You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/49281201

Fundamental concepts of soil reinforcement - An overview

Article  in  International Journal of Geotechnical Engineering · July 2009


DOI: 10.3328/IJGE.2009.03.03.329-342 · Source: OAI

CITATIONS READS
28 4,603

3 authors, including:

Sanjay Kumar Shukla Nagaratnam Sivakugan


Edith Cowan University James Cook University
190 PUBLICATIONS   1,713 CITATIONS    205 PUBLICATIONS   2,864 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Behaviour of geosynthetic reinforced foundation sand for a strip footing View project

Soil improvement View project

All content following this page was uploaded by Nagaratnam Sivakugan on 28 November 2016.

The user has requested enhancement of the downloaded file.


329
Sanjay K. Shukla *, Nagaratnam Sivakugan , Braja M. Das
1 2 3

Fundamental concepts of soil reinforcement


— an overview

Abstract: In most of the practical applications, reinforced soils are obtained by incorporating continuous reinforcement
inclusions (for example, strip, bar, sheet, mat or net) within a soil mass in a definite pattern. The reinforcing mechanisms
for such reinforced soils have been explained in different forms in the literature since the first modern form of soil reinforce-
ment was developed by Henry Vidal in 1966. In the past three decades, randomly distributed fiber-reinforced soil mechanisms
have also been studied extensively suggesting a new emerging market for them in geotechnical engineering applications. It is
timely to critically review the basic concepts of soil reinforcement, considering different types of reinforcement inclusions. In
this review, reinforced soils are classified into two main groups: systematically reinforced soil, and randomly distributed fiber-
reinforced soil. It is emphasized that a single reinforcing mechanism cannot be used to explain the behavior of all reinforced
soils, in fact, it is highly dependent on the type of reinforcement inclusions; however, the basic concept of soil reinforcement
remains the same for all types of reinforcement.

Keywords: Extensible reinforcement, inextensible reinforcement, fibers, geogrid, geosynthetic, geotextile, metal strips,
reinforced soil, unreinforced soil

1. INTRODUCTION be called ‘systematically reinforced soil’, whereas latter one is


called ‘randomly distributed/oriented fiber-reinforced soil’
Reinforced soil is a composite construction material formed or simply ‘fiber-reinforced soil’. Although the reinforced soil
by combining soil and reinforcement. This material pos- has been in practice in crude form since the ancient times, it
sesses high compressive and tensile strength similar, in prin- is being used more frequently in the civil engineering appli-
ciple, to the reinforced cement concrete. It can be obtained cations since the development of the modern form of soil
by either incorporating continuous reinforcement inclusions reinforcement in 1966 by Henry Vidal, a French architect
(for example, strip, bar, sheet, mat or net) within a soil mass and engineer.
in a definite pattern or mixing discrete fibers randomly with a In most of the current civil engineering applications,
soil fill before placement. The term ‘reinforced soil’ generally the reinforcement generally consists of geosynthetic sheets
refers to the former one, although it may more appropriately or strips of galvanized steel, arranged horizontally or in the
directions in which the soil is subject to the undesirable
*Corresponding Author tensile strains. Compared to the geosynthetic sheets, metal
1
Senior Lecturer in Civil Engineering, School of Engineering, strips are assumed to be relatively inextensible at the stress
Edith Cowan University, Joondalup, Perth, WA 6027, levels experienced in civil engineering applications. In the
Tel: +61 8 6304 2632, Fax: +61 8 6304 5811, early days, the metal strips were used as reinforcement, and
E-mail: s.shukla@ecu.edu.au,
the composite material so obtained was termed ‘Reinforced
Also as Adjunct Associate Professor, School of Engineering and
Physical Sciences, James Cook University, Townsville, QLD 4811, and Earth’ by Henry Vidal (1966, 1969) who first presented the
Reader, Department of Civil Engineering, Banaras Hindu University, concept of improving the strength of a soil mass by inclusion
Varanasi – 221 005, India of reinforcements within it. The soil should preferably be
Associate Professor and Head, Discipline of Civil and Environmental
2 cohesionless, characterized by high frictional properties, in
Engineering, School of Engineering and Physical Sciences, order to prevent the slip between the soil and the reinforce-
James Cook University, Townsville, Queensland 4811, ment. The surface texture of the reinforcement should also
Tel: +61 7 4781 4431, Fax: +61 7 4781 6788,
be as rough as possible for similar reasons.
E-mail: siva.sivakugan@jcu.edu.au
The apparently simple mechanism of reinforced soil and
Professor and Dean Emeritus, California State University, Sacramento,
3
the economy in cost and time have made it an instant suc-
2505 Anthem Village Drive, # E 208, Henderson, Nevada 89052, USA,
Tel.: 702 616 2161, Cell: 702 686 8024, E-mail: brajamdas@gmail.com cess in geotechnical and highway engineering applications
for temporary as well as permanent structures. Reinforcing

International Journal of Geotechnical Engineering (2009) 3: (329-342) J. Ross Publishing, Inc. © 2009
DOI 10.3328/IJGE.2009.03.03.329-342
330  International Journal of Geotechnical Engineering

Figure 2.  Basic reinforcement mechanism resulting in increase in the


stability of the soil mass

ment, inextensible or extensible, has the main task of resist-


ing the applied tensile stresses or preventing inadmissible
(a) deformations in geotechnical structures such as retaining
walls, soil slopes, bridge abutments, foundation rafts, etc. In
this process, the reinforcement acts as a tensile member (see
Fig. 2) coupled to the soil/fill material by friction, adhesion,
interlocking or confinement, and thus improves the stability
of the soil mass.
The concept of reinforcing soil with fibers, especially
natural ones, originated in the ancient times. Applications
of reinforced soils using clayey soils and natural fibers can
be seen even today in the rural areas of India for making
containers, ovens, toys, etc. However, randomly distrib-
uted fiber-reinforced soils have recently attracted increas-
ing attention in geotechnical engineering. In comparison
with systematically reinforced soils, randomly distributed
fiber-reinforced soils exhibit some advantages. Preparation
of randomly distributed fiber-reinforced soils mimics soil
stabilization by admixtures. Discrete fibers are simply added
(b)
and mixed with soil, much like cement, lime, or other addi-
tives. Randomly distributed fibers offer strength isotropy and
Figure 1.  An example of soil reinforcement application as an alterna- limit potential planes of weakness that can develop parallel
tive to a conventional geotechnical structure: (a) conventional concrete/ to the oriented reinforcement as included in systematically
brick masonry/stone masonry wall; (b) reinforced soil retaining wall reinforced soil.
Since 1966, efforts have been made to present the
basic concept of soil reinforcement in different forms in
soil-like materials such as coal ashes and other waste materi- the literature considering both inextensible and extensible
als by continuous inclusions is also an economical means of types of reinforcement (Vidal 1966, 1969; Long et al., 1972;
improving their mechanical properties. One of the common Yang, 1972; Schlosser and Long, 1974; Hausmann, 1976;
applications of soil reinforcement is a reinforced soil retain- Hausmann and Vagneron, 1977; McGown and Andrawes,
ing wall, which is an alternative to a conventional heavy 1977; McGown et al., 1978; Jewell, 1980; Narain, 1985) as well
concrete/brick masonry/stone masonry retaining wall (Fig. as fibers (Waldron 1977; Gray and Ohashi, 1983; Mahar and
1). Reinforcement improves the mechanical properties of a Gray, 1990; Shewbridge and Sitar, 1990; Ranjan et al., 1994a;
soil mass as a result of its inclusion. In fact, any reinforce- Ranjan et al., 1996). Specific aspects of soil reinforcement
Fundamental concepts of soil reinforcement — an overview  331

have also been discussed in some of the books (For example, deformation of the reinforcement but smaller than the lateral
Hausmann, 1990; Koerner, 2005; Das, 1999; Sawicki, 2000; deformation of the soil that might occur in the absence of fric-
Shukla, 2002; Shukla and Yin, 2006). However, in spite of all tion and/or adhesion bonding between both the constituents.
these efforts, no review work exists for understanding, in a This means that in case of perfect friction and/or adhesion
simple manner, the concept of soil reinforcement consider- bonding between reinforcement and soil, the reinforcement
ing both inextensible and extensible types of reinforcements will be extended resulting in a mobilized tensile force T, and
as well as the fibers as described above. An attempt is there- the soil will be compressed by additional compressive lateral
fore made in this paper to present an overview of the funda- stress as reinforcement restraint σR (= Δσ3), introduced into
mental concepts of the soil reinforcement. it in the direction of the reinforcement as shown in Fig. 3(c).
The stress state in soil represented by the Mohr circle ‘b’ in
Fig. 3(d) is no more tangent to the failure envelope lU, and the
2. SYSTEMATICALLY REINFORCED SOIL reinforced specimen is able to sustain greater stresses than
those in the case of unreinforced soil.
Systematically reinforced soil is a soil reinforced with geo- Consider that the reinforced soil specimen shown in
synthetic (woven geotextile/geogrid/geocomposite) sheets Figure 3(b) is expanding horizontally due to decrease in
or strips of galvanized steel in desired directions, and is cur- applied horizontal stress σ3 with constant vertical stress σ1
rently widely used in civil engineering practice. It is mainly and assume that failure occurs by rupture of the reinforce-
because such a reinforced soil possesses many novel charac- ment, that is, the lateral restraint σR is limited to a maximum
teristics, which render it eminently suitable for construction value σRCmax depending on the strength of the reinforcement.
of geotechnical structures. The reinforcement can easily be This state of stress is represented by the Mohr circle ‘c’ in
handled, stored and installed. The soil that constitutes most Figure 3(d). The strength increase can be characterized by
of its bulk may be locally available and can be placed in posi- a constant cohesion intercept cR as an apparent cohesion.
tion in limited time in an economical way by modern hauling Most of these early works were carried out in France, and
and compaction equipment. The flexible nature of reinforced they include research on the material properties of reinforced
soil mass enables it to withstand vibrations caused by earth- earth and its application in retaining walls and abutments.
quakes and large differential settlements without significant These studies indicate that reinforced earth can be consid-
distress. Systematically reinforced soil thus permits construc- ered as a cohesive material with anisotropic cohesion, intro-
tion of geotechnical structures over poor and difficult sub- duced due to reinforcement, being a function of strength and
soil conditions. density of reinforcement (Schlosser and Vidal, 1969). Results
The principle of reinforced soil is analogous to that of of both the triaxial tests and the direct shear tests on sand
reinforced cement concrete; however, their basic reinforc- specimens reinforced with tensile inclusions have also shown
ing mechanisms differ significantly. If the reinforced soil that the apparent cohesion of the reinforced soil material is
is assumed as a homogeneous but anisotropic material, the a function of the orientation of the inclusions with respect to
Mohr-Coulomb failure criterion can be applied to explain the direction of the maximum extension in the soil (Long et
the basic mechanism of reinforced soil. Consider a simplified al., 1972; Schlosser and Long, 1974; Jewell, 1980; Gray and
situation, shown in Figs. 3(a) and (b), where two cylindrical Refeai, 1986). Thus, the strength envelope for reinforced
specimens of a cohesionless soil are subjected to the same cohesionless soil for reinforcement rupture condition can be
triaxial loading. The first specimen is not reinforced, and the interpreted in terms of Mohr-Coulomb failure envelope lRC
second is reinforced with horizontal reinforcement layers. for the homogeneous cohesive soil as shown in Figure 3(d).
Figure 3(c) shows a magnified view of the reinforced soil For Mohr circle ‘a’, the principal stresses σ1 and σ3 are
element PQRS as indicated in Fig. 3(b). Assume that that related to each other as:
the Mohr-Coulomb failure criterion has been attained in the
unreinforced specimen. For this case, the stress state in the σ1 = σ3tan2(45° + ϕ/2) (1)
soil can be represented, in the normal stress (σ) and shear
where ϕ is the angle of shearing resistance (or the friction
stress (τ) space, by a Mohr circle ‘a’ as shown in Fig. 3(d),
angle) of the unreinforced soil.
which is tangent to the Mohr-Coulomb failure envelope lU for
For Mohr circle ‘c’, representing the stress state of a
unreinforced soil. If the reinforced soil specimen is subjected
reinforced soil at failure, the principal stresses σ1 and σ3 are
to the same stress state, then due to friction and/or adhesion
related to each other as:
bonding between both constituents, the lateral deformation/
strain of the specimen will be reduced. This lateral deforma- σ1 = σ3mintan2(45° + ϕ/2) + 2cRtan2(45° + ϕ/2) (2)
tion of the composite material will be greater than the lateral
332  International Journal of Geotechnical Engineering

(a) (b) (c)

(d)

Figure 3.  Basic mechanism of reinforced soil in triaxial loading: (a) unreinforced cylindrical cohesionless soil specimen; (b)
reinforced cylindrical cohesionless soil specimen; (c) magnified view of a reinforced soil element PQRS as indicated in (b); (d)
Mohr circles for reinforced and unreinforced cases [Note: σ1 is the major principal stress and σ3 is the minor principal stress.]

Since σ3min = (σ3 – σRCmax) as seen in Fig. 3(d), Eq. (2) Ka = tan2(45° – ϕ/2) (5)
becomes:
and
σ1 = (σ3 – σRCmax)tan (45° + ϕ/2) + 2cRtan(45° + ϕ/2) (3)
2
KP = tan2(45° + ϕ/2) (6)
Combining Eqs. (1) and (3) leads to: are the Rankine’s coefficients of active and passive lateral
earth pressures respectively. Thus, it is found that the aniso-
 RC max tan( 450   / 2)  RC max K p  RC max tropic cohesion is produced in the direction of reinforce-
cR    (4)
2 2 2 Ka ment, and this concept is based on the behavior of laboratory
tests (shear tests) on reinforced soil samples. It has, however,
where
not been possible to define this cohesion in a way as to enable
its use in the design of reinforced earth structures.
Fundamental concepts of soil reinforcement — an overview  333

Now, consider that the reinforced soil specimen shown


in Figure 3(b) is expanding horizontally due to decrease in
applied horizontal stress σ3 = σ30 with constant vertical stress
σ1 = σ10 as represented by the Mohr circle ‘d’, and assume
that failure occurs by slippage between the reinforcement
and soil, that is, lateral restraint σR is limited to σRF, which is
proportional to σ10, that is,
σRF = σ10F (7)
where F is a friction factor that depends on the cohesionless
(a)
soil – reinforcement interface characteristics. This concept is
based on the Yang’s experimental results (Yang, 1972) as pre-
sented by Hausmann and Vagneron (1977). The failure state
of stress is represented by the Mohr circle ‘e’ in Figure 3(d).
The strength increase can be characterized by an increased
friction angle ϕR. Thus, the strength envelope for reinforced
cohesionless soil for reinforcement slippage condition can
be interpreted in terms of Mohr-Coulomb failure envelope
lRF for the homogeneous cohesionless soil as shown in Figure
3(d).
For Mohr circle ‘d’, the principal stresses σ10 and σ30 are
related by:
σ10 = σ30tan2(45° + ϕ/2) (8)
Substituting Eq. (5) in Eq. (8) yields:
(b)
σ
σ10 = 30 (9)
Ka
For Mohr circle ‘e’, the principal stresses σ10 and σ30 are
related by:
σ10 = σ30mintan2(45° + ϕR/2) (10)
Since σ30min= σ30 – σRF as seen in Fig. 3(d), Eq. (10)
becomes:
σ10 = (σ30 – σRF)tan2(45° + ϕR/2) (11)
Substituting Eq. (7) in Eq. (11) yields:
σ10 = (σ30 – σ10F)tan2(45° + ϕR/2) (12)
Combining Eqs. (8) and (12) leads to: (c)

1 = (Ka – F)
� �
1 + ϕR
1 – ϕR
Figure 4.  Postulated behavior of a unit cell in plane strain conditions
with and without inclusions: (a) unit cell; (b) dense sand with inclusions;
(c) loose sand with inclusions (adapted from McGown et al., 1978)
or
1 + F – Ka (13) by rupture of the reinforcement or reinforcement slippage.
sin ϕR=
1 – F + Ka These failure states of stress are represented by the Mohr
Now, consider that the reinforced soil specimen shown circles ‘f’ or ‘g’ in Fig. 3(d) respectively. It can be noted that
in Figure 3(b) is expanding horizontally due to increase in the reinforcement increases the compressive strength of the
applied σ1 with constant σ3 and assume that failure occurs
334  International Journal of Geotechnical Engineering

exists in that both inhibit the development of internal and


boundary deformations of the soil mass by developing ten-
sile stresses in the reinforcement. In other words, both the
ply soil and the reinforced earth are tensile strain inclusion
systems.
Fluet (1988) subdivided the reinforcement, based on its
function, into the following two categories:
(a) 1. A tensile member, which supports a planar load, as
shown in Figure 5(a).
2. A tensioned member, which supports not only a
planar load but also a normal load, as shown in
Figure 5(b).

Jewell (1996) and Koerner (2005) consider not two but


three mechanisms for soil reinforcement, because when the
geosynthetic works as a tensile member it might be due to
two different mechanisms: shear and anchorage. Therefore,
the three reinforcing mechanisms, concerned simply with
(b)
the types of load that are supported by the geosynthetic, are
the following:
Figure 5.  Reinforcement function: (a) tensile member; (b) tensioned
member (adapted from Fluet, 1988) 1. Shear, also called sliding: The geosynthetic sup-
ports a planar load due to slide of the soil over it.
2. Anchorage, also called pullout: The geosynthetic
soil by Δσ1 or Δσ10 depending on the type of failure mode of supports a planar load due to its pullout from the
the reinforced soil. soil.
A different concept of the influence of reinforcement 3. Membrane: The geosynthetic supports both a pla-
on the behavior of reinforced soil mass was described by nar and a normal load when placed on a deform-
Basset and Last (1978). It is suggested that introduction of able soil.
reinforcement modifies the dilatancy characteristics of soil
with possible rotation of principal strain directions. This Shukla (2002, 2004) and Shukla and Yin (2006) describe
concept is based on the fact that if the dilation of the soil is reinforcing mechanisms that take into account the reinforce-
restricted, the shear strength mobilized will be higher than ment action of the geosynthetic, in other words, how the
for the case of no restriction. The presence of reinforcement geosynthetic reinforcement takes the stresses from the soil
in soil imposes a condition of restricted dilatancy. It also pre- and which type of stresses are taken by it. This concept can
determines the principal incremental strain directions and be observed broadly in terms of the following roles of geo-
rotates them relative to the unreinforced case, resulting in a synthetics:
redistribution of stresses. 1. A geosynthetic layer reduces the outward hori-
The behavior of the soil reinforced with extensible zontal stresses (shear stresses) transmitted from
reinforcements, such as geosynthetics, does not fall entirely the overlying soil/fill to the top of the underlying
within the concepts as described above. The difference, foundation soil. This action of geosynthetics is
between the influences of inextensible and extensible rein- known as shear stress reduction effect. This effect
forcements, is significant in terms of the load-settlement results in a general-shear, rather than a local-shear
behavior of the reinforced soil system as shown in Fig. 4 failure (Figure 6(a)), thereby causing an increase
(McGown et al., 1978). The soil reinforced with extensible in the load-bearing capacity of the foundation soil
reinforcement, termed ply-soil by McGown and Andrawes (Bourdeau et al., 1982; Guido et al., 1985; Love et
(1977), has greater extensibility and smaller losses of post al., 1987; Espinoza, 1994; Espinoza and Bray, 1995;
peak strength compared to soil alone or soil reinforced with Adams and Collin, 1997). Through the shear inter-
inextensible reinforcement, termed reinforced earth by Vidal action mechanism the geosynthetic can therefore
(1966, 1969). In spite of some differences in the behavior improve the performance of the system with very
of ply soil and reinforced earth, a similarity between them little or no rutting. In fact, the change in the failure
335  International Journal of Geotechnical Engineering

mode as a result of reduction in shear stress is the


primary benefit of the geosynthetic layer at small
deformations.
2. A geosynthetic layer redistributes the applied sur-
face load by providing restraint of the granular
fill if embedded in it, or by providing restraint
of the granular fill and the soft foundation soil, if
placed at their interface, resulting in reduction of
applied normal stress on the underlying founda-
tion soil (Fig. 6(b)). This is referred to as slab effect
(a) or confinement effect of geosynthetics (Bourdeau
et al., 1982; Giroud et al., 1984; Madhav and
Poorooshasb, 1989; Sellmeijer, 1990; Hausmann,
1990). The friction mobilized between the soil and
the geosynthetic layer plays an important role in
confining the soil.
3. The deformed geosynthetic, sustaining normal and
shear stresses, has a membrane force with a verti-
cal component that resists applied loads, i.e. the
deformed geosynthetic provides a vertical support
to the overlying soil mass subject to loading. This
(b)
action of geosynthetics is popularly known as its
membrane effect (Figure 6(c)) (Giroud and Noiray,
1981; Bourdeau et al., 1982; Sellmeijer et al.,
1982; Love et al., 1987; Madhav and Poorooshasb,
1988; Bourdeau, 1989; Sellmeijer, 1990; Shukla and
Chandra 1994, 1995). Depending upon the type
of stresses — normal stress and shear stress, sus-
tained by the geosynthetic during their action, the
membrane support may be classified as ‘normal
stress membrane support’, and ‘interfacial shear
stress membrane support’ respectively (Espinoza
(c)
and Bray, 1995). Edges of the geosynthetic layer
are required to be anchored in order to develop
the membrane support contribution resulting
from normal stresses, whereas membrane support
contribution resulting from mobilized interfacial
membrane shear stresses does not require any
anchorage. The membrane effect of geosynthetics
(d) causes an increase in the load-bearing capacity of
the foundation soil below the loaded area with a
downward loading on its surface to either side of
the loaded area, thus reducing its heave potential.
Figure 6.  Roles of a geosynthetic reinforcement: (a) causing change It is to be noted that both the woven geotextile and
of failure mode (shear stress reduction effect); (b) redistribution of the the geogrid can be effective in membrane action in
applied surface load (confinement effect); (c) providing vertical support
case of high-deformation systems.
(membrane effect) (adapted from Bourdeau et al., 1982 & Espinoza,
1994); (d) providing passive resistance through interlocking of the soil
4. The use of geogrids has another benefit owing to
particles (interlocking effect) the interlocking of the soil through the apertures
(openings between the longitudinal and transverse
ribs, generally greater than 6.35 mm of the grid
known as interlocking effect (Guido et al., 1986)
336  International Journal of Geotechnical Engineering

(Fig. 6(d)). The transfer of stress from the soil


to the geogrid reinforcement is made through
bearing (passive resistance) at the soil to the grid
cross-bar interface. It is important to underline
that because of the small surface area and large
apertures of geogrids, the interaction is due mainly
to interlocking rather than to friction. However, an
exception occurs when the soil particles are small.
In this situation the interlocking effect is negligible
because no passive strength is developed against
the geogrid (Pinto, 2004).

3. RANDOMLY DISTRIBUTED FIBRE-


REINFORCED SOIL
Concept of the randomly distributed fiber-reinforced soil
has been reported in the literature in the past few decades.
A large number of experimental studies have been carried
out to observe the characteristics of fiber-reinforced soils
(Hoare, 1979; Gray and Ohashi, 1983; Freitag, 1986; Gray
and Al-Refeai, 1986; Lindh and Eriksson, 1990; Maher and
Gray, 1989; Shewbridge and Sitar, 1990; Al-Refeai, 1991;
Ranjan et al., 1994b; Bauer and Oancea, 1996; Michalowski
and Zhao, 1996; Consoli et al., 1998; Kumar et al., 1999;
Figure 7.  Model for flexible, elastic reinforcement extending across the
Santoni et al., 2001; Tingle et al., 2002; Consoli et al., 2002; shear zone of thickness z (adapted from Gray and Ohashi, 1983)
Michalowski and Cermak 2003; Yetimoglu and Salbas, 2003;
Gosavi et al., 2004; Heineck et al., 2005; Gupta et al., 2006; of maximum principal tensile strain in a direct shear test;
Kumar and Singh, 2008). Most of the basic characteristics of and (3) Shear strength increases as a result of fiber reinforce-
fiber-reinforced soils can be found in the experimental works ment are approximately the same for loose and dense sand.
described below. However, larger strains were required to reach the peak shear
Gray and Ohashi (1983) conducted direct shear tests on resistance in the loose case. Shear strength envelopes for fiber
a dry sand reinforced with different types of fibers in speci- reinforced sand clearly showed the existence of a threshold
fied orientation. Both natural and synthetic fibers including confining stress below which the fibers tended to slip or pull
metal fibers were tested for purposes of comparison. Test out. The strength envelop also indicates that the fibers do not
results show that relatively low modulus, fiber reinforce- affect the angle of internal friction of the sand. Increasing the
ments behave as “ideally extensible” inclusions. They do length of fiber reinforcements increased the shear strength of
not rupture during shear. Their main role is to increase the fiber-reinforced composite but only up to a point. A lim-
peak shear strength and to limit the magnitude of post peak iting or asymptotic shear strength was soon reached beyond
reduction in shear resistance in dense sand. The experimen- which any further increase in fiber length had no effect. The
tal behavior was compared with the theoretical predictions same or higher shear strength increases could be achieved
by presenting a force equilibrium model of fiber reinforced with a modest increase in the number of fibers or area ratio
sand as shown in Fig. 7, which has been explained later in of relatively low modulus reed fibers compared to stiffer and
detail in this paper. The developed model predicts correctly stronger copper wires with the same initial area ratio.
the influence of various soil-fiber parameters, namely: (1) Gray and Al-Refeai (1986) conducted triaxial compres-
Shear strength increases are directly proportional to fiber sion tests on a dry sand reinforced with randomly distrib-
area ratios, at least up to 1.7%, which is a probable upper uted, discrete fibers and oriented, continuous fabric layers.
limit for fiber concentrations across potential shear planes The sand was a clean, uniform, medium-grained sand with
in a root permeated soil; (2) Shear strength increases are the following properties: specific gravity of solids, Gs = 2.65;
greatest for initial fiber orientations of 60° with respect to the effective grain size, D10 = 0.28 mm; median grain diameter,
shear surface. This orientation coincides with the direction D50 = 0.41 mm; coefficient of uniformity, Cu = 1.50; minimum
Fundamental concepts of soil reinforcement — an overview  337

void ratio, emin = 0.50; maximum void ratio, emax = 0.78; and Second, an optimum fiber length of 51 mm was identified for
angle of shearing resistance (triaxial tests), ϕ = 39° and 32° at the reinforcement of sand specimens. Third, a maximum per-
relative density, Dr = 86 and 21% respectively. Both natural formance was achieved at a fiber dosage rate between 0.6 and
and synthetic glass fibers varying from 13 to 38 mm in length 1% dry weight. Fourth, specimen performance was enhanced
and from 0.3 to 1.75 mm in diameter were used. The results in both wet and dry of optimum conditions. Finally, the
showed that both types of reinforcement systems increased inclusion of up to 8% of silt does not affect the performance
strength and modified the stress-deformation behavior of the of the fiber reinforcement. These data do not appear reliable
sand. It was observed that continuous, oriented fabric inclu- to authors because how can one think to conduct unconfined
sions markedly increased the ultimate strength, increased compression tests, especially on dry sands.
the axial strain at failure, and in most cases limited reduc- Consoli et al. (2002) described an experimental study of
tions in post-peak loss of strength. The discrete, randomly the utilization of polyethylene terephthalate (PET) fibers in
distributed fibers increased both the ultimate strength and the reinforcement of uncemented and artificially cemented
the stiffness of reinforced sand. The decrease in stiffness sand. The sand was classified as non-plastic uniform fine
at low strains (less than 1%), observed with oriented, con- sand (SP) according to the Unified Soil Classification System,
tinuous fabric inclusions, did not occur with the fibers. The and was having the following properties: specific gravity of
increase in strength with fiber content varied linearly up to a solids, Gs = 2.62; effective grain size, D10 = 0.16 mm; coef-
fiber content of 2% by weight, and thereafter approached an ficient of uniformity, Cu = 1.9; coefficient of curvature, Cc =
asymptotic upper limit. The rate of increase was roughly pro- 1.2; minimum void ratio = 0.57; maximum void ratio = 0.85.
portional to the fiber aspect ratio (length to diameter ratio). PET fibers of cylindrical shape with different lengths were
At the same aspect ratio, confining pressure, and weight frac- obtained buy chopping long filaments that produces recycled
tion and roughness (not stiffness) of fibers tended to be more PET plastic waste (specific gravity = 1.06; tensile strength =
effective in increasing strength. 207 – 230 MN/m2; elastic modulus = 7 GN/m2, linear strain
Ranjan et al. (1994b) conducted a series of triaxial com- at failure = 20 – 30%). The cement was rapid-hardening
pression tests to study the stress-strain behavior of plastic- Portland cement. The stress-strain-strength response, under
fiber-reinforced sand and increase in shear strength of sand axisymmetric loading, was evaluated by a comprehensive
due to fiber inclusions. The sand used was poorly graded fine experimental program, which included unconfined com-
sand (SP-SM) as per the Unified Soil Classification system. pression tests, splitting tensile tests, and drained triaxial
Plastic fibers were cut to length from continuous fibers. Their compression tests with local strain measurement. For the
study shows that the inclusion of fibers causes an increase in uncemented sand, the inclusion of PET fiber increased the
peak shear strength and reduction in the loss of post-peak peak strength, ultimate strength, and energy absorption
strength. Thus, the residual strength of fiber-reinforced sand capacity. The cohesive intercept did not change whereas the
is higher as compared to that of unreinforced sand. The prin- friction angle increased with the inclusion of fiber and fiber
cipal stress envelopes for fiber-reinforced sand are bilinear length. Reinforcement did not affect the initial stiffness or
having a break at a confining stress, called critical confining ductility of the uncemented sand. As expected, the addition
stress, below which the fibers tend to slip or pull out. An of cement to the sand significantly increased the stiffness and
increase in fiber aspect ratio results in a lower critical confin- peak strength, and changed the soil behavior from ductile
ing stress and higher contribution to shear strength. Shear to a noticeable brittle one. The cement content increased
strength increases approximately linearly with increasing both the peak friction angle and the cohesive intercept. For
fiber content up to 2% by weight, beyond which the gain in an intermediate cement content of 5%, an increase in fiber
strength is not appreciable. content from 0.1 to 0.9% caused average increases in uncon-
Santoni et al. (2001) conducted laboratory unconfined fined compressive strength and in tensile strength of about
compression tests on sand specimens reinforced with ran- 40 and 78% respectively. The initial stiffness of the cemented
domly oriented discrete fibers to isolate the effect of each sand was not affected by fiber inclusion, since it is basically
variable on the performance of the fiber-reinforced material. a function of cementation. The efficiency of the fiber rein-
The test involved six different sands ranging from a fine sand forcement when applied to cemented sand was found to be
to a coarse sand. Four different types of fibers as synthetic dependent on the fiber length. The greatest improvements
monofilament, fibrillated, tape and mesh fibers used in this in triaxial strength, ductility, and energy absorption capacity
investigation were made of polypropylene. Five primary were observed for the longer, 36 mm fiber. The positive effect
conclusions were obtained from this investigation. First, the of fiber length was not detected by either the unconfined
inclusion of randomly oriented discrete fibers significantly compression tests or the split cylinder tests, clearly indicating
improved the unconfined compressive strength of sands. the major influence of confining pressure and the necessity
338  International Journal of Geotechnical Engineering

of carrying out triaxial tests to fully show fiber reinforced soil as a composite material in which roots of relatively high
behavior. tensile strength were embedded in a matrix of lower tensile
Yetimoglu and Salbas (2003) studied the shear strength strength. The model consists of vertical flexible, elastic roots
of a clean, oven dried, uniform quartz river sand (specific of uniform diameter extending an equal distance on either
gravity = 2.64; maximum dry unit weight = 17.48 kN/m3; side of horizontal sheer plane. This model considers only
minimum dry unit weight = 14.92 kN/m3; effective grain partial mobilization of fiber tensile strength depending upon
size, D10 = 0.2 mm; coefficient of uniformity, Cu = 1.65; coef- the amount of fiber elongation during shear. The model does
ficient of curvature, Cc = 1.02; angle of shearing resistance at not place any constraint on the distribution or location of the
relative density of 70% = 42°) reinforced with randomly dis- reinforcing fibers.
tributed polypropylene fibers (diameter = 0.05 mm, length The concept of the Waldron’s model was extended by
= 20 mm; specific surface = 8.5 m2/N, density = 9.1 kN/ Gray and Ohashi (1983) to describe the deformation and fail-
m3, tensile strength = 320 – 400 MPa, modulus of elasticity ure mechanism of systematically distributed fiber-reinforced
= 3500 – 3900 MPa) by carrying out direct shear tests. The soil and to estimate the contribution of fiber reinforcement
fiber content was varied from 0.10 to 1.0% of the weight of in increasing the shear strength of soil. The model consists of
the dry sand. The results of the tests indicate that the peak a long, elastic fiber extending an equal length over either side
shear strength and initial stiffness of the sand are not affected of a potential shear plane in sand (Fig. 7). The fiber may be
significantly by the fiber reinforcements. Hence, the values oriented initially perpendicular to the shear plane or at some
of peak shear strength angle can be considered identical for arbitrary angle, i. Shearing causes the fiber to distort, thereby
reinforced and unreinforced sand. Thus, Mohr-Coulomb mobilizing tensile resistance in the fiber. The tensile force in
shear envelopes for fiber-reinforced sands are similar to that the fiber can be divided into components normal and tan-
for unreinforced sand, and are linear with a zero cohesion gential to the shear plane. The normal component increases
intercept. The horizontal displacements at failure are also the confining stress on the failure plane thereby mobilizing
found comparable for reinforced and unreinforced sands additional shear resistance in the sand, whereas the tangen-
under the same vertical normal stress. It was explained that tial component directly resists shear. The fiber is assumed to
the horizontal displacement, that is, strain level at failure be thin enough that it offers little if any resistance to shear
seemed to be not high enough to mobilize operative tensile displacement from bending stiffness. If many fibers are pres-
stresses in the reinforcement inclusions. Most likely, due to ent, their cross-sectional areas are computed, and the total
this limited strain compatibility, fiber contributions to peak fiber concentration is expressed in terms of a fiber area ratio
strength and strain at failure were insignificant in the tests. (AR/A), where AR is the total cross-sectional area of fibers in
Fiber reinforcements, however, can reduce soil brittleness the shear plane, and A is the total area of the shear plane. The
providing smaller loss of post-peak strength. Thus, there shear strength increase ΔSR from the fiber reinforcement in
appears to be an increase in residual shear strength angle of sand thus can be estimated from the following expressions:
the sand by adding fiber reinforcements.
ΔSR = tR(sinθ + cosθtanϕ) (14)
A large number of experimental studies as mentioned
above have been reported in literature to observe the char- for fibers oriented initially perpendicular to shear plane, and
acteristics of fiber-reinforced soils, and it has now been
established that the strength and deformation behavior of the ΔSR = tR[(sin(90° – ψ) + cos(90° – ψ)tanϕ)] (15a)
fiber-reinforced soils is governed by the soil characteristics for fibers oriented initially at some arbitrary angle i, where
(e.g., gradation and particle size and shape) and the fiber
properties (weight ratio, aspect ratio, and modulus). In spite 1
ψ = tan–1 −1 (15b)
of this fact, a limited number of basic models (Waldron 1977; (
k + tan −1 i )
Gray and Ohashi, 1983; Mahar and Gray, 1990; Shewbridge
and tR is the mobilized tensile strength of fibers per unit area
and Sitar, 1990; Ranjan et al., 1996) have been suggested to
of soil; ϕ is the angle of internal friction of the soil; θ is the
explain the mechanism of fiber reinforcements in soils. It
angle of shear distortion; x is the horizontal shear displace-
is probably because the modeling of the states of stress and
ment; z is the thickness of shear zone; and k (= x/z) is shear
strain in fiber-reinforced soil during deformation and failure
distortion ratio.
is complex and difficult.
The mobilized tensile strength per unit area of soil (tR)
Waldron (1977) proposed a force-equilibrium model to
can be estimated as:
estimate the increase in strength of soil reinforced with plant

� �
roots, taking into account the tensile force developed in the AR
tR = σR (16)
reinforcement. The root-reinforced soil mass was considered A
Fundamental concepts of soil reinforcement — an overview  339

where σR is the tensile stress developed in the fiber at the shear


plane, which depends upon a number of parameters and test
variables. The fibers must be long enough and frictional
enough to avoid pullout; conversely the confining stress
must be high enough so that pullout forces do not exceed
skin friction (shear forces) along the fiber. It is also necessary
to assume some sort of tensile stress distribution along the
length of the fiber. Two likely or reasonable possibilities are
a linear or parabolic distribution, respectively, with tensile
stress a maximum at the shear plane and decreasing to zero
at the fiber ends. The resulting tensile stress at the shear plane
for these two distributions is given (Waldron, 1977) by the
following expressions:

� �
4ERτR 1/2
σR = {z(secθ – 1)}1/2 (17) Figure 8.  Effect of the fiber inclusion in sand on its principal stress
DR envelope obtained from triaxial compression tests (adapted from Maher
and Gray, 1990) [Note: Below critical confining stress, the stress enve-
for linear distribution, and lopes are either linear or nonlinear depending on the types of sand and
reinforcement.]

� �
1/2
8ERτR
σR = {z(secθ – 1)}1/2 (18)
3DR
soils, Mahar and Gray (1990) have reported that the failure
for parabolic distribution, where ER is the modulus or longi- surfaces in triaxial compression tests are planer and oriented
tudinal stiffness of the fiber; τR is the skin friction stress along in the same manner as predicted by the Mohr-Coulomb
the fiber; DR is the diameter of fiber; and z is the thickness of theory. This finding suggests an isotropic reinforcing action
the shear zone. with no development of preferred planes of weakness or
Direct shear tests conducted by Shewbridge and Sitar strength. The principal stress envelops have been found to be
(1989) on the dry sand reinforced with oriented fibers show either curved-linear or bilinear, with the transition or break
that the thickness z and shape of a shear zone is altered by the occurring at a confining stress, called the critical confining
reinforcement concentration, stiffness, and reinforcement- stress, σ3crit (Fig. 8). The increase in shear strength from the
soil bond strength. The actual deformation pattern of the fiber reinforcement can be estimated form the following
reinforcement-soil composite can be described by a smooth expressions:
asymptotic curve of the form as:
x1 = B(1 – e–b|x2|) (19) ΔSR = Ns
� �
πD2R
4
(2σ3av tanδ)(sinθ + cosθ tanϕ)k (20a)

where B is one-half of the shear displacement (= x/2); x1 and


for 0 <σ3av <σ3crit, and
x2 are the coordinate directions parallel and perpendicular
to the direction of shear respectively; and b is a curve fitting
parameter describing the shape of the curve. This defor-
mation pattern differs substantially from the simple shear
ΔSR = Ns
� �
πD2R
4
(2σ3crit tanδ)(sinθ + cosθ tanϕ)k (20b)

deformation assumed in Fig. 7. The shear zone thickness z is for σ3av >σ3crit, where Ns is the average number of fibers
greater in reinforced soil than in soil alone, and it increases intersecting a unit area; σ3av is the average confining stress
with increasing stiffness of the reinforced soil due to increased in triaxial chamber; δ is the fiber skin friction angle, and k
reinforcement concentration, stiffness, or bond strength. is an empirical coefficient depending on soil characteristics
Though the reinforcement concentration and type affect the (median grain size, D50, particle sphericity, S, and coefficient
strength of the reinforced soil, but a non-linear relationship of uniformity, CU) and fiber parameters (aspect ratio and skin
between reinforcement concentration and increased strength friction). The value of σ3crit in Eqs. 20(a & b) can be deter-
was observed by Shewbridge and Sitar (1989). This observa- mined empirically from the experimental measurements,
tion contradicts the findings reported by Gray and Ohashi thus depending on soil characteristics and fiber properties.
(1983), Mahar and Gray (1990), and Ranjan et al. (1996). The theoretical model proposed by Maher and Gray
Based on the observations made in triaxial tests and (1990) have predicted reasonably well the increase in the
statistical analysis of randomly distributed fiber-reinforced strength of randomly distributed fiber reinforced soil.
However, the width of shear zone, z, which significantly
340  International Journal of Geotechnical Engineering

affects the increase in strength (Shewbridge and Sitar, 1989, of major influence of confining pressure, the necessity of
1990) has not been determined for reinforced soil. Also, the carrying out more triaxial tests, especially on large-size speci-
average expected orientation of fibers is statistically predicted mens, is felt to fully show the fiber reinforced soil behavior.
to be perpendicular to the plane of shear failure. But it is dif-
ficult to determine experimentally the orientation of fibers.
Keeping these limitations in view, Ranjan et al. (1996) pre- REFERENCES
sented a regression analysis of a large number of triaxial tests
on randomly distributed fiber reinforced soil, and reported Al-Refeai, T.O. (1991). “Behaviour of granular soils rein-
that their analysis is accurate in predicting the increase in forced with discrete randomly oriented inclusions.”
shear strength due to inclusions of fibers in cohesionless Geotextiles and Geomembranes, 10, 319-333.
soils. Adams, M.T. and Collin, J.G. (1997). “Large model spread
footing load tests on geosynthetic reinforced soil foun-
dations.” Journal of Geotechnical and Geoenvironmental
CONCLUSIONS Engineering, 123(1), 66-72.
Bassett, R.H. and Last, N.C. (1978). “Reinforcing earth
The concept of reinforcing soils by introducing tension- below footings and embankments.” Proceedings of the
resisting elements such as strips, bars, sheets, meshes, fibers, Symposium on Earth Reinforcement, ASCE, Pittsburgh,
etc. has become a common practice in geotechnical engineer- 202-231
ing lately. The flexible nature of reinforced soil mass enables Bauer, G.E., Oancea, A. (1996). “Triaxial testing of granular
it to withstand a large differential settlement without any soils reinforced with discrete polypropylene fibers.”
major distress. In certain difficult situations, the concept of Proc., First European Geosynthetics Conference on
reinforced soil is the only valid technical solution. Although Geosynthetics: Applications, Design and Construction,
the soil reinforcing mechanism differs for different types of A.A. Balkema, Rotterdam, The Netherlands, 407-410.
reinforcements, the soil-reinforcement friction/adhesion is Bourdeau, P.L. (1989). “Modeling of membrane action in
fundamental to the concept of all types of the reinforced soil a two-layer reinforced soil system.” Computers and
described in this paper. The inclusion of reinforcement in a Geotechnics, 7, 19-36.
soil mass inhibits the tensile strains and develops the strength Bourdeau, P.L., Harr, M.E., and Holtz, R.D. (1982). “Soil-
of the soil. The strength increase of the soil due to presence fabric interaction — an analytical model.” Proc., 2nd
of reinforcement is reflected either by an apparent cohesion International Conference on Geotextiles. Las Vegas,
intercept cR (failure due to rupture of reinforcement) or by U.S.A., 387-391.
increased friction angle ϕR (failure due to slippage). The Consoli, N.C. Prietto, P.D.M., and Ulbrich, L.A. (1998).
increase in strength is a function of shear strength of soil, and “Influence of fiber and cement addition on behavior of
the tensile strength and distribution of reinforcement in the sandy soil.” Journal of Geotechnical and Geoenvironmental
soil mass. If geosynthetics are used as reinforcement in a soil Engineering, ASCE, 124(12), 1211-1214.
mass, they improve the strength and settlement characteristic Consoli, N.C., Montardo, J.P., Prietto, P.D.M. and Pasa,
of the soil mass by showing the following effects: shear stress G.S. (2002). “Engineering behavior of a sand rein-
reduction effect, confinement effect, membrane effect, and forced with plastic waste.” Journal of Geotechnical and
interlocking effect. The previous studies indicate that the Geoenvironmental Engineering, ASCE, 128(6), pp. 462-
stress-strain properties of randomly distributed fiber-rein- 472.
forced soils are functions of fiber content, aspect ratio and Das, B.M. (1999). Principles of Foundation Engineering,
skin friction along with the soil and fiber index and strength Fourth Edition, PWS Publishing Company, Boston.
characteristics. For field applications of fiber-reinforced soils, Espinoza, R.D. (1994). “Soil-geotextile interaction: evaluation
much care is required to obtain reasonably uniform distribu- of membrane support.” Geotextiles and Geomembranes,
tion of the fibers within the soil mass. Since the influences of 13, 281-293.
engineering properties of soil and fiber, and the scale effects Espinoza, R.D. and Bray, J.D. (1995). “An integrated approach
on the stress-strain characteristics of fiber-reinforced soils to evaluating single-layer reinforced soils.” Geosynthetics
have not been investigated fully on a large scale, the actual International, 2(4), 723-739.
behavior of fiber-reinforced soils is not yet well known. It is Fluet, J.E. (1988). “Geosynthetics for soil improvement: a
expected that further studies including especially large-scale general report and keynote address.” Proc., Symposium
tests with fiber of larger diameter can explain in better way on Geosynthetics for Soil Improvement, Tennessee,
the behavior of fiber-reinforced soils. It is noted that because U.S.A., 1-21.
Fundamental concepts of soil reinforcement — an overview  341

Freitag, D.R. (1986). Soil randomly reinforced with fibers. Kumar, P. and Singh, S.P. (2008). “Fibre-reinforced fly ash
Journal of Geotechnical Engineering, ASCE, 112(8), 823- subbases in rural roads.” Journal of Transportation
826. Engineering, ASCE, 134(4), 171-180.
Giroud, J. P. and Noiray, L. (1981). “Geotextile-reinforced Kumar, R., Kanaujia, V.K., and Chandra, D. (1999).
unpaved road design.” Journal of Geotechnical “Engineering behavior of fiber-reinforced pond ash and
Engineering, ASCE, 107(9), 1233-1254. silty sand.” Geosynthetics International, 6(6), 509-518.
Giroud, J.P., Ah-Line, A. and Bonaparte, R. (1984). “Design of Lindh, E. and Eriksson, L. (1990). “Sand reinforced with
unpaved roads and trafficked areas with geogrids.” Proc., plastic fibers: a field experiment.” Proc., International
Symposium on Polymer Grid Reinforcement, London, Reinforced Soil Conference, Glasgow, U.K., 471-474.
116-127. Long, N.T., Guegan, Y. and Legeay, G. (1972). Etude de la
Gosavi, M., Patil, K.A., Mittal, S. and Saran, S. (2004). terre armee a l’appareil triaxial. Rapport de Recherche No.
“Improvement of properties of black cotton soil sub- 17, Laboratoire Central des Ponts et Chaussees, Paris,
grade through synthetic reinforcement.” Journal of the France.
Institution of Engineers (India), 84, 257-262. Love, J.P., Burd, H.J., Milligan, G.W.E. and Houlsby (1987).
Gray, D.H. and Ohashi, H. (1983). Mechanics of fiber rein- “Analytical and model studies of reinforcement of a
forcement in sand. Journal of Geotechnical Engineering, layer of granular fill on soft clay subgrade.” Canadian
ASCE, 109(3), pp. 335-353. Geotechnical Journal, 24, 611-622.
Gray, D.H. and Al-Refeai, T. (1986). “Behaviour of fabric Madhav, M.R. and Poorooshasb, H.B. (1988). “A new
versus fiber reinforced sand.” Journal of Geotechnical model for geosynthetic-reinforced soil.” Computers and
Engineering, ASCE, 112(8), 804-820. Geotechnics, 6, 277-290.
Guido, V.A., Dong, K.G. and Sweeny, A. (1986). “Comparison Madhav, M.R. and Poorooshasb, H.B. (1989). “Modified
of geogrid and geotextile reinforced earth slabs.” Pasternak model for reinforced soil.” Mathematical and
Canadian Geotechnical Journal, 23(1), 435-440. Computational Modelling, 12, 1505-1509.
Gupta, P.K., Saran, S. and Mittal, R.K. (2006). “Behaviour of Maher, M.H. and Gray, D.H. (1990). “Static response of sands
fiber reinforced sand in different test conditions.” Indian reinforced with randomly distributed fibers.” Journal of
Geotechnical Journal, 36(3), 272-282. Geotechnical Engineering, ASCE, 116(11), 1661-1677.
Hausmann, M.R. (1976). “Strength of reinforced soil.” Proc. McGown, A. and Andrawes, K.Z. (1977). “The influence of
Australian Road Research Board Conference, 8, 1-8. nonwoven fabric inclusions on the stress-strain behavior
Hausmann, M.R. (1990). Engineering Principles of Ground of a soil mass.” Proc., International Conference on the Use
Modification. McGraw-Hill, New York. of Fabrics in Geotechnics, Paris, 161-166.
Hausmann, M.R. and Vagneron, J. (1977). “Analysis of soil- McGown, A., Andrawes, K.Z. and Al-Hasani, M.M. (1978).
fabric interaction.” Proc., Int. Conf. on the Use of Fabrics “Effect of inclusion properties on the behavior of sand.”
in Geotechnics, Paris, 139-144. Geotechnique, 28(3), 327-346.
Heineck, C.S., Consoli, N.C. and Coop, M.R. (2005). “Effect of Michalowski, R.L. and Cermak, J. (2003). “Triaxial com-
microreinforcement of soils from very small to large shear pression of sand reinforced with fibers.” Journal of
strains.” Journal of Geotechnical and Geoenvironmental Geotechnical and Geoenvironmental Engineering, ASCE,
Engineering, ASCE, 127(3), 258-268. 129(2), 125-135.
Hoare, D.J. (1979). “Laboratory study of granular soils rein- Michalowski, R.L. and Zhao, A. (1996). “Failure of fiber-
forced with randomly oriented discrete fibers.” Proc., reinforced granular soils.” Journal of Geotechnical
International Conference on Soil Reinforcement, Paris, Engineering, ASCE, 122(3), 226-234.
47-52. Narain, J. (1985). “Reinforced earth.” Indian Geotechnical
Jewell, R.A. (1980). Some effects of reinforcement on the Journal, 15(1), 1-25.
mechanical behavior of soils. Ph.D. thesis, University of Pinto, M.I.M. (2004). Reply to the discussion of “Applications
Cambridge, England. of geosynthetics for soil reinforcement” by M.I.M. Pinto,
Jewell, R.A. (1996). Soil Reinforcement with Geotextiles. Ground Improvement, 7(2), 2003, 61–72 by S.K. Shukla.
Construction Industry Research and Information Ground Improvement, UK, 8, 180-181.
Association, London, CIRIA, Special Publication 123. Ranjan, G., Vasan, R.M. and Charan, H.D. (1994a).
Koerner, R. M. (2005). Designing with geosynthetics. 5th edi- “Mechanism of fiber-reinforced soil – State of-the-
tion, Prentice Hall, New Jersey, USA. art.” Behaviour of plastic-fiber reinforced sand. Indian
Geotechnical Journal, 24(3), 241-262.
342  International Journal of Geotechnical Engineering

Ranjan, G., Vasan, R.M. and Charan, H.D. (1994b). Vidal, H. (1966). La terre Armée. Annales de l’Institut
“Behaviour of plastic-fiber reinforced sand.” Geotextiles Technique de Batiment et de Travaux Publics, France.
and Geomembranes, 13, 555-565. Vidal, H. (1969). “The principle of reinforced earth.” Highway
Ranjan, G. Vasan, R.M. and Charan, H.D. (1996). Research Record 282, 1-16.
“Probabilistic analysis of randomly distributed fiber- Waldron, L.J. (1977). “Shear resistance of root-permeated
reinforced soil.” Journal of Geotechnical Engineering, homogeneous and stratified soil.” Proc., Soil Science
ASCE, 122(6), 419-426. Society of America, 41(5), 843 – 849.
Sanatoni, R.L., Tingle, J.S. and Webster, S.L. (2001). Yang, Z. (1972). Strength and deformation characteristics of
“Engineering properties of sand –fiber mixtures reinforced sand. Ph.D. thesis, University of California,
for road construction.” Journal of Geotechnical and Los Angeles.
Geoenvironmental Engineering, ASCE, 127(3), 258-268. Yetimoglu, T. and Salbas, O. (2003). “A study on shear
Sawicki, A. (2000). Mechanics of Reinforced Soil. A.A. strength of sands reinforced with randomly distributed
Balkema, Rotterdam discrete fibers.” Geotextiles and Geomembranes, 21, 103-
Schlosser, F. and Long, N.-T. (1974). “Recent results in 110.
French research on reinforced earth.” Journal of the
Construction Division, ASCE, 100(3), 223-237.
Schlosser, F. and Vidal, H. (1969). “Reinforced earth.” Bulletin
de Liaison des Laboratoires des Ponts et Chaussees, 41,
France.
Sellmeijer, J.B., Kenter, C.J. and Van den Berg, C. (1982).
“Calculation method for fabric reinforced road.” Proc.,
2nd International Conference on Geotextiles. Las Vegas,
USA, 393-398.
Sellmeijer, J.B. (1990). “Design of geotextile reinforced
unpaved roads and parking areas.” Proc., 4th International
Conference on Geotextiles, Geomembranes and Related
Products. The Hague, Netherlands, 177-182
Shewbridge, S.E. and Sitar, N. (1989). “Deformation char-
acteristics of reinforced sand in direct shear.” Journal of
Geotechnical Engineering, 115(8), 1134-1147.
Shewbridge, S.E. and Sitar, N. (1990). “Deformation based
model for reinforced sand.” Journal of Geotechnical
Engineering, 116(7), 1153-1170.
Shukla, S.K. (ed.) (2002). Geosynthetics and Their Applications.
Thomas Telford, London.
Shukla, S.K. (2004). Discussion of “Applications of geosyn-
thetics for soil reinforcement” by M.I.M. Pinto, Ground
Improvement, 7(2), 2003, 61 – 72. Ground Improvement,
UK, 8, 179-181
Shukla, S.K. and Yin, J.-H. (2006). Fundamentals of
Geosynthetic Engineering. Taylor and Francis, London.
Shukla, S.K. and Chandra, S. (1994). “A generalized mechan-
ical model for geosynthetic-reinforced foundation soil.”
Geotextiles and Geomembranes, 13, 813-825.
Shukla, S.K. and Chandra, S. (1995). “Modelling of geosyn-
thetic-reinforced engineered granular fill on soft soil”.
Geosynthetics International, USA, 2(3), 603-618.
Tingle, J.S., Sanatoni, R.S. and Webster, S.L. (2002). “Full
scale field tests on discrete fiber reinforced sand.” Journal
of Transportation Engineering, ASCE, 128(1), 9-16.

View publication stats

You might also like