You are on page 1of 12

Geotextiles and Geomembranes xxx (xxxx) xxx

Contents lists available at ScienceDirect

Geotextiles and Geomembranes


journal homepage: www.elsevier.com/locate/geotexmem

Centrifuge modeling of geosynthetic-encased stone column-supported


embankment over soft clay
Jian-Feng Chen a, Liang-Yong Li b, *, Zhen Zhang a, **, Xu Zhang a, Chao Xu a,
Sathiyamoorthy Rajesh c, Shou-Zhong Feng d
a
College of Civil Engineering, Tongji University, Shanghai, 200092, China
b
College of Civil Engineering and Architecture, Hainan University, Haikou, 570228, China
c
Department of Civil Engineering, Indian Institute of Technology Kanpur, Kanpur, 208016, India
d
Wuhan Guangyi Transportation Science and Technology Co., Ltd, Wuhan, 430074, China

A R T I C L E I N F O A B S T R A C T

Keywords: Geosynthetic-encased stone column (GESC) has been proven as an effective alternative to reinforcing soft soils. In
Geosynthetics this paper, a series of centrifuge model tests were conducted to investigate the performance of GESC-supported
Encased stone column embankment over soft clay by varying the stiffness of encasement material. The enhancement in the performance
Embankment
of stone columns encased with geosynthetic materials was quantified by comparing the test with ordinary stone
Soft soil
Centrifuge
columns (OSCs) under identical test conditions. The test results reveal that by encasing stone columns with
Load transfer geosynthetic material, a significant reduction in the ground settlement, relatively faster dissipation of excess pore
pressure and enhanced stress concentration ratio was noticed. Moreover, with the increase in the encasement
stiffness from 450 kN/m to 3300 kN/m, the stress concentration ratio increased from 4 to 6.5, which signifies the
importance of encasement stiffness. In addition, a relatively lower value of soil arching ratio observed for GESCs
compared to OSCs indicate the formation of a relatively strong soil arch in the GESC-supported embankment.
Interestingly, under embankment loading, GESCs fail by bending while OSCs fail by bulging. The stress reduction
method can be used to calculate the settlement of GESC-supported embankment with larger stress reduction
factor than that in the OSC-supported embankment. Finally, the limitation of the construction of the embank­
ment at 1 g was addressed.

1. Introduction 2005; Murugesan and Rajagopal, 2009; Castro and Sagaseta, 2011; Ali
et al., 2012; Hong, 2012; Yoo and Lee, 2012; Almeida et al., 2013, 2014;
Stone columns are considered as one of the cost-effective and Dash and Bora, 2013; Elsawy, 2013; Wu and Hong, 2014; Hosseinpour
environmentally-friendly methods to improve soft soils. The load- et al., 2015; Yoo, 2015; Gu et al., 2016; Rajesh and Jain, 2015; Hong
carrying capacity of stone columns primarily depends on the lateral et al., 2016, 2017; Miranda and Da Costa, 2016; Mohapatra and Raja­
confinement of surrounding soil. When the stone column is installed in gopal, 2017; Rajesh, 2017; Schnaid et al., 2017; Cengiz and Güler, 2018;
soils having low undrained shear strength (e.g., cu < 15 kPa), it could Kadhim et al., 2015; Alkhorshid et al., 2019; Cengiz et al., 2019; Yoo and
deform laterally under vertical stress, and some granular materials may Abbas, 2020).
squeeze into the surrounding soil (Weber et al., 2006, 2010), resulting in There have been a number of studies focusing on single and group
bulging failure at the upper section of the column (Hughes et al., 1975). GESCs under rigid loading. Gu et al. (2017), Chen et al. (2018) and Xue
To solve or mitigate such a problem, geosynthetic-encased stone column et al. (2019) conducted unit cell tests to investigate the behavior of
(GESC) is proposed. The encasement provides additional confinement to GESCs under rigid loading. Murugesan and Rajagopal (2009), Ali et al.
the stone column, resulting in increasing bearing capacity, reducing (2014) and Chen et al. (2018) investigated the bearing capacity of single
settlement and mitigating bulging failure of stone column (Raithel et al., and group of GESCs under rigid loading based on small-scale (1g) model

* Corresponding author.
** Corresponding author.
E-mail addresses: jf_chen@tongji.edu.cn (J.-F. Chen), liliangyong200@163.com (L.-Y. Li), dyzhangzhen@gmail.com (Z. Zhang), 1710016@tongji.edu.cn
(X. Zhang), c_axu@tongji.edu.cn (C. Xu), hsrajesh@iitk.ac.in (S. Rajesh), fsz63@vip.163.com (S.-Z. Feng).

https://doi.org/10.1016/j.geotexmem.2020.10.021

0266-1144/© 2020 Elsevier Ltd. All rights reserved.

Please cite this article as: Jian-Feng Chen, Geotextiles and Geomembranes, https://doi.org/10.1016/j.geotexmem.2020.10.021
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

test. Gu et al. (2017) indicated that the interlocking between aggregate 1.85 and 1.60 g/cm3, respectively. The relative density was controlled to
and geogrid influenced the behavior of the geogrid-encased stone col­ 70% with a density of 1.77 g/cm3. The peak friction angle of the silica
umn under rigid loading. For a group of GESCs under rigid footing sand at 70% relative density was 38◦ based on the direct shear test.
loading, the columns near the edges of footing carried higher stress, For replicating the prototype response in a small-scale model, suit­
while the columns near the center area were better laterally confined able scaling relationships, which link the model behavior to that of the
(Castro, 2017). GESCs subjected to rigid loading can be modeled using prototype, need to be adopted (Schofield, 1980). To access the actual
the concept of the cylindrical unit cell and equal strain assumption performance of the GESC supported embankment, the prototype stresses
(Raithel and Kempfert, 2000; Pulko et al., 2011; Castro and Sagaseta, were generated by performing centrifuge tests at 25 gravities (i.e., N =
2013; Pulko and Logar, 2017). 25). The dimensions of the GESC-supported embankment have been
As expected, equal strain condition exists between columns and soil worked out by considering the scale factor of 25. As the tensile strength
under rigid loading but not for embankment loading. The embankment and the stiffness of the geosynthetic material dictates the performance of
loading is regarded as a loading between rigid and flexible ones, in GESC, model geosynthetic material needs to be selected by considering
which, columns and soil may deform under a condition between the suitable scaling relations. According to the scaling laws of geosynthetic,
equal strain and equal stress. Han (2015) and Zhang et al. (2019) the tensile strength and the stiffness of model geogrid should be 1/N
pointed out that load stiffness influences the load transfer behavior times as that of prototype geogrid (Madabhushi, 2014). Two
between columns and soil. The column behavior is also influenced by the scaled-down geogrids, namely polyamide grid (labeled as G1) and nylon
column position under loading (Chen et al. 2015, 2018; Castro, 2017). grid (labeled as G2) were used to model the geosynthetic encasement.
The applied stress on the foundation by embankment fill is not uniform. Based on the wide-width tensile test, G1 and G2 had the ultimate tensile
As a result, GESCs at different locations under the embankment base strengths of 13.6 and 2.5 kN/m, and stiffness of 132 and 18 kN/m at 5%
could behave differently (Mohapatra et al., 2017). Fattah et al. (2016) strain, respectively. The column had a diameter of 32 mm, corre­
examined the influence of embankment height, column spacing, and the sponding to 800 mm in the prototype. In the prototype, the ratio of
ratio of column length to column diameter on the bearing capacity of encasement stiffness to column diameter is similar to that common in
GESC-reinforced soil. Kadhim et al. (2015) and Mohapatra et al. (2017) the field, which corresponds to encasement stiffness of 3300 and 450
investigated the stability of GESC-supported embankment over soft clay kN/m. Fig. 1 presents the prepared model encasements for the GESCs.
through numerical modeling. The geosynthetics were wrapped to form a hollow tube with an overlap
Chen et al. (2015) performed a series of small-scale (1g) model test of 10 mm. The overlap area was sewed using nylon threads and glued
and found that GESCs under embankment slope experienced a large with the geosynthetics by special glue. The wide-width tensile test was
bending deformation instead of shear failure. As the scaled-down 1g also conducted to determine the ultimate seam strength with a specimen
model test has a limitation of low-stress level when compared with the having a horizontal seam at mid-length. The seams of geosynthetic G1
full-scale prototypes, and the conclusions were drawn by considering and G2 had ultimate strengths of 12.2 and 2.8 kN/m, and stiffness of 120
one type of geosynthetic encasement, the findings may need to be and 21 kN/m at 5% strain, respectively, which were comparable to those
further verified by well-instrumented test in prototype/actual stress of unpatched geosynthetics.
levels. Hence, in the present study, a series of centrifuge model tests are
conducted to investigate the performance of GESC-supported embank­
2.2. Model test preparation
ment over soft clay. Two types of geosynthetic material were used to
model encasement. A test with an ordinary stone column (OSC) was also
The centrifuge model tests were carried out using the geotechnical
performed for comparison. The vertical stresses and settlements gener­
centrifuge at Tongji University. The model tests were conducted in a
ated on top of the column and surrounding soil, the pore water pressures
model box with internal dimensions of 900 mm (length) × 700 mm
experienced in the soil, and the displacement profile obtained due to
(width) × 700 mm (height). The model box was made of steel on three
embankment loading were monitored during the test. The performance
sides, and the front side of the box was made of toughened glass, which
of the GESC-supported embankment was assessed in terms of consoli­
allowed the observation of the deformations of the embankment during
dation characteristic and load transfer mechanism.
the test. To minimize the side effect due to the friction of the steel
sidewall of the model box, a thin layer of petroleum grease was put on
2. Centrifuge model tests the inside steel sidewalls, followed by placing two polythene sheets.
Fig. 2 shows the schematic view of a centrifuge model test for GESC-
2.1. Materials supported embankment over soft soil. Due to the symmetry of the cross-
section, half the embankment was modeled. The model embankment
The commercially available kaolin clay was used to model soft with a height of 200 mm, a crest width of 200 mm, and a slope of 1:1.25
ground. The liquid limit and plastic limit of kaolin clay was found to be was constructed on a 400-mm thick soft clay. Stone columns with or
54.2% and 34.3%, respectively. The shear strength parameters of kaolin without geosynthetic encasement were installed in the soft clay. The
clay were evaluated from the consolidated undrained triaxial tests. The columns had a length of 400 mm and a diameter of 32 mm and were
effective cohesion and internal frictional angle of kaolin clay were found installed in a square pattern at a center-to-center spacing of 100 mm.
to be zero and 27.7◦ , respectively. The model embankment was made of
iron ore sand with particle sizes ranging from 0.5 to 2 mm. Due to the
high density of the iron ore sand, the embankment would apply larger
stress on the ground than the embankment formed by conventional sand
with equal height. The maximum and minimum dry densities of the sand
were found to be 2.4 and 1.8 g/cm3, respectively. The sand was com­
pacted to a relative density of 70% with a density of 2.2 g/cm3 to form
the embankment. The peak frictional angle of the sand at 70% relative
density was 35◦ based on direct shear test under the normal stresses of
25, 50, 100 and 150 kPa considering the self-weight of the embankment
in the centrifuge. Silica sand with particle sizes in a range of 2.5–3 mm
was used as the granular material for stone columns. The mean particle
size of the silica sand was 2.64 mm, and the uniformity coefficient was
1.891. The maximum and minimum dry densities of the silica sand were Fig. 1. Model encasement with G1 and G2.

2
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

Fig. 3. Variation of undrained shear strength of model ground with the depth.

Fig. 2. Schematic of model test: (a) plan view; (b) cross-section (unit: mm).

The column tips were directly placed on the base of the strong-box to
simulate the condition that the columns were seated on a firm soil layer,
like bed-rock.
The fully saturated remolded kaolin clay slurry was poured into the
strong box and pre-consolidated under the self-weight at a centrifuge
acceleration of 25 g for three hours. Then, the centrifuge was stopped.
The model ground was trimmed to 400 mm in thickness. Mini-cone
penetrometer test was conducted to determine the undrained shear
strength of the soft clay (Chen et al., 2012). Fig. 3 shows the profiles of
the undrained shear strength of the soft clay with depth for all test cases.
A gradual increase in the undrained shear strength of the soft clay with
depth can be noticed. The average undrained shear strength of the soft
clay with depth can be determined using the equation

Su = 0.004z + 4.93 (1)

in which, Su is the undrained shear strength (kPa) at depth z (mm).


After the model foundation was prepared, a hollow stainless steel
tube with an outer diameter of 32 mm and a wall thickness of 0.4 mm
was driven into the foundation. Auger extruder was used to remove the
soil inside the tube. The premanufactured geosynthetic encasement with
32 mm diameter and designed length (i.e., 400 mm or 200 mm) was
placed inside the tube. Silica sand was poured into the geosynthetic
encasement with a lift thickness of 50 mm. A repeated withdrawal and
penetration of the tube was carried out to densify the sand inside the
Fig. 4. Model preparation: (a) after the installation of columns, (b) after the
geosynthetic encasement to achieve the desired density of 1.77 g/cm3. construction of embankment.
Fig. 4(a) shows the prepared model ground with columns. Once the

3
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

column-reinforced ground was prepared, the model embankment was converge when the gravity acceleration increased to 6.5 g, corre­
formed above it adhering to the mentioned dimensions. In the stone sponding to 0.26 times the ultimate gravity acceleration (i.e., 6.5/25 =
column system, a horizontal drainage blanket is an essential component. 0.26). Fig. 5 shows the contours of the shear strain increments of the
The embankment was made of sand, and it can be functioned as a hor­ numerical model at 6.5 g. It is clear to see a globe failure taking place.
izontal drainage blanket. Fig. 4(b) shows the model after the embank­ On the other hand, the bearing capacity of the soft soil was 29.4 kPa (i.e.,
ment was completed. pu = 5.14Su = 5.14 × 5.73 = 29.4 ​ kPa, in which, Su is the undrained
′ ′

The performance of GESC-supported embankment was monitored shear strength at the mid-depth of the soil layer.) according to the
using several calibrated sensors like mini-pore pressure transducers Terzaghi method, however, the vertical stress induced by the proposed
(PPTs), linear variable displacement transducers (LVDTs), and mini- embankment is found to be 107.8 kPa, which is greater than the bearing
earth pressure cells (EPC). All the sensors were installed near the cen­ capacity of the foundation, indicating that bearing capacity failure
tral area of the model embankment. The layout of sensors installed in the would happen during the construction of the embankment. The results
centrifuge model test is shown in Fig. 2. Two LVDTs were installed to from these analyses indicate that the foundation soil would fail under
measure the settlement on the soil at midspan between two columns and embankment loading in terms of slope stability and bearing capacity due
the column head. Four EPCs were placed to measure the stresses to the low shear strength of the foundation, and hence ground
generated at different locations; in which, two on the column head and improvement in the form of OSC or GESC was attempted in this study. As
two on the soil at midspan between two columns. PPTs were used to the results are obvious for soft clays, the centrifuge model test for the
monitor the pore water pressure generated in the soft clay during the embankment resting on soft clay was not carried out in this study.
test; two PPTs were installed in soft clay between four columns at a
depth of 100 (P2) and 200 mm (P3) from the ground surface, and one in- 3. Performance of GESC-supported embankment
between two columns at a depth of 100 mm from the ground surface
(P1). 3.1. Column deformation after the test

2.3. Loading procedure After the centrifuge test, a portion of the foundation soil was care­
fully removed to observe the deformed columns. Fig. 6(a) presented the
After the installation of sensors at the pre-defined locations and the status of the ordinary stone columns after the test. The dash lines
completion of model column-supported embankment, the model box sketched the columns. Consistent with other reported literature (Hughes
was placed in the centrifuge. The embankment construction was simu­ et al., 1975), bulging at the upper portion of the columns was noticed in
lated by increasing the centrifugal acceleration linearly to 25 g within the test of OSCs and some sands were squeezed into the surrounding soil.
300 s to simulate a fast rate of embankment construction, which is On the other hand, lateral movements occurred for the columns posi­
termed as the loading stage in this study. And then the model test was tioned under the embankment slope.
run in 25 g centrifugal acceleration field for 600 s to simulate the Fig. 6(b) and (c) show the status of the GESCs after the test. Column
consolidation of the reinforced ground under the embankment loading, Nos. 2 to 4 bent outwards the embankment and the largest bending
which is called the consolidation stage in this study. The variations in occurred at column No. 4. For the columns close to the embankment
the settlement, pore water pressure, and load transfer were monitored centerline, they mainly suffered compression. Moreover, due to the
continuously in both the stages using the calibrated sensors and asso­ higher stiffness of the encasement, the bending deformations of the
ciated data acquisition system. Once the consolidation was allowed for columns in Test GESCs-G1F were less than those in Test GESCs-G2F.
the pre-defined duration, the centrifuge was stopped. Chen et al. (2015) indicated that bending failure could be the failure
mode for geosynthetic-encased stone columns under an embankment
instead of shear or bulging failures as the case for ordinary stone col­
2.4. Test program
umns. Li et al. (2016) presented a centrifugal test using a mixture of
sodium bentonite powder and glycerin to model surrounding soil and
Table 1 lists the information of the centrifuge model tests conducted
found that the bending shape was dependent on the strength of the
in this study, which includes the model size and the corresponding
surrounding soil. From a practical point of view, the flexural strength of
prototype size. In general, three model tests were designed: (i) ordinary
GESC may be considered while assessing the embankment stability,
stone column with no geosynthetic encasement (labeled as OSCs); (ii)
since the encased stone column could fail by bending.
stone column encased with geosynthetic material G1 (GESCs-G1F); (iii)
stone column encased with geosynthetic type G2, (GESCs-G2F).
The centrifuge test on embankment resting on soft clay (i.e. without
any ground improvement) was deliberately not carried out in the pre­
sent study for the following reasons. A two-dimensional numerical
analysis was conducted to simulate the embankment placed on the soft
soil, in which the geometry of the numerical modeling and the gravity
acceleration change were the same as the centrifugal test. The soft soil
and the embankment fill were modeled as Mohr-Coulomb failure
criteria, and their properties were the same as the materials used in the
centrifugal modeling. The undrained shear strength of the soft clay
increased with depth followed Eq. (1). The numerical model cannot

Table 1
Model test program (model/prototype).
Test label Lc (mm) s (mm) D (mm) Le (mm) G1/G2

OSCs 400/10,000 100/2500 32/800 N/A N/A


GESCs-G1F 400/10,000 100/2500 32/800 400/10,000 G1
GESCs-G2F 400/10,000 100/2500 32/800 400/10,000 G2

Note: Lc = column length; s = column spacing; D = column diameter; Le =


encasement length. Fig. 5. Contour of shear strain increments.

4
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

Fig. 6. Deformed columns after excavation: (a) OSCs; (b) GESCs-G1F; (c) GESCs-G2F.

3.2. Ground settlement (GESCs-G1F) and 12% (GESCs-G2F) on surrounding soil, respectively,
when compared with OSCs. This demonstrates the significance of
Fig. 7 shows the variation of settlements on the top of the column and encasing the stone columns with geosynthetic material in settlement
the surrounding soil with time. It is noticed that the ground settlements, reduction.
both on the column heads and the surrounding soils, increased at a fast Several analytical models available to analyze the performance of
rate during the loading stage, and then gradually attained constant GESC assumed that the settlements on the top of the column and the soil
values during the consolidation stage. The cases with encased stone are equal (equal strain approach), which is applicable to the condition
columns had relatively lesser ground settlements than the case with under a rigid loading (Raithel and Kempfert, 2000; Alexiew et al., 2005;
ordinary stone columns. Test GESCs-G1F had relatively lesser ground Zhang et al., 2012; Pulko and Logar, 2017). Differential vertical strain,
settlements than Test GESCs-G2F, resulting from the fact that the geo­ which is defined as the differential settlement between column head and
synthetic reinforcement G1 had larger stiffness than G2. Zhou and Kong the surrounding soil divided by the thickness of soft soil (i.e., 400 mm in
(2019) also theoretically indicated that encasement with high stiffness this study), is proposed to assess whether the equal strain condition is
plays a significant role in the settlement of GESC-supported embank­ valid for embankment loading. As shown in Fig. 8, the differential ver­
ment. At the end of the consolidation stage, settlements reduced by 73% tical strains increased fast during the loading stage and gradually
(GESCs-G1F) and 45% (GESCs-G2F) on column head, and by 27% reached constant values during the consolidation stage. The increase of

Fig. 7. Variation of settlement with time. Fig. 8. Differential vertical strain with time.

5
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

the differential settlement was mainly due to the development of the for high stiffness geosynthetic material than lower stiffness geosynthetic
settlement on the soil as shown in Fig. 7. At the end of the test, differ­ material.
ential vertical strains in the case with OSCs was 1.7%; while large dif­
ferential vertical strains of 6.5% and 5.3% were noticed in the cases with 3.4. Stresses on column and soil
encased stone columns (i.e., tests GESCs-G1F and GESCs-G2F), respec­
tively. Zhou and Kong (2019) also indicated the differential settlement Fig. 10 presents the variations of the average vertical stresses
between the encased stone column and the surrounding soil under generated on the top of the column and surrounding soil with time. As
embankment loading. The centrifugal test results reveal that the equal
strain condition does not exist for encased stone column-reinforced soil
under embankment loading. While for the ordinary stone
column-reinforced soil under embankment loading, this approach may
be used with precautions.

3.3. Excess pore water pressure

Fig. 9 shows the variation of the excess pore water pressure at


different monitoring points with time. The excess pore water pressure at
each monitoring point displays a similar trend, as it increased rapidly
during the embankment loading stage and dissipated at a fast rate during
the consolidation stage. Noted from Fig. 7, during consolidation the
settlement was mainly on the soft soil, which is consistent with the
dissipation of excess pore water pressure. The maximum excess pore
pressure generated at the end of the embankment loading stage was
found to be lower for encased stone columns when compared with the
OSC case, irrespective of the type of geosynthetic material. The Fig. 10. Variation of vertical stresses on embankment base with time.
maximum reduction in the generation of excess pore pressures is noticed

Fig. 9. Variation of excess pore water pressure at monitoring point of: (a) P1; (b) P2; (c) P3.

6
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

expected, the vertical stresses generated on the column were higher than benefit of the geosynthetic encasement. Based on the test results, the
those on the surrounding soil and increased obviously with the increase improvement factor was 2.8 and 1.7 for the tests of GESCs-G1F and
of encasement stiffness. Meanwhile, the vertical stresses generated on GESCs-G2F, respectively. Some previous studies showed an improve­
the surrounding soil in Test GESCs-G1F and GESCs-G2F were smaller ment factor around 2.5 to 5 under a rigid loading (Gniel and Bouazza,
than that in Test OSCs; and decreased with the increase of encasement 2009; Miranda et al., 2017). As the improvement factor is a dimen­
stiffness. The above phenomenon can be used to explain the character­ sionless index, the comparison has certain guiding meaning for design:
istics of excess pore water pressure in the former section. As the encased lower stress concentration ratio than that under rigid loading shall be
stone column had a larger modulus than the ordinary stone column, considered in the design of GESC-supported embankment, but the
more embankment load was carried by the encased stone columns. The improvement factor would also be influenced by other factors, including
reduction of the embankment load carried by the surrounding soil properties of soil, column, and encasement.
resulted in a reduction of the excess pore water pressure in the soil. Soil arching ratio (SAR) is used to evaluate the degree of soil arching
During the consolidation stage, the vertical stress on the column head in effect, which is defined as a ratio of the applied stress on the soil to the
Test GESCs-G1F slightly increased, but other tests display a reduction total overburden stress. Stress on the soil all transferred onto adjacent
more or less in vertical stress on column head after peak values. columns (i.e., SAR = 0) is considered complete soil arching, whereas the
stress on the soil equal to the overburden stress (i.e., SAR = 1) represents
no soil arching. Iglesia et al. (2013) used the curve of soil arching ratio
3.5. Load transfer with the relative displacement (called Ground Reaction Curve (GRC)) to
characterize the progressive development of soil arching with four
The stress concentration ratio is an index to evaluate the load stages: initial soil arching, maximum soil arching, stress recovery, and
transfer between column and soil, which is defined as a ratio of applied ultimate state. The GRC has been used by researchers (Han et al., 2017;
vertical stress on the column head to that on the soil. It is usually equal King et al., 2017) to describe soil-arching development in real projects.
to or greater than unity as the column has a greater stiffness than soil. Fig. 12 presents the progressive development of the soil arching ratio
Increase of stress concentration ratio indicates the embankment load is with the normalized displacement. The normalized displacement was
transferred from soil to column, whereas its decrease represents the determined as the differential settlement between the column and the
embankment load is transferred from column to soil. Fig. 11 presents the surrounding soil divided by the clear spacing between two diagonal
variation of the stress concentration ratio with time for all the tests. In columns. The soil arching curves in the tests of GESCs-G1F and
the tests with encased stone columns, the stress concentration ratio GESCs-G2F approximately followed the GRC defined by Iglesia et al.
increased to peak values in the loading stage. However, the stress con­ (2013). The soil arching ratio dropped sharply after a small normalized
centration ratio in Test OSCs started to decrease before the end of the displacement happened, and then the decrease rate gradually slowed
loading stage, indicating that the column might yield during the loading down until a minimum value. In the tests of GESCs-G1F and GESCs-G2F,
stage. In the consolidation stage, the stress concentration ratio decreased the minimum soil arching ratios of 0.692 and 0.784 were obtained at
with time, and the reduction became larger in Test OSCs. While for Test 8.6% and 4.7% of the normalized displacement, followed by a slight
GESCs-G1F, the stress concentration ratio had a slight increase during recovery to 0.694 and 0.807 at the end of the test. However, due to the
the consolidation stage. bulging of column in the test of OSCs, its ultimate soil arching ratios of
The tests of OSCs, GESCs-G1F, and GESCs-G2F had steady stress 0.903 unveiled a low soil arching effect.
concentration ratios of 2.3, 6.5, and 4.0, respectively. The results
demonstrated that the encasement significantly increased the column
3.6. Strain-stress relation of column
stiffness. The use of encasement with high stiffness increased the stress
concentration ratio. Based on the analytical method by Raithel and
Fig. 13 presents the relation between the vertical stress and the
Kempfert (2000), the steady stress concentration ratio was 10.4 in the
vertical strain of columns during the loading stage. The vertical strain is
case of GESCs-G1F and 4.6 in the case of GESCs-G2F, which are both
defined as a ratio of the measured settlement at the top of the column to
higher than the measured results, as this method was proposed for
the length of the column. It reveals that the columns in GESCs-G2F
GESC-reinforced soft soil under a rigid loading. In the calculations, the
yielded after 25 g centrifugal acceleration was reached, while the ordi­
model scale was transferred to the prototype scale with the scaling factor
nary stone column yielded earlier during the process of centrifugal ac­
of 25, and the input properties are included in the Appendix.
celeration increasing due to the lower confinement. The encasement
The improvement factor in terms of stress concentration ratio, which
increased the yield stress of the column. On the other hand, the vertical
is defined as a ratio of the stress concentration ratio of the encased stone
column to that of an ordinary stone column, is introduced to assess the

Fig. 11. Variation of stress concentration ratio with time. Fig. 12. Development of soil arching ratio with normalized displacement.

7
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

embankment at 1 g would cause additional foundation deformation as


compared with in-flight placement of the embankment. However, it
should be noted that the interaction between the soft soil and the stone
column was not considered in the analysis. It is reasonable to think that
during the reactivation of self-weight at 25 g, the compression of stone
column was not significant as it was compacted to a dense state at 1 g.
The frictional interaction between the column and the soft soil during
self-weight reactivation could reduce the overall recompression of the
soft soil (see Fig. 15, Δs’<Δs). Thus, neglecting the existence of columns
is a conservative way to consider the influence of the embankment
construction at 1 g.
The compression and swelling indexes of the kaolin clay, Cc and Cs
were 0.317 and 0.028 from the consolidation test, respectively. If the
interaction between the columns and the foundation is neglected, the
extra soil settlement induced by stopping the centrifuge can be conser­
Fig. 13. Stress-strain of encased and non-encased stone columns. vatively estimated as
( )
H0 p2 400
stress on the column in the test GESCs-G2F dropped more rapidly than Δs = Cs lg = × 0.028 × lg25 = 7.6mm (2)
1 + eC p1 1 + 1.05
those on the columns in the tests of OSCs and GESCs-G1F. The column in
the test of GESCs-G1F did not yield during the test. Taking the initial in which, H0 is the thickness of the soft clay, and other notations have
tangent modulus in each test for comparison, as expected, the encased been explained previously. As a result, the measured settlement on the
column had larger modulus than the ordinary stone column, and it soft soil can be corrected by subtracting the extra settlement. Table 2
increased with the increase of encasement stiffness. presents the settlements on column and soil and their corresponding
differential vertical strain. Clearly, the differential vertical strain
4. Discussion reduced after the foundation heaving part was subtracted. There was
almost no differential settlement for Test OSCs, while relatively larger
4.1. Assessment of construction of embankment at 1 g differential settlements in the tests with encased stone columns were
observed. This can be explained by a fact that the encased stone column
As the friction effect between the inside wall of the model box and has a larger modulus than the soft soil.
the ground was minimized by placing two polythene sheets between
them, the soft soil rebounded almost under one-dimensional (1-D) 4.2. Stress reduction factor
condition after the centrifuge was stopped for the pre-consolidation of
soft soil at 25 g. Thus, the difference between in-flight construction of Following the stress-reduction method by Aboshi (1979), the set­
embankment and embankment constructed at 1 g can be explained tlements of ordinary or encased stone column-reinforced foundations
through a plot of the void ratio versus vertical effective stress as shown can be expressed as,
in Fig. 14. When the vertical pressure on a soil element increased from
the initial overburden pressure, p1 (point A) to p2 (point B) by activating ′
Sos = μos S, μos =
1
(3a)
acceleration from 1 g to 25 g, the void ratio decreased from eA to eB 1 + as (nos − 1)
following the compression line AB. After the completion of the consol­
1
idation under self-weight, the centrifuge was stopped, thus the vertical (3b)

Ses = μes S, μes =
1 + as (nes − 1)
pressure decreased to p1. During this process, the soil element reboun­
ded through the line from point B to C (at void ratio, eC). Clearly, eC is
in which, S is the settlement of natural foundation; S′ is the settlement of
less than eA. After the construction of the embankment at 1 g and the
column-reinforced foundation; μ is the stress reduction factor; as is the
activation of acceleration from 1 g to 25 g again, the vertical pressure
area replacement ratio of column; n is the steady stress concentration
could increase from p1 (point C) to p3 (point D), the induced void ratio
ratio; and the subscripts os and es denote ordinary stone column and
would be ΔeBC+ΔeBD following the recompression line CB and the
encased stone column, respectively. From Eqs. (3a)–(3b), the following
compression line BD. In other words, the construction of the
relations can be obtained:

(4a)
′ ′ ′
Ses = μ Sos

μes 1 + as (nos − 1)
(4b)

μ= =
μos 1 + as (nes − 1)

Fig. 15. Ground settlement influenced by the embankment construction at 1 g:


Fig. 14. Effective stress versus void ratio. (a) after self-weight consolidation; (b) reactivation of self-weight at 25 g.

8
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

Table 2 Table 4
Settlement and differential vertical strain. Maximum excess pore pressure generated at the monitoring points for all the test
Test Settlement (mm) Differential vertical
cases (kPa).
labels strain (%) Test label P1 P2 P3
Column Soil Soil Measured Corrected OSCs 33.1 37.7 53.4
(Measured) (Measured) (Corrected) GESCs-G1F 26.7 27.2 31.9
GESCs-G2F 30.3 34.6 45.4
OSCs 45.24 52.18 44.58 1.7 − 0.2
GESCs- 12.03 38.21 30.61 6.5 4.6
G1F
GESCs- 24.99 46.05 38.45 5.3 3.4
8crm π2 cvm
G2F β= + (5b)
de2 Fm′ 4h2dr

in which, μ′ is the stress-reduction factor from OSC-reinforced founda­ in which, ut is the pore water pressure at time t; uinc is the excess pore
tion to GESC-reinforced foundation. water pressure increment due to loading; crm and cvm are the modified
Table 3 shows the μ′ calculated in terms of settlement (i.e., Eq. (4a)) coefficients of consolidation in the radial and vertical directions; de is the
and stress (i.e., Eq. (4b)). Ses and Sos were determined based on the area-
′ ′
diameter of influence zone for a single column; hdr is the longest
weighted average of the settlements of the column head and the sur­ drainage path due to vertical flow; and Fm is the factor considering

rounding soft soils at the end of the test. It is noted that μ′ calculated smear and well resistance effects of stone column. Letting λ = crm/cvm,
based on Eqs. (4a) and (4b) agreed well for the encased stone column- Eq. (5b) can be rearranged in the following forms:
reinforced foundations. The settlements calculated by Raithel and ( ) ( )
Kempfert (2000) which was developed based on equal strain assump­ 8λ π2 8 π2
β = cvm 2 ′ + 2 = crm 2 ′ + (5c)
tion, were 9.28 mm for Test GESCs-G1F and 12.04 mm for Test de Fm 4hdr de Fm 4λh2dr
GESCs-G2F. The less settlement calculations than the measured Ses are

From Eq. (5), the excess pore water pressure at arbitrary time ti and ti-
attributed to the significant differential settlements between column and (i.e., Δt = ti − ti− 1 ) after embankment load was maintained for soil
1
soil under embankment loading, in which equal strain condition cannot consolidation can be given by the relation,
be strictly satisfied.
ui = ui− 1 e− βΔt
(6)

4.3. Consolidation rate where, ui and ui-1 are the average excess pore water pressures at time ti
and ti-1, respectively. Assuming a constant β, Eq. (6) reveals that a plot of
Table 4 summarizes the measured maximum increments of excess (ui-1, ui) at different times should be a straight line passing the original.
pore water pressure at each monitoring point during the loading stage. Aboshi (1979) proposed a method to back calculate β based on the
Test GESCs-G1F yielded the lowest increment of excess pore water measured settlement during consolidation. Following the Aboshi (1979)
pressure, while Test OSCs the highest one among the three tests. For the method, β can be determined by
case of an encased stone column, due to its higher stiffness, the encased
stone column carried more embankment load than the ordinary stone β= −
ln(α1 )
(7)
column, resulting in a reduction in the excess pore water pressure Δt
generated in the soil during the loading stage.
in which, α1 is the slope of the best-fit line through the parallel settle­
The excess pore water pressure generated in the soft clay was found
ments of si at current time step and si-1 at the previous time step, and Δt is
to be increased with the increases in the horizontal distance from the
the time interval for the successive settlements.
column and depth from the ground surface. After stone columns or
The predicted pore water dissipation curves using Eq. (5a) are pre­
encased stone columns were installed, the excess pore water pressures in
sented in Fig. 9, and reasonable agreements were found with the mea­
the soil generated by the embankment loading were dissipated in both
surements. Table 5 shows that the back-calculated β based on the pore
vertical and horizontal directions. The vertical drainage path in the
water pressure yielded reasonable agreements with those based on the
shallow portion was shorter than that in the deep portion, and the
settlement, indicating that the dissipation of excess pore pressures
discharge capacity of the stone column reduced after it deformed or
during the consolidation stage was compatible with the measured set­
partially clogged by the surrounding soil. Another reason for the in­
tlement during consolidation. The deviations of back-calculated β were
crease in the excess pore water pressure with depth could be due to
also noted. This can be explained by facts that the method the Aboshi
geostatic stress condition.
(1979) method was developed based on one-dimensional consolidation
The pore water pressure at time t in the dissipation process can be
theory, while the lateral deformation of soil occurred under the
expressed in a general format:
embankment loading. Castro and Sagaseta (2011) and Jiang et al.
ut = uinc e− βt
(5a) (2013) also found that it affected the consolidation rate of
column-reinforced foundation after allowing the lateral deformation in
according to Han and Ye (2002), where the unit cell model. On the other hand, the back-calculated β based on
the pore water pressure indicated the β of soft soil at the measuring
Table 3 point, while the back-calculated β based on the settlement indicated the
Calculation of μ′
Test label S’(mm) Sos or

nos or μ′
nes
Table 5
Ses (mm)

Soil Column by Eq. by Eq. Back calculation of β (10− 3/s).


(4a) (4b)
Test label Pore water pressure Settlement
OSCs 52.18 45.24 51.62 2.34 N/A N/A
P1 P2 P3
GESCs- 38.21 12.03 36.12 6.51 0.70 0.77
G1F OSCs 2.40 2.08 1.28 3.10
GESCs- 46.05 24.99 44.37 3.99 0.86 0.89 GESCs-G1F 3.38 3.21 2.42 4.79
G2F GESCs-G2F 2.79 2.52 1.61 3.51

9
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

overall β of soil layer. Table 6


For the β determined either by the pore water pressures or the set­ Calculated.ces os
vm /cvm
tlements, Tests GESCs-G1F and GESCs-G2F had larger β than Test OSCs, Test label Pore water pressure Settlement Eq. (8)
and the stiffer encasement resulted in larger β. Miranda et al. (2017)
P1 P2 P3
found that the excess pore pressures dissipated faster in the soils with
encased stone columns than in those with non-encased columns. Castro GESCs-G1F 1.41 1.55 1.88 1.54 1.30
GESCs-G2F 1.16 1.21 1.25 1.13 1.12
and Sagaseta (2011) also indicated that consolidation is faster for stiffer
encasement from a theoretical point of view.
Assuming λ = crm/cvm and Fm to be equal in all tests, the ratio of ces 6. Conclusions

rm /
os es os
crm or cvm /cvm , in which, superscripts es and os in crm and cvm denote
encased stone column and ordinary stone column, respectively, can be This paper presented a series of centrifugal modeling on the perfor­
obtained by its corresponding back-calculated β. Based on the simplified mance of geosynthetic-encased stone column-supported embankment
method proposed by Han and Ye (2002), ces os
rm /crm ratio can also be ob­ over soft clay. Based on the analyses and discussions, the following
tained in terms of stress concentration ratio: conclusions can be drawn:
( )
es es 1 + Nn2es− 1 1. Using the encased stone columns to support embankment signifi­
crm cvm
= os = ( ) (8) cantly reduced the ground settlement, enhanced the stress concen­
cos cvm
rm
1 + Nn2os− 1 tration ratio, accelerated consolidation, and increased column
stiffness when compared with the OSCs. The benefits became more
prominent when adopting encasement with higher stiffness. A rela­
in which, N is the diameter ratio, N = de/dc; and de and dc are the di­
tively strong soil arch was noticed in the tests with GESCs as
ameters of column and its influence zone, respectively. Other notations
compared with the test with OSCs. Based on the test results, Tests
have been explained previously. Table 6 shows the calculated ces os
vm /cvm
GESCs-G1F and GESCs-G2F had improvement factor of 2.8 and 1.7,
based on back-calculated β, and Eq. (8) using the measured stress con­
respectively; and back-calculated modified coefficients of consoli­
centration ratios. The calculated ces os
vm /cvm were greater than 1.0, indi­
dation of 1.1–1.9 times those in Test OSCs.
cating that encased stone column-reinforced foundation accelerates the
2. Under the embankment loading, encased stone columns could fail by
consolidation rate as compared with ordinary stone column-reinforced
bending instead of bulging or shear as noticed in the case of ordinary
foundation. It is noted that the results by Eq. (8) were smaller than
stone columns. From a practical point of view, the bending strength
those by back-calculated β. This could be explained by a possible reason
of GESC shall be considered when assessing the embankment
that the simplified method proposed by Han and Ye (2002) was
stability.
appropriate to the condition under rigid loading, and the proposed crm
3. The stress reduction method can be used to calculate the settlement
and cvm represent the overall modified coefficients of consolidation for
of encased stone column-supported embankment with larger stress
subsoil layer rather than the one at a certain point in subsoil.
reduction factor than the ordinary stone column-supported
embankment.
5. Limitation
4. The calculations based on equal strain assumption yielded less
ground settlements and larger stress concentration ratios than the
In the centrifuge modeling, the columns and the embankments were
measurements due to the differential settlements between column
prepared at 1 g and then increased to 25 g to simulate a prototype. Thus,
and soil under embankment loading. Additionally, the construction
this study was not designed to capture the effect of the construction
of the embankment at 1 g could increase the differential settlements.
stages of embankment on the treated foundation soil. As discussed
earlier, the construction of the embankment at 1g would generate
Acknowledgements
additional foundation deformation as compared with the in-flight con­
struction of the embankment. However, by applying suitable corrections
The authors appreciate the financial support provided by the Natural
in the computation of the settlement, a reasonable estimate of differ­
Science Foundation of China (NSFC) (Grant No. 41572266, No.
ential settlement can be obtained even when the embankment was
51508408 & No. 41772289), and by the Shanghai International Science
constructed at 1g. Moreover, this approach of constructing model
and Technology Cooperation Fund (Grant No. 18230742700). The fifth
embankment at 1 g has been adopted by several researchers because of
author Dr. Sathiyamoorthy Rajesh from the Indian Institute of Tech­
the non-availability of the sophisticated installation tools (like inflight
nology Kanpur is thankful for the Talented Young Scientist Program
sand hopper). Having said that, the actual behavior of the embankment
supported by China Science and Technology Exchange Center. This
resting on the soft ground during the construction of the embankment
study was also financially supported by the Key Research and Devel­
can be accurately captured by the inflight construction of the embank­
opment Project of Chinese Ministry of Science and Technology (Grant
ment, which is beyond the scope of the present study.
No. 2016YFE0105800).

Notation list

Basic SI units are given in parentheses


as area replacement ratio of column
crm, cvm modified coefficients of consolidation in the radial and vertical directions (m2/s)
dc, de diameters of column and its influence zone
Fm factor considering smear and well resistance effects of stone column

hdr longest drainage path due to vertical flow (m)


N diameter ratio, N = de/dc
ns steady stress concentration ratio of column-reinforced foundation
nos, nes steady stress concentration ratios for OSC-reinforced foundation and GESC-reinforced foundation

10
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

S settlements of a natural foundation (m)


Sos , Ses settlements of ordinary stone column-reinforced foundation and geosynethetic-encased stone column-reinforced foundation (m)
′ ′

Δt time interval, Δt = ti − ti− 1


ut pore water pressure at time t (kPa)
uinc excess pore water pressure increment due to loading (kPa)
ui, ui-1 average excess pore water pressures at time ti and ti-1 (kPa)
μos, μes stress reduction factors of ordinary stone column-reinforced foundation and geosynethetic-encased stone column-reinforced foundation
μ′ stress reduction factor from OSC-reinforced foundation to GESC-reinforced foundation
λ ratio of modified coefficients of consolidation, crm/cvm

APPENDIX

In the calculations, the model scale was transferred to the prototype scale following the scaling laws with scaling factor of 25. The soft soil had the
following properties: constraint modulus with stress range of 0.1–0.2 MPa, Ds = 2520 kPa; saturated unit weight, γ s = 15.3 kN/m3; Poisson’s ratio, vs
= 0.35; and effective friction angle, φs = 27.7◦ . The groundwater table was at the ground surface. The embankment had a height of 5 m and density of

3
2.2 g/cm . GESC with a diameter of 800 mm were installed in a square pattern with spacing of 2.5 m which corresponded to area replacement ratio, as
= 0.08. The sand columns had density of 1.77 g/cm3 and effective friction angle, φc = 38◦ . The tensile stiffness of the geosynthetic encasement was, J

= 3300 kN/m for G1 and 450 kN/m for G2, respectively.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.geotexmem.2020.10.021.

References Gu, M., Han, J., Zhao, M., 2017. Three-dimensional dem analysis of single geogrid-
encased stone columns under unconfined compression: a parametric study. Acta
Geotechnica 12 (3), 559–572.
Aboshi, H., 1979. The compozer, a method to improve characteristics of soft clays by
Han, J., 2015. Principles and Practice of Ground Improvement. John Wiley & Sons, Inc.,
inclusion of large diameter sand columns. In: Proceedings of 1st International
Hoboken, New Jersey.
Conference on Soil Reinforcement, 1. E.N.P.C., Paris, pp. 211–216.
Han, J., Ye, S.L., 2002. A theoretical solution for the rate of consolidation of a stone
Alexiew, D., Brokemper, D., Lothspeich, S., 2005. Geotextile encased columns (GEC):
column reinforced foundation accounting for smear and well resistance. Int. J.
load capacity, geotextile selection and pre-design graphs. In: Proceedings of the Geo-
GeoMech. 2 (2), 135–151.
Frontiers Conference, Austin, Texas, USA. ASCE, Reston, VA, USA, pp. 497–510.
Han, J., Wang, F., Al-Naddaf, M., Xu, C., 2017. Progressive development of two-
Ali, K., Shahu, J.T., Sharma, K.G., 2012. Model tests on geosynthetic-reinforced stone
dimensional soil arching with displacement. Int. J. GeoMech. 17 (12), 04017112.
columns: a comparative study. Geosynth. Int. 19 (4), 292–305.
Hong, Y.S., 2012. Performance of encased granular columns considering shear-induced
Ali, K., Shahu, J.T., Sharma, K.G., 2014. Model tests on single and groups of stone
volumetric dilation of the fill material. Geosynth. Int. 19 (6), 438–452.
columns with different geosynthetic reinforcement arrangement. Geosynth. Int. 21
Hong, Y.S., Wu, C.S., Yu, Y.S., 2016. Model tests on geotextile-encased granular columns
(2), 103–118.
under 1-g and undrained conditions. Geotext. Geomembranes 44 (1), 13–27.
Alkhorshid, N.R., Araujo, G.L.S., Palmeira, E.M., Zornberg, J.G., 2019. Large-scale load
Hong, Y.S., Wu, C.S., Kou, C.M., Chang, C.H., 2017. A numerical analysis of a fully
capacity tests on a geosynthetic encased column. Geotext. Geomembranes 47 (5),
penetrated encased granular column. Geotext. Geomembranes 45 (5), 391–405.
631–641.
Hosseinpour, I., Almeida, M.S.S., Riccio, M., 2015. Full-scale load test and finite-element
Almeida, M.S.S., Hosseinpour, I., Riccio, M., 2013. Performance of a geosynthetic-
analysis of soft ground improved by geotextile-encased granular columns. Geosynth.
encased column (GEC) in soft ground: numerical and analytical studies. Geosynth.
Int. 22 (6), 428–438.
Int. 20 (4), 252–262.
Hughes, J.M.O., Withers, N.J., Greenwood, D.A., 1975. A field trial of the reinforcing
Almeida, M.S.S., Hosseinpour, I., Riccio, M., Alexiew, D., 2014. Behavior of geotextile-
effect of a stone column in soil. Geotechnique 25 (1), 31–44.
encased granular columns supporting test embankment on soft deposit. J. Geotech.
Iglesia, G.R., Einstein, H.H., Whitman, R.V., 2013. Investigation of soil arching with
Geoenviron. Eng. 141 (3), 04014116.
centrifuge tests. J. Geotech. Geoenviron. Eng. 140 (2), 04013005.
Castro, J., 2017. Groups of encased stone columns: influence of column length and
Jiang, Y., Han, J., Zheng, G., 2013. Numerical analysis of consolidation of soft soils fully-
arrangement. Geotext. Geomembranes 45 (2), 68–80.
penetrated by deep-mixed columns. KSCE Journal of Civil Engineering 17 (1),
Castro, J., Sagaseta, C., 2011. Deformation and consolidation around encased stone
96–105.
columns. Geotext. Geomembranes 29 (3), 268–276.
Kadhim, S., Parsons, R.L., Han, J., 2015. March. Stability analysis of embankments
Castro, J., Sagaseta, C., 2013. Influence of elastic strains during plastic deformation of
supported by geosynthetic encased stone columns. In: IFCEE 2015 International
encased stone columns. Geotext. Geomembranes 37, 45–53.
Association of Foundation DrillingDeep Foundation InstitutePile Driving Contractors
Cengiz, C., Güler, E., 2018. Seismic behavior of geosynthetic encased columns and
AssociationAmerican Society of Civil Engineers, pp. 2318–2327.
ordinary stone columns. Geotext. Geomembranes 46 (1), 40–51.
King, D.J., Bouazza, A., Gniel, J.R., Rowe, R.K., Bui, H.H., 2017. Serviceability design for
Cengiz, C., Kilic, I.E., Guler, E., 2019. On the shear failure mode of granular column
geosynthetic reinforced column supported embankments. Geotext. Geomembranes
embedded unit cells subjected to static and cyclic shear loads. Geotext.
45 (4), 261–279.
Geomembranes 47 (2), 193–202.
Li, L.-Y., Chen, J.-F., Xu, C., Feng, S.-Z., 2016. Centrifuge modeling of geosynthetic-
Chen, J.F., Liu, J.X., Ma, J., 2012. Calibration of a miniature cone penetrometer for
encased stone columns in soft clay under embankment. In: Transportation Research
geotechnical model test. J. Tongji Univ.: Nat. Sci. 40 (4), 549–552, 588. (in Chinese).
Congress 2016: Innovations in Transportation Research Infrastructure - Proceedings
Chen, J.F., Li, L.Y., Xue, J.F., Feng, S.Z., 2015. Failure mechanism of geosynthetic-
of the Transportation Research Congress 2016, pp. 486–495 v 2018-January.
encased stone columns in soft soils under embankment. Geotext. Geomembranes 43
Madabhushi, G., 2014. In: Centrifuge Modelling for Civil Engineers. CRC Press, London,
(5), 424–431.
UK. Barksdale, R.D., Bachus, R.C., 1983. Design and Construction of Stone Columns,
Chen, J.F., Wang, X.T., Xue, J.F., Zeng, Y., Feng, S.Z., 2018. Uniaxial compression
FHWA/RD-83/026.
behavior of geotextile encased stone columns. Geotext. Geomembranes 46 (3),
Miranda, M., Da Costa, A., 2016. Laboratory analysis of encased stone columns. Geotext.
277–283.
Geomembranes 44 (3), 269–277.
Dash, S.K., Bora, M.C., 2013. Influence of geosynthetic encasement on the performance
Miranda, M., Costa, A.D., Castro, J., Sagaseta, César, 2017. Influence of geotextile
of stone columns floating in soft clay. Can. Geotech. J. 50 (7), 754–765.
encasement on the behaviour of stone columns: laboratory study. Geotext.
Elsawy, M.B.D., 2013. Behaviour of soft ground improved by conventional and geogrid-
Geomembranes 45 (1), 14–22.
encased stone columns, based on FEM study. Geosynth. Int. 20 (4), 276–285.
Mohapatra, S.R., Rajagopal, K., 2017. Undrained stability analysis of embankments
Fattah, M.Y., Zabar, B.S., Hassan, H.A., 2016. Experimental analysis of embankment on
supported on geosynthetic encased granular columns. Geosynth. Int. 24 (5),
ordinary and encased stone columns. Int. J. GeoMech. 16 (4), 04015102.
465–479.
Gniel, J., Bouazza, A., 2009. Improvement of soft soils using geogrid encased stone
Mohapatra, S.R., Rajagopal, K., Sharma, J., 2017. 3-Dimensional numerical modeling of
columns. Geotext. Geomembranes 27 (3), 167–175.
geosynthetic-encased granular columns. Geotext. Geomembranes 45 (3), 131–141.
Gu, M., Zhao, M., Zhang, L., Han, J., 2016. Effects of geogrid encasement on lateral and
Murugesan, S., Rajagopal, K., 2009. Studies on the behavior of single and group of
vertical deformations of stone columns in model tests. Geosynth. Int. 23 (2),
geosynthetic encased stone column. J. Geotech. Geoenviron. Eng. 136 (1), 129–139.
100–112.

11
J.-F. Chen et al. Geotextiles and Geomembranes xxx (xxxx) xxx

Pulko, B., Logar, J., 2017. Fully coupled solution for the consolidation of poroelastic soil Weber, T.M., Plotze, M., Laue, J., Peschke, G., Springman, S.M., 2010. Smear zone
around geosynthetic encased stone columns. Geotext. Geomembranes 45 (6), identification and soil properties around stone columns constructed in-flight in
616–626. centrifuge model tests. Geotechnique 60 (3), 197–206.
Pulko, B., Majes, B., Logar, J., 2011. Geosynthetic-encased stone columns: analytical Wu, C.S., Hong, Y.S., 2014. A simplified approach for evaluating the bearing
calculation model. Geotext. Geomembranes 29 (1), 29–39. performance of encased granular columns. Geotext. Geomembranes 42 (4), 339–347.
Raithel, M., Kempfert, H.G., 2000. Calculation models for dam foundations with Xue, J., Liu, Z., Chen, J., 2019. Triaxial compressive behaviour of geotextile encased
geotextile coated sand columns. In: Proceedings of International Conference on stone columns. Comput. Geotech. 108, 53–60.
Geotechnical and Geological Engineering, GeoEng 2000, pp. 347–352. Melbourne, Yoo, C., 2015. Settlement behavior of embankment on geosynthetic-encased stone
Australia. column installed soft ground - a numerical investigation. Geotext. Geomembranes 43
Raithel, M., Kirchner, A., Schade, C., 2005. Foundation of constructions on very soft soils (6), 484–492.
with geotextile encased columns -state of the art. In: Proceedings of GeoFrontiers Yoo, C., Abbas, Q., 2020. Laboratory investigation of the behavior of a geosynthetic
2005, pp. 1–11. Austin, Texas, USA. encased stone column in sand under cyclic loading. Geotext. Geomembranes 48 (4),
Rajesh, S., 2017. Time dependent behaviour of fully and partially penetrated 431–442.
geosynthetic encased stone columns. Geosynth. Int. 24 (1), 60–71. Yoo, C., Lee, D., 2012. Performance of geogrid-encased stone columns in soft ground:
Rajesh, S., Jain, P., 2015. Influence of permeability of soft clay on the efficiency of stone full-scale load tests. Geosynth. Int. 19 (6), 480–490.
columns and geosynthetic encased stone columns-a numerical study. Int. J. Geotech. Zhang, Y., Chan, D., Wang, Y., 2012. Consolidation of composite foundation improved by
Eng. 9 (5), 483–493. geosynthetic-encased stone columns. Geotext. Geomembranes 32, 10–17.
Schnaid, F., Winter, D., Silva, A.E.F., Alexiew, D., Küster, V., 2017. Geotextile encased Zhang, Z., Ye, G.B., Cai, Y.S., Zhang, Z., 2019. Centrifugal and numerical modeling of
columns (GEC) used as pressure-relief system. Instrumented bridge abutment case stiffened deep mixed column-supported embankment with a slab over soft clay. Can.
study on soft soil. Geotext. Geomembranes 45 (3), 227–236. Geotech. J. 56 (10), 1418–1432.
Schofield, A.N., 1980. Cambridge geotechnical centrifuge operations. Geotechnique 30 Zhou, Y., Kong, G., 2019. Deformation analysis of geosynthetic-encased stone
(3), 227–268. column–supported embankment considering radial bulging. Int. J. GeoMech. 19 (6),
Weber, T.M., Laue, J., Springman, S.M., 2006. Centrifuge modelling of sand compaction 04019057.
piles in soft clay under embankment load. In: Proceedings of the VI International
Conference on Physical Modelling in Geotechnics Hong Kong, 1, pp. 603–608.

12

You might also like