, if results are to be directly applied with any to determine the design CSR profile with depth. A simplified
confidence, it is important to conduct cyclic laboratory tests method to estimate CSR was also developed by Seed and
on reconstituted samples with a structure similar to that in Idriss (1971) based on the maximum ground surface acceler-

© 1998 NRC Canada


Robertson and Wride 447

Fig. 5. Comparison between three CPT-based charts for tails of the SPT to avoid or at least minimize the effects of
estimating cyclic resistance ratio (CRR) for clean sands (after some of the major factors, the most important of which is
Ishihara 1993). D50, average grain size; FC, fines content. 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 pro-
posed to estimate CRR for clean sands and silty sands using
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

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 cor-
rected 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
⎛q ⎞
[3] q c1N = ⎜⎜ c ⎟ CQ = q c1
⎟ Pa2
⎝ Pa2 ⎠
For personal use only.

where qc is the measured cone tip penetration resistance; CQ


= (Pa/σʹvo)n is a correction for overburden stress; the expo-
nent n is typically equal to 0.5; Pa is a reference pressure in
the same units as σʹvo (i.e., Pa = 100 kPa if σʹvo 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 nor-
malized cone penetration resistance, qc1N, is dimensionless.
ation (amax) at the site. This simplified approach can be sum- Robertson and Campanella (1985) developed a chart for
marized as follows: estimating CRR from corrected CPT penetration resistance
(qc1) based on the Seed et al. (1985) SPT chart and
τav ⎛ a ⎞⎛ σ ⎞ SPT–CPT conversions. Other similar CPT-based charts were
[2] CSR = . ⎜ max ⎟ ⎜ vo ⎟ rd
= 065
σʹvo ⎜ g ⎟ ⎜ σʹ ⎟ also developed by Seed and de Alba (1986), Shibata and
⎝ ⎠ ⎝ vo ⎠
Teparaska (1988), and Mitchell and Tseng (1990). A com-
where τav is the average cyclic shear stress; amax is the maxi- parison between three of these CPT charts is shown in
mum horizontal acceleration at the ground surface; g = Fig. 5. In recent years, there has been an increase in avail-
9.81 m/s2 is the acceleration due to gravity; σvo and σʹvo are able field performance data, especially for the CPT (Ishihara
the total and effective vertical overburden stresses, respec- 1993; Kayen et al. 1992; Stark and Olson 1995; Suzuki et al.
tively; and rd is a stress-reduction factor which is dependent 1995b). The recent field performance data have shown that
on depth. Details about the rd factor are summarized by the existing CPT-based correlations to estimate CRR are
Robertson and Wride (1998), based on the recommendations generally good for clean sands. The recent field performance
by Seed and Idriss (1971) and Liao and Whitman (1986). data show that the correlation between CRR and qc1N by
The CSR profile from the earthquake can be compared to Robertson and Campanella (1985) for clean sands provides a
the estimated CRR profile for the soil deposit, adjusted to reasonable estimate of CRR. Based on discussions at the
the same magnitude using eq. [1]. At any depth, if CSR is 1996 NCEER workshop, the curve by Robertson and
greater than CRR, cyclic softening (liquefaction) is possible. Campanella (1985) has been adjusted slightly at the lower
This approach is the most commonly used technique in most end to be more consistent with the SPT curve. The resulting
parts of the world for estimating soil liquefaction due to recommended CPT correlation for clean sand is shown in
earthquake loading. Fig. 6. Included in Fig. 6 are suggested curves of limiting
The approach based on the SPT has many problems, pri- shear strain, similar to those suggested by Seed et al. (1985)
marily due to the inconsistent nature of the SPT. The main for the SPT. Occurrence of liquefaction is based on level
factors affecting the SPT have been reviewed (e.g., Seed et ground observations of surface manifestations of cyclic liq-
al. 1985; Skempton 1986; Robertson et al. 1983) and are uefaction. For loose sand (i.e., qc1N < 75) this could involve
summarized by Robertson and Wride (1998). It is highly large deformations resulting from a condition of essentially
recommended that the engineer become familiar with the de- zero effective stress being reached. For denser sand (i.e.,

© 1998 NRC Canada


448 Can. Geotech. J. Vol. 35, 1998

Fig. 6. Recommended cyclic resistance ratio (CRR) for clean Fig. 7. Summary of variation of cyclic resistance ratio (CRR)
sands under level ground conditions based on CPT. γl, limiting with fines content based on CPT field performance data (after
shear strain. Stark and Olson 1995).
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
For personal use only.

hence, the authors consider the suggested curve to be consis-


qc1N > 75) this could involve the development of large pore tent with field observations. It is important to note that the
pressures, but the effective stress may not fully reduce to simplified approach based on either the SPT or the CPT has
zero and deformations may not be as large as those in loose many uncertainties. The correlations are empirical and there
sands. Hence, the consequences of liquefaction will vary de- is some uncertainty over the degree of conservatism in the
pending on the soil density as well as the size and duration correlations as a result of the methods used to select repre-
of loading. An approximate equation for the clean sand CPT sentative values of penetration resistance within the layers
curve shown in Fig. 6 is given later in this paper (eq. [8]). assumed to have liquefied. A detailed review of the CPT
Based on data from 180 sites, Stark and Olson (1995) also data, similar to those carried out by Liao and Whitman
developed a set of correlations between CRR and cone tip (1986) and Fear and McRoberts (1995) on SPT data, would
resistance for various sandy soils based on fines content and be required to investigate the degree of conservatism con-
mean grain size, as shown in Fig. 7, in terms of qc1N. The tained in Figs. 6 and 7. The correlations are also sensitive to
CPT combined database is now larger than the original the amount and plasticity of the fines within the sand.
SPT-based database proposed by Seed et al. (1985). For the same CRR, SPT or CPT penetration resistance in
The field observation data used to compile the CPT data- silty sands is smaller because of the greater compressibility
base are apparently based on the following conditions, simi- and decreased permeability of silty sands. Therefore, one
lar in nature to those for the SPT-based data: Holocene age, reason for the continued use of the SPT has been the need to
clean sand deposits; level or gently sloping ground; magni- obtain a soil sample to determine the fines content of the
tude M = 7.5 earthquakes; depth range from 1 to 15 m (3–45 soil. However, this has been offset by the poor repeatability
ft) (84% is for depths < 10 m (30 ft)); and representative av- of SPT data. With the increasing interest in the CPT due to
erage CPT qc values for the layer that was considered to its greater repeatability, several researchers (e.g., Robertson
have experienced cyclic liquefaction. and Campanella 1985; Olsen 1988; Olsen and Malone 1988;
Caution should be exercised when extrapolating the CPT Olsen and Koester 1995; Suzuki et al. 1995a, 1995b; Stark
correlation to conditions outside of the above range. An im- and Olson 1995; Robertson and Fear 1995) have developed
portant feature to recognize is that the correlation appears to a variety of approaches for evaluating cyclic liquefaction po-
be based on average values for the inferred liquefied layers. tential using CPT results. It is now possible to estimate grain
However, the correlation is often applied to all measured characteristics such as apparent fines content and grain size
CPT values, which include low values below the average, as from CPT data and incorporate this directly into the evalua-
well as low values as the cone moves through soil layer in- tion of liquefaction potential. Robertson and Fear (1995)
terfaces. Therefore, the correlation can be conservative in recommended an average correction, which was dependent
variable deposits where a small part of the CPT data could on apparent fines content, but not on penetration resistance.
indicate possible liquefaction. Although some of the re- This paper provides modifications to and an update of the
corded case histories show liquefaction below the suggested CPT approach suggested by Robertson and Fear (1995). The
curve in Fig. 6, the data are based on average values and, proposed equation to obtain the equivalent clean sand nor-

© 1998 NRC Canada


Robertson and Wride 449

Fig. 8. Normalized CPT soil behaviour type chart, as proposed Based on extensive field data and experience, it is possi-
by Robertson (1990). Soil types: 1, sensitive, fine grained; 2, ble to estimate grain characteristics directly from CPT re-
peats; 3, silty clay to clay; 4, clayey silt to silty clay; 5, silty sults using the soil behaviour type chart shown in Fig. 8.
sand to sandy silt; 6, clean sand to silty sand; 7, gravelly sand to The boundaries between soil behaviour type zones 2–7 can
dense sand; 8, very stiff sand to clayey sand (heavily be approximated as concentric circles (Jefferies and Davies
overconsolidated or cemented); 9, very stiff, fine grained 1993). The radius of each circle can then be used as a soil
(heavily overconsolidated or cemented). OCR, overconsolidation behaviour type index. Using the CPT chart by Robertson
ratio; ϕʹ, friction angle. (1990), the soil behaviour type index, Ic, can be defined as
follows:

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


. ) 2 ]0.5
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

n
⎛ q − σvo ⎞ ⎛ Pa ⎞
where Q = ⎜ c ⎟⎜ ⎟
⎜ P ⎟ ⎜ σʹ ⎟
⎝ a2 ⎠ ⎝ vo ⎠
is the normalized CPT penetration resistance
(dimensionless); the exponent n is typically equal to 1.0; F =
[fs/(qc – σvo)]100 is the normalized friction ratio, in percent;
fs is the CPT sleeve friction stress; σvo and σʹvo are the total
and effective overburden stresses, respectively; Pa is a refer-
ence pressure in the same units as σʹvo (i.e., Pa = 100 kPa if
σʹvo is in kPa); and Pa2 is a reference pressure in the same
units as qc and σvo (i.e., Pa2 = 0.1 MPa if qc and σvo are
in MPa).
The soil behaviour type chart by Robertson (1990) uses a
For personal use only.

normalized cone penetration resistance (Q) based on a sim-


ple 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 resis-
tance (qc1N) based on a stress exponent n = 0.5 (see eq. [3]).
Olsen and Malone (1988) correctly suggested a normaliza-
tion 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 recom-
malized CPT penetration resistance, (qc1N)cs, is a function of mended for soil classification in clay type soils when Ic >
both the measured penetration resistance, qc1N, and the grain 2.6. However, in sandy soils when Ic ≤ 2.6, it is recom-
characteristics of the soil, as follows: mended that data being plotted on the Robertson chart be
modified by using n = 0.5. Hence, the recommended proce-
[4] (q c1N ) cs = Kc q c1N
dure is to first use n = 1.0 to calculate Q and, therefore, an
where Kc is a correction factor that is a function of the grain initial value of Ic for CPT data. If Ic > 2.6, the data should be
characteristics of the soil, as described later in this paper. plotted directly on the Robertson chart (and assume qc1N =
Q). However, if Ic ≤ 2.6, the exponent to calculate Q should
Grain characteristics from the CPT be changed to n = 0.5 (i.e., essentially calculate qc1N using
In recent years, charts have been developed to estimate eq. [3], since σvo < < qc) and Ic should be recalculated based
soil type from CPT data (Olsen and Malone 1988; Olsen and on qc1N and F. If the recalculated Ic remains less than 2.6, the
Koester 1995; Robertson and Campanella 1988; Robertson data should be plotted on the Robertson chart using qc1N
1990). Experience has shown that the CPT friction ratio (ra- based on n = 0.5. If, however, Ic iterates above and below a
tio of the CPT sleeve friction to the cone tip resistance) in- value of 2.6, depending which value of n is used, a value of
creases with increasing fines content and soil plasticity. n = 0.75 should be selected to calculate qc1N (using eq. [3])
Hence, grain characteristics such as apparent fines content of and plot data on the Robertson chart. Note that if the in situ
sandy soils can be estimated directly from CPT data using effective overburden stresses are in the order of 50–150 kPa,
any of these soil behaviour charts, such as that by Robertson the choice of normalization has little effect on the calculated
(1990) shown in Fig. 8. As a result, the measured penetra- normalized penetration resistance.
tion resistance can be corrected to an equivalent clean sand The boundaries of soil behaviour type are given in terms
value. The addition of pore-pressure data can also provide of the index, Ic, as shown in Table 1. The soil behaviour type
valuable additional guidance in estimating fines content. index does not apply to zones 1, 8, or 9. Along the normally
Robertson et al. (1992) suggested a method for estimating consolidated region in Fig. 8, soil behaviour type index in-
fines content based on the rate of pore-pressure dissipation creases with increasing apparent fines content and soil plas-
(t50) during a pause in the CPT. ticity, and the following simplified relationship is suggested:

© 1998 NRC Canada


450 Can. Geotech. J. Vol. 35, 1998

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 7 Gravelly sand to dense sand
1.31 < Ic < 2.05 6 Sands: clean sand to silty sand
2.05 < Ic < 2.60 5 Sand mixtures: silty sand to sandy silt
2.60 < Ic < 2.95 4 Silt mixtures: clayey silt to silty clay
2.95 < Ic < 3.60 3 Clays: silty clay to clay
Ic > 3.60 2 Organic soils: peats

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

eq. [6] to sands that plot in the region defined by 1.64 < Ic <
[6a] if I c < 126
. apparent fines content FC (%) = 0
2.36 and F < 0.5% in Fig. 8 so as not to confuse very loose
[6b] if 126
. ≤ I c ≤ 35
. clean sands with denser sands containing fines. In this zone,
it is suggested that the apparent fines content is set equal to
. I c3.25 − 3.7
apparent fines content FC (%) = 175 5%, such that no correction will be applied to the measured
CPT tip resistance when the CPT data plot in this zone. To
[6c] if I c > 35
. apparent fines content FC (%) = 100 evaluate the correlation shown in Fig. 9 it is important to
show the complete soil profile (CPT and samples, e.g., see
The range of potential correlations is illustrated in Fig. 9,
Fig. 13), since comparing soil samples with an adjacent CPT
which shows the variation of soil behaviour type index (Ic)
at the same elevation can be misleading due to soil strati-
with apparent fines content and the effect of the degree of
graphic changes and soil heterogeneity.
plasticity of the fines. The recommended general relation-
Based on the above method for estimating grain character-
ship given in eq. [6] is also shown in Fig. 9. Note that this
istics directly from the CPT using the soil behaviour index
equation is slightly modified from the original work by Rob-
(Ic), the recommended relationship between Ic and the cor-
ertson and Fear (1995) to increase the prediction of apparent
rection factor Kc is shown in Fig. 10 and given by the fol-
FC for a given value of Ic.
lowing equations:
The proposed correlation between CPT soil behaviour in-
dex (Ic) and apparent fines content is approximate, since the [7a] if I c ≤ 164
. , Kc = 10
.
CPT responds to many other factors affecting soil behaviour,
such as soil plasticity, mineralogy, sensitivity, and stress his- [7b] if I c ≤ 164
. , Kc = −0.403I c4 + 5581
. I c3 − 2163
. I c2
tory. However, for small projects, the above correlation pro-
vides a useful guide. Caution must be taken in applying +33.75I c − 1788
.
© 1998 NRC Canada
Robertson and Wride 451

Fig. 10. Recommended grain characteristic correction to obtain clean sand equivalent CPT penetration resistance in sandy soils.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
For personal use only.

The proposed correction factor, Kc, is approximate, since Figure 11 shows the resulting equivalent CRR curves for
the CPT responds to many factors, such as soil plasticity, Ic values of 1.64, 2.07, and 2.59 which represent approxi-
fines content, mineralogy, soil sensitivity, and stress history. mate apparent fines contents of 5, 15, and 35%, respectively.
However, for small projects or for initial screening on larger
projects, the above correlation provides a useful guide. Cau- Influence of thin layers
tion must be taken in applying the relationship to sands that A problem associated with the interpretation of penetra-
plot in the region defined by 1.64 < Ic < 2.36 and F ≤ 0.5% tion tests in interbedded soils occurs when thin sand layers
so as not to confuse very loose clean sands with sands con- are embedded in softer deposits. Theoretical as well as labo-
taining fines. In this zone, it is suggested that the correction ratory studies show that the cone resistance is influenced by
factor Kc be set to a value of 1.0 (i.e., assume that the sand the soil ahead of and behind the penetrating cone. The cone
is a clean sand). will start to sense a change in soil type before it reaches the
Note that the relationship between the recommended cor- new soil and will continue to sense the original soil even
rection factor, Kc, and soil behaviour type index, Ic, is shown when it has entered a new soil. As a result, the CPT will not
in Fig. 10 as a broken line beyond an Ic of 2.6, which corre- always measure the correct mechanical properties in thinly
sponds to an approximate apparent fines content of 35%. interbedded soils. The distance over which the cone tip
Soils with Ic > 2.6 fall into the clayey silt, silty clay, and senses an interface increases with increasing soil stiffness. In
clay regions of the CPT soil behaviour chart (i.e., zones 3 soft soils, the diameter of the sphere of influence can be as
and 4). When the CPT indicates soils in these regions (Ic > small as two to three cone diameters, whereas in stiff soils
2.6), samples should be obtained and evaluated using criteria the sphere of influence can be up to 20 cone diameters.
such as those shown in Fig. 4. It is reasonable to assume, in Hence, the cone resistance can fully respond (i.e., reach full
general, that soils with Ic > 2.6 are nonliquefiable and that value within the layer) in thin soft layers better than in thin
the correction Kc could be large. Soils that fall in the lower stiff layers. Therefore care should be taken when interpret-
left region of the CPT soil behaviour chart (Fig. 8), defined ing cone resistance in thin sand layers located within soft
by Ic > 2.6 and F ≤ 1.0%, can be very sensitive and, hence, clay or silt deposits. Based on a simplified elastic solution,
possibly susceptible to both cyclic and (or) flow liquefac- Vreugdenhil et al. (1994) have provided some insight as to
tion. Soils in this region should be evaluated using criteria how to correct cone data in thin layers. Vreugdenhil et al.
such as those shown in Fig. 4 combined with additional test- have shown that the error in the measured cone resistance
ing. within thin stiff layers is a function of the thickness of the
© 1998 NRC Canada
452 Can. Geotech. J. Vol. 35, 1998

Fig. 11. CPT base curves for various values of soil behaviour (qc1N)cs, directly from the measured CPT data. Then, using
index, Ic (corresponding to various apparent fines contents, as the equivalent clean sand normalized penetration resistance
indicated). (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
⎡(q ) ⎤
if 50 ≤ (q c1N ) cs < 160, CRR = 93⎢ c1N cs ⎥ + 0.08
⎣ 1000 ⎦

[8b]
3
⎡(q ) ⎤
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

if (q c1N ) cs < 50, . ⎢ c1N cs ⎥ + 0.05


CRR = 0833
⎣ 1000 ⎦

In summary, eqs. [4]–[7] and [8] can be combined to pro-


vide an integrated method for evaluating the cyclic resis-
tance (M = 7.5) of saturated sandy soils based on the CPT. If
thin layers are present, corrections to the measured tip resis-
tance 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
For personal use only.

in using the proposed integrated method based on the CPT


for evaluating CRR and indicates the recommended equa-
tions for each step of the process.
Although the proposed approach provides a method to
layer as well as the stiffness of the layer relative to that of correct CPT results for grain characteristics, it is not in-
the surrounding softer soil. The relative stiffness of the lay- tended to remove the need for selected sampling. If the user
ers is reflected by the change in cone resistance from the has no previous CPT experience in the particular geologic
soft surrounding soil to the stiff soil in the layer. region, it is important to take carefully selected samples to
Vreugdenhil et al. validated the model with laboratory and evaluate the CPT soil behaviour type classification.
field data. Site-specific modifications are recommended, where possi-
Based on the work by Vreugdenhil et al. (1994), Robert- ble. However, if extensive CPT experience exists within the
son and Fear (1995) suggested a conservative correction fac- geologic region, the modified CPT method with no samples
tor for cone resistance. The corrections apply only to thin can be expected to provide an excellent guide to evaluate
sand layers embedded in thick, fine-grained layers. A full liquefaction potential.
description of the proposed correction factor is given by
Robertson and Wride (1998) in the NCEER report (Youd Application of the proposed CPT method
and Idriss 1998). Thin sand layers embedded in soft clay de- An example of this proposed modified CPT-based method
posits are often incorrectly classified as silty sands based on is shown in Fig. 13 for the Moss Landing site that suffered
the CPT soil behaviour type charts. Hence, a slightly im- cyclic liquefaction during the 1989 Loma Prieta earthquake
proved classification can be achieved if the cone resistance in California (Boulanger et al. 1995, 1997). The measured
is first corrected for layer thickness before applying the clas- cone resistance is normalized and corrected for overburden
sification charts. stress to qc1N and F and the soil behaviour type index (Ic) is
calculated. The final continuous profile of CRR at N = 15
Cyclic resistance from the CPT cycles (M = 7.5) is calculated from the equivalent clean sand
In an earlier section, a method was suggested for estimat- values of qc1N (i.e., (qc1N)cs = Kcqc1N) and eq. [8]. Included
ing apparent fines content directly from CPT results, using in Fig. 13 are measured fines content values obtained from
eq. [6]. Following the traditional SPT approach, the esti- adjacent SPT samples. A reasonable comparison is seen be-
mated apparent fines content could be used to estimate the tween the estimated apparent fines contents and the mea-
correction necessary to obtain the clean sand equivalent pen- sured fines contents. Note that for Ic > 2.6 (i.e., FC > 35%;
etration resistance. However, since other grain characteris- see eq. [6]), the soil is considered to be nonliquefiable; how-
tics also influence the measured CPT penetration resistance, ever, this should be checked using other criteria (e.g.,
it is recommended that the necessary correction be estimated Marcuson et al. 1990) (see Fig. 4). The estimated zones of
from the soil behaviour type index, as described above. soil that are predicted to experience cyclic liquefaction are
Hence, eqs. [4], [5], and [7] can be combined to estimate the very similar to those observed and reported by Boulanger et
equivalent clean sand normalized penetration resistance, al. (1995, 1997).

© 1998 NRC Canada


Robertson and Wride 453

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

The predicted zones of liquefaction are slightly conserva- tion zone and the method predicts a very small zone of pos-
tive compared to the observed ground response. For exam- sible liquefaction. However, given the continuous nature of
ple, when the CPT passes from a sand into a clay (e.g., at a the CPT and the high frequency of data (typically every
depth of 10.5 m), the cone resistance decreases in the transi- 20 mm), these erroneous predictions are easy to identify. In

© 1998 NRC Canada


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
454
during the 1989 Loma Prieta earthquake in California.

For personal use only.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
Can. Geotech. J. Vol. 35, 1998

© 1998 NRC Canada


Robertson and Wride 455

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). σʹv , vertical effective stress.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
For personal use only.

general, given the complexity of the problem, the proposed to determine the normalization. The method by Suzuki et al.
methodology provides a good overall prediction of the po- uses the qc1N normalization suggested in this paper (eq. [3]
tential for liquefaction. Since the methodology should be ap- with n = 0.5). The Olsen and Koester method is very sensi-
plied to low-risk projects and in the initial screening stages tive to small variations in measured friction ratio and the
of high-risk projects, it is preferred that the method pro- user is not able to adjust the correlations based on
duces, in general, slightly conservative results. site-specific experience. The friction sleeve measurement for
the CPT can vary somewhat depending on specific CPT
Comparison with other CPT methods equipment and tolerance details between the cone and the
Olsen (1988), Olsen and Koester (1995), and Suzuki et al. sleeve and, hence, can be subject to some uncertainty. The
(1995b) have also suggested integrated methods to estimate method proposed in this paper is based on field observations
the CRR of sandy soils directly from CPT results with the and is essentially similar to those of Olsen and Koester and
correlations presented in the form of soil behaviour charts. Suzuki et al.; however, the method described here is slightly
The Olsen and Koester method is based on SPT–CPT con- more conservative and the process has been broken down
versions plus some laboratory-based CRR data. The method into its individual components.
by Suzuki et al. is based on limited field observations. The Built into each of the CPT methods for estimating CRR is
methods by Olsen and Koester and Suzuki et al. are shown the step of correcting the measured cone tip resistance to a
in Fig. 14. The Olsen and Koester method uses a variable clean sand equivalent value. It is the size of this correction
normalization technique, which requires an iterative process that results in the largest differences between predicted val-
© 1998 NRC Canada
456 Can. Geotech. J. Vol. 35, 1998

Fig. 15. Approximate comparison of various methods for correcting CPT tip resistance to clean sand equivalent values, based on soil
behaviour type.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15
For personal use only.

ues of CRR from the various methods. Figure 15 provides and Olson are generally smaller than those of the other
an approximate comparison between the methods by Stark methods.
and Olson (1995), Olsen and Koester (1995), and Suzuki et Figure 15 indicates that the recommended correction is
al. (1995b), in terms of Kc from soil behaviour type index, generally more conservative than the corrections proposed
Ic. Also superimposed in Fig. 15 is the recommended rela- by the other authors. The other methods generally predict
tionship, based on the equations given in this paper. The higher values of Kc and suggest that corrections should be
comparisons are approximate because the different authors applied beginning at lower values of Ic, particularly for
use different normalizations when correcting CPT tip resis- higher values of CRR. Note that the recommended relation-
tance for plotting data on a soil classification chart. How- ship between Kc and Ic is shown in Fig. 15 as a broken line
ever, for effective overburden stresses in the order of beyond Ic = 2.6, which corresponds to an apparent FC of
100 kPa (!1 tsf (tsf= ton-force per square foot)), all of the 35% (eq. [6]). This shows that the integrated CPT method
normalization methods should give similar values of normal- for evaluating CRR, as outlined here, does not apply to soils
ized penetration resistance for the same value of measured that would be classified as clayey silt, silty clay, or clay. As
penetration resistance. The boundaries between different soil explained earlier, when interpretation of the CPT indicates
behaviour types are also indicated in Fig. 15, as a guide to that these types of soils are present, samples should be ob-
the type of soil in which corrections of certain magnitudes tained and evaluated using other criteria, such as those given
are suggested by the various methods. in Fig. 4 (Marcuson et al. 1990). It is logical that in
The methods by Suzuki et al. (1995b) and Olsen and nonliquefiable clay soils, the equivalent correction factor,
Koester (1995) appear to be consistent with each other, indi- Kc, could be very large for Ic > 2.6.
cating 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 For low-risk, small-scale projects, the potential for cyclic
trend of the combined Suzuki et al. and Olsen and Koester liquefaction can be estimated using penetration tests such as
lines. The magnitudes of the corrections suggested by Stark the CPT. The CPT is generally more repeatable than the SPT

© 1998 NRC Canada


Robertson and Wride 457

Fig. 16. Summary of liquefaction potential on soil classification eral corrections provide a useful guide. Correcting CPT re-
chart by Robertson (1990). Zone A, cyclic liquefaction possible, sults in thin sand layers embedded in softer fine-grained de-
depending on size and duration of cyclic loading; zone B, posits may also be appropriate. The CPT is generally limited
liquefaction unlikely, check other criteria; zone C, flow to sandy soils with limited gravel contents. In soils with high
liquefaction and (or) cyclic liquefaction possible, depending on gravel contents, penetration may be limited.
soil plasticity and sensitivity as well as size and duration of A summary of the CPT method is shown in Fig. 16, which
cyclic loading. 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
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

in this region should be evaluated using other criteria. Cau-


tion should also be exercised when extrapolating the sug-
gested 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 lay-
ers. However, the correlations are often applied to all mea-
sured CPT values, which include low values below the
average for a given sand deposit and at the interface be-
tween different soil layers. Hence, the correlations could be
conservative in variable stratified deposits where a small
part of the penetration data could indicate possible liquefac-
tion.
As mentioned earlier, it is clearly useful to evaluate CRR
For personal use only.

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 bore-
hole. The CPT provides detailed profiles of cone tip resis-
tance, 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
and is the preferred test, where possible. The CPT provides CPT provides a shear wave velocity profile typically aver-
continuous profiles of penetration resistance, which are use- aged over 1 m intervals and, therefore, contains less detail
ful for identifying soil stratigraphy and for providing contin- than the cone tip resistance profile. However, shear wave ve-
uous profiles of estimated cyclic resistance ratio (CRR). locity is less influenced by grain characteristics and few or
Corrections are required for both the SPT and CPT for grain no corrections are required (Robertson et al. 1992; Andrus
characteristics, such as fines content and plasticity. For the and Stokoe 1998). Shear wave velocity should be measured
CPT, these corrections are best expressed as a function of with care to provide the most accurate results possible, since
soil behaviour type index, Ic, which is affected by a variety the estimated CRR is sensitive to small changes in shear
of grain characteristics. wave velocity. There should be consistency in the liquefac-
For medium- to high-risk projects, the CPT can be useful tion evaluation using either method. If the two methods pro-
for providing a preliminary estimate of liquefaction potential vide different predictions of CRR profiles, samples should
in sandy soils. For higher risk projects, and in cases where be obtained to evaluate the grain characteristics of the soil.
there is no previous CPT experience in the geologic region, A final comment to be made is that a key advantage of the
it is also preferred practice to drill sufficient boreholes adja- integrated CPT method described here is that the algorithms
cent to CPT soundings to verify various soil types encoun- can easily be incorporated into a spreadsheet. As illustrated
tered and to perform index testing on disturbed samples. A by the Moss Landing example presented here, the result is a
procedure has been described to correct the measured cone straightforward method for analyzing entire CPT profiles in
resistance for grain characteristics based on the CPT soil be- a continuous manner. This provides an useful tool for the en-
haviour type index, Ic. The corrections are approximate, gineer to review the potential for cyclic liquefaction across a
since the CPT responds to many factors affecting soil behav- site using engineering judgement.
iour. 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 The authors appreciate the contributions of the members
evaluated and modified to suit a specific site and project. of the 1996 NCEER Workshop on Evaluation of Liquefac-
However, for small-scale, low-risk projects and in the initial tion Resistance of Soils (T.L. Youd, Chair) which was held
screening process for higher risk projects, the suggested gen- in Salt Lake City, Utah. Particular appreciation is extended

© 1998 NRC Canada


458 Can. Geotech. J. Vol. 35, 1998

to the following individuals: W.D.L. Finn, I.M. Idriss, J. ings of the 2nd International Conference on Soil Mechanics and
Koester, S. Liao, W.F. Marcuson III, J.K. Mitchell, R. Foundation Engineering, Rotterdam, Holland. pp. 89–96.
Olsen, R. Seed, and T.L. Youd. Ladd, R.S. 1974. Specimen preparation and liquefaction of sands.
Journal of the Geotechnical Engineering Division, ASCE,
110(GT10): 1180–1184.
Liao, S.S.C., and Whitman, R.V. 1986. A catalog of liquefaction
Andrus, R.D., and Stokoe, K.H. 1998. Guidelines for evaluation of and non-liquefaction occurrences during earthquakes. Research
liquefaction resistance using shear wave velocity. In Proceed- Report, Department of Civil Engineering, Massachusetts Insti-
ings of the 1996 NCEER Workshop on Evaluation of Liquefac- tute of Technology, Cambridge, Mass.
tion Resistance of Soils, Salt Lake City, Utah. Edited by T.L. Marcuson, W.F., III, Hynes, M.E., and Franklin, A.G. 1990. Evalu-
Youd and I.M. Idriss. Report No. NCEER-97-0022. ation and use of residual strength in seismic safety analysis of
Bishop, A.W. 1973. The stability of tips and spoil heaps. Quarterly embankments. Earthquake Spectra, 6(3): 529–572.
Journal of Engineering Geology, 6: 335–376. Mitchell, J.K., and Tseng, D.-J. 1990. Assessment of liquefaction
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

Boulanger, R.W., Idriss, I.M., and Mejia, L.H. 1995. Investigation potential by cone penetration resistance. In Proceedings of the
and evaluation of liquefaction related ground displacements at H. Bolton Seed Memorial Symposium, Berkeley, Calif. Edited
Moss Landing during the 1989 Loma Prieta earthquake. Center by J.M. Duncan. Vol. 2. Bitech Publishers. pp. 335–350.
for Geotechnical Modeling, Department of Civil and Environ- Mogami, T., and Kubo, K. 1953. The behaviour of soil during vi-
mental Engineering, College of Engineering, University of Cali- bration. In Proceedings of the 3rd International Conference on
fornia at Davis, Report UCD/CGM-95/02. Soil Mechanics and Foundation Engineering, Vol. 1.
Boulanger, R.W., Mejia, L.H., and Idriss, I.M. 1997. Liquefaction pp. 152–153.
at Moss Landing during Loma Prieta earthquake. Journal of Mulilis, J.P., Seed, H.B., Chan, C.K., Mitchell, J.K., and
Geotechnical and Geoenvironmental Engineering, 123(5): Arulanandan, K. 1977. Effects of sample preparation on sand
453–467. liquefaction. Journal of the Geotechnical Engineering Division,
Casagrande, A. 1965. The role of the “calculated risk” in earth- ASCE, 103(GT2): 99–108.
work and foundation engineering. The Terzaghi Lecture. Journal Olsen, R.S. 1988. Using the CPT for dynamic response character-
of the Soil Mechanics and Foundations Division, ASCE, 91(4): ization. In Proceedings of the Earthquake Engineering and Soil
1–40. Dynamics II Conference, American Society of Civil Engineers,
For personal use only.

Fear, C.E., and McRoberts, E.C. 1995. Reconsideration of initia- New York. ASCE. pp. 111–117.
tion of liquefaction in sandy soils. Journal of Geotechnical Engi- Olsen, R.S., and Koester, J.P. 1995. Prediction of liquefaction re-
neering, ASCE, 121(3): 249–261. sistance using the CPT. In Proceedings of the International Sym-
Finn, W.D.L. 1981. Liquefaction potential: developments since posium on Cone Penetration Testing, CPT’95, Linkoping,
1976. In Proceedings of the 1st International Conference on Re- Sweden, Vol. 2. SGS. Oct. pp. 251–256.
cent Advances in Geotechnical Earthquake Engineering and Soil
Olsen, R.S., and Malone, P.G. 1988. Soil classification and site
Dynamics, St. Louis, Mo. Vol. 2. Edited by S. Prakash. Univer-
characterization using the cone penetrometer test. In Penetration
sity of Missouri-Rolla. May. pp. 655–681.
testing 1988. De Ruiter Balkema, Rotterdam. ISOPT-1, Vol. 2.
Hofmann, B.A. 1997. In-situ ground freezing to obtain undisturbed
pp. 887–893.
samples of loose sand for liquefaction assessment. Ph.D. thesis,
Pando, M., and Robertson, P.K. 1995. Evaluation of shear stress re-
University of Alberta, Edmonton, Alta.
versal due to earthquake loading for sloping ground. In Proceed-
Hofmann, B.A., Sego, D.C., and Robertson, P.K. 1995. Undis-
ings of the 48th Canadian Geotechnical Conference, Vancouver,
turbed sampling of a deep loose sand deposit using ground
B.C. Vol. 2. CGS. Sept. pp. 955–962.
freezing. In Proceedings of the 47th Canadian Geotechnical
Conference, Halifax, N.S. CGS. Sept. pp. 287–296. Poorooshasb, H.B., and Consoli, N.C. 1991. The ultimate state. In
Ishihara, K. 1993. Liquefaction and flow failure during earth- Proceedings of the 9th Pan-American Conference.
quakes. The 33rd Rankine Lecture. Géotechnique, 43(3): pp. 1083–1090.
351–415. Robertson, P.K. 1990. Soil classification using the cone penetration
Ishihara, K., and Koseki, J. 1989. Cyclic shear strength of test. Canadian Geotechnical Journal, 27(1): 151–158.
fines-containing sands. In Earthquake Geotechnical Engi- Robertson, P.K. 1994. Suggested terminology for liquefaction. In
neering, Discussion Session on Influence of Local Conditions Proceedings of the 47th Canadian Geotechnical Conference,
on Seismic Response, Proceedings of the 12th International Halifax, N.S. CGS. Sept. pp. 277–286.
Conference on Soil Mechanics and Foundation Engineering, Rio Robertson, P.K., and Campanella, R.G. 1985. Liquefaction poten-
de Janeiro. A.A. Balkema, Amsterdam. pp. 101–106. tial of sands using the cone penetration test. Journal of
Jefferies, M.G., and Davies, M.P. 1993. Use of CPTu to estimate Geotechnical Engineering, ASCE, 22(3): 298–307.
equivalent SPT N60. Geotechnical Testing Journal, 16(4): Robertson, P.K., and Campanella, R.G. 1988. Design manual for
458–467. use of CPT and CPTu. Pennsylvania Department of Transporta-
Kayen, R.E., Mitchell, J.K., Lodge, A., Seed, R.B., Nishio, S., and tion.
Coutinho, R. 1992. Evaluation of SPT-, CPT-, and shear Robertson, P.K., and Fear, C.E. 1995. Liquefaction of sands and its
wave-based methods for liquefaction potential assessment using evaluation. Keynote lecture. In IS Tokyo ‘95, Proceedings of the
Loma Prieta data. In Proceedings of the 4th Japan–U.S. Work- 1st International Conference on Earthquake Geotechnical Engi-
shop on Earthquake Resistant Design of Lifeline Facilities and neering. Nov. Edited by K. Ishihara. A.A. Balkema, Amsterdam.
Countermeasures for Soil Liquefaction. Edited by M. Hamada Robertson, P.K., and Wride (née Fear), C.E. 1998. Cyclic liquefac-
and T.D. O’Rourke. Technical Report NCEER-94-0019, Vol. 1, tion and its evaluation based on the SPT and CPT. In Proceed-
pp. 177–204. ings of the 1996 NCEER Workshop on Evaluation of
Koppejan, A.W., Van Wamelen, B.M., and Weinberg, L.J.H. 1948. Liquefaction Resistance of Soils. Edited by T.L. Youd and I.M.
Coastal landslides in the Dutch province of Zealand. In Proceed- Idriss. NCEER-97-0022.

© 1998 NRC Canada


Robertson and Wride 459

Robertson, P.K., Campanella, R.G., and Wightman, A. 1983. Suzuki, Y., Tokimatsu, K., Koyamada, K., Taya, Y., and Kubota, Y.
SPT–CPT correlations. Journal of Geotechnical Engineering, 1995b. Field correlation of soil liquefaction based on CPT data.
ASCE, 109: 1449–1459. In Proceedings of the International Symposium on Cone Pene-
Robertson, P.K., Woeller, D.J., and Finn, W.D.L. 1992. Seismic tration Testing, CPT ’95, Linkoping, Sweden. Vol. 2. SGS. Oct.
cone penetration test for evaluating liquefaction potential under pp. 583–588.
cyclic loading. Canadian Geotechnical Journal, 29: 686–695. Tatsuoka, F., Ochi, K., Fujii, S., and Okamoto, M. 1986. Cyclic un-
Seed, H.B. 1979. Soil liquefaction and cyclic mobility evaluation drained triaxial and torsional shear strength of sands for differ-
for level ground during earthquakes. Journal of the Geotechnical ent sample preparation methods. Soils and Foundations, 26(3):
Engineering Division, ASCE, 105(GT2): 201–255. 23–41.
Seed, H.B., and de Alba, P. 1986. Use of SPT and CPT tests for Terzaghi, K., and Peck, R.B. 1967. Soil mechanics in engineering
evaluating the liquefaction resistance of sands. In Use of in situ practice. 2nd ed. John Wiley & Sons, Inc., New York.
tests in geotechnical engineering. Edited by S.P. Clemence. Tokimatsu, K., and Hosaka, Y. 1986. Effects of sample disturbance
American Society of Civil Engineers, Geotechnical Special Pub- on dynamic properties of sand. Soils and Foundations, 26(1):
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

lication 6, pp. 281–302. 53–64.


Seed, H.B., and Idriss, I.M. 1971. Simplified procedure for evalu- Vreugdenhil, R., Davis, R., and Berrill, J. 1994. Interpretation of
ating soil liquefaction potential. Journal of the Soil Mechanics cone penetration results in multilayered soils. International Jour-
and Foundations Division, ASCE, 97(SM9): 1249–1273. nal for Numerical Methods in Geomechanics, 18: 585–599.
Seed, H.B., Tokimatsu, K., Harder, L.F., and Chung, R. 1985. In- Wang, W. 1979. Some findings in soil liquefaction. Water Conser-
fluence of SPT procedures in soil liquefaction resistance evalua- vancy and Hydroelectric Power Scientific Research Institute,
tions. Journal of Geotechnical Engineering, ASCE, 111(12): Beijing, China.
1425–1445. Yoshimi, Y., Richart, F.E., Prakash, S., Balkan, D.D., and Ilyichev,
Sego, D.C., Robertson, P.K., Sasitharan, S., Kilpatrick, B.L., and Y.L. 1977. Soil dynamics and its application to foundation engi-
Pillai, V.S. 1994. Ground freezing and sampling of foundation neering. In Proceedings of the 9th International Conference on
soils at Duncan Dam. Canadian Geotechnical Journal, 31(6): Soil Mechanics and Foundation Engineering, Tokyo, Vol. 2,
939–950. pp. 605–650.
Shibata, T., and Teparaska, W. 1988. Evaluation of liquefaction po- Yoshimi, Y., Hatanaka, M., and Oh-Oka, H. 1978. Undisturbed
tentials of soils using cone penetration tests. Soils and Founda- sampling of saturated sands by freezing. Soils and Foundations,
For personal use only.

tions, 28(2): 49–60. 18(3): 105–111.


Skempton, A.W. 1986. Standard penetration test procedures and Yoshimi, Y., Tokimatsu, K., and Hosaka, Y. 1989. Evaluation of
the effects in sands of overburden pressure, relative density, par- liquefaction resistance of clean sands based on high-quality un-
ticle size, aging and overconsolidation. Géotechnique, 36(3): disturbed samples. Soils and Foundations, 29(1): 93–104.
425–447. Yoshimi, Y., Tokimatsu, K., and Ohara, J. 1994. In situ liquefaction
Stark, T.D., and Olson, S.M. 1995. Liquefaction resistance using resistance of clean sands over a wide density range.
CPT and field case histories. Journal of Geotechnical Engi- Géotechnique, 44(3): 479–494.
neering, ASCE, 121(12): 856–869. Youd, T.L., and Idriss, I.M. (Editors). 1998. Proceedings of the Na-
Suzuki, Y., Tokimatsu, K., Taye, Y., and Kubota, Y. 1995a. Corre- tional Center for Earthquake Engineering Research (NCEER)
lation between CPT data and dynamic properties of in situ Workshop on Evaluation of Liquefaction Resistance of Soils,
frozen samples. In Proceedings of the 3rd International Confer- Salt Lake City, Utah, January 1996. NCEER-97-0022.
ence on Recent Advances in Geotechnical Earthquake Engi-
neering and Soil Dynamics, St. Louis, Mo. Vol. 1. Edited by S.
Prakash. University of Missouri-Rolla. pp. 210–215.

© 1998 NRC Canada


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

SkyMed: 2012 Emilia (Italy) earthquake. International Journal of Applied Earth Observation and Geoinformation 39, 65-78.
[CrossRef]
5. E. Subaşı 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 2010–2011 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
For personal use only.

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 Pisanò. 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. Efraín Ovando Shelley, Vanessa Mussio, Miguel Rodríguez, José G. Acosta Chang. 2015. Evaluation of soil liquefaction from
surface analysis. Geofísica 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]
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
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

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. Subası 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]
For personal use only.

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

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]
For personal use only.

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]
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. Győri, L. Tóth, Z. Gráczer, 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
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

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 earthquake-
induced soil liquefaction and lateral displacement. Engineering Applications of Artificial Intelligence 24, 142-153. [CrossRef]
For personal use only.

81. Aslı Kurtuluş. 2011. Pipeline Vulnerability of Adapazarı 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, FernandoRodríguez-RoaF. Rodríguez-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.GourvèsR. Gourvès. 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]
94. Dawn A.ShuttleD.A. Shuttle, JohnCunningJ. Cunning. 2008. Reply to the discussion by Robertson on “Liquefaction potential
of silts from CPTu”Appears in Canadian Geotechnical Journal, 45: 140–141.. 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]
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by Curtin University on 07/14/15

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]
For personal use only.

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

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]
For personal use only.

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 Küpper,
T Law, E RF Lord, P A Monahan, N R Morgenstern, R Phillips, R Piché, 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