You are on page 1of 11

Determination of Viscoelastic Properties of Soil and

Prediction of Static and Dynamic Response


Abhijeet Swain1 and Priyanka Ghosh, Ph.D., A.M.ASCE2
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

Abstract: This paper employs a general and direct method to determine the time-dependent stress–strain behavior of soil using a Prony series
parameter fitting. The rheological stress–strain data obtained from the relaxation test of soil are used to evaluate the Prony series parameters
employing the method of nonlinear curve fitting. The optimized extraction of Prony series parameters for four different Indian soils is pre-
sented, which depicts the viscoelastic behavior of different soils considered in the study. Later, finite-element models are developed to simulate
and validate the time-dependent constitutive relationship of the soil under both static and dynamic loading conditions. Using such time-
dependent constitutive relationships of the soil, stress relaxation, creep, and steady-state dynamic response of the footing-soil system are
explored through some example problems. DOI: 10.1061/(ASCE)GM.1943-5622.0001456. © 2019 American Society of Civil Engineers.
Author keywords: Creep; Dynamic response; FE analysis; Stress relaxation; Viscoelastic.

Introduction dependent constitutive models have been effectively applied to


study the stress–strain behavior of soil subjected to static loads, such
The advancement in the field of geotechnical engineering has been as soil consolidation problems (Dey and Basudhar 2010; Mishra and
constantly engaged in explaining the science governing the com- Patra 2018; Taherkhani and Jalali 2018). Sun and Zhang (2007)
plex behavior of soil. This complexity involved in the material used the Voigt model to model the time-dependent behavior of soft
behavior of soil makes it quite unreliable in predicting its behavior soils. Zhu et al. (2012) have reported the settlement analysis of foun-
in various geotechnical engineering problems. Researchers over dation by modeling the soil with the use of the Kelvin-Voigt model.
time have developed different approaches (continuum and model- Later, the same Voigt model was adopted by various researchers
ing) for predicting the mechanical behavior of soil under different (Kakar 2016; Zheng et al. 2016; Kakar and Kakar 2017; Manna
loading conditions. The continuum approach is fast, adaptable, and et al. 2018) to solve some dynamic problems. Yin and Graham
accurate but limited to ideal cases. With increasing complexities of (1989) investigated the time-dependent behavior of clay to study the
soil behavior, this approach tends to involve more complex mathe- stress–strain behavior at different strains using the viscous–elastic–
matics. In contrast to this, the modeling approach is found to be sim- plastic model. Later, the elastic viscoplastic behavior of soft clay
pler, involving simple mathematics for determining the modeling was explored by different researchers (Yin 2015; Yin et al. 2015;
parameters. In several occasions, the soil is assumed to show a lin- Nash and Brown 2015; Azari et al. 2016; Zhang et al. 2018). Kimoto
ear or nonlinear elastic stress–strain behavior (Winkler’s model, et al. (2015) presented a cyclic elastoviscoplastic constitutive model
Hetenyi’s model, Pasternak model) neglecting the effect of time de- for clayey soil considering the nonlinear kinematic hardening rules
pendency in the constitutive relationship. This assumption simpli- and the structural degradation. Feng et al. (2017) reported an elabo-
fies the theoretical and the numerical modeling of soil, which yields rate study on the effect of time and strain rate on the viscous stress–
an acceptable prediction of the bearing capacity and the settlement strain behavior of plasticine materials (similar to soft clay). The diffi-
characteristics of soil in various soil–structure interaction problems. culty involved in the lumped parameter models is in the estimation
However, it is important to understand the behavior of soil at differ- of the correct values of different component parameters to replicate
ent time intervals, most critically in the case of expansive soils sub- the true mechanical behavior of the material under diversified load-
jected to static and dynamic loading. To capture such time- ing conditions.
dependent mechanical behavior of soil, several researchers (Tan The present study focuses on general but more versatile proce-
dures to determine the rheological properties of soil using a Prony
1957; Kerr 1961; Schiffman et al. 1966; Lyakhov 1968) have
series parameter fitting. Prony series is a generalized mathematical
adopted the concept of viscoelasticity. The idea of viscoelasticity
representation of a bunch of Maxwell elements (spring and dashpot)
is mainly limited to the Maxwell model (two-parameter), Kelvin-
in series and a spring in parallel to the whole array. The time-
Voigt model (two-parameter), Poynting-Thompson model (three-
dependent stress–strain test data (creep/relaxation data) of soil is
parameter), and Burger model (four-parameter). These time-
used to evaluate the Prony series parameters through curve fitting.
Relaxation tests have been performed on four different types of
1
Research Scholar, Dept. of Civil Engineering, Indian Institute of Indian soil, such as local soil from Kanpur, marine clay from
Technology Kanpur, Kanpur 208016, India. Vishakhapatnam, black cotton soil from Sagar, and kaolin clay. The
2
Professor, Dept. of Civil Engineering, Indian Institute of Technology relaxation test data obtained from different soils were used to
Kanpur, Kanpur 208016, India (corresponding author). Email: priyog@
extract the Prony-series parameters mathematically, which define
iitk.ac.in
Note. This manuscript was submitted on May 7, 2018; approved on
the stress–strain behavioral model of the respective soil in both time
January 24, 2019; published online on May 3, 2019. Discussion period and frequency domain. Commercially available finite-element (FE)
open until October 3, 2019; separate discussions must be submitted for software ABAQUS 6.14 was used to obtain two-dimensional (2D)
individual papers. This paper is part of the International Journal of and three-dimensional (3D) FE models for static and steady-
Geomechanics, © ASCE, ISSN 1532-3641. state dynamic analysis of the soil–structure interaction problems,

© ASCE 04019072-1 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


respectively. Finally, the time-dependent constitutive relationship modulus (ER). The linear viscoelastic system may consist of simple
obtained for the soil was validated with the experimental results models composed of springs and dashpots. The basic arrangement
obtained from a creep test (using oedometer test setup) and the of these components (spring and dashpot) in the Maxwell, Kelvin-
block vibration tests (Bhoumik 1989; Swain and Ghosh 2016) per- Voigt, and three-parameter models along with the constitutive stress
formed in Kanpur local soil. The simulated steady-state response –strain relationship is presented in Fig. 1(a).
curves for all four types of soil were compared to capture the diver- In Fig. 1(a), the three-parameter model has a Maxwell element
sity in their dynamic behavior under similar loading conditions. attached in parallel with a spring element. Prony series is defined as
the mathematical expression representing n Maxwell elements
arranged in series as presented in Fig. 1(b). This arrangement of
Rheology of Soil and Prony Series Maxwell elements in parallel with a spring element represents a
general model for linear viscoelastic stress–strain behavior for non-
Rheology is a science devoted to studying the variation of the stress– homogenous solids. From Fig. 1(b), the relaxation modulus and
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

strain state of real materials over time. Soil generally exhibits viscos- Prony series can be expressed as
ity, which is responsible for the time-dependent stress–strain charac-
teristics (Graham et al. 1983; Mitchell 1993). In the creep test, the X
n
t
=ti
soil experiences strain increment with time at a constant stress (s 0 Þ, ER ðtÞ ¼ E þ Ei  ðeÞ (1a)
and the ratio of the output strain, e ðtÞ, to the input constant stress i¼1

(s 0 Þ is defined as the creep compliance (Cc Þ. On the contrary, the


reduction in stress over time due to the application of a constant X
n
t
=ti
strain (e 0 Þ defines a relaxation test, and the ratio of the output stress, Prony series ¼ Ei  ðeÞ (1b)
s ðtÞ, to the input constant strain (e 0 Þ is defined as the relaxation i¼1

Fig. 1. (a) Basic lumped parameter models for soil; and (b) generalized linear viscoelastic lumped parameter model.

© ASCE 04019072-2 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


In terms of the shear modulus of soil, the shear relaxation modu- Experimental Program
lus (GR ) can be expressed as
This study involves the analysis of four different Indian soils. Soil 1
X
n
t
=ti
GR ðtÞ ¼ G þ Gi  ðeÞ (2) represents Kanpur local soil explored from the geotechnical in situ
i¼1 laboratory site at Indian Institute of Technology (IIT) Kanpur, Uttar
Pradesh, India. Soil 2 represents the marine clay obtained from the
8 eastern coast of the country, Vishakhapatnam, Andhra Pradesh,
GR ðtÞ G1 X n
Gi t
< G1 ¼ lim GR ðtÞ India. Soil 3 represents the black cotton soil explored from Sagar,
=ti
) ¼ þ  ðeÞ where t!1
G0 G0 G : Madhya Pradesh, India. Soil 4 represents the standard kaolin clay.
i¼1 0
G0 ¼ GR ðt ¼ 0Þ The basic index properties of different types of soil have been
obtained from the routine laboratory experiments and are reported
! in Table 1.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

X
n X
n
t
=ti
) gR ðtÞ ¼ 1 gi þ gi  ðeÞ
i¼1 i¼1 Relaxation Test

X
n  t
 Because the relaxation test involves a constant strain application,
) gR ðtÞ ¼ 1  gi  1  ðeÞ =ti (3) strain-controlled, unconsolidated undrained (UU) triaxial tests were
i¼1 performed on the soil specimens of 50 mm in diameter and 100 mm
in height. A fully automatic and computer-controlled low-confining
The shear modulus terms GR ðtÞ and Gi are normalized with pressure triaxial shear test apparatus was used for conducting the
respect to G0 and are represented by gR ðtÞ and gi , respectively, where relaxation experiments. Once the soil specimen is assembled and in-
gR ðtÞ is defined as the normalized relaxation modulus. It can be con- stalled in the triaxial cell, all the input parameters, such as confining
ceived that a general model with n number of Maxwell elements will pressure and loading strain rate, are programmed in the dedicated
have (2n þ 1) unknown parameters (spring elements = n þ 1; damp- computer software interface. However, there is a manual override
ing elements = n). After normalization, gi and ti are the material pa- for controlling the rate of loading in the digital load frame, which is
rameters that need to be determined. Similarly, the normalized creep the key feature in this system to perform a relaxation test. The triax-
compliance (jc Þ can be computed as a function of gi and ti . The pa- ial cell is designed for different pressure ratings ranging from 5 to
rameters gi and ti can be obtained from Eq. (3) by performing a non- 300 kPa. A linear variable differential transformer (LVDT) is used
linear least square fit for the creep/relaxation test data. The relation- to record the variations in deformation.
ship between creep compliance and relaxation modulus can be The rheological properties of clay are the functions of its mois-
established by a convolution integral (Ferry 1980), given in Eq. (4) ture content (Franklin and Krizek 1969), and in the dry state, clay
ðt does not generally show any viscosity. However, an increase in the
ER ðt  t ÞCc ðt Þ  dt ¼ t; for t > 0 (4) amount of moisture in clay alters its consistency and makes it more
0 viscous in nature. Soil specimens considered in this study are all
clayey in nature; hence, the viscosity of soil specimens is expected
Performing the Laplace transform of Eq. (4), the following rela- to vary with the moisture content. However, in this study, optimum
tion can be instituted: moisture content (OMC) is considered as the reference standard for
1 the preparation of the viscous soil specimens. To introduce parity in
ER ðsÞCc ðsÞ ¼ (5) different soil specimens irrespective of their compositions and prop-
s2
erties, all the soil specimens have been prepared at their OMC, main-
where f ðsÞ = Laplace transform of f ðtÞ; and s = transform parame- taining a density at their respective maximum dry density (MDD)
ter. There are various approximate methods for the interconversion values based on standard Proctor compaction test of soil (IS 1980).
of the relaxation modulus and the creep compliance in the time do- After the specimen is installed in the triaxial setup, the desired
main (Park and Kim 1999; Park and Schapery 1999; Schapery and confining (cell) pressure is programmed and set. It is worth men-
Park 1999). However, for materials exhibiting very weak visco- tioning here that the soil specimen has not been saturated prior to
elastic behavior, ER ðtÞ  Cc ðtÞ ffi 1 (quasi-static interrelationship) the shearing. The objective is to shear the specimen to a constant
provides a good approximation (Park and Kim 1999). strain (e 0 ) and record the relaxation of stress, s ðtÞ, with time. To

Table 1. Index properties of soils

Index property Soil 1 Soil 2 Soil 3 Soil 4 References


Liquid limit (percentage) 28.88 76.32 57.4 42.42 IS (1985b)
Plastic limit (percentage) 18.51 33.1 30 23.05 IS (1985b)
Plasticity index 10.37 43.22 27.4 19.37 IS (1985b)
Specific gravity 2.62 2.63 2.32 2.6 IS (1963)
Sand fraction (percentage) 15 9 12 — IS (1985a)
Silt fraction (percentage) 71 29 37 46 IS (1985a)
Clay fraction (percentage) 13 62 51 54 IS (1985a)
OMC (percentage) 14 29 26 24 IS: 2720-(VII)-1980
MDD (kg/m3) 1,800 1,440 1,560 1,550 IS: 2720-(VII)-1980
Classification CL CH OH CL USCSa
a
ASTM (2006).

© ASCE 04019072-3 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


achieve the desired e 0 , the soil specimen needs to be sheared instan-
taneously. This instantaneous shearing can be achieved by linearly
ramping the strain to e 0 at some strain rate. If the strain rate is rela-
tively small, some fraction of the stress developed in the soil with
time does relax concurrently with time before the required strain
value is reached (Ömer Bilgin 2014). Hence, it may not be possible
to observe and record the true relaxation of stress in the soil. On the
contrary, with very high strain rate, the specimen may exhibit frac-
ture. The trials were performed with different strain rates, and the
strain rate at which the relative error in the relaxation curve reaches
a tolerance of 5% was chosen as the optimum strain rate to perform
the relaxation test. Following this procedure, the strain rate was cho-
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

sen as 0.8 mm/min for Soil 1, whereas the same is 1.25 mm/min for
Soil 2 to Soil 4. The strain rate for loading was programmed in the
computer interface, and the shearing of the specimen was initiated.
The LVDT measured the progress of strain, and the load cell pro-
vided the reaction load (in kilograms) experienced by the specimen
in real time. Once the desired strain (e 0 ) was reached, the digital
load frame was manually switched off, keeping the load frame
locked at the same strain. Because the load cell, LVDT, and the con-
fining pressure system were independent of the load frame system,
they continued to work as per their programmed nature and the out-
put was recorded by the computer system. Figs. 2(a and b) present
the variation of strain and stress with time, respectively, for Soil 1.
The data were recorded at 1 s time intervals, and the stress relaxa-
tion was recorded for a total time of 800 s in all the tests. With dif-
ferent trials for Soil 1, it was observed from Fig. 2(c) that the stress
varied almost linearly with strain up to 23 kPa, and the modulus of
elasticity was also observed to increase with an increase in the con-
fining pressure (Table 2). Therefore, choosing a single strain value
may result in stress to increase beyond the linear elastic region. The
aim of this study is to capture the linear viscoelastic nature of the
soil. Therefore, the strain (e 0 ) corresponding to the maximum stress
was recorded for all types of soil. The magnitudes of e 0 selected for
conducting the relaxation tests at different confining pressures are
given in Table 2.
The stress data, s ðtÞ, obtained from the relaxation test were con-
verted to the normalized relaxation modulus, gR ðtÞ; as explained in
Eq. (3), and the variations of gR ðtÞ with time for different soils are
presented in Fig. 3. Fig. 3 indicates that the stress relaxation of soil is
not much affected by the variation of confining pressure. The maxi-
mum variation from the average relaxation curves (Fig. 4) was found
to be 3.94%, 6.93%, 6.81%, and 6.07% for Soil 1 to Soil 4, respec-
tively. Hence, the relaxation property of soil may be assumed inde-
pendent of the depth. A similar hypothesis for the frozen Ottawa
sand was reported by Ladanyi and Benyamina (1995), suggesting
that the confining pressure (100–300 kPa) had no significant influ-
ence on the stress relaxation behavior. Therefore, for the analysis
purpose, the average value of gR ðtÞ was obtained for different con-
fining pressures, and the variation of average gR ðtÞ with time is pre-
sented in Fig. 4. These average curves represent the general relaxa-
tion behavior of the respective soils in the linear elastic range. It is
worth noting here that Lade and Karimpour (2015) reported that the
stress relaxation behavior of soil may be different at high confining
pressure. However, the effect of high confining pressure on the stress
relaxation behavior has not been explored in the present study.
Fig. 2. Variation of (a) input strain with time; (b) output stress with
time; and (c) stress with strain for Soil 1 at different confining pressures.
Prony Series Parameter Fitting

The relaxation behavior of different soils can be made more useful number of Maxwell elements, which provide the best fit to the
by fitting and extracting the general behavior in a simple mathemat- relaxation test data. A detailed procedure to find out the optimum
ical model formulation such as the Prony series. The method of fit- number of Maxwell elements required for all the soils is presented
ting the Prony series involves the determination of a minimum in this study. For Soil 1, the fitting curves for three-parameter (one

© ASCE 04019072-4 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


Table 2. Magnitude of strain and elastic modulus at different confining pressures

Parameter Confining pressure (kPa) Soil 1 Soil 2 Soil 3 Soil 4


3
Strain [e 0 (10 )] 20 2.2 8.0 10.0 5.5
40 1.7 7.3 9.0 4.5
80 1.4 6.3 7.5 4.1
Modulus of elasticity [E (MPa)] 20 10.19 6.24 5.09 9.08
40 13.14 6.84 5.55 11.10
80 16.11 7.93 6.66 12.18

Maxwell element), five-parameter (two Maxwell elements), and control boundary condition (Uy = e 0  100) at a time interval of
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

seven-parameter (three Maxwell elements) models are presented in 1 s (applied with stable time increments). A new analysis step was
Fig. 5, and the fitting parameters for all the cases along with the defined for a total time of 800 s to calculate and store the variation
coefficient of determination (R2) are tabulated in Table 3. The pa- of relaxation stress over time. In this study, the linear viscoelasticity
rameters reported in Table 3 are obtained by performing a nonlinear material model available in the material module of ABAQUS is
curve fitting (nonlinear least-squares fitting) on the experimentally considered to define the small strain linear viscoelastic property in
obtained relaxation data gR ðtÞ with the mathematical expression for the system.
the Prony series. A statistical analysis was performed to obtain the
best fit curve for the respective soil data. In Fig. 5, the variation of Relaxation Test
residual error (difference between the predicted and the actual Soil 1 is considered to validate the numerical modeling with the ex-
value) with time is also presented in the inset. The variation of the perimental observations. The seven-parameter Prony series parame-
total error was obtained and used to calculate the coefficient of ters given in Table 3 are considered the direct input to the time do-
determination (R2). It is worth noting from Table 3 that based on the main viscoelastic material model available in ABAQUS. The
R2 values, the three-parameter model provides an unsatisfactory fit- Poisson’s ratio of soil is assumed to be 0.33. The values of E0 and
ting of the data irrespective of the type of soil. On the contrary, the e 0 corresponding to different confining pressures were selected
five-parameter model provides a good agreement and nicely cap- from Table 2. The model was discretized with 100 numbers of
tures the behavior for all the soils with R2 value varying from 0.970 eight-noded biquadratic axisymmetric quadrilateral (CAX8R) ele-
to 0.989. However, looking at the distribution of the residual errors ments. The stress output obtained from the FE analysis was plotted
presented in Fig. 5, the deviation of the actual values from the pre- along with the experimental data in Fig. 7(a). It is worth mentioning
dicted values is very high as compared to that observed in the that only one simulation output curve represents the stress relaxa-
seven-parameter model. The residuals in the seven-parameter fitting tion of Soil 1 for three different confining pressures because all
are normally distributed about the predicted values with a magni- three curves exactly overlap onto a single curve showing no effect
tude as high as 0.0075, which may be considered insignificant as of the confining pressure on the output stress. Similar overlap of the
compared to the other two models. Therefore, for all the soils, the output stress at different confining pressures can be observed in the
seven-parameter model is considered to result in fitting the relaxa- experimental results as well. The FE results were found to establish
tion data reasonably (R2 > 0.997). a good agreement with the experimental results; hence, the seven-
parameter model is considered to work satisfactorily for the stress
relaxation problem.
Numerical Modeling
Creep Test
The validation of the stress–strain material model defined in the pre- Because the creep test involves a constant load application, a fixed-
vious section was performed by modeling and solving some physi- ring-type oedometer test was performed on the collected soil
cal problems through the FE method. Two different types of prob- specimen measuring 6 cm in diameter and 2 cm in height. A load in-
lems were considered: one was solved in the time domain (static), tensity (s 0) of 10 kPa was applied on the soil specimen, and the cor-
and the other was solved in the frequency domain (steady-state responding strain at different time intervals was recorded. The tan-
dynamics). gent modulus obtained from the test was found to be approximately
0.9 MPa. Because the oedometer test is confined radially, the modu-
Static Problem lus of elasticity is therefore considered to be the constrained modu-
lus (Eyy: modulus of elasticity in vertical direction), so the value of
As mentioned previously, the seven-parameter model describes the isotropic modulus of elasticity (E) turns out to be approximately
stress–strain constitutive behavior of the four different soils reason- 0.6 MPa, assuming the Poisson’s ratio of the soil as 0.33. The same
ably well. In order to verify the model parameters, a 2D axisymmet- axisymmetric triaxial model previously discussed is considered to
ric FE model was developed in ABAQUS 6.14 for the relaxation simulate the creep test as well. The boundary conditions remain the
test similar to the experimental test setup [Fig. 6(a)]. The modeling same as that of the previous model. However, the free sides were
space chosen for the soil was a 2D planar deformable shell. Because kept free without the application of confining stress, and a constant
the soil in the triaxial test setup only moves in the downward and stress of 10 kPa was applied to the top surface instead of the strain
the radial directions, the base of the soil in the FE model was kept rate. The stress was applied in the first step, and then the creep
fixed in the vertical (Uy = 0) as well as in the horizontal (Ux = 0) response of the soil was computed in the second step of the analysis.
directions, whereas the left vertical side (axis of symmetry) was The creep output obtained from the FE analysis is presented in
provided with the roller support (Ux = 0). The top and the right-side Fig. 7(b) along with the experimental data extracted from the oed-
surfaces of the model were free and provided with the desired con- ometer test. The creep output determined from the FE analysis also
fining pressure. Because the soil needs to be sheared with a constant shows a good concurrence with the data obtained experimentally,
strain rate, the top surface was set to the required displacement ensuring the calibration of the seven-parameter model with the

© ASCE 04019072-5 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Variation of average gR with time.

rheology of the soil. It is evident from the preceding examples that


the seven-parameter linear viscoelastic model is capable of effec-
tively capturing the rheology of soil for the static load response.

Dynamic Problem
The same seven-parameter constitutive stress–strain relation
adopted for Soil 1 was implemented in a 3D FE model to capture
the response of the soil subjected to dynamic loading. Bhoumik
(1989) and Swain and Ghosh (2016) reported the dynamic
response of machine foundations performing block vibration tests
in Soil 1. The dynamic response of such a machine foundation
(0:65  0:65  0:2 m) reported by Swain and Ghosh (2016) was
modeled with finite elements considering a hemispherical soil do-
main, as presented in Fig. 6(b). Sensitivity analysis was per-
formed based on the response of the footing with different diame-
ters of the hemispherical soil domain and resulted in an optimum
domain size of 5.2 m in diameter. The footing was placed exactly
at the center of the top circular surface of the soil domain, as pre-
sented in Fig. 6(b). The material properties considered for the
dynamic analysis are given in Table 4. The footing and the soil
are assumed to be rigid and viscoelastic in nature, respectively,
where the viscoelastic properties of soil are taken from Table 3.
The value of E0 was chosen from Swain and Ghosh (2016). The
contact between the soil and the footing was completely con-
strained using a tie constraint, which ensured no separation and
sliding of the footing relative to the soil surface. The soil domain
was discretized in to finite (inner) and infinite (outer) zones, as
presented in Fig. 6(b). The footing and the finite zone were discre-
tized with eight-noded linear brick elements of smaller size with
reduced integration and hourglass control (C3D8R), whereas the
infinite zone was discretized with 3D eight-noded linear contin-
uum one-way infinite elements (CIN3D8) (Ni et al. 2017a, b). It
is worth mentioning that the infinite elements act as an absorbing
boundary condition to represent the semi-infinite soil domain.
In the block vibration test, the basic assumption is that the foun-
dation resting on the soil is relatively rigid and thus anticipating
rigid body motion. Therefore, a rigid body constraint is defined and
applied to the footing in the FE analysis, which allows the constitu-
Fig. 3. Variation of gR ðtÞ with time at different confining pressures for ent elements not to deform but undergo large rigid body motions.
various soils: (a) Soil 1; (b) Soil 2; (c) Soil 3; and (d) Soil 4. The maximum grid spacing (Dh) can be expressed as (Kuhlemeyer
and Lysmer 1973)

© ASCE 04019072-6 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


study, the maximum value of Dh is calculated to be 0.125 m;
hence, the mesh size of the soil domain is varied from 0.05 to
0.12 m, generating 8,960 elements. As proposed by Swain and
Ghosh (2016), the footing was excited with a harmonic force as
given in Eq. (7).
 
F ¼ F0  sin v t ¼ me  e  v 2  sin v t
 
u
¼ 0:11904 v 2  sin  sin v t (7)
2

where me = total unbalanced rotating mass in the excitation system;


Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

e = radius of rotation of the eccentric mass; and u = eccentric angle


setting between two unbalanced rotating masses, which is equal to
24 as proposed by Swain and Ghosh (2016). To evaluate the
dynamic response of the footing-soil system in the frequency do-
main, the steady-state dynamic procedure was chosen in the analy-
sis, which performed a frequency sweep in the desired frequency
range to obtain the steady-state amplitude response of the system at
every frequency increment due to the harmonic excitation. A fre-
quency range of 0 to 44 Hz was chosen with an increment sweep of
1 Hz to capture the response of the foundation-soil system. Fig. 8(a)
presents the variation of vertical displacement amplitude response
with frequency as obtained from the present FE analysis. The reso-
nant frequency (fmr) as obtained from the response curve was found
to be 34 Hz, and the natural frequency (fn) was estimated as
32.8 Hz. In Fig. 8(b) and Table 5, the steady-state dynamic response
extracted from the present analysis was compared to that reported
by Bhoumik (1989) and Swain and Ghosh (2016) at the same inves-
tigation site. It is worth noting that Bhoumik (1989) conducted
block vibration tests at the same location (Soil 1) and presented the
response curves for a block foundation (1  0:75  0:7 m) vibrat-
ing on the soil surface. In Fig. 8(b), the variation of magnification
factor (M) with frequency ratio (fr) is presented as proposed by
Richart et al. (1970), where the frequency ratio and the magnifica-
tion factor can be expressed as

f
fr ¼ (8)
fn

A
M¼   (9)
me  e
m

where A = vertical displacement amplitude of vibration; f = oper-


ating frequency; and fn = natural frequency of the system. It is evi-
dent from the comparison that the present FE analysis reveals a
reasonably close match with the available experimental results. In
Fig. 8(b), the dynamic response obtained from the present analy-
sis matches reasonably well with that of Swain and Ghosh (2016)
but deviates from the results of Bhoumik (1989) at lower frequen-
cies. This may be attributed to the Prony series parameters
Fig. 5. Prony series fitting for Soil 1: (a) three-parameter fitting; extracted from a static relaxation test rather than a dynamic vibra-
(b) five-parameter fitting; and (c) seven-parameter fitting. tional test. However, the region of resonance is well captured
with acceptable predictions of the resonating frequency of the
system and M, which are vital in this class of problem. From
v Table 5, the difference between the present results and that of
Dh  (6)
8fmax Swain and Ghosh (2016) was found to be 6.4% for vertical dis-
placement at resonance, 6.3% for fmr, 7.3% for fn, and 13.6% for
where v = lowest wave velocity, that is, the shear wave velocity damping. This difference may be attributed to various factors
in this case; and fmax = maximum operating frequency, and for such as heterogeneity and nonlinearity of soil deposit and uncer-
most of the industrial machines, it can be taken as 50 Hz. In this tainties involved in the experiments.

© ASCE 04019072-7 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


Table 3. Prony series fitting parameters

Soil Parameters g1 g1 t1 (s) g2 t2 (s) g3 t3 (s) R2


1 3 0.573 0.427 77.183 — — — — 0.670
5 0.573 0.229 9.108 0.198 207.930 — — 0.989
7 0.566 0.138 2.910 0.128 26.385 0.168 264.630 0.998
2 3 0.250 0.750 209.029 — — — — 0.477
5 0.262 0.327 19.299 0.411 334.469 — — 0.986
7 0.222 0.210 5.867 0.209 61.044 0.359 551.419 0.999
3 3 0.573 0.427 164.146 — — — — 0.349
5 0.580 0.202 16.538 0.218 368.062 — — 0.970
7 0.577 0.119 4.479 0.126 45.360 0.178 442.021 0.999
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

4 3 0.710 0.290 112.671 — — — — 0.557


5 0.715 0.146 13.606 0.139 259.627 — — 0.983
7 0.704 0.093 4.613 0.091 45.722 0.112 393.984 0.997

Fig. 7. Comparison of (a) stress relaxation output; and (b) creep output
Fig. 6. FE modeling for (a) triaxial test; and (b) dynamic response of obtained from FE analysis with experimental data for Soil 1.
footing.

physical parameters, such as domain size, footing size, and load


With the confidence gained in the present numerical analysis, type, were kept the same as that considered by Swain and Ghosh
the same FE model was used to predict the steady-state dynamic (2016). The dynamic response of footing (0:65  0:65  0:2 m)
response for all other soils considered in this study. The other resting on different types of soil was predicted from the FE analysis.

© ASCE 04019072-8 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


Table 4. Input parameters for 3D FE analysis Table 5. Output comparison with experimental results

Material property Soil 1 Footing Current Swain and Bhoumik


3 Output parameter study Ghosh (2016) (1989)
Mass density (kg/m ) 1,832 2,500
Elastic modulus [E0 (MPa)] 12 2,500 Resonant displacement (m) 161  106 172  106 127  106
Poisson’s ratio 0.33 0.20 Resonant frequency [fmr (Hz)] 34.00 36.28 40.98
Natural frequency [fn (Hz)] 32.80 35.40 39.13
Damping (percentage) 17.60 15.50 14.50
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Steady-state dynamic response obtained from FE analysis for


different soils.

Table 6. Predicted dynamic response parameters for different soils

Output parameter Soil 1 Soil 2 Soil 3 Soil 4


Resonant 161  106 220  106 180  106 181  106
displacement (m)
Resonant frequency 34.0 29.5 23.0 30.5
[fmr (Hz)]
Natural frequency 32.8 29.0 22.4 29.7
[fn (Hz)]
Damping 17.6 13.0 16.0 16.1
(percentage)

Fig. 8. (a) Dynamic response of foundation-soil system obtained from


the real-time behavior of materials. However, it is quite challenging
FE analysis; and (b) comparison of isolated footing response with ex- to extract the rheological properties of soil-like material correctly
perimental results in Soil 1. with precision. In this paper, the rheology of four different types of
soil (local soil, marine clay, black cotton soil, and kaolin clay) was
explored. A generalized lumped parameter model representing lin-
The variation of the vertical displacement amplitude of the footing ear viscoelastic stress–strain behavior is presented. The data
with frequency for different soil deposits is presented in Fig. 9. obtained from the relaxation test were converted to normalized
Table 6 presents the magnitude of various output parameters relaxation modulus ðgR Þ and used to estimate the unknown Prony
obtained for different soil deposits. In Table 6, Soil 2 experiences series parameters gi and ti by performing a nonlinear curve fitting.
the maximum vertical displacement, which implies that Soil 2 pos- An elaborate fitting was performed with three-, five-, and seven-
sesses the least damping as compared to other three soils. parameter models to find the best fit, where the seven-parameter
model was found to be satisfactory in fitting the relaxation data of
all soils. To check the accuracy of the lumped parameter model, 2D
Conclusions and 3D FE models were developed. The 2D-axisymmetric FE
viscoelastic model developed for predicting the stress relaxation
Rheology is an important and critical aspect of any material. captured the stress versus time response quite efficiently. The same
Understanding rheology opens up a new dimension (time) to predict model correctly predicted the variation of strain with time for the

© ASCE 04019072-9 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


creep test performed at s 0 ¼ 10 kPa. The stress relaxation was Ladanyi, B., and M. B. Benyamina. 1995. “Triaxial relaxation testing of a
observed to be invariant of the confining pressure up to a range of frozen sand.” Can. Geotech. J. 32 (3): 496–511. https://doi.org/10.1139
80 kPa. The same Prony series parameters were used in a 3D FE /t95-052.
Lade, P. V., and H. Karimpour. 2015. “Stress relaxation behavior in
model to capture the displacement response of a vibrating square
Virginia Beach sand.” Can. Geotech. J. 52 (7): 813–835. https://doi.org
footing under a rotating mass-type excitation. The dynamic /10.1139/cgj-2013-0463.
response of the vibrating footing compares reasonably well with the Lyakhov, G. M. 1968. “Viscous properties of soil.” J. Appl. Mech. Tech.
experimental results reported in the literature. A comparison of the Phys. 9 (4): 412–416. https://doi.org/10.1007/BF00912740.
dynamic response for four different types of soil is presented to Manna, S., S. Kundu, and J. C. Misra. 2018. “Theoretical analysis of tor-
understand the dynamic behavior of different soils under similar sional wave propagation in a heterogeneous aeolotropic stratum over a
physical conditions. Voigt-type viscoelastic half-space.” Int. J. Geomech. 18 (6): 04018050.
https://doi.org/10.1061/(ASCE)GM.1943-5622.0001144.
Mishra, A., and N. R. Patra. 2018. “Time-dependent settlement of pile
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

References foundations using five-parameter viscoelastic soil models.” Int. J.


Geomech. 18 (5): 04018020. https://doi.org/10.1061/(ASCE)GM
ASTM (American Society of Testing and Materials). 2006. Standard prac- .1943-5622.0001122.
tice for classification of soils for engineering purposes (Unified Soil Mitchell, J. K. 1993. Fundamentals of soil behavior. 2nd ed. New York:
Classification System). D2487–06. ASTM. John Wiley & Sons.
Azari, B., B. Fatahi, and H. Khabbaz. 2016. “Assessment of the elastic- Nash, D., and M. Brown. 2015. “Influence of destructuration of soft clay on
viscoplastic behavior of soft soils improved with vertical drains captur- time-dependent settlements: Comparison of some elastic viscoplastic
ing reduced shear strength of a disturbed zone.” Int. J. Geomech. 16 (1): models.” Int. J. Geomech. 15 (5): A4014004. https://doi.org/10.1061
B4014001. https://doi.org/10.1061/(ASCE)GM.1943-5622.0000448. /(ASCE)GM.1943-5622.0000281.
Bhoumik, K. 1989. “Coupled vibration of footings on elastic half-space.” Ni, P., S. Mangalathu, G. Mei, and Y. Zhao. 2017a. “Permeable piles: An al-
M.Tech. thesis, Dept. of Civil Engineering, Indian Institute of Technology. ternative to improve the performance of driven piles.” Comput. Geotech.
Dey, A., and P. K. Basudhar. 2010. “Applicability of Burger model in pre- 84 (Apr): 78–87. https://doi.org/10.1016/j.compgeo.2016.11.021.
dicting the response of viscoelastic soil beds.” In GeoFlorida 2010: Ni, P., G. Mei, and Y. Zhao. 2017b. “Numerical investigation of the uplift
Advances in Analysis, Modeling & Design, Geotechnical Special performance of prestressed fiber-reinforced polymer floating piles.”
Marine Georesour. Geotechnol. 35 (6): 829–839. https://doi.org/10
Publication 199, 2611–2620. Reston, VA: ASCE.
.1080/1064119X.2016.1255688.
Feng, W., J. Yin, X. Tao, F. Tong, and W. Chen. 2017. “Time and strain-
Ömer Bilgin, P. E. 2014. “Modeling viscoelastic behavior of polyethylene
rate effects on viscous stress–strain behavior of plasticine material.” Int.
pipe stresses.” J. Mater. Civ. Eng. 26 (4): 676–683. https://doi.org/10
J. Geomech. 17 (5): 04016115. https://doi.org/10.1061/(ASCE)GM
.1061/(ASCE)MT.1943-5533.0000863.
.1943-5622.0000806.
Park, S. W., and Y. R. Kim. 1999. “Interconversion between relaxation
Ferry, J. D. 1980. Viscoelastic properties of polymers. 3rd ed. New York:
modulus and creep compliance for viscoelastic solids.” J. Mater.
Wiley.
Civ. Eng. 11 (1): 70–82. https://doi.org/10.1061/(ASCE)0899
Franklin, A. G., and R. J. Krizek. 1969. “Complex viscosity of a kaolin
-1561(1999)11:1(76).
clay.” Clays Clay Miner. 17 (2): 101–110. https://doi.org/10.1346
Park, S. W., and R. A. Schapery. 1999. “Methods of interconversion
/CCMN.1969.0170208.
between linear viscoelastic material functions. Part I—A numerical
Graham, J., J. H. A. Crooks, and A. L. Bell. 1983. “Time effects on the
method based on Prony series.” Int. J. Solids Struct. 36 (11): 1653–
stress-strain behaviour of natural soft clays.” Geotechnique 33 (3):
1675. https://doi.org/10.1016/S0020-7683(98)00055-9.
327–340. https://doi.org/10.1680/geot.1983.33.3.327. Richart, F. E., J. R. Hall, and R. D. Woods. 1970. Vibrations of soils and
IS (Indian Standards Institution). 1963. Method of test for aggregates for foundations. Englewood Cliffs, NJ: Prentice-Hall.
concrete. IS 2386 Part III (Reaffirmed 2002). Manak Bhavan, New Schapery, R. A., and S. W. Park. 1999. “Methods of interconversion
Delhi, India: IS. between linear viscoelastic material functions. Part II—An approximate
IS (Indian Standards Institution). 1980. Method of test for soils. 2nd analytical method.” Int. J. Solids Struct. 36 (11): 1677–1699. https://doi
Revision. IS 2720 Part VII (Reaffirmed 2011). Manak Bhavan, New .org/10.1016/S0020-7683(98)00060-2.
Delhi, India: IS. Schiffman, R. L., C. C. Ladd, and A. T. F. Chen, 1966. “The secondary con-
IS (Indian Standards Institution). 1985a. Method of test for soils. 2nd solidation of clay.” In Rheology and Soil Mechanics International
Revision. IS 2720 Part IV (Reaffirmed 2006). Manak Bhavan, New Union of Theoretical and Applied Mechanics, edited by J. Kravtchenko
Delhi, India: IS. and P. M. Sirieys, 273–303. Berlin: Springer-Verlag.
IS (Indian Standards Institution). 1985b. Method of test for soils. 2nd Sun, H. Z., and W. Zhang. 2007. “Analysis of soft soil with viscoelastic
Revision. IS 2720 Part V (Reaffirmed 2006). Manak Bhavan, New fractional derivative Kelvin model.” Rock Soil Mech. 28 (9): 1983–
Delhi, India: IS. 1986.
Kakar, R. 2016. “Love waves in Voigt-type viscoelastic inhomogeneous Swain, A., and P. Ghosh. 2016. “Experimental study on dynamic interfer-
layer overlying a gravitational half-space.” Int. J. Geomech. 16 (3): ence effect of two closely spaced machine foundations.” Can. Geotech.
04015068. https://doi.org/10.1061/(ASCE)GM.1943-5622.0000582. J. 53 (2): 196–209. https://doi.org/10.1139/cgj-2014-0462.
Kakar, R., and S. Kakar. 2017. “Love wave in a Voigt-type viscoelastic het- Taherkhani, H., and M. Jalali. 2018. “Viscoelastic analysis of geogrid-
erogeneous layer overlying heterogeneous viscoelastic half-space.” Int. reinforced asphaltic pavement under different tire configurations.”
J. Geomech. 17 (1): 06016009. https://doi.org/10.1061/(ASCE)GM Int. J. Geomech. 18 (7): 04018060. https://doi.org/10.1061/(ASCE)GM
.1943-5622.0000675. .1943-5622.0001183.
Kerr, A. D. 1961. “Viscoelastic Winkler foundation with shear interactions.” Tan, T. K. 1957. “Three-dimensional theory on the consolidation and flow
J. Eng. Mech. Div. 87 (3): 13–30. https://cedb.asce.org/CEDBsearch of the clay-layers.” Sci. Sin. 6 (1): 203–215. https://doi.org/10.1360
/record.jsp?dockey=0012494. /ya1957-6-1-203.
Kimoto, S., B. S. Khan, M. Mirjalili, and F. Oka. 2015. “Cyclic elasto- Yin, J. 2015. “Fundamental issues of elastic viscoplastic modeling of
viscoplastic constitutive model for clay considering nonlinear kine- the time-dependent stress–strain behavior of geomaterials.” Int. J.
matic hardening rules and structural degradation.” Int. J. Geomech. Geomech. 15 (5): A4015002. https://doi.org/10.1061/(ASCE)GM
15 (5): A4014005. https://doi.org/10.1061/(ASCE)GM.1943-5622 .1943-5622.0000485.
.0000327. Yin, J., and J. Graham. 1989. “Viscous-elastic-plastic modelling of one-
Kuhlemeyer, R. L., and J. Lysmer. 1973. “Finite element method accuracy dimensional time-dependent behaviour of clays.” Can. Geotech. J. 26
for wave propagation problems.” J. Soil Dyn. Div. 99 (sm5): 421–427. (2): 199–209. https://doi.org/10.1139/t89-029.

© ASCE 04019072-10 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072


Yin, Z., Q. Xu, and C. Yu. 2015. “Elastic-viscoplastic modeling for Zheng, C., X. Ding, and Y. Sun. 2016. “Vertical vibration of a pipe pile in
natural soft clays considering nonlinear creep.” Int. J. Geomech. viscoelastic soil considering the three-dimensional wave effect of soil.”
15 (5): A6014001. https://doi.org/10.1061/(ASCE)GM.1943-5622 Int. J. Geomech. 16 (1): 04015037. https://doi.org/10.1061/(ASCE)GM
.0000284. .1943-5622.0000529.
Zhang, S., G. Ye, and J. Wang. 2018. “Elastoplastic model for overconsoli- Zhu, H. H., L. C. Liu, H. F. Pei, and B. Shi. 2012. “Settlement analysis of
dated clays with focus on volume change under general loading condi- viscoelastic foundation under vertical line load using a fractional
tions.” Int. J. Geomech. 18 (3): 04018005. https://doi.org/10.1061 Kelvin-Voigt model.” Geomech. Eng. 4 (1): 67–78. https://doi.org/10
/(ASCE)GM.1943-5622.0001101. .12989/gae.2012.4.1.067.
Downloaded from ascelibrary.org by Universidad Nacional De Ingenieria on 08/27/19. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04019072-11 Int. J. Geomech.

Int. J. Geomech., 2019, 19(7): 04019072

You might also like