You are on page 1of 15

Technical Note

Stability Analysis of Geocell-Reinforced Slopes Using the


Limit Equilibrium Horizontal Slice Method
Iman Mehdipour, S.M.ASCE1; Mahmoud Ghazavi2; and Reza Ziaie Moayed3
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Given the three-dimensional confinement, the geocell reinforcement can act as a beam that can carry both bending and membrane
stresses compared with only the planar membrane effect of planar reinforcement. This study presents an analytical approach to determine the
factor of safety (FOS) of geocell-reinforced slope using the horizontal slice method (HSM). The geocell reinforcement composed of geocell
and infill soil was modeled as a composite beam filled with soil. The improvement in the FOS of the geocell-reinforced slope was expressed in
terms of vertical stress dispersion, vertical frictional resistance, and structural mechanisms provided by geocell reinforcement. The effect of
geocell pocket diameter on the soil confinement and FOS of geocell-reinforced slope was considered using hoop tension theory. The results of
the proposed analytical model were compared with those determined by numerical modeling using the strength reduction method (SRM). A
series of parametric analyses were also performed to evaluate the influence of the placement depth of the geocell layer, the number of geocell
layers, the pocket diameter and the height of geocell, and the slope angle and soil shear strength, on the stability of the geocell-reinforced slope.
The results indicate that the reinforcing mechanism of geocell reinforcement is considerably related to the geocell thickness. The mobilized
flexural strength and vertical frictional resistance magnify as the height of the geocell increases. Compared with planar reinforcement, a
smaller quantity of geocell reinforcement is required to achieve an equivalent FOS value. The results predicted from the present analytical
approach were found to be in good agreement with SRM results. DOI: 10.1061/(ASCE)GM.1943-5622.0000935. © 2017 American Society
of Civil Engineers.
Author keywords: Composite beam model; Geocell; Horizontal slice method (HSM); Numerical modeling; Planar reinforcement;
Slope stability.

Introduction Sitharam 2015b; Krishnaswamy et al. 2000; Latha and Murthy


2007; Leshchinsky and Ling 2013b; Sireesh et al. 2009; Sitharam
Generally, geosynthetic reinforcements are used to enhance the ten- and Hegde 2013; Tafreshi and Dawson 2010, 2012; Zhang et al.
sion resistance of the soil and reduce the soil deformation. In recent 2010a). Given the 3D confinement, the behavior of a geocell mat-
years, various research studies have been undertaken to evaluate the tress can be similar to a foundation beam or plate, which can carry
performance of reinforced walls and slopes, including the experi- both bending and membrane stresses (Dash et al. 2001a, 2007;
mental study (Chen et al. 2007; El-Emam and Bathurst 2007; Pokharel et al. 2010; Tang and Yang 2013; Thallak et al. 2007;
Kazimierowicz-Frankowska 2005; Latha and Krishna 2008; Yang Zhang et al. 2009, 2010b). For example, Dash et al. (2007) reported
et al. 2009), numerical modeling (Hatami and Bathurst 2000; that the geocell mattress behaves as a flexural member in which
Huang and Wu 2006; Rowe and Ho 1998; Rowe and Skinner deep beam behavior becomes predominant as the height of the geo-
2001), and analytical models (Ahmadabadi and Ghanbari 2009; cell increases. Zhang et al. (2009, 2010b) modeled the geocell-
Ghanbari and Taheri 2012; Nouri et al. 2006, 2008; Shahgholi reinforced mattress as a beam on a Winkler foundation to analyze
2001; Shekarian et al. 2008). The more recent advancement of soil the deformation behavior of the foundation. The authors found that
reinforcement is to provide three-dimensional (3D) confinement by the modulus and the height of the geocell contribute to the rigidity
using geocell reinforcements. The geocell mattress consists of a se- of the beam, which can result in a lower soil deformation compared
ries of interlocking cells that can effectively confine the soil within with the planar reinforcement. Dash (2010) pointed out that the geo-
its pockets. The effectiveness of using geocell to enhance load- cell mattress acts as a wide slab that transmits the footing pressure
carrying capacity of the soil and reduce the footing settlement has to the underlying soil and redistributes over a wider area. Tang and
been thoroughly investigated (Dash et al. 2003, 2004; Hegde and Yang (2013) investigated the flexural behavior of the geocell rein-
forcement by means of three-layered beam model testing. In the an-
1
Graduate Research Assistant, Dept. of Civil, Architecture, and alytical study, the geocell reinforcement was modeled as a continu-
Environmental Engineering, Missouri Univ. of Science and Technology, ous beam sandwiched between two identical polywood beams.
Rolla, MO 65409 (corresponding author). E-mail: iman.mehdipour@mst Their results showed that the geocell reinforcement provides con-
.edu siderable resistance to flexural deformation, especially for the
2
Professor, Faculty of Civil Engineering, K.N. Toosi Univ. of higher height-to-diameter ratio (aspect ratio) of a geocell. Zhang
Technology, Tehran 1969764499, Iran. E-mail: ghazavi_ma@kntu.ac.ir et al. (2009) reported that the lateral resistance, vertical stress dis-
3
Associate Professor, Faculty of Civil Engineering, Imam Khomeini persion, and membrane effects are the main characteristics of geo-
International Univ., Qazvin 3414916818, Iran. E-mail: ziaie@eng.ikiu.ac.ir
Note. This manuscript was submitted on August 31, 2016; approved
cell reinforcement that can lead to a reduction in the deformation of
on February 8, 2017; published online on April 24, 2017. Discussion pe- embankment.
riod open until September 24, 2017; separate discussions must be submit- In addition to the improvement in the bearing capacity, the geo-
ted for individual papers. This technical note is part of the International cell reinforcement can be effectively used to reduce deformation of
Journal of Geomechanics, © ASCE, ISSN 1532-3641. the slope and walls. For example, Ling et al. (2009) reported that

© ASCE 06017007-1 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


the use of geocells resulted in a reduction in the lateral deformation beam filled with soil by considering its main characteristics, includ-
of gravity walls. Mehdipour et al. (2013) conducted numerical anal- ing vertical stress dispersion, vertical frictional resistance, and
yses to investigate the stability of geocell-reinforced slope using the structural mechanisms. The analytical model was validated with
strength reduction method (SRM). The authors found that the geo- SRM numerical analysis performed by Mehdipour et al. (2013). A
cell reinforcement acts as a slab or beam at which the mobilized series of comparative analyses were performed to evaluate the influ-
flexural resistance of geocell reinforcement can reduce the lateral ence of various factors, such as the placement depth of the geocell
deformation of the slope. Therefore, under the geocell placement, layer, the number of geocell layers, the pocket diameter and the
the lateral deformation and shear strain values of the slope decrease. height of geocell, and the slope angle and soil shear strength proper-
Chen et al. (2013) indicated that an increase in the geocell length at ties, on the FOS of geocell-reinforced slope.
certain depths of the retaining structure can result in lower wall
movement and settlement of the backfill. Latha (2000) proposed the
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

equivalent composite approach to design the geocell-reinforced Analytical Analysis and Assumptions
embankment at which the geocell reinforcement is modeled as a
soil layer with additional cohesive strength and stiffness. A sum- In the present study, the FOS of the geocell-reinforced slope was
mary of the research studies performed to examine the performance determined using the limit equilibrium HSM proposed by
of geocell reinforcement for different geotechnical applications is Shahgholi (2001). In this technique, the sliding wedge is divided
presented in Table 1. Despite numerous studies aiming to evaluate into a number of horizontal slices that do not intersect the reinforce-
the bearing capacity and settlement behavior of geocell-reinforced ments; hence, the reinforcements have no direct influence on the
soil, the potential benefit of using geocell reinforcement to improve interslice forces. A number of different HSM formulations have
the slope stability has not been well documented (Bush et al. 1990; been investigated by Nouri et al. (2006, 2008) to analyze the planar-
Chen et al. 2013; Cowland and Wong 1993; Krishnaswamy et al. reinforced slope. Based on their results, the (2n þ 1) formulation
2000; Latha 2000; Latha et al. 2006; Latha and Rajagopal 2007; (where n is the number of horizontal slices) by considering moment
Mehdipour et al. 2013; Sitharam and Hegde 2013; Song et al. equilibrium equation for the sliding wedge can provide acceptable
2014). results with adequate accuracy (Nouri et al. 2006). Therefore, in the
The subject of this study is to examine the effectiveness of using study reported in this paper, the (2n þ 1) formulation was used to
geocell reinforcement to improve the stability of geocell-reinforced determine the FOS of the geocell-reinforced slope. The list of equa-
slope. A simple analytical model was proposed to determine the fac- tions and unknowns of the (2n þ 1) formulation is presented in
tor of safety (FOS) of slope reinforced with geocell reinforcement Table 2.
using the horizontal slice method (HSM). In the proposed analytical A log spiral failure mechanism was used to model the potential
approach, the geocell reinforcement was modeled as a composite failure surfaces of the geocell-reinforced slope. In this method, the

Table 1. Summary of Investigations on the Application of Geocell Reinforcement for Various Geotechnical Applications

Numerical modeling
Equivalent composite
Scope of study Experimental study Analytical study 3D numerical model model Beam model
Foundation (Biswas et al. 2016; Dash 2012; (Hegde and Sitharam (Dash et al. 2003; Han (Latha et al. 2008, (Tang and Yang
systems Dash et al. 2001a, b, 2003, 2004; 2015a–c, Tafreshi et al. et al. 2008; Hegde and 2009) 2013; Zhang et al.
Dash and Bora 2013; Latha and 2015; Tang and Yang Sitharam 2015c; Latha 2009, 2010b, 2012)
Murthy 2007; Latha and 2013; Zhang et al. 2009, and Somwanshi 2009;
Somwanshi 2009; Ling et al. 2010b, 2012) Saride et al. 2009; Yang
1997; Pokharel et al. 2010; et al. 2010)
Rajagopal et al. 1999; Sireesh
et al. 2009; Tafreshi and Dawson
2010, 2012; Yang et al. 2012;
Zhou and Wen 2008)
Subballast and (Biabani et al. 2016a, b; (Indraratna et al. 2015) (Biabani et al. 2016a, b; — —
railway Indraratna et al. 2015; Leshchinsky and Ling
Leshchinsky and Ling 2013a) 2013a, b)
Protection of (Bathurst and Knight 1998; — (Hegde and Sitharam — (Bathurst and
buried pipeline Hegde et al. 2014; Hegde and 2015d; Tavakoli Knight 1998)
Sitharam 2015d; Mehrjardi et al. Mehrjardi et al. 2015)
2012, 2013; Sireesh et al. 2009;
Tavakoli Mehrjardi et al. 2015)
Retaining wall (Chen and Chiu 2008; Chen et al. — (Chen et al. 2013) — —
2013; Leshchinsky et al. 2009;
Ling et al. 2009; Song et al.
2014; Xie and Yang 2009)
Slope stability (Bush et al. 1990; Cowland and (Sitharam and Hegde — (Latha 2011; Latha et (Mehdipour et al.
and supporting Wong 1993; Krishnaswamy et al. 2013; Zhang et al. al. 2006; Latha and 2013)
embankment 2000; Latha et al. 2006; Sitharam 2010a) Rajagopal 2007)
and Hegde 2013; Zhang et al.
2010a)

© ASCE 06017007-2 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Table 2. List of Equations and Unknowns in the ð2n þ 1Þ Formulation respectively. t t(x) and t b(x) represent the interface shear strengths
at the top and bottom of the geocell layer, respectively. Si and Ni
Unknowns Number Equations Number
P are the shear and normal forces acting on base of the ith slice,
Normal forces at the n Fy ¼ 0 (for each slice) n respectively, as indicated in Fig. 1(a).
base of each slice (Ni) Given 3D confinement, the geocell mattress can provide a beam
t
Shear forces at the base n t r ¼ FOS
f
(along the base n or slab mechanism compared with only the tensile membrane effect
of each slice (Si) of each slice) for planar reinforcement (Mehdipour et al. 2013; Tang and Yang
P
FOS 1 M ¼ 0 (for entire 1 2013; Zhang et al. 2009, 2010b). In the study presented in this pa-
wedge) per, the geocell reinforcement composed of soil and geocell mat-
Summation 2n þ 1 — 2n þ 1 tress was assumed to act as a composite beam filled with soil that is
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

embedded in the slope. It is important to note that the composite


beam model used in this study to simulate the geocell reinforcement
was different from the equivalent composite approach at which geo-
cell reinforcement is modeled as a soil layer alone with additional
cohesive strength and stiffness. In this study, the performance
improvement of geocell-reinforced slope is primarily attributed to
three characteristics: vertical stress dispersion, vertical frictional re-
sistance, and structural mechanisms. The structural performance of
the geocell reinforcement refers to the mobilized tensile and flexural
resistances. A detailed elaboration of these three characteristics of
geocell reinforcement is discussed in the following section.

Characteristics of Geocell Reinforcement

Vertical Stress Dispersion Mechanism


The 3D geocell-soil composite can behave as a rigid slab (Pokharel
et al. 2010; Tang and Yang 2013). This can lead to a redistribution
of the applied load over a wider area, resulting in a reduction in
pressure induced by surcharge loads acting on the slope. The sche-
matic view of the vertical stress dispersion mechanism of geocell
reinforcement is illustrated in Fig. 2. As a result of the vertical stress
dispersion effect, a value of transferred pressure of P(y) at the bot-
tom of the geocell-soil composite is lower than q(y) acting on the
superface of the geocell layer. The influence of the vertical stress
Fig. 1. Schematic view of the (a) log spiral failure mechanism and dispersion mechanism of geocell reinforcement on the reduction of
(b) acting forces on geocell reinforcement pressure for underlying soil can be expressed as follows:

b
PðyÞ ¼ qy (2)
failure surface is defined by a center point (Point O) and an initial ðb þ 2hg tan λÞ ð Þ
radius (r0), as shown in Fig. 1(a). The radius of the spiral (r) varies
with the angle of rotation (u ), with respect to the center of the spiral,
where b = width of the acting pressure q(y) above the geocell layer
as follows: caused by the surcharge load; l = load dispersion angle within geo-
cell reinforcement, which typically varies between 30 and 60°
r ¼ r0 eu tan f (1)
(Dash 2012); and hg = height of geocell reinforcement. It is impor-
tant to note that the vertical stress dispersion mechanism of geocell
where f = internal friction angle of soil. For a homogenous slope reinforcement is mobilized when geocell-reinforced slope is sub-
(which is the case in this study), the log spiral failure mechanism is jected to a surcharge loading. If there are no surcharge loads, then
theoretically the widely used limit equilibrium method to determine the vertical stress dispersion mechanism associated with geocell
the critical failure surface (Duncan and Wright 2005; Leshchinsky reinforcement can be ignored.
and Boedeker 1989). The base of the slope was assumed to be hard
and dense; hence, the failure surface passes through the toe of the
slope, as shown in Fig. 1(a). A spreadsheet solver tool was devel- Vertical Frictional Resistance Mechanism
oped to determine the critical failure surface in which u and r0 pa- The 3D behavior of geocell reinforcement can arrest the lateral
rameters were adjusted to achieve the minimum possible FOS value spreading and provide all around confinement to the infill soil. The
for a given slope. active earth pressure on the cell wall generates hoop stress within
Fig. 1(b) demonstrates the acting forces on the geocell- the wall and passive earth pressure on the adjacent walls
reinforced slope. Tig and Mig refer to the tensile force and bending (Emersleben and Meyer 2009). The confinement effect of the geo-
moment, respectively, of the ith geocell reinforcement. Qig is the cell is illustrated in Fig. 3(a). The confinement effect of the geocell
vertical frictional resistance between infill soil and wall of the ith is based on three mechanisms: active earth pressure within the cell
geocell reinforcement. Because of the overburden pressure and (s 3), additional confining pressure due to passive earth pressure in
possible surcharge loads, the pressures q(y) and P(y) act on the the adjacent cells (Ds 3), and the hoop stress (s c) within the cell
superface and the subgrade reaction of the geocell layer, wall (Hegde and Sitharam 2015a). The additional confinement

© ASCE 06017007-3 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


tensile stress-strain response. The additional confinement of the geo-
cell layer is significantly influenced by the geocell diameter; a geocell
with a lower pocket diameter results in a higher confining pressure.
When infill soil comes in contact with the geocell wall, vertical fric-
tional resistance will be mobilized between the infill soil and the geocell
wall, as indicated in Fig. 3(b). The vertical frictional resistance (Qig)
between the geocell wall and infill soil can be expressed as follows:

Qig ¼ t ig  hg ¼ ðs 3  Ds 3 Þtan d  hg
¼ ðka g zig  Ds 3 Þtan d  hg (4)
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Schematic illustration of the vertical stress dispersion effect of


geocell reinforcement where Qig = vertical frictional resistance of the ith geocell reinforce-
ment; t ig = vertical frictional strength between infill soil and wall of
the ith geocell reinforcement; s 3 = active earth pressure exerted by
the infill soil on the geocell wall; and d = interface friction angle
between geocell wall and infill soil. In this study, d was assumed to
be 2/3 of the soil friction angle ( f ). This value is consistent with
other research studies (Biabani et al. 2016a; Leshchinsky and Ling
2013b; Saride et al. 2009; Yang et al. 2010). zig and g = placement
depth of the ith geocell reinforcement and unit weight of soil,
respectively; and ka = coefficient of active earth pressure, which is
calculated as follows:

ka ¼ tan 2 ð45  f =2Þ (5)

Structural Mechanism
Because of the shear and bending rigidity, the behavior of a geocell
reinforcement is similar to a beam filled with soil (Dash et al. 2007;
Pokharel et al. 2010; Thallak et al. 2007). The force analysis of
geocell-soil composite is illustrated in Fig. 4(a). The tensile force is
resisted by the geocell mattress, and this force is positioned in the mid-
dle height of the geocell reinforcement, regardless of the geocell bend-
ing moment. The mobilized tensile force in the ith geocell layer (Tig)
can be determined using a liner distribution of forces in the reinforce-
ments proposed by Ling et al. (1997). The assumption of triangular
distribution for reinforcement loads along the height of the slope has
been commonly used to evaluate the stability of reinforced slopes and
walls (Ahmadabadi and Ghanbari 2009; Ehrlich and Mirmoradi 2013;
Ghanbari and Taheri 2012; Hatami and Bathurst 2000; Nouri et al.
2006, 2008). In accordance with the AASHTO (1998) simplified
method, the mobilized tensile force in reinforcement can be estimated
as follows:
  
zt þ zb
Tig ¼ Sv ðs a Þave ¼ Sv ka g (6)
2

where Sv = vertical spacing between reinforcements; and (s a)ave =


Fig. 3. Schematic illustration of vertical frictional resistance mecha- average of horizontal soil stresses in the ith slice between depths of zt
nism of geocell reinforcement for (a) the confinement effect of geocell and zb, as indicated in Fig. 4(b). The geocell-soil composite at the slip
and (b) the mobilization of vertical frictional resistance surface is subjected to displacement, as shown in Fig. 1(b). The dis-
placement of geocell can lead to the mobilization of the bending
moment in the geocell-soil composite. Because of the flexural defor-
effect offered by the geocell can be determined using the hoop ten- mation induced in the geocell and assuming the neutral axis in the mid-
sion theory proposed by Henkel and Gilbert (1952). By considering dle height of the geocell (hg/2), the top and bottom halves of the com-
Hooke’s law, the additional confining pressure (Ds 3) provided by posite beam filled with soil are subjected to compression and tension,
the geocell can be obtained as follows: respectively. Because of the tensionless nature of soil, it is hypothe-
 pffiffiffiffiffiffiffiffiffiffiffiffiffi sized that the flexural deformation is resisted by the bending moment
2s c 2M 1  1  ɛa resulting from the tensile force provided by the beam element (i.e.,
Ds 3 ¼ ¼ (3)
dg dg 1  ɛa geocell mattress) and compressive stresses due to the infill soil, as illus-
trated in Fig. 4(c). Because soil is typically capable of resisting com-
where s c and dg = circumferential stress (hoop stress) in the geocell pressive stresses, the mobilized (s 3 − Ds 3) distribution is acting only
and the diameter of geocell pocket, respectively; and M = secant mod- in the upper half of the geocell (i.e., above the neutral axis). This
ulus of geocell material corresponding to the axial strain of « a in the behavior of soil is also referred to as a tensionless foundation

© ASCE 06017007-4 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Schematic illustration of the structural mechanism of geocell reinforcement for (a) the force analysis in geocell-soil composite, (b) the deter-
mination of tensile force in geocell reinforcement using internal equilibrium, and (c) the determination of the bending moment in geocell reinforce-
ment using internal equilibrium

( "  pffiffiffiffiffiffiffiffiffiffiffiffi#)
(Kaschiev and Mikhajlov 1995; Zhang and Murphy 2004). Therefore, hg 2M 1 1ɛa hg hg
the resisting bending moment (Mig) can be determined using the tensile Mig ¼ Tig  þ ka g zig    (8)
8 dg 1ɛa 2 4
stresses along the cell wall of the geocell mattress and compressive
stresses resulting from the infill soil with respect to the neutral axis of
the geocell-soil composite beam. The bending moment in the geocell- Substituting the value of Tig determined from Eq. (6) into Eq.
soil composite can be obtained using internal equilibrium, as follows: (8), the bending moment can be determined.
Tig hg hg hg hg
Mig ¼   þ ðs 3  Ds 3 Þ   (7)
hg 2 4 2 4 Determination of FOS Using Limit Equilibrium HSM

where Tig and Mig = tensile force and bending moment in the ith Fig. 5 plots the acting forces on the ith horizontal slice containing
geocell reinforcement, respectively. Substituting Eq. (3) into Eq. geocell reinforcement. The reinforcing mechanism of geocell rein-
(7), the resisting bending moment provided by the geocell-soil com- forcement and the (2n þ 1) HSM formulation proposed by Nouri et
posite can be expressed as follows: al. (2006) were used to determine the FOS of the geocell-reinforced

© ASCE 06017007-5 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Schematic illustration of the anchorage pull-out length of the


geocell reinforcement

Fig. 5. Forces acting on the single horizontal slice containing geocell X


reinforcement MO ¼ 0 ðfor entire wedgeÞ
X
m X
m X
m X
n
) Tig ðRTi Þ þ Mig þ Qig ðRQi Þ  Wi ðRwi Þ
slope. In the (2n þ 1) HSM formulation approach, the vertical i¼1 i¼1 i¼1 i¼1
forces equilibrium for each horizontal slice, moment equilibrium X
n
for the entire sliding mass, and Mohr-Coulomb failure criteria along þ ½ðSi sin ai þ Ni cos ai ÞðXNS;O Þ
the base of each slice are applied, as presented in Table 2. The fol- i¼1
lowing assumptions were made to determine the FOS of the þ ðSi cos ai  Ni sin ai ÞðYNS;O Þ ¼ 0 (11)
geocell-reinforced slope:
1. The potential failure surface is log spiral and it passes through where n and m = number of horizontal slices and geocell reinforce-
the toe of the slope. This method is limited to homogeneous ments, respectively; RTi, RQi, and Rwi = coordinates of geocell ten-
slopes. sile force, vertical frictional resistance, and weight of the ith slice
2. Surcharge loading is not considered on the slope. with respect to the slip surface center; and YðNS;OÞi and XðNS;OÞi =
3. Ni and Si act at the middle of base of the ith slice. coordinates of the point, where Ni and Si act on the base of the ith
4. Geocell-soil composite reinforcement behaves as a composite slice, which are calculated as follows:
beam, and the location of neutral axis is assumed to be in the
middle of geocell height. hi
XðNS;OÞi ¼ ri cos u i  (12)
5. The distribution of tensile forces in geocell layers is linear 2 tan ai
along the height of the slope.
6. A minimum of three slices is needed to determine the FOS with hi
satisfactory accuracy. YðNS;OÞi ¼ ri sin u i þ (13)
2
To determine the FOS of the geocell-reinforced slope, the
vertical equilibrium equation for each slice and Mohr-Coulomb where ri, u i, and hi = radius of log spiral, angle of rotation with
failure criteria along the base of each slice were satisfied as respect to horizontal axis, and height of the ith slice, respectively.
follows: The Mohr-Coulomb failure criteria along the base of each slice
X are a function of Ni and Si. By substituting Eq. (10) into Eq. (9), the
Fy ¼ 0 ðfor each sliceÞ value of Ni is derived as a function of FOS. Then, by substituting
) Viþ1  Vi  Wi þ Si sin ai þ Ni cos ai þ Qig ¼ 0 Eq. (9) into Eq. (11), the value of FOS for a given geocell-
reinforced slope is determined. A detailed calculation of the FOS
(9) for a given geocell-reinforced slope is presented in the Appendix.
In all analyses presented in this study, the required length of geo-
cell reinforcement was adjusted to prevent pull-out failure of the
tf geocell reinforcement. The total geocell length required to prevent
tr ¼ ðalong the base of each sliceÞ
FOS pull-out failure is the summation of the anchorage length and length
falling within the shear zone. The total length of the ith geocell rein-
1 forcement can be expressed as follows:
) Si ¼ ðcli þ Ni tan f Þ (10)
FOS
LðigÞt ¼ LðigÞa þ LðiÞc (14)

where Vi, Wi, li, and ai = vertical interslice force, weight of the ith
slice, length of the ith slice base, and inclination angle of the slice Tig
LðigÞa ¼ (15)
with respect to the horizontal axis, respectively; c and f = cohesive 2 ms vi tan f
strength and internal friction angle of soil; and t f and t r = available
and required shear strengths, respectively. The axial tension of geo- where LðigÞt , LðigÞa , and LðiÞc = total length required to prevent pull-out
cell reinforcement (Tig) and vertical frictional resistance (Qig) were failure, anchorage length to mobilize tensile resistance, and length
included in the Ni and Si. The moment equilibrium for the entire located in the shear zone (distance between critical failure surface and
sliding wedge was then satisfied with respect to the slip surface cen- slope face) for the ith geocell reinforcement, respectively, as sche-
ter (Point O in Fig. 5), as follows: matically illustrated in Fig. 6; s vi = overburden pressure acting on the

© ASCE 06017007-6 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Model description of the geocell-reinforced slope used for SRM calculations

ith geocell layer; and m = soil-geocell interface coefficient, which typ- Table 3. Comparison between FOS Results Obtained by the Analytical
ically varies from 0.6 to 0.8 (Biabani et al. 2016a; Leshchinsky and Method and SRM
Ling 2013b; Saride et al. 2009; Yang et al. 2010). Number of FOS difference
hg geocell FOS (analytical FOS between two
Case (m) layers method) (SRM) methods (%)
Numerical Modeling
1 — 0 1.21 1.24 2.48
In the present study, the FLAC2D program is used to examine the 2 0.02 3 1.49 1.64 10.07
stability of the geocell-reinforced slope and compare the results 3 0.02 4 1.68 1.73 2.98
with those determined from the proposed analytical approach. 4 0.02 5 1.74 1.79 2.87
FLAC has been used by several researchers (Cheng et al. 2007; Wei 5 0.02 6 1.82 1.87 2.75
and Cheng 2009, 2010) to investigate the FOS of the reinforced 6 0.02 7 1.90 1.96 3.16
slope using the SRM technique. In the numerical simulation, differ- 7 0.02 8 1.97 2.06 4.57
ent failure modes such as pull-out failure and tensile failure were 8 0.02 9 2.00 2.10 5.00
also assessed, ensuring that the FOS of the modeled geocell- 9 0.2 3 1.64 1.73 5.49
reinforced slope is not dominated by such failure modes. A 30 m 10 0.2 4 1.78 1.84 3.37
high and 50 m long grid was considered to model slope in plane 11 0.2 5 1.92 2.05 6.77
strain analysis, as illustrated in Fig. 7. The bottom and lateral boun- 12 0.2 6 2.02 2.11 4.46
daries of the slope domain were taken long enough from the rein-
forcement to avoid any boundary effects. The bottom boundary was
fixed against movements in all directions, whereas the vertical and soil, respectively; and Fn = normal force at the interface. The
boundaries were restrained in the horizontal direction. friction and adhesion of the geocell-soil interface for both vertical
The elastic-perfectly plastic Mohr-Coulomb model was used to and horizontal directions were considered to be 2/3 f and 2/3c,
simulate the behavior of the soil. Geocell mattress filled with soil pro- respectively. This assumption is similar to other existing research
duces a reinforced composite layer that acts as a beam with significant studies (Biabani et al. 2016a; Leshchinsky and Ling 2013b; Saride
improved structural properties. In this study, therefore, the geocell et al. 2009; Yang et al. 2010). As mentioned previously, the linearly
reinforcement composed of geocell and infill soil was modeled as a elastic plate element was used to model the geocell-soil rein-
composite beam infilled with soil to simulate the geocell reinforce- forcement. The normal (kn) and the shear (ks) stiffness of the com-
ment as a flexible slab that can carry both bending and membrane posite beam element were calculated as follows:
stresses (Mehdipour et al. 2013; Tang and Yang 2013; Zhang et al.  
2009, 2010b). The beam element is a two-dimensional element with K þ 4=3 G
kn ¼ ks ¼ 10  max (17)
three degrees of freedom (x-translation, y-translation, and rotation) at zmin
each end node. In this study, the composite beam was assumed to
behave as a linearly elastic material with both axial tensile and com- where K and G = bulk and shear moduli, respectively; and Zmin =
pressive failure limits. Because of the thickness, a moment of inertia smallest width of an adjoining zone in the normal direction.
was assigned to the composite beam, which in turn acts as a flexible The numerical simulation presented in this paper was verified
member that takes flexural moment. The input parameters for the through laboratory model tests for geocell-supported embankments
beam element in FLAC include thickness (hg), second moment of conducted by Krishnaswamy et al. (2000). In this verification, the
area (Ig), elastic modulus (Eg), axial peak tensile yield strength embankment was constructed above a layer of geocell on top of a
[Fyg(ten)], axial compressive yield strength [Fyg(comp)], and density soft clay bed with a 600-mm thickness and the unit weight of 17
( g g). The interface shear stress-strain relationships between soil- kN/m3. A uniform settlement was applied on the surface of the
geocell at the top and bottom of the geocell mattress as well as embankment by applying equal vertical displacement at all nodes
between the geocell wall and infill soil were modeled using on the crest of the embankment. The comparison between experi-
Mohr-Coulomb sliding criterion according to Eq. (3) mental and numerical results exhibited good agreement in terms of
Fsmax ¼ ci  A þ Fn tan f i (16) lateral deformation and surface heave of geocell-reinforced slope
modeled by the composite beam. A more detailed explanation of
where Fsmax = shear force at the geocell-soil interface; ci, f i, and the simulated geocell-reinforced slope and verification procedure
A = adhesion, angle of friction, and contact area between geocell can be found in Mehdipour et al. (2013).

© ASCE 06017007-7 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Verification of the Present Analytical Approach analytical method were found to be in good agreement with
those determined by numerical modeling using FLAC2D.
The proposed analytical approach was validated by the results
from numerical analysis performed by Mehdipour et al. (2013)
to examine the influence of geocell reinforcement on the FOS Parametric Studies
and critical failure surface of slopes using SRM analysis. In the
numerical modeling, the geocell-soil reinforcement was mod- A series of comparative analyses were performed to evaluate the
eled as a composite beam model filled with soil that can carry influence of geocell reinforcement on the FOS of slope. The investi-
both bending and membrane stresses. To model geocell reinforced gated parameters included placement depth of the first geocell layer
slope, a 30 m high and 50 m long grid with a slope angle of 30° from the top of the slope (u), number of geocell layers (m), geocell
was considered, as shown in Fig. 7. The cohesion, friction angle, diameter (dg), geocell height (hg), and geocell secant modulus (M).
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

unit weight, and elastic modulus of the soil were taken as 2 kPa, To simplify the representations of results, all parameters were
30°, 20 kN/m3, and 40 MPa, respectively. Two different geocell expressed in nondimensional terms with respect to the slope height:
heights of 0.02 and 0.2 m with a secant modulus of 100 kN/m u/H = depth ratio for the first reinforcement layer, dr/H = reinforced
and geocell pocket diameter of 0.2 m were considered in all zone ratio, and hg/H = geocell height ratio. The influence of slope
analyses. The placement depth of the first geocell layer-to-slope angle ( b ) and soil shear strength properties on the performance of
height ratio (u/H), the reinforced zone-to-slope height ratio the geocell-reinforced slope was also investigated. Furthermore, the
(dr/H), and geocell length-to-slope height ratio (L/H) were con- performances of geocell and planar reinforcements were compared
sidered as 0.2, 0.9, and 2.0, respectively. The comparison of to evaluate the effect of different forms of reinforcements on the
FOS values determined by the analytical model and SRM analy- FOS of reinforced slopes. The performance improvement in terms
ses is presented in Table 3. The results from the present of the increase in the FOS value, due to the inclusion of reinforce-
ment, was quantified using a nondimensional improvement factor
Table 4. Characteristics of Soil and Geocell Used in This Study (IF), which is expressed as follows:
Material Property Quantity FOSðreinforcedÞ
IF ¼ (18)
3 FOSðunreinforcedÞ
Soil Unit weight ( g ) 20 kN/m
Young’s modulus (E) 40 MPa where FOS(reinforced) and FOS(unreinforced) = FOS for reinforced and
Poisson’s ratio 0.3 unreinforced slopes, respectively. The soil properties and characteris-
Shear strength parameter c: 2, 10, and 20 kPa tics of geocell reinforcement used in this investigation are summarized
f : 10, 20, and 30 degrees in Table 4. To have a more realistic representation of geocell reinforce-
Slope angle ( b ) 30 and 60 degrees ment, the majority of values used in the parametric analysis (Table 4)
Geocell Height (hg) 1, 10, 100, 200, and 500 mm were approximated by numerical and experimental published studies
reinforcement Secant modulus (M) 100, 200, and 300 kN/m
performed on geocell reinforcement, such as Hegde and Sitharam
Axial strain (« a) 2%
(2015b), Latha et al. (2008, 2009), Latha et al. (2006), and Latha and
Interface friction angle 2/3 f
Rajagopal (2007). The values of FOS and location of critical failure
between geocell wall
surfaces determined by the analytical model were compared with SRM
and infill soil (d )
results. For SRM analyses, the geocell layer was modeled as a compos-
Pocket diameter (dg) 200 and 500 mm
ite beam using FLAC2D (Mehdipour et al. 2013).

Fig. 8. Variation in IF with placement depth of the first geocell layer for (a) m = 1 and (b) m = 3

© ASCE 06017007-8 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Effect of Placement Depth of the First Geocell Layer to 1.45, whereas a further increase in the placement depth of the first
geocell layer (u/H > 0.6) resulted in lower IF values from 1.45 to
The variation in IF with u/H for m = 1of the geocell-reinforced slope 1.28. A similar tendency was obtained for SRM analysis.
is shown in Fig. 8(a). The cohesion and friction angle of the soil For a given number of reinforcement layers, the use of planar rein-
were set at 2 kPa and 30°, respectively, and the slope angle was con- forcements resulted in lower IF values compared with the geocell rein-
sidered as 30°. M, « a, and dg were assumed to be 100 kN/m, 2%, forcement. For instance, in the case of m = 3 and u/H = 0.6, the geo-
and 200 mm, respectively. The IF value increased by increasing the cell-reinforced slope exhibited a 75% higher IF value compared with
placement depth of the geocell reinforcement up to a certain thresh- the planar-reinforced slope. In other words, compared with the planar
old value (u/H = 0.6), beyond which the IF sharply decreased. In the reinforcement, a lower quantity of geocell mattress is required to
case of u/H = 1, the inclusion of the geocell layer resulted in no achieve a similar improvement in the FOS of a given slope. The loca-
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

improvement in the FOS of the geocell-reinforced slope. In this tion of slip surfaces and maximum bending moment of geocell rein-
case, the geocell layer is located out of the effective zone and is not forcement obtained by the analytical method and SRM are illustrated
interested by the slip surface; thus, the reinforced slope behaves as in Fig. 9. Generally, good agreement was observed for the location of
an unreinforced slope. slip surfaces determined by the present method and SRM.
The variation in IF with u/H for m = 3 and dr/H = 0.9 is plotted in
Fig. 8(b). For m = 3, the second geocell layer was placed at the mid-
dle of first and third geocell reinforcements along the height of the Effect of Number and Thickness of Geocell Layers
slope. Depending on the placement depth of the first geocell layer,
the use of geocell resulted in 28–45% improvement in the IF values. The variation in IF with the number of reinforcement layers is plot-
As in the case of single geocell layer, for u/H < 0.6, the inclusion of ted in Fig. 10(a). The slope geometry and soil properties were set
three geocell layers led to a significant increase in the IF from 1.26 similar to the previous section. Generally, using a higher number of

Fig. 9. Comparison between the SRM and analytical method for u/H = 0.6

Fig. 10. Variations in IF with (a) number of reinforcement layers and (b) geocell thickness

© ASCE 06017007-9 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Table 5. Comparison between FOS Results for Different Soil Shear Strength Properties and Slope Angles

FOS difference between


c U FOS (unreinforced FOS (analytical FOS FOS (planar FOS difference between different reinforcement
Slope angle Case (kPa) (degrees) slope) method) (SRM) reinforcement) two methods (%) form (%)
b ¼ 30 1 2 10 0.49 0.91 0.97 0.72 6.67 20.99
2 2 20 0.88 1.41 1.52 1.13 7.32 19.56
3 2 30 1.28 1.92 2.05 1.69 6.34 13.61
4 10 10 0.85 1.13 1.26 1.00 10.29 11.48
5 10 20 1.30 1.60 1.75 1.35 8.47 15.78
6 10 30 1.73 1.99 2.13 1.72 6.57 15.70
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

7 20 10 1.21 1.53 1.65 1.31 7.30 14.17


8 20 20 1.71 1.96 2.18 1.75 9.79 10.61
9 20 30 2.11 2.35 2.53 2.02 7.11 16.34
b ¼ 60 10 2 10 0.28 0.68 0.79 0.58 14.12 14.86
11 2 20 0.48 0.97 1.06 0.77 8.70 20.43
12 2 30 0.61 1.58 1.72 1.08 8.14 46.30
13 10 10 0.53 0.86 0.97 0.68 11.43 21.12
14 10 20 0.73 1.16 1.25 0.98 7.41 15.23
15 10 30 0.93 1.49 1.66 1.25 10.24 19.20
16 20 10 0.81 1.17 1.28 0.96 8.70 17.52
17 20 20 1.15 1.33 1.51 1.19 11.66 10.85
18 20 30 1.28 1.81 1.96 1.47 7.65 23.13

geocell layers resulted in greater IF values. The comparison


between geocell and planar-reinforced slopes indicates that the
improvement in FOS is significantly affected by the form of
reinforcement. The difference in IF values for planar and geocell
reinforcements was shown to increase by increasing the number
of reinforcements. For example, in the case of m = 3, the geocell-
reinforced slope exhibited a 12% higher IF value compared with the
planar reinforcement, whereas an increase in the number of rein-
forcements from 3 to 10 resulted in a 16% higher FOS value. As
illustrated in Fig. 10(a), the number of planar reinforcements
required to achieve the IF value of 1.55 was eight layers compared
with the five layers for the geocell reinforcement. Fig. 10(b) shows
the variation in IF with geocell height. Geocell with a higher height
resulted in greater IF values. This is attributed to the mobilization of
vertical frictional resistance and structural mechanisms of geocell
reinforcement. For m = 3, an increase in the geocell height ratio from Fig. 11. Location of slip surfaces for geocell-reinforced slope for vari-
hg/H = 0.0001 to hg/H = 0.05 resulted in a 20% improvement in the ous soil shear strength properties
IF value. In the case of relatively low geocell height, the geocell rein-
forcement tends to behave as a thin beam, such as planar reinforce-
ment, as indicated in Fig. 10(b). cohesive strength of 20 kPa resulted in a 147% higher FOS value
compared with a slope with a cohesive strength of 2 kPa. The differ-
ence between geocell and planar reinforcements is considerably
Effect of Soil Shear Strength Properties and affected by varying soil shear properties, as presented in Table 5.
Slope Angle Depending on the cohesion and friction angle of the soil, the difference
in the FOS values for geocell and planar reinforcements varied from 10
Table 5 presents the FOS values determined by analytical method and to 46%.
SRM for a slope with various slope angles and soil shear strength Generally, good agreement was obtained for FOS values deter-
properties. The u/H, hg/H, and number of geocell layers were set at mined by the analytical method and SRM. The FOS values deter-
0.2, 0.02, and 5, respectively. The investigated soil shear strength mined by SRM analysis were 6–13% higher than those obtained by
properties included cohesive strength ranging between 2 and 20 kPa, the analytical method, especially in the case of a slope with higher
and the friction angle varied from 10 to 30°. The FOS of geocell- cohesion. The locations of the slip surfaces for the present method
reinforced slope was shown to be significantly affected by the soil and SRM for various soil shear strength properties are compared in
shear strength properties and slope angle. Generally, a higher friction Fig. 11. It is noted that the locations of critical failure surfaces for
angle and cohesion resulted in greater FOS values. For instance, in the SRM were obtained using FLAC2D. In the case of a slope with low
case of b = 30° and f = 10°, increasing cohesive strength from 2 to cohesion, the location of critical failure surface determined by the
20 kPa led to a 70% improvement in the FOS of geocell-reinforced SRM and present method were close, but a larger difference was
slope. The results of FOS for unreinforced slope are also presented in found for a slope with a higher cohesion, as indicated in Fig. 11. In
Table 5. The cohesive strength of soil had a significant influence on the case of c = 20 kPa and f = 10°, the upper endpoint of the failure
the FOS. For example, for b = 30°and f = 10°, a slope with a surface derived from SRM was 2.2 m away from the right endpoint

© ASCE 06017007-10 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


of the critical failure surface obtained by the analytical method. compared with the case with a lower cohesion and higher friction
Generally, by increasing the friction angle of the soil, the critical angle.
failure surface is shallower and the upper end of the critical failure
surface moves closer to the slope crest. On the other hand, by
increasing the cohesion of soil, the critical failure surface becomes Conclusions
deeper and the upper end of the slip surface is far away from the
slope crest. This can lead to a larger potential of failure volume An analytical model was proposed to determine the factor of safety
(FOS) of the geocell-reinforced slope using horizontal slice method
(HSM). In this approach, the geocell reinforcement composed of geo-
cell mattress and infill soil was modeled as a composite beam filled
with soil by considering its main characteristics, including vertical
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

stress dispersion, vertical frictional resistance, and structural mecha-


nisms. A series of parametric analyses were performed to evaluate
the influence of various key factors, such as placement depth of the
geocell layer, the number of geocell layers, the pocket diameter and
height of geocell, the slope angle, and the soil shear properties on the
FOS of geocell-reinforced slope. Based on the results presented in
this paper, the following conclusions can be drawn:
1. The results of FOS and locations of critical failure surfaces
determined by the analytical method and SRM were found to
be in good agreement. In most cases, the FOS values deter-
mined by the present analytical method were 3–12% lower than
those obtained by SRM analysis.
2. The form of reinforcement is shown to have significant influ-
ence on the stability of reinforced slope. For a given FOS value,
a lower quantity of geocell reinforcement is required compared
Fig. 12. Geocell-reinforced slope model used for the investigated with the planar reinforcement.
example 3. The reinforcing mechanism of geocell reinforcement is signifi-
cantly related to the geocell thickness. As the geocell height
elevates, the effectiveness of vertical frictional resistance and
Table 6. Parameters Used for the Investigated Example structural mechanisms to improve the stability of the reinforced
Soil properties Geocell characteristics slope increases.
3
4. The FOS of geocell-reinforced slope was found to be signifi-
g = 20 kN/m hg = 200 mm cantly affected by the soil shear strength properties and slope
f = 30° dg = 200 mm angle. Depending on the cohesion and friction angle of soil, the
c = 2 kPa « a = 2% difference in the FOS for geocell and planar reinforcements
H = 10 m M = 100 kN/m varied from 10 to 46%.
n=3 d = 2/3 f = 20 degrees 5. The geocell layer with a higher modulus and lower pocket
b = 60° m=3 diameter resulted in higher FOS values. A smaller pocket
size of the geocell offers higher confinement to the infill
materials.
Table 7. Soil Weights and Geocell Forces of Horizontal Slices for the
Although the results derived from this study provide encourage-
Investigated Example
ment for the use of the limit equilibrium method to estimate the FOS
Slice of geocell-reinforced slope, in a two-dimensional framework, this
number Wi (kN) Ds 3 (kPa) t ig (kPa) Tig (kN) Qig (kN) Mig (kN·m) method has certain limitations. The proposed analytical approach
1 332.12 10.25 0.52 82.50 0.11 11.22 for calculating the FOS of a geocell-reinforced slope underestimated
2 327.32 10.25 8.28 210.20 1.66 26.44 the stability of the geocell-reinforced slope compared with numeri-
3 212.83 10.25 16.08 347.10 3.21 51.13 cal modeling using the SRM approach; however, both methods
offered similar trends with changes in geocell characteristics. It

Fig. 13. (a) Free-body diagram and (b) force polygon of the second slice for investigated example (Note: All forces are expressed in kilonewtons)

© ASCE 06017007-11 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


should be noted that the present analytical/numerical results are c ¼ cohesion of soil;
based on the assumptions of using a composite beam model filled ci ¼ adhesion between geocell and soil;
with soil and fully mobilized tensile strength of geocell reinforce- dg ¼ diameter of geocell pocket;
ment. Furthermore, the effect of soil compaction and surcharge load- dr ¼ reinforced zone;
ing on the performance of geocell reinforcement were not taken into E ¼ Young’s modulus of soil;
consideration in this study. Construction procedure results in higher Eg ¼ elastic modulus of composite beam;
confining pressure at the lower part of the slope, leading to a pre- Fn ¼ normal force at interface between geocell
stressing of the reinforcement. Because of the vertical stress disper- and soil;
sion provided by the geocell, it is expected that the effect of sur- FOS(reinforced) ¼ factor of safety for reinforced slope;
charge loading on reduced slope stability is less pronounced for FOS(unreinforced) ¼ factor of safety for unreinforced slope;
geocell reinforcement compared with the planar-reinforced slope. Fsmax ¼ shear force at the geocell-soil interface;
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Moreover, the critical slip surface was assumed to be tangential to Fyg(comp) ¼ axial compressive yield strength of
the base, whereas in some cases slip surfaces can pass above the composite beam;
base or below the base (deep-seated failure). Finally, a better under- Fyg(ten) ¼ axial peak tensile yield strength of
standing of the limitations of the proposed limit equilibrium method composite beam;
would be helpful in developing a more comprehensive analytical G ¼ shear modulus of beam;
method that can accurately predict the stability of the geocell- H ¼ height of slope;
reinforced slope. Further research is required to validate the analyti- hg ¼ thickness of geocell reinforcement;
cal approach with full-scale and/or field test data to apply research hi ¼ height of the ith horizontal slice;
findings of this study to practical situations. Ig ¼ second moment of area of composite
beam;
IF ¼ improvement factor;
Appendix. An Example of Calculation
K ¼ bulk modulus of beam;
ka ¼ coefficient of active earth pressure;
A sample calculation for determining the FOS of a geocell-
kn ¼ normal stiffness of composite beam;
reinforced slope using the analytical model is provided in Eq. (19).
ks ¼ shear stiffness of composite beam;
A geocell-reinforced slope with three layers of geocell reinforce-
L(i)c ¼ length falling between slope face and
ment and slope height of 10 m were considered. The slope geometry
critical failure surface for the ith geocell
and location of the critical failure surface are illustrated in Fig. 12.
The soil properties and geocell characteristics used in this example reinforcement;
are listed in Table 6. The detailed calculation of geocell forces and L(ig)a ¼ anchorage length of the ith geocell
soil weights for each horizontal slice are presented in Table 7. By reinforcement;
substituting Eq. (10) into Eq. (9), the Ni associated with each hori- L(ig)t ¼ total length required to prevent pull-out
zontal slice can be derived as a function of the FOS as follows: failure for the ith geocell reinforcement;
li ¼ length of the ith slice base;
7:5 M ¼ secant modulus of geocell material;
88:2  Mig ¼ bending moment for the ith geocell
N1 ¼ FOS
0:44 reinforcement;
0:61 þ m ¼ number of geocell layers;
FOS
Ni ¼ normal force acting on the base of the ith
5:12
174:6  slice;
N2 ¼ FOS n ¼ number of horizontal slices;
0:34 P(y) ¼ subgrade reaction at bottom of geocell
0:61 þ
FOS reinforcement;
3:4 Qig ¼ vertical frictional resistance for the ith
648  geocell reinforcement;
N3 ¼ FOS (19)
0:1 q(y) ¼ acting pressure on superface of geocell
0:98 þ reinforcement;
FOS
RTi and RQi ¼ coordinate of the ith geocell tensile force
The determination of geocell forces and derived equations for Ni and vertical frictional resistance with
(as s function of FOS) enables the calculation of the FOS for geo- respect to the slip surface center;
cell-reinforced slope using moment equilibrium for the entire slid- RWi ¼ coordinates of the weight of the ith slice
ing wedge [Eq. (11)]. For this particular example, the value of FOS with respect to the slip surface center;
was calculated to be 1.21. To validate the equilibrium of derived r ¼ radius of the log spiral;
forces, the free-body diagram and force polygon of the second slice ri ¼ radius of the log spiral for the ith slice;
are illustrated in Fig. 13. The force polygon closure indicated an r0 ¼ initial radius of the log spiral;
appropriate accuracy. Si ¼ shear force acting on the base of the ith
slice;
Notation Sv ¼ vertical reinforcement spacing;
Tig ¼ tensile force for the ith geocell
The following symbols are used in this paper: reinforcement;
A ¼ contact area between geocell and soil; u ¼ placement depth of the first geocell
b ¼ width of acting pressure on top of geocell reinforcement from top of slope;
reinforcement; Vi ¼ vertical interslice force of the ith slice;

© ASCE 06017007-12 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Wi ¼ weight of the ith slice; Chen, R. H., and Chiu, Y. M. (2008). “Model tests of geocell retaining struc-
YNS,O and XNS,O ¼ coordinates of the point where Ni and Si tures.” Geotext. Geomembr., 26(1), 56–70.
act on the base of the slice; Chen, R. H., Wu, C. P., Huang, F. C., and Shen, C. W. (2013). “Numerical
analysis of geocell-reinforced retaining structures.” Geotext. Geomembr.,
Zmin ¼ smallest width of an adjoining zone in
39(Aug), 51–62.
the normal direction; Cheng, Y. M., Lansivaara, T., and Wei, W. B. (2007). “Two-dimensional
zig ¼ placement depth of the ith geocell slope stability analysis by limit equilibrium and strength reduction meth-
reinforcement; ods.” Comput. Geotech., 34(3), 137–150.
Ds 3 ¼ additional confining pressure provided Cowland, J. W., and Wong, S. C. K. (1993). “Performance of a road
by geocell reinforcement; embankment on soft clay supported on a geocell mattress foundation.”
ai ¼ inclination angle of the base of the ith slice; Geotext. Geomembr., 12(8), 687–705.
b ¼ inclination angle of slope; Dash, S. K. (2010). “Influence of relative density of soil on performance of
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

g ¼ unit weight of soil; geocell-reinforced sand foundations.” J. Mater. Civ. Eng., 10.1061
/(ASCE)MT.1943-5533.0000040, 533–538.
g g ¼ density of composite beam;
Dash, S. K. (2012). “Effect of geocell type on load-carrying mechanisms of
d ¼ interface friction angle between geocell geocell-reinforced sand foundations.” Int. J. Geomech., 10.1061
wall and infill soil; /(ASCE)GM.1943-5622.0000162, 537–548.
« a ¼ geocell axial strain at failure; Dash, S. K., and Bora, M. C. (2013). “Improved performance of soft clay
u ¼ angle of rotation with respect to r0; foundations using stone columns and geocell-sand mattress.” Geotext.
u i ¼ angle of rotation with respect to Geomembr., 41(Nov), 26–35.
horizontal axis for the ith slice; Dash, S. K., Krishnaswamy, N. R., and Rajagopal, K. (2001a). “Bearing
l ¼ load dispersion angle within geocell capacity of strip footings supported on geocell-reinforced sand.”
reinforcement; Geotext. Geomembr., 19(4), 235–256.
Dash, S. K., Rajagopal, K., and Krishnaswamy, N. R. (2001b). “Strip foot-
m ¼ soil-geocell interface coefficient; ing on geocell reinforced sand beds with additional planar reinforce-
(s a)ave ¼ average of horizontal soil stresses in the ment.” Geotext. Geomembr., 19(8), 529–538.
ith slice; Dash, S. K., Rajagopal, K., and Krishnaswamy, N. R. (2004). “Performance
s c ¼ circumferential stress in geocell of different geosynthetic reinforcement materials in sand foundations.”
mattress; Geosynth. Int., 11(1), 35–42.
s vi ¼ overburden pressure acting on the ith Dash, S. K., Rajagopal, K., and Krishnaswamy, N. R. (2007). “Behaviour of
geocell reinforcement; geocell-reinforced sand beds under strip loading.” Can. Geotech. J.,
s 3 ¼ active earth pressure exerted by the infill 44(7), 905–916.
Dash, S. K., Sireesh, S., and Sitharam, T. G. (2003). “Model studies on cir-
soil on the geocell wall;
cular footing supported on geocell reinforced sand underlain by soft
t b(x) ¼ interface shear strength at bottom of clay.” Geotext. Geomembr., 21(4), 197–219.
geocell; Duncan, J. M., and Wright, S. G. (2005). Soil strength and slope stability,
t f ¼ available shear strength; John Wiley and Sons, Hoboken, NJ.
t ig ¼ vertical frictional strength between infill Ehrlich, M., and Mirmoradi, S. H. (2013). “Evaluation of the effects of fac-
soil and wall of the ith geocell ing stiffness and toe resistance on the behavior of GRS walls.” Geotext.
reinforcement; Geomembr., 40(Oct), 28–36.
t r ¼ required shear strength; El-Emam, M. M., and Bathurst, R. J. (2007). “Influence of reinforcement
parameters on the seismic response of reduced-scale reinforced soil
t t(x) ¼ interface shear strength at top of geocell;
retaining walls.” Geotext. Geomembr., 25(1), 33–49.
f ¼ internal friction angle of soil; and Emersleben, A., and Meyer, N. (2009). “Interaction between hoop stresses
f i ¼ angle of friction between geocell and soil. and passive earth resistance in single and multiple geocell structures.”
Proc., GIGSA GeoAfrica 2009 Conf., Geosynthetic Interest Group of
South Africa, Edenglen, South Africa, 1–10.
References FLAC2D [Computer software]. Itasca Consulting Group, Inc., Minneapolis.
Ghanbari, A., and Taheri, M. (2012). “An analytical method for calculating
AASHTO. (1998). Bridge design specifications, Washington, DC. active earth pressure in reinforced retaining walls subject to a line sur-
Ahmadabadi, M., and Ghanbari, A. (2009). “New procedure for active earth charge.” Geotext. Geomembr., 34(Oct), 1–10.
pressure calculation in retaining walls with reinforced cohesive- Han, J., Yang, X. M., Leshchinsky, D., Parsons, R. L., and Rosen, A.
frictional backfill.” Geotext. Geomembr., 27(6), 456–463. (2008). “Numerical analysis for mechanisms of a geocell-reinforced
Bathurst, R. J., and Knight, M. A. (1998). “Analysis of geocell reinforced-soil base under a vertical load.” Proc., 4th Asian Regional Conf. on
covers over large span conduits.” Comput. Geotech., 22(3), 205–219. Geosynthetics, Springer, New York, 741–746.
Biabani, M. M., Indraratna, B., and Ngo, N. T. (2016a). “Modelling of geo- Hatami, K., and Bathurst, R. J. (2000). “Effect of structural design on funda-
cell-reinforced subballast subjected to cyclic loading.” Geotext. mental frequency of reinforced-soil retaining walls.” Soil Dyn.
Geomembr., 44(4), 489–503. Earthquake Eng., 19(3), 137–157.
Biabani, M. M., Ngo, N. T., and Indraratna, B. (2016b). “Performance eval- Hegde, A., Kadabinakatti, S., and Sitharam, T. G. (2014). “Protection
uation of railway subballast stabilised with geocell based on pull-out of buried pipelines using a combination of geocell and geogrid
testing.” Geotext. Geomembr., 44(4), 579–591. reinforcement: Experimental studies.” Ground Improvement and
Biswas, A., Krishna, A., and Dash, S. K. (2016). “Behavior of geosynthetic Geosynthetics, Geotechnical Special Publication 238, ASCE, Reston,
reinforced soil foundation systems supported on stiff clay subgrade.” VA, 289–298.
Int. J. Geotech., 10.1061/(ASCE)GM.1943-5622.0000559, 04016007. Hegde, A., and Sitharam, T. G. (2015a). “Joint strength and wall deforma-
Bush, D. I., Jenner, C. G., and Bassett, R. H. (1990). “The design and con- tion characteristics of a single-cell geocell subjected to uniaxial com-
struction of geocell foundation mattresses supporting embankments pression.” Int. J. Geomtech., 10.1061/(ASCE)GM.1943-5622.0000433,
over soft grounds.” Geotext. Geomembr., 9(1), 83–98. 04014080.
Chen, H. T., Hung, W. Y., Chang, C. C., Chen, Y. J., and Lee, C. J. (2007). Hegde, A., and Sitharam, T. G. (2015b). “Experimental and analytical stud-
“Centrifuge modeling test of a geotextile-reinforced wall with a very ies on soft clay beds reinforced with bamboo cells and geocells.” Int. J.
wet clayey backfill.” Geotext. Geomembr., 25(6), 346–359. Geosynth. Ground Eng., 1(2), 1–11.

© ASCE 06017007-13 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Hegde, A., and Sitharam, T. G. (2015c). “3-Dimensional numerical modelling Mehrjardi, G. T., Tafreshi, S. M., and Dawson, A. R. (2012).
of geocell reinforced sand beds.” Geotext. Geomembr., 43(2), 171–181. “Combined use of geocell reinforcement and rubber–soil mixtures
Hegde, A. M., and Sitharam, T. G. (2015d). “Experimental and numerical to improve performance of buried pipes.” Geotext. Geomembr.,
studies on protection of buried pipelines and underground utilities using 34(Oct), 116–130.
geocells.” Geotext. Geomembr., 43(5), 372–381. Mehrjardi, G. T., Tafreshi, S. M., and Dawson, A. R. (2013). “Pipe response
Henkel, D. J., and Gilbert, G. D. (1952). “The effect measured of the rubber in a geocell-reinforced trench and compaction considerations.”
membrane on the triaxial compression strength of clay samples.” Geosynth. Int., 20(2), 105–118.
Geotechnique, 3(1), 20–29. Nouri, H., Fakher, A., and Jones, C. (2006). “Development of horizontal
Huang, C. C., and Wu, S. H. (2006). “Simplified approach for assessing slice method for seismic stability analysis of reinforced slopes and
seismic displacements of soil-retaining walls. Part I: Geosynthetic rein- walls.” Geotext. Geomembr., 24(3), 175–187.
forced modular block walls.” Geosynth. Int., 13(6), 219–233. Nouri, H., Fakher, A., and Jones, C. (2008). “Evaluating the effects of the
Indraratna, B., Biabani, M. M., and Nimbalkar, S. (2015). “Behavior of geo-
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

magnitude and amplification of pseudo-static acceleration on reinforced


cell-reinforced subballast subjected to cyclic loading in plane-strain con- soil slopes and walls using the limit equilibrium horizontal slices
dition.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)GT.1943-5606 method.” Geotext. Geomembr., 26(3), 263–278.
.0001199, 04014081. Pokharel, S. K., Han, J., Leshchinsky, D., Parsons, R. L., and Halahmi, I.
Kaschiev, M. S., and Mikhajlov, K. (1995). “A beam resting on a tension- (2010). “Investigation of factors influencing behavior of single geocell-
less Winkler foundation.” Comput. Struct., 55(2), 261–264. reinforced bases under static loading.” Geotext. Geomembr., 28(6),
Kazimierowicz-Frankowska, K. (2005). “A case study of a geosynthetic re- 570–578.
inforced wall with wrap-around facing.” Geotext. Geomembr., 23(1), Rajagopal, K., Krishnaswamy, N. R., and Latha, G. M. (1999). “Behaviour
107–115. of sand confined with single and multiple geocells.” Geotext.
Krishnaswamy, N. R., Rajagopal, K., and Latha, G. M. (2000). “Model Geomembr., 17(3), 171–184.
studies on geocell supported embankments constructed over a soft clay Rowe, R. K., and Ho, S. K. (1998). “Horizontal deformation in reinforced
foundation.” Geotech. Test. J., 23(1), 45–54. soil walls.” Can. Geotech. J., 35(2), 312–327.
Latha, G. M., Dash, S., and Rajagopal, K. (2009). “Numerical simulation of Rowe, R. K., and Skinner, G. D. (2001). “Numerical analysis of geosyn-
the behavior of geocell reinforced sand in foundations.” Int. J. thetic reinforced retaining wall constructed on a layered soil founda-
Geomech., 10.1061/(ASCE)1532-3641(2009)9:4(143), 143–152. tion.” Geotext. Geomembr., 19(7), 387–412.
Latha, G. M., Rajagopal, K., and Krishnaswamy, N. R. (2006). “Experimental Saride, S., Gowrisetti, S., Sitharam, T. G., and Puppala, A. J. (2009).
and theoretical investigations on geocell-supported embankments.” Int. J. “Numerical simulation of geocell-reinforced sand and clay.” Proc. Inst.
Geomech., 10.1061/(ASCE)1532-3641(2006)6:1(30), 30–35.
Civ. Eng. Ground Improv., 162(4), 185–198.
Latha, G. M. (2000). “Investigations on the behaviour of geocell supported
Shahgholi, M. (2001). “Horizontal slice method of analysis.” Geotechnique,
embankments.” Ph.D. dissertation, Indian Institute of Technology
51(10), 881–886.
Madras, Chennai, India.
Shekarian, S., Ghanbari, A., and Farhadi, A. (2008). “New seismic parame-
Latha, G. M. (2011). “Design of geocell reinforcement for supporting
ters in the analysis of retaining walls with reinforced backfill.” Geotext.
embankments on soft ground.” Geomech. Eng., 3(2), 117–130.
Geomembr., 26(4), 350–356.
Latha, G. M., Dash, S. K., and Rajagopal, K. (2008). “Equivalent continuum
Sireesh, S., Sitharam, T. G., and Dash, S. K. (2009). “Bearing capacity of
simulations of geocell reinforced sand beds supporting strip footings.”
circular footing on geocell–sand mattress overlying clay bed with void.”
Geotech. Geol. Eng., 26(4), 387–398.
Geotext. Geomembr., 27(2), 89–98.
Latha, G. M., and Krishna, A. M. (2008). “Seismic response of reinforced
Sitharam, T. G., and Hegde, A. (2013). “Design and construction of geocell
soil retaining wall models: Influence of backfill relative density.”
foundation to support the embankment on settled red mud.” Geotext.
Geotext. Geomembr., 26(4), 335–349.
Latha, G. M., and Murthy, V. S. (2007). “Effects of reinforcement form on Geomembr., 41(Nov), 55–63.
Song, F., Xie, Y. L., Yang, Y. F., and Yang, X. H. (2014). “Analysis of fail-
the behavior of geosynthetic reinforced sand.” Geotext. Geomembr.,
25(1), 23–32. ure of flexible geocell-reinforced retaining walls in the centrifuge.”
Latha, G. M., and Rajagopal, K. (2007). “Parametric finite element analyses Geosynth. Int., 21(6), 342–351.
of geocell-supported embankments.” Can. Geotech. J., 44(8), 917–927. Tafreshi, S. M., and Dawson, A. R. (2010). “Comparison of bearing
Latha, G. M., and Somwanshi, A. (2009). “Effect of reinforcement form on capacity of a strip footing on sand with geocell and with planar
the bearing capacity of square footings on sand.” Geotext. Geomembr., forms of geotextile reinforcement.” Geotext. Geomembr., 28(1),
27(6), 409–422. 72–84.
Leshchinsky, B., and Ling, H. (2013a). “Effects of geocell confinement on Tafreshi, S. M., and Dawson, A. R. (2012). “A comparison of static and
strength and deformation behavior of gravel.” J. Geotech. Geoenviron. cyclic loading responses of foundations on geocell-reinforced sand.”
Eng., 10.1061/(ASCE)GT.1943-5606.0000757, 340–352. Geotext. Geomembr., 32(Jun), 55–68.
Leshchinsky, B., and Ling, H. I. (2013b). “Numerical modeling of behavior Tafreshi, S. M., Shaghaghi, T., Mehrjardi, G. T., Dawson, A. R., and
of railway ballasted structure with geocell confinement.” Geotext. Ghadrdan, M. (2015). “A simplified method for predicting the settle-
Geomembr., 36(Feb), 33–43. ment of circular footings on multi-layered geocell-reinforced non-
Leshchinsky, D., and Boedeker, R. H. (1989). “Geosynthetic reinforced soil cohesive soils.” Geotext. Geomembr., 43(4), 332–344.
structures.” J. Geotech. Eng., 10.1061/(ASCE)0733-9410(1989)115: Tang, X., and Yang, M. (2013). “Investigation of flexural behavior of geo-
10(1459), 1459–1478. cell reinforcement using three-layered beam model testing.” Geotech.
Leshchinsky, D., Ling, H. I., Wang, J. P., Rosen, A., and Mohri, Y. (2009). Geol. Eng., 31(2), 753–765.
“Equivalent seismic coefficient in geocell retention systems.” Geotext. Tavakoli Mehrjardi, G., Moghaddas Tafreshi, S. N., and Dawson, A. R.
Geomembr., 27(1), 9–18. (2015). “Numerical analysis on Buried pipes protected by combination
Ling, H. I., Leshchinsky, D., Wang, J. P., Mohri, Y., and Rosen, A. (2009). of geocell reinforcement and rubber-soil mixture.” Int. J. Civ. Eng.,
“Seismic response of geocell retaining walls: Experimental studies.” J. 13(2), 90–104.
Geotech. Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2009)135: Thallak, S. G., Saride, S., and Dash, S. K. (2007). “Performance of surface
4(515), 515–524. footing on geocell-reinforced soft clay beds.” Geotech. Geol. Eng.,
Ling, H. I., Leshchinsky, D., and Perry, E. B. (1997). “Seismic design and 25(5), 509–524.
performance of geosynthetic-reinforced soil structures.” Geotechnique, Wei, W. B., and Cheng, Y. M. (2009). “Strength reduction analysis for slope
47(5), 933–952. reinforced with one row of piles.” Comput. Geotech., 36(7), 1176–1185.
Mehdipour, I., Ghazavi, M., and Moayed, R. Z. (2013). “Numerical study Wei, W. B., and Cheng, Y. M. (2010). “Soil nailed slope by strength reduc-
on stability analysis of geocell reinforced slopes by considering the tion and limit equilibrium methods.” Comput. Geotech., 37(5),
bending effect.” Geotext. Geomembr., 37(Apr), 23–34. 602–618.

© ASCE 06017007-14 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007


Xie, Y., and Yang, X. (2009). “Characteristics of a new-type geocell flex- Zhang, L., Zhao, M., Shi, C., and Zhao, H. (2012). “Nonlinear analysis of a
ible retaining wall.” J. Mater. Civ. Eng., 10.1061/(ASCE)0899 geocell mattress on an elastic–plastic foundation.” Comput. Geotech.,
-1561(2009)21:4(171), 171–175. 42(May), 204–211.
Yang, G., Zhang, B., Lv, P., and Zhou, Q. (2009). “Behaviour of geogrid re- Zhang, L., Zhao, M., Zou, X., and Zhao, H. (2009). “Deformation analysis
inforced soil retaining wall with concrete-rigid facing.” Geotext. of geocell reinforcement using Winkler model.” Comput. Geotech.,
Geomembr., 27(5), 350–356. 36(6), 977–983.
Yang, X., et al. (2012). “Accelerated pavement testing of unpaved roads Zhang, L., Zhao, M., Zou, X., and Zhao, H. (2010b). “Analysis of geocell-
with geocell-reinforced sand bases.” Geotext. Geomembr., 32(Jun), reinforced mattress with consideration of horizontal–vertical coupling.”
95–103. Comput. Geotech., 37(6), 748–756.
Yang, X., Han, J., Parsons, R. L., and Leshchinsky, D. (2010). “Three- Zhang, Y., and Murphy, K. D. (2004). “Response of a finite beam in contact
dimensional numerical modeling of single geocell-reinforced sand.” with a tensionless foundation under symmetric and asymmetric load-
Front. Archit. Civ. Eng. China, 4(2), 233–240. ing.” Int. J. Solids Struct., 41(24), 6745–6758.
Downloaded from ascelibrary.org by "Indian Institute of Technology, Hyderabad" on 10/06/23. Copyright ASCE. For personal use only; all rights reserved.

Zhang, L., Zhao, M., Shi, C., and Zhao, H. (2010a). “Bearing capacity of Zhou, H., and Wen, X. (2008). “Model studies on geogrid-or geocell-
geocell reinforcement in embankment engineering.” Geotext. Geomembr., reinforced sand cushion on soft soil.” Geotext. Geomembr., 26(3),
28(5), 475–482. 231–238.

© ASCE 06017007-15 Int. J. Geomech.

Int. J. Geomech., 2017, 17(9): 06017007

You might also like