Kar Gar 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

European Journal of Environmental and Civil Engineering

ISSN: 1964-8189 (Print) 2116-7214 (Online) Journal homepage: http://www.tandfonline.com/loi/tece20

Influence of reinforcement stiffness and strength


on load-settlement response of geocell-reinforced
sand bases

Mohsen Kargar & S. Majdeddin Mir Mohammad Hosseini

To cite this article: Mohsen Kargar & S. Majdeddin Mir Mohammad Hosseini (2016):
Influence of reinforcement stiffness and strength on load-settlement response of geocell-
reinforced sand bases, European Journal of Environmental and Civil Engineering, DOI:
10.1080/19648189.2016.1214181

To link to this article: http://dx.doi.org/10.1080/19648189.2016.1214181

Published online: 28 Jul 2016.

Submit your article to this journal

Article views: 2

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tece20

Download by: [La Trobe University] Date: 02 August 2016, At: 03:30
European Journal of Environmental and Civil Engineering, 2016
http://dx.doi.org/10.1080/19648189.2016.1214181

Influence of reinforcement stiffness and strength on load-settlement


response of geocell-reinforced sand bases
Mohsen Kargar and S. Majdeddin Mir Mohammad Hosseini*

Department of Civil and Environmental Engineering, Amirkabir University of Technology,


Tehran, Iran
(Received 4 October 2015; accepted 6 July 2016)
Downloaded by [La Trobe University] at 03:30 02 August 2016

Improvement in the performance of a strip footing supported on geocell-reinforced


sand bed is investigated through series of experimental tests on a reduced scale phys-
ical model. Geocells with scaled-down mechanical and geometrical properties pro-
duced from various material types consisting of woven geotextile, nonwoven
geotextile and biaxial geogrid are used as the reinforcement and the influence of geo-
cell strength and stiffness on the enhancement of bearing capacity and settlement of
the surface footing is described. The results indicate that stiffness of geocell makes a
significant contribution on the performance of geocell-reinforced soil. For the geo-
cells used in the tests with the height equal to footing width, for example at footing
settlement ratio of 6%, improvement in the bearing capacity of geocell with the stiff-
est material is found to be 1.5 times the geocell with the softest material. The failure
mechanism of geocell-reinforced foundation system has been shown to depend
highly on geocell modulus and height. It is also concluded that ultimate tensile
strength of geocell material is generally not influential at load-settlement behaviour
of the footing at practical settlement levels. However, if flexural rupture of reinforce-
ment controls the failure mechanism, geocell strength affects the ultimate bearing
capacity.
Keywords: bearing capacity; settlement; reinforced sand; geocell strength; geocell
stiffness

1. Introduction
Geosynthetics are extensively used as soil reinforcement to improve soil behaviour.
Many investigations have been carried out to evaluate the behaviour of different types
of geosynthetic-reinforced soil (e.g. Davarci, Ornek, & Turedi, 2014; Ferreira, Vieira,
Lopes, & Carlos, 2015; Zhuang, Wang, Liu, & Chu, 2013). One of the most recent
types of geosynthetics is geocell that can be effectively used as soil reinforcement for
embankments, foundations, retaining walls and slopes. Since geocell has a three-
dimensional geometry, it provides a great lateral confinement for the infill material
without dependence on the friction or interlocking with the infill soil (Yang, 2010). If
geocell-reinforced soil is used under the footing or embankments, the bearing capacity
of subgrade increases and settlements decreases as a result of lateral confinement,
membrane action and redistribution of pressure to a wider area in the underlying soil
(Zhang, Zhao, Shi, & Zhao, 2010).

*Corresponding author. Email: smmirhos@aut.ac.ir

© 2016 Informa UK Limited, trading as Taylor & Francis Group


2 M. Kargar and S.M. Mir Mohammad Hosseini

Several investigations have been carried out on stress–strain behaviour of geocell


encased soil samples on triaxial condition (e.g. Chen, Huang, & Huang, 2013; Khedkar
& Mandal, 2009; Nair & Madhavi Latha, 2015). The results of large-scale triaxial tests
on samples of geocell-reinforced sand have shown that shear strength of geocell encased
sand increases since an apparent cohesion is induced in the sand due to confinement
while the internal friction of the infill sand remains unchanged (Bathurst & Karpurapu,
1993; Rajagopal, Krishnaswamy, & Madhavi Latha, 1999). The induced apparent cohe-
sion is related to axial strain of geocell at failure, diameter of geocell pockets, geocell
material modulus and internal friction of the infill sand. Based on findings on strength
and stiffness of the geocell sand composite from triaxial tests, Madhavi Latha, Dash,
and Rajagopal (2009) proposed a design method to analyse the geocell-reinforced
foundations numerically regarding equivalent properties of the geocell-reinforced sand
composite.
Several researchers have conducted laboratory tests to study the improvement effect
Downloaded by [La Trobe University] at 03:30 02 August 2016

of geocell reinforcements under vertical loading via small-scale and large-scale physical
models (e.g. Bathurst & Jarret, 1988; Biswas, Murali Krishna, & Dash, 2013; Dash,
Krishnaswamy, & Rajagopal, 2001; Gurbuz & Mertol, 2012; Indraratna, Biabani, &
Nimbalkar, 2015; Leshchinsky & Ling, 2013; Madhavi Latha & Somwanshi, 2009;
Moghaddas Tafreshi & Dawson 2012; Moghaddas Tafreshi, Khalaj, & Dawson, 2014;
Sireesh, Sitharam, & Dash, 2009; Tanyu, Aydilek, Lau, Edil, & Benson, 2013).
Moghaddas Tafreshi and Dawson (2010) compared the improvement in bearing capacity
of planar geotextiles and three-dimensional geotextiles and concluded that the geocell
reinforcement system behaves much stiffer, carries greater loading and settles less than
the equivalent planar reinforcement system. Pokharel, Han, Leshchinsky, Parsons, and
Halahmi (2010) investigated the parameters affecting the behaviour of single geocell-
reinforced soil under static loading and reported that performance of geocell-reinforced
sand depends on the elastic modulus of the geocell.
Dash (2010) investigated the influence of relative density of soil on the performance
of geocell-reinforced sand foundations and suggested that for the effective utilisation of
geocell reinforcement, the foundation soil should be compacted to higher density. Dash
(2012) concluded that the strength, stiffness, aperture size and orientation of the ribs of
geocell prepared using geogrids influence the performance of the reinforced sand foun-
dations. Avesani Neto, Bueno, and Futai (2013) proposed an analytical approach to pre-
dict the bearing capacity of geocell-reinforced soil by taking into account the geometric
characteristics of geocell reinforcement. Moghaddas Tafreshi, Shaghaghi, Tavakoli
Mehrjardi, Dawson, and Ghadrdan (2015) presented a simplified method for predicting
the load-settlement response of geocell-reinforced sand considering the effects of soil
and reinforcement stiffness.
Despite the extensive research carried out on the performance of geocell-reinforced
sand foundations, there is a lack of quantitative investigation about the influence of rein-
forcement strength and stiffness on the load-settlement response of footing and interac-
tion mechanisms of geocell and infill soil for different material types. Results of
previous researches on geowebs and geogrid cells (Bathurst & Jarret, 1988; Dash, 2012)
as well as loading tests on commercial single pocket geocells (Pokharel et al., 2010)
imply the importance of reinforcement tensile properties. However, some of the pro-
posed equations to calculate bearing capacity of geocell-reinforced soil foundations do
not take the significance of geocell stiffness into account (e.g. Avesani Neto et al.,
2013; Koerner, 2005).
European Journal of Environmental and Civil Engineering 3

In this research, series of experimental tests have been conducted on a reduced scale
physical model of a foundation on geocell-reinforced sand. Different types of geocell
material with wide ranges of mechanical properties in terms of strength and stiffness
fabricated by woven geotextile, nonwoven geotextile and biaxial geogrid are considered
in the experimental programme to quantitatively evaluate the significance of geocell
strength and stiffness on bearing capacity and settlement of the footing. Furthermore,
unlike several former researches, in current paper factory-made geowebs or geocells
formed by commercial geogrids are not used. Instead, hand-made geocells of scaled-
down geosynthetics are considered in the experimental design to account for scale
effects.

2. Physical model
A physical model is developed to study the pressure-settlement behaviour of footings on
Downloaded by [La Trobe University] at 03:30 02 August 2016

reinforced sand. The schematic diagram of test set-up is shown in Figure 1. Main parts
of the laboratory apparatus include:

• Footing model: A thick plate made of aluminium measuring B = 50 mm in width


and L = 340 mm in length is used to represent a strip footing. The thickness of
the footing model is high enough to be considered a rigid footing and the base of
the footing is made rough by sticking sand paper.
• Soil tank: The length of soil container is considered 800 mm (16B) which is ade-
quate to overcome boundary effects. Observations of the soil deformation during

Figure 1. Schematic diagram of physical model set-up.


4 M. Kargar and S.M. Mir Mohammad Hosseini

the tests and after failure verify that side walls of the test tank do not affect the
footing performance. The width of the container should be equal to the length of
the footing in order to establish the plane strain conditions. However, 1-mm wide
gap is provided to avoid any contact between the footing and side walls and as a
result soil box width is measured 342 mm.
The height of sand sample is considered 560 mm which is large enough (more
than 10B) to eliminate the effect of bottom boundary on the result. The bottom
and a lateral wall of the soil tank in the longitudinal direction is made of a thick
steel plate and the other faces of the tank are made of plexiglass sheets which are
restrained by steel frames to provide rigidity of the tank. The plexiglass sheets in
the sides of the tank are used to reduce the effect of friction between soil particles
and tank walls. In addition, they allow visualisation of soil particles movement
and failure zones during loading tests.
• Raining system: The sand sample in this research is prepared using the air pluvia-
Downloaded by [La Trobe University] at 03:30 02 August 2016

tion technique (Kolbsuzewski, 1948) to achieve uniform sand bed. The height and
rate of raining are pre-calibrated to obtain the desired relative density.
• Loading equipment: A pneumatic cylinder attached to a compressed air tank pro-
vides loading. It is able to apply monotonic loading in pressure-controlled condi-
tion. The capacity of the loading system is sufficient to reach the ultimate bearing
capacity of the foundation in all tests.
• Data acquisition: During the tests, loading and settlement of the footing are mea-
sured by an S-shaped load cell having 25 kN capacity with an accuracy of ±.02%
full scale and a linear variable displacement transducer having stroke of 100 mm
with an accuracy of ±.05% full range. The data are recorded using a data acquisi-
tion system.

3. Materials
3.1. Sand
The soil used in the experimental tests is a dry silica sand with grain sizes between .6
and 2.36 mm and specific gravity of 2.67 (Gs = 2.67). Particle size distribution of the
soil is demonstrated in Figure 2. It indicates that the relatively uniform soil used in this
research can be classified as poorly-graded sand (SP) according to Unified Soil Classifi-
cation System. Maximum and minimum dry densities of soil are determined as 16.9 and

100

D10=0.75 mm
80
D30=1.29 mm
Passing Percentage (%)

D50=1.54 mm
60 D60= 1.63 mm
Cu=2.18
Cc =1.38
40

20

0
0.1 1 10
Grain Size (mm)

Figure 2. Soil particle size distribution curve.


European Journal of Environmental and Civil Engineering 5

14.2 kN/m3, respectively. All of the experimental tests in this paper have been con-
ducted on sand with 72% relative density that corresponds to dry unit weight equal to
16.2 kN/m3.

3.2. Reinforcement
In this research three different materials consisting of woven geotextile, nonwoven geo-
textile and biaxial geogrid are used to prepare geocell reinforcements. Table 1 provides
the engineering properties of the materials used in the tests. Figure 3 also demonstrates
the load-elongation behaviour of the geosynthetic samples in-isolation under uniaxial
tensile load with the rate of 10% strain per minute. It should be noted that depending
on the polymeric material of the reinforcement, the rate of tensile loading might have
considerable effect on the modulus and strength (e.g. HDPE) as reported by Bathurst
and Cai (1994).
Downloaded by [La Trobe University] at 03:30 02 August 2016

To fabricate geocell from the geotextile, strips of geotextile are measured and cut in
particular dimensions and are stitched together in the location of connections using an
industrial sewing machine. This procedure provides uniform and stiff seam at the junc-
tion that is much stronger than the parent material under the applied loading. The geo-
cell made from biaxial geogrid is prepared by attaching the geogrid strips via plastic
belt fasteners at opening apertures. It should be noted that none of the geosynthetic
materials utilised in the experimental tests have real-world practical applications as soil
reinforcement. They are selected of materials with lower mechanical properties com-
pared to commercial reinforcements to comply with the model scale rules in the reduced
scale physical model. They are also chosen to have varying physical, geometrical and
mechanical properties to satisfy the objective of this research programme.
The photos of hand-made geocells of the three different material types are illustrated
in Figure 4. The woven geotextile reinforcement is selected from curtains typically used
in office windows. Nonwoven geotextile is used for the purpose of drainage and geogrid
type application is usually for fencing. In all of the reinforcing samples, geocell pockets
are in diamond shape with equilateral diameter designated by “d”. Height and total
width of the geocell are shown by “h” and “b”, respectively, and are assumed constant

Table 1. The engineering properties of the reinforcement materials used in the tests.
Description Value
Geocell designation W-GT NW-GT B-GG
Geocell material type Woven geotextile Nonwoven geotextile Biaxial geogrid
Polymer PET1 PP2 HDPE3
Thickness, t (mm) .81 3.46 2.06
Ultimate tensile strength, Tu (kN/m) 21.8 19.7 2.8
Secant modulus at 2% strain, J2% (kN/m) 275 18 110
Secant modulus at 5% strain, J5% (kN/m) 72 18 –
Elongation at failure, εu (%) 25.0 54.7 3.7
Aperture opening size (mm) – – 6×6
Aperture opening shape – – Square
1
Polyester (polyethylene terephthalate).
2
Polypropylene.
3
High density polyethylene.
6 M. Kargar and S.M. Mir Mohammad Hosseini

Figure 3. Load-strain behaviour of the reinforcement materials.


Downloaded by [La Trobe University] at 03:30 02 August 2016

Figure 4. Photographic view of geocell reinforcement types used in this research. From left to
right: W-GT, NW-GT, B-GG.

in this research. The length of the geocell reinforcement perpendicular to the footing
width is considered approximately equal to the width of soil tank (length of strip foot-
ing model) to achieve plane strain conditions.

4. Experimental design
4.1. Experimental test schedule
The general configuration of the experimental test and the geometrical parameters used
in this research are presented in Figure 5. All the experimental tests are performed using
a 50-mm width rigid loading plate (B = 50 mm) and the depth of placement of geocell
layer is kept constant. The distance between bottom of footing and top of reinforcement
layer is considered u = 5 mm (equivalent to u/B = .1) which have been shown to be the
optimum value for geocell placement in the literature (e.g. Dash et al., 2001;
Moghaddas Tafreshi & Dawson, 2010). Three different heights are considered for
geocell samples (h/B = .5, 1, 1.5) simulating small, medium and large geocell layers’
height to account for the influence of aspect ratio (h/d) on geocell mattress behaviour
and to observe the effect of geocell strength and stiffness at different geocell heights.
The diameter of geocell pockets with respect to footing width is kept constant equal to
d/B = 1 in all geocell samples. Width of all the geocells used in this study is also con-
sidered five times the footing width (b/B = 5) that is found to be the maximum value
beyond which the improvement effect is negligible (Kargar & Mir Mohammad Hosseini,
in press; Sitharam & Sireesh, 2005).
European Journal of Environmental and Civil Engineering 7

Figure 5. Geometry of the geocell-reinforced foundation bed: (a) 3D view, (b) Cross section.
Downloaded by [La Trobe University] at 03:30 02 August 2016

In the experimental tests, the sand bed is prepared with the desired density (relative
density, Dr = 72%) by raining sand with the calibrated height and rate in the test tank.
At the particular depth, sand raining is paused and the geocell reinforcement is located
on the surface of sand while the geocell pockets are stretched to its desired diameter.
Then, the infill sand is rained in the pockets of the geocell and raining continues until
the tank is fully filled. The surface of soil is levelled and footing model is placed on the
soil surface in the centre of test tank. The static load is then applied at a constant rate
of 1 kPa per second until failure is reached.

4.2. Experimental model scale effects


To account for the scaling effects in this research the commercial geocells, geogrids and
geotextiles are not utilised in the tests and other types of materials with applications in
lighter works rather than soil reinforcements are used. To quantitatively extend the
results from the model tests in a reduced scale experimental work to the field, scaling
rules should be taken into account. According to dimensional analysis, the modulus of
the soil and the reinforcement in the field should be, respectively, N and N2 times of the
corresponding properties in the model if N is the geometrical model scale (Fakher &
Jones, 1996).
Figure 6 shows the idealised relationship between model and prototype conditions in
reduced scale physical models (Viswanadham & König, 2004). In this Figure, J and Tu
are the modulus and ultimate strength of reinforcement in tension test and K and qu are
the stiffness and ultimate bearing capacity of geocell-reinforced sand bed. Indexes m
and p represent model and prototype, respectively. If the characteristics of reinforcement
in-isolation tensile behaviour in model and prototype is according to Figure 6(a), the rel-
evance of model and prototype (scale factor = N) pressure-settlement response of the
footing is anticipated to be identical to Figure 6(b). Therefore, quantitative conclusions
of the results of the experimental tests to be applicable in full-scale field can be
extracted based on Figure 6 provided that scaling requirements are fulfilled.
Table 2 indicates the properties of geosynthetic materials used in this research scaled
to field by assuming scale factor of N = 4. As can be seen, the modulus of geocell in
real-life projects should be 4400 kN/m (275 kN/m × 16) for W-GT, 288 kN/m
(18 kN/m × 16) for NW-GT and 1760 kN/m (110 kN/m × 16) for B-GG which are much
8 M. Kargar and S.M. Mir Mohammad Hosseini

(a) (b)
Prototype Prototype
(Tu)p (q u)p

(Tu)p=N2.(Tu)m (qu)p=N.(q u)m


Tensile Strength, T

Bearing Pressure, q
Jp=N2.Jm Kp=N.Km Kp
1
(q u)m
Model
Model
(Tu)m
Jp Km
1
1
Jm
1
εu (s/B)u
Strain, ε (%) Relative Settlement, s/B (%)

Figure 6. Idealised model and prototype relationship for (a) geosynthetic tensile behaviour and
Downloaded by [La Trobe University] at 03:30 02 August 2016

(b) footing pressure-settlement response.

Table 2. Properties of geosynthetics used in this study scaled to prototype (scale factor, N = 4).
Geocell W-GT NW-GT B-GG
Height, h (mm) Model 25-50-75 25-50-75 25-50-75
Prototype 100-200-300 100-200-300 100-200-300
Diameter, d (mm) Model 50 50 50
Prototype 200 200 200
Secant modulus at 2% strain, J(ε = 2%) (kN/m) Model 275 18 110
Prototype 4400 288 1760
Secant modulus at 5% strain, J(ε = 5%) (kN/m) Model 72 18 –
Prototype 1152 288 –
Ultimate tensile strength, Tu (kN/m) Model 21.8 19.7 2.8
Prototype 349 315 45

larger than the modulus of the geocell in the experimental model. It is obvious that
using reinforcement with lower stiffness in practice will result in less improvement fac-
tors compared to the results of this study. In other words, the quantitative results pre-
sented in this paper overestimate the beneficial effect of geocell provided that the
modulus of experimental model geocell is higher than the reduced-scaled modulus of
considered geocell in the field.

4.3. Limitations and applications of experimental tests


In this research, the experimental tests are conducted on the reduced scale model of strip
footing in plane strain conditions. Therefore, for the cases of other footing shapes and
dimensions like square or circular footing with larger sizes, a three-dimensional physical
model would be very useful. Furthermore, as mentioned in the previous section, the
quantitative validity of the test results depends highly on scaling rules. For example,
considering width of footing equal to 800 mm in real life, the scale factor would be
800/50 = 16. Then the stiffness of geosynthetics to be used at full scale should be 256
times as great as in the model test which is generally not in the range of conventional
geosynthetics stiffness (except for the case of NW-GT). Hence, the results are valid for
low ranges of scale factors.
European Journal of Environmental and Civil Engineering 9

Despite the above-mentioned scaling limitations, the general mechanisms and


comparative parametric study with different types of geocell material are utterly reliable.
Adams and Collin (1997) indicated that general behaviour of the reduced scale model
tests on the reinforced soil is similar to large-scale tests. Nevertheless, it should be noted
that the experimental tests are scheduled to be conducted in a particular geometrical
conditions of geocell relative to the footing i.e. u/B, d/B and b/B are assumed constant.
The validity of the results for different geometrical conditions should also be investi-
gated. The tests are conducted on only one type of soil and the results for other soil
types may be different. Additionally, only surface footing is used in the experimental
work and the effect of footing embedment is not investigated in this paper.
The results obtained from the current experimental work can establish a basis to
design highly-instrumented large-scale model tests and to conduct numerical simula-
tions. The comprehensive analysis of the results from these researches provides more
information regarding the variation of stresses in the soil and strains in the reinforce-
Downloaded by [La Trobe University] at 03:30 02 August 2016

ment (which are not investigated in this paper) to fully understand the behaviour of
geocell-reinforced sand foundation systems in order to develop a design method.

5. Results and discussion


In order to evaluate the performance of geocell-reinforced sand foundation, improve-
ment in the bearing capacity and settlement of the footing is expressed by definition of
nondimensional improvement factors with respect to the ultimate loading capacity as
well as different levels of footing settlement. However, design of most foundations is
based on the allowable level of settlement rather than the ultimate bearing capacity in
practice. The parameters used in the analysis of pressure-settlement behaviour of the
footing are defined in Figure 7. In this Figure, “q” is the bearing pressure at a footing
settlement of “s”, qu and su are ultimate bearing capacity and ultimate settlement of
footing. Indexes r and ur represent reinforced and unreinforced, respectively, and the
improvement factors are described as follows:

• BCRu: The ratio of ultimate bearing capacity of reinforced to unreinforced soil.


• BCRs: The ratio of bearing pressure of reinforced to unreinforced soil at a given
settlement level.

(a) Reinforced (b) Reinforced


(q u)r
(qu)r (su)r
BCRu=(q ) SRu=(s )
u ur
u ur
Bea ring Pressure
Bea ring Pressure

qr sr
BCRs= q PRS=(1- )×100
ur
sur

Unreinforced Unreinforced
(q u)ur

qr
q ur

sur sr (su)ur (su)r


si
Relative Settlement Relative Settlement

Figure 7. Definition of parameters to determine improvements of the reinforced soil in terms of


(a) bearing capacity and (b) settlement.
10 M. Kargar and S.M. Mir Mohammad Hosseini

• SRu: The ratio of ultimate settlement of reinforced to unreinforced soil.


• PRS: Percentage reduction in settlement at a given pressure level.

In this section, the results of the experimental tests are presented and the effect of
mechanical properties of geocell wall material on the performance of the reinforced soil
is discussed through comparison of the above-mentioned improvement factors.

5.1. Bearing pressure-settlement responses


Three series of experimental tests are conducted using different geocell materials at dif-
ferent heights (h/B = .5, 1, 1.5) which are all infilled with dense sand and the other geo-
cell geometrical dimensions are similar in all tests (d/B = 1, b/B = 5). To verify
repeatability of the test data, some tests are performed two or three times and the maxi-
mum difference between the results are found less than 10%. The pressure-settlement
Downloaded by [La Trobe University] at 03:30 02 August 2016

responses of geocells made from different types of materials are presented in Figure 8
for different geocell heights.
Generally, it is evident that the geocell reinforcement significantly increases the ulti-
mate bearing capacity of the footing compared to the unreinforced soil. It also provides
a stiffer foundation resulting in less settlement at a specified pressure. The geocell foun-
dation mattress has turned the pressure-settlement response of the footing into a highly
linear behaviour up to high levels of settlement in comparison with the unreinforced soil
beyond which the slope of pressure-settlement response inclines to gradually decrease
until it reaches the peak value (i.e. ultimate bearing capacity).
Figure 9 shows the soil surface deformation before and after tests. Visual inspection
of surface deformation after failure shows that failure zones in the reinforced soil has

(a) 0 100
Bearing Pressure, q (kPa)
200 300 400 500
(b) 0 100 200
Bearing Pressure, q (kPa)
300 400 500 600 700 800
0 0
Unreinforced Unreinforced
NW-GT 5 NW-GT
5
B-GG B-GG
Footing Settlement, s/B (%)
Footing Settlement, s/B (%)

10
W-GT W-GT
10
15

15 20

25
20
30
25 h/B=0.5 h/B=1
35

30 40

(c) 0 100 200 300


Bearing Pressure, q (kPa)
400 500 600 700 800 900 1000 1100
0
Unreinforced
5
NW-GT
10 B-GG
Footing Settlement, s/B (%)

15 W-GT
20
25
30
35
40
45
h/B=1.5
50
55

Figure 8. Variation of bearing pressure with footing settlement for different geocell material
types: (a) h/B = .5, (b) h/B = 1, (c) h/B = 1.5.
European Journal of Environmental and Civil Engineering 11

(a) (b) (c)

Figure 9. Soil surface deformation (a) before test, (b) after test for unreinforced soil and (c) after
test for reinforced soil (W-GT).

extended to approximately 6B beyond the edge of strip footing in each side while the
corresponding value in the unreinforced soil is found to be 2.5B. It can be concluded
Downloaded by [La Trobe University] at 03:30 02 August 2016

that the geocell mattress acts as a secondary foundation that distribute stresses from the
footing to the bottom of geocell cushion with a dispersion angle (Dash, Rajagopal, &
Krishnaswamy, 2007). Therefore, it transmits the footing load over a wider area giving
rise to better performance. Observation of failure surfaces and heave zones after the
tests also confirm the sufficiency of the soil tank geometry (Figure 9).
According to Figure 8(a), (b) and (c), different geocells have shown distinctive
pressure-settlement responses in spite of their geometrical similarities. The difference in
the behaviour is mostly attributed to the diversity of geocell walls’ stiffness. The W-GT
geocell has the stiffest pocket walls. High stiffness of this type of reinforcement leads to
better confinement of the infill soil, which restrains the lateral movements of soil parti-
cles inside the geocell. At low levels of settlement, the pressure induced by footing
loads tends to make the infill sand rearrange in a denser packing and the frictional resis-
tance of the soil-reinforcement interface restrains the movements of soil particles thanks
to the rough texture of the geocell walls. However, the confining contribution of the
geocell is not fully mobilised prior to densification of the infill soil. Therefore, the
improvement in the performance of geocell-reinforced soil should not be anticipated to
be so significant at low levels of deformation.
At higher ranges of settlement, the geocell performance would be enhanced as cate-
nary shape deformation occurs and the membrane effect of the geocell reinforcements
develops a tensile force in the reinforcement. The vertical component of this force
resists the downward movement of the footing and increases the bearing pressure. By
increasing the level of footing pressure and settlement, the sand in the geocell pockets
beside the footing width starts to move upward as it overcomes the frictional resistance
and the excessive bending of the reinforced sand cushion causes high levels of horizon-
tal tensile strains in the bottom axis of geocell walls located beneath the centre of foot-
ing width. At this stage, the points below the middle axis of geocell walls under the
footing centre undergo large strains up to the ultimate strain and rupture happens in the
geocell. Consequently, a sudden shear failure takes place leading to a large heave in
the soil surface beside the footing width and the infill soil of the geocell moves out of
the pockets.
The B-GG geocell has a modulus less than W-GT and more than NW-GT geocell.
According to Figure 8, at low levels of settlement the improvement in the performance
of B-GG is lower than the W-GT and higher than NW-GT as a result of the stiffness. It
should be mentioned that since the walls of B-GG geocell are not solid, the interlocking
of soil through the apertures of the geogrid make a noticeable influence on the response
12 M. Kargar and S.M. Mir Mohammad Hosseini

of the footing (Dash, 2012). At high levels of settlements, the general behaviour is
similar to W-GT geocell. The comparatively inferior performance of B-GG geocell is
attributed to the less modulus of the reinforcing material. When the footing settlement
rises, the flexural deflexion of the geocell mattress leads to strain mobilisation in the
geocell wall under the footing centre. However, unlike W-GT the B-GG geocell is not
able to survive at high levels of strain (failure strain = 5.9%). Therefore, it starts to rup-
ture in the lowest part of the geogrid wall under the footing centre which continuously
extends to upper parts. Subsequently, a general shear failure occurs immediately.
The NW-GT geocell is extremely flexible and sufficient settlement of the footing is
essential to give the geocell a catenary shape and to develop tensile stress in geocell
walls to resist the pressure transmitted from the footing. Due to low stiffness of the
NW-GT geocell, the geocell wall is incapable of providing sufficient confinement for
the infill soil and the soil undergoes large lateral expansions. Therefore, the soil will fail
prior to reaching the strain in the geocell wall to its ultimate failure strain. So in this
Downloaded by [La Trobe University] at 03:30 02 August 2016

case, no distinctive damage is observed in the geocell material and only local buckling
in the cell wall occurs as a result of footing penetration and development of compres-
sive vertical strain in geocell wall. For that reason, NW-GT geocell with low modulus
is not a suitable material in the improvement of bearing capacity.
According to the results, the stiffer geocells provide more confining stress to the
infill soil and develop higher apparent cohesion so the shear strength of the geocell-soil
composite increases that subsequently leads to increase in bearing capacity (Madhavi
Latha et al., 2009). The additional confinement due to geocell provision also increases
the modulus of granular soil because of the stress dependency behaviour of granular
soil. Therefore, the modulus of the geocell material can be considered the most influen-
tial parameter of geocell mechanical characteristics in the pressure-settlement behaviour
of geocell-reinforced foundation.
The results also show that the ultimate tensile strength of the reinforcing material is
generally not an important factor in the load-settlement behaviour of footing at practical
levels of settlement. Although W-GT and NW-GT materials have almost similar ultimate
strength, the load-settlement responses are completely different (see Figure 8). On the
other hand, the B-GG geocell (with the ultimate strength 9 times smaller than NW-GT)
shows a considerably better improvement in the bearing capacity and for example in the
case of h/B = 1, the ultimate bearing capacity of soil reinforced with B-GG is 30%
higher than NW-GT.
According to the type of failure that occurs for geocell-reinforced sand, it can be
concluded that for the cases where shear in soil takes place at failure, not the reinforce-
ment rupture (NW-GT at all heights and W-GT and B-GG at h/B = .5), the ultimate
strength of reinforcement does not play any role in the response of footing. However,
for the cases where rupture of reinforcement in bending controls the failure mechanism
(B-GG and W-GT at h/B = 1, 1.5), the ultimate tensile strength of the reinforcement can
have a slight influence on the ultimate bearing capacity of the reinforced soil founda-
tion. It should be noted that the geocell strength is only influential at failure level and at
very large levels of settlements that are usually out of serviceability of the footing in
practice. Therefore, in engineering point of view, the effect of geocell ultimate strength
is not as significant as geocell stiffness.
It is also noteworthy that besides the stiffness and strength of the geogrid type rein-
forcement, aperture opening size and orientation of the ribs of geocell material influence
the performance of the geocell-reinforced sand foundations. Application of geocell with
smaller aperture opening size and square/rectangular aperture opening with ribs oriented
European Journal of Environmental and Civil Engineering 13

parallel and perpendicular to the footing result in better performance of the geocells
made from geogrid (Dash, 2012).

5.2. Improvement factors


Figure 10 shows the influence of geocell material stiffness on the improvement in ultimate
bearing capacity and ultimate settlement. Figure 10(a) indicates that the ultimate bearing
capacity of geocell-reinforced sand depends highly on geocell material modulus. Accord-
ing to Figure 10(a), ultimate bearing capacity of geocell-reinforced sand with for example
h/B = 1 is increased 5.1, 3.4 and 2.6 times of the unreinforced sand for W-GT, B-GG and
NW-GT geocells, respectively. It shows an approximately linear relationship between
BCRu and geocell stiffness which is assessed by tensile modulus of reinforcement at 2%
strain (Jε = 2%). It should also be noted that soil confinement may have an enormous
impact on the modulus of geosynthetics especially for nonwoven geotextile material
Downloaded by [La Trobe University] at 03:30 02 August 2016

(Walters, Allen, & Bathurst, 2002). Nevertheless, the modulus of geocell wall material
determined from the in-isolation load-extension tests is considered in these charts.
It is also evident from Figure 10(a) that geocells with larger heights show consider-
ably higher bearing capacity at ultimate failure level due to mobilisation of higher
lateral and vertical confinement as well as the effect of stress distribution (Dash et al.,
2007). For the case of NW-GT, ultimate bearing capacity has increased from 1.63 for
h/B = .5 to 4.17 for h/B = 1.5. Although general failure mechanism for this type of geo-
cell at all heights is shear failure in the underlying soil (due to inadequacy of material
modulus), the frictional resistance at the vertical cell surface is higher for geocell with
larger h/B, so larger footing load is required to overcome this friction leading to move
the sand particles out.
For B-GG and W-GT at h/B = .5 similar failure mechanism is observed as the geocell
mattress bends like a centrally loaded shallow beam with the large deflection resulting in
large shear stress mobilisation. For B-GG and W-GT at h/B = 1 and h/B = 1.5, high rigid-
ity of geocell-sand composite results in behaving like a slab that application of excessive
loads causes failure in bending (Avesani Neto et al., 2013). However, since W-GT geo-
cell has higher modulus, it shows higher bearing capacity compared to B-GG geocell.
Figure 10(b) shows that the ultimate settlement ratio increases significantly by inclu-
sion of geocell reinforcement. It implies that using geocells causes the footing to be
stable over larger levels of settlement without failure and the stable settlement threshold

(a) 8 (b) 4
h/B=0.5
h/B=0.5
7 3.5 h/B=1
h/B=1
6 h/B=1.5
h/B=1.5 3
5
BCRu

SRu

2.5
4
2
3

2 1.5

1 1
0 50 100 150 200 250 300 0 50 100 150 200 250 300
J (kN/m) J (kN/m)

Figure 10. Influence of geocell stiffness on (a) ultimate bearing capacity ratio and (b) ultimate
settlement ratio.
14 M. Kargar and S.M. Mir Mohammad Hosseini

is typically larger for geocells made from stiffer materials. As can be seen from
Figure 10(b), SRu for the case of h/B = 1 is about 2.3 for both NW-GT and B-GG geo-
cells and is 2.75 for W-GT geocell. For the case of h/B = 1.5, however, SRu is 3.6 for
NW-GT, 2.9 for B-GG and 3.3 for W-GT. It is due to the fact that for NW-GT geocell,
high level of settlement is required for shear failure in the underlying soil and for B-GG
rupture happens in the bottom of geocell walls because of inadequate tensile strength.
Figure 11 illustrates the deformed shape of different types of geocells with h/B = 1
after the tests. According to the photos, for geocells with higher modulus (W-GT &
B-GG) rupture occurs at the bottom of geocell layer directly beneath the footing centre.
Similar failure pattern (bending failure) is also observed for these geocells at h/B = 1.5.
On the other hand, for NW-GT geocell with low modulus, despite the flexural deforma-
tion type, no rupture has occurred in geocell wall and failure is as a result of lack of
bearing capacity in the soil under the geocell layer. Similar mechanism takes place for
NW-GT at h/B = .5, 1.5 and W-GT and B-GG at h/B = .5. It is due to the fact that the
Downloaded by [La Trobe University] at 03:30 02 August 2016

stiffness and load-distributing characteristics of the geocell layer is inadequate to prevent


shear failure of the underlying soil.
Figure 12 shows the influence of geocell material in terms of its stiffness on the
improvement in bearing capacity and settlement reduction of footing at different levels
of settlement. The maximum settlement investigated in Figure 12 is considered 14% of
footing width which is approximately the ultimate settlement of the unreinforced sand.
This value is selected because the improvement is assessed with respect to unreinforced
sand and there is no data after this settlement ratio for the unreinforced sand. In addi-
tion, levels of settlement larger than this value are usually not allowed in engineering
practice (Moghaddas Tafreshi & Dawson, 2010).
It is evident from Figure 12 that the efficiency of geocell provision in increasing the
bearing capacity increases with the level of settlement. At lower footing settlement, the
footing load is solely for the densification of the infill soil and hence no axial strain is
mobilised in the geocell walls. As the footing settlement increases, the horizontal pres-
sure distributed to the surrounding perimeter of geocell walls results in mobilisation of
tensile strain in the geocell (Yang, 2010). Therefore, the confinement is provided which
controls the lateral movements of soil particles.
According to Figure 12, for the NW-GT geocell with low stiffness walls, the
improvement in bearing capacity ratio up to settlement levels of s/B < 6% is negligible.
The reason is that soft flexible reinforcement is incapable of generating resistance to
movement of soil particles and reinforcement will deform like soil particles. For the
B-GG and W-GT geocell, great improvement in bearing capacity is shown. At settle-
ment level of s/B = 14%, increase in bearing capacity of W-GT, B-GG and NW-GT geo-
cells (h/B = 1) relative to unreinforced soil are 25, 80 and 120%, respectively. It shows
that modulus of the geocell material makes immense contribution on increasing the stiff-
ness of geocell-reinforced foundation.

(a) (b) (c)

Figure 11. Photo of geocells with h/B = 1 after tests: (a) W-GT, (b) B-GG, (c) NW-GT.
European Journal of Environmental and Civil Engineering 15

(a) 2 s/B=2%
(b) 2.5 s/B=2%
h/B=0.5 h/B=1
s/B=4% s/B=4%
s/B=6% s/B=6%
s/B=8% s/B=8%
2
s/B=10% s/B=10%
BCRs

BCRs
s/B=12% s/B=12%
1.5
s/B=14% s/B=14%

1.5

1 1
0 50 100 150 200 250 300 0 50 100 150 200 250 300
J (kN/m) J (kN/m)

(c) 3
s/B=2%
h/B=1.5
s/B=4%
s/B=6%
2.5
s/B=8%
s/B=10%
BCRs

s/B=12%
2
Downloaded by [La Trobe University] at 03:30 02 August 2016

s/B=14%

1.5

1
0 50 100 150 200 250 300
J (kN/m)

Figure 12. Influence of geocell stiffness on bearing pressure improvement: (a) h/B = .5,
(b) h/B = 1, (c) h/B = 1.5.

(a) 50 (b) 60
W-GT h/B=0.5 W-GT h/B=1
B-GG 50 B-GG
40
NW-GT NW-GT
40
30
PRS (%)

PRS (%)

30
20
20

10
10

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
s/B (%) s/B (%)

(c) 70 W-GT
h/B=1.5
60 B-GG
NW-GT
50

40
PRS (%)

30

20

10

0
0 2 4 6 8 10 12 14
s/B (%)

Figure 13. Influence of geocell stiffness on settlement reduction at different footing settlement
level: (a) h/B = .5, (b) h/B = 1, (c) h/B = 1.5.

The variation of percentage reduction in settlement of footing at different settlement


levels of the unreinforced soil for different geocell materials are also presented in
Figure 13. As can be seen, the percentage reduction in footing settlement is larger for
16 M. Kargar and S.M. Mir Mohammad Hosseini

geocells with higher modulus and height which increases with the footing settlement
level. For example, in the case of h/B = 1 at settlement ratio s/B = 6% there are 38, 21
and 4% reduction in footing settlement for W-GT, B-GG and NW-GT, respectively. It
implies the low efficiency of NW-GT geocell in reduction of footing settlement while
stiffer reinforcements show greater improvement in reduction of footing settlement.

6. Conclusions
In this research, series of laboratory model tests are scheduled to investigate the
improvement effect of geocell-reinforced sand under strip footing static loading. Differ-
ent types of reinforcement material including woven geotextile (W-GT), nonwoven geo-
textile (NW-GT) and biaxial geogrid (B-GG) with various specifications are used to
prepare geocells, and the effects of geocell mechanical properties on the bearing pres-
sure-settlement response of the footing are studied. The improvement in the performance
Downloaded by [La Trobe University] at 03:30 02 August 2016

of geocell-reinforced sand is assessed in terms of increasing the bearing pressure and


percentage reduction in settlement for practical levels of settlements and for the ultimate
failure level. By comparing the behaviour of different geocell-reinforced sand beds
under footing load, the following conclusions can be drawn:

• The ultimate bearing capacity of geocell-reinforced sand increases significantly


with geocell stiffness and height. For example at h/B = 1, ultimate bearing capac-
ity increases 2.6, 3.4 and 5.1 times of the unreinforced sand for NW-GT, B-GG
and W-GT geocells, respectively.
• Stiffness of geocell makes a significant contribution on the performance of
geocell-reinforced soil, although several proposed approaches to determine bearing
capacity of geocell-reinforced soil does not consider the importance of material
modulus.
• Higher improvement in bearing pressure and settlement reduction is obtained by
increase in geocell modulus. This beneficial effect also increases substantially with
footing settlement.
• Failure mechanism for geocell with inadequate modulus and geocells with small
heights is found to be shear failure in the underlying soil while for the stiffer geo-
cells with larger height failure in bending happens with the cells tearing apart at
the bottom.
• As opposed to material modulus, tensile strength of geocell material does not
make a significant contribution on the performance of footings over geocell-
reinforced sand in serviceability levels of footing settlement. However, if rupture
of reinforcement controls the failure mechanism (like the B-GG geocell in this
research) this parameter influences the ultimate bearing capacity of footing.

Acknowledgements
The authors would like to appreciate Professor Richard Bathurst for his valuable comments and
suggestions during the first author’s visiting period in geo-engineering centre at Queen’s-RMC.

Disclosure statement
No potential conflict of interest was reported by the authors.
European Journal of Environmental and Civil Engineering 17

References
Adams, M. T., & Collin, J. G. (1997). Large model spread footing load tests on geosynthetic rein-
forced soil foundations. Journal of Geotechnical and Geoenvironmental Engineering, ASCE,
123, 66–72.
Avesani Neto, J. O., Bueno, B. S., & Futai, M. M. (2013). A bearing capacity calculation method
for soil reinforced with a geocell. Geosynthetics International, 20, 129–142.
Bathurst, R. J., & Cai, Z. (1994). In-isolation cyclic load-extension behavior of two geogrids.
Geosynthetics International, 1(1), 1–19.
Bathurst, R. J., & Jarret, P. M. (1988). Large-scale model tests of geocomposite mattresses over
peat subgrades. Transportation Research Record, 1188, 28–36.
Bathurst, R. J., & Karpurapu, R. (1993). Large-scale triaxial compression testing of geocell-
reinforced granular soils. Geotechnical Testing Journal, ASTM, 16, 296–303.
Biswas, A., Murali Krishna, A., & Dash, S. K. (2013). Influence of subgrade strength on the per-
formance of geocell-reinforced foundation systems. Geosynthetics International, 20, 376–388.
Chen, R.-H., Huang, Y.-W., & Huang, F.-C. (2013). Confinement effect of geocells on sand sam-
ples under triaxial compression. Geotextiles and Geomembranes, 37, 35–44.
Downloaded by [La Trobe University] at 03:30 02 August 2016

Dash, S. K. (2010). Influence of relative density of soil on performance of geocell-reinforced sand


foundations. Journal of Materials in Civil Engineering, ASCE, 22, 533–538.
Dash, S. K. (2012). Effect of geocell type on load carrying mechanism of geocell reinforced sand
foundations. International Journal of Geomechanics, ASCE, 12, 537–548.
Dash, S. K., Krishnaswamy, N. R., & Rajagopal, K. (2001). Bearing capacity of strip footings
supported on geocell-reinforced sand. Geotextiles and Geomembranes, 19, 235–256.
Dash, S. K., Rajagopal, K., & Krishnaswamy, N. R. (2007). Behaviour of geocell-reinforced sand
beds under strip loading. Canadian Geotechnical Journal, 44, 905–916.
Davarci, B., Ornek, M., & Turedi, Y. (2014). Model studies of multi-edge footings on geogrid-
reinforced sand. European Journal of Environmental and Civil Engineering, 18, 190–205.
doi:10.1080/19648189.2013.854726
Fakher, A., & Jones, C. (1996). Discussion on BEARING capacity of rectangular footings on
geogrid-reinforced sand, by Yetimoglu T., Wu, J. T. H., Saglamer, A., 1994. Journal of
Geotechnical Engineering, ASCE, 122, 326–327.
Ferreira, F. B., Vieira, C. S., Lopes, M. L., & Carlos, D. M. (2015). Experimental investigation
on the pullout behaviour of geosynthetics embedded in a granite residual soil. European Jour-
nal of Environmental and Civil Engineering, 1–34. doi:10.1080/19648189.2015.1090927
Gurbuz, A., & Mertol, H. C. (2012). Interaction between assembled 3D honeycomb cells pro-
duced from high density polyethylene and a cohesionless soil. Journal of Reinforced Plastics
& Composites, 31, 828–836.
Indraratna, B., Biabani, M. M., & Nimbalkar, S. (2015). Behavior of geocell-reinforced subballast
subjected to cyclic loading in plane-strain condition. Journal of Geotechnical and Geoenviron-
mental Engineering, ASCE, 141, 04014081-1–04014081-16. doi:10.1061/(ASCE)GT.1943-
5606.0001199
Kargar, M., & Mir Mohammad Hosseini, S. M. (in press). Effect of reinforcement geometry on
the performance of a reduced-scale strip footing model supported on geocell reinforced sand.
Scientia Iranica. Retrieved from http://www.scientiairanica.com/en/ManuscriptDetail?mid=
2183
Khedkar, M. S., & Mandal, J. N. (2009). Behaviour of cellular reinforced sand under triaxial load-
ing conditions. Geotechnical and Geological Engineering, 27, 645–658.
Koerner, R. M. (2005). Designing with geosynthetics. Upper Saddle River, NJ: Prentice Hall.
Kolbsuzewski, J. (1948). General investigation of the fundamental factors controlling loose pack-
ing of sands. Proceedings of the Second International Conference on Soil Mechanics and
Foundation Engineering, Rotterdam, Vol. VII, pp. 47–49.
Leshchinsky, B., & Ling, H. I. (2013). Effects of geocell confinement on strength and deformation
behavior of gravel. Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 139,
340–352.
Madhavi Latha, G., Dash, S. K., & Rajagopal, K. (2009). Numerical simulation of the behavior of
geocell reinforced sand in foundations. International Journal of Geomechanics, ASCE, 9,
143–152.
Madhavi Latha, G., & Somwanshi, A. (2009). Effect of reinforcement form on the bearing capac-
ity of square footings on sand. Geotextiles and Geomembranes, 27, 409–422.
18 M. Kargar and S.M. Mir Mohammad Hosseini

Moghaddas Tafreshi, S. N., & Dawson, A. R. (2010). Comparison of bearing capacity of a strip
footing on sand with geocell and with planar forms of geotextile reinforcement. Geotextiles
and Geomembranes, 28, 72–84.
Moghaddas Tafreshi, S. N., & Dawson, A. R. (2012). A comparison of static and cyclic loading
responses of foundations on geocell-reinforced sand. Geotextiles and Geomembranes, 32,
55–68.
Moghaddas Tafreshi, S. N., Khalaj, O., & Dawson, A. R. (2014). Repeated loading of soil con-
taining granulated rubber and multiple geocell layers. Geotextiles and Geomembranes, 42,
25–38.
Moghaddas Tafreshi, S. N., Shaghaghi, T., Tavakoli Mehrjardi, G. H., Dawson, A. R., & Ghadrdan,
M. (2015). A simplified method for predicting the settlement of circular footings on multi lay-
ered geocell-reinforced non-cohesive soils. Geotextiles and Geomembranes, 43, 332–344.
Nair, A. M., & Madhavi Latha, G. (2015). Large diameter triaxial tests on geosynthetics-rein-
forced granular subbases. Journal of Materials in Civil Engineering, ASCE, 27, 04014148-1–
04014148-8. doi:10.1061/(ASCE)MT.1943-5533.0001088
Pokharel, S. K., Han, J., Leshchinsky, D., Parsons, R. L., & Halahmi, I. (2010). Investigation of
factors influencing behavior of single geocell-reinforced bases under static loading. Geotextiles
Downloaded by [La Trobe University] at 03:30 02 August 2016

and Geomembranes, 28, 570–578.


Rajagopal, K., Krishnaswamy, N. R., & Madhavi Latha, G. (1999). Behaviour of sand confined
with single and multiple geocells. Geotextiles and Geomembranes, 17, 171–184.
Sireesh, S., Sitharam, T. G., & Dash, S. K. (2009). Bearing capacity of circular footing on
geocell-sand mattress overlying clay bed with void. Geotextiles and Geomembranes, 27, 89–98.
Sitharam, T. G., & Sireesh, S. (2005). Behavior of embedded footings supported on geogrid cell
reinforced foundation beds. Geotechnical Testing Journal, 28, 452–463.
Tanyu, B. F., Aydilek, A. H., Lau, A. W., Edil, T. B., & Benson, C. H. (2013). Laboratory evalua-
tion of geocell-reinforced gravel subbase over poor subgrades. Geosynthetics International,
20, 47–61.
Viswanadham, B. V. S., & König, D. (2004). Studies on scaling and instrumentation of a geogrid.
Geotextiles and Geomembranes, 22, 307–328.
Walters, D. L., Allen, T. M., & Bathurst, R. J. (2002). Conversion of geosynthetic strain to load
using reinforcement stiffness. Geosynthetics International, 9, 483–523.
Yang, X. (2010). Numerical analyses of geocell-reinforced granular soils under static and
repeated loads (PhD Thesis), University of Kansas, Lawrence, KS.
Zhang, L., Zhao, M., Shi, C., & Zhao, H. (2010). Bearing capacity of geocell reinforcement in
embankment engineering. Geotextiles and Geomembranes, 28, 475–482.
Zhuang, Y., Wang, K. Y., Liu, H. L., & Chu, J. (2013). The contribution of the subsoil in a rein-
forced piled embankment. European Journal of Environmental and Civil Engineering, 17
(sup1), s269–s281. doi:10.1080/19648189.2013.834599

You might also like