You are on page 1of 18

Analytical and Experimental Studies on Cemented Stone

Columns for Soft Clay Ground Improvement


Yadav Sambhaji Golait, Ph.D.1; and Amit Harihar Padade, Ph.D.2
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: This paper reports a novel concept of ground improvement technique using cemented stone columns for improvement of the soft
clay ground. The experimental investigations and analytical studies led to evolution of the technology of soft ground improvement using such
stone columns almost to the fullest extent. The study included the following aspects: relative degree of effectiveness of cemented stone col-
umns over other reported approaches, theoretical analysis for predicting the ultimate load-carrying capacity of soft clay ground treated with
cemented stone columns, analysis to estimate the probable short- and long-term settlements of a given structure founded on the treated ground
under the applied foundation pressure intensity, and an illustrative design problem using this technology. The failure mode and the load-transfer
mechanism for this system were conceptualized on the basis of the unit-cell concept. The mathematical formulation resulted in equations ena-
bling quantitative evaluation of the required design aspects related to the actual design of cemented stone column–soft clay systems. DOI:
10.1061/(ASCE)GM.1943-5622.0000779. © 2016 American Society of Civil Engineers.
Author keywords: Soft clay ground; Cemented stone columns; Load-carrying capacity; Settlement; Unit cell.

Introduction column or group of stone columns (Ranjan and Rao 1983), geotex-
tile encasement of the column (Raithel et al. 2002; Murugesan and
The stone column technique is becoming increasingly popular for Rajagopal 2006; Gniel and Bouazza 2008; Deb et al. 2010; Yoo
improvement of soft clay ground such as soft coastal clays, expan- 2010), natural-bamboo-encased stone column (Dutta et al. 2012),
sive soil, and relatively looser silty clays in river valley regions. The use of several reinforcing geofabric layers placed horizontally at
conventional practice of stone column construction is to install the several levels or providing a rigid casing pipe in the top portion of
straight shafted columns with fill material in the form of an unce- stone column (Madhav 1983; Rao and Nayak 1995), incorporation
mented mixture of crushed stones, gravel, and coarse sand of randomly distributed reinforcing fibers in sandy fill material
(Barksdale and Bachus 1983). The ultimate load-carrying capacity (Maher and Gray 1990; Samadhiya et al. 2008), reinforcement of the
of such column-treated ground is governed, besides other factors, by stone column with vertical nails (small-diameter steel bars) driven
the lateral passive resistance the surrounding soft clay can offer to along the circumference (Babu et al. 2010; Nayak et al. 2011), and
the lateral bulging of stone column within the upper length of about stone columns strengthened by additives (Fattah et al. 2013).
3 to 4 times the diameter of the column under vertical applied loads. Experimental studies on the settlement and deformation behav-
Several researchers have done extensive work on stone columns ior of stone columns is also carried out ( e.g., Alamgir et al. 1996;
in different directions, such as testing and installation of stone col- Zhang et al. 2013; Castro et al. 2014; Killeen and McCabe 2014;
umns (Datye and Nagaraju 1977), performance of a stone column as Zhang and Zhao 2014), and, recently, mathematical modeling for a
a foundation (Mitchell and Huber 1985), load-carrying capacity similar approach was reported by Yoo (2015). Mathematical and
improvement in soft soils (Shahu and Reddy 2011; Al-Saoudi et al. numerical studies on cement mixed stone columns have also been
2015), yield strength of stone columns (Hassen et al. 2013), failure reported in the recent past (e.g., Ng and Tan 2014; Jamsawang et al.
mechanism of single and grouped stone columns (Hanna et al. 2013; 2015; Gueguin et al. 2015).
A new concept consisting of use of a semirigid-type cemented
Chen et al. 2015), and different structures supported by stone col-
stone column for enhancing the performance of stone columns in
umns (Deb 2010; Deb et al. 2011; Castro 2014; Das and Deb 2014).
treated soft clay ground has been recently introduced by Golait
The literature reports some approaches to improve the effective-
et al. (2009). The construction of such columns does not pose much
ness of a stone column by restraining its lateral bulging under axial
difficulty, and not much additional cost is involved because the fill
loads. Some of these techniques are as follows: cemented compacted
material used in the formation of the columns is the usual granular
stone column (Juran and Riccobono 1991), soil–cement column
material, except that it is mixed with a marginal amount of cement-
(Farouk and Shahien 2013), provision of skirting around the stone
ing admixture (3–4% cement and 6–7% fly ash by weight). It is
reported that the ultimate bearing capacity of soft clay ground is sig-
1
Ex. Professor Emeritus, Civil Engineering Dept., Ramdeobaba nificantly increased by use of such cemented stone columns instead
College of Engineering and Management, Katol Rd., Nagpur 440013, of the conventional uncemented stone columns. Similarly, enhanced
India. E-mail: yadavgolait@gmail.com improvement in ground stiffness was also observed in the model
2
Assistant Professor, Civil Engineering Dept., Ramdeobaba College of studies reported by Golait et al. (2009).
Engineering and Management, Katol Rd., Nagpur 440013, India (corre-
This paper presents the analytical and experimental studies
sponding author). E-mail: amit51085@gmail.com
Note. This manuscript was submitted on December 2, 2015; approved undertaken to develop this cemented stone column technology
on June 30, 2016; published online on September 12, 2016. Discussion pe- almost in its entirety. It includes such aspects as idealized load-
riod open until February 12, 2017; separate discussions must be submitted transfer mechanism for the soil–column system, analysis to evolve
for individual papers. This paper is part of the International Journal of an equation for predicting the ultimate load-carrying capacity of the
Geomechanics, © ASCE, ISSN 1532-3641. cemented stone column in treated soft clay ground, and the analysis

© ASCE 04016100-1 Int. J. Geomech.

Int. J. Geomech., 04016100


for short- and long-term settlements under the design load of struc-
ture. The experimental investigation to assess the relative degree of
effectiveness of these cemented stone columns with respect to other
main approaches is also highlighted.

Basic Principle

The lateral bulging of the conventional stone column under the action
of axial load limits its load-carrying capacity. It can thus be inferred
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

that if a nonbulging type of fill material is used in forming the stone col-
umn, it is expected to result in the enhanced effect of treatment of soft
clay ground. The use of concrete as fill material and in forming the
highly rigid structural elements is naturally the ideal solution. But this
becomes a problem in the provision of deep pile foundations and as
such cannot remain a cost-effective ground-improvement technique.
In the cemented stone column approach, the conventionally used
coarse fill material (i.e., crushed stones with certain gravel content)
was converted into a semirigid-type porous mass. This was accom-
plished by mixing it with a marginal amount of cementitious admix- Fig. 2. Conceptual nature of stress–strain curves for cemented stone
tures, such as cement and fly ash. The admixture that was found to column and soft clay soil in unit cell
give the desired effect is a mixture of 3–4% cement and 6–7% fly
ash by weight of the stone particles. A wet workable mix of crushed
stone and gravel with this admixture was therefore used as a fill ma-
terial in the formation of the stone column. After curing under wet Table 1. Composition by Weight of Different Constituents in the Mix
clay ground conditions, the result is a porous mass with cementation Symbol Constituent of mix Specification (mm) Proportion (%)
bonds between particles at their points of contact, as shown in Fig. 1.
Thus, the cemented stone column was found to exhibit insignifi- a Coarse crushed stones 15–30 50
cant lateral bulging and to transfer the applied load to a larger depth, b Fine crushed stones 5–15 50
thereby effecting larger participation of soil within its unit cell in c Coarse sand and gravel 2–5 25 (of a þ b)
the load-transfer mechanism. The conceptual stress–strain curves d Cement 3–4 (of a þ b þ c)
for two different media in unit cells—namely, soft clay and the stiff e Fly ash 6–7 (of a þ b þ c)
stone column—are schematically represented in Fig. 2.

Preferable Composition of Mix for Cemented Stone


Column

The main criteria in deciding the composition of mix is that it


should be possible to obtain the porous fill material of required
properties—namely, the modulus of elasticity (Ec) and compressive
strength (s f)—with the least amount of cement content. From the
primary studies, it was found that the crushed stones of moderate
size (10–30 mm) with 25% coarse sand–gravel are better suited
in this respect. Based on the detailed investigation, the composition
as given in Table 1 was found desirable. The grain-size distribution
for this mix is shown in Fig. 3. Fig. 3. Particle-size distribution of specified cemented stone column
The mixing was done with 5% water with respect to total dry mix
weight of the dry mixture. It should be noted that the cement

content is 18–20% for commonly used concrete pile construc-


tion, whereas it is only 3–4% in cemented stone columns.
A cylindrical sample of 10 cm in diameter and 20 cm in height
was prepared in a mold by using the specified mix as mentioned in
Table 1. It was cured in water for 7 days and tested to find its com-
pressive strength. Fig. 4 presents the stress–strain curve of
cemented stone column sample observed in the compression test. It
exhibited an almost linear relationship between stress and strain up
to a stage of cracking. The sample was found to get crushed soon
thereafter. The compressive strength and strain at failure for the
sample were found to be 2,200 kPa and 1.15% respectively, and the
modulus of elasticity (Ec) was 15  104 kPa. The lateral deforma-
Fig. 1. Matrix of cemented stone column material
tion for the sample was found to be insignificant.

© ASCE 04016100-2 Int. J. Geomech.

Int. J. Geomech., 04016100


The cemented stone column formed in soft clay ground has to Considering the laboratory test results, the specified wet mix af-
be highly pervious to enable free draining of water. A constant- ter proper compaction and 7 days of curing is expected to result in
head permeability test was conducted on an identical cylindrical the formation of a semirigid-type cemented stone column in the
sample. Appropriate modifications were made in the conventional field with the following property ranges:
test setup considering the large-size cylindrical stiff sample, as Compressive strength ðs f Þ ¼ 2;000–2;500 kPa
shown in Fig. 5. The coefficient of permeability (k) was found to Modulus of elasticity ðEc Þ ¼ ð18–22Þ  104 kPa
be 1.5  10 −1 cm/s. Coefficient of permeability ðkÞ ¼ ð1:4–1:5Þ  101 cm=s
These properties indicate that after curing, the cemented stone
column is a strong, stiff, and porous element, and it will have per-
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

meability equivalent to that of a free-draining sand–gravel mix-


ture. The schematic structure of the cemented stone column is
shown in Fig. 1. The photographic view of the cylindrical sample
prepared for the composition as indicated in Table 1 is depicted in
Fig. 6.

Relative Degree of Effectiveness of Cemented Stone


Columns

Laboratory model tests were conducted to observe the effectiveness


of various modified stone columns, including cemented stone col-
umns. The selected clay–column systems considered were as
follows:
1. Conventional stone column of uncemented granular material
(CU)
2. Skirted uncemented stone column (SCU)
3. Uncemented stone column with partial encasement by geotex-
tile tube (PEU)
4. Uncemented stone column with full-length encasement by geo-
Fig. 4. Observed stress–strain curve for cemented stone column textile tube (FEU)
sample 5. Uncemented stone column reinforced by horizontal geotextile
discs (HRU)

Fig. 5. Constant-head permeability test setup for cemented stone column sample

© ASCE 04016100-3 Int. J. Geomech.

Int. J. Geomech., 04016100


test tank, the bore hole was made by auger, an appropriate stone column
was formed, and the assembly was kept in the water-filled saturation
tank for 4–5 days, during which the soil attained near-saturation state.
The saturation tank was dewatered before the load test. The soil under
test conditions exhibited a dry unit weight of 1.23–1.25 g/cm3 and water
content of 40–42%. The undrained cohesion (Cu) of clay in this state was
found to be 20 kPa. A circular rigid steel plate of 10 cm in diameter was
placed over the column, and the load test was performed by the incre-
mental load method. A schematic representation of the experimental
setup for the load test is shown in Fig. 8.
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

The load-settlement curves obtained from the model tests are


shown in Fig. 9. The characteristic features of all the curves indicate
an almost linear relationship between load and settlement initially
over a considerable load range, which was then followed by a non-
linear relationship leading to failure of the soil–column system. The
failure load (Qf) for the unit cell was determined by the double-
tangent method. A procedure for the same is illustrated, with
respect to a typical PSC system, in Fig. 10. Table 2 shows the
observed values of failure pressure for different soil–column sys-
tems, along with the analysis of test results. The following fea-
tures were observed from this series of tests:
1. Compared with the load-carrying capacity of the unit cell with
conventional uncemented stone columns, the other systems
resulted in enhancement in the load-carrying capacity ranging
from 11 to 57%. Notably, the cemented stone column system
gave the highest enhancement, as indicated in the Table 2.
2. The increase in the stiffness of the unit cell was observed to be
in the range of 23–75%. Here also, the cemented stone columns
were observed to be the most effective.
The study thus indicated that the soft clay ground treated with
the cemented stone column is the most effective way of such ground
improvement compared with any other method or approach to stone
column modification in terms of load-carrying capacity and the
stiffness of the treated ground.
Fig. 6. Photographic view of the cylindrical sample of cemented stone
column
Analysis for Predicting Load-Carrying Capacity of
Cemented Stone Column–Treated Soft Clay Ground
6. Part-length cemented stone column (PSC)
7. Full-length cemented stone column (FSC) The behavior of a cemented stone column–soft clay soil system and its
In addition to these systems, an additional eighth case of a highly idealized load-transfer mechanism was conceptualized to mathemati-
rigid stone column in the form of concrete pile element (RCP) was cally formulate the ultimate load-carrying capacity of the treated soft
also included in the investigation for comparison of stone column clay ground. These aspects are explained in the following sections.
systems with concrete pile foundations.
Idealized Load-Transfer Mechanism
Experimentation and Observations A simple idealization for interaction between tributary soft clay in
The model unit cell used in the laboratory investigation was selected the unit cell and the cemented stone column is described with
to represent the field stone column layout of 40-cm-diameter stone respect to Fig. 11.
columns of 4-m length spaced at 0.95 m in a triangular pattern. A unit cell of diameter Du incorporating a floating cemented stone
Selecting the scale reduction factor of 10, the geometry was as fol- column of diameter ds and length L is acted upon by a uniform pres-
lows: stone column of diameter ds = 4 cm; length L = 40 cm; and di- sure s through a rigid circular base of area A [Fig. 11(a)]. Because of
ameter of unit cell Du = 10 cm. The prototype and model pertained the different stiffness values of the tributary soil and stone column, the
to an area replacement ratio (ar = a/A) of 16% and a length ratio (rl = stresses on top of unit cell are stress on soil s s and stress on column
L/ds) of 10, where ds = diameter of stone column; L = length of col- s c [Fig. 11(b)]. It has been pointed out by Hughes and Withers (1974)
umn; a = cross-sectional area of stone column; and A = area of unit that very little of the applied load reaches the bottom of an isolated
cell. The unit cells for the aforementioned eight cases of soil–column conventional stone column if the column length is larger than about 3
systems are shown in Figs. 7(a–h). to 4 times its width. In view of nonbulging nature of the cemented
The experimental setup comprised of a perforated mild steel (M.S.) stone column, the axial stress in column, s c, will be maximum at top
tank (having a diameter and height of 50 cm), loading frame fitted with level and will decrease with depth. Accordingly, the vertical down-
screw jack and proving ring, laboratory auger for 4-cm-diameter bore ward displacement of column d will be decreasing with depth and
holes in compacted clay, and dial gauges for settlement observations. may become negligible at certain level [Figs. 11(c and d)]. The stress
The field stone column installation procedure was simulated in the labo- in soil compresses the soft clay within a zone-of-pressure bulb whose
ratory model preparation. The wet clay was uniformly compacted in the depth is considered to be of the size of loading base, that is, Du.

© ASCE 04016100-4 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Schematic representation of investigated eight unit cells of soft clay–stone column systems: (a) CU case; (b) SCU case; (c) PEU case; (d) FEU
case; (e) HRU case; (f) PSC case; (g) FSC case; (h) RCP case

© ASCE 04016100-5 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

© ASCE
04016100-6
Fig. 7. (Continued.)

Int. J. Geomech., 04016100


Int. J. Geomech.
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Experimental test setup

Fig. 9. Load-settlement curves for different unit cells of clay–column system Fig. 10. Determination of failure load for a typical PSC test

Beyond this depth there is practically no soil compression or its verti- there will be relative displacement between column and soil up to a
cal downward movement. Thus, considering the nature of movements certain depth Ze below ground surface. The magnitude of the relative
of soil and stone column as just conceptualized, it can be realized that displacement of column with respect to soil will be large at depth of,

© ASCE 04016100-7 Int. J. Geomech.

Int. J. Geomech., 04016100


Table 2. Comparative Results for Various Soil–Stone Column Systems

Failure Failure pressure Increase in qf w. r. t. Increase in qf w. r. t. Increase in stiffness w. r. t.


Unit cell system load (kN) qf (kPa) CU case (kN) CU case (%) CU case (%) Remarks
CU 0.971 123.6 — — — Stone column systems
SCU 1.226 156.2 0.26 26.26 23.08
PEU 1.128 143.6 0.16 16.16 33.23
FEU 1.314 167.3 0.35 35.35 35.36
HRU 1.079 137.3 0.11 11.11 27.46
PSC 1.403 178.5 0.44 44.44 37.40
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

FSC 1.530 194.8 0.57 57.57 75.59


RCP 1.815 231.0 0.86 86.87 135.10 Concrete pile foundation
Note: w.r.t. CU = with respect to conventional uncemented.

Fig. 11. Conceptualized stress and displacements in cemented stone column: (a) soil–stone column system of unit cell; (b) downward displacement d
of stone column under stress s c and stresses on unit cell; (c) distribution of axial stress s c in stone column; (d) variation of displacement d with depth z

say, Zt, and thereafter it will decrease gradually. Beyond Zc (hereafter The conceptual load-transfer mechanism just described leads to
referred to as Le), there may not be any relative displacement between three component forces resisting the ultimate applied axial force Qf
column and soil [Fig. 12(a)]. on unit cell, as expressed by
It is well established from experience and studies on axially
loaded piles that the full shaft resistance develops during an early Q f ¼ R s þ Rf þ Rt (1)
stage of loading or at small pile penetration, whereas a considerable where Qf = failure load on unit cell; Rs = soil resistance offered by
pile displacement is required for full development of tip resistance. tributary soil of area (A – Ac) at local shear failure of soft clay; Rf =
The cemented stone column in a unit cell is also visualized to ex- ultimate shaft surface resistance mobilized over length L (L ≤ Le)
hibit similar behavior with respect to development of shaft surface (Le is the effective length of cemented stone column); and Rt =
resistance and tip resistance. Considering the nature of relative mobilized tip resistance of column when L ≤ Le.
downward displacement of the column as shown in Fig. 12(a),
the shaft resistance may be assumed to develop fully over its entire
length up to the Ze level because this requires small relative move- Analysis for Load-Carrying Capacity
ment of the stone column. The development of tip resistance, how- Basic Assumptions
ever, will depend on the magnitude of column displacement at the The resisting forces in Eq. (1) are mathematically formulated with
bottom. At a depth of about 3ds, which is the depth of probable bulg- the following basic assumptions:
ing under high stress level and where the column movement is 1. The unit-cell concept is valid for estimating the performance of
large, full development of tip resistance is expected, and at depth of cemented stone column–treated soft clay ground, and the
Ze, where practically no column movement takes place, negligible approach of static load equilibrium between applied load and
or zero tip resistance will be mobilized. Thus, there will be reduc- the resisting forces within a unit cell holds well. This unit-cell
tion in the ultimate bearing resistance from a depth 3ds to depth Ze, concept is extensively used in the analysis of conventional
that is, from 0 to 100% reduction [Fig. 12(b)]. stone column system (Hughes and Withers 1974; Barksdale

© ASCE 04016100-8 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Scheme of load transfer at the tip of cemented stone column: (a) relative displacement between soil and column; (b) reduction in ultimate
bearing resistance; (c) reduction factor

and Bachus 1983). It is intended to take into account the three- theories of Prandtl (1920), Terzaghi (1943), Mayerhof (1951), and
dimensional effect within the boundary of unit cell, that is, the Skempton (1951). Eq. (3) thus leads to
cylindrical composite system of soil and column. qf ¼ 4:04 Cu (4)
2. Soil is saturated soft clay whose angle of internal friction w is very
small (w < 12°), and undrained cohesion Cu is less than 25 kPa.
3. The effective length of the column, Le, over which shaft resist-
ance and tip resistance develop pertains to length ratio (rl) of Unit-Cell Analysis for Soft Clay with Cemented
23, where rl = L/ds. Stone Column
4. The reduction in ultimate tip resistance for any depth or length Fig. 14 represents a single unit cell with various resisting forces at
of column is expressed by the tip bearing-capacity reduction fac- failure.
tor (rd). This factor is assumed to vary from 0 at depth of 3ds to 1 • Rs is the resisting force effected by soft clay soil for local shear
at depth of 23ds, as shown in Fig. 12(c). Thus, the value of rd at failure under a rigid circular base of area (A – a). It can be
any depth level z below ground surface can be expressed as expressed as
 
rd ¼ 0:05
z
 0:15 (2) Rs ¼ 4:04 Cu  ðA  aÞ (5)
ds • Rf is the shaft resistance over length L of the column and is
given by
5. Only the top layer of soil within the depth of Du participates in
the local shear failure of soil below the circular load-transferring Rf ¼ ðp  a  ds :  LÞ Cu (6)
rigid base. The undrained cohesion within this zone, Cu0 , is
assumed to be 10% less than the average undrained cohesion Cu where L  Le ; and a = adhesion factor, whose value is assumed
for the entire depth below ground surface up to (L þ 3ds). The to be 0.95 for soft clay with Cu ≤ 25 kPa.
reduction of 10% is to approximately take into account the rela- • The tip resistance Rt will depend on the tip bearing-capacity
tively looser or softer nature of ground near the surface. reduction factor for the column tip level, that is, rdt. Using
Eq. (2), its value is
Unit-Cell Analysis for Untreated Soft Clay  
A rigid circular base of diameter Du having area A resting on soft L
rdt ¼ 0:05  0:15 (7)
clay is expected to create local shear failure in soil within depth ds
equal to Du, as shown in Fig. 13.
The ultimate bearing capacity qf can be expressed as where L  Le .
 
2 0 Hence, mobilized tip resistance, Rt is expressed as
qf ¼ 1:2 Cu  Nc (3)
3
Rt ¼ 9Cu ð1  rdt Þ  a
where Cu0 ¼ 0:9Cu ; and Nc = bearing-capacity factor. Its value is
fixed as 5.61, which is the average of four different values from the where a = cross-sectional area of column.

© ASCE 04016100-9 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Failure in untreated soft clay

Thus
  
L
Rt ¼ 10:35  0:45 a  Cu (8)
ds

Adding Eqs. (5), (6), and (8) and dividing by area A of the unit
cell, the expression for failure load-carrying capacity of treated soft
clay is obtained as
   
4:04 ð A  aÞ p  a  ds :  L
ð f Þt
q ¼ C u þ Cu
A A
  
L a
þ 10:35  0:45 Cu (9)
ds A

The nondimensional parameters of area replacement ratio (ra)


and length ratio (rl) are introduced in the Eq. (9) as
  Fig. 14. Unit cell of cemented stone column at failure
a L ds 4 ra
¼ ra ; ¼ rl and ¼ 
A ds A p ds
The analysis for load capacity qf is presented in the paper for a
Hence, the final form of equation for (qf)t is single unit cell in a group of a number of cemented stone columns
covering the entire area under the proposed structure. Therefore, it
ðqf Þt ¼ ½f4:04 ð1  ra Þg is necessary to incorporate the group effect while estimating the
design value of the overall failure pressure intensity qf for the group
þ f4  a  ra  rl g þ fð10:35  0:45rl Þ  ra g Cu (10) of columns. This can be expressed as

The bearing-capacity analysis leading to formulation of Eqs. (4) qf ðfor groupÞ ¼ Cg  qf ðfor single columnÞ
and (10) is based on the soft clay strength properties under the
undrained condition. This is viewed to be more relevant to highly where Cg ¼ correction factor for group effect:
plastic clayey soils in the field because the pore-water-pressure dis- For the conventional stone columns of uncemented stones/
sipation in a deep soft clay deposit is a very slow process. As such, course granular materials, Barksdale and Bachus (1983) have
for the design of structures on these deposits, the bearing capacity pointed out that the group correction factor is slightly larger than
under the undrained condition is a more critical case. The long-term 1.0. Each stone column in a group is found to have slightly higher
capacity in the drained condition will have a larger value because of load capacity compared with an isolated single column. They
complete dissipation of excess pore-water pressures and effective note that this is because, as the surrounding columns are added to
stresses becoming equal to total stresses. The footings and rafts on form a group, the interior columns are confined and hence some-
clays are commonly analyzed based on these considerations. what stiffened by surrounding columns. The small-scale model
However, for the cemented stone column element, the undrained studies they carried out by them (on columns of diameter = 3 cm
condition is irrelevant in view of its very high permeability, which and length ratio = 6.3 in soft clay with undrained cohesion of
is almost equal to that of free-draining granular soil material. 17 kPa) showed that for groups having one and two rows of stone

© ASCE 04016100-10 Int. J. Geomech.

Int. J. Geomech., 04016100


columns, the load capacity increased almost linearly with number replacement ratio ra and length ratio rl. The variation of Fb with
of columns in a group. An individual column in a group of seven ra and rl is shown in Fig. 15.
columns showed an increase in capacity of about 16%. As an illustration, the analytical results point out that for a typi-
In the present case of soft clay ground improved by cemented cal layout of cemented stone columns with ra = 15% and rl = 20, the
stone columns, the mechanistic behavior was found to be similar to load-carrying capacity may increase 3.75 times.
the aforesaid phenomenon as conceived for a conventional unce-
mented stone column group. However, for arriving at a definite con-
clusion on Cg in this case, further extensive investigations are
Experimental Investigation for Verification of
needed. It may be realized that this aspect is beyond the scope of the the Analysis
present paper.
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

To verify the validity of the previously described theoretical analy-


sis, laboratory model tests were conducted on a unit cell of
Bearing-Capacity Improvement Factor cemented stone column in soft clay. The test included columns with
diameters of 3 and 4 cm and lengths of 10, 35, and 40 cm, with unit-
Assessment of the degree of improvement of soft clay ground as a
cell diameters of 10 and 15 cm. The model configurations repre-
result of installation of cemented stone columns can be made by
sented cemented stone column layouts of area replacement ratios of
using the bearing-capacity improvement factor Fb. It is defined as a 4, 9, and 16% and length ratios of 10.0, 11.67, and 13.33. The soil
ratio of the failure pressure of soft clay treated with cemented stone used in the investigation was clay of high compressibility (CH)
columns to the failure pressure of untreated soft clay ground. Fb can black cotton soil as per BIS-1498 (BIS 1970) in the neighborhood
be obtained by dividing Eq. (10) by Eq. (4). of Nagpur City in India. It had liquid limit wL = 65% and a plasticity
It is thus realized that the bearing-capacity improvement factor index Ip = 35%. The cemented granular mix used to construct the
for the case under consideration will depend on two nondimen- small-size model stone columns comprised 60% coarse sand and
sional parameters characterizing the column layout system: area 40% fine sand mixed with 2% cement and 6% fly ash. Water content
of 5% was found appropriate to get a homogeneous workable moist
mix that could be easily filled and compacted in the small-size bore
hole made in the compacted clay.
The results of triaxial tests on cured cylindrical samples of the
mix showed the following properties: cohesion (C) = 68 kPa; angle
of internal friction w = 37°. The experimental setup, model prepara-
tion, and test procedure were the same as described for the earlier
series of test program. Seven unit cells, with details as shown in
Table 3, were load tested. A nearly identical physical state for the
soil used in different tests was ensured with dry unit weight of
1.23 g/cm3 and water content of 42%. The undrained cohesion (Cu)
in this state was found to be 19 kPa from the unconfined compres-
sive strength test.
The load-settlement curves for seven categories of unit cells are
shown in Fig. 16. The curves indicate almost direct proportionality
between load and settlement initially over a considerable load
range, which is then followed by nonlinear nature leading to plastic
yielding of the soil–column system, as reflected in the constant high
rate of settlement under load increments. The failure load of the sys-
tem was determined by the double-tangent method. The procedure
is illustrated for a typical case of unit cell (No. 6) in Fig. 17. The
failure load values, in kilograms, thus determined for all seven tests
were as follows: (1) 62, (2) 135, (3) 110, (4) 117, (5) 120, (6) 181,
and (7) 156. The corresponding failure pressure intensities, qf, and
their comparison with the computed values using analytical
Fig. 15. Variation of bearing-capacity improvement factor Fb with Eq. (10) are given in Table 3.
area replacement ratio ra and length ratio rl It can be seen that there was close agreement between the
observed and computed values of qf. For the cases studied, the

Table 3. Comparison of Observed and Computed Capacity

qf (kPa)
Unit cell ds (cm) Du (cm) L (cm) ra (%) rl Observed Computed Error in predicting qf (%)
1 0 10 0 0 0 79 77 −2.53
2 0 15 0 0 0 76 77 þ1.32
3 3 10 30 9 10.00 140 145 þ3.57
4 3 10 35 9 11.67 149 154 þ3.35
5 3 10 40 9 13.33 153 164 þ7.19
6 3 15 30 4 10.00 102 107 þ4.90
7 4 10 40 16 10.00 197.3 197.7 þ0.20

© ASCE 04016100-11 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 16. Load-settlement curves for different unit cells of cemented


stone column

Fig. 17. Determination of failure load for a typical case of unit cell
(No. 6) Fig. 18. Stress–strain curve from UCS test on soft clay sample: (a)
observed stress–strain curve; (b) hyperbolic model parameters a and b

theoretically predicted values of qf using the described analysis had


marginal errors ranging between only 0.2 and 7.2%. both the materials (i.e., soft clay and the cemented stone col-
umn) were conceptualized as shown in Fig. 2.
4. The pressure coming on the top of the tributary soil of the unit
Settlement Analysis cell was transferred within the depth of the pressure bulb below
the circular base, which was assumed to be 2Du, where Du =
Basic Considerations diameter of unit cell.
5. The axial stress in the column beyond the depth of 23ds was
The analysis for settlement of a structure constructed on cemented assumed to be negligible.
stone column–treated soft clay ground was carried out with certain 6. The uniform settlement, r , of the rigid loading base was
assumptions and idealizations, as follows: assumed to be the same as the vertical compression of tributary
1. The concept of the unit cell was assumed to be valid for settle- soil, r s, and also the same as vertical compression of a stone
ment of a rigid circular base transferring uniform applied pres- column, r c. Thus, r = r s = r c.
sure, s , to the treated ground.
2. The nonlinear nature of the soil with respect to stress–strain
Short-Term Settlement Analysis
behavior was taken into account, and this was characterized
by a hyperbolic model as proposed by Duncan and Chang The short-term analysis involved the stress–strain behavior of soft
(1970). This model was used for analysis of short-term settle- clay soil under undrained conditions as obtained from the uncon-
ment. For long-term settlement analysis, the consolidation fined compressive strength (UCS) test conducted on the clay sample
behavior of soil as obtained from consolidation test was con- [Fig. 18(a)]. The stress–strain curve from the test was assumed to be
sidered valid. hyperbolic in nature.
3. The stress–strain behavior of cemented stone columns was Duncan and Chang (1970) expressed such a curve by the
assumed to be linear up to failure. The stress–strain curves for equation

© ASCE 04016100-12 Int. J. Geomech.

Int. J. Geomech., 04016100


Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 19. Unit cell of soil–column system: (a) vertical pressures on top surface; (b) variation of vertical pressure in clay soil; (c) variation of vertical
pressure in cemented stone column

ɛv
sv ¼
a þ b  ɛv

where s v ¼ axial ðverticalÞ stress, ɛv ¼ axial ðverticalÞ strain, and


a and b are hyperbolic soil model parameters as determined from
Fig. 18(b).
Thus, the vertical strain can be given as
a  sv
ɛv ¼ (11)
1  b  sv

The unit cell of soil–column system under consideration is


shown in Fig. 19. Also represented therein are the variations of ver-
tical stresses in tributary soil and in the cemented stone column ele-
ment. The notations used in Fig. 19 are as follows: s = applied uni-
form pressure intensity on rigid circular base of diameter Du; s st =
vertical pressure acting on the top of tributary soft clay soil of area
(A – a); s sz= vertical pressure in soil at any depth level z within
pressure bulb of approximate depth 2Du; s ct = vertical pressure act-
ing on the top of cemented stone column of area a; and s cz = verti-
cal pressure in cemented stone column at any depth level z within
effective length of column Le (Le = 23ds).
Considering an elemental soil ring at depth z and of thickness dz,
the vertical pressure in soil at z will be,
 
z
s sz ¼ s st 1  (12)
2Du

The change in thickness of elemental ring Dz = « vz  dz.


Fig. 20. Stress–strain curve from consolidation test on soft clay sam- By making use of Eq. (11)
ple: (a) observed stress–strain curve; (b) hyperbolic model parameters
a  s sz
a0 and b0 Dz ¼ dz
1  b:  s sz

© ASCE 04016100-13 Int. J. Geomech.

Int. J. Geomech., 04016100


   
Thus, the total vertical compression of soil within the 2Du depth 2Du 1 1
I¼  log e 1
zone of the unit cell, r s, can be obtained as b b  s st 1  b  s st
ðu
2D ðu
2D
a  s sz Hence, the final form of the equation for settlement of soil ( r s)
rs ¼ Dz ¼  dz is
1  b  s sz
0 0    
2Du  a 1 1
rs ¼  log e 1 (13)
Substituting the values of s sz from Eq. (12) results in b b  s st 1  b  s st
 
z Two cases, namely, (i) when L ≥ Le and (ii) when L < Le, where
ðu s st 1 
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

2D
2Du Le = 23ds, need to be considered while evaluating the vertical com-
rs ¼ a     dz
z pression of the stone column, r c.
0 1  b s st 1 
2Du

Let (i) qc when L ≥ 23ds



 Consider an elemental column disc of thickness dz at depth z within
z
ðu
2D s st 1  the effective length of column as shown in Figs. 11(a and c). Under
2Du
I¼     dz the action of vertical stress s cz at this level, the reduction in thick-
z
0 1  b s st 1  ness of the elemental disc, Dz, will be Dz = « vz  dz, where « vz =
2Du
axial strain in column at depth z.
The stress–strain relationship for the cemented stone column
The important steps in solving the integral I in the previous equa- over a very large stress level below failure stress is almost linear.
tion are Hence
 
z s cz
ðu
2D s st 1  Ec ¼
 
2Du
  dz ɛvz

z
0 1  b s st 1 
2Du where Ec = Young’s modulus of elasticity for the column.
Therefore

ðu
2D s vz
1 2Du  z ɛvz ¼
I¼  dz Ec
b 2Du
0  2Du þ z
b: s st Hence
s vz
1 2Du Dz ¼  dz
I ¼  ðI1 Þ  Ec
b b
Thus
The integral I1 involved in the previous equation is
ðs
23d ðs
23d
ðBþ2D
ð uÞ s vz
1 rc ¼ Dz ¼  dz
I 1 ¼ ð A þ BÞ   dx Ec
x 0 0
B

The variation of vertical pressure in the cemented stone column


where as shown in Fig. 11(c) is
x¼Bþz  
z
s cz ¼ s ct 1:15 
20ds
2Du
B¼  2Du Substituting this value in equation for r c results in
b  s st
ðs
23d  
A ¼ 2Du rc ¼
s ct
1:15 
z
 dz
Ec 20ds
0
The value of I1 works out as

2Du
 2 3
2Du bs st ðs
23d ðs
23d
I1 ¼  log e s ct 6 z 7
b  s st 2Du rc ¼ 4 1:15 dz  dz5
 2Du Ec 20ds
b  s st 0 0

Thus Solving the integrals and simplifying result in

© ASCE 04016100-14 Int. J. Geomech.

Int. J. Geomech., 04016100


 
s ct  ds An elemental soil ring at depth z and having thickness dz under-
r c ¼ 13:225 (14)
Ec goes reduction in its thickness, Dz, as a result of full consolidation of
soil under the action of vertical pressure s sz as

Dz ¼ ɛvz  dz
(ii) qc when L < 23ds or rl < 23

ðL ðL   Substituting the value of « vz from Eq. (17) results in


s vz 1 z
rc ¼ dz ¼ s ct 1:15   dz s sz
Ec Ec 20ds Dz ¼  dz
0 0 a0 þ b0  s sz
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

Solving the integrals and simplifying results in Substituting the value of s sz from Eq. (12) results in
 
L  s ct rl ðu
2D
rc ¼ 1:15  (15)
Ec 40 rs ¼ Dz  dz
0
Thus, the settlement r of the unit cell is given by the following
equation:  
z
    ðu
2D s st 1 
2 Du  a 1 1 2Du
r¼  loge 1 rs ¼     dz
b b  s st 1  b  s st z
0 a0 þ b0 s st 1 
s ct :  ds 2Du
¼ 13:225 : : : when L  23ds (16a)
Ec
ðu
2D
    s
2 Du  a 1 1 rs ¼  st   dz
r¼  loge 1 z
b b  s st 1  b  s st 0 a0 þ b0
s st 1 
  2Du
L  s ct rl
¼ 1:15  : : : when L < 23ds (16b) ðu
2D
Ec 40 s st
=
  2Du   z  dz (18)
z
0 a0 þ b0 s st 1 
2Du
Long-Term Settlement Analysis
Solving the integrals and simplifying, the following equation for
The long-term settlement of a structure built on a cemented stone
r s is obtained:
column–treated soft clay deposit is considered to take place as a
    
result of primary consolidation of tributary soil in the unit cell. In 2 Du a0 a0
this case, the pore water in soil drains radially toward the pervious rs ¼ 0 1 þ 0  log e 0 (19)
b b  s st a þ b0  s st
cemented stone column, resulting in consequent vertical compres-
sion of clay. For the stress–strain relationship to be taken into
Hence, for the condition r = r s = r c, the governing relations for
account in the analysis pertaining to this field behavior of the unit
predicting the long-term settlement of cemented stone column–
cell, the data from the oedometer tests on undisturbed soil samples
treated soft clay are
were used.
    
The clay sample of initial height H was subjected to various con- 2 Du a0 a0
solidating pressure intensities s v1, s v2, s v3 . . ., s v(max) in stages r ¼ 0 1þ 0  log e 0
b b  s st a þ b0  s st
during oedometer tests. The final changes in height (DH1, DH2,
DH3 . . .) were noted under each stage of loading. s ct  ds
¼ 13:225 : : : when L  23ds (20a)
The vertical strain « v under any consolidating pressure intensity Ec
s v was obtained as « v = DH/H. The s v – « v plot was obtained and is     
shown in Fig. 20(a). This curve is analogous to the e (void ratio) – p 2 Du a0 a0
r¼ 1 þ  log e
(consolidating pressure) curve that is usually referred to in the con- b0 b0  s st a0 þ b0  s st
solidation process. The nonlinear relationship of the s v − « v plot is  
L  s ct rl
assumed to be hyperbolic and can be expressed as ¼ 1:15  : : : when; L < 23ds (20b)
Ec 40
sv
ɛv ¼ (17)
a0 þ b0  s v
Stepwise Procedure to Develop Load-Settlement Curve
where a and b = hyperbolic soil model parameters for consolida-
tion behavior; a0 has unit of pressure; and b0 is dimensionless. The set of equations formed in the analysis [as given in Eqs. (16)
The parameters a0 and b0 can be obtained by plotting the test and (20)] enables development of the load-settlement curve for soft
data, as shown in Fig. 20(b). clay ground provided with cemented stone column installation.
In this case, the vertical compression r s of tributary soil in a unit Then, the settlement under any desired loading intensity can be
cell was also formulated by considering the same load-transfer determined from the curves. The stepwise procedure to accomplish
mechanism as was idealized for the short-term settlement analysis. this is as follows:

© ASCE 04016100-15 Int. J. Geomech.

Int. J. Geomech., 04016100


1. Estimate the maximum value of s st. (It may be taken slightly • Thus, with a small area replacement ratio of 12.56%, the
more than the ultimate bearing capacity of untreated soft clay bearing-capacity improvement factor Fb is
deposit, which was formulated earlier as 4.04 Cu for a circular
load base; that is, s st(max) = 5 Cu). ðqf Þt 286:2
Fb ¼ ¼ ¼ 3:15
2. Divide s st(max) into a suitable 8–10 parts. qf 90:8
3. For each value of s st, obtain s ct using the relevant equation.
From these two values, calculate the stress concentration factor • The factor of safety (FOS) for the load-carrying capacity of
h (where, h = s ct/s st). treated ground for a given structure is
4. Find the value of s using the following basic equation:
ðqf Þt 286:2
5. s ¼ f½1 þ ð h  1Þra = h gs ct FOS ¼ ¼ ¼ 2:29 ðOKÞ
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

sd 125
6. Obtain the value of r . Repeat this process for all values of s st.
From the set of values of s and r thus obtained, plot the load- • Le = 23 ds = 23  0.4 = 9.2 m.
settlement curve. The settlement r under any design pressure • Hence, L < Le. Thus, using Eqs. (16b) and (20b), the load-
intensity s can be read. settlement curves can be constructed. The following settle-
ment values for a design pressure intensity of 125 kNm2 are
obtained from the curve: short-term settlement = 15.8 mm; long
Illustrative Design Problem
term settlement = 18.7 mm.
The following illustrative field problem of the design of a cemented
stone column layout for improvement of the soft clay ground is solved Summary and Conclusions
to highlight the use of the methodology described in the paper.
A new technology was developed for soft clay ground improvement
Problem by installation of cemented stone columns. Practically, its imple-
An oil storage tank of 8 m diameter is to be constructed on a site of a mentation is easy and involves not much additional cost. In addi-
deep soft clay deposit 9 m thick underlain by a firm stratum. The tion, the model studies carried out indicate its degree of effective-
design foundation intensity (s d) is 125 kN/m2. The laboratory inves- ness to be the largest compared with the other main stone column
tigations (UCS test and consolidation test) were conducted on undis- systems suggested in the literature.
turbed samples of clay. The design soil parameters were obtained as The purpose of actual planning and designing of the stone col-
follows: undrained cohesion, C u = 22.5 kN/m2 ; a = 2.15  10–4 umn layout for the intended improvement and analytical treatment
m2/kN; b = 8  10–3 m2/kN; a0 = 21  102 kN/m2; and b0 = 1.73. For based on the unit-cell concept was conceptualized with respect to
the proposed cemented stone column, Ec = 21  104 kN/m2. the load-transfer mechanism, load-carrying capacity, and the for-
The aim is to design a suitable layout of cemented stone columns mulation of equations for short- and long-term settlement of struc-
for a factor of safety of 2.0–2.5 for the bearing capacity and estimate tures founded on cemented stone column–treated soft clays. The
the probable short- and long-term settlements of the structure. idealized form of the load-transfer mechanism and its mathematical
formulation incorporate the nonlinear stress–strain behavior of
Solution clays, the strength properties, and the geometrical features of the
• Cemented stone columns of 0.4 m diameter (ds) spaced at 1.0
layout (i.e., diameter, length, pattern and spacing of columns).
m center to center (s) in a square pattern are proposed. The Eq. (10) enables easy estimation of the failure load-carrying
length of these floating columns (L) is decided as 7.5 m. capacity of treated soft clay ground, whereas Eqs. (16a), (16b),
Hence, Du = 1.13s = 1.13  1 = 1.13 m (20a), and (20b) facilitate computation of the settlement of a struc-
 2  2 ture under its design foundation pressure intensity. The results of
ds 0:4
ra ¼ 0:785 ¼ 0:785 ¼ 0:1256 the laboratory investigation program (model studies on unit cells)
s 1
confirm the validity of proposed bearing-capacity analysis. Solution
and of an illustrative design problem indicates a high degree of effec-
L 7:5 tiveness of the cemented stone column with respect to bearing
rl ¼ ¼ ¼ 18:75 capacity and settlement.
ds 0:4
The analysis presented in the paper is the first attempt for the sys-
• Using Eq. (4), the ultimate bearing capacity of untreated tem of cemented stone columns in soft clay material. Certain
ground is assumptions made in the study, especially with respect to the effec-
tive length of the cemented stone column, need verification and
ðqf Þt ¼ 4:04:Cu ¼ 4:04  22:5 ¼ 90:8 kN=m2 additional investigation. Further research in this respect may refine

the analysis.
Using Eq. (10), the ultimate bearing capacity of column-
treated ground is
ðqf Þt ¼ ½f4:04 ð1  ra Þg þ f4  a  ra  rl g Acknowledgments
þ fð10:35  0:45rl Þ  ra g Cu
The assistance rendered by R. Gautam, V. Satyanarayana, Ch. S.
ðqf Þt ¼ ½f4:04 ð1  0:1256Þg þ f4  a  ra :  rl g S. V. Raju, R. P. Kinikar, and M. N. Chute in the laboratory
investigation is highly acknowledged. The comments and
þ fð10:35  0:45rl Þ  ra g Cu suggestions of the three peer reviewers enabled the authors to
greatly improve the presentation and the technical quality of the
ðqf Þt ¼ 286:2 kN=m2 paper.

© ASCE 04016100-16 Int. J. Geomech.

Int. J. Geomech., 04016100


Notation r c ¼ settlement of stone column in unit cell;
r d ¼ dry density of soil;
The following symbols are used in this paper: r s ¼ settlement of tributary soil in unit cell;
A ¼ area of unit cell; s ¼ applied uniform pressure on unit cell/axial
A ¼ cross-sectional area of column; stress/stress;
Ac ¼ area of cemented stone column; s c ¼ stress on column;
a and b ¼ hyperbolic stress–strain parameters of soil with s cb ¼ stress on column at top level;
respect to UCS test; s cz ¼ stress on column at depth z;
a0 and b0 ¼ hyperbolic stress–strain parameters of soil with s d ¼ design foundation pressure intensity;
respect to consolidation test; s f ¼ compressive strength of cemented stone
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

C ¼ cohesion; column;
C0 ¼ effective cohesion; s s ¼ stress on soil;
Cg ¼ correction factor for group effect; s st ¼ stress on tributary soil at top level;
Cu ¼ undrained cohesion; s sz ¼ stress on tributary soil at depth z from top;
D ¼ particle size in mm; s v ¼ axial (vertical) stress; and
Du ¼ diameter of unit cell;
w ¼ angle of internal friction of soil.
ds ¼ diameter of stone column;
dz ¼ thickness;
Ec ¼ modulus of elasticity of cemented stone column References
material;
Es ¼ modulus of elasticity of soil; Alamgir, M., Miura, N., Poorooshasb, H. B., and Madhav, M. R. (1996).
F ¼ bearing-capacity improvement factor; “Deformation analysis of soft ground reinforced by columnar inclu-
H ¼ initial height of clay sample in consolidation sions.” Comput. Geotech., 18(4), 267–290.
test; Al-Saoudi, N. K. S., Rahil, F., and Abbawi, Z. (2015). “Soft soil improved
Ip ¼ plasticity index; by stone columns and/or ballast layer.” Ground Improv., 168(3),
179–186.
k ¼ coefficient of permeability;
Babu, M. R. D., Shivshankar, R., Nayak, S. S., and Majid, J. A. (2010).
L ¼ length of cemented stone column; “Load settlement behavior of stone column with circumferential nails.”
Le ¼ effective length of cemented stone column; Proc., Indian Geotechnical Conf. 2010, B. V. S. Vishwanadham and D.
N ¼ percentage fines; Choudhury, eds., Vol. 2, MacMillan, New Delhi, India, 579–582.
Q ¼ load; Barksdale, R. D., and Bachus, R. C. (1983). “Design and construction of
Qf ¼ failure load; stone columns.” FHWA Rep. No. RD. 83/026, Vol. 1, Federal Highway
Qt ¼ ultimate failure load; Administration, Washington, DC.
qf ¼ ultimate load-carrying capacity of untreated BIS (Bureau of Indian Standards). (1970). “Classification and identification
clay; of soils for general engineering purposes.” BIS-1498, New Delhi, India.
qf ¼ ultimate load-carrying capacity of untreated Castro, J. (2014). “Numerical modelling of stone columns beneath a rigid
clay; footing.” Comput. Geotech., 60, 77–87.
Castro, J., Karstunen, M., and Sivasithamparam, N. (2014). “Influence of
(qf)t ¼ ultimate load-carrying capacity of column-
stone column installation on settlement reduction.” Comput. Geotech.,
treated clay; 59, 87–97.
Rf ¼ shaft resistance; Chen, J. F., Li, L. Y., Xue, J. F., and Feng, S. Z. (2015). “Failure mechanism
Rs ¼ soil resistance; of geosynthetic-encased stone columns in soft soils under embank-
Rt ¼ tip resistance; ment.” Geotext. Geomembr., 43(5), 424–431.
ra ¼ area replacement ratio; Das, A. K., and Deb, K. (2014). “Modeling of uniformly loaded circular raft
rd ¼ bearing-capacity reduction factor; resting on stone column-improved ground.” Soils Found., 54(6),
rdt ¼ tip bearing-capacity reduction factor; 1212–1224.
rl ¼ length ratio; Datye, K. R., and Nagaraju, S. S. (1977). “Installation and testing of
S ¼ settlement; rammed stone columns.” Proc., 5th Asian Regional Conf. on Soil
Mechanics and Foundation Engineering, McGraw-Hill, New York,
s ¼ spacing of columns;
Vol. 1, 101–104.
w ¼ water content; Deb, K. (2010). “A mathematical model to study the soil arching effect in
wL ¼ liquid limit; stone column-supported embankment resting on soft foundation soil.”
Ze ¼ effective depth of soil; Appl. Math. Modell., 34(12), 3871–3883.
Zt ¼ depth of zone of probable bulking of column; Deb, K., Basudhar, P. K., and Chandra, S. (2010). “Extensible geosynthetics
z ¼ depth; and stone-column-reinforced soil.” Ground Improv., 163(4), 231–236.
a ¼ adhesion factor for cemented stone column and Deb, K., Samadhiya, N. K., and Jagtap, B. N. (2011). “Laboratory model
clay interface; studies on unreinforced and geogrid-reinforced sand bed over stone
DH ¼ change in height of clay sample in consolidation column-improved soft clay.” Geotext. Geomembr., 29(2), 190–196.
test; Duncan, J. M., and Chang, C. Y. (1970). “Non-linear analysis of stress and
strain in soil.” J. Soil Mech. Found. Div., 96(5), 1629–1653.
Dz ¼ change in thickness;
Dutta, S., Padade, A. H., and Mandal, J. N. (2012). “Experimental study on
d ¼ vertical downward displacement of cemented natural bamboo geogrid encased stone column.” Proc., 5th Asian
stone column; Regional Conf. on Geosynthetics, Geosynthetics Asia 2012, Indian
« v ¼ axial (vertical) strain; Geosynthetics Society, New Delhi, 417–426.
« vz ¼ axial (vertical) strain in column at depth z; Farouk, A., and Shahien, M. M. (2013). “Ground improvement using soil–
h ¼ stress concentration factor; cement columns: Experimental investigation.” Alexandria Eng. J.,
r ¼ uniform settlement of unit cell; 52(4), 733–740.

© ASCE 04016100-17 Int. J. Geomech.

Int. J. Geomech., 04016100


Fattah, M. Y., Shlash, K. T., and Al-Waily, M. J. (2013). “Experimental Murugesan, S., and Rajagopal, K. (2006). “Geosynthetic-encased stone col-
evaluation of stress concentration ratio of model stone columns strength- umns: Numerical evaluation.” Geotext. Geomembr., 24(6), 349–358.
ened by additives.” Int. J. Phys. Model. Geotech., 13(3), 79–79. Ng, K. S., and Tan, S. A. (2014). “Design and analyses of floating stone col-
Gniel, J., and Bouazza, A. (2008). “Improvement of soft soil using geogrid umns.” Soils Found., 54(3), 478–487.
encased stone column.” Geotext. Geomembr., 27(3), 167–175. Nayak, S., Shivashankar, R., and Babu, M. R. D. (2011). “Performance of
Golait, Y. S., Satyanarayana, V., and Raju, S. S. V. (2009). “Concept of stone columns with circumferential nails.” Ground Improv., 164(2),
under reamed cemented stone columns for soft clay ground improve- 97–106.
ment.” Proc., Indian Geotechnical Conf. 2009, Vol. 1, Allied, Mumbai, Prandtl, L. (1920). Uber die Harte Plastischer Korper, Nachrichten van
India, 356–360. der Koniglichen Gesellschaft der Wissenschaften zu Gottingen,
Gueguin, M., Hassen, G., and de Buhan, P. (2015). “Stability analysis of ho- Berlin, 74–85.
mogenized stone column reinforced foundations using a numerical yield Raithel, M., Kempfert, H. G., and Kirchner, A. (2002). “Geotextile-encased
Downloaded from ascelibrary.org by Missouri University of Science and Technology on 09/15/16. Copyright ASCE. For personal use only; all rights reserved.

design approach.” Comput. Geotech., 64, 10–19. columns for foundation as a dike on very soft soils.” Proc., 7th Int. Conf.
Hanna, A., Etezad, M., and Ayadat, T. (2013). “Mode of failure of a group on Geosynthetics, Indian Geosynthetics Society, New Delhi, India.
of stone columns in soft soil.” Int. J. Geomech., 10.1061/(ASCE)GM Ranjan, G., and Rao, B. G. (1983). “Skirted granular piles for ground
.1943-5622.0000175, 87–96. improvement.” Proc., 8th European Conf. on Soil Mechanics and
Hassen, G., Gueguin, M., and de Buhan, P. (2013). “A homogenization Foundation Engineering, CRC Press, Boca Raton, FL.
approach for assessing the yield strength properties of stone column re- Rao, N. B. S., and Nayak, I. G. (1995). “Model studies on partially con-
inforced soils.” Eur. J. Mech. A. Solids, 37, 266–280. fined sand columns using geogrid tubes.” Indian Geotech. J., 25(3),
Hughes, J. M. O., and Withers, N. J. (1974). “Reinforcing of soft cohesive 365–378.
soils with stone columns.” Ground Eng., 7(3), 42–49. Samadhiya, N. K., Maheshwari, P., Basu, P., and Kumar, M. B. (2008).
Jamsawang, P., Voottipruex, P., Boathong, P., Mairaing, W., and “Load settlement characteristics of granular piles with randomly mixed
Horpibulsuk, S. (2015). “Three-dimensional numerical investigation on fibres.” Proc., Indian Geotechnical Conf. 2008, Vol. 3, Indian
lateral movement and factor of safety of slopes stabilized with deep Geosynthetics Society, New Delhi, India, 345–354.
cement mixing column rows.” Eng. Geol., 188, 159–167. Shahu, J., and Reddy, Y. (2011). “Clayey soil reinforced with stone column
Juran, I., and Riccobono, O. (1991). “Reinforcing soft soils with artifi- group: Model tests and analyses.” J. Geotech. Geoenviron. Eng., 10
cially cemented compacted sand columns.” J. Geotech. Eng., .1061/(ASCE)GT.1943-5606.0000552, 1265–1274.
10.1061/(ASCE)0733-9410(1991)117:7(1042), 1042–1060. Skempton, A. W. (1951). The bearing capacity of clays, Building Research
Killeen, M. M., and McCabe, B. A. (2014). “Settlement performance of pad Congress, Institute of Civil Engineers sDiv-I, London.
footings on soft clay supported by stone columns: A numerical study.” Terzaghi, K. (1943). Theoretical soil mechanics, John Wiley and Sons,
Soils Found., 54(4), 760–776. New York, Inc., 118–143.
Madhav, M. R. (1983). “Recent developments in the use and analysis of Yoo, C. (2010). “Performance of geosynthetic-encased stone columns in
granular piles.” Proc., Int. Symp. on Recent Developments in Ground embankment construction: Numerical investigation.” J. Geotech.
Improvement Techniques, A. A. Balkema, Rotterdam, Netherlands, Geoenviron. Eng., 10.1061/(ASCE)GT.1943-5606.0000316, 1148–1160.
117–129. Yoo, C. (2015). “Settlement behavior of embankment on geosynthetic-
Maher, M. H., and Gray, D. H. (1990). “Static response of sand reinforced encased stone column installed soft ground—A numerical investiga-
with randomly distributed fibres.” J. Geotech. Eng., 10.1061/(ASCE)0733 tion.” Geotext. Geomembr., 43(6), 484–492.
-9410(1990)116:11(1661), 1661–1677. Zhang, L., and Zhao, M. (2014). “Deformation analysis of geotextile-
Mayerhof, G. G. (1951). “The ultimate bearing capacity of foundations.” encased stone columns.” Int. J. Geomech., 10.1061/(ASCE)GM.1943
Geotechnique, 2(4), 301–331. -5622.0000389, 04014053.
Mitchell, J., and Huber, T. (1985). “Performance of a stone column founda- Zhang, L., Zhao, M., Shi, C., and Zhao, H. (2013). “Settlement calculation
tion.” J. Geotech. Eng., 10.1061/(ASCE)0733-9410(1985)111:2(205), of composite foundation reinforced with stone columns.” Int. J.
205–223. Geomech., 10.1061/(ASCE)GM.1943-5622.0000212, 248–256.

© ASCE 04016100-18 Int. J. Geomech.

Int. J. Geomech., 04016100

You might also like