Вы находитесь на странице: 1из 12

Centrifuge Modeling for Liquefaction Mitigation

Using Colloidal Silica Stabilizer


Carolyn T. Conlee1; Patricia M. Gallagher2; Ross W. Boulanger3; and Ronnie Kamai4

Abstract: This paper reports the results of two centrifuge tests that were conducted to evaluate the effectiveness of colloidal silica for lique-
faction mitigation. Colloidal silica has been selected as a stabilizer material in soils because of its permanence and ability to increase the strength
of soils over time. The centrifuge model geometry was selected to study the effects of lateral spreading in a 4.8-m-thick liquefiable layer overlain
by a silty clay sloping toward a central channel. The centrifuge test evaluates the response of untreated loose sands versus loose sands treated
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

with 9, 5, and 4% colloidal silica concentrations (by weight). The models were subjected to a series of peak horizontal base accelerations ranging
from 0.007 up to 1:3g (prototype) with a testing centrifugal acceleration of 15g. The results show a reduction in both lateral spreading and
settlement in colloidal silica–treated sands versus untreated sands. The shear modulus at low strains was determined from shear wave velocity
measurements for the untreated and treated loose sands. The hysteretic response during cyclic loading was also determined for various levels of
shaking. The results from the centrifuge tests show an increase in cyclic resistance ratios and a decrease in cyclic shear strains for increasing
colloidal silica concentrations. DOI: 10.1061/(ASCE)GT.1943-5606.0000703. © 2012 American Society of Civil Engineers.
CE Database subject headings: Soil liquefaction; Centrifuge models; Strain.
Author keywords: Colloidal silica; Liquefaction; Centrifuge testing; Shear strain.

Introduction characterized in terms of accelerations, displacements, and the hys-


teretic stress-strain response.
Liquefaction is a phenomenon marked by a rapid and dramatic loss
of soil strength that occurs in loose, saturated soil deposits sub- Colloidal Silica Properties
jected to earthquake motions. At developed sites, traditional
ground improvement methods to mitigate the liquefaction risk are CS is an aqueous dispersion of negatively charged, spherical, silica
often infeasible because of site constraints or the potential for nanoparticles (7–22 nm). Gelation is induced by reducing the repulsive
adverse effects on adjacent structures as a result of vibration, forces, resulting in the formation of a coherent network of siloxane (Si-
densification, and increased lateral stresses. Passive site stabilization O-Si) bonds (Iler 1979) that bind the soil particles together and im-
is a noninvasive technique in which a stabilizer is injected at the edge mobilize the pore fluid (Fig. 2). In dilute solution, CS has a low initial
of a site and delivered to target locations through augmented or natural viscosity and a wide range of controllable gel times. It is permanent,
groundwater flow (Fig. 1). As the stabilizer material flows through nontoxic, biologically and chemically inert, and has excellent dura-
the formation, it displaces the existing groundwater within the bility characteristics (Iler 1979; Whang 1995). Gel time decreases with
pore spaces and forms a permanent gel that binds to soil particles, increasing percent silica, decreasing particle size, and increasing ionic
resulting in a stronger, stiffer formation (Gallagher 2000). strength. Higher ionic strengths decrease gel time; cations cause the
Colloidal silica (CS) has been shown to stabilize soils and double layer around the particles to shrink, increasing the probability of
mitigate the effects associated with liquefaction-induced damage. interparticle collisions. With respect to pH, a broad minimum gel time
This paper reports the results of centrifuge model tests to compare is exhibited in a neutral range of 5 , pH , 7 (Gallagher 2000).
the dynamic response of sands treated with 4, 5, and 9% (by weight) The shape of the gel time curve is similar regardless of gel time
of CS with untreated sands. The effects of improvement levels are (Fig. 3). Persoff et al. (1999) provide a chart of gel state descriptions
with corresponding viscosities that range from Gel State 1 (2 cP)
1
Geotechnical Engineer, Geocomp Corporation, 125 Nagog Park, to Gel State 11 (2,500 cP). Gel State 1 is the period during which
Acton, MA 07120 (corresponding author). E-mail: ctc5681@gmail.com the viscosity of the catalyzed solution remains virtually unchanged
2
Professor, Civil, Architectural, and Environmental Engineering, Drexel (2 cP) and behaves as a Newtonian fluid. During Gel State 2, the
Univ., 3141 Chestnut St., Philadelphia, PA 19104. E-mail: pmg24@drexel.edu viscosity gradually begins to increase (up to 10 cP) and becomes
3
Professor, Civil and Environmental Engineering, Univ. of California– a non-Newtonian fluid. At the initiation of Gel State 3, the solution
Davis, One Shields Ave., Davis, CA 95616. E-mail: rwboulanger@ begins a rapid progression through several states until it finally
ucdavis.edu becomes a rigid gel.
4
Graduate Research Assistant, Civil and Environmental Engineering, The strength of sands treated with CS increases with increasing
Univ. of California–Davis, One Shields Ave., Davis, CA 95616. E-mail: concentration of silica in the grout (Persoff et al. 1999; Gallagher and
rkamai@ucdavis.edu
Mitchell 2002). After gelation, the strength continues to increase.
Note. This manuscript was submitted on June 21, 2010; approved on
February 6, 2012; published online on October 15, 2012. Discussion period Persoff et al. (1999) measured the strength for a year after CS
open until April 1, 2013; separate discussions must be submitted for treatment and found it continued to increase gradually during that
individual papers. This paper is part of the Journal of Geotechnical and period. However, Gallagher and Lin (2005) found that the strength
Geoenvironmental Engineering, Vol. 138, No. 11, November 1, 2012. gain occurred during the initial curing or within about four times the
©ASCE, ISSN 1090-0241/2012/11-1334–1345/$25.00. time required to reach the resonating gel state (Gel State 10).

1334 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


The feasibility of CS for liquefaction mitigation was investigated
in centrifuge model tests by Gallagher et al. (2007b) and Pamuk et al.
(2007). For peak accelerations of 0.2 and 0:25g, Gallagher et al.
(2007b) showed significantly lower levels of strains (1/2–1%)
for 6% (by weight) CS soils versus strains recorded in similar
centrifuge tests done on untreated models (3–6%). Pamuk et al.
(2007) compared the effects of 6% by weight CS treatment on pile
foundations embedded in sloping liquefiable sands and showed CS
provided significant liquefaction resistance, greatly reduced the
free-field lateral deformation, and reduced the surface settlement
Fig. 1. Passive site stabilization concept by 90%. A full-scale field test was performed (Gallagher et al.
2007a) to treat the upper 20% of a liquefiable sand layer with 8%
CS and a controlled blasting technique was implemented to induce
liquefaction. The results showed a 30% reduction in settlement in
the treated area compared with that of an adjacent, untreated area.
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

Purpose and Configuration of the Test

The purpose of the centrifuge model tests was (1) to compare the
liquefaction and deformation response of loose sands treated with
various CS concentrations (4, 5, and 9% by weight) with untreated
sands in the liquefiable layer and (2) to characterize the effects of
improvement in terms of accelerations, displacements, and the hys-
teretic stress-strain response. Each centrifuge model consisted of two
Fig. 2. Formation of siloxane bonds as a colloidal silica particle gel slopes that sloped 3 toward a 3-m-wide central channel (Fig. 4). In
(after Moridis et al. 1995) this way, four configurations could be investigated using two models.
The slopes were constructed with the following three distinct layers:
(1) a bottom layer of dense Monterey No. 0/30 sand with a prototype
thickness of 0.75 m overlain by (2) a 4.8-m-thick liquefiable layer of
loose Nevada sand that was overlain by (3) a 1.0-m-thick layer of
compacted Yolo loam (Fig. 4). He data and results are presented in
prototype units hereafter, unless otherwise noted.
In Test 1 (CTC01), one slope was treated with 9% CS (denoted as
CS-9), while the second slope was left untreated (Fig. 4). In Test 2
(CTC02), one slope was treated with 4% CS (denoted as CS-4) and
the second slope was treated with 5% CS (denoted as CS-5) (Fig. 4).
Laboratory tests using CS concentrations of less than 5% have been
shown to gel; however, they do not achieve a rigid gel (i.e., they do
not reach Gel State 11). The purpose of using 4% CS in CTC02
was to compare the effectiveness of liquefaction mitigation for
a weak gel to a strong gel that had achieved the majority of its
strength gain.
Fig. 3. Typical gel time curve of colloidal silica at various normalities

Model Preparation
Use of Colloidal Silica for Liquefaction Mitigation
Soil and Sensor Placement
Laboratory, bench-scale model, and full-scale field tests have
demonstrated the effectiveness of CS treatment against liquefaction Both models were built inside a flexible shear-beam container. The
and liquefaction-induced damage. Gallagher and Mitchell (2002) sand layers were placed in several lifts using a dry pluviation tech-
conducted cyclic triaxial shear tests on loose sand treated with 5– nique with a calibrated pluviator. Each sensor was placed at the
10% by weight CS and the results showed a dramatic increase in desired location during model construction and the model container
deformation resistance. Kodaka et al. (2005) compared the hysteretic was weighed at each lift to confirm the relative density of the soil.
stress-strain behavior of CS-treated sand and untreated sands using The average relative densities in the loose sand layer were estimated
cyclic torsional shear tests and showed the CS-treated sand initially to be 35 and 45% for CTC01 and CTC02, respectively.
demonstrated behavior similar to that of dense sand with significant Accelerometers, pore pressure transducers, linear potentiometers
damping and cyclic mobility with continued loading. (LPs), and linear variable differential transformers (LVDTs) were
The small-strain dynamic behavior of CS sands from resonant placed at the locations shown in Fig. 4. The instrumentation in the
column tests showed higher G for treated soils compared with un- liquefiable layer consisted of 10 piezoelectric accelerometers (PCB
treated soils with minimal differences between CS concentrations Piezotronics, Inc.) and nine pore pressure transducers (PDCR-81
(Spencer et al. 2008). The G for each CS concentration appeared to Series by Druck, Inc.) placed in a vertical array on each side of the
be unaffected up to strains of about 1 3 1023 % and then gradually model. The LPs (Duncan 600 Series by BEI Technologies, Inc.) and
decreased at higher strain amplitudes from about 65 to 45 MPa. the LVDTs (DCTH Series by RDP Group) were placed on the

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1335

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Model geometry and sensor layout

surface of the model to measure the surface settlement and lateral Table 1. Colloidal Silica Properties
spreading. Bender elements (PiezoSystems, Inc.) were placed within Colloidal silica
the liquefiable layer to measure the shear wave velocity. Open steel concentration CS-9a CS-5b CS-4c
tubes were also placed vertically on each side of the model, which
were filled with colored sand and subsequently used to observe the Normality (N) 0.24 0.33 0.27
lateral movement of the soil profile, including the relative movements Time at Gel State 3 (h) 13 15 84
at the interface between the crust and the loose sand after testing. Resonating time (h)b 16 26 120
N* gel time at maximum 4 4 NA
strength gainc
Colloidal Silica Preparation and Saturation N* gel time at time of 10 12 NA
The target gel state of the CS was selected to allow ample time to shaking
saturate the treated zones of the model before the viscosity increased Density (g/cc) 1.06 1.05 1.04
too much and to provide sufficient curing time prior to model testing. Initial viscosity (cP) 1.3 1.15 1.1
a
The gel times of the CS solution were controlled by altering the ionic Viscosity of solution measured by the Cannon Ubbelohde viscometer.
b
strength through the addition of sodium chloride (NaCl). Prior to Refers to Gel State 10 (Persoff et al. 1999).
c
model preparation, CS gel time tests were conducted to determine Defined as 4* resonating time or 4* Gel State 10 (Gallagher and Mitchell
the normality that resulted in the desired gel time for each CS con- 2002).
centration. The curing time to reach Gel State 10 was determined via
visual and manual inspection of laboratory specimens. The prop-
deionized water upon completion of CS-9 and placement of the crust
erties of each CS grout solution are summarized in Table 1.
(described subsequently). In CTC02, the preparation and saturation
The CS solution was prepared in a saturation tank. First, deion-
of CS-5 and CS-4 followed a similar approach as CS-9.
ized water was mixed with the appropriate amount of granular NaCl
Once saturation was complete, the vacuum pressure was slowly
(Fisher Scientific) using an air circulation pump. Red food dye was
released from the model and the height of the fluid was carefully
added to enable visual inspection of the treatment coverage during
monitored. While the vacuum pressure was released, the solution
model dissection. The appropriate volume of Ludox SM-30 was
elevation decreased, indicating the solution was continuing to fill
added and the entire batch was mixed.
in dry pockets within the liquefiable layer. The vacuum release was
Prior to saturation, a vacuum pressure of approximately 90 kPa
stopped and more CS solution was added to the model. The process
was applied and the model was flooded with carbon dioxide gas to
was repeated until no further drop in fluid height was observed to
displace any air trapped in the soil voids. Troughs with saturation
ensure complete saturation. Saturation occurred at 1g; however,
tubes were mounted on the north and south ends of the container. The
grouting with CS under increasing levels of confinement resulted in
saturation tubes extended to the bottom of the model container. The
increasing strength in the sands (Kodaka et al. 2005). The strength
fluid was introduced in the troughs under vacuum pressure from
gains provided by CS likely would have been higher had they been
a deaired water chamber. For CTC01, the plan was to drip the CS
allowed to cure under the testing acceleration of 15g.
solution into the north trough, after which it would flow through the
saturation tubes such that saturation would occur from the bottom up.
When saturation was near completion, the injection rate decreased
Crust Placement and Saturation of the Untreated Side
dramatically and the saturation procedure was modified such that the
solution flowed over the lip of the trough and saturated the model from The crust was constructed using Yolo loam sieved to pass a #10 sieve.
the top down. The untreated side of CTC01 was saturated with The crust was placed at the optimum water content of about 12–15%

1336 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


in several lifts. Each lift was compacted with a model-sized sheepsfoot load cells; thus, the tip force could be measured in addition to the
roller that created a normal stress on the surface of approximately combined effect of the tip resistance and sleeve friction.
9 kPa such that compaction pressure would not exceed the prototype The data from both tests are plotted in Fig. 5 as cone tip re-
pressure on the top of the sand layer. The crust was trimmed to the sistance, qc , which is the penetration force divided by the tip cross-
appropriate sloped geometry and the surface was marked with colored sectional area. The value of qc from CTC01 was greater than the true
sand lines (running east/west) to evaluate the uniformity of the lateral value because it included the shaft friction force; however, the error
deformations and boundary effects on the container after testing. The was estimated to be less than about 1% based on the ratio of the shaft
container was mounted on the arm of the centrifuge and the untreated friction to the tip force in CTC02. The CS treatment in CTC01
side was saturated as described previously. increased qc of the loose sand layer from about 2.5 to 5.0 MPa.
CTC02 shows that qc increased from about 10 MPa on CS-4 to
about 23 MPa on CS-5. The difference in qc between these two sides
Soil Properties was again a factor of about 2, which is attributed to the fact that CS-4
only reached Gel State 7 whereas CS-5 achieved a fully resonated
gel. For similar effective stresses (10–40 kPa), typical qc values of
Overview saturated sands were estimated using the empirical correlations from
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

The soil properties are summarized in Table 2. The mmnimum and


maximum relative densities for Nevada sand were determined using Table 2. Soil Properties
Japanese Standard Method Number JIS A1244. The cyclic stress ratio
(CSR)—defined as the ratio of the cyclic shear stress amplitude to the Parameter Dense sand Loose sand Crust
initial effective confining stress required to cause an excess pore Soil type Monterey 0/30 Nevada sand Yolo loam
pressure ratio of 1.0 in 15 uniform loading cycles for the saturated Specific gravity 2.64 2.64a —
loose sand layer (Dr  40%)—was estimated to be between 0.09 and Mean grain size, D50 (mm) 0.4b 0.17 0.32
0.13 based on torsional hollow cylinder tests (Kano, personal com- Coefficient of uniformity — 1.64 10
munication, 2010) and direct simple shear tests (Doygun 2009). The gmax;dry (kN/m3 ) 16.81b 17.5 —
wet unit weight of the Yolo loam crust was defined as the average unit gmax;dry (kN/m3 ) 13.96 (CTC01) 14.5 —
weight of the Yolo loam samples collected during model preparation. 13.74 (CTC02)b
The water content of the crust layer after shaking was determined by Relative density (%) 95 (CTC01) 35 (CTC01) —
averaging the wet and dry weights of the samples collected. 98 (CTC02) 45 (CTC02)
Wet unit weight (kN/m3 ) 19.6 (CTC01) 19 21.8 (CTC01)
19.7 (CTC02)c 21.5 (CTC02)d
Cone Penetration Tests
Permeability (cm/s) — 2 3 1023 —
One cone penetration test (CPT) was performed in each slope at 15g Water content — — 15% (CTC01)
prior to shaking. A hydraulic actuator was used to push the 0.095-m- 12% (CTC02)
diameter (prototype) cone into the sand at a constant rate of ap- a
Chen (1995) and Cruz (1995).
proximately 1 cm/s. In CTC01, the CPT had a rod length of 16 cm b
Kammerer et al. (2004).
(model scale) and one working load cell at the back to measure the c
Based on emax 5 0:821 and emin 5 0:631 (Polito 1999).
combined effect of the tip resistance and sleeve friction. In CTC02, d
Estimation based on the average of two samples collected during
the CPT had a rod length of 32 cm (model scale) and two separate placement.

Fig. 5. Cone penetration test

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1337

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Salgado et al. (1997) and were found to range from 2 to 4 MPa (Dr 5 The crust layer on the treated side remained virtually intact and
35%) and 3 to 5 MPa (Dr 5 45%). The values of qc for the untreated experienced minimal settlement. The posttesting measurements
sand in CTC01 agree with these estimates; however, the overall revealed cracks on the treated side up to 1.5 cm in the direction
increase in qc between CTC01 and CTC02 is likely a result of the perpendicular to shaking.
combined effect of increased relative density and CS treatment. The visual observations indicated that the surface lines on all CS-
treated sides moved laterally in a uniform manner, suggesting the
crust surface moved as a continuous block with minimal boundary
Model Testing effects. On the untreated side, the surface lines were washed away
when the crust was submerged and movement could not be observed
The tests were conducted on the 9-m-radius centrifuge at the Uni- by visual inspection (Fig. 6). Visual inspection of the vertical colored
versity of California at Davis. Eight shaking events were applied to sand markers showed a portion of the lateral movement occurred as
the models in the north-south (longitudinal) direction at a centrifugal localized shear strains at the interface between the liquefiable sand
acceleration of approximately 15g. Each motion consisted of 20 and crust layer (Fig. 7).
cycles of a sinusoidal wave at a frequency of 2 Hz. The testing and The LP and LVDT readings recorded during each shaking event
shaking sequence are shown in Table 3. were used to compute the settlements and lateral spreading on the
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

crust surface. Figs. 8 and 9 compare the cumulative settlement and


lateral spreading for the treated and untreated zones measured on the
Lateral Spreading and Settlement surface of the model for Shakes 1–5. The positive values for lateral
spreading indicate movement toward the central water channel and
Immediately after testing CTC01, visual inspection showed the crust the negative values for settlement indicate downward movement.
on the untreated side was completely submerged in water as a result The data analysis reveals the surface settlements to be minimal
of the occurrence of large settlements when the underlying sand with CS treatment when compared with no treatment. At the end of
layer liquefied. The sloped crust was completely washed away and
experienced significant lateral spreading and settlement [Fig. 6(b)].

Table 3. Testing Sequence


CTC01 CTC02
a b a
Event PBA Event PBAb
CTC01_01 (Shake 1) 0.007 CTC02_01 (Shake 1) 0.007
CTC01_02 (Shake 2) 0.03 CTC02_02 0.03
(pulse wave test) (Shake 2) (bender test)
CTC01_03 (Shake 3) 0.1 CTC02_03 0.15
(pulse wave test) (Shake 3) (bender test)
CTC01_04 (Shake 4) 0.2 CTC02_04 0.25 Fig. 7. Model profile near the open channel after the test
(pulse wave test) (Shake 4) (bender test)
CTC01_05 (Shake 5) 0.56 CTC02_05 0.69
(pulse wave test) (Shake 5) (bender test)
CTC01_06 (Shake 6) 0.03 CTC02_06 0.03
(pulse wave test) (Shake 6) (bender test)
CTC01_07 (Shake 7) 0.2 CTC02_07 0.25
(pulse wave test) (Shake 7) (bender test)
CTC01_08 (Shake 8) 1.28 CTC02_08 1.37
(pulse wave test) (Shake 8) (bender test)
a
Rest time between events is approximately 15 min and shaking time is 10 s.
b
Peak base acceleration determined by average recordings at the north and
south bases of the container.

Fig. 6. Model surface: (a) before the test; (b) after the test Fig. 8. Cumulative surface settlement for Shakes 1–5

1338 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Shake 5, the cumulative surface settlements for the untreated side end of Shake 5 cumulative settlements ranged from about 0.05
ranged from 0.2 to 0.9 m (Fig. 8), with the greatest settlements to 0.08 m for CS-5 and 0.07 to 0.1 m for CS-4.
occurring closest to the container wall and the smallest settlements Similarly, lateral spreading was reduced with CS treatment
occurring closest to the central water channel. The reason for such compared with no treatment (Fig. 9). At the end of Shake 5 in
large settlements along the wall and smaller settlements near the CTC01, CS-9 reduced the cumulative lateral spreading on the
water channel was a result of the fact that the entire crust layer was model surface from 0.9 to 0.15 m closest to the open channel. In
sloped 3 toward the central water channel. When liquefaction in the CTC02, the lateral spreading was greater for CS-4 compared with
untreated side occurred, water migrated to the surface and slope that of CS-5. The lateral spreading for CS-5 ranged from about 0.02
failure of the crust occurred. At this point, the entire crust layer to 0.1 m while the lateral spreading for CS-4 ranged from about 0.1 to
moved laterally toward the central water channel and the visual 0.2 m. It should be noted that the maximum lateral spreading at the
observations revealed the crust layer was completely flattened channel and midcrust locations for CS-9 was greater than for CS-5.
and filled into the central water channel (Fig. 6). The surface set- This is primarily a result of the fact that in CTC02 the pore fluid on
tlements for CS-9 were relatively uniform, and at the end of Shake 5 both sides of the model consisted of gel while in CTC01 the water on
cumulative settlements of about 0.05, 0.07, and 0.05 m occurred for the untreated side allowed for greater movement throughout the
the boundary, midcrust, and channel locations, respectively. In overall model. Fig. 9 also shows that at the end of Shake 5 the lateral
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

CTC02, the surface settlements were minimal overall because of the spreading on the boundary of CS-4 is slightly greater than that of the
fact that liquefaction did not occur on either side of the model and untreated sand as a result of the significant cracking that occurred
slope failure did not occur on the crust. However, the surface set- along the boundary where the sensor was located.
tlements were slightly greater in CS-4 compared with CS-5. At the
Pore Pressure and Acceleration

The pore pressure was measured on the untreated side of the model
within the liquefiable layer to identify the occurrence of liquefaction.
The pore pressure ratio, rU , was computed for the untreated sand as the
ratio of excess pore pressure to the initial vertical effective stress. The
pore pressure data of the CS soils are not reported in this paper. The
mechanism for generating pressure on CS gel in the sand pores is not
yet well understood and requires further study. Although liquefaction
is conventionally defined as the time when rU 5 1:0, the significance
of rU for CS sands does not represent the same physical meaning.
The acceleration and pore pressure responses at the midpoint of the
untreated liquefiable layer are shown in Fig. 10 for Shakes 2–5. In
Shake 2 the model was subjected to a peak base acceleration (PBA) of
0:03g. Fig. 10 shows that rU remains low and the acceleration reveals
a uniform response similar to the base shaking motion. During Shake 3
(PBA 5 0:1g), liquefaction was initiated 3 s into the shaking event. At
this same time, amplification was observed in the acceleration record
followed by deamplification in the subsequent cycles. This same
pattern was observed during Shakes 4 and 5; however, the amplifi-
cation and deamplification responses became more dramatic. Based on
the acceleration and pore pressure responses in Fig. 10, the untreated
loose sand layer continued to liquefy for successive shaking events.
From the acceleration records, the amplification in the acceleration
response occurred when rU 51:0, followed by deamplification.
Fig. 9. Cumulative lateral spreading for Shakes 1–5 Fig. 11 compares the acceleration response at the midpoint of the
untreated and treated liquefiable layers for Shakes 2–5. Fig. 11 shows

Fig. 10. Acceleration and pore pressure response at the midpoint of the untreated liquefiable layer for Shakes 1–5

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1339

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Acceleration response at the midpoint of the untreated liquefiable layer, CS-9, CS-5, and CS-4 for Shakes 2–5

a uniform response for each soil type during Shake 2 when the applied In CTC01, the VS measurements were planned to be determined
shaking amplitude was small (PBA 5 0:03g). When the shaking using bender elements. However, the bender elements failed; there-
amplitudes were increased in Shakes 3 and 4, the acceleration response fore, an alternative method utilizing the centrifuge shaker was pro-
in the treated soils showed a relatively uniform response in terms of the posed. The shaker was used to generate a single 0.1 s (model scale) sine
acceleration amplitudes compared with that of the untreated soil. On the wave at the base of the model container, which was recorded in vertical
other hand, the response in Shake 5 showed nonuniformity and high- accelerometer arrays (Fig. 4) with a sampling rate of 20,000 Hz. The
frequency amplification spikes in CS-4 and CS-5 compared with CS-9. pulse wave had an average PBA of 0:025g. As the shear wave traveled
In general, the deamplification of the acceleration response ob- through the liquefiable soil, the vertical array of accelerometers picked
served in the untreated sand is directly related to a change in the ratio up the signal and captured the wave propagation through the profile
between input frequency (which was constant) and the soil profile’s (Fig. 12). The distance between the two signal measurement points, L,
natural frequency (which progressively decreased). The significance is the difference in elevation between accelerometers. The travel time,
of these trends indicates that the stiffness in the untreated sand is t, was taken as the time shift of the pulse wave from successive
continuously degrading for each loading cycle. On the other hand, accelerometers in the array using cross-correlation methods. Prior to
the trends in the acceleration response for CS-9 suggest strength and cross correlation, a seventh-order band-pass Butterworth filter was
stiffness is being maintained for cyclic loading up to 0:56g. The applied to the original records to remove higher-frequency noise.
amplification that occurred in CS-5 and CS-4 in Shake 5 suggests In CTC02, bender elements were used to estimate small-strain VS .
changes in the behavior of the grouted medium. The bender element tests used the signal stacking procedure by
Brandenberg et al. (2008). Ten pulses were sent out in quick suc-
Shear Wave Velocity cession and the signals were stacked to improve the signal-to-noise
ratio (SNR). Once the SNR was satisfactory, the multiple waves were
The shear modulus degradation curve is characterized using both the stacked into a single signal. The data were transmitted and received
maximum shear modulus, Gmax , and the variation in modulus ratio, through a data acquisition system and processed from a built-in
G/Gmax , with cyclic strain amplitude; Gmax can be obtained by software program (i.e., LabVIEW). A square source wave (9 V am-
measuring the shear wave velocity (VS ) at low strain amplitudes by plitude) was sent and sampled at a frequency of 90,000 Hz. Here, L is
the tip-to-tip distance from the source to the receiving bender and
Gmax ¼ r × Vs2 ð1Þ t is the time of first arrival of the receiving signal (Fig. 13). The source
and receiver benders were placed approximately 100 mm apart at four
where VS is measured as different elevations in the liquefiable soil (Fig. 4). The duration of the
bender element test was approximately 2 ms.
The values of VS from CTC01 and CTC02 cannot be directly
VS ¼ L ð2Þ
t compared because of the differences in the measurement methods.
The bender element tests typically induced strains lower than 3 3
where L 5 effective distance between two signal measurement points 1024 % while the shear strains induced from the pulse wave were
and t 5 travel time for the shear wave to travel through the medium. determined to range about 0.01%. This difference in strains can

1340 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Table 4. Shear Wave Velocity Summary
After PBA
Type shake (g) Depth (m) s90 (kPa) Vs (m/s) G (MPa)
Untreateda 0.007 2.6 30 174 59
CS-9a 0.007 2.6 30 185 66
CS-5b 0.007 2.66 32 176 60
CS-4b 0.007 2.66 32 151 44
a
Measurement from pulse wave.
b
Measurement from Bender elements.

Stress-Strain Analysis

The interpretation of the stress-strain response based on the accel-


erometer data are based on the method presented in Zeghal et al.
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

(1995) with the appropriate modifications as presented in Kamai and


Fig. 12. Normalized signals of the vertical accelerometer array in Boulanger (2010). The analysis assumes a one-dimensional (1D)
untreated soil shear-beam response, with upward propagation of the shear waves.
Other forms of waves, such as P waves reflected off the container
walls, are assumed to be negligible with respect to the shaking-
induced, upward propagating shear waves. This assumption is be-
lieved to be reasonable, based on the following two main factors: (1)
the test was performed within a flexible shear-beam container, which
deforms with the soil and, hence, is expected to produce smaller
inertia loads on the soil; and (2) the vertical array is located in the
middle of the soil mass, at a prototype distance of about 6 m from the
container wall, to reduce the boundary effects.
The accuracy of the inverse analysis techniques require in-
strument spacing to be less than about one-eighth of the shortest
wavelength. Based upon estimates of VS during dynamic loading, the
maximum wavelength ranged from about 115 m for the smaller
shakes to a minimum of about 10 m for the larger shaking events as
a result of softening. This equated to a maximum sensor spacing of
14 m for the smaller shakes and 1.25 m for the larger shakes. The
maximum sensor spacing was measured at 1.2 m with an average
sensor spacing of about 0.65 m (prototype) and, therefore, con-
sidered dense enough to capture wave transmission accurately.

Fig. 13. Travel time determination from the Bender element tests Calculation of Shear Stress
The shear stresses were determined at the midpoints between the
accelerometers in the vertical arrays (Fig. 14). The calculations fol-
cause a 30% difference in the shear modulus as demonstrated by lowed 1D shear-beam behavior where the shear stress at any depth, z,
Spencer et al. (2008). The step-wave technique used in CTC01 also can be obtained by integration of the soil density, r, times acceler-
had lower resolution, in part because of limits on the sampling ation, a
frequency and input wave frequency. In general, both methods can
ðz
be used to assess trends in G; however, the bender elements offer
a more accurate means to obtain Gmax . tðzÞ ¼ r × a × dz ð3Þ
The VS and G measurements are summarized in Table 4 for CTC01 0
and CTC02, respectively, after a small shaking event (PBA 5 0:007g)
was applied to the model. For CTC01, VS of the untreated sand at 30 Unlike field data, surface accelerations (when z 5 0) cannot be
kPa confinement was measured as 174 m/s. The results are generally directly measured in centrifuge models because the instruments re-
consistent with the studies by Kasantikul (2009), which reported VS  quire sufficient contact with the soil and must be buried within the
160–180 m/s for loose-to-medium sand at 30 kPa confinement. The model. As a result, rigid body motion in the centrifuge was assumed
G for CS-9 showed an increase of about 12% compared with the for the first soil element in the array. This is a reasonable assumption
untreated sand. These results are also consistent with the initial small- because the distance from the surface to the first accelerometer was
strain G obtained through the resonant column tests for loose Nevada only about 0.8 m. Based on rigid body motion, the acceleration at the
sand treated with CS (Spencer et al. 2008). The values of VS from surface is considered equal to the acceleration measured at Node 1
CTC02 indicated an increase of about 14% in G between CS-4 and from Fig. 14. The equations used for calculation of shear stress are
CS-5. Although strength gains in soils are reported to increase with presented in Kamai and Boulanger (2010). The shear stresses were
increasing CS concentration, such large differences between the two normalized by the initial effective overburden pressure at the be-
treatment concentrations are not expected to occur. The discrepancy is ginning of each shake to establish the CSR. The CSR was determined
likely a result of the fact that CS-4 represents treatment with a weak as the ratio of shear stress at the midpoints of each element by the
grout characteristic of a highly deformable, nonflowing gel. corresponding initial effective vertical stress at a similar depth.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1341

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Calculation of Shear Strains applied to the accelerometer records before the integration. Differ-
entiation of the displacements utilized a weighted residual method
The shear strains were obtained from the vertical accelerometer
following Kamai and Boulanger (2010) to determine the shear
array. The acceleration records were double integrated in time to
strains at the midpoints between the sensor locations.
obtain transient displacements and then differentiated in space to
Figs. 15–17 show the stress-strain response for (1) small base
obtain shear strains. A sixth-order Butterworth band-pass filter was
shaking amplitude; (2) medium base shaking amplitude when rU 51:0
occurs in the untreated sand; and (3) large base shaking amplitude. Fig.
15 shows the stress-strain response for Shakes 2–5 at the midpoint of
the liquefiable layer (D 5 3:1 m). However, CS-4 exhibited erratic
stress-strain behavior and was omitted from the analysis. The erratic
behavior is likely attributed to minor rotations of the sensors along with
the poor quality of the signals in one or more sensors, which become
extremely sensitive to the analysis (Kamai and Boulanger 2010). For
a CSR of approximately 0.1, the shear strain (g) is about 0.025% for
each soil type and does not indicate development of additional shear
strain for additional loading cycles. The values of g for Shake 2 were
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

also consistent for liquefiable soils with g  0:01% (Drnevich and


Richart 1970; Dobry et al. 1982). The response for Shake 3 showed
a rapid development of g in the untreated soil of up to 1%. Compared
with the untreated soil, gmax was reduced to 0.3 and 0.6% for CS-9 and
CS-5, respectively. Similar reductions in g occurred for Shakes 4 and 5
when the model was subjected to stronger base shaking, and the
reductions appear to correlate well with the CS concentrations (Fig. 15).
In the untreated soil, liquefaction was triggered (rU 5 1:0) during
Shake 3. Based on the pore pressure response from Fig. 12, rU 5 1:0
occurred about 3 s into the shaking event, corresponding to the sixth
shaking cycle. Fig. 16 shows the stress-strain profiles at specific
cycles in the shaking event to capture the behavior during (1) the
beginning of shaking (Cycle 2); (2) the middle of shaking (Cycle 6);
and (3) the end of shaking (Cycle 19). In Cycle 2 (prior to rU 5 1:0)
there was development of g  0:5% for the untreated soil while the
Fig. 14. Schematic of the 1D shear-beam model for interpreting the treated soils during Cycle 2 developed g  0:2%.
vertical array data (after Kamai and Boulanger 2010) As rU approached 1.0 in the untreated soil, g developed rapidly up
to 0.5% during Cycle 6 for a CSR of 0.3 (Fig. 16). At this time, the

Fig. 15. CSR versus shear strain for Shakes 3–5 at the midpoint of the liquefiable layer (D 5 3:1 m) for untreated, 9% CS, and 5% CS

1342 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


untreated soil experienced zero effective stress and reached the movement of the sensors in the loose sand layer as it liquefies and
failure envelope, during which the sand underwent dilative behavior experiences large deformations. On the other hand, CS-9 from Fig. 16
and was accompanied by transient stiffening as well as temporary shows that the response during Cycle 19 is nearly identical to that from
drops in rU . Comparatively, the CS sands showed reductions in Cycle 6. The characteristic response for CS-9 shows that for a CSR of
g with increasing CS concentrations. For example, at the midpoint of about 0.2, a maximum value of g is reached (0.3%) early in the
the liquefiable soil layer, g from Cycles 2 to 6 increased by 0.1 and shaking sequence at which point there is no further increase with
0.3% for CS-9 and CS-5, respectively. increasing loading cycles. On the other hand, for a CSR of 0.25, CS-5
Toward the end of the shaking event (Cycle 19), the differences in experienced continuously incremental increases in shear strain with
behavior between the grouted and ungrouted soils were more pro- further loading cycles. For example, at the midpoint of the layer, CS-5
nounced. Fig. 16 shows significant loss in stiffness in the untreated soil showed an increase in g from 0.2% (Cycle 2) up to 0.65% (Cycle 19).
indicated by dramatic differences in the stress-strain path between Overall, the maximum CSRs and shear strains that developed were
Cycles 2 and 19. The unusual stress-strain path during Cycle 19 was both reduced for the CS sands and the reductions correlated well with
primarily attributed to further soil softening by additional loading the treatment levels. For example, the averages of the maximum CSRs
cycles applied to the liquefied soils, thus precluding transmission of determined along the profile were 0.3, 0.25, and 0.2 and the average
motions along the profile. Another contributing factor to the unusual values of gmax that developed along the profile were determined to be
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

stress-strain path observed from Cycles 6 and 19 may be a result of the 0.7, 0.4, and 0.3% for the untreated sand, CS-5, and CS-9, respectively.

Fig. 16. CSR versus shear strain for Shake 3 (PBA 5 0.1–0:15g)

Fig. 17. CSR versus shear strain for Shake 5 (PBA 5 0.56–0:69g)

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1343

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


The stress-strain response for Shake 5 represents an applied PBA Acknowledgments
of 0.56 and 0:69g for CTC01 and CTC02, respectively, and is shown
in Fig. 17, which also shows that g  2:7% along the profile of the Model tests were performed at the Center for Geotechnical Modeling,
untreated soil during Cycle 2 of the shake sequence. Soon after Cycle University of California at Davis. The writers would like to acknowl-
2, there was a limited response in the untreated soil during Cycle 6 as edge the assistance of Dan Wilson, Chad Justice, Peter Rojas, Ray
a result of further soil softening that impeded motions from prop- Gerhard, Nick Sinikas, and Lars Pederson. This research was per-
agating up through the profile. formed as part of a NEES Grand Challenge project, and benefitted
The behavior of the grouted soils indicated quick development from discussions and collaborations with Glenn Rix, Ellen Rathje,
of g in the early stages of cyclic loading with smaller g developed Rachelle Howell, and Tony Marinucci. This material is based on
for increasing CS concentrations. During Cycle 2, both concen- work supported by the National Science Foundation (NSF) under
trations showed a CSR 5 0.8 corresponding to g  1.2 and 1.8% Grant No. CMS-0530478. Any opinions, findings, and conclusions
along the profile for CS-9 and CS-5, respectively. CS-9 exhibited expressed in this material are those of the writers and do not neces-
no further change in CSR or g for up to 20 additional loading sarily reflect the views of NSF. Recent upgrades have been funded
cycles. CS-5 underwent cyclic mobility with only minor increases by NSF Award No. CMS 0086566 through the George E. Brown,
in g. Jr., Network for Earthquake Engineering Simulation (NEES).
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

In general, the stress-strain loops determined from the model


tests demonstrate the cyclic mobility behavior of CS sands is very References
similar to that associated with dense-of-critical sands in undrained
cyclic loading. Additionally, the hysteretic response determined
Brandenberg, S. J., Kutter, B. L., and Wilson, D. W. (2008). “Fast stacking
for increasing applied horizontal base accelerations in subsequent and phase corrections of shear wave signals in a noisy environment.” J.
shakes confirms the ability of CS sands to provide adequate cyclic Geotech. Geoenviron. Eng., 134(8), 1154–1165.
shear resistance despite the occurrence of the increasing shear Chen, Y. R. (1995). “Behavior of fine sand in triaxial, torsional and rotational
strains that were observed. In this way, the gelled sand matrix shear tests.” Ph.D. thesis, Univ. of California–Davis, Davis, CA.
continued to provide significant shear resistance, which also Cruz, R. D. (1995). “Rate dependent shear and consolidation of remolded San
translates well with the expanded dilation region observed for CS Francisco Bay mud.” Ph.D. thesis, Univ. of California–Davis, Davis, CA.
sands during phase transformation under monotonic loading Dobry, R., Ladd, R. S., Yokel, F. Y., Chung, R. M., and Powell, D. (1982).
(Kodaka et al. 2005). Prediction of pore water pressure buildup and liquefaction of sands
during earthquakes by the cyclic strain method, Building Science Series
183, National Bureau of Standards, Gaithersburg, MD.
Doygun, O. (2009). “Monotonic and cyclic undrained loading behavior of
intermediate soils.” M.S. thesis, Technische Univ. Darmstadt, Darm-
Conclusions stadt, Germany.
Drnevich, V. P., and Richart, F. E., Jr. (1970). “Dynamic prestraining of dry
The centrifuge tests provided comparisons of the dynamic re- sands.” J. Soil Mech. and Found. Div., 96(2), 453–469.
sponse for untreated, loose sands, and loose sands treated with 4, Gallagher, P. M. (2000). “Passive site remediation for mitigation of liq-
5, and 9% CS (by weight) subjected to a sequence of dynamic uefaction risk.” Ph.D. dissertation, Virginia Polytechnic Institute and
shaking events. The dynamic responses were characterized by State Univ., Blacksburg, VA.
surface displacements, accelerations, and the hysteretic behavior Gallagher, P. M., Conlee, C. T., and Rollins, K. M. (2007a). “Full-scale field
characterized through integration and inverse analysis methods to testing of colloidal silica grouting for mitigation of liquefaction risk.” J.
Geotech. Geoenviron. Eng., 133(2), 186–196.
obtain stress-strain responses. After comparing the response be-
Gallagher, P. M., and Lin, Y. (2005). “Column testing to determine colloidal
tween a liquefiable sand layer saturated with various concen- silica transport mechanisms.” Innovations in grouting and soil im-
trations of CS and a liquefiable sand layer saturated with only provement, Geotechnical Special Publication No. 136, E. M. Rathje, ed.,
water, the key findings from the test show the following: ASCE, Reston, VA.
• Ground deformations were progressively reduced with increas- Gallagher, P. M., and Mitchell, J. K. (2002). “Influence of colloidal silica
ing levels of CS treatment. grout on liquefaction potential and cyclic undrained behavior of loose
• CS treatment can dramatically reduce lateral spreading in slopes sand.” Soil Dyn. Earthquake Eng., 22(9-12), 1017–1026.
resting on liquefiable soil layers. Gallagher, P. M., Pamuk, A., and Abdoun, T. (2007b). “Stabilization of
• Sands treated with CS yield higher cone tip resistance and shear liquefiable soils using colloidal silica grout.” J. Mater. Civ. Eng., 19(1),
wave velocity. 33–40.

Iler, R. K. (1979). The chemistry of silica: Solubility, polymerization, colloid
The stress-strain responses of the treated sands were consistent
and surface properties, and biochemistry, Wiley, New York.
with the laboratory element test data in that they exhibited a cyclic Kamai, R., and Boulanger, R. W. (2010). “Characterizing localization
mobility behavior consistent with the undrained response of processes during liquefaction using inverse analyses of instrumentation
dense-of-critical sands. arrays.” Meso-scale shear physics in earthquake and landslide me-
• Increasing treatment levels produced greater cyclic resistance chanics, Y. H. Hatzor, J. Sulem, and I. Vardoulakis, eds., CRC, Leiden,
ratios and lower cyclic shear strains. These results are consistent Netherlands, 219–238.
with reductions of ground deformations determined for increas- Kammerer, A., Wu, J., Riemer, M., Pestana, J., and Seed, R. (2004). “Shear
ing concentrations of CS treatment. strain development in liquefiable soil under bidirectional loading con-
CS grouting offers the potential advantage of enabling passive dition.” Proc., 13th World Conf. on Earthquake Engineering (WCEE),
stabilization of soils that underlay existing structures that may be Paper No. 2081, Vancouver, Canada.
Kasantikul, P. (2009). “Resonant column and torsional shear testing to
otherwise difficult to treat by other methods. The results of these
evaluate soil properties deposited using dry pluviation and hydraulic
centrifuge tests support the findings from the laboratory element fill.” M.S. thesis, Rensselaer Polytechnic Institute, Troy, NY.
studies in demonstrating the treated soil’s improved resistance to Kodaka, T., Oka, F., Ohno, Y., Takyu, T., and Yamasaki, N. (2005).
seismic deformations. The archived experimental data also provide “Modeling of cyclic deformation and strength characteristics of silica
a basis for future evaluation of numerical models that may be used treated sand.” Geomechanics: Testing, modeling, and simulation, Geo-
in the design or analysis of CS-treated sites. technical Special Publication No. 143, ASCE, Reston, VA, 205–216.

1344 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.


Moridis, G. J., Pruess, K., Persoff, P., and Apps, J. A. (1995). “Performance Salgado, R., Boulanger, R. W., and Mitchell, J. K. (1997). “Lateral stress
and properties of colloidal silica and polysiloxane grouts.” Proc., Int. effects on CPT liquefaction resistance correlations.” J. Geotech. Geo-
Containment Technology Workshop, Baltimore. environ. Eng., 123(8), 726–735.
Pamuk, A., Gallagher, P. M., and Zimmie, T. (2007). “Remediation of pile Spencer, L. M., Rix, G. J., and Gallagher, P. M. (2008). “Colloidal silica gel
foundations against lateral spreading by passive site stabilization and sand mixture dynamic properties.” Proc., Conf. of Geotechnical
technique.” Soil Dyn. Earthquake Eng., 27(9), 864–874. Earthquake Engineering and Soil Dynamics IV, Geotechnical Special
Persoff, P., Apps, J., Moridis, G., and Whang, J. M. (1999). “Effect of Publication No. 181, ASCE, Reston, VA.
dilution and contaminants on strength and hydraulic conductivity of sand Whang, J. M. (1995). “Chemical-based barrier materials.” Assessment of
grouted with colloidal silica gel.” J. Geotech. Geoenviron. Eng., 125(6), barrier containment technologies for environmental remediation appli-
461–469. cations, R. R. Rumer and J. K. Mitchell, eds., National Technical In-
Polito, C. P. (1999). “The effects of non-plastic and plastic fines on the formation Service, Springfield, VA, 211–247.
liquefaction of sandy soils.” Ph.D. dissertation, Dept. of Civil Engi- Zeghal, M., Elgamal, W. A., Tang, H. T., and Stepp, J. C. (1995). “Evaluation
neering, Virginia Polytechnic Institute and State Univ., Blacksburg, VA. of soil nonlinear properties.” J. Geotech. Engrg., 121(4), 363–378.
Downloaded from ascelibrary.org by Xavier Vera on 11/10/12. Copyright ASCE. For personal use only; all rights reserved.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2012 / 1345

J. Geotech. Geoenviron. Eng. 2012.138:1334-1345.

Вам также может понравиться