You are on page 1of 54

Page 1 of 54

For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Centrifugal and numerical modeling of stiffened deep mixed

column-supported embankment with a slab over soft clay

Zhen Zhang1; Guanbao Ye2*; Yongsheng Cai3; Zhao Zhang4


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

1Assistant Professor. Department of Geotechnical Engineering, Tongji University, Shanghai 200092, China.

dyzhangzhen@gmail.com
2Professor. Department of Geotechnical Engineering, Tongji University, 1239 Siping Road, Shanghai 200092, China.

Email: guanbaoye@gmail.com(*corresponding author)

3Engineer. SGIDI Engineering Consulting (Group) Co., Ltd. Shanghai 200093, China. Email: cys036@163.com

4Master student. Department of Geotechnical Engineering, Tongji University, Shanghai 200092, China. Email:

1395483534@qq.com

Abstract: A stiffened deep mixed (SDM) column can significantly increase the bearing capacity,

reduce settlement, and enhance the slope stability of soft clays as compared with a conventional deep

mixed (DM) column. This technique involves inserting plain concrete, reinforced concrete, or a steel

pile into the center of the DM column after the DM column is constructed. In this paper, a series of

centrifugal modeling tests were conducted to investigate the performance of an SDM

column-supported embankment over soft clay. A model embankment supported only by DM columns

was constructed for comparison. Two ideal numerical models of column-reinforced soil under equal

stress and equal strain conditions were established, respectively, to explore the role the column played

in accelerating soil consolidation. A parametric study was conducted to investigate the influence

factors of the length of the core pile, column spacing, thickness of the underlying soil, modulus and

thickness of the cushion, and modulus of the slab on the load transfer of the system, and some

recommendations were proposed for its application. The load transfer mechanism of an SDM

column-supported embankment system with a slab was established based on the development of the

volumetric strains and the principal stresses in the numerical models.


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 2 of 54

Keywords: Centrifugal modeling, stiffened deep mixed column, embankment, soft clay,

consolidation, load transfer


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

INTRODUCTION

With the rapid growth of highways and high-speed railways in China in recent decades, some

transportation lines have to be constructed across regions consisting of thick saturated soft soil. Such

unfavorable geotechnical conditions lead to geotechnical problems such as low bearing capacity,

excessive settlement, and slope instability. Among the techniques used to stabilize the soft soil (e.g.,

jet grouting, stone columns, deep mixed (DM) columns and vertical drains), the DM column has been

a commonly used alternative around the world (Lorenzo and Bergado 2006; Ye et al. 2013). However,

since the increase in soil strength is limited through mixing soil with cement, DM column-reinforced

soft soil cannot sufficiently meet the desired bearing capacity and/or allowable deformation, especially

when heavy loads and strict deformation control are involved.

To solve the limitations of DM columns in application, an innovative technology called the

stiffened deep mixed (SDM) column was proposed (Dong et al. 2002). An SDM column is formed by

installing plain concrete, reinforced concrete, or a steel pile into the center of the DM column

immediately after the DM column is constructed (Zheng et al. 2009). Installing a core pile into a DM

column significantly increases the bearing capacity and reduces the deformation of soft clay owing to

the high strength and stiffness of the core pile (Tanchaisawat et al. 2008; Jamsawang et al. 2011;

Voottipruex et al. 2011a; Raongjant and Jing 2013; Wonglert and Jongpradist 2015). Voottipruex et

al. (2011a) and Wang et al. (2014) stated that the bearing capacity of a single SDM column was 3 to

3.6 times that of a conventional DM column with the same column geometry. Jamsawang et al. (2011)

further indicated that the length of a concrete core pile significantly affected the axial ultimate bearing

capacity of an SDM column, while the cross-sectional area of a concrete core column significantly

affected its lateral ultimate bearing capacity. As compared to the conventional DM column, an SDM

column has high bending stiffness, and it can reduce the lateral deformation of soft soil (Voottipruex
Page 3 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

et al. 2011b; Jamsawang et al. 2011). Voottipruex et al. (2011b) and Raongjant and Jing (2013)

indicated that the lateral bearing capacity of an SDM column was 11 to 15 times and 3 to 4 times

greater than that of a DM column under a static load and cyclic load, respectively.

In a system of SDM column-reinforced soft soil, the external load is mostly concentrated onto the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

core pile head, and then the axial load carried by the core pile is transferred to the soil-cement column

socket through the skin friction along the pile shaft. The additional stress is spread to the surrounding

soil. The gradual reduction in stiffness from the core pile and DM column to the soft soil becomes a

reasonable load transfer system to effectively play the bearing performances of the individual

components. Tanchaisawat et al. (2008) investigated the interface behavior between a DM column and

a concrete core pile, and indicated that the adhesion intercept was sensitive to the cement content

compared to the interfacial friction angle. Voottipruex et al. (2011a) and Wonglert and Jongpradist

(2015) observed three possible failure modes for an SDM column when it was subjected to static axial

load: soil failure, pile failure at the core tip, and pile failure in the DM socket at the top part of the

pile. Ye et al. (2017) investigated the load transfer effect of SDM column-reinforced ground under an

embankment using three-dimensional finite element analysis.

Owing to the large strength and stiffness of an SDM column, the spacing of SDM columns can be

enlarged to make a single column carry a higher embankment load. The commonly used area

replacement ratio of an SDM column is in the range of 5 to 10% in practice, which is significantly

smaller than those used for conventional DM columns (typically in a range of 15 to 30%). The large

spacing of SDM columns results in a large differential settlement between the column and soil. Such a

large differential settlement is reflected on the embankment surface when the embankment height is

not high enough. Therefore, an adequate embankment height is required in design. (For example,

BS8006 (2010) requires that the embankment height is greater than 1.4 times the clearing spacing of

the column.) In a column-supported embankment, geosynthetic reinforcement is usually involved to

promote the load transfer onto the pile and reduce the differential settlement on the embankment base.

However, the required embankment height is still needed to meet the critical height for a complete soil

arch. In some regions such as the east costal area in China, the resources of high-quality backfilling
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 4 of 54

material (i.e., crushed stone or gravel) are scarce and must be transported from other places, resulting

in an increase in the construction cost. To effectively increase the load carried by the SDM column

and reduce the embankment height, a combination of SDM columns with concrete slabs is proposed to

support embankments over soft clay (see Fig. 1). A gravel cushion (typically 20-50 cm in thickness) is
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

placed above the SDM column head, followed by placing a slab on top of the gravel cushion. Only a

relatively thin embankment is necessary for embankment construction. The concrete slab is used to

transfer more embankment load onto the SDM column head, and the gravel cushion acts to adjust the

stress distribution between the SDM column and the surrounding soil. This embankment system has

been used to support the embankments of high-speed railways in China (for example, Qu-Ning

high-speed railway in Fujian province). However, the behavior of this new system of SDM

column-supported embankments with a slab over soft clay has not been well investigated so far.

This paper presents a series of centrifugal modeling tests of SDM column-supported embankments

over soft clay. A model embankment was constructed in a strongbox, and mini-earth pressure cells,

displacement transducers, strain gauges, and mini-piezometers were installed to monitor the

performance of the SDM column-supported embankment. For comparison, a model test of the DM

column-supported embankment was also performed. A parametric study based on a unit cell concept

model was conducted to investigate the influence factors of the lengths of the DM columns and

concrete core piles, slabs, and bearing stratum conditions (i.e., end bearing column and floating

column) on the load transfer behavior. The load transfer mechanism of the embankment system with a

slab was discussed and established based on the development of the volumetric strains and the

principal stress vectors in the numerical models.

CENTRIFUGE APPARATUS

The centrifugal model tests were conducted in the Geotechnical Centrifuge TLJ-150 at Tongji

University (see Fig. 2). The centrifuge had a radius of 3 m, maximum acceleration of 200 g, and

maximum capacity of 150 g-tons. The details of the centrifuge were introduced by Ye et al. (2015).

The model tests were conducted in a strongbox with inside dimensions of 800 mm (length) × 500 mm
Page 5 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

(width) × 500 mm (height). The strongbox was made of stainless steel plates.

MODEL MANUFACTURING

Model ground preparation


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

To minimize the side effects owing to the friction of the steel side walls of the box, two layers of

polytetrafluoroethene (PTFE) membrane were placed on the inside of the steel sidewalls. One layer of

PTFE membrane was pasted on the inside of the steel sidewall, and the other layer was directly placed

on the first layer so that it was allowed to move freely along with the model foundation. Four drainage

pipes were installed at the four corners of the strongbox to collect the water drained from the soil

during the flight.

The third stratum in Shanghai, called a soft silty clay layer, was sampled from the field to

manufacture the model ground. Two subsoil conditions were manufactured for the model ground: (a)

subsoil consisting of a 300-mm-thick soft soil layer and (b) subsoil consisting of a 200-mm-thick soft

soil layer, underlain by a 100-mm sand layer. The model ground was prepared using the following

procedures. For the model ground with a 300-mm-thick soft soil layer, the soft silty clay was air-dried,

coarsely ground, and sieved to pass a 2-mm screen. The soil powder was mixed with water to

manufacture uniform slurry with a water content of 200 % of its liquid limit, and then the clay slurry

was deaired for 48 h in the strongbox. The model ground was preconsolidated under the self-weight at

a centrifuge acceleration of 50 g for about 18,000 s. For the model ground with two soil layers, a fine

sand layer with a thickness of 100 mm was placed at the bottom of the strongbox. The preparation of

the upper soft soil layer was the same as that introduced previously.

To assess the consolidation effect under the self-weight, mini piezometers were installed at three

depths (150, 250, and 350 mm from the ground surface) to monitor the dissipation of excess pore

water pressure during preconsolidation. Fig. 3 shows the variation of the excess pore water pressures

with time during the flight. It can be seen that the excess pore water pressures at the three depths were

almost dissipated after 18,000 s since the ratios of the excess pore water pressure at the end of

consolidation to the initial excess pore water pressure ( ut u0 ) were small (4.3%, 7.1%, and 18.8%
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 6 of 54

for the three depths, respectively). After completion of the model ground, laboratory tests were

conducted to test the soil properties. Table 1 tabulates the main properties of the materials used in the

centrifugal tests. It can be seen that the model ground was almost identical in each test.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Model column preparation

The DM column was manufactured using a mixture of soft silty clay and Portland cement. The

core pile was modeled by an aluminum tube with an outer diameter of 8 mm and an inner diameter of

6 mm. The soil-cement slurry was filled into a hollow polyvinyl chloride (PVC) tube with an internal

diameter of 16 mm. Before the soil-cement was hardened, the aluminum tube was inserted into the

center of the soil-cement column. The ratio of the cross-sectional area of the core pile to that of the

SDM column was 0.25, which approximately equals the average value of the typical ranges of 0.04 to

0.4 used in the field (Zhao et al. 2010). After curing for 28 days, the model SDM column was

extracted from the PVC tube. The unconfined compression strengths of the soil-cement samples at 28

days varied from 0.97 to 1.15 MPa with an average of 1.10 MPa. The unconfined compressive

strength of the DM column obtained from the field typically ranged from 0.5 to 1.5 MPa (Voottipruex

et al. 2011b; Zhang et al. 2014). The aluminum tube had a compressive rigidity of 1.52 × 106 N, which

followed the compressive rigidity in a manner similar to the prototype concrete core pile.

After the self-weight consolidation of the model foundation soil, the centrifuge was stopped to

produce the SDM column-reinforced ground. The model ground was trimmed to 300 mm in thickness.

A plastic plate with prepunched apertures was placed on top of the model ground to precisely locate

the column positions. Apertures with a diameter of 18 mm were arranged in a square grid pattern at a

spacing of 58 mm. A thin cylindrical wall tube with an outer diameter of 18 mm was vertically pushed

into the soil through the apertures. The soil inside the tube was carefully removed, and a model SDM

column was inserted after removing the tube. This procedure was repeated until all of the model

columns were installed in the soil. After that, the improved model ground was brought up again to the

centrifuge at an acceleration of 50 g to minimize the soil disturbance effect and mitigate the gaps

between the columns and the subsoil.


Page 7 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Model embankment preparation

After completion of the ground improvement work, the model embankment was placed on the

model foundation. The model embankment had a base width of 340 mm, a slope of 1.5H:1V, and a
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

height of 80 mm. The embankment consisted of a gravel cushion, and a slab and embankment fill

from bottom to top. The model gravel cushion was made of medium sand. The model slab was made

of a high-density polyethylene (HDPE) plate with a thickness of 8 mm. The use of the HDPE plate

followed the flexural rigidity in a manner similar to the prototype concrete slab. The model

embankment fill was made of fine sand with a side slope of 1.5H:1V. Fig. 4 shows the grain size

distribution curves of the adopted sand. The uniformity coefficients (Cu) and the coefficients of

curvature (Cc) of the sand used to model the gravel cushion and the embankment fill were 3.80 and

1.64, and 2.60 and 1.06, respectively.

MODEL TEST PROGRAMS

Table 2 shows the model test program. Six model tests were conducted to investigate the influence

factors of column length (i.e., DM column length and core pile length), slab in the embankment, and

bearing stratum conditions (i.e., end bearing or floating column) on the performance of the SDM

column-supported embankment over soft clay. For comparison purposes, the ground improved only

by DM columns was included as well. For DM column-supported embankments, geosynthetic

reinforcement is commonly included in the gravel cushion. However, to make a direct comparison of

the performances between the DM column with a slab and the SDM with a slab, the inclusion of

geosynthetic reinforcement in the gravel cushion is not considered in the centrifugal test.

Both the SDM columns and the DM columns had a diameter of 16 mm and were installed in a

square pattern at a spacing of 58 mm. The area replacement ratio of the column (i.e., SDM column or

DM column), which is defined as a ratio of the cross-sectional area of a single column to its

corresponding influence zone, was 6% for all of the centrifugal model tests. In the following

discussion, the model tests are denoted as F/E-SDM/DM-S/NS(LX/X), in which F/E represents the
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 8 of 54

end-bearing column or the floating column, SDM/DM represents the SDM column or the DM column,

S/NS represents the case with a slab or without a slab, and LX/X represents the length of the core pile

(former X) and DM column (latter X). For instance, F-SDM-S(L160/200) denotes a model with a

floating SDM column with a 160-mm-long core pile, 200-mm-long DM column, and a slab on the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

embankment base.

Fig. 5 shows a schematic of the model tests. The monitoring instruments were all installed near the

central area of the model embankment. Mini-earth pressure cells were placed at midspan in the

surrounding soil. A displacement transducer was installed on the embankment base to monitor the

ground settlement. Mini piezometers were installed at depths of 50, 150, and 250 mm from the ground

surface.

After completion of the model SDM column-supported embankment, the strongbox was installed

in the centrifuge. The embankment construction was simulated by increasing the centrifugal

acceleration linearly to 50 g within 800 s to simulate a fast rate of embankment construction

(equivalent to 23 days in the prototype). Then, a model test was run in a 50-g centrifugal acceleration

field for 50,000 s to simulate a four-year soil consolidation under the embankment load in the

prototype. Since the vertical and lateral bearing capacities of the SDM column and ground

deformation were significantly improved as compared to the conventional DM column (Voottipruex et

al. 2011; Jamsawang et al. 2011), the possibility of instability in SDM column-supported

embankments was significantly mitigated. Therefore, the construction time can be reduced by a fast

rate of embankment construction. This is another advantage of the application of SDM columns.

TEST RESULTS AND ANALYSES

Settlement and differential settlement

Fig. 6 presents the ground settlements on the centerline of the embankment in the process of

embankment construction and consolidation. Fig. 6(a) shows that the SDM column significantly

reduced the ground settlement as compared to the DM column (i.e., at the end of the test, the ground

settlements were 14.2 mm and 21.6 mm for the model tests denoted as F-SDM-S(L160/200) and
Page 9 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

F-DM-S(L200), respectively). This can be explained by the fact that the SDM column had a modulus

that was much greater than that of the DM column (the concrete core pile commonly had a modulus

greater than 10 GPa, and the DM column had a modulus typically in a range of 30 to 150 MPa). Since

the area replacement ratio was only 6%, the composite modulus of the SDM column-reinforced zone
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

was greater than that of DM column-reinforced zone. Thus, the model test with a DM column had a

larger ground settlement than the model test with an SDM column. Fig. 6(b) shows that increasing the

length of the DM column can more effectively reduce the ground settlement as compared with

increasing the length of the core pile. At the end of testing, the model tests denoted as F-DM-S(L200),

F-SDM-S(L160/200), and F-SDM-S(L120/200) had ground settlements of 21.6, 14.2, and 15.0 mm,

respectively. In the floating-type-column condition, the SDM model tests denoted as

F-SDM-S(L160/200) and F-SDM-S(L120/200) had ground settlements that were 34.3% and 30.6%

less than that of the DM model test (F-DM-S(L200)). Voottipruex et al. (2011a) stated that

end-bearing SDM columns reduced the settlement by 40% as compared to the DM columns based on

a full-scale embankment test.

Fig. 6(c) shows that the floating-type SDM column-reinforced soil with and without a slab had

settlements of 14.2 mm and 20.1 mm, respectively. The inclusion of a slab significantly reduced the

soil compression within the column length, as more load was concentrated on the stiffer inclusion. On

the other hand, the inclusion of a slab in the embankment fill significantly reduced the differential

settlement between the column and the surrounding soil (see Fig. 7). Fig. 6(d) shows that the

end-bearing-type SDM column significantly mitigated the ground settlement as compared to the

floating-type SDM column. In the end-bearing case, most loads were concentrated onto the SDM

column head through the slab and were transmitted to the underlying firm soil through the pile shaft.

Owing to the large modulus of the SDM column, the compression of soil within the column length

was low. As a result, only 2.3 mm of total settlement was observed in the centrifugal test of the

end-bearing case.

It is interesting to note that in Fig. 6(a) and 6(c), the settlements in the model test of

F-SDM-S(L160/200) were greater than those of F-DM-S(L200) and F-SDM-NS(L160/200) in the


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 10 of 54

initial stage. In the process of column installation, the model column was inserted into the hole after

the soil inside the hole was removed. The large initial settlement could result from the fact that the

hole cleaning was not conducted very well and some disturbed soil remained at the bottom of the hole.

With the increase of the centrifugal acceleration up to 50 g within 800 s, the column with a certain
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

thickness of disturbed soil under the column tip experiences a quick settlement, resulting in a large

settlement occurring in the initial stage under the embankment load.

Stress on surrounding soil

Fig. 8 presents the variation of the earth pressures at the midspan surrounding soil between four

columns in the process of consolidation. In all of the model tests, the earth pressures at the midspan

surrounding soil decreased with time, which indicates that with a process of soil consolidation, more

embankment load was transferred onto the columns. This stress transfer process from soil to columns

is called the “stress concentration.” Han and Ye (2001) theoretically demonstrated this load transfer

mechanism. Ye et al. (2012) and Huang and Han (2009) also found the additional total stress in the

soil decreased and the additional total stress in the column increased during the consolidation process.

The earth pressure at the midspan surrounding soil in the model test E-SDM-S(L160/200) was

smallest among all of the model tests and decreased rapidly to a constant value. In the model test with

a floating-type column, the earth pressures gradually reduced to a constant value. As compared with

the DM column model, the earth pressure at the midspan surrounding soil in the SDM column model

(i.e., F-SDM-S(L160/200)) significantly decreased since the SDM column had higher stiffness.

However, the SDM column model without a slab (i.e., F-SDM-NS(L160/200)) had higher earth

pressure than the SDM column model with a slab (i.e., F-SDM-S(L160/200)). Therefore, the

installation of a rigid slab on the top of the embankment base helped transfer more embankment load

onto the columns.

To further investigate the load transfer effect, the load transfer between the columns and the

surrounding soil was evaluated by the efficacy of the load transfer, which is defined as a proportion of

the total vertical load carried by the columns. Low et al. (1994) determined that the efficacy can be
Page 11 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

expressed as

  
Efficacy  1  s 1  m    100% (1)
  hm 
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

where  s is the earth pressure on the surrounding soil,  is the unit weight of the embankment fill,

hm is the height of the embankment, and m is the area replacement ratio of the column. Fig. 9

presents the variation of the load transfer efficacy with time. With the elapse of time, the efficacy of

the load transfer gradually increased to a constant value. The model of the end-bearing SDM column

(i.e., E-SDM-S(L160/200)) rapidly increased to a constant value and achieved the maximum final

efficacy of the load transfer (i.e, 89.3%). A comparison between model tests F-SDM-S(L160/200) and

F-SDM-NS(L160/200) shows that the inclusion of a slab significantly increased the efficacy of the

load transfer from 33.2% to 63.9%. Chevalier et al. (2010) investigated the load transfer in a

pile-supported thin-fill platform and indicated that when a rigid slab was applied, the load transfer was

much greater than that without a slab. A comparison between F-DM-S(L160/200) and

F-SDM-S(L160/200) shows that a greater difference in stiffness between the column and the

surrounding soil resulted in a greater amount of load transferred to the columns.

The load transfer mechanism of the SDM column-supported embankment system without a slab is

a result of the soil arching effect, which was termed by Terzaghi (1943). The differential settlement

between the soil and the column is restrained by the shear stress developing in the embankment fill,

which reduces the load acting on the surrounding soil and increases the load acting on the column. Ye

et al. (2017) proposed a modified method for the use of EBGEO (2011) to calculate the load transfer

efficacy of SDM column-supported embankments over soft soil. It was expressed as follows:

Efficacy  f ( c )  Efficacy0  (2a)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 12 of 54

0.882  0.039 ln( c ) hm  1.5( s  d c )



f ( c )   [0.118  0.039 ln( c )]  hm (2b)
1  1.5( s  d c )
hm  1.5( s  d c )

where Efficacy0 is the original efficacy of the load transfer calculated by EBGEO (2011), f ( c ) is
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

the modified factor, αc is the area ratio of the concrete core pile, s is the spacing of the SDM column,

dc is the diameter of the core pile, and hm is the embankment height. In the centrifugal modeling test of

F-SDM-NS(L160/200), the ratio of the embankment height to the clearing spacing between the core

piles was 1.6, which was greater than 1.5. Thus, the modified factor was 0.828 based on Eq. (2b). The

original efficacy of the load transfer by EBGEO (2011) was 44%. Thus, the calculated load transfer

efficacy was 36% after multiplying by the modified factor of 0.828, which agreed with the modeling

test result of 33%. The above analysis demonstrates that the modified method proposed by Ye et al.

(2017) is feasible for the calculation of the load transfer efficacy of SDM column-supported

embankments over soft soil.

Excess pore water pressure

Fig. 10 presents the variations of the measured excess pore water pressures with time at depths of

50, 150, and 250 mm along the centerline of the embankment. In all of the tests, the excess pore water

pressures increased in the process of increasing the centrifugal acceleration until the centrifugal

acceleration was 50 g, and then they dissipated gradually with time. Fig. 10(a) shows that the

maximum excess pore water pressure in the model test with DM columns (i.e., F-DM-S(L200)) was

higher than that in the model test with SDM columns (i.e., F-SDM-S(L160/200)) at depths of 50 mm

and 150 mm. This can be explained by the fact that as the SDM column had a larger modulus than the

DM column, a greater embankment load was carried by the SDM columns as compared with the

model test with a DM column. The reduction of the embankment load carried by the surrounding soil

resulted in a reduction of the excess pore water pressure in the soil. Han and Ye (2001) and Ye et al.

(2012) found this phenomenon when investigating the consolidation characteristics of soft soil with
Page 13 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

stone columns or DM columns. However, below the column base, the increment of excess pore water

pressure in the model test of SDM columns was slightly larger than that in the model test of a DM

column. Since the overall stiffness of the SDM column-reinforced area was higher than that of the DM

column-reinforced area, additional stress was concentrated in the SDM column-reinforced area and
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

then transferred to the deep soil.

Fig. 10(b) shows the dissipation of the excess pore water pressure at a depth of 50 mm. In the

model tests denoted as F-SDM-NS(L160/200) and F-SDM-NS(L120/200), increasing the core pile

from 120 to 160 mm had a minor effect on changing the rate of pore water dissipation. A possible

explanation is that since the length of the DM column and the thickness of the unreinforced soft soil

below the column base were the same in the two tests, the stiffness difference between the SDM

column and the surrounding soil did not experience a significant change after increasing the core pile

from 120 to 160 mm under the floating column condition. As a result, the load carried by the

surrounding soil was similar. However, the maximum excess pore water pressure decreased by 5.8

kPa after the length of the DM column was increased from 200 to 240 mm by comparing the model

tests of F-SDM-NS(L160/200) and F-SDM-NS(L160/240). When the SDM column was seated on

firm soil (i.e., the model denoted as E-SDM-S(L160/200)), the increment of excess pore water

pressure was small. This confirms that increasing the improvement depth in a thick soft soil stratum

has a stronger effect on reducing the maximum excess pore water pressure in the soil between the

piles than increasing the length of the core pile but maintaining the total length of the SDM column.

Fig. 10(c) shows the influence of the slab on the excess pore water pressure in the soil when it is

subjected to the embankment load. As compared with the model without a slab, the maximum

increments of excess pore water pressure in the soil decreased by 7.5 kPa and 5.9 kPa at depths of 50

mm and 250 mm, respectively, when the slab was installed. It can be speculated that the inclusion of a

slab enhanced the load transfer effect onto the SDM column, resulting in a low stress carried by the

soil between the columns. The load transfer mechanism will be discussed in a later section.

CONSOLIDATION CHARACTERISTICS OF SYSTEM WITH A SLAB


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 14 of 54

Back analysis of consolidation parameter

It is assumed that the soil consolidation at different depths in the model tests followed the

one-dimensional consolidation theory under one-way drainage conditions by Terzaghi (1943). The

excess pore pressure u  z , t  at depth z at any time t can be expressed as


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18


2u0  Mz   M 2Tv
u  z , t    sin  e , M   2m  1  2, m  0, 1, 2... (3)
m  0 Mπ  Hd 

where Tv  Cv t H d2 is the time factor for vertical drainage, Cv is the vertical coefficient of

consolidation, u0 is the initial excess pore water pressure, and H d is the drainage path. Bromwell

and Lambe (1968) proposed an incremental time method to back calculate the coefficient of

consolidation of soil based on the measured pore water pressure in the field. Based on Eq. (3), the

ratio of the excess pore pressure at two consecutive elapsed times can be expressed as

π 2 Cv
ui  t t 
2 i i 1
 Fv e 4 H d (4a)
ui 1

 π2 
2u0  Mz   M  4 Tvi
2


m0 M
sin  e
Fv   Hd  (4b)
 2 π2 

2u0  Mz   M  4 Tvi1

m0 M
sin  e
 Hd 

where ui and ui -1 are the excess pore water pressures, and Tvi and Tvi-1 are the time factors at

time ti and ti 1 , respectively. Cao et al. (2001) demonstrated that the value of Fv converges to unity

for Tv  0.2 . As a result, the preceding equation can be simplified as

Cv π 2  ti  ti 1 

4 H d2
ui  e ui 1 (5)
Page 15 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Thus, the measured excess pore water pressures at different depths during a later preloading period

(e.g., a degree of consolidation greater than 60%) were divided into a number of data items at

successive equal time intervals ( t =800 c. in this study), and then the excess pore water pressure for
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

the current time step ui was plotted against the excess pore water pressure for the previous time

step ui -1 . We draw a best-fit line from the origin through these data points, and the vertical

coefficient of consolidation is determined based on the slope of the best-fit line denoted as β. Cv can

be back calculated using the following equation:

4 H d2
Cv   ln  (6)
t 2

As the soil consolidation at different depths was assumed to be under one-dimensional consolidation

with one-way drainage conditions, the drainage path ( H d ) is the thickness of the model foundation.

Fig. 11 shows the measured excess pore water pressure and the best-fit equations for ui vs. ui -1 .

Fig. 12 presents the back-calculated vertical coefficients of consolidation based on the measured

pore water pressures. It is worth noting that the back-calculated Cv represents the value at a certain

location, which varied in a narrow range from 0.015 to 0.023 cm2/s with an average of 0.0188 cm2/s.

The soft soil in the centrifugal test was relatively homogeneous. Since the back-calculated values of

Cv were almost independent for depth, the back-analysis results validate the feasibility of the

hypothesis of the one-dimensional consolidation of SDM column-reinforced soil under embankments.

Similarly, based on the measured excess pore water pressures in the model ground during

preconsolidation under the self-weight, the back-calculated values of Cv were from 0.020 to 0.027

cm2/s. The coefficients of consolidation of the model foundation experienced a small reduction owing

to the soil compression.

Consolidation characteristics of column-reinforced soil


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 16 of 54

To further investigate the effect of columns on soil consolidation, three ideal numerical models

based on the axisymmetric unit cell concept were established using the software FLAC2D 5.0 (see

Fig. 13): (a) unreinforced soil deforms under an equal stress-flexible loading, (b) column-reinforced

soil deforms under an equal stress-flexible loading, and (c) column-reinforced soil deforms under an
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

equal strain-rigid loading. The DM column was selected for the analysis and was modeled using the

Mohr-Coulomb model. The soft soil was modeled using a modified Cam-Clay model. Most of the

properties of the soft soil and column were the same as those used in the centrifugal test, and the

parameters of the modified Cam-Clay model referred to a study on the remolded third stratum in

Shanghai by Sheng (2012). The elastic modulus of the DM column was estimated based on the typical

relationship of E = 100qu, where qu = unconfined compressive strength (1.1 MPa in this study).

Poisson’s ratios of the soils and DM columns were assumed based on typical values. The permeability

of the DM column was assumed as 10 times less than the soil. Table 3 presents the parameter

properties used in the numerical analysis.

Under flexible loading, a uniform pressure was directly applied on the ground surface. Under the

rigid loading condition, a uniform pressure was exerted on the soil through a rigid plate without mass

to maintain the equal strain condition. A frictionless contact was created in the interface between the

rigid plate and the ground. In all of the models, a uniform pressure of 30 kPa was applied

instantaneously, followed by a long consolidation for 360 days. The boundary conditions of the model

were set as follows: the bottom boundary was fixed in both the horizontal and vertical directions, and

the two side boundaries were fixed in the horizontal direction but free in the vertical direction. The

bottom and side boundaries were defined as impermeable, and a free drainage boundary was defined

at the ground surface.

Fig. 14 shows the variations of the excess pore water pressure at the monitored locations (see Fig.

13) with the elapse of time in the three models. It can be seen that the excess pore water pressures in

the soil dissipated faster when a column was installed, especially in the case under the equal strain

condition. The soil properties related to soil consolidation (i.e., modulus and permeability) were kept

constant in the numerical models. The drainage paths for the monitored soil elements were the vertical
Page 17 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

distance from the monitored locations to the ground surface, as the three cases were almost under a

one-dimensional consolidation condition. At the end of preloading for 360 days, the vertical distances

had small reductions (i.e, 7.8 mm, 4.8 mm, and 3.4 mm) in the unreinforced soil model, the

column-reinforced soil model under flexible loading, and the column-reinforced soil model under
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

rigid loading, respectively. Therefore, the faster dissipation of the pore water pressure in the soil is not

attributed to the changes in the soil consolidation properties and the drainage path.

Fig. 15 shows the variations of the additional total vertical stress at the monitored locations in the

soil and in the column with time. It can be seen that the total vertical stress applied to the soil element

was equal to the applied external pressure in the unreinforced soil model. However, when a column

was included, the total vertical stress applied to the soil element decreased with time, while the total

vertical stress applied to the column element increased with time. This load transfer phenomenon is

more significant under the equal strain condition. In the column-reinforced soil, the shear stress was

generated in the soil owing to the uneven deformation between the soil and the column under the

equal stress condition. The shear stress transferred the additional stress in the soil onto the stiffer

column. Under the equal strain condition, owing to the great difference in stiffness between the

column and the soil, the applied load was concentrated onto the column through the rigid plate during

the soil consolidation. The above two load transfer mechanisms result in a reduction in the applied

additional total stress in the soil, which helps accelerate the dissipation of the excess pore water

pressure in the soil when a column is included. Based on the numerical analyses, the equal strain

condition had more efficiency in load transfer than the equal stress condition. Therefore, the inclusion

of a slab over the column-reinforced soil has a function to accelerate the soil consolidation.

PARAMETRIC STUDY ON LOAD TRANSFER BEHAVIOR

Numerical modeling validation

As presented in the centrifugal test, the inclusion of a slab significantly increased the load transfer

efficacy of the SDM column as compared with that without a slab. Owing to the complexity of the

SDM column-supported embankment system, many factors such as the gravel cushion, core pile, and
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 18 of 54

bearing stratum conditions (i.e., end bearing column and floating column) also affect the load transfer

behavior of the system. A parametric study based on the unit cell concept was performed in this

section to investigate the load transfer behavior of the SDM column-supported embankment system.

The load transfer efficacy of the column was used to evaluate the degree of load transfer.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

To validate the feasibility of the unit cell model, the results from the unit cell of the prototype

model were compared with the test results of the corresponding centrifugal model. Considering the

acceleration of 50 g used in the centrifuge test, a scale factor of 50 was applied to establish the unit

cell model. The geometries of the prototype model were equal to 50 times those of the centrifuge

model. The diameter of the unit cell model was equal to the diameter of the influence zone of a single

SDM column, which was 1.13 times the column spacing for a square column pattern. The baseline

model was established according to the centrifugal test denoted as F-SDM-S(L160/200), as shown in

Fig. 16. Other models were modified based on the baseline model (see Table 4).

The core pile and the slab were simplified as elastic materials. The DM column, gravel cushion,

and embankment fill were modeled as linearly elastic-perfectly plastic materials with the

Mohr-Coulomb failure criteria. The soft soil was modeled using a modified Cam-Clay model. For

granular materials (i.e., the gravel cushion and the embankment fill), the dilation behavior was

considered based on the empirical correlation (i.e., ψ = φ-30º, in which ψ is the dilation angle and φ is

the friction angle of the soil). Most of the properties used in the numerical models are the same as

those used in the centrifugal test. The compressive rigidity of the core pile and the flexural rigidity of

the slab of the unit cell model followed the similarity theory to that in the corresponding centrifugal

test. Table 5 lists the properties used in the baseline numerical model.

An interfacial model with the Mohr-Coulomb failure criterion was created on the interface

between the core pile shaft and the DM column. The frictional angle was 25º, and the adhesion was 50

kPa, which approximates the average values reported by Tanchaisawat et al. (2008). The values of the

interfacial normal and shear stiffness were estimated by the following formula (Itasca Consulting

Group Inc. 2006):


Page 19 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

10  K  1.33G 
k n  ks  (6)
z

where kn and ks = interface normal and shear stiffness, respectively; K and G = bulk and shear moduli,

respectively; and Δz = smallest width of an adjoining zone near the interface. The bottom boundary
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

was fixed in both the radial and vertical directions. The side boundaries were fixed in the radial

direction but were allowed to move freely in the vertical direction.

Table 6 lists the load transfer efficacies obtained from the unit cells of the prototype models and

their corresponding results by the centrifugal models at the end of testing. It can be seen that the

numerical results of the prototype models and the centrifugal modeling agreed well. For the end

bearing condition (i.e., E-SDM-S(L160/200)), the numerical result underestimated the load transfer

efficacy. The gravel layer might have a large deformation in the end bearing condition as the gravel

material is likely to be squeezed to the area above the surrounding soil owing to the upward

penetration of the column. This was unable to be simulated well using the Mohr-Coulomb failure

behavior. In general, however, the comparisons demonstrate the reliability of the unit cell of the

prototype model.

Length of core pile

Fig. 17(a) shows the influence of the core pile length on the load transfer efficacy, in which the

length of the core pile changes to 0, 1, 5, 8, and 10 m. A length of the core pile equal to 0 m represents

the soil improved by the conventional DM column, and a length of the core pile equal to 10 m means

the core pile and the DM column have equal lengths. The SDM column had an obviously higher load

transfer efficacy than the DM column. The load transfer efficacy increased with an increase in the

length of the core pile, and gradually achieved a stable value after the length ratio (i.e., the ratio of the

core pile length to the DM column length) was greater than 0.5. From a practical point of view,

blindly increasing the length of the core pile cannot effectively exploit the load transfer performance

of an SDM column. Based on the numerical results, a length ratio from 0.5 to 0.8 is recommended in

terms of enhancing the load transfer efficacy.


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 20 of 54

Column spacing

Fig. 17(b) shows the influence of column spacing on the load transfer efficacy. The column

spacing in the baseline model was changed to 2.6, 3.3, and 4 m, which corresponds to area
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

replacement ratios of 0.04, 0.06, and 0.09, respectively. The load transfer efficacy increased with an

increase in the area replacement ratio, as expected. When the area replacement ratio increased from

0.06 to 0.09, the improvement in load transfer efficacy was not that significant as compared with the

results with an area replacement ratio from 0.04 to 0.06.

Thickness of underlying soft soil

Fig. 17(c) shows the influence of the thickness of the underlying soft soil on the load transfer

efficacy. A thickness of the underlying soft soil equal to 0 m means the SDM column was seated on a

firm soil layer (for example, bedrock). The load transfer efficacy had a significant decrease with an

increase in the thickness of the underlying soft soil. Caution should be taken when using a

floating-type SDM column in a thick soft soil layer owing to the relatively low load transfer efficacy.

Modulus of slab

Fig. 17(d) shows the influence of the slab modulus on the load transfer efficacy. A model without

a slab is also included for comparison. It can be seen that when a slab was involved in the SDM

column-supported embankment system, the load transfer efficacy had a significant enhancement as

compared to the model without a slab. When the modulus of the slab increased from 1 GPa and 10

GPa to 30 GPa, the load transfer efficacy had a slight increase owing to the already significantly high

stiffness of the slab as compared with the soft clay in the system. From a practical point of view, the

commonly used concrete grade C30 for the slabs (i.e, elastic modulus is 30 GPa) has sufficient

stiffness to enhance the load transfer effect.

Thickness and modulus of gravel cushion


Page 21 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

The gravel cushion acts to adjust the stress distribution between the SDM column and the

surrounding soil. Fig. 17(e)(f) show that the load transfer efficacy decreased with an increase in the

cushion thickness and a decrease in the cushion modulus. Therefore, it is important to design a

cushion with appropriate thickness and modulus to give full play to the bearing performances of the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

SDM column and the surrounding soil.

LOAD TRANSFER MECHANISM OF SYSTEM WITH A SLAB

It is interesting to note from the centrifugal tests that when a rigid slab was involved, the load

transfer efficacy of the SDM column was significantly almost twice that of the SDM column without a

slab (i.e., 33.2% in F-SDM-NS(L160/200) and 63.9% in F-SDM-S(L160/200)). The significantly

higher load transfer efficacy might be owing to the fact that involving a slab in the SDM

column-supported embankment system had a different load transfer mechanism as compared to that

without a slab. To investigate the slab’s role in the load transfer of the SDM column-supported

embankment, the above unit cells of the prototype models corresponding to the centrifugal tests

denoted as F-SDM-NS(L160/200), F-SDM-S(L160/200), and E-SDM-S(L160/200) were further

investigated.

Fig. 18 shows the volumetric strains of the models with and without a slab. For better presentation,

the results in the upper portion of the model are examined herein. Owing to the inclusion of a slab, the

relatively large volumetric strains of the gravel cushion were concentrated mostly above the SDM

column head, especially above the core pile head. The strain of the embankment fill above the slab

was equal at elevation as compared to the case without a slab. The relatively stiff slab cuts off the

reflection of the differential settlement between the column and the surrounding soil to the

embankment fill. As a result, the differential settlement in the embankment fill is effectively mitigated

by the slab, and the embankment height can be reduced.

Fig. 19 illustrates the principal stress vectors of the model with a slab. Following the deflections of

the principal stress vectors starting from the edges of the DM column and core pile, more stress was

concentrated onto the SDM column, especially onto the core pile through the slab. The upper surface
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 22 of 54

of the slab above the column and its lower surface above the surrounding soil generated tensile

stresses. Therefore, the slab behaves like a rigid beam that carries the embankment load and transfers

the embankment fill load onto the column head through the gravel cushion. The concept of the stress

distribution angle was used to evaluate the load transfer behavior of the system with a slab. As shown
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

in Fig. 19, α and α’ represent the distribution angles along the edges of the DM column and core pile

in the gravel cushion, and β and β’ represent the distribution angles in the slab, respectively.

The stress distribution angles (i.e., α and α’ in Fig. 19) were determined by integrating the vertical

stresses distributed on a series of interface lengths (see bb’ in Fig. 19) until the integral force on the

interface of bb’ was equal to the total vertical forces on the SDM column and the core pile. Then, a

similar procedure was followed to determine the stress distribution angles β and β’. Involving a slab

above the gravel cushion significantly increases the distribution angle (i.e., β > α and β’ > α’),

resulting in more of the embankment load being transferred on the column through the slab. As

compared with the floating-type column, the end bearing column had a high distribution angle in the

gravel cushion. This explains why the end bearing column had a maximum load transfer efficacy in

the centrifugal modeling.

Based on the preceding analysis, the load transfer mechanism of the column-supported

embankment with a slab can be established as shown in Fig. 20: when a slab is involved, the slab

behaves like a rigid beam that carries the embankment load. Owing to the great difference in the

stiffness between the column and the surrounding soil, most of the embankment load is concentrated

onto the localized area of the slab above the columns and is transferred to the column, especially the

core pile, through the soil wedge between the column and the slab.

Conclusions

This paper presented a series of centrifugal modeling tests and numerical modeling to investigate

the performance of an SDM column-supported embankment over soft clay. The influence factors of

the length of the DM column and concrete core pile, the slab in the embankment, and the bearing

stratum conditions (i.e., end bearing column and floating column) were considered in this study.
Page 23 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Based on analyses and discussion, the following conclusions can be drawn:

1. For the floating-type column, installation of a slab significantly reduced the total and differential

settlement and increased the load carried by the SDM columns as compared to that without a slab.

2. The SDM column-supported embankment with a slab minimized the maximum increment of the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

excess pore water pressure in the surrounding soil under the embankment load. The numerical

analyses demonstrated that the installation of a slab accelerates the soil consolidation as compared

with the unreinforced case and the reinforced case without a slab.

3. Based on numerical analyses, a length ratio in the range of 0.5 to 0.8 is recommended in terms of

the load transfer effect of an SDM column. The commonly used concrete slab with a modulus of

30 GPa has sufficient stiffness to transfer the embankment load to the SDM column. Appropriate

selections of the thickness and modulus of the gravel cushion are necessary to take full advantage

of the bearing capacities of SDM columns and subsoil.

4. The slab behaves as a rigid beam to carry the embankment load. Most of the embankment load is

concentrated onto the localized area of the slab above the SDM columns and is transferred to the

column, especially the core pile, through the soil wedge between the column and the slab.

Acknowledgments

The authors appreciate the financial support provided by the Natural Science Foundation of China

(NSFC) (grant no. 51508408 and no. 41772281) and by the Fundamental Research Funds for the

Central Universities (grant no. 22120180106) for this research.

References

Asaoka, A. 1978. Observational procedure for settlement prediction. Soils and Foundations., 18(4):

87-101.

Bromwell, L.G., and Lambe, T.W. 1968. Comparison of laboratory and field values of Cv for Boston

Blue Clay. Highway Research Record, 243, pp. 23-37.

Carlsson, B. 1987. Reinforced soil, principles for calculation. Terratema AB, Linköping. (In Swedish).
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 24 of 54

Cao, L.F., Chang, M.F., Teh, C.I., and Na, Y.M. 2001. Back-calculation of consolidation parameters

from field measurements at a reclamation site. Canadian geotechnical journal, 38(4): 755-769.

Chevalier, B., Villard, P., and Combe, G. 2010. Investigation of load-transfer mechanisms in

geotechnical earth structures with thin fill platforms reinforced by rigid inclusions. International
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Journal of Geomechanics, 11(3): 239-250.

Dong, P., Chen, Z. Z., and Qin, R. 2002. Use of concrete-cored DCM pile in soft ground. Chinese

Journal of Geotechnical Engineering, 24(2): 204-207 (in Chinese).

EBGEO. 2011. Recommendations for Design and Analysis of Earth Structures using Geosynthetic

Reinforcements–EBGEO, Wiley, Berlin, Germany.

Han, J., and Ye, S.L. 2001. Simplified method of consolidation rate of stone column reinforced

foundation. Journal of Geotechnical and Geoenvironmental Engineering, 127(7): 597-603.

Huang, J., and Han, J. 2009. 3D coupled mechanical and hydraulic modeling of a

geosynthetic-reinforced deep mixed column-supported embankment. Journal of Geotextiles and

Geomembranes, 27(4): 272-280.

Itasca Consulting Group Inc. 2006. FLAC2D User's Guide, Version 5.0.

Jamsawang, P., Bergado, D. T., and Voottipruex, P. 2011. Field behaviour of stiffened deep cement

mixing piles. Proceedings of the Institution of Civil Engineers-Ground Improvement, 164(1):

33-49.

Lorenzo, G.A., and Bergado, D.T. 2006. Fundamental characteristics of cement-admixed clay in deep

mixing. ASCE Journal of Materials in Civil Engineering, 18(2): 161-174.

Low, B. K., Tang, S. K., and Choa, V. 1994. Arching in piled embankments. Journal of Geotechnical

Engineering, 120(11): 1917-1938.

Øiseth, E., Svanø, G., Watn, A., and Emdal, A. 2002. A computer program for designing reinforced

embankments. Proceedings of the 7th international conference on geosynthetics, France, vol. 1, pp.

201-204.

Raongjant, W., and Jing, M. 2013. Field testing of stiffened deep cement mixing piles under lateral

cyclic loading. Earthquake Engineering and Engineering Vibration, 12(2): 261-265.


Page 25 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Tanchaisawat, T., Suriyavanagul, P., and Jamsawang, P. 2008. Stiffened Deep Cement Mixing

(SDCM) pile: Laboratory investigation. In Proceedings of the International Conference on Concrete

Construction, Kingston University, United Kingdom, 9-10 September 2008. CRC Press,

Netherlands, pp. 39-48.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Sheng, J.-R. 2012. Laboratory tests and constitutive modeling on the mechanical behavior of shanghai

clays. Master of Engineering thesis, School of Naval Architecture, Ocean and Civil Engineering

(NAOCE), Shanghai Jiao Tong University, Shanghai.

Terzaghi, K. 1943. Theoretical soil mechanics, Wiley, New York, 66-75.

Voottipruex, P., Bergado, D.T., Suksawat, T., Jamsawang, P., and Cheang, W. 2011a. Behavior and

simulation of deep cement mixing (DCM) and stiffened deep cement mixing (SDCM) piles under

full scale loading. Soils and Foundations, 51(2): 307-320.

Voottipruex, P., Suksawat, T., Bergado, D. T., and Jamsawang, P. 2011b. Numerical simulations and

parametric study of SDCM and DCM piles under full scale axial and lateral loads. Computers and

Geotechnics, 38(3): 318-329.

Wang, C., Xu, Y. F., and Dong, P. 2014. Working characteristics of concrete-cored deep cement

mixing piles under embankments. Journal of Zhejiang University (Science A), 15(6): 419-431

Wonglert, A., and Jongpradist, P. 2015. Impact of reinforced core on performance and failure

behavior of stiffened deep cement mixing piles. Computers and Geotechnics, 69: 93-104.

Ye, G., Zhang, Q., Zhang, Z., and Chang, H. 2015. Centrifugal modeling of a composite foundation

combined with soil–cement columns and prefabricated vertical drains. Soils and Foundations,

55(5): 1259-1269.

Ye, G.B., Cai, Y.S., and Zhang, Z. 2017. Numerical Study on Load Transfer Effect of Stiffened Deep

Mixed Column-supported Embankment over soft soil. KSCE Journal of Civil Engineering, 21(3):

703-714.

Ye, G.B., Zhang, Z., Han, J., Xing, H.F., Huang, M.S., and Xiang, P.L. 2013. Performance Evaluation

of an Embankment on Soft Soil Improved by Deep Mixed Columns and Prefabricated Vertical

Drains. ASCE Journal of Performance of Constructed Facilities, 27(5): 614-623.


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 26 of 54

Zhang, Z., Han, J., and Ye, G.B. 2014. Numerical analysis of failure modes of deep mixed

column-supported embankments on soft soils. Ground Improvement and Geosynthetics: In

Proceedings of the 2014 GeoShanghai International Conference, Shanghai, 24-26 May 2014,

Geotechnical Special Publication, n 238 GSP, pp. 78-87.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Zhao, X., Wu, M., Chen, S., and Kong, D. 2010. Study on bearing behaviors of single axially loaded

SDCM pile. Deep Foundations and Geotechnical In Situ Testing - Proceedings of the 2010

GeoShanghai International Conference, 3-5 June 2010, Geotechnical Special Publication, n 205

GSP, pp. 277-284.

Zheng, G., Liu, S.Y., and Chen, R.P. 2009. State of advancement of column-type reinforcement

element and its application in China. Advances in Ground Improvement: Research to Practice in the

United States and China-Proceedings of the 2009 US-China Workshop on Ground Improvement

Technologies, March 14, 2009, Orlando, Florida, Geotechnical Special Publication, n 188, p 12-25.
Page 27 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure Captions

Fig. 1 Schematic of SDM column-supported embankment over soft clay


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Fig. 2 TLJ-150 Geotechnical Centrifuge apparatus

Fig.3 Variation of excess pore water pressure at depths

Fig. 4 Grain size distribution of the embankment material

Fig. 5 Schematic of model test setup (unit: mm)

Fig. 6 Variation of ground settlement at embankment centerline

Fig. 7 Ground settlement after test: (a) F-SDM-S(L120/200); (b) F-SDM-NS(L120/200)

Fig. 8 Variation of earth pressure at the mid-span surrounding soil between four columns with time

Fig. 9 Variation of column efficacy of load transfer with time

Fig. 10 Variation of excess pore water pressure with time: (a) at depth of 5 cm; (b) at depth of 25 cm

Fig. 11 Variation of ui versus ui-1

Fig. 12 Back-calculated Cv from pore pressures

Fig. 13 Ideal models for comparison of consolidation process (not to scale): (a) unreinforced soil under a

flexible loading; (b) column-reinforced soil under a flexible loading; (c) column-reinforced soil under a

rigid loading.

Fig. 14 Variation of excess pore water pressure in the three models

Fig. 15 Variation of additional total vertical stress with time in soil and in column

Fig. 16 Unit cell of baseline model (Unit: m, not to scale)

Fig. 17 Effect of influence factors on load transfer efficacy

Fig. 18 Volumetric strain of the unit cell models corresponding to (unit: m): (a) F-SDM-NS(L160/200); (b)

F-SDM-S(L160/200)

Fig. 19 Principal stress vectors of the numerical model: (a) F-SDM-S(L160/200); (b) E-

SDM-S(L160/200)

Fig. 20 Load transfer mechanism of SDM column-supported embankment with a slab


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Slab

Soft soil

Firm soil
Gravel cushion
Embankment fill

SDM pile
Core pile
DM column

Fig. 1 Schematic of SDM column-supported embankment over soft clay


Page 28 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 29 of 54

Fig. 2 TLJ-150 Geotechnical Centrifuge apparatus


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Excess pore water pressure, kPa

0
40
80
120
160
200

0.0
0.5
350
250
150
Depth(mm)

1.0
7.1
4.3

18.8

Time, 104s
ᇞut/ᇞu0(%)

1.5
350 mm
250 mm
150 mm

2.0

Fig.3 Variation of excess pore water pressure at depths


Page 30 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 31 of 54

Percent passing, %

0
20
40
60
80
100

110
0.1
Particle size, mm
0.01
Sand cushion
Embankment fill

1E-3

Fig. 4 Grain size distribution of the embankment material


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

 
 
Soft soil

SDM pile

SDM pile

Embankment
Embankment
transducer
Displacement

100
340
800

(a)

(b)
1.5
1
Piezometer

Settlement plate

500 300 80
Earth pressure cell

Fig. 5 Schematic of model test setup (unit: mm)


500
Page 32 of 54
Page 33 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

0 0

8
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

8
Settlement, mm

Settlement, mm
16 16

24 24

32 32 F-SDM-S (L160/200)
F-DM-S (L200) F-SDM-S (L120/200)
F-SDM-S (L160/200) F-SDM-S (L160/240)
40 40
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time, 104s   Time, 104s  

(a) (b)

0 0

8 8
Settlement, mm

Settlement, mm

16 16

24 24

32 32
F-SDM-S (L160/200) F-SDM-S (L160/200)
F-SDM-NS (L160/200) E-SDM-S (L160/200)
40 40
0 1 2 3 4 5 6 0 1 2 3 4 5 6
4
   Time, 10 s   Time, 104s  

(c) (d)

Fig. 6 Variation of ground settlement at embankment centerline

   
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

(a)
 
 
 

(b)

Fig. 7 Groound settlemennt after test: ((a) F-SDM-S((L120/200); ((b) F-SDM-N


NS(L120/200)
 
Page 34 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 35 of 54

Earth pressure, kPa

0
20
40
60
80
100

0
1
2
3
Time, 104s
4
F-DM-S (L200)

5
F-SDM-S (L160/200)
E-SDM-S (L160/200)
F-SDM-NS (L160/200)

Fig. 8 Variation of earth pressure at the mid-span surrounding soil between four columns with time
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

 
Efficacy, %

0
20
40
60
80
100

 
1
2
3
Time, 104s
4
5
F-DM-S (L200)
89%

75%

47%
50%

F-SDM-S (L160/200)
E-SDM-S (L160/200)
F-SDM-NS (L160/200)

Fig. 9 Variation of column efficacy of load transfer with time


Page 36 of 54
Page 37 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

50 50
F-SDM-S (L160/200)

Excess pore water pressure, kPa


F-DM-S (L200) F-SDM-S (L160/200)
Excess pore water pressure, kPa

50 mm 50 mm 50 mm
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

40 150 mm 150 mm 40 F-SDM-S (L120/200)


250 mm 250 mm 50 mm
F-SDM-S (L160/240)
30 30 50 mm
E-SDM-S (L160/200)
50 mm
20 20

10 10

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time, 10 s 4
    Time, 104s

(a) (b)
60
F-SDM-S (L160/200)
Excess pore water pressure, kPa

50 50 mm
250 mm
F-SDM-NS (L160/200)
40 50 mm
250 mm
30

20

10

0
0 1 2 3 4 5 6
Time, 104s

(c)

Fig. 10 Variation of excess pore water pressure with time: (a) at depth of 5 cm; (b) at depth of 25 cm

   
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

 
 

∆ui, kPa

0
5
10
15
20
25

0
5
10

 
∆ui-1, kPa
15
150 mm: ∆ui=0.956∆ui-1, R2=0.9989
50 mm: ∆ui=0.9599∆ui-1, R2=0.9991

20
250 mm: ∆ui=0.9625∆ui-1, R2=0.9997

50 mm

250 mm
150 mm

Fig. 11 Variation of ui versus ui -1


F-SDM-S(L160/200)

25
Page 38 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 39 of 54

 
 

Depth from ground surface, mm

300
250
200
150
100
50
0

0.00
0.02
2
Ave. Cv

0.04

Cv, cm /s
F-DM-S(L200)

0.06
F-SDM-S(L160/240)
F-SDM-S(L120/200)
F-SDM-S(L160/200)

E-SDM-S(L160/200)
F-SDM-NS(L160/200)

Fig. 12 Back-calculated Cv from pore pressures


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 40 of 54

 
 

z z z P
P P
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Rigid plate
2.5 m

2.5 m
2.5 m
0.2 m 0.2 m

a B b C c
1.02 m 1.02 m 1.02 m
8m

Soft soil Column Soft soil Column Soft soil

O O O
r r r

1.64 m 0.4 m
(a) (b) (c)

Fig. 13 Ideal models for comparison of consolidation process (not to scale): (a) unreinforced soil

under a flexible loading; (b) column-reinforced soil under a flexible loading; (c) column-reinforced

soil under a rigid loading.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 41 of 54

Excess pore water pressure, kPa

0
10
20
30
40

0
100
200
Time, Day
300
In soil

c
b
a

400

Fig. 14 Variation of excess pore water pressure in the three models


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Additional total stress in soil, kPa

0
30
60
90

0
100
200
Time, Day
In soil

c
b
a

300
B
C
In column

400
0
30
60
90

-30
120
150

Additional total stress in column, kPa


Fig. 15 Variation of additional total vertical stress with time in soil and in column
Page 42 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 43 of 54

15 4

10
z
O

0.8

3.27
Slab

Soft soil
8
Gravel cushion
r
Embankment fill

SDM
column

Fig. 16 Unit cell of baseline model (Unit: m, not to scale)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 44 of 54

100 100

80 80
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Efficacy, %

Efficacy, %
60 60

40 40
(DM column)

20 20
-2 0 2 4 6 8 10 12 0.02 0.04 0.06 0.08 0.10
Length of core pile, m Area replacement raito

(a) (b)

100 100

90 80
(End bearing column)
Efficacy, %
Efficacy, %

80 60

70 40
(No slab)

60 20

50 0
-1 0 1 2 3 4 5 6 0 10 20 30 40
Thickness of underlying soft soil, m Modulus of slab, GPa

(c) (d)

100 100

80
80
Efficacy, %
Efficacy, %

60

60
40

20 40
100 200 300 400 500 20 40 60 80 100 120
Cushion thickness, mm Modulus of gravel cushion, MPa

(e) (f)

Fig. 17 Effect of influence factors on load transfer efficacy


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 45 of 54

     

(b) F-SDM-S(L160/200
 

Fig. 18 Volumetric strain of the unit cell models corresponding to (unit: m): (a) F-SDM-NS(L160/200);
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

(a)

(b)

E-SDM-S(L160/200)
 

Fig. 19 Principal stress vectors of the numerical model: (a) F-SDM-S(L160/200); (b)
Page 46 of 54
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 47 of 54

 
 
Soil
Slab

Granular
Embankment

β' β

α' α
 fill

SDM Column

Fig. 20 Load transfer mechanism of column-supported embankment


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 48 of 54

Table Captions
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Table 1 Material properties

Table 2 Model test programs

Table 3 Main properties used in the numerical study

Table 4 Influence factors considered in the parametric study

Table 5 Main properties of baseline case in the parametric study

Table 6 Efficacies by numerical analysis and centrifugal test


Page 49 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Table 1 Material properties

Material Properties
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Foundation soil ρ= 1.81 g/cm3 (CV=0.008), w=37.5%(CV=0.039),


ϕ=4.3º(CV=0.065), c=9.66kPa(CV=0.045)
Embankment fill ρ= 1.65 g/cm3, ϕ=35º
Sand cushion ρ= 1.70 g/cm3, ϕ=37º
DM column ρ= 1.93 g/cm3, qu=1.1MPa(CV=0.094), D=16mm
Core pile ρ= 2.70 g/cm3, EA=1.52×106N, D=8mm, t=1mm
Slab ρ= 0.96 g/cm3, t=8mm, EI=46N·m2
Note: ρ=density, w=water content, ϕ=friction angle, c=cohesion, CV=coefficient of variation, v=Poisson’s ratio, E=
elastic modulus, qu=unconfined compressive strength, EA=compressive rigidity, EI=flexural rigidity, t=thickness,
D=diameter.
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 50 of 54

Table 2 Model test programs

Test No. 1 2 3 4 5 6
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Improvement method SDM SDM SDM SDM SDM DM


Bearing stratum Floating Floating Floating End bearing Floating Floating
With/without slab With With With With Without With
DM column length (mm) 200 200 240 200 200 200
Core pile length (mm) 160 120 160 160 160 N/A
Page 51 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Table 3 Main properties used in the numerical study


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Material Constitutive Properties


model
Foundation soil Modified ρ= 1.8 g/cm3, E=4MPa, ν=0.35, κ=0.0084,
Cam-clay λ=0.1005, M=1.38, pc=80kPa, υ0=2.18,
k=3×10-4cm/d
DM column Mohr-Coulom ρ= 1.93 g/cm3, E=110MPa, ν=0.3, c=550kPa,
b D=800mm, k=3×10-5cm/d
Note: ρ=density, E=elastic modulus, ν=Poisson's ratio, κ= slope of swelling line, λ= slope of normal consolidation line, M= frictional constant,
pc=pre-consolidation pressure, υ0=initial specific volume, ϕ = friction angle, c=cohesion, D=diameter, t=thickness, k=permeability.
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 52 of 54

Table 4 Influence factors considered in the parametric study

Influence factor Variable


Length of core pile 0/1/5/8/10m
Column spacing 2.6/3.3/4m
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Thickness of underlying soil 0/3/5m


Modulus of slab 1/10/30GPa
Modulus of cushion 40/60/80/100MPa
Thickness of cushion 200/300/400mm
Page 53 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Table 5 Main properties of baseline case in the parametric study

Material Constitutive Properties


model
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18

Foundation soil Modified ρ= 1.8 g/cm3, E=4MPa, ν=0.35, κ=0.0084,


Cam-clay λ=0.1005, M=1.38, pc=80kPa, υ0=2.18
Embankment fill Mohr-Coulom ρ= 1.65 g/cm3, E=15MPa, ν=0.3, ϕ=35º, ψ=5º
b
Sand cushion Mohr-Coulom ρ= 1.7 g/cm3, E=40MPa, ν=0.3, ϕ=37º, ψ=7º,
b t=300mm
DM column Mohr-Coulom ρ= 1.93 g/cm3, E=110MPa, ν=0.3, ϕ=25º,
b c=550kPa, D=800mm
Core pile Elastic ρ= 2.4 g/cm3, E=30GPa, ν=0.2, D=400mm
Slab Elastic ρ= 2.4 g/cm3, E=30GPa, ν=0.2, t=500mm
Note: ρ=density, E=elastic modulus, ν=Poisson's ratio, κ= slope of swelling line, λ= slope of normal consolidation line, M= frictional constant,
pc=pre-consolidation pressure, υ0=initial specific volume, ϕ = friction angle, c=cohesion, D=diameter, t=thickness.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by GLASGOW UNIVERSITY LIBRARY on 10/28/18
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Test case
F-DM-S(L160/200)
F-SDM-S(L160/200)
E-SDM-S(L160/200)
F-SDM-N(L160/200)
33.2
89.3
63.9
37.2
Centrifuge,%

30.3
78.3
64.8
38.6
Numerical,%
Table 6 Efficacies by numerical analysis and centrifugal test
Page 54 of 54

You might also like