Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Geotextiles and Geomembranes xxx (xxxx) xxx

Contents lists available at ScienceDirect

Geotextiles and Geomembranes


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

Stress-dependent method for calculating the modulus improvement factor


in geocell-reinforced soil layers
R.S. Garcia, J.O. Avesani Neto *
Department of Structural and Geotechnical Engineering, Escola Politécnica da Universidade de São Paulo (EP-USP), São Paulo, SP, Brazil

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

Keywords: This paper presents an analytical method for determining the modulus improvement factor (MIF) in geocell-
Geosynthetics reinforced soil layers. Using a modified version of the hyperbolic soil model as a constitutive model, the
Soil improvement method is developed based on the soil-reinforcement interaction relating nonlinear elastic soil behavior to the
Modulus of elasticity
linear elastic response of the reinforcement. The proposed method, in an original way, explicitly takes into ac­
Pavement reinforcement
count the geometry of the geocell pocket, effects of soil and geocell-reinforcement stiffness, compaction-induced
stresses, soil strength and strain compatibility. The method can be used both analytically and using simple and
presented non-dimensional charts. Parametric analyses show that the reinforcement, soil relation and the stresses
induced during the compaction procedure are the major factors influencing MIF. An evaluation using data from
several laboratory, full-scale and field experiments in works is presented showing good predictive capability of
proposed method. An application procedure for calculating MIF is presented.

1. Introduction the membrane effect (Avesani Neto et al., 2013; Bathurst and Jarrett,
1988; Zhang et al., 2010). Several studies, using varied approaches, such
Geocell is perhaps the most versatile product in the geosynthetics as membrane theory, limit equilibrium, theory of plasticity, limit state
family with different applications, such as reinforcement of several analysis and semi-empirical models, have provided a way to quantify the
structures, such as embankments, foundations, paved and unpaved improvement provided by the geocell (Arvin and Beigi, 2018; Avesani
roads, railways, heavy duty traffic areas and storage yards; in infra­ Neto et al, 2013, 2015; Bathurst and Karpurapu, 1993; Indraratna et al.,
structure protection for buried pipelines; in reducing vibration in 2015; Koerner, 1994; Latha et al., 2006; Presto, 2008; Rajagopal et al.,
structures; as a retaining wall; and as an element of erosion control in 1999; Sitharam and Hegde, 2013; Song et al., 2019; Zhang et al., 2009).
slopes and water channels (Ahmed and Hany, 2020; Fazeli Dehkordi However, these studies present concepts and equations specially focused
et al., 2019; Guo et al., 2015; Halder and Chakraborty, 2020; Hegde, on the ultimate limit state (ULS) - typical condition in foundation rein­
2017; Lalima et al., 2020; Latha and Rajagopal, 2007; Liu et al., 2018; forcement, for example.
Moghaddas Tafreshi et al., 2020; Siabil et al., 2020; Sitharam and Especially in pavement reinforcement (e.g. paved and unpaved
Hegde, 2013; Song et al, 2018, 2019; Tafreshi et al., 2020; Tavakoli roads, railways and port and airport areas), the condition is typically of
Mehrjardi et al., 2019; Venkateswarlu and Hegde, 2019; Wu and Austin, serviceability limit state (SLS), in which elastic parameters are more
1992). In all of them, the improvement the geocell promotes in the soil is important than the resistance ones in the analyses carried out - as occurs
especially due to the increased confinement generated by its inter­ with the mechanistic-empirical method (AASHTO, 2015; George et al.,
connected cell honeycomb shape (Latha and Somwanshi, 2009; Punetha 2019; Yang and Han, 2013). In these applications, the modulus of
et al., 2020; Zhang et al., 2010). elasticity of the geocell-reinforced soil layer is an important input
In the reinforcement application, there was a great understanding in parameter and difficult to determine. Few but significant researches
the bearing capacity of geocells-reinforced soils due to the several have been carried out in this area to quantify the modulus of elasticity of
resistance mechanisms, such the increased confinement generated by its the soil-geocell layer, using confinement increase in the reinforced soil
interconnected cell honeycomb shape, the stress dispersion effect and (Latha, 2000; Punetha et al., 2020), application of Winkler model

* Corresponding author. Escola Politécnica – Universidade de São Paulo (EP-USP), 05508-010, Av. Professor Almeida Prado, travessa 2 nº 83, LMS, São Paulo, SP,
Brazil.
E-mail addresses: robinson.garcia@usp.br (R.S. Garcia), avesani@usp.br (J.O. Avesani Neto).

https://doi.org/10.1016/j.geotexmem.2020.09.009
Received 1 April 2020; Received in revised form 17 September 2020; Accepted 27 September 2020
0266-1144/© 2020 Elsevier Ltd. All rights reserved.

Please cite this article as: R.S. Garcia, J.O. Avesani Neto, Geotextiles and Geomembranes, https://doi.org/10.1016/j.geotexmem.2020.09.009
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

concepts (Priti et al., 2017; Zhang et al, 2009, 2018) and the layered sea level), K is the lateral earth pressure coefficient (defined as a simple
elastic theory (Avesani Neto, 2019; Moghaddas Tafreshi et al., 2015). ratio between the horizontal and vertical stresses), Kaa is the equivalent
The increase in the soil elastic modulus due to the geocell is trans­ active Rankine earth pressure coefficient and asymptotic value of the
lated into quantitative terms in the parameter called modulus minor to the major principal-stress ratio on the hyperbolic stress-strain
improvement factor (MIF), determined as the ratio of the elastic curve and σ′ 3 is the minor principal stress inside the geocell pocket.
modulus of geocell-reinforced soil by the elastic modulus of unrein­ According to Duncan et al. (1980), the ku/k ratio varies between 1.2
forced soil, usually being reported in a range of 1.5–4.0 by laboratory and 3.0. According to Ehrlich and Mitchell (1994), Kaa can be defined as:
and retro-analysis of works in the field (Al-Qadi and Hughes, 2000; Han,
Ka
2015; Kief et al., 2015; Livneh and Livneh, 2014; Pokharel et al, 2010, Kaa = ( ) (3)
2018; Rajagopal et al, 2012, 2014; Ruge et al., 2020; Tanyu et al., 2013;

c
(1 − Ka ) σ′ tan ϕ
′ + 1 ⋅R1f + Ka
Vega et al., 2018). The previously reported studies only superficially
3

take into account important factors that influence the MIF and, there­ where: Ka is the Rankine active earth pressure coefficient, c’ is the infill
fore, are not able to predict its values with adequate accuracy. soil effective cohesion intercept, ϕ’ is the soil effective friction angle and
In the present paper an analytical method for determining MIF is Rf is the Duncan et al.’s (1980) failure ratio.
presented. The method is developed focusing on the SLS condition, a Considering that usually the materials used to fill the geocells in
situation in which deformability is the main property of analysis. It reinforcement applications are granular, the cohesion will be adopted as
indirectly considers two reinforcement mechanisms, confinement effect zero simplifying the model by ignoring the term σ′ 3 in equation (3).
due to the geocell wall stiffness, and vertical stress dispersion, due to the According to several authors (Huang, 1993; Hunt, 2005; Lambe and
layered elastic system (LET), and does not consider the membrane effect, Whitman, 1991; Look, 2014; Moffatt, 2019; NCHRP, 2004; Nikolaides,
which requires large deformations to contribute to improvement. The 2014; Thom, 2013; Witczak and Yoder, 1975) the Poisson’s ratio of
model presented analyzes the improvement provided to the filling ma­ granular materials commonly used in transport infrastructure varies in a
terial elastic modulus due to the increase in confinement. The method range from 0.25 to 0.40 - with typical values of 0.30–0.35. In the case of
can be used both analytically and using simple non-dimensional charts. granular materials reinforced with geocells, Poisson’s ratio is reported in
It considers the geometry of the geocell pocket, effects of soil and a range of 0.20–0.30 - with a typical value of 0.25 (Avesani Neto, 2019;
geocell-reinforcement stiffness, compaction-induced stresses and strain Mhaiskar and Mandal, 1996; Saride et al., 2016). As these values are
compatibility. The soil-reinforcement interaction was modeled by very close and as a small variation in this parameter influences the
relating the nonlinear elastic soil behavior to the linear elastic response analysis only marginally (De Barros, 1966; Hunt, 2005), for convenience
of the reinforcement. Full interface adherence between soil and geocell the Poisson’s ratio of the soil-geocell composite was adopted as equal to
wall in the strain compatibility was assumed. The soil constitutive model that of the infill soil under the following conditions: i. constant and equal
employed was the modified version of the hyperbolic soil model (Dun­ to the value corresponding to at rest condition during the initial soil
can et al., 1980; Ehrlich and Mitchell, 1994). The compaction stress loading process (equation (4)); ii. constant and equal to the corre­
model (Duncan and Seed, 1986) was applied in a condition similar to sponding value for unloading under at rest condition during soil
that adopted by Ehrlich and Mitchell (1994), allowing analysis using unloading (equation (5)). Poisson’s ratio in the unloading phase was
closed-form equations. The accuracy of the method proposed herein was based on the methods of Duncan and Seed (1986) and Monnet (1994) as
evaluated using data from several laboratory and full-scale plate load used by Ehrlich and Mitchell (1994). According to these authors, this
tests and field pavement retro-analysis for which measured values for analytical soil model permits to analyze compaction-induced stresses
MIF determination were available. A simple suggestion procedure for using closed-form equations and the initial geostatic-soil stresses need
MIF calculation is presented using this method. not correspond to those for K0-conditions.

2. Model development νo =
K0
(4)
1 + K0
The proposed model development considers the compatibility of
Kd
deformations between the soil and the geocell reinforcement and fol­ νd = (5)
1 + Kd
lowed the same line of reasoning as the model of reinforced soil walls
presented by Ehrlich and Mitchell (1994). Some parameters and co­
where: v0 and vd are, respectively, Poisson’s ratio for load and unloading
efficients were the same adopted by those authors, or had adaptations to
conditions, K0 is the at-rest earth-pressure coefficient and Kd is the at-
better fit the modeling of soil confinement by the geocell.
rest decremental K for unloading and defined as:

2.1. Soil model Kd = K0 ×


OCR − OCRα
(6)
OCR − 1
The soil model used is a modified version of the nonlinear elastic
where: OCR and α are, respectively, the overconsolidation ratio between
hyperbolic constitutive model proposed by Duncan et al. (1980). The
vertical stresses and the unloading coefficient reported by Mayne and
tangent modulus was used for convenience as rearranged by Ehrlich and
Kulhawy (1982), given by:
Mitchell (1994) in the loading and unloading condition, respectively:

⎛ ⎞2 σvc
( ′ )n OCR = (7)
σ3 Kaa
⎜1− K ⎟ σ ′v0
Es = k ⋅ Pa ⋅⎝ ⎠ (1)
Pa 1 − Kaa
(8)

α = senϕ
( ′ )n
σ3 where: σ′ vc is the maximum vertical stress due to compaction, σ′ v0 is the
Eun = ku ⋅Pa (2)
Pa vertical stress due to the geocell-reinforcement and above soil layers and
ϕ′ is the effective friction angle of the geocell infill material.
where: Es and Eun are, respectively, the tangent soil deformation
modulus for loading and unloading conditions, k, ku and n are Duncan
et al.’s (1980) modulus number for loading, unloading, and modulus
exponent, respectively, Pa is the atmospheric pressure (= 101.325 kPa at

2
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

2.2. Geocell and soil-geocell interaction For calculating the horizontal stresses in the soil inside the full
loaded cell (σ′ hi), the presence of the geocell wall reinforcement was
The geocell was modeled as a thin cylinder under pressure (Fig. 1) considered. Due to its tensile rigidity and resistance, it contributes by
composed of a linear elastic material with perfect adherence to the increasing the soil confinement inside the cell. Considering the loading
filling and outside surrounding soil throughout the range of stress and and unloading stresses path, the internal horizontal stresses in the cell
deformation considered. Thus, there is complete strain compatibility can be calculated, respectively, using the loading (Kc) and residual (Kr)
between the geocell and the soil without relative sliding between them - lateral earth pressure coefficients, as developed in item 2.3.
a reasonable hypothesis for the interface between geosynthetic For the soil located outside the directly loaded cells - as those
reinforcements-soils under working conditions of stresses and strain bordering the cells fully loaded - the stress path composed of a loading
(Dyer and Milligan, 1984; Jewel, 1980; Palmeira, 2009). due to compaction, with subsequent unloading, was also considered. In
Each structure (geocell wall) of each cell was assumed to be this partially loaded region, the stresses were calculated without the
responsible for its internal horizontal balance. In other words, the hor­ contribution of the reinforcement using the Kaa condition in both
izontal stresses induced within each cell are balanced by the cell walls loading and unloading situations.
after subtraction by the horizontal stresses outside the cell as: To simplify the strain calculations, the minor principal stress was
( ′ ′ ) adopted as the horizontal stress in the loading phase. After the unloading
σ − σ he deq phase, the minor principal stress will be the lowest between the vertical
T = hi (9)
2 and horizontal stresses. The only load applied considered on the cell wall
where: T is the tensile force on the cell wall (hoop stress), σ′ hi is the was the tensile force (hoop stress) due to the internal horizontal stress
internal horizontal stress in the cell, σ′ he is the external horizontal stress inside the cell according to rubber membrane theory (Chen et al., 2013;
in the cell and deq is the equivalent cell diameter obtained considering Henkel and Gilbert, 1952).
the unit cell area of a geocell (AGCE) as an equivalent circumference,
according to the equation below. 2.3. Stress path
√̅̅̅̅̅̅̅̅̅̅̅̅
deq =
4AGCE
(10) The stresses in the geocell filling soil were considered with the stress
π path shown in Fig. 2, the compaction process being simulated as a single
Several studies consider the geocell cell format to be equivalent to loading cycle. The trajectory proposed is based on previously published
circular (Hegde and Sitharam, 2015) or square (Avesani Neto et al, methods of conventional retaining walls and reinforced soil walls
2013, 2015; Leshchinsky and Ling, 2013; Saride et al., 2009). The real (Campanella and Vaid, 1972; Duncan and Seed, 1986; Ehrlich and
shape of industrialized geocells (honeycomb), which corresponds to the Mitchell, 1994).
vast majority of geocells used in transport infrastructure, is very close to The initial stress state (I in Fig. 2) is described as a null condition of
the circular one, justifying the adoption of an equivalent diameter for vertical and horizontal stresses. The resulting increase in stresses due to
the cell. In fact, even for geocells in triangular shape made from geo­ the cell filling material self-weight plus the stresses induced by the
grids, an equivalent diameter is adopted in their modeling (Krishnasw­ compaction process result in the path described between points I and II.
amy et al., 2000; Latha, 2011; Latha et al., 2009). In point II, the vertical (σ′ vc) and horizontal (σ′ hc) maximum stresses
The triaxial state inside the cells was adopted with vertical and acting on the filling material are reached during the execution stage, due
horizontal stresses as the principal stresses and with horizontal stresses to the compaction process. The path between points I and II follows the
being considered of the same magnitude in both directions according to line of the loading lateral earth pressure coefficients (Kc) and its incli­
an axial symmetry (σ′ z = σ′ v σ′ x = σ′ y = σ′ h.). The geocell wall strain is nation, for a given soil, depends on the geocell wall rigidity. For high
then calculated as: rigidity geocell walls, the tendency is for greater lateral strain re­
( ′ ′ )
strictions, generating a Kc-line closer to the K0-condition. However, the
Δεgh =
ΔT
=
Δσ hi − Δσhe deq
(11) lower the geocell walls stiffness, the greater the horizontal strains
J 2J experienced by the soil inside the cell will be, conducting to Kc-line
inferior slopes with limit in the active condition - Kaa-line. For a situation
where: Δεgh is the geocell wall strain and J is the geocell wall stiffness.
of zero reinforcement stiffness (i.e. without reinforcement), the soil will

Fig. 1. Geocell-reinforced soil layer internal equilibrium.

3
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

and dσ′ h, in the triaxial state of stresses of a soil with tangent modulus
and coefficient of Poisson Es and νs, and for the convention of
compressive stresses as positive:
νs − 1 + νs ′
(12)

dεh = dσ v + dσ h
Es Es
Integrating equation (12) and using the formulations and conditions
adopted for the soil tangent modulus and Poisson’s ratio in the loading
and unloading condition, the horizontal strain at the end of the loading
stage (Δεshc) and at the end of the unloading step (Δεshr) are respectively
obtained - points II and III of Fig. 2.

Kc2 (1 − Kaa )2 [νo − Kc (1 − ν0 )]



σ OCR
Δεshc = (v0′ )n (13)
σv0 OCR (Kc − Kaa )2 1− n
k⋅Pa Pa
Kcn

[ ( ) ]
′ [
1− n 1− n ] νd 1 + K12 − 1
σ v0 Kr − (Kc OCR)
Δεshr = ( ′ )n (14)
σ 1− n
ku Pa Pv0a

where: Kc is the loading lateral earth pressure coefficient corresponding


to the stress state including the equivalent increase in stresses induced
by the compaction operations, Kr is the residual lateral earth pressure
Fig. 2. Assumed stress-path for geocell-reinforced soil layers. coefficient corresponding to the end of construction, after the unload of
the compaction operation and K2 is the earth pressure coefficient for the
follow the Kaa-line (Uzan, 1985). unloading path defined as:
After the compaction process has ended, the stress path follows from ′ ′

point II to III in Fig. 2. This path coincides with the decremental K-line σ hc − σ hr (Kc OCR − Kr )
K2 = = (15)
coefficient (K2) until it reaches point III, where a vertical stress occurs σ ′vc − σ ′v0 (OCR − 1)
due to the geocell infill material and any overlying layers self-weight Also considering the equations and conditions related with the soil/
(σ′ v0). The geocell presence induces a residual horizontal stress (σ′ hr) geocell interaction and the stress path (with the concepts of the coeffi­
in the infill soil that, associated with the vertical stress σ′ v0, generates cient of earth pressure Kc and Kr), the geocell wall strains are determined
the stress and strain compatibility point between the soil and the rein­ for loading (Δεghc) and unloading (Δεghr), respectively, after integration
forcement in the line defined by the residual lateral earth pressure co­ as:
efficients (Kr). Again, the greater the geocell walls stiffness, the more ( ′ )
inclined the Kr-line is (i.e., greater the value of the Kr coefficient and the

σv0 Kc OCR − σ v0 Kaa OCR deq
Δεghc = (16)
residual horizontal stress), with a limit on the passive condition - Kp-line. 2J
The Kr value increase implies a higher confinement, improving the soil ′
elastic modulus (Duncan and Bursey, 2013; Hicks and Monismith, 1971) Δεghr =
σv0 [(Kr − Kc OCR) − Kaa (1 − OCR)]deq
(17)
and, therefore, the MIF. 2J
For a situation with no vertical stress induction due to the compac­ The hypothesis of perfect adherence at the interface between the
tion process, there will be no stress path from point II to III. In this geocell walls and the soil leads to a strain compatibility between these
condition, the soil will not experience the unloading and only the materials, for both loading and unloading situations. Thus:
stresses path between points I to II will have been induced, with σ′ vc =
σ′ v0 and point III = point II (OCR = 1). Δεsh = Δεgh (18)
Unlike what was considered by Ehrlich and Mitchell (1994), who Replacing in (18), respectively, the couple equations (13) and (16)
modeled both the loading and unloading process in two stages for and (14) and (17), and rearranging the terms:
analytical convenience (i.e. initial stage without horizontal deformation ( )n
and posterior stage in horizontal deformation under conditions of con­ kPa deq σ vc K 2 (1 − Kaa )2 [νo − Kc (1 − ν0 )]
= c (19)
stant horizontal stress), in our modeling, both trajectories were 2J Pa Kcn (Kc − Kaa )3 1− n
considered direct, coupling the stresses and strains increase/decrease [ ( ) ]
conditions at the same stages during the integrations. [ 1− n ]
( )n Kr − (Kc OCR)1− n νd 1 + (K(OCR− 1)
− 1
Note that Fig. 2 also indicates K-line = 1. A lateral earth pressure kPa deq σ v0 c OCR− Kr )
= (20)
coefficient greater than one situation indicates that the horizontal stress 2J Pa (ku /k)(1 − n)[(Kr − Kc OCR) − Kaa (1 − OCR)]
is higher than the vertical one and, therefore, there is an inversion in the
The terms on the left side of the above equations are defined as the
principal planes. For convenience, the major and minor principal planes
relative extensibility index in loading (β1) and unloading (β2) condition.
were assumed to be constant, horizontally and vertically, respectively,
In these parameters, there is also a relationship between the reinforce­
along the loading path (points I to II). In the unloading condition (point
ment and the soil stiffness, which is defined as the relative stiffness soil-
II to III), the principal planes can be inverted if the residual horizontal
reinforcement index (Si). Equations (19) and (20) can be rewritten as:
stress is high enough.
Kc2 (1 − Kaa )2 [νo − Kc (1 − ν0 )]
β1 = (21)
2.4. Compatibility of soil-geocell deformations Kcn (Kc − Kaa )3 1− n

Considering Hook’s law for the strain increase in the horizontal di­
rection dεh due to the increase of vertical and horizontal stresses dσ′ v

4
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

[ ( ) ]
[ 1− n ] composed of a portion referring to the Es confined soil and another
Kr − (Kc OCR)1− n νd 1 + (K(OCR− 1)
− 1
c OCR− Kr ) portion referring to the Eg geocell according to equation (28) (Dowling,
β2 = (22)
(ku /k)(1 − n)[(Kr − Kc OCR) − Kaa (1 − OCR)] 2013):
Es As Eg Ag
where: Esg = + (28)
A A
( )n
1 σ v0 OCR
β1 = (23) where: A is the sum of the geocell (Ag) and soil (As) areas and Eg is the
Si Pa
geocell wall modulus, given by:
( )n
β2 =
1 σ v0 β
= 1 n (24) As = hds (29)
Si Pa OCR
Ag = ht (30)
2J
Si = (25)
kPa deq J
Eg = (31)
t
Typical values of Si assuming usual granular materials and depend­
ing on the raw material of the geocells wall are: where: h is the geocell height, ds is the soil width inside the cell and t is
the geocell wall thickness.
• HDPE and PP geotextile geocells: 0.005 to 0.15 (Gurbuz and Mertol, Assuming that the soil width inside the cell is equal to the equivalent
2012; Han et al., 2008; Pokharel et al., 2010); width of the geocell - deq (equation (10)):
• NPA geocells: 0.01 to 0.50 (Leshchinsky and Ling, 2013; Pokharel
et al., 2010; PRS, 2020); Es hdeq + Jh
Esg = ( ) (32)
• Metallic geocells (for comparative purposes only) 2.0 to 20.0 deq + t h
(Webster, 1979, 1981; Webster and Alford, 1978). Since deq is much greater than t (usually for industrialized geocells
deq/t > 10 to 15), then the geocell thickness wall in the denominator can
Although there is a complexity in the formulations developed, be neglected, obtaining the modulus of elasticity of the soil-geocell
requiring interactive calculations to determine its parameters, it is composite:
possible to extract some properties that can facilitate the calculations for
specific situations: Esg = Es +
J
(33)
deq
• When the maximum vertical loading stress in compaction is equal to It should be noted that if the flexural stiffness concept from the plate
the vertical stress due to the soil self-weight (i.e., σ′ vc/σ′ v0 = OCR = theory is used (Timoshenko and Woinowsky-Krieger, 1959; Ventsel and
1), coefficients Kc and Kr are equal. Krauthammer, 2001), adopting, by simplification, an equal value for the
• Kc remains constant at the same value of β1. soil and geocell wall Poisson’s ratio, the same result is obtained (equa­
• When the geocell wall stiffness (J) tends to infinity (i.e., β1 and β2 tion (33)).
tend to zero), Kc tends to K0 and Kr tends to the at rest lateral earth
pressure coefficient of an overconsolidated soil (K0, OCR), according
to Mayne and Kulhawy (1982), given by equation (26) - situation 2.6. Determination of the modulus improvement factor
equivalent to a retaining wall with a nondeflecting face (Duncan and
Seed, 1986). Consequently, for this situation, the value of K2 As stated in the introduction, several studies into physical, analytical
(equation (15)) tends to the value of Kd (equation (6)). and computational models and into instrumented works have shown
• When the geocell wall stiffness (J) tends to zero (i.e., β1 and β2 tend to that the presence of the geocell generates an increase in soil stiffness. It
infinity), Kc and Kr tend to Kaa - situation equivalent to the lack of depends on several factors, such as type and condition of the filling soil;
reinforcement that leads to the active condition in the soil (Uzan, geocell dimensions, geometry and raw material; and state of stresses
1985). applied to the reinforced soil layer. The relationship between the
• There is an OCR value in which the increase in strain due to the modulus of elasticity of a geocell-reinforced soil by the modulus of that
unloading Δεshr and Δεghr (equations (14) and (17)) is null. same soil without the reinforcement is called the modulus improvement
Furthermore, this implies that the external and internal horizontal factor, MIF.
stress reduction will be the same. This OCR value must obey the To determine the MIF, it is necessary to initially calculate the
relation presented in equation (27). modulus of elasticity in the unreinforced soil. Applying the equations
• OCR values below this previous limit will lead to external stress and concepts from items 2.1 and 2.3 and using a reinforcement stiffness
reduction in higher magnitude than the internal ones, generating J = 0 (β1 and β2 → ∞):
positive Δεsh strain (i.e., cell expansion). The opposite phenomenon ( ′
σ Kaa
)n
will occur when OCR is above this limit. Eun = ku Pa v0 (34)
Pa
1− n
• All the strain to be obtained is proportional to the (σ vc /Pa ) .

The modulus of elasticity of the geocell-reinforced soil layer is then


K0,OCR = K0 OCR α
(26) determined by equation (33), rewritten below, and must observe the
( ) conditions imposed on items shown so far:
K0 (OCR − OCRα ) 1
1 + =1 (27) ( ′
σ Kr
)n
J
OCR − 1 + K0 (OCR − OCRα ) Kaa Esg = ku Pa v0 + (35)
Pa deq

2.5. Geocell-soil composite modulus of elasticity The MIF is calculated as the relationship between equations (35) and
(34) that can be rewritten as:
Considering the geocell-reinforced soil layer as a composite material, ( )n ( )n
Esg Kr 0, 5 1
and the reinforcement as an inclusion of a material with a higher tensile MIF = = + (36)
Eun Kaa β2 ku /k Kaa
elastic modulus and resistance than the soil, a composite elastic modulus
(Esg) is obtained. The composite modulus can be considered as

5
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

3. Parametric analysis and MIF charts proportional to it, the increase in Si generates a reduction in β2 and thus
an increase in MIF. For low OCR and Si conditions, an increase of Si by
Based on the model developed in the previous item, and due to its about 10 times is verified to generate MIF increment in the order of
interactive character, a parametric analysis was carried out to verify the 5–20%, while for higher Si, the MIF increase is greater than 60% for low
MIF major influence factors - i. soil shear resistance (ϕ′ ); ii. Stress state OCR. This result is expected since an increase in the geocell stiffness
(σ′ v0, σ′ vc and OCR); and iii. relative stiffness soil-reinforcement index generates higher confinement values, especially for less rigid soils.
(Si). In addition to the parametric analysis, it was possible to generate Greater confinements in granular soils lead to greater gains in strength
charts to facilitate the proposed model application. and stiffness. The soil-reinforcement relative stiffness index (Si) can be
As a result of this analysis, and considering the values of Rf = 0.8, ku/ increased by using a geocell with high wall rigidity – achieved with a
k = 1.2 and n = 0.4, a range usually reported in the literature for higher raw material modulus of elasticity or by a thicker wall, or even by
granular materials normally used in pavement structures (Duncan et al., a cell with a smaller area - or by using an infill soil with low stiffness.
1980; Ehrlich and Mitchell, 1994), Fig. 3 shows the nondimensional Reducing the equivalent cell size (pocket size) also increases Si
charts for determining MIF as a function of β2 and OCR for filling ma­ Second parameter of greatest influence is compaction, represented
terial friction angles from 30◦ to 45◦ - usual range of granular materials by the OCR parameter which is the relationship between the maximum
used in paving (Nicks and Adams, 2013; Nicks et al., 2015). In the vertical stress applied to the geocell layer, resulting from the compaction
charts, the lines corresponding to β2 → ∞ indicate the situation without process, divided by the vertical stress due to self-weight. For low Si
reinforcement (J = 0) and, therefore, MIF = 1. In addition, changes in values, a situation in which the stiffness ratio between the geocell and
the slope of the curves under higher OCR conditions can be observed, the soil is low (more rigid soils or more flexible geocells), an increase
indicating situations with an inversion of the principal stresses, i.e. σ′ v0 from 1 to 10 in OCR leads to increments in the range of 50% in the MIF.
becomes the minor principal stress and the model adopts Kr = 1. In general, both parameters are complementary and it is possible to
The chart analysis in Fig. 3 leads to the conclusion that the effect of obtain higher MIF values by working with these two parameters: raising
soil-reinforcement relative stiffness index (Si), characterized by the the OCR in lower Si conditions or increasing the Si in low OCR situations.
relationship between the relative reinforcement stiffness (J/deq) and the Note that the condition of limiting the Kr value equal to one also implies
soil stiffness (k), is the most sensitive parameter independently of the that an increase in the compaction stresses, for high OCR situations, does
friction angle. The Si parameter directly relates the geocell walls stiffness not modify the MIF significantly. Furthermore, the MIF also depends on
to the soil stiffness. As β2 depends directly on Si, being inversely the vertical geostatic stresses (layers self-weights). If the geocell-

Fig. 3. Charts for MIF determination in geocell-reinforces soil layers.

6
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

reinforced soil layer is installed at a high depth, the compaction effect is The analyses in the next sub-items are grouped according to the
attenuated due to the OCR decrease. experiments condition: laboratory plate load tests (PLT) on single cell;
The infill material friction angle is the third most influential on multiple cells; and field tests or instrumented works to allow the
parameter, since the lateral earth pressure coefficients (Kc and Kr) are proposed model evaluations under different conditions of geocell
directly dependent on it. Soils with high friction angles positively in­ application. The MIF values calculated with the proposed model were
fluence the stiffness increase, this result being more pronounced for high compared with the results reported in each study and are shown in Fig. 4
OCR values situations. Keeping all the other parameters, increasing the and in Table 2.
friction angle by 5◦ leads to about 10% in MIF gain.

4. Case studies

The model developed herein was used to determine MIF values from
several previously published geocell-reinforcement soil studies and with
enough information to allow the calculations. Ten publications were
selected involving different parameters, such as type of test, loading
condition, filling soil, compacting condition and geocells with several
polymers, wall stiffness and sizes - most common models in the geo­
synthetics market. Table 1 presents some details of the eleven cases
including a total of 24 experiments.
In many publications, the necessary information for MIF calculation
using the proposed model was not available in its entirety and some
were adopted (and indicated) under consistent premises with the ex­
periments characteristics. The vertical stresses used in the proposed
model were considered in the middle of the geocell-reinforced soil layer,
where the compaction stresses, when existing, were calculated using the
Ehrlich and Mitchell (1994) methodology with equipment and
compaction energy information provided in each study. Some publica­
tions did not provide the compaction methodologies or equipment, and
conservatively admitted that no compaction occurred or that the plu­
viation technique was used to build the physical models. In this condi­
Fig. 4. Comparison of the modulus improvement factor (MIF) reported in the
tion, the layers were considered not to be subjected to vertical stresses
Case Studies and calculated by the method proposed.
greater than their self-weight, leading to an OCR = 1 condition.

Table 1
Case studies used for evaluating the proposed model.
Case Situation Condition Load Reference Geocell h (mm) Soil classification (USCS)

1 Laboratory Single cell Static plate load Han et al. (2008) HDPE 50 SP

2 Laboratory Single cell Static plate load Pokharel et al. (2010) HDPE 100 SP
NPA 100 SP
NPA 100 SP
NPA 100 SP

3 Laboratory Single cell Cyclic plate load Pokharel et al. (2018) NPA 100 SP-SC
NPA 100 GW-GC

4 Laboratory Multiple cells Static plate load Biswas et al. (2013, 2016) and Biswas and Krishna (2018, 2019) GG 80 SP
GG 158 SP
GG 236 SP
GG 314 SP

5 Laboratory Multiple cells Cyclic plate load Tanyu et al. (2013) HDPE 150 GP-GM
HDPE 200 GP-GM
HDPE 150 GP-GM
HDPE 200 GP-GM

6 Laboratory Multiple cells Cyclic plate load Pokharel et al. (2018) NPA 100 SP
NPA 100 SP-SC
NPA 100 GW-GC

7 Laboratory Multiple cells Cyclic plate load Pokharel (2010) NPA 150 GW-GC
NPA 300 GW-GC

8 Laboratory Multiple cells Cyclic plate load George et al. (2019) HDPE 100 SW

9 Field Multiple cells FWD Al-Qadi and Hughes (2000) NPA 100 SP-SC

10 Field Multiple cells Plate loada Rajagopal et al. (2012) NPA 100 GW-GC

11 Field Multiple cells Cyclic plate load Saride et al. (2016) HDPE 150 GW ou GP

HDPE: High-density polyethylene.


NPA: Novel polymeric alloy.
GG: Geocell made from geogrid strips.
h: Geocell height.
a
No information if it was cyclic or static loading.

7
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

Table 2
MIF calculation by the method proposed using parameters from Case Studies and experimentally obtained.
Case deq (mm)a J (kN/m)b ϕ’ (◦ )c kc nc OCRd β2 MIFexp MIFcalc

1 183 250 45 555 0.4 1 2.7 1.88 1.62

2 205 465 40 555 0.4 1 2.0 1.50 1.63


205 390 40 555 0.4 1 2.4 1.60 1.57
205 385 40 555 0.4 1 2.5 1.70 1.57
205 484 40 555 0.4 1 2.0 2.00 1.65

3 175 390 45 1200 0.4 1 4.7 1.26 1.44


175 390 45 900 0.4 1 3.6 1.32 1.52

4 135 240 40 195 0.4 1 0.8 3.40 2.40


135 240 40 195 0.4 1 1.0 2.00 2.10
135 240 40 195 0.4 1 1.2 1.70 2.00
135 240 40 195 0.4 1 1.3 1.50 1.90

5 242 250 45 1200 0.4 30 13.3 3.05 2.25


242 250 45 1200 0.4 30 13.3 2.13 2.25
192 250 45 1200 0.4 30 10.6 1.54 2.26
192 250 45 1200 0.4 30 10.6 2.17 2.26

6 175 390 40 555 0.4 1 2.1 2.04 1.65


175 390 45 1200 0.4 1 4.8 1.46 1.44
175 390 45 900 0.4 1 3.6 1.73 1.52

7 183 391 45 900 0.4 20.8 4.1 2.50 2.41


183 391 45 900 0.4 10.4 5.4 1.60 2.28

8 242 250 45 900 0.4 34.0 6.8 2.50 2.31

9 169 250 45 1500 0.4 40 18.4 2.00 2.23

10 183 960 45 1500 0.4 96 3.7 2.75 2.43

11 192 250 45 1500 0.4 148 11.5 2.18 2.26

J: Geocell wall stiffness.


deq: equivalent cell diameter (pocket size).
ϕ’: Filling soil friction angle.
k, n: Hyperbolic model parameters.
OCR: Overconsolidation ratio.
β2: relative extensibility index.
MIFexp: MIF obtained in experiments.
MIFcalc: MIF calculated by the proposed method.
a
Calculated according to cell area or dimensions.
b
When not provided, estimated as a function of the elastic modulus of the geocell polymer and cell wall thickness.
c
When not provided, estimated according to the granular filling material and based on Duncan et al. (1980) and Trautmann et al. (1983).
d
Estimated based on the compaction data provided.

From Fig. 4 and in general, it can be seen that for all test conditions Han et al. (2008) and Pokharel et al. (2010) - Cases 1 and 2,
evaluated, the proposed model calculated MIF values in good agreement respectively - carried out tests in static load conditions using the Kansas
to those obtained in the experiments. Complementing the analyses with River sand in different strength situation for each article as the filling
Table 2 and for all the results, the absolute value of MIF difference be­ material. The results of tests described by Yang et al. (2010) were used
tween the one calculated with those obtained was between 0.02 and for the hyperbolic model parameters of this sand. The experiments were
1.00 (respectively for Cases 6 and 4), with an average of 0.29. The error carried out with several configurations of geocells in terms of di­
range obtained was from 1 to 47% (respectively for Cases 6 and 5) with mensions, polymers and cell wall stiffness. Tables 1 and 2 show the
an average of 15%. In half of the Cases, the proposed model calculated a variables used in the experiments and the MIF values obtained by the
value lower than that obtained in the test. Considering the number of authors above and calculated by the proposed method. Case 3 is
variables involved, the need to estimate some parameters and dis­ described by Pokharel et al. (2018), who performed tests with cyclic
regarding the effect of compaction in most cases, the proposed model is load condition, using quarry waste and crushed aggregate as infill ma­
concluded to satisfactorily predict the improvement obtained by the terials and an NPA geocell reinforcement in a condition similar to one of
reinforcement in the different conditions of experiments, infill soils, the geocells employed by Pokharel et al. (2010).
compaction loads and geocells. In all these researches, the authors reported the MIF values obtained
in the experiments, yet no information about the compaction process
4.1. Laboratory plate load tests on single cell – cases 1-3 and equipment was provided; the proposed model considered OCR = 1
for these cases. For Case 3, the hyperbolic soil model parameters were
Laboratory plate load tests are good references as they are carried out adopted according to the range suggested by Duncan et al. (1980) and
under controlled conditions with minor variations in parameters and Trautmann et al. (1983), abiding by the USCS classification for each
conditions. The experiments with single cell are interesting to evaluate infill material.
the proposed model because they physically represent the analytical Table 2 and Fig. 4 (open chart markers such as “+“) show that, for the
modeling used herein. The tests were carried out in a small apparatus single cell condition over a rigid base, the proposed model calculated
(usually less than 1.0 m) with and without the presence of reinforcement MIF values consistent with those obtained in the experiments. The dif­
in the granular material, placed directly on a rigid base, using a plate ference between the values calculated with those obtained was between
load with a dimension smaller than the cell opening and positioned in its 0.03 and 0.35 (average of 0.18) generating an error of 2–18% (average
center. of 11%). In most cases, the proposed model calculated a value lower

8
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

than that obtained in the assay - possibly by considering OCR = 1. in the soil confinement condition but also due to the membrane effect.
Considering all the seven experiments in Cases 1 to 3, the average MIF However, as it was not possible to evaluate all the assembly and
value obtained in the tests and calculated by the present method were execution conditions of the tests at these specific points (geocell instal­
1.61 and 1.57, respectively. lation and fill process, compaction, positioning of the loading plate over
the geocell mattress, etc.), more conclusions about these results are
4.2. Laboratory plate load tests on multi cell – cases 4-8 compromised, understanding that the error obtained is within the range
observed in the method proposed.
Laboratory tests with multiple cells are important to expand the
proposed model evaluation under conditions similar to the application 4.3. Field tests – cases 9-11
of reinforcement in the field, but in situations of greater control on the
variables involved. Unlike single cell tests, the apparatus involved is Despite having less control on the variables and parameters involved,
larger - usually greater than 1.0 m. Again, the tests were carried out with field tests and instrumented works are interesting because they allow
and without the reinforcement in the granular material, this time placed comparing results under real conditions of geocell reinforcement
directly over a subgrade with less resistance/stiffness in order to simu­ application. In field conditions, different types of tests were used, and
late a common field condition. again the tests carried out with and without reinforcement in the
Biswas et al. (2013, 2016) and Biswas and Krishna (2019, 2018) – granular material were performed directly on a weak subgrade with less
Case 4 - carried out tests under static load conditions on a layer of poorly strength/stiffness than the reinforced layer.
graded sand over a subgrade of clay material under different strength Al-Qadi and Hughes (2000) – Case 9 - performed a series of FWD tests
conditions. The geocell reinforcement was produced using geogrid strips on a road rehabilitation in the state of Pennsylvania/USA. The pavement
assembled in the “Chevron” format. Tests under similar conditions but structure consisted of an 88-thickness hot-mixed asphalt layer, 200 mm
with cyclic loading were performed by Tanyu et al. (2013), Pokharel of bituminous concrete base and 150 mm of granular sub-base on a
et al. (2018), Pokharel (2010) and George et al. (2019) - Cases 5, 6, 7 and non-woven geotextile directly disposed over the subgrade composed of a
8 respectively. Tanyu et al. (2013) simulated a weak subgrade using EPS sandy silt with organic content and CBR equal to 4%. The work used
blocks and used four HDPE geocells with different dimensions filled with several techniques for rehabilitation including 100-mm high HDPE
poorly graded gravel as reinforcement. Pokharel et al. (2018) performed geocells.
experiments with multiple cells with the same materials used for the Case 10 is reported by Rajagopal et al. (2012) who performed plate
single cell test (Case 3), also using the Kansas River sand used in load tests on an unpaved road. The structure consisted of a clay subgrade
Pokharel et al. (2010) - here with a friction angle of 41◦ reported by the with 4% of CBR, 400 mm of a compacted in situ soil with 6% of CBR,
authors. Only the experiments with NPA geocell filled with crushed 400 mm of granular sub-base and 225 mm of granular sub-base rein­
aggregate from Pokharel (2010) were here analyzed. George et al. forced with a 150 mm high NPA geocell.
(2019) performed experiments using a HDPE geocell with multiples cells Case 11 is described by Saride et al. (2016), already analyzed by
filled with reclaimed asphalt pavement (RAP) material. Avesani Neto (2019). Cyclic loading plate tests were performed directly
The compaction methodology was detailed in Tanyu et al. (2013), on the soft clayey subgrade and on a 150 mm granular sub-base layer
Pokharel (2010) and George et al. (2019), which allowed estimating the (GSB) without and with a HDPE geocell of the same height.
compaction stresses for Cases 5, 7 and 8. For Cases 4 and 6, OCR = 1 was The compaction equipment was mentioned only in Saride et al.
adopted. For all the Cases, the hyperbolic model parameters were esti­ (2016) and considered to be the same equipment in other cases - vertical
mated according to the range suggested by Duncan et al. (1980) and dynamic roller drum force of 200 kN and roller drum length equal to 2.3
Trautmann et al. (1983) abiding by the USCS classification for each infill m. The compaction stress was obtained equal to 250 kPa according to
material. The MIF values were reported only in Cases 6 and 8. For Cases the procedure suggested by Ehrlich and Mitchell (1994). For all the
4 and 5, the MIF values from the experiments were determined using the Cases, the hyperbolic model parameters were estimated according to the
layered elastic theory according to the methodologies proposed by range suggested by Duncan et al. (1980) and Trautmann et al. (1983)
Avesani Neto (2019) and Garcia and Avesani Neto (2020). In Case 7 the according to the USCS classification for each infill material. The MIF
elastic moduli form reinforced and unreinforced situations were re­ values from the field tests were reported in Cases 9 and 10, and for Case
ported allowing the MIF calculation. The information described in the 11, determined using the layered elastic theory according to method­
studies and that adopted in the calculations of the proposed model are ologies proposed by Avesani Neto (2019) and Garcia and Avesani Neto
presented in Tables 1 and 2 (2020).
Table 2 and Fig. 4 (shapes chart markers such as “□“) show that, for Table 2 and Fig. 4 (filled shapes chart markers such as “■“) show
the multiple cell condition tests on a weak subgrade, the proposed model that, for field tests, the proposed model was again able to calculate MIF
was also able to calculate MIF values consistent with those registered in values consistent with those obtained in the experiments. The difference
the experiments. The difference between the values calculated with between the values calculated with those obtained was from 0.08 to 0.32
those obtained was between 0.02 and 1.00 (average of 0.37) generating (average of 0.21) generating an error of 4–12% (average of 9%), and in
an error of 1–47% (average of 18%). Half of the times, the proposed more than half of the times, the proposed model calculated an MIF value
model calculated a value lower than that obtained in test. Considering higher than that obtained in the test - possibly due to the type of
all the fourteen experiments in Cases 4 to 8, the average MIF values compaction considered in the model. Considering the three tests in Cases
obtained in the experiments and calculated by the present method were 9 to 11, the average MIF obtained in the experiments and calculated by
2.09 and 2.07, respectively. A higher MIF is expected for this test con­ the present method were the same and equal to 2.31. A higher MIF value
dition due to the presence of multiple cells and the existence of a sub­ is expected for this test condition due to the presence of multiple cells,
grade with less rigidity, leading to greater improvements provided by the existence of a subgrade with less rigidity and the effect of field
the reinforcement effect. compaction, all leading to greater improvements provided by the rein­
Still from this Figure and Table, it is possible to observe two points forcement effect.
(Cases 4 and 5) where the MIF generated in the experiments were visibly
higher and the proposed model was not able to make a better prevision, 5. Suggestion of application
generating an error of about 25%. In Case 4, this point refers to a con­
dition of a small height of the layer over the subgrade (with and without As verified in the previous item, the proposed model can be applied
the geocell) of 80 mm, a situation in which, even with few deformation satisfactorily to different conditions of filling granular material and
magnitudes, the geocell reinforcement layer might not be working only geocells, explicitly taking into account properties, such as soil and

9
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

reinforcement stiffness, geocell geometry and compaction effect. conditions of reduced confinement stress). This limitation can be solved
The improvement in determining the modulus of elasticity due to the by considering a higher friction angle value for the infill soil (Bolton,
geocell can be made analytically using the equations in item 2 or by the 1986). vi. The hyperbolic model uses the minor principal stress in the
charts in Fig. 3. Seven soil parameters are needed: one physical – unit tangent elastic modulus for unloading and reloading, which can be
weight (γ); two shear resistances (c‘ and ϕ′ ); and four from the hyper­ related to the resilient modulus (Dunlap, 1963; Hicks and Monismith,
bolic model (Rf, n, k, ku); two parameters from the geocell reinforce­ 1971). However, recent models of resilient modulus can consider other
ment: one geometric (deq) and one from stiffness (J); finally, two from relationship of stress, such as bulk stress and octahedral stresses
state of stresses: self-weight vertical stresses from geocell and any above (NCHRP, 2004; Uzan, 1992). The concept of the MIF developed here
layer and compaction stress (σ′ v0 and σ′ vc). may allow its use in these conditions, since it is a direct proportion be­
In the analytical procedure, using the soil-reinforcement strain tween the moduli in the reinforced and non-reinforced condition. The
compatibility equations in item 2.4, equations (21) and (22) can be method of MIF calculation proposed here has a good approximation with
iteratively solved for determining Kc and Kr and then calculating the MIF the experimental data, which would (in practice) enable its direct
according to item 2.6. For the chart in Fig. 3, prepared considering the application to the reinforced modulus determination independent of the
filling soil characteristics as described in item 3, the following procedure soil constitutive model.
is recommended: (i) computing σ′ v0 in the geocell layer center; (ii)
determining σ′ vc (for example, according to the Ehrlich and Mitchell 7. Conclusion
1994 methodology) and OCR; (iii) choosing a geocell with characteris­
tics of deq and J; (iv) computing Si and β2; (v) with OCR and β2, choosing This paper presented an analytical method for determining the
the appropriate chart for the respective filling material friction angle modulus improvement factor (MIF) in geocell-reinforced soil layers. The
and determining the MIF - the results from two charts can be interpo­ methodology was based on the soil-reinforcement interaction modeled
lated for an intermediate friction angle. An illustrative calculation of a by relating nonlinear elastic soil behavior, using a modified version of
hypothetical example is performed below. the hyperbolic soil model as constitutive model to the linear elastic
response of the reinforcement. The model developed considers the ge­
5.1. Example ometry of the geocell pocket, effects of soil and geocell-reinforcement
stiffness, compaction-induced stresses, soil strength and strain compat­
Unpaved road with a granular material reinforced with geocell ibility. The accuracy of the proposed method was evaluated using data
placed directly on the subgrade. Use of a granular sub-base with the from several laboratory and full-scale plate load tests and field pavement
following parameters: γ = 20 kN/m3, c’ = 0, ϕ’ = 45◦ , Rf = 0.8, k = 600, retro analysis for which measured values for MIF determination were
ku/k = 1.2, n = 0.4. Geocell height h = 200 mm, individual cell area available. The proposed method can be used both analytically and using
equal to 250 cm2, 1.5-mm wall thickness and elastic modulus of the wall simple non-dimensional charts and a suggested procedure for calcu­
(polymer) material equal to 600 MPa. Compaction equipment with 130 lating MIF was presented. Based on the results, the following conclusions
kN of load and 2.1 m roll width. can be drawn:

5.2. Solution • The method developed is original since it considers the geometry of
the geocell pocket, effects of soil and geocell-reinforcement stiffness,
(i) With the geocell height and the granular sub-base unit weight σ′ v0 compaction-induced stresses, soil strength and strain compatibility,
= 2 kPa is calculated. (ii) Using the Ehrlich and Mitchell (1994) and it is simple and easy to apply, especially when using non-
compaction methodology σ′ vc = 201 kPa and OCR = 100 are deter­ dimensional charts;
mined. (iii) With the cell wall modulus of elasticity and its thickness J = • Parametric analysis showed that the MIF is more significantly
900 kN/m is calculated and with the cell area, deq = 178.5 mm is influenced by the relative soil-reinforcement stiffness index (Si). An
determined. (iv) Si = 0.17 and β2 = 1.25 are computed. (v) With 1/OCR increase by about 10 times in Si, which can be obtained by adjusting
= 0.01 and β2 = 1.25, a MIF = 2.9 is determined using the ϕ’ = 45◦ chart geocell types and geometry configurations with different filling
in Fig. 3. materials, generates MIF increment in the order of 5–20% for low
OCR and Si conditions. For higher Si values, the MIF increase is
6. Reflection points and limitations greater than 60% for low OCR conditions.
• The other parameters of greatest influence in the MIF are OCR, which
As with any analytical method in the geotechnical engineering field, is a function of the compaction process and the depth of installation
the one presented here is not free from limitations as it was developed of the geocell-reinforced soil layer, and the soil friction angle (ϕ′ ). An
for specific boundary conditions. Its application by engineers must be increase from 1 to 10 in OCR leads to increments in the range of 50%
accompanied by a careful assessment of the problem, whereby the in the MIF. Keeping all the other parameters, increasing the friction
limitations can be circumvented by design criteria. angle by 5◦ leads to about 10% in the MIF gain.
Some points of important reflection on the application of geocell • The good predictive ability of the method proposed was demon­
reinforcement in soils and that refer to possible limitations of the present strated by comparison with MIF data from several laboratory, full-
method are: i. the soils and the geocell elastic parameters must be known scale and field experiments in works in various configurations of
for the typical range of stresses and strains in the application. ii. The type, raw material and geometry of geocells, filling soils and
hypothesis of cohesion equal to 0 could lead to a different value of MIF compaction procedures. Therefore, the proposed method has the
when the infill material used is not granular or cohesionless, as the potential to be promising in geocell applications for designing
reclaimed asphalt pavement material (RAP). iii. In large life-spam infrastructure and transport engineering, targeting roads and pave­
works, in which the pavement is partially reinforced or reconstructed, ments built with economy and efficiency.
permanent deformations can be significant. In this condition, the geocell
layer can also start working with the membrane effect, generating im­
provements not accounted for by this method. iv. The model does not Declaration of competing interest
directly predict the geocell stiffness degradation by the stress relaxation
effect related to the polymeric creep. For applications with large life- The authors declare that they have no known competing financial
spam, the value of J may require some reduction. v. The hyperbolic interests or personal relationships that could have appeared to influence
model does not take into account the dilatancy (especially acting in the work reported in this paper.

10
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

Acknowledgements Improvement of Higher Education Personnel (CAPES – 88887.463870/


2019–00) for funding the first author, the Escola Politécnica at Uni­
The authors would like to acknowledge the Coordination for the versidade de São Paulo (EP-USP) for the support.

Notations

Basic SI units are given in parentheses.

Ag geocell cell wall area (m2)


AGCE geocell unit cell area (m2)
As soil inside cell area (m2)
c interception of cohesion of the infill soil (kPa)

deq equivalent cell diameter (m)


ds soil width inside the cell (m).Eg geocell wall modulus of elasticity (Young’s modulus) (kPa)
Es infill soil modulus of elasticity (Young’s modulus) (kPa)
ESG geocell-soil composite modulus of elasticity (Young’s modulus) (kPa)
Eun unreinforced soil modulus of elasticity (Young’s modulus) (kPa)
h geocell height (m)
J geocell wall stiffness (kN/m)
k Duncan et al.’s (1980) modulus number for loading (dimensionless)
K lateral earth pressure coefficient (dimensionless)
K0 at-rest lateral earth pressure coefficient (dimensionless)
K0,OCR at-rest lateral earth pressure coefficient for overconsolidated soils (dimensionless)
K2 lateral earth pressure coefficient in the unloading path (dimensionless)
Ka active lateral earth pressure coefficient (dimensionless)
Kaa equivalent active lateral earth pressure coefficient (dimensionless)
Kc loading lateral earth pressure coefficient (dimensionless)
Kd unloading lateral earth pressure coefficient (dimensionless)
Kr residual lateral earth pressure coefficient (dimensionless)
Kp passive lateral earth pressure coefficient (dimensionless)
ku Duncan et al.’s (1980) modulus number for unloading (dimensionless)
MIF Modulus Improvement Factor (dimensionless)
n Duncan et al.’s (1980) modulus exponent (dimensionless)
OCR overconsolidation ratio - σ vc /σv0 (dimensionless)
′ ′

Pa atmospheric pressure (dimensionless)


Rf Duncan et al.’s (1980) failure ratio
Si soil-reinforcement relative stiffness index (dimensionless)
t geocell wall thickness (m)
T tensile force on the cell wall (hoop stress) (kN/m)
α unloading coefficient (dimensionless)
β1 relative extensibility index in loading (dimensionless)
β2 relative extensibility index in unloading (dimensionless)
Δεgh geocell wall strains (dimensionless)
Δεghc geocell wall strains for loading (dimensionless)
Δεghr geocell wall strains are determined for unloading (dimensionless)
εh horizontal strain (dimensionless)
Δεsh soil horizontal strain (dimensionless)
Δεshc soil horizontal strain for loading (dimensionless)
Δεshr soil horizontal strain for unloading(dimensionless)
ν0 at-rest Poisson ratio (dimensionless)
νd at-rest Poisson ratio for unloading (dimensionless)
νs soil Poisson ratio (dimensionless)
effective horizontal stress (kPa)

σh
effective vertical stress (kPa)

σv
effective horizontal stress inside the cell (kPa)

σ hi
effective horizontal stress outside the cell (kPa)

σ he
minor principal stress (kPa)

σ3
maximum horizontal stress in loading (kPa)

σ hc
residual horizontal stress (kPa)

σ hr
self-weight vertical stress (kPa)

σ v0
maximum vertical stress in loading (kPa)

σ vc
effective soil friction angle (◦ )

ϕ

11
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

References Han, J., 2015. Principles and Practice of Ground Improvement. John Wiley & Sons,
Hoboken, NJ, USA.
Han, J., Yang, X., Leshchinsky, D., Parsons, R.L., 2008. Behavior of geocell-reinforced
AASHTO, 2015. Mechanistic-empirical Pavement Design Guide–A Manual of Practice.
sand under a vertical load. Transp. Res. Rec. J. Transp. Res. Board 2045 95–101.
Ahmed, M., Hany, E.N., 2020. Coupled TDA–geocell stress-bridging system for buried
https://doi.org/10.3141/2045-11.
corrugated metal pipes. J. Geotech. Geoenviron. Eng. 146, 4020052 https://doi.org/
Hegde, A., 2017. Geocell reinforced foundation beds-past findings, present trends and
10.1061/(ASCE)GT.1943-5606.0002279.
future prospects: a state-of-the-art review. Construct. Build. Mater. 154, 658–674.
Al-Qadi, I.L., Hughes, J.J., 2000. Field evaluation of geocell use in flexible pavements.
https://doi.org/10.1016/j.conbuildmat.2017.07.230.
Transp. Res. Rec. J. Transp. Res. Board 1709, 26–35. https://doi.org/10.3141/1709-
Hegde, A., Sitharam, T.G., 2015. Joint strength and wall deformation characteristics of a
04.
single-cell geocell subjected to uniaxial compression. Int. J. GeoMech. 15, 04014080
Arvin, M.R., Beigi, Z., 2018. A design scheme for geocell-reinforced foundations based on
https://doi.org/10.1061/(ASCE)GM.1943-5622.0000433.
the lower bound limit analysis method. Comput. Geotech. 97, 69–77. https://doi.
Henkel, D.J., Gilbert, G.D., 1952. The effect measured of the rubber membrane on the
org/10.1016/J.COMPGEO.2018.01.001.
triaxial compression strength of clay samples. Geotechnique 3, 20–29. https://doi.
Avesani Neto, J.O., 2019. Application of the two-layer system theory to calculate the
org/10.1680/geot.1952.3.1.20.
settlements and vertical stress propagation in soil reinforcement with geocell.
Hicks, R.G., Monismith, C.L., 1971. Factors influencing the resilient response of granular
Geotext. Geomembranes 47, 32–41. https://doi.org/10.1016/j.
materials. Highw. Res. Rec. 345, 15–31.
geotexmem.2018.09.003.
Huang, Y.H., 1993. Pavement Analysis and Design.
Avesani Neto, J.O., Bueno, B.S., Futai, M.M., 2015. Evaluation of a calculation method
Hunt, R.E., 2005. Geotechnical Engineering Investigation Handbook. CRC Press.
for embankments reinforced with geocells over soft soils using finite-element
Indraratna, B., Biabani, M.M., Nimbalkar, S., 2015. Behavior of geocell-reinforced
analysis. Geosynth. Int. 22, 439–451. https://doi.org/10.1680/jgein.15.00024.
subballast subjected to cyclic loading in plane-strain condition. J. Geotech.
Avesani Neto, J.O., Bueno, B.S., Futai, M.M., 2013. A bearing capacity calculation
Geoenviron. Eng. 141, 1–16. https://doi.org/10.1061/(ASCE)GT.1943-
method for soil reinforced with a geocell. Geosynth. Int. 20, 129–142. https://doi.
5606.0001199.
org/10.1680/gein.13.00007.
Jewel, R.A., 1980. Some Effects of Reinforcement on the Mechanical Behaviour of Soils.
Bathurst, R.J., Jarrett, P.M., 1988. Large-scale model tests of geocomposite mattresses
University of Cambridge.
over peat subgrades. Transp. Res. Rec. No. 1188 Eff. Geosynth. soil Prop. Environ.
Kief, O., Schary, Y., Pokharel, S.K., 2015. High-modulus geocells for sustainable highway
pavement Syst. 28–36.
infrastructure. Indian Geotech. J. 45, 389–400. https://doi.org/10.1007/s40098-
Bathurst, R.R.J., Karpurapu, R., 1993. Large-scale triaxial compression testing of geocell-
014-0129-z.
reinforced granular soils. Geotech. Test J. 16, 296. https://doi.org/10.1520/
Koerner, R.M., 1994. Design with Geosynthetics, third ed. Prentice Hall, Englewood
GTJ10050J.
Cliffs, NJ, USA.
Biswas, A., Krishna, A.M., 2019. Behaviour of circular footing resting on layered
Krishnaswamy, N., Rajagopal, K., Latha, G.M., 2000. Model studies on geocell supported
foundation: sand overlying clay of varying strengths. Int. J. Geotech. Eng. 13, 9–24.
embankments constructed over a soft clay foundation. Geotech. Test J. 23, 45.
https://doi.org/10.1080/19386362.2017.1314242.
https://doi.org/10.1520/GTJ11122J.
Biswas, A., Krishna, A.M., 2018. Behaviour of geocell–geogrid reinforced foundations on
Lalima, B., Sowmiya, C., Kumar, D.S., 2020. Performance evaluation of coal mine
clay subgrades of varying strengths. Int. J. Phys. Model. Geotech. 18, 301–314.
overburden as a potential subballast material in railways with additional
https://doi.org/10.1680/jphmg.17.00013.
improvement using geocell. J. Mater. Civ. Eng. 32, 4020200 https://doi.org/
Biswas, A., Krishna, A.M., Dash, S.K., 2016. Behavior of geosynthetic reinforced soil
10.1061/(ASCE)MT.1943-5533.0003269.
foundation systems supported on stiff clay subgrade. Int. J. GeoMech. 16, 04016007
Lambe, T.W., Whitman, R.V., 1991. Soil Mechanics. John Wiley & Sons.
https://doi.org/10.1061/(ASCE)GM.1943-5622.0000559.
Latha, G.M., 2011. Design of geocell reinforcement for supporting embankments on soft
Biswas, A., Murali Krishna, A., Dash, S.K., 2013. Influence of subgrade strength on the
ground. In: 12 Th International Conference of International Association for
performance of geocell-reinforced foundation systems. Geosynth. Int. 20, 376–388.
Computer Methods and Advances in Geomechanics (IACMAG). Techno-Press,
https://doi.org/10.1680/gein.13.00025.
pp. 117–130.
Bolton, M.D., 1986. The strength and dilatancy of sands. Geotechnique 36, 65–78.
Latha, G.M., 2000. Investigations on the Behaviour of Geocell Supported Embankments.
https://doi.org/10.1680/geot.1986.36.1.65.
Indian Institute of Technology Madras, Chennai.
Campanella, R.G., Vaid, Y.P., 1972. A simple K o triaxial cell. Can. Geotech. J. 9,
Latha, G.M., Kumar, D.S., K, R., 2009. Numerical simulation of the behavior of geocell
249–260. https://doi.org/10.1139/t72-029.
reinforced sand in foundations. Int. J. GeoMech. 9, 143–152. https://doi.org/
Chen, R.H., Huang, Y.W., Huang, F.C., 2013. Confinement effect of geocells on sand
10.1061/(ASCE)1532-3641(2009)9:4(143).
samples under triaxial compression. Geotext. Geomembranes 37, 35–44. https://doi.
Latha, G.M., Rajagopal, K., 2007. Parametric finite element analyses of geocell-supported
org/10.1016/j.geotexmem.2013.01.004.
embankments. Can. Geotech. J. 44, 917–927. https://doi.org/10.1139/T07-039.
De Barros, S.T., 1966. Deflection factor charts for two-and three-layer elastic systems.
Latha, G.M., Rajagopal, K., Krishnaswamy, N.R., 2006. Experimental and theoretical
Highw. Res. Rec.
investigations on geocell-supported embankments. Int. J. GeoMech. 6, 30–35.
Dowling, N.E., 2013. Mechanical Behavior of Materials : Engineering Methods for
https://doi.org/10.1061/(ASCE)1532-3641(2006)6:1(30).
Deformation, Fracture, and Fatigue, fourth ed. Pearson, Boston.
Latha, G.M., Somwanshi, A., 2009. Effect of reinforcement form on the bearing capacity
Duncan, J.M., Bursey, A., 2013. Soil modulus correlations. In: Foundation Engineering in
of square footings on sand. Geotext. Geomembranes 27, 409–422. https://doi.org/
the Face of Uncertainty, Proceedings. American Society of Civil Engineers, Reston,
10.1016/j.geotexmem.2009.03.005.
VA, pp. 321–336. https://doi.org/10.1061/9780784412763.026.
Leshchinsky, B., Ling, H., 2013. Effects of geocell confinement on strength and
Duncan, J.M., Byrne, P., Wong, K.S.K., Mabry, P., 1980. Strength, Stress–Strain and Bulk
deformation behavior of gravel. J. Geotech. Geoenviron. Eng. 139, 340–352. https://
Modulus Parameters for Finite Element; Analyses of Stresses and Movements in Soil
doi.org/10.1061/(ASCE)GT.1943-5606.0000757.
Masses -, Geotechnical Engineering - Research Report UCB/GT/80-01. California.
Liu, Y., Deng, A., Jaksa, M., 2018. Three-dimensional modeling of geocell-reinforced
Duncan, J.M., Seed, R.B., 1986. Compaction-induced earth pressures under K0-
straight and curved ballast embankments. Comput. Geotech. 102, 53–65. https://
conditions. J. Geotech. Eng. 112, 1–22. https://doi.org/10.1061/(ASCE)0733-9410
doi.org/10.1016/j.compgeo.2018.05.011.
(1986)112:1(1).
Livneh, M., Livneh, N.A., 2014. Design of railway trackbeds with geocells. In:
Dunlap, W.S., 1963. A Report on a Mathematical Model Describing the Deformation
Proceedings of the International Conference on Road and Rail Infrastructure CETRA.
Characteristics of Granular Materials. Tex. Technical Report 1, Project 2-8-62-27.
Look, B.G., 2014. Handbook of Geotechnical Investigation and Design Tables. CRC Press.
Dyer, M.R., Milligan, G.W.E., 1984. A photoelastic investigation of the interaction of a
Mayne, P.M.W., Kulhawy, F.H., 1982. Ko-OCR relationship in soil. J. Geotech. Eng. Div.
cohesionless soil with reinforcement placed at different orientations. In:
108, 851–872.
Renforcement En Place Des Sols et Des Roches. Colloque International, pp. 257–262.
Mhaiskar, S.Y., Mandal, J.N., 1996. Investigations on soft clay subgrade strengthening
Ehrlich, M., Mitchell, J.K., 1994. Working stress design method for reinforced soil walls.
using geocells. Construct. Build. Mater. 10, 281–286. https://doi.org/10.1016/
J. Geotech. Eng. 120, 625–645. https://doi.org/10.1061/(ASCE)0733-9410(1994)
0950-0618(95)00083-6.
120:4(625).
Moffatt, M., 2019. Guide to Pavement Technology: Part 2: Pavement Structural Design,
Fazeli Dehkordi, P., Ghazavi, M., Ganjian, N., Karim, U.F.A., 2019. Effect of geocell-
4.3. Austroads Ltd., Sidney, Australia.
reinforced sand base on bearing capacity of twin circular footings. Geosynth. Int. 26,
Moghaddas Tafreshi, S.N., Joz Darabi, N., Dawson, A.R., Azizian, M., 2020. Experimental
224–236. https://doi.org/10.1680/jgein.19.00047.
evaluation of geocell and EPS geofoam as means of protecting pipes at the bottom of
Garcia, R.S., Avesani Neto, J.O., 2020. Accuracy of the theory of equivalent thickness in
repeatedly loaded trenches. Int. J. GeoMech. 20, 4020023.
two-layer system Applications with geocell-reinforced soils. In: Geoamericas - 4th
Moghaddas Tafreshi, S.N., Shaghaghi, T., Tavakoli Mehrjardi, G., Dawson, A.R.,
Pan American Conference on Geosynthetics. Rio de Janeiro.
Ghadrdan, M., 2015. A simplified method for predicting the settlement of circular
George, A.M., Banerjee, A., Puppala, A.J., Saladhi, M., 2019. Performance evaluation of
footings on multi-layered geocell-reinforced non-cohesive soils. Geotext.
geocell-reinforced reclaimed asphalt pavement (RAP) bases in flexible pavements.
Geomembranes 43, 332–344. https://doi.org/10.1016/j.geotexmem.2015.04.006.
Int. J. Pavement Eng. 1–11. https://doi.org/10.1080/10298436.2019.1587437.
Monnet, A., 1994. Module de réaction, coefficient de décompression, au sujet des
Guo, J., Han, J., Schrock, S.D., Parsons, R.L., 2015. Field evaluation of vegetation growth
paramètres utilisés dans la méthode de calcul élasto-plastique des soutènements.
in geocell-reinforced unpaved shoulders. Geotext. Geomembranes 43, 403–411.
Rev. Fr. Geotech. 67–72. https://doi.org/10.1051/geotech/1994066067.
https://doi.org/10.1016/j.geotexmem.2015.04.013.
NCHRP, 2004. Guide for Mechanistic–Empirical Design of New and Rehabilitated
Gurbuz, A., Mertol, H.C., 2012. Interaction between assembled 3D honeycomb cells
Pavement Structures., Final Rep., NCHRP Project 1-37A. Transportation Research
produced from high density polyethylene and a cohesionless soil. J. Reinforc. Plast.
Board, Washington, DC.
Compos. 31, 828–836. https://doi.org/10.1177/0731684412447529.
Nicks, J., Adams, M., 2013. Friction Angles of Open-Graded Aggregates from Large-Scale
Halder, K., Chakraborty, D., 2020. Influence of soil spatial variability on the response of
Direct Shear Testing (FHWA-HRT-13-068).
strip footing on geocell-reinforced slope. Comput. Geotech. 122, 103533 https://doi.
org/10.1016/j.compgeo.2020.103533.

12
R.S. Garcia and J.O. Avesani Neto Geotextiles and Geomembranes xxx (xxxx) xxx

Nicks, J.E., Gebrenegus, T., Adams, M.T., 2015. Strength Characterization of Open- Song, F., Liu, H., Yang, B., Zhao, J., 2019. Large-scale triaxial compression tests of
Graded Aggregates for Structural Backfills (FHWA-HRT-15-034). Federal Highway geocell-reinforced sand. Geosynth. Int. 26, 388–395. https://doi.org/10.1680/
Administration. Office of Infrastructure, United States. jgein.19.00019.
Nikolaides, A., 2014. Highway Engineering: Pavements, Materials and Control of Tafreshi, S.N.M., Darabi, N.J., Dawson, A.R., 2020. Combining EPS geofoam with geocell
Quality. CRC Press. to reduce buried pipe loads and trench surface rutting. Geotext. Geomembranes.
Palmeira, E.M., 2009. Soil–geosynthetic interaction: modelling and analysis. Geotext. https://doi.org/10.1016/j.geotexmem.2019.12.011.
Geomembranes 27, 368–390. https://doi.org/10.1016/J. Tanyu, B.F., Aydilek, A.H., Lau, A.W., Edil, T.B., Benson, C.H., 2013. Laboratory
GEOTEXMEM.2009.03.003. evaluation of geocell-reinforced gravel subbase over poor subgrades. Geosynth. Int.
Pokharel, S.K., 2010. Experimental Study on Geocell-Reinforced Bases under Static and 20, 47–61. https://doi.org/10.1680/gein.13.00001.
Dynamic Loading. University of Kansas. Tavakoli Mehrjardi, G., Behrad, R., Moghaddas Tafreshi, S.N., 2019. Scale effect on the
Pokharel, S.K., Han, J., Leshchinsky, D., Parsons, R.L., 2018. Experimental evaluation of behavior of geocell-reinforced soil. Geotext. Geomembranes 47, 154–163. https://
geocell-reinforced bases under repeated loading. Int. J. Pavement Res. Technol. 11, doi.org/10.1016/j.geotexmem.2018.12.003.
114–127. https://doi.org/10.1016/j.ijprt.2017.03.007. Thom, N., 2013. Principles of Pavement Engineering, second ed. Thomas Telford Ltd.
Pokharel, S.K., Han, J., Leshchinsky, D., Parsons, R.L., Halahmi, I., 2010. Investigation of https://doi.org/10.1680/ppe.58538. second ed.
factors influencing behavior of single geocell-reinforced bases under static loading. Timoshenko, S.P., Woinowsky-Krieger, S., 1959. Theory of Plates and Shells, second ed.
Geotext. Geomembranes 28, 570–578. https://doi.org/10.1016/j. McGraw-hill.
geotexmem.2010.06.002. Trautmann, C.H., Beech, J.F., O’Rourke, T.D., McGuire, W., Wood, W.A., Capano, C.,
Presto, 2008. Geoweb Load Support System – Technical Overview [WWW Document]. 1983. Transmission-line Structure Foundations for Uplift-Compression Loading.
https://www.geofabrics.co/sites/default/files/technicaldata/Geoweb-Load-Techn Final Report.
ical-OverviewNZ.pdf. Uzan, J., 1992. Resilient characterization of pavement materials. Int. J. Numer. Anal.
Priti, M., Sivakumar, B.G.L., Maheshwari, P., Babu, G.L.S., Priti, M., Sivakumar, B.G.L., Methods GeoMech. 16, 453–459.
2017. Nonlinear deformation analysis of geocell reinforcement in pavements. Int. J. Uzan, J., 1985. Characterization of granular material. Transport. Res. Rec. 1022, 52–59.
GeoMech. 17, 04016144 https://doi.org/10.1061/(ASCE)GM.1943-5622.0000854. Vega, E., van Gurp, C., Kwast, E., 2018. Geokunststoffen Als Funderingswapening in
PRS, 2020. PRS NEOLOY® GEOCELLS CATEGORY D - Cellular Confinement System Ongebonden Funderingslagen (Geosynthetics for Reinforcement of Unbound Base
-SPECIIFIICATIIONS [WWW Document]. https://www.prs-med.com/wp-content/up and Subbase Pavement Layers). SBRCUR/CROW, Nethersland.
loads/2017/10/PRS-Geotech-Data-Spec-Neoloy-Category-D-v8.2.pdf. accessed Venkateswarlu, H., Hegde, A.M., 2019. Effect of infill materials on vibration isolation
3.6.20. efficacy of geocell reinforced soil beds. Can. Geotech. J. cgj-2019-0135. https://doi.
Punetha, P., Nimbalkar, S., Khabbaz, H., 2020. Evaluation of additional confinement for org/10.1139/cgj-2019-0135.
three-dimensional geoinclusions under general stress state. Can. Geotech. J. 57, Ventsel, E., Krauthammer, T., 2001. Thin Plates and Shells : Theory , Analysis , and
453–461. https://doi.org/10.1139/cgj-2018-0866. Applications, first ed. CRC Press.
Rajagopal, K., Chandramouli, S., Parayil, A., Iniyan, K., 2014. Studies on geosynthetic- Webster, S.L., 1981. Investigation of Beach Sand Trafficability Enhancement Using Sand-
reinforced road pavement structures. Int. J. Geotech. Eng. 8, 287–298. https://doi. Grid Confinement and Membrane Reinforcement Concepts. Report 2. Sand Test
org/10.1179/1939787914Y.0000000042. Sections 3 and 4. Vicksburg, Mississippi.
Rajagopal, K., Krishnaswamy, N.R., Latha, G.M., 1999. Behaviour of sand confined with Webster, S.L., 1979. Investigation of Beach Sand Trafficability Enhancement Using Sand-
single and multiple geocells. Geotext. Geomembranes 17, 171–184. https://doi.org/ Grid Confinement and Membrane Reinforcement Concepts. Report 1. Sand Test
10.1016/S0266-1144(98)00034-X. Sections 1 and 2. Vicksburg, Mississippi.
Rajagopal, K., Veeragavan, A., Chandramouli, S., 2012. Studies on geocell reinforced Webster, S.L., Alford, S.J., 1978. Investigation of Construction Concepts for Pavements
road pavement structures. In: GA 2012 - 5th Asian Regional Conference on across Soft Ground. Vicksburg, Mississippi.
Geosynthetics: Geosynthetics for Sustainable Adaptation to Climate Change. Witczak, M.W., Yoder, E.J., 1975. Principles of Pavement Design. Wiley.
Bangkok, Thailand, pp. 497–502. Wu, K.J., Austin, D.N., 1992. Three-dimensional polyethylene geocells for erosion
Ruge, J.C., Gomez, J.G., Moreno, C.A., 2020. Analysis of the creep and the influence on control and channel linings. Geotext. Geomembranes 11, 611–620. https://doi.org/
the modulus improvement factor (MIF) in polyolefin geocells using the stepped 10.1016/0266-1144(92)90035-9.
isothermal method. In: Gomez, J.G. (Ed.), Geopolymers and Other Geosynthetics. Yang, X., Han, J., 2013. Analytical model for resilient modulus and permanent
IntechOpen, Rijeka. https://doi.org/10.5772/intechopen.88518. Ch. 9. deformation of geosynthetic-reinforced unbound granular material. J. Geotech.
Saride, S., Gautam, D., Madhav, M.R., Vijay, K.R., 2016. Performance evaluation of Geoenviron. Eng. 139, 1443–1453. https://doi.org/10.1061/(ASCE)GT.1943-
geocell reinforced granular subbase (GSB) layers through field trials. J. Indian Roads 5606.0000879.
Congr. 76, 249–257. Yang, X., Han, J., Parsons, R.L., Leshchinsky, D., 2010. Three-dimensional numerical
Saride, S., Gowrisetti, S., Sitharam, T.G., Puppala, A.J., 2009. Numerical simulation of modeling of single geocell-reinforced sand. Front. Architect. Civ. Eng. China 4,
geocell-reinforced sand and clay. Proc. Inst. Civ. Eng. - Gr. Improv. 162, 185–198. 233–240. https://doi.org/10.1007/s11709-010-0020-7.
https://doi.org/10.1680/grim.2009.162.4.185. Zhang, L., Ou, Q., Zhao, M., 2018. Double-beam model to analyze the performance of a
Siabil, S.M.A.G., Tafreshi, S.N.M., Dawson, A.R., 2020. Response of pavement pavement structure on geocell-reinforced embankment. J. Eng. Mech. 144, 6018002
foundations incorporating both geocells and expanded polystyrene (EPS) geofoam. https://doi.org/10.1061/(ASCE)EM.1943-7889.0001453.
Geotext. Geomembranes 48, 1–23. https://doi.org/10.1016/j. Zhang, L., Zhao, M., Shi, C., Zhao, H., 2010. Bearing capacity of geocell reinforcement in
geotexmem.2019.103499. embankment engineering. Geotext. Geomembranes 28, 475–482. https://doi.org/
Sitharam, T.G., Hegde, A., 2013. Design and construction of geocell foundation to 10.1016/J.GEOTEXMEM.2009.12.011.
support the embankment on settled red mud. Geotext. Geomembranes 41, 55–63. Zhang, L., Zhao, M., Zou, X., Zhao, H., 2009. Deformation analysis of geocell
https://doi.org/10.1016/j.geotexmem.2013.08.005. reinforcement using Winkler model. Comput. Geotech. 36, 977–983. https://doi.
Song, F., Liu, H., Ma, L., Hu, H., 2018. Numerical analysis of geocell-reinforced retaining org/10.1016/J.COMPGEO.2009.03.005.
wall failure modes. Geotext. Geomembranes 46, 284–296. https://doi.org/10.1016/
j.geotexmem.2018.01.004.

13

You might also like