You are on page 1of 8

Technical Note

Small-Strain Stiffness, Shear-Wave Velocity,


and Soil Compressibility
Minsu Cha, Ph.D., A.M.ASCE1; J. Carlos Santamarina, Ph.D., M.ASCE2; Hak-Sung Kim3;
and Gye-Chun Cho, Ph.D.4

Abstract: The small-strain shear modulus depends on stress in uncemented soils. In effect, the shear-wave velocity, which is often used to
Downloaded from ascelibrary.org by Colorado School of Mines on 09/29/14. Copyright ASCE. For personal use only; all rights reserved.

calculate shear stiffness, follows a power equation with the mean effective stress in the polarization plane Vs 5 aðs9m=1 kPaÞb , where the a
factor is the velocity at 1 kPa, and the b exponent captures the velocity sensitivity to the state of stress. The small-strain shear stiffness, or
velocity, is a constant-fabric measurement at a given state of stress. However, parameters a and b are determined by fitting the power equation to
velocity measurements conducted at different effective stress levels, so changes in both contact stiffness and soil fabric are inherently involved.
Therefore, the a and b parameters should be linked to soil compressibility CC . Compiled experimental results show that the a factor decreases
and the b exponent increases as soil compressibility CC increases, and there is a robust inverse relationship between a and b for all sediments:
b  0:73 2 0:27 log½a=ðm=sÞ. Velocity data for a jointed rock mass show similar trends, including a power-type stress-dependent velocity and
inverse correlation between a and b; however, the a-b trend for jointed rocks plots above the trend for soils. DOI: 10.1061/(ASCE)GT.1943-
5606.0001157. © 2014 American Society of Civil Engineers.
Author keywords: Shear-wave velocity; Velocity-stress power relations; Contact effects; Compression index; Granular fabric; Coordination
number.

Introduction a factor (m=s) 5 shear-wave velocity at a mean effective stress


s9m 5 1 kPa; and the b exponent captures the sensitivity of the
Soils and fractured rocks are granular materials. The small-strain skeletal shear stiffness to the applied stress. This equation applies
shear stiffness G reflects contact-level deformation and exhibits when capillary forces are significantly smaller than stress-induced
a Hertzian-type power relation with effective stress. Shear-wave skeletal forces.
propagation is the most versatile and portable method to assess the Although wave propagation is a small-strain constant-fabric stiff-
small-strain shear stiffness in the laboratory (e.g., bender element ness measurement, velocity-stress relations such as Eq. (1) are fitted to
and resonant column) and in the field (e.g., cross hole, seismic cone, wave-velocity data gathered at different effective stress levels that
and surface waves). The shear-wave velocity VS (m=s) is determined inherently involve fabric changes. Therefore, the velocity-stress power
by the mean effective stress in the polarization plane s9m (kPa) relationship captures both contact behavior and fabric changes. [Note
(Hardin and Drnevich 1972; Knox et al. 1982; Petrakis and Dobry that the correction factor for void ratio e frequently used in other forms
1987; Santamarina and Cascante 1996) of Eq. (1) is implicitly included in the a factor (Hardin and Richart
rffiffiffiffi   1963; Hardin and Drnevich 1972; Lo Presti et al. 1995; Shibuya et al.
G s9’ þ s9== b 1997).]
VS ¼ ¼a (1)
r 2 kPa Soil compressibility CC measured in oedometer cells is primarily
the result of interparticle relative displacement and fabric changes.
where s9’ and s9== 5 effective stresses in the direction of particle Thus, a causal link between small-strain a and b parameters and the
motion and the direction of wave propagation, respectively; the compressibility CC of soils is anticipated. The authors compiled
a database to explore relations between these parameters; the
database includes a wide range of soils, from fine to coarse, both
1
Postdoctoral Fellow, School of Civil and Environmental Engineering, natural and freshly remolded specimens, and both normally and
Georgia Institute of Technology, Atlanta, GA 30332-0355 (corresponding overconsolidated sediments.
author). E-mail: mcha678@gmail.com
2
Professor, School of Civil and Environmental Engineering, Georgia
Institute of Technology, Atlanta, GA 30332-0355. E-mail: jcs@gatech.edu Compressibility and Shear-Wave Velocity
3
Doctoral Candidate, Dept. of Civil and Environmental Engineering,
Korea Advanced Institute of Science and Technology (KAIST), Yuseong-gu, The sediment compressibility for a normally consolidated soil can
Daejeon 305-701, Republic of Korea. E-mail: shield5200@kaist.ac.kr
4
be characterized using the compression index CC obtained by fitting
Professor, Dept. of Civil and Environmental Engineering, Korea the standard void ratio e versus vertical effective stress s9z relation to
Advanced Institute of Science and Technology (KAIST), Yuseong-gu,
laboratory data
Daejeon 305-701, Republic of Korea. E-mail: gyechun@kaist.edu
Note. This manuscript was submitted on July 13, 2013; approved on  
s9z
June 2, 2014; published online on June 27, 2014. Discussion period open e ¼ e0 2 CC log (2)
until November 27, 2014; separate discussions must be submitted for s9z0
individual papers. This technical note is part of the Journal of Geo-
technical and Geoenvironmental Engineering, © ASCE, ISSN 1090- Fig. 1 shows the correlation between the a factor and b exponent
0241/06014011(4)/$25.00. with compressibility. Data are collected from previous studies

© ASCE 06014011-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2014.140.


Downloaded from ascelibrary.org by Colorado School of Mines on 09/29/14. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Inverse relationship between the b exponent and the a factor;


data for jointed rocks are shown for comparison (equation defines the
central trend for the soil data shown as a continuous line; dotted lines
show 61 SD from the central trend)

a ¼ 13:5ðm=sÞ × CC20:63 (3)

b ¼ 0:17 log CC þ 0:43 (4)

Statistical parameters for fitted trends are shown in the figures and
include the coefficient of determination R, the standard error (SE),
and the number n of data points used in the regression.
The compression index CC appears in both expressions; this
suggests a direct relationship between the a factor (m=s) and the b
exponent; combining Eqs. (3) and (4) produces
 
a
b ¼ 0:73 2 0:27 log , 1 m=s # a # ∼ 500 m=s (5)
m=s

The database plotted in Fig. 1 and additional data gathered from


published studies on wave velocity by other researchers are plotted
in Fig. 2. (Note that soil types and all references can be found in
Fig. 1. Relationships between small-strain parameters: (a) a factor Fig. S1 in the Supplemental Data.) The data trend is well pre-
(i.e., shear-wave velocity at 1 kPa) and (b) b exponent (i.e., sensitivity of dicted by Eq. (5) (superimposed on Fig. 2), which supersedes an
the shear-wave velocity to changes in mean effective stress on the earlier relationship that was based on a limited data set [ b 5 0:36
polarization plane) and the sediment compression index CC (equations 2 a=ð700 m=sÞ in Santamarina et al. (2001)]. Underlying local
define the central trend indicated by the continuous lines; dotted lines trends within the full data set presented in Fig. S2 of the Supple-
define 61 SD from the central trend) mental Data are consistent with differences in compressibility CC :
denser sandy soils, rounder particles, lower plasticity, and higher
cementation are associated with higher a factors and smaller b
exponents than their counterparts.
published by the authors; in most cases, the data were obtained
in oedometer cells instrumented with bender elements. All values Discussion: Physical Interpretation of Observed
correspond to stress levels below the yield stress when grain Trends—Implications
crushing governs sediment compaction. The range of mean ef-
fective pressures was from 10 up to 1,200 kPa. In the case of
lightly cemented soils, plotted compressibility and velocity Small- and Large-Strain Stiffness
parameters were determined before the onset of cementation The tangent-constrained modulus M can be computed from the
breakage. derivative of Eq. (2), i.e., the local tangent to the e-s9z compression
Soils range from lightly cemented dense sands (low CC , high a, curve for a normally consolidated soil
and low b values) to soft, high-specific-surface clays (high CC , low
a, and high b values). It can be observed that highly compressible ds9z 1 þ e0
M¼ ¼ 2:3 s9z (6)
fine-grained sediments have low shear-wave velocity at low con- dɛ z CC
finement (a factor) and high stiffness sensitivity to changes in mean
effective stress on the polarization plane (b exponent). Trends are This is the tangent to a large-strain compressibility trend that
properly fitted with the following expressions: involves mostly plastic strains; indeed, the linear dependency with

© ASCE 06014011-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2014.140.


vertical effective stress sz9 underscores the underlying Mohr- Etan 1 r
¼  c (8)
Coulomb frictional resistance to deformation. In contrast, the small- Gg 1 2 ng R
strain shear stiffness is a constant-fabric measurement that reflects
the stiffness of contacts and follows a power relation with effective where R 5 radius of spherical grains. This form of the equation
stress [Eq. (1)]. highlights that the sediment skeletal stiffness increases whenever the
contact area increases, either from the applied stress (in hertz) or from
any diagenetic process such as creep, pressure-solution/precipitation,
a Factor
cementation, and salt precipitation (analyses in Dvorkin et al. 1991;
The small-strain stiffness Etan of regular arrangements of monosize Dvorkin and Yin 1995; Fernandez and Santamarina 2001). Conse-
spherical particles can be computed assuming Hertzian contact quently, natural soils tend to be stiffer and less compressible than their
behavior; for example, for a simple cubic packing (Richart et al. remolded counterparts (Dudas 1981; Burland 1990; Abduljauwad and
1970), see other cases in Santamarina et al. (2001) Al-Amoudi 1995; Houston et al. 2001; Herrera et al. 2007). Fully
lithified sediments such as shales and sandstones can reach shear-
2 vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3
Downloaded from ascelibrary.org by Colorado School of Mines on 09/29/14. Copyright ASCE. For personal use only; all rights reserved.

u wave velocities as high as 2,500e3,000 m=s. Microfissures can cause


u G2g
Etan ¼ 4 t   5s
3 3 1=3 stress-dependent stiffness even in lithified sediments, and the expo-
(7) nent may be greater (b $ 0).
2 1 2 ng 2

Preloading
where Gg and ng 5 shear modulus and Poisson’s ratio, respectively,
of the mineral that makes the grains. The factor in brackets in Eq. (7) Similar to normally consolidated soils, the small-strain shear stiffness
is equivalent to the a factor. The analysis of this and similar of preloaded sediments is determined by the current state of stress and
expressions for other packings confirms the link between the a factor the void ratio (i.e., interparticle coordination). Changes in the void
and grain packing or fabric, as observed in the data (Duffy and ratio during loading are only partially recovered during unloading,
Mindlin 1957; Deresiewicz 1974; Petrakis and Dobry 1987; Chang and residual stresses may remain when zero lateral strain conditions
et al. 1991; Santamarina et al. 2001). prevail; hence, the void ratio and the horizontal effective stress under
k0 conditions depend on the maximum past pressure s9. p Accordingly,
several possible predictive equations can be suggested (e.g., Hardin
b Exponent and Black 1969; Houlsby and Wroth 1991; Viggiani and Atkinson;
1995; Shibuya et al. 1997; Vardanega and Bolton 2013).
The b exponent reflects both the nature of interparticle contacts and
fabric changes during loading. Contact deformation at constant
fabric can justify the following b values (Cascante and Santamarina Jointed Rocks
1996): The shear and longitudinal stiffnesses of jointed rock masses depend
• b 5 0 for an ideal solid, and b  0 for cemented granular media on stress resulting from contacts between asperities or between grains
(below the transition stress); that form the gouge material. The global a-b trend for jointed rock
• b 5 1=6 for Hertzian contacts (elastic spherical particles); masses resembles that for soils, but it is shifted above the soil trend
• b 5 1=4 for cone-to-plane contacts (rough or angular particles); [open rhombuses in Fig. 2; data from Fratta and Santamarina (2002)
• b 5 1=4 for particles that experience contact yield; and and Cha et al. (2009)]. Trends within this data set reveal that a
• b 5 3=4 for Coulomb’s electrical force between charges, and, in increases and b decreases when (1) the stiffness of the intact blocks and
general, b is variable with interparticle distance for contacts the joint separation increase, (2) joint surfaces are rougher, (3) gouge
dominated by double-layer effects. thickness decreases, and (4) the gouge sediment is less plastic [detailed
In addition to contact deformation, data in Fig. 1 show that more analyses in Fratta and Santamarina (2002) and Cha et al. (2009)].
compressible soils exhibit a higher b exponent even for the same
contact characteristics because of fabric changes and increased Engineering Applications
coordination number during loading.
The stiffness of the granular skeleton controls the analysis of defor-
mations in all geosystems, from foundations, tunnels, and pavements
Inverse Relationship between a and b to reservoir compaction during oil production. Stress-dependent
The trend in Fig. 2 is observed from very soft, high-specific-surface stiffness adds complexity to such analyses. Causally linked trends
soils b → 0:7 [e.g., sedimentation from slurry controlled by elec- reported in this paper can help to discern soil type from velocity-
trical Coulombian interactions where shear-wave velocities can be versus-depth profiles (e.g., low a and high b imply high-plasticity
as low as Vs 5 0:5 m=s (data in Klein and Santamarina 2005)] to clays, high a and low b suggest rounded sands, and b  0 indicates
stiff, lightly cemented sediments where skeletal stiffness exhibits cemented media), assess the extent of diagenetic cementation (b → 0),
vanishing sensitivity to changes in effective stress b → 0 [e.g., and preselect and corroborate parameters for analysis. Results show
lightly cemented sands (Fernandez and Santamarina 2001; Yun and that shear-wave velocity can be used to estimate shear stiffness and
Santamarina 2005)]. For highly cemented sediments, the exponent compressibility; this is particularly important in aged soils and coarse-
remains b  0 in the absence of microfissures or debonding, and the grained deposits, where sampling effects and testing difficulties
a factor increases with the extent of cementation. preclude the application of standard testing methods.

Conclusions
Cementation-Diagenesis-Lithification
The macroscale stiffness estimated for a granular medium where The small-strain stiffness of the granular skeleton is a measure of
grains interact through Hertzian contacts [Eq. (7)] can be written in state; it is determined at constant fabric and reflects contact-level
terms of the radius of the contact area between grains rc as deformation.

© ASCE 06014011-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2014.140.


Shear-wave propagation is a versatile and robust method to de- Dvorkin, J., and Yin, H. (1995). “Contact laws for cemented grains:
termine the small-strain shear stiffness of sediments in the laboratory Implications for grain and cement failure.” Int. J. Solids Struct.,
and the field. The shear-wave velocity follows a power relation with 32(17–18), 2497–2510.
the mean effective stress on the polarization plane Vs 5 aðs9mÞb . The Fernandez, A. L., and Santamarina, J. C. (2001). “Effect of cementation on
the small-strain parameters of sands.” Can. Geotech. J., 38(1), 191–199.
a factor (Vs at s9m 5 1 kPa) and the b exponent reflect contact
Fratta, D., and Santamarina, J. C. (2002). “Shear wave propagation in jointed
behavior and changes in fabric associated with effective stress rock: State of stress.” Géotechnique, 52(7), 495–505.
changes. Hardin, B. O., and Black, W. L. (1969). “Closure of ‘Vibration modulus of
Less-compressible soils exhibit higher a factors and lower b normally consolidated clay’.” J. Soil Mech. and Found. Div., 95(6),
exponents, i.e., clays of lower plasticity, denser sands made of 1531–1537.
rounder grains, and soils that have experienced diagenetic cemen- Hardin, B. O., and Drnevich, V. P. (1972). “Shear modulus and damping in
tation. There is a robust inverse relationship between a and b for all soils: Measurement and parameter effects.” J. Soil Mech. and Found.
sediments: b  0:73 2 0:27 log½a=ðm=sÞ. Data for jointed-rock Div., 98(6), 603–624.
masses agree with the inverse a-b trend, but the data deviate from Hardin, B. O., and Richart, F. E., Jr. (1963). “Elastic wave velocities in
granular soils.” J. Soil Mech. and Found. Div., 89(1), 33–65.
Downloaded from ascelibrary.org by Colorado School of Mines on 09/29/14. Copyright ASCE. For personal use only; all rights reserved.

the trend obtained for sediments.


Herrera, M. C., Lizcano, A., and Santamarina, J. C. (2007). “Colombian
volcanic ash soils.” Characterisation and engineering properties of
natural soils, T. S. Tan, K. K. Phoon, D. W. Hight, and S. Leroueil, eds.,
Acknowledgments Vol. 3, Taylor & Francis, Singapore, 2385–2409.
Houlsby, G. T., and Wroth, C. P. (1991). “The variation of shear modulus
Support for this research was provided by the Department of Energy of a clay with pressure and overconsolidation ratio.” Soils Found.,
Savannah River Operations Office (Aiken, South Carolina) and the 31(3), 138–143.
Goizueta Foundation (Atlanta, Georgia). The authors are grateful to Houston, S. L., Houston, W. N., Zapata, C. E., and Lawrence, C. (2001).
the anonymous reviewers for their valuable comments and insight. “Geotechnical engineering practice for collapsible soils.” Geotech. Geol.
Eng., 19(3–4), 333–355.
Klein, K., and Santamarina, J. C. (2005). “Soft sediments: Wave-based
characterization.” Int. J. Geomech., 10.1061/(ASCE)1532-3641(2005)
Supplemental Data 5:2(147), 147–157.
Knox, D. P., Stokoe, K. H., and Kopperman, S. E. (1982). “Effect of state
Figs. S1 and S2 are available online in the ASCE Library (www of stress on velocity of low-amplitude shear waves propagating along
.ascelibrary.org). principal stress directions in dry sand.” Technical Rep. GR 82-23, Geo-
technical Engineering Center, Univ. of Texas–Austin, Austin, TX.
Lo Presti, D. C. F., Pallara, O., Costanzo, D., and Impavido, M. (1995). “Small
References strain measurements during triaxial tests: Many problems, some sol-
utions.” Pre-failure deformation of geomaterials, S. Shibuya, T. Mitachi,
Abduljauwad, S. N., and Al-Amoudi, O. S. B. (1995). “Geotechnical be- and S. Miura, eds., Balkema, Rotterdam, Netherlands, 1067–1088.
haviour of saline sabkha soils.” Géotechnique, 45(3), 425–445. Petrakis, E., and Dobry, R. (1987). “Micromechanical modeling of granular
Burland, J. B. (1990). “On the compressibility and shear strength of natural soil at small strain by arrays of elastic spheres.” Rep. No. CE-87-02, Dept.
clays.” Géotechnique, 40(3), 329–378. of Civil Engineering, Rensselaer Polytechnic Institute, Troy, NY.
Cascante, G., and Santamarina, J. C. (1996). “Interparticle contact behavior Richart, F. E., Hall, J. R., and Woods, R. D. (1970). Vibrations of soils and
and wave propagation.” J. Geotech. Engrg., 10.1061/(ASCE)0733-9410 foundations, Prentice-Hall, Englewood Cliffs, NJ.
(1996)122:10(831), 831–839. Santamarina, J. C., and Cascante, G. (1996). “Stress anisotropy and wave
Cha, M., Cho, G.-C., and Santamarina, J. C. (2009). “Long-wavelength propagation: A micromechanical view.” Can. Geotech. J., 33(5), 770–782.
P-wave and S-wave propagation in jointed rock masses.” Geophysics, Santamarina, J. C., Klein, K. A., and Fam, M. A. (2001). Soils and waves:
74(5), E205–E214. Particulate materials behavior, characterization and process moni-
Chang, C. S., Misra, A., and Sundaram, S. S. (1991). “Properties of granular toring, Wiley, Chichester, U.K.
packings under low amplitude cyclic loading.” Soil. Dyn. Earthquake Shibuya, S., Hwang, S. C., and Mitachi, T. (1997). “Elastic shear modulus of
Eng., 10(4), 201–211. soft clays from shear wave velocity measurement.” Géotechnique, 47(3),
Deresiewicz, H. (1974). “Bodies in contact with applications to granular 593–601.
media.” R.D. Mindlin and applied mechanics, B. Herrman, ed., Per- Vardanega, P. J., and Bolton, M. D. (2013). “Stiffness of clays and silts:
gamon Press, New York, 105–147. Normalizing shear modulus and shear strain.” J. Geotech. Geoenviron.
Dudas, M. J. (1981). “Long-term leachability of selected elements from fly Eng., 10.1061/(ASCE)GT.1943-5606.0000887, 1575–1589.
ash.” Environ. Sci. Technol., 15(7), 840–843. Viggiani, G., and Atkinson, J. H. (1995). “Stiffness of fine-grained soil at
Duffy, J., and Mindlin, R. D. (1957). “Stress-strain relations and vibrations very small strains.” Géotechnique, 45(2), 249–265.
of a granular medium.” J. Appl. Mech., 24(11), 585–593. Yun, T. S., and Santamarina, J. C. (2005). “Decementation, softening,
Dvorkin, J., Mavko, G., and Nur, A. (1991). “The effect of cementation and collapse: Changes in small-strain shear stiffness in k0 loading.”
on the elastic properties of granular material.” Mech. Mater., 12(3–4), J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2005)131:
207–217. 3(350), 350–358.

© ASCE 06014011-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng. 2014.140.


SUPPLEMENTAL DATA

ASCE Journal of Geotechnical and Geoenvironmental Engineering

Small-Strain Stiffness, Shear Wave Velocity, and Soil Compressibility


Minsu Cha, J. Carlos Santamarina, Hak-Sung Kim, and Gye-Chun Cho

DOI: 10.1061/(ASCE)GT.1943-5606.0001157

Supplemental Data

Figure S1. Inverse relationship between the β-exponent and the α-factor. List of soil types and references.

0.6

0.5

0.4
β exponent

0.3

0.2

0.1

0
5 50 500
α factor [m/s]
Lee et al. 2007: Mica + sand
Shin and Santamarina 2013: Round and blasting sand
Kim and Santamarina 2008: Sand + rubber
Cha and Santamarina (unpublished): Clayey and sandy soils
Cha and Cho 2007: Sand - varying densities
Santamarina et al. 2001: Soils and other particulate materials
Cascante and Santamarina 1996: Steel ball, lead shot, silica flour-kaolinite pellet, and sand
Fernandez and Santamarina 2001: Sand - cemented
Yun and Santamarina 2005: Sand - cemented
Cha et al. 2009: Jointed rocks
Kwon et al. 2011: Clayey soils
Patel et al. 2009: mainly sand and glass bead
Lee et al. 2010: Sand-rubber mixtures
Stokoe et al. 1991: Sand
Belloti et al. 1996: Sand
Yamashita et al. 2003: Sand
Jamiolkowski et al. 1995: Clay
Ng et al. 2004: Decomposed tuff
Santagata et al. 2005: Clay
Kim et al. 2013: Mostly clayey soils
Oh 2008: Mostly clayey soils

www.ascelibrary.org © ASCE 2014 / S1


SUPPLEMENTAL DATA

Figure S2. Local trends within the full dataset. Denser sandy soils, rounder particles, lower plasticity and
higher cementation are associated to higher α-factors and smaller β-exponents than their counterparts. (a)
Density. (b) Plasticity. (c) Particle angularity. (d) Mica content and cementation.

0.5 0.6

0.4 0.5

Increasing

β exponent
0.4
β exponent

Increasing plasticity
0.3
density
0.3
0.2
0.2
0.1
0.1
(a) (b)
0 0
5 50 500 5 50 500
α factor [m/s] α factor [m/s]

0.3 0.4
: Increasing mica
contents
0.3
β exponent

β exponent

Increasing particle
angularity
0.2

0.1
: Increasing
cementation
(c) (d)
0.2 0
50 500 5 50 500
α factor [m/s] α factor [m/s]

S2 / © 2014 ASCE www.ascelibrary.org


SUPPLEMENTAL DATA

References

Bellotti, R., Jamiolkowski, M., LoPresti, D. C. F., and Oneill, D. A. (1996). "Anisotropy of small strain
stiffness in Ticino sand." Geotechnique, 46(1), 115-131.

Cascante, G., and Santamarina, J. C. (1996). "Interparticle contact behavior and wave propagation." J.
Geotech. Eng.-ASCE, 122(10), 831-839.

Cha, M., and Santamarina, J. C. (unpublished). Laboratory data of Savannah River Site specimens –
clayey and sandy soils.

Cha, M. S., and Cho, G. C. (2007). "Shear strength estimation of sandy soils using shear wave velocity."
Geotech. Test. J., 30(6), 484-495.

Cha, M., Cho, G. C., and Santamarina, J. C. (2009). "Long-wavelength P-wave and S-wave propagation
in jointed rock masses." Geophysics, 74(5), E205-E214.

Fernandez, A. L., and Santamarina, J. C. (2001). "Effect of cementation on the small-strain parameters of
sands." Canadian Geotechnical Journal, 38(1), 191-199.

Jamiolkowski, M., Lancellotta, R., and Lo Presti, D. C. F. (1995). "Remarks on the stiffness at small
strains of six Italian clays." Proc., Pre-Failure Deformation Characteristics of Geomaterials, Balkema,
817-836.

Kim, H. S., Cho, G. C., Lee, J. Y., and Kim, S. J. (2013). "Geotechnical and geophysical properties of
deep marine fine-grained sediments recovered during the second Ulleung Basin Gas Hydrate expedition,
East Sea, Korea." Marine and Petroleum Geology, 47, 56-65.

Kim, H. K., and Santamarina, J. (2008). "Sand-rubber mixtures (large rubber chips)." Canadian
Geotechnical Journal, 45(10), 1457-1466.

Ku, T., Mayne, P. W., and Gutierrez, B. J. "Hierarchy of Vs Modes and Stress-Dependency in
Geomaterials." Proc., International Symposium on Deformation Characteristics of Geomaterials,
September 1~3, 2011, 533-540.

Kwon, T. H., Lee, K. R., Cho, G. C., and Lee, J. Y. (2011). "Geotechnical properties of deep oceanic
sediments recovered from the hydrate occurrence regions in the Ulleung Basin, East Sea, offshore Korea."
Marine and Petroleum Geology, 28(10), 1870-1883.

Lee, C., Truong, Q. H., Lee, W., and Lee, J. S. (2010). "Characteristics of Rubber-Sand Particle Mixtures
according to Size Ratio." J. Mater. Civ. Eng., 22(4), 323-331.

Lee, J. S., Guimaraes, M., and Santamarina, J. C. (2007). "Micaceous sands: Microscale mechanisms and
macroscale response." Journal of Geotechnical and Geoenvironmental Engineering, 133(9), 1136-1143.

Ng, C. W. W., Leung, E. H. Y., and Lau, C. K. (2004). "Inherent anisotropic stiffness of weathered
geomaterial and its influence on ground deformations around deep excavations." Canadian Geotechnical
Journal, 41(1), 12-24.

www.ascelibrary.org © ASCE 2014 / S3


SUPPLEMENTAL DATA

Oh, T. M. (2008). "Undrained shear strength estimation of marine clay using electrical resistivity and
shear wave velocity." M. Sc. Thesis, KAIST, Korea.

Patel, A., Bartake, P. P., and Singh, D. N. (2009). "An Empirical Relationship for Determining Shear
Wave Velocity in Granular Materials Accounting for Grain Morphology." Geotech. Test. J., 32(1), 1-10.

Santagata, M., Germaine, J. T., and Ladd, C. C. (2005). "Factors affecting the initial stiffness of cohesive
soils." Journal of Geotechnical and Geoenvironmental Engineering, 131(4), 430-441.

Santamarina, J. C., Klein, K. A., and Fam, M. A. (2001). Soils and waves - Particulate materials
behavior, characterization and process monitoring, Wiley, Chichester, England.

Shin, H., and Santamarina, J. C. (2013). "Role of Particle Angularity on the Mechanical Behavior of
Granular Mixtures." Journal of Geotechnical and Geoenvironmental Engineering, 139(2), 353-355.

Stokoe, K. H., Lee, J. N. K., and Lee, S. H. H. (1991). "Characterization of soil in calibration chambers
with seismic waves." Proc., International Symposium on Calibration Chamber Testing, Postdam, New
York.

Yamashita, S., Hori, T., and Suzuki, T. (2003). Effects of fabric anisotropy and stress condition on small
strain stiffness of sands, A a Balkema Publishers, Leiden.

Yun, T. S., and Santamarina, J. C. (2005). "Decementation, Softening, and Collapse: Changes in Small-
Strain Shear Stiffness in k0 Loading." Journal of Geotechnical & Geoenvironmental Engineering, 131,
350-358.

S4 / © 2014 ASCE www.ascelibrary.org

You might also like