Sei sulla pagina 1di 11

Small-Strain Behavior of Granular Soils.

I: Model
for Cemented and Uncemented Sands and Gravels
Juan M. Pestana, M.ASCE1; and Lynn A. Salvati, M.ASCE2

Abstract: A simple formulation is presented that predicts the nonlinear small strain behavior of cemented and uncemented granular soils.
Its performance is evaluated through the comparison of model predictions to results from laboratory tests. A companion paper evaluates
the performance of this model implemented in a site response analysis code through comparison with the measured response at two sites.
The formulation for the maximum shear modulus, Gmax, which is selected through the evaluation of existing formulations and data, is
presented with the hysteretic model developed to describe the shear modulus reduction and damping increase with increasing strains. Few
parameters are needed to predict the small strain response, and correlations between model parameters and index properties of granular
materials are presented when possible. The model, SimSoil, is shown to capture the cyclic response for sands and gravels with varying
densities over a wide range of pressures measured in laboratory tests, including cases when cementation is present.
DOI: 10.1061/共ASCE兲1090-0241共2006兲132:8共1071兲
CE Database subject headings: Shear modulus; Constitutive models; Cementation; Granular materials.

Introduction medium dense to dense sand with gravel and stiff clay, was
230– 300 m deep. The 1989 Loma Prieta earthquake and the 1994
Site response analysis is an essential element in geotechnical Northridge earthquake provided more examples of deep, stiff soil
earthquake engineering. The equivalent linear procedure in the deposits significantly amplifying ground motion. These events
program SHAKE 共Schnabel et al. 1972兲 has been used in practice have shown the need for methods to predict the response at deep,
to approximate the nonlinear properties of soil and perform site stiff sites, since many cities are located on such deposits. There-
response analyses for over 30 years. However, as the field of fore a simple formulation is developed that can describe the maxi-
mum shear modulus, shear modulus reduction, and damping of
earthquake engineering progresses, the limitations of SHAKE
granular materials over a wide range of densities and pressures
become more restrictive. As a result, the use of nonlinear site
that is appropriate for use in nonlinear site response analysis. This
response analyses has increased. However, to obtain the benefits
formulation provides more control over the shear modulus reduc-
from performing a nonlinear analysis, a model that can accurately
tion and damping curves than hyperbolic models employed in
represent the nonlinear behavior of the soil must be used. nonlinear site response analysis codes such as DESRA 共Lee and
Most response analyses of stiff soil sites do not include depths Finn 1978兲, while requiring only a few parameters unlike the
exceeding 50 m. As an example, the 1997 Uniform Building models employed in nonlinear site response codes such as
Code only considers the top 30 m 共⬃100 ft兲 when developing SUMDES 共Li et al. 1992兲.
characteristic soil profiles for stiff materials, following the re- The first section of the paper describes the selected formula-
commendations by Borcherdt 共1994兲. Similar to deep, soft clay tion for the maximum shear modulus, Gmax. The model predic-
deposits, deep deposits of primarily dense granular material can tions are compared with available data for several soils over a
significantly amplify ground motions. The 1967 Caracas, Venezu- wide range of conditions, including very high pressures, and data
ela earthquake, which provided undisputed evidence of the effect for cemented sands. In the second section of the paper, the hys-
of “local soil conditions” on structural response, was also the first teretic formulation that describes the reduction in shear modulus
event to focus attention on the amplification potential of stiff and increase in damping with increasing shear strain is detailed.
soils. Much of the damage from that event occurred in the Palos The model predictions are again validated with laboratory test
Grandes area, where the thickness of alluvium, which consisted of data for sands and gravels, cemented and uncemented, and corre-
lations between model parameters and index properties of the
1
Associate Professor, Dept. of Civil and Environmental Engineering, materials are developed.
Univ. of California, Berkeley, CA 94720.
2
Clare Boothe Luce Assistant Professor, Dept. of Civil Engineering
and Geological Sciences, 156 Fitzpatrick Hall, Univ. of Notre Dame, Small Strain Stiffness
Notre Dame, IN 46556.
Note. Discussion open until January 1, 2007. Separate discussions
must be submitted for individual papers. To extend the closing date by Previous Investigations
one month, a written request must be filed with the ASCE Managing
Editor. The manuscript for this paper was submitted for review and pos- As the database of laboratory test results has increased over the
sible publication on July 11, 2003; approved on September 29, 2005. This years, many empirical relationships have been developed to pre-
paper is part of the Journal of Geotechnical and Geoenvironmental dict the maximum shear modulus, Gmax. For sand, Gmax is prima-
Engineering, Vol. 132, No. 8, August 1, 2006. ©ASCE, ISSN 1090- rily controlled by the density and confining pressure, which has
0241/2006/8-1071–1081/$25.00. been represented in formulations for Gmax dating from the 1960s,

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1071


Fig. 1. Effect of confining pressure on Gmax for 共a兲 Monterey No. 0
sand; 共b兲 selected gravels

e.g., Hardin and Richart 共1963兲; Seed and Idriss 共1970兲. If aniso-
tropy is not considered, a general formulation for Gmax can be Fig. 2. 共a兲 Effect of particle angularity on Gmax for selected uniform
proposed cohesionless soils; 共b兲 effect of density on Gmax for selected sands
Gmax
pat
冋 册
= Gb · f 1关e兴·f 2
p
pat
共1兲
and gravels

where p = mean effective stress; pat = atmospheric pressure; have problems describing the shear modulus of sands that can
e = void ratio; and Gb = material constant. All of the formulations have high formation void ratios, such as Dog’s Bay sand.
developed from 1960 to 2000 that were reviewed 共Salvati 2002兲
use a power law to describe the effect of confining pressure on
Selected Formulation
Gmax. Formulations that prescribe the exponent for the power law
are in a well-bounded range of 0.4–0.6 with the exception of the Fig. 1共a兲 compares predictions of Gmax using three different
value of 0.33. The 0.33 value suggested by Pestana and Whittle power-law exponents with measured values of Gmax for Monterey
共1995兲 and later used by others 共Assimaki et al. 2000; Kausel and sand. The power-law exponent, n = 0.5 tends to fit the data for
Assimaki 2002兲 is based on the Hertzian contact theory for elastic uniform sands better than n = 0.4, which was suggested by Iwasaki
compression of spheres 共Mindlin and Deresiewicz 1953兲. The ef- and Tatsuoka 共1977兲, or n = 0.33 共Pestana and Whittle 1995兲.
fect of void ratio is not as uniformly described, however. It is Based on its wide range of applicability 共many other materials
desirable for Gmax to approach zero as the void ratio increases to examined were not included here for conciseness兲 a power-law
a limiting value. Some of the relations do not uphold this, and exponent n = 0.5 can be used as a default value in the Gmax rela-

1072 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006


Table 1. Selected Model Parameters and Index Properties for Sands
Sand Source ANGa Cu Gb ␻s ␻a
Antewerp 共Yoon and VanImpe 1995兲 R 1.5 420–460
Dog’s Bay 共Jovicic and Coop 1997兲 A 2.4 1,800
Ham River 共Jovicic and Coop 1997兲 SR 1.6 550
Hokksund 共Carriglio 1989兲 SA 2.0 560–570
Iruma 共Iwasaki and Tatsuoka 1977兲 690
Mol 共Yoon and VanImpe 1995兲 SA-A 1.6 655–670
Monterey No. 0 共Chung et al. 1984兲 R-SR 1.5 420–500 1.5
共Saxena and Reddy 1989兲
Mortar 共Laird and Stokoe 1993兲 2.0 625 1.75
Ottawa 共Alarcon-Guzman et al. 1989兲 R 1.2 475–500 1.25 3.0
共Hardin et al. 1994兲
Quiou 共LoPresti et al. 1993兲 A 4.4 660–700 3.5 2.0
共LoPresti et al. 1997兲
共Fioravante et al. 1994兲
Reid-Bedford 共Skoglund et al. 1976兲 SA-SR 1.7 580
Sacramento 共Salvati 2002兲 SA-SR 1.3 400
Sand, 0–2 mm 共Souto et al. 1994兲 2.8 370–390 2 2
Ticino 共Carriglio 1989兲 SR 1.6 580 1.5
共LoPresti et al. 1993兲
Toyoura 共LoPresti et al. 1997兲 SA 1.5 700-720 1.0 1.0
共Bellotti et al. 1997兲
共Iwasaki and Tatsuoka 1977兲
共Kokusho 1980兲
共Yasuda and Matsumoto 1993兲
a
ANG= angularity; A = angular; SA= subangular; SR= subrounded; and R = rounded.

tionship for sands, unless there is data that indicates a different Jamiolkowski et al. 共1991兲, f 1关e兴. The f 2关e兴 function does not
value would be more appropriate. Gravels and materials with a describe the trend of the data as well and f 3关e兴 does not match
significant amount of gravel tend to have higher values of n the data at higher void ratios as well. Using the formulation
共n = 0.5– 0.75兲, as shown in Fig. 1共b兲. f 1关e兴 = e−1.3, all of the different materials can be described with a
To isolate the effect of void ratio on Gmax, only values of Gmax material constant, Gb, that is within a well-bounded range. Fur-
measured at a mean effective stress of within 10% of 101 kPa thermore, the value of Gb seems to be well-correlated with the
共1 atm兲 were analyzed initially. In Fig. 2共a兲, the test data is shown angularity of the material for fairly uniform sands; the sands with
with three functions used to describe the effect of void ratio on more angular grains tend to have higher values of Gb. Sands and
Gmax: f 1关e兴 = e−1.3 共Jamiolkowski et al. 1991兲, f 2关e兴 = 共1 + e兲 / e gravels, which are very well graded 共Cu ⬎ 7兲 tend to have lower
共Pestana and Whittle 1995兲, and f 3关e兴 = 共2.17− e兲2 / 共1 + e兲 共Hardin values of Gb 共Gb = 200– 400兲. In Fig. 2共b兲, reported Gmax values at
and Richart 1963兲. Both f 2关e兴 and f 3关e兴 are shown with typical all measured confining pressures are shown with the formulation
values of the material constant for those formulations, whereas proposed by Jamiolkowski, f 1关e兴 = e−1.3, after normalizing the data
f 1关e兴 is shown with a range of Gb values from 400 to 800. The by the confining pressure and Gb for the material. Almost all of
trend of the data is best matched by the formulation presented by the data is within the 20% error bars shown. Based on this ex-

Table 2. Selected Model Parameters and Index Properties for Gravels


Gravel Source ANGa Cu Gb ␻s ␻a
Folsom 共Hynes-Griffin 1988兲 75.8 430 4.0 2.0
Livermore 共Seed et al. 1984兲 R 62.5 235 5.0 4.0
NM 共Hatanaka et al. 1999兲 15.0 245–260
Oroville 共Seed et al. 1984兲 SR-SA 27.2 360 3.0 2.0
Pyramid 共Seed et al. 1984兲 A 7.1 740 5.0 2.0
R-1 and R-2 共Yasuda et al. 1996兲 R-SR 45.0 160–180 3.0 1.0
R-3 共Yasuda et al. 1996兲 R-SR 114.0 230 3.0 1.0
Rockfill 共Yasuda and Matsumoto 1993兲 7.0 470–550 4.5 1.0
Venado 共Seed et al. 1984兲 A 750 5.0 5.0
a
ANG= angularity; A = angular; SA= subangular; SR= subrounded; and R = rounded.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1073


冤 冉 冊 冉 冊冉 冊 冥
amination of the available data the maximum shear modulus, Kmax e1.3␦␪b ␳c共1 − ␦␪b兲 −1

Gmax, for uncemented cohesionless materials is best represented = n + 共3兲


pat p p e
by the following formulation: Kb
pat pat 1 + e

Gmax
pat
= Gbe−1.3 冉 冊
p
pat
n
共2兲 where ␦b = 1 − p / pb; pb = pref共1 / e兲1/␳c; and pref, ␪, ␳c, and
Kb = material constants. The equivalent stress on the limiting com-
pression curve 共LCC兲 at a given void ratio is pb, and the dimen-
sionless measure of distance between the current mean effective
Gb values for the proposed formulation were determined for many
stress and the stress on the LCC is given by ␦b.
commonly used test sands and gravels based on published data, Jovicic and Coop 共1997兲 performed bender element tests on
and the values are listed in Tables 1 and 2. Dog’s Bay sand and Ham River sand isotropically compressed to
very high pressures. When the compression curves are plotted in
Shear Stiffness at High Pressures log e − log p space, as in Fig. 3, it can be seen that both of these
materials have reached the LCC during compression. Therefore
The previous section discussed the elastic shear modulus, which the maximum shear modulus cannot be determined simply using
describes deformation in the soil that is solely due to the elastic the elastic formation, explaining the nonlinearity of the plot of
compression of the soil skeleton. However, as the confining stress Gmax in log G-log p space. Using the compression curves that
increases, particle rearrangement begins, and plastic strains begin were presented by Jovicic and Coop 共1997兲 and the data pre-
to accrue. Then, as the increase in stresses continues, the particles sented by Pestana 共1994兲, ␳c, pref, and ␪ were determined for both
begin to crush and the strains become increasingly dominated by sands and are shown in the figures. Kb was determined from the
the plastic contribution. Using the formulation discussed above to reported response at low stress levels. Then the compression
describe the elastic shear modulus of the soil within the frame- model was used to predict Kmax values with increasing confining
work of the compression model developed by Pestana and Whittle pressure for the lowest and highest initial void ratios tested for
共1995兲 the plastic and elastoplastic stiffness can be described as each of these sands. The Gmax values are calculated from the Kmax
well as the elastic stiffness. The functions used to describe the values based on the Poisson’s ratio. The two sets of predictions
effect of confining pressure and void ratio on the elastic bulk 共for each void ratio兲 are shown in Fig. 3: one in which the Pois-
modulus, K, have been modified from those used by Pestana and son’s ratio is assumed to be constant, and a second in which the
Whittle 共1995兲, as shown below, based on the formulation pre- Poisson’s ratio,␮, is assumed to vary with the void ratio as shown
sented in the previous section in the equation below

Fig. 3. Comparison of measured and predicted Gmax values over a large stress range

1074 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006


␮ = ␮ref关exp共1 − e兲兴 共4兲 cementation bonds are broken. Another important consequence of
cementation is that the influence of confining pressure on the
where ␮ref = value of ␮ at e = 1. The model predictions for Ham maximum shear modulus decreases as the level of cementation
River sand and Dog’s Bay sand bound the results over the large increases. If there is enough cementation, confining pressure does
range in confining pressures well. The model predictions based on not have a significant effect on the shear modulus, unless it ex-
the variable Poisson ratio are closer to the measured values, and ceeds the pressure necessary to break the cementation bonds.
indicate the importance of the changing void ratio when the load- Therefore to describe the maximum shear modulus for cemented
ing pressures encompass several orders of magnitude. granular materials, the following formulation is provided:

Cemented Sands and Gravels


Cementation can occur naturally or can be the result of a soil
Gmax
pat
= Gbe−1.3
p

pat
+ accCC2 冊 n
共5兲

improvement technique. There is only limited data, in either case,


to describe the effect that cementation has on the dynamic prop- where CC= percent cement by weight; and acc = constant to de-
erties of sand. Much of the testing on cemented sands has been scribe the cementing agent and process. The gradation properties
performed on resonant column devices. However, comparison of of the sand and the density of the sand may also affect Gmax for
the measured maximum shear modulus and Young’s modulus for cemented sands 共Chang and Woods 1992兲, but this effect is small
a given sample has shown that unreasonable values of the Pois- compared to the level of cementation, and therefore is not in-
son’s ratio result unless the samples are adequately fixed to the cluded in at this time. Fig. 4共a兲 illustrates the decreasing influence
end platens 共Lovelady and Picornell 1990; Baig et al. 1997兲. This of confining pressure on Gmax for sands with increasing levels of
indicates that Gmax values presented in much of the early research cementation predicted by the formulation, and Fig. 4共b兲 compares
are not accurate, especially at lower confining pressures. There- model predictions for Gmax with increasing confining pressure
fore the database of results that can be used to describe the effects with test data for cemented Ottawa sand at two relative densities
of cementation is extremely limited. Clearly, cemented sands are 共Baig et al. 1997兲. For these cemented sands, the density actually
stronger and stiffer than their uncemented counterparts until the has a larger effect on Gmax than the confining pressure. This data
set was selected for comparison because it is more complete with
respect to the influence of density and confining pressure on the
small strain shear modulus of cemented sands than the studies
presented by Huang and Airey 共1998兲 or Delfosse-Ribay et al.
共2004兲.

Small Strain Nonlinearity

Describing the response of a soil to cyclic loading, such as an


earthquake, also requires an accurate description of the reduction
of shear modulus and the increase in damping as the strain in-
creases. For uncemented granular materials the confining pressure
has the largest effect on the shear modulus and damping. As the
confining pressure increases, the linear elastic region for granular
materials will extend to larger strains, which can be an important
phenomenon when dealing with deep deposits. The density of the
granular material also plays a role in the small strain nonlinearity,
but the effect is not nearly as pronounced.

Previous Formulation
Pestana and Whittle 共1999兲 used the perfectly hysteretic formula-
tion in the generalized constitutive model, MIT-S1, to predict the
modulus reduction and damping curves. In the perfectly hysteretic
formulation proposed by Hueckel and Nova 共1979兲 the isotropic
tangential moduli are a function of the most recent stress reversal
state. The concept was developed to describe unload-reload
cycles for clay, and based on this relation the reversibility of
strain was guaranteed over a stress loop between two stress rever-
sal points. The modulus reduction at a constant mean effective
stress in MIT-S1 is described by

Gtan 1
= 共6兲
Gmax 共1 + ␻␰s兲共1 + ␻s␰s兲
Fig. 4. Effect of light cementation on confining stress dependence of
Gmax: 共a兲 model predictions; 共b兲 model predictions and measured where ␻ and ␻s = constants for a given material; and
values for Ottawa sand ␰s = dimensionless stress measure. ␰s is given by

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1075


␰s = 储␩ − ␩rev储 共7兲
where ␩ = stress ratio, which is the ratio of deviatoric to mean
effective stress; and ␩rev = stress ratio at the most recent stress
reversal point. Since the soil nonlinearity can be described best by
the strain history 共Hardin and Drnevich 1972; Hight et al. 1983兲,
the stress reversal point 共SRP兲 and loading condition are defined
by the strain direction. The stress reversal point is set at the tran-
sition between loading and unloading, and the loading/unloading
condition is based on the vector product of the accumulated strain
共from the last reversal point兲, ␹, and the incremental strain, ␹˙ , as
shown below

␹:␹˙ = 再 艌0 loading
⬍0 unloading
冎 共8兲

This formulation describes the unload-reload cycles in monotonic Fig. 5. Translation of hysteresis loop
loading well, and can predict reasonable behavior for cyclic tests
on soils. However, when describing the response to cyclic load-
ing, the parameters do not provide adequate control of the modu- is capped so that it cannot exceed 0.005 to limit its effect in the
lus reduction and damping curves, since the parameters do not intermediate to large strain range. Masing rules can be used to
provide independent control of the response at larger strains. describe the hysteresis loops, and the loading/unloading condition
Therefore it is difficult to model the hysteretic behavior of a wide is the same as given above.
range of soils well. Since the objective of this paper is to describe a simple model
that can be used to evaluate the site response at deep sites due to
vertically propagating shear waves, the formulation can be sim-
New Formulation
plified for this purpose. The analytical solutions, which can be
A new formulation was developed to describe the nonlinear re- used to determine shear modulus reduction and damping curves,
sponse of soils. This formulation, which is shown below, provides are developed and shown below for the case in which pore pres-
better control of the shear modulus reduction curves without an sure changes are not considered. For one-dimensional 共1D兲 site
increase in the number of parameters. It also provides improved response analyses, ␰s = 冑2␶ / p, where ␶ = horizontal shear stress;
prediction of damping at very low strains, which is a difficulty and p = mean effective stress, which is considered to be constant.
with hysteretic models, such as the one described above. These The factor of 冑2 is a result of the transformed variables that are
improvements result in the ability to model a much wider range of used by Pestana 共1994兲 and the MIT family of models 共Potts and
materials with more accuracy than with the previous formulation. Zdravkovic 1999兲. To find the damping and shear modulus reduc-
Also, the new formulation provides better predictions of the shear tion curves, it is convenient to translate the hysteresis loop so that
modulus and damping at high pressures, which are necessary to it begins at the origin as shown in Fig. 5. For the backbone curve,
model deep deposits accurately the reversal state is set at zero 共␰rev = 0 and ␶rev = 0兲, and Eq. 共9兲
Gtan 1 can be rewritten as
= 共9兲
Gmax 1 + ␻s␰s0.75
* + ␻ s␰ s + ␻ a␰ s
2 2 Gtan 1
= 共10兲
Gmax 1 + ␻s共c␶*兲 + ␻sc␶ + ␻2a共c␶兲2
0.75
where ␻s and ␻a = material parameters, which describe the non-
linearity. The parameter, ␻a, controls the intermediate to large where c = 冑2 / p, and ␶* is limited to a maximum value of 0.005/c.
strain behavior; ␻s controls the small to intermediate strain be- The definition of tangent shear modulus, Gtan = d␶ / d␥, is used to
havior. The definition of ␰s is given above, and the parameter ␰s* solve for the shear strain, which is given in Eq. 共11兲

冋 册
冦 冧
␻s共c␶*兲1.75 ␻s ␻2a
c1 c␶ + + 共c␶兲2 + 共c␶兲3 if 共c␶兲 艋 0.005
1.75 2 3

冋 册
␥= 共11兲
␻s共c␶*兲1.75 ␻s ␻2a
c1 c␶ + + 共c␶兲2 + 共c␶兲3 + ␻s共c␶*兲0.75c␶** otherwise
1.75 2 3

where c1 = p / 冑2Gmax and ␶** = ␶ − ␶*. Since the hysteresis loop is translated and the reversal state is set at zero, the shear strain calculated
in Eq. 共11兲 is taken as the double amplitude shear strain. The equivalent viscous damping ratio 共␭兲 is a function of the ratio of energy
dissipation 共⌬W兲 in one hysteretic cycle to the maximum stored energy 共W兲. For models in which the shear strain is expressed in terms
of shear stress, damping can be evaluated as 共Ishihara 1996兲


冤 冥
␶a
2 ␥共␶兲d␶
1 ⌬W 2 0
␭= = 1− 共12兲
4␲ W ␲ ␶a␥共␶a兲

1076 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006


冉 冊再 冋 冉 冊 册冎
冦 冧
2 2c1 c␶ ␻s共c␶*兲1.75 ␶* ␻s共c␶兲2 ␻2a共c␶兲3
1− + + + if 共c␶兲 艋 0.005
␲ ␥ 2 1.75␶ 2.75 6 12

冉 冊再 冋冉 册冎
␭= 共13兲
2

1−
2c1

c␶ ␻s共c␶兲2 ␻2a共c␶兲3
2
+
6
+
12

␶+
␻s共c␶*兲1.75 ␶*
1.75␶

2.75
+ ␶** + 冊
␻s共c␶*兲0.75c␶2**
2␶
otherwise

With Eqs. 共11兲 and 共13兲 the damping curves can be developed, range of pressures and/or void ratios. However, if the shear modu-
and the secant shear modulus needed for the shear modulus re- lus is not measured at small enough strains to be considered Gmax,
duction curves is given by Gsec = ␶a / ␥a, as shown in Fig. 5. then the Gb value described earlier can be used to calculate Gmax
Fig. 6共a兲 shows the influence of ␻s on the shear modulus reduc- and the model predictions of shear modulus versus shear strain
tion and damping curves. As ␻s increases, the shear modulus can be used to compare with the available data. Then an appro-
decreases and the damping increases for a given strain. In many priate value of ␻s can be determined simply by comparing the
of the following figures, the ratio of shear modulus to a reference
SimSoil model predictions to the measured data. The parameter
shear modulus, Gref, is shown. Gref is the shear modulus at a shear
␻a is selected in a similar manner to ␻s, if there is data at large
strain of 0.0001%. The influence of ␻a is similar to ␻s, but the
enough strains to compare to the predictions.
effects are seen only at larger strains as shown in Fig. 6共b兲. With
these two parameters, the model can predict the behavior over a
wide range of stresses 关see Fig. 6共c兲兴. Also, since the parameters Model Evaluation
are based only on the material, the effects of varying density can
be described with the same single set of parameters as shown in Fig. 7 shows the model predictions, measured shear modulus, and
Fig. 6共d兲. This hysteretic formulation together with the formula- damping values for Monterey No. 0 sand over a range of pres-
tion for Gmax comprises a model, SimSoil, that can predict the sures. Considering the range in the results reported by Chung
small strain behavior of granular materials. et al. 共1984兲 and Saxena and Reddy 共1989兲, the SimSoil model
fits the data well on average. Comparisons of the model predic-
Parameter Selection tions with available test data for Toyoura sand from Kokusho
The parameters ␻s and ␻a can be determined using modulus re- 共1980兲, as seen in Fig. 8, demonstrate that the effect of confining
duction and/or damping data from any test if the general form of pressure is much greater than the effect of density on the small
the model is used. If the simplified version of the model is used strain nonlinearity, and the SimSoil model captures that trend.
关Eqs. 共11兲–共13兲兴, it is most appropriate to select the parameters The model can match shear modulus reduction and damping data
from resonant column, torsional, or simple shear tests. Normaliz- ranging over almost two orders of magnitude in pressure in the
ing the curves 共Gsec / Gmax versus shear strain兲 simplifies the de- case of Mortar sand 共Laird and Stokoe 1993兲, as shown in
termination of parameters for a material with test data over a Fig. 9共a兲. This formulation can also be used to describe the be-

Fig. 6. 共a,b兲 Effect of material parameters describing the nonlinearity in shear and 共c,d兲 predicted effect of confining pressure and density on the
shear modulus and damping

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1077


havior of gravels. Fig. 9共b兲 shows modulus reduction and damp-
ing data measured for rockfill by Yasuda and Matsumoto 共1993兲
with the model predictions.
The test data discussed above and the parameters determined
for them are listed in Tables 1 and 2. Although the comparison of
model predictions to measured laboratory results were not shown
for some materials for conciseness, the model parameters for
these materials are still included in Tables 1 and 2. Overall, a
wide range of granular materials with differing angularities and
grain sizes are well-described by the model with a small range of
parameters. The SimSoil model predictions for small strain non-
linearity of granular materials highlight the flexibility of the for-
mulation and its ability to simulate materials of varying density
over a range of confining pressure with a single set of parameters.
Preliminary findings indicate that ␻s is well-correlated to the uni-
formity coefficient, Cu, of the material, as shown in Fig. 10, for
fairly uniform materials. Although the correlation does not work
as well with gravels or sands with a Cu greater than 7, it does
provide a way to estimate ␻s in the absence of test data.

Cemented Soils
As the cementation level of a material increases, it becomes more
brittle, and therefore begins to degrade at a much lower strain
level than an uncemented material. The increasing cementation
also results in an increase in damping at lower strain levels than
in the uncemented material. These trends are illustrated in the
work done by Saxena et al. 共1988兲 and the model predictions as
shown in Fig. 11. Further evidence of the model’s applicability to
cemented materials can be shown in conjunction with the results
Fig. 7. Comparison of measured and predicted values of shear from Yasuda et al. 共1996兲. In this study both undisturbed and
modulus and damping for Monterey No. 0 sand reconstituted samples of gravel were tested. From the shear
modulus measurements for the gravel samples during loading and
unloading, it is evident that the lower gravel samples have some
level of cementation, most probably due to aging effects. Since
the gradation of the reconstituted sample, R-3, is the most similar
to the lower gravel, it is used for comparison with the undisturbed

Fig. 8. Comparison of measured and predicted values of shear modulus and damping for Toyoura sand

1078 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006


samples, and the properties determined for it are used to represent
the undisturbed samples as well. Assuming a value of acc = 16 and
CC= 0.5%, the SimSoil model would predict similar results to
those observed in the lower level gravel. Fig. 12 compares the
model predictions for cement contents of 0 and 0.50% to the
shear modulus and damping values measured for R-3 共reconsti-
tuted兲, UU-1 共undisturbed upper layer兲, and UL-1 共undisturbed
lower level兲. It is not surprising that the measured values for
UU-1 are very close to those measured for R-3 and the predic-
tions for CC= 0%, given that there was little evidence of cemen-
tation for the upper level gravel samples from the stiffness values
measured in loading and unloading, which is consistent with the
age of the deposit. The predicted values of the shear modulus at
the confining pressures of 290 and 590 kPa are reasonable for all
of the samples, but are higher than the measured values for UL-1

Fig. 10. Empirical correlation for model parameter, ␻s, for poorly
graded granular soils

at 50 kPa. Significantly higher damping values are noticeable in


the test results for the cemented gravels tested at 50 kPa only, but
are predicted at all pressures. As was the case with the cemented
Monterey No. 0 sand, the SimSoil model predictions capture the
trend of the measured results, but there is a clear need for addi-
tional data to verify this aspect of the model.

Fig. 9. Comparison of measured and predicted values of shear Fig. 11. Comparison of measured and predicted values of shear
modulus and damping for 共a兲 mortar sand; 共b兲 rockfill modulus and damping for lightly cemented Monterey No. 0 sand

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1079


can be a valuable component of site response analyses at deep,
stiff soil sites. Overall it provides a simple but accurate way of
describing soil behavior that can be used to predict site response
analysis in a code such as AMPLE2000 共Pestana and Nadim
2000兲, which will be discussed in the companion paper.

Acknowledgments

The writers would like to acknowledge the support provided by a


National Science Foundation Graduate Research Fellowship and
the National Science Foundation, CAREER award CMS-963979.

Notation

The following symbols are used in this paper:


acc ⫽ material constant describing the cementing agent
and process;
CC ⫽ cement content 共by weight兲;
Cu ⫽ uniformity coefficient;
e ⫽ void ratio;
Gb ⫽ material constant describing the maximum shear
modulus;
Gmax ⫽ maximum shear modulus;
Gref ⫽ shear modulus at 1 ⫻ 10−4% strain;
Gsec ⫽ secant shear modulus;
Gtan ⫽ tangent shear modulus;
Kb ⫽ material parameter to describing the maximum bulk
modulus;
Kmax ⫽ maximum bulk modulus;
n ⫽ material parameter describing the effect of pressure
on maximum shear modulus;
p ⫽ mean effective stress;
pat ⫽ atmospheric pressure;
pref ⫽ material parameter that gives the value of the LCC
at e = 1;
Fig. 12. Comparison of measured and predicted values of shear ␥ ⫽ shear strain;
modulus and damping for lightly cemented riverbed gravel ␦b ⫽ parameter describing the dimensionless distance to
the LCC;
␩ ⫽ stress ratio tensor, the ratio of deviatoric to mean
effective stress;
Summary
␩rev ⫽ stress ratio tensor at the reversal point;
␪ ⫽ material parameter describing the transition from
A new model, SimSoil, is developed that can describe the small
low to high stresses;
strain behavior of a wide range of granular materials with differ-
␭ ⫽ equivalent viscous damping ratio;
ing angularities and grain sizes. The form for the maximum shear
␮ ⫽ Poisson’s ratio;
stiffness was selected after an extensive review of available data
␮ref ⫽ value of Poisson’s ratio at e = 1;
and relationships. The robustness of the formulation is evidenced
␰s ⫽ dimensionless stress measure;
by its applicability to very high pressures and the ease with which
␰s* ⫽ minimum of ␰s and 0.005;
cemented sands can also be represented with the same formula-
␳c ⫽ slope of the LCC in log p – log e space;
tion. The reduction of the shear modulus and the increase in
␶ ⫽ horizontal shear stress;
damping with increasing shear strain for granular materials is
␶rev ⫽ shear stress at the reversal point;
described with a hysteretic formulation that is shown to match
␹ ⫽ strain accumulated since the last reversal point;
measured values over a wide range of confining pressures and
densities with a single set of parameters for a material. With only ␹˙ ⫽ increment of strain;
a small addition to the formulation, the SimSoil model can predict ␻a ⫽ material constant to describe soil nonlinearity at
the effect of cementation on the shear modulus and damping, as intermediate to large strains; and
well. The ability of the model to capture the response of a large ␻s ⫽ material constant to describe soil nonlinearity at the
range of materials is complemented by correlations that can be small to intermediate strains.
used to estimate the model parameters based index properties of
the soil when necessary. Therefore, if small strain test data is not References
available, reasonable estimates of the small strain nonlinearity of
the material can be made. Since this model can match the shear Alarcon-Guzman, A., Chameau, J. L., Leonards, G. A., and Frost, J. D.
modulus and damping for soils over a wide range of pressures, it 共1989兲. “Shear modulus and cyclic undrained behavior of sands.”

1080 / JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006


Soils Found., 29共4兲, 105–119. Lee, M. K. W., and Finn, W. D. L. 共1978兲. “Dynamic effective stress
Assimaki, D., Kausel, E., and Whittle, A. J. 共2000兲. “Model for dynamic response analysis of soil deposits with energy transmitting boundary
shear modulus and damping for granular soils.” J. Geotech. Geoenvi- including assessment of liquefaction potential.” Rep. 36 Soil Mechan-
ron. Eng., 126共10兲, 859–869. ics Series, Univ. of British Columbia, Faculty of Applied Science,
Baig, S., Picornell, M., and Nazarian, S. 共1997兲. “Low strain shear Vancouver, B.C.
moduli of cemented sands.” J. Geotech. Geoenviron. Eng., 123共6兲, Li, X. S., Wang, Z. L., and Shen, C. K. 共1992兲. “SUMDES: A nonlinear
540–545. procedure for response analysis of horizontally layered sites subjected
Bellotti, R., Benoit, J., Fretti, C., and Jamiolkowski, M. 共1997兲. “Stiffness to multi-directional earthquake loading.” Dept. of Civil Engineering,
of Toyoura sand from dilatometer tests.” J. Geotech. Geoenviron. Univ. of California, Davis, Calif.
Eng., 123共9兲, 836–846. LoPresti, D. C. F., Jamiolkowski, M., Pallara, O., Cavallaro, A., and
Borcherdt, R. D. 共1994兲. “Estimates of site dependent response spectra Pedroni, M. 共1997兲. “Shear modulus and damping of soils.” Geotech-
for design 共methodology and justification兲.” Earthquake Spectra, nique, 47共3兲, 603–617.
10共4兲, 617–653. LoPresti, D. C. F., Pallara, O., Lancellotta, R., and Maniscalco, R. 共1993兲.
Carriglio, F. 共1989兲. “Caratteristiche sforzi-deformazioni-resistenza delle “Monotonic and cyclic loading behavior of two sands at small
sabbie.” Ph.D. thesis, Politecnico di Torino 共in Italian兲. strains.” Geotech. Test. J., 16共4兲, 409–424.
Chang, T. S., and Woods, R. D. 共1992兲. “Effect of particle contact bond
Lovelady, P. L., and Picornell, M. 共1990兲. “Sample coupling in resonant
on shear modulus.” J. Geotech. Eng., 118共8兲, 1216–1233.
column testing of cemented sands.” Dynamic elastic modulus mea-
Chung, R. M., Yokel, F. Y., and Drnevich, V. P. 共1984兲. “Evaluation of
dynamic properties of sands by resonant column testing.” Geotech. surements in materials, A. Wolfenden, ed., ASTM, West Consho-
Test. J., 7共2兲, 60–69. hocken, Pa, 180–193.
Delfosse-Ribay, E., Djeran-Maigre, I., Cabrillac, R., and Gouvenot, D. Mindlin, R. D., and Deresiewicz, H. 共1953兲. “Elastic spheres in contact
共2004兲. “Shear modulus and damping ratio of grouted sand.” Soil under varying oblique forces.” J. Appl. Mech., 20共3兲, 327–344.
Dyn. Earthquake Eng., 24, 461–471. Pestana, J. M. 共1994兲. “A unified constitutive model for clays and sands.”
Fioravante, V., Jamiolkowski, M., and LoPresti, D. C. F. 共1994兲. “Stiff- ScD thesis, Massachusetts Institute of Technology, Cambridge, Mass.
ness of carbonitic Quiou sand.” XIII Int. Conf. on Soil Mechanics and Pestana, J. M., and Nadim, F. 共2000兲. “Nonlinear site response analysis of
Foundation Engineering, ICSMFE, London, 163–167. submerged slopes.” Rep. UCB/GT/2000-04, University of California,
Hardin, B. O., and Drnevich, V. P. 共1972兲. “Shear modulus and damping Berkeley, Calif.
in soils.” J. Soil Mech. Found. Div., 98共7兲, 667–692. Pestana, J. M., and Whittle, A. J. 共1995兲. “Compression model for cohe-
Hardin, K. O., Drnevich, V. P., Wang, J., and Sams, C. E. 共1994兲. “Reso- sionless soils.” Geotechnique, 45共4兲, 611–631.
nant column testing at pressures up to 3.5 MPa 共500 psi兲.” Dynamic Pestana, J. M., and Whittle, A. J. 共1999兲. “Formulation of a unified con-
geotechnical testing II, R. J. Ebelhar, V. P. Drnevich, and B. L. Kutter, stitutive model for clays and sands.” Int. J. Numer. Analyt. Meth.
eds., ASTM, West Conshohoken, Pa. Geomech., 23共12兲, 1215–1243.
Hardin, B. O., and Richart, F. E. 共1963兲. “Elastic wave velocities in Potts, D. M., and Zdravkovic, L. 共1999兲. Finite element analysis in geo-
granular soils.” J. Soil Mech. Found. Div., 89共1兲, 33–65. technical engineering, Thomas Telford, London.
Hatanaka, M., Uchida, A., Taya, Y., Hagiwara, T., and Terui, N. 共1999兲. Salvati, L. A. 共2002兲. “Seismic response and cyclic behavior of sands.”
“Some factors affect the initial elastic modulus of sandy and gravelly Ph.D. thesis, Univ. of California, Berkeley, Calif.
soils measured in triaxial cell.” Proc., Second Int. Conf. on Earth- Saxena, S., Avramidis, A., and Reddy, K. 共1988兲. “Dynamic moduli and
quake Geotechnical Engineering, A. A. Balkema, The Netherlands, damping ratios for cemented sands at low strains.” Can. Geotech. J.,
59–64. 25共2兲, 353–368.
Hight, D. W., Gens, A., and Jardine, R. J. 共1983兲. “Evaluation of geo- Saxena, S., and Reddy, K. 共1989兲. “Dynamic moduli and damping ratios
technical parameters from triaxial tests on offshore clay.” Proc., So- for Monterey No. 0 sand by resonant column tests.” Soils Found.,
ciety for Underwater Technology Conf. on Offshore Site Investigation, 29共2兲, 37–51.
SUT, London, 253–268. Schnabel, P. B., Lysmer, J., and Seed, H. B. 共1972兲. “SHAKE: A com-
Huang, J. T., and Airey, D. W. 共1998兲. “Properties of artificially cemented puter program for earthquake response analysis of horizontally lay-
carbonate sand.” J. Geotech. Geoenviron. Eng., 124共6兲, 492–499. ered sites.” Rep. UCB/EERC 72-12, Univ. of California, Berkeley,
Hueckel, T., and Nova, R. 共1979兲. “Some hysteresis effects of the behav- Calif.
ior of geologic media.” Int. J. Solids Struct., 15共8兲, 625–642. Seed, H. B., and Idriss, I. M. 共1970兲. “Soil moduli and damping factors
Hynes-Griffin, M. E. 共1988兲. “Pore pressure generation characteristics for dynamic response analysis.” Rep. UCB/EERC-70/10, Univ. of
of gravel under undrained cyclic loading.” Ph.D. thesis, Univ. of California, Berkeley, Calif.
California, Berkeley, Calif. Seed, H. B., Wong, R. T., Idriss, I. M., and Tokimatsu, K. 共1984兲.
Ishihara, K. 共1996兲. Soil behavior in earthquake geotechnical engineer- “Moduli and damping factors for dynamic analyses of cohesionless
ing, Clarendon Press, Oxford. soils.” Rep.. UBC/EERC 84-14, Univ. of California, Berkeley,
Iwasaki, T., and Tatsuoka, F. 共1977兲. “Effects of grain size and grading on Calif.
dynamic shear moduli of sands.” Soils Found., 17共3兲, 19–35. Skoglund, G. R., Cunny, R. W., and Marcuson, W. F. 共1976兲. “Evaluation
Jamiolkowski, M., Leroueil, S., and LoPresti, D. C. F. 共1991兲. “Design of resonant column test devices.” J. Geotech. Eng. Div., Am. Soc. Civ.
parameters from theory to practice.” Proc., Int. Conf. on Geotechnical Eng., 102共11兲, 1147–1158.
Engineering for Coastal Development: Geo-Coast 1991, Coastal Souto, A., Hartikainen, J., and Ozudogru, K. 共1994兲. “Measurement of
Development Institute of Technology, Yokohama, Japan, 877–917. dynamic parameters of road pavement materials by the bender ele-
Jovicic, V., and Coop, M. R. 共1997兲. “Stiffness of coarse grained soils at ment and resonant column tests.” Geotechnique, 44共3兲, 519–526.
small strains.” Geotechnique, 47共3兲, 545–561. Yasuda, N., and Matsumoto, N. 共1993兲. “Dynamic deformation character-
Kausel, E., and Assimaki, D. 共2002兲. “Seismic simulation of inelatic soils istics of sands and rockfill materials.” Can. Geotech. J., 30共5兲,
via frequency-dependent moduli and damping.” J. Eng. Mech., 747–757.
128共1兲, 34–47. Yasuda, N., Ohta, N., and Nakamura, A. 共1996兲. “Dynamic deformation
Kokusho, T. 共1980兲. “Cyclic triaxial test of dynamic soil properties for characteristics of undisturbed riverbed gravels.” Can. Geotech. J.,
wide strain range.” Soils Found., 20共2兲, 45–60. 33共2兲, 237–247.
Laird, J. P., and Stokoe, K. H. 共1993兲. “Dynamic properties of remolded Yoon, Y. W., and VanImpe, W. F. 共1995兲. “Dynamic behavior of crushable
and undisturbed soil samples tested at high confining pressures.” sand.” First Int. Conf. on Earthquake Geotechnical Engineering,
Rep.GR93-6, Electric Power Research Institute, Palo Alto, Calif. A. A. Balkema, The Netherlands, 227–232.

JOURNAL OF GEOTECHNICAL AND GEOENVIRONMENTAL ENGINEERING © ASCE / AUGUST 2006 / 1081

Potrebbero piacerti anche