Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Assessing Liquefaction Resistance of Fiber-Reinforced

Sand Using a New Pore Pressure Ratio


Xidong Zhang 1 and Adrian R. Russell, Ph.D. 2

Abstract: Mixing discrete flexible fibers into sand is a reinforcement technology that increases the sand’s strength as well as ability to resist
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

liquefaction. This study demonstrates the mechanisms by which the applied loads are distributed across and shared by a sand’s skeleton, the
fibers, and the pore water in drained and undrained triaxial compression tests. It is shown how the effective stress on the sand skeleton may be
quantified by invoking the rule of mixtures and separate constitutive laws for the fibers and sand skeleton. It is also shown how the fibers alter
the load paths experienced by the sand skeleton. Most notably it is shown how the fibers prevent liquefaction and that the conventionally
defined pore water pressure ratio may incorrectly indicate otherwise. Alternate and more suitable pore water pressure ratios are introduced,
accounting for the transversely isotropic orientation distribution of fibers that prevails in most fiber-reinforced soils, that gives an accurate
indication of how far away the reinforced sand skeleton is from a liquefied state. The new pore pressure ratios, which are dependent on the
direction of the principal stresses and account for the fiber orientations, are for use in any situation when the potential liquefaction of fiber-
reinforced sand is a concern. Use of the conventional pore pressure ratio, which until now has been the only one employed in the literature,
may make the fiber-reinforcement technology appear less effective at suppressing liquefaction than it actually is. This may, incorrectly, hinder
its uptake in industry. DOI: 10.1061/(ASCE)GT.1943-5606.0002197. © 2019 American Society of Civil Engineers.
Author keywords: Fiber reinforcement; Static liquefaction; Effective stress; Fiber stress; Pore water pressure ratio.

Introduction ratio, which approaches unity for an unreinforced loose soil as


liquefaction develops, does not apply. In fact, the conventionally
Mixing discrete flexible fibers into soils to improve their mechani- defined pore pressure ratio may reach unity for a fiber-reinforced
cal properties has aroused the interest of geotechnical engineers soil but the soil is quite far from a liquefied state. It remains very
for many years. The benefits include making a soil stronger, in stable and will continue to remain so with further undrained load-
some cases failure resistant, as well as more able to resist liquefac- ing. Assuming that the conventional pore pressure ratio of unity
tion. The benefits seem too good for the geotechnical engineering signifies liquefaction (e.g., Krishnaswamy and Isaac 1994, 1995;
industry to ignore. Even so, very few industry applications have Boominathan and Hari 2002; Wang and Brennan 2015) may make
been reported (e.g., Santoni and Webster 2001; Tingle et al. 2002; the fiber-reinforcement technology appear less appealing and less
Zornberg and Kavazanjian 2002; Park and Tan 2005; Hazirbaba effective than it actually is.
and Gullu 2010; Russell et al. 2017). One of the factors that There have been many academic studies on the fiber-reinforced
may hinder the industry’s full appreciation of the benefits, espe- soil technology; however, there are too many for a complete review
cially the enhanced liquefaction resistance, surrounds a common in this paper. Many involve soil element tests and undrained loading.
underappreciation in the literature of how the stresses applied to It is usually observed that liquefaction of a sand, when induced by
a soil–fiber composite are distributed to the fibers, the soil skeleton cyclic loading, is delayed with the inclusion of fibers. This is dem-
and the pore water pressure. The effective stress of a soil–fiber onstrated by an increased number of loading cycles required for
composite is rarely quantified in the literature leading to misinfor- liquefaction to occur and the decreased cyclic strain accumulation
mation around what constitutes a liquefied state. Furthermore, there (e.g., Krishnaswamy and Isaac 1994, 1995; Unnikrishnan et al.
are no documented established methods outlining how the fibers 2002; Noorzad and Amini 2014; Eskisar et al. 2016). Shaking table
are best mixed into soil at field scale. There are no relevant stan- tests also show that fiber reinforcement in sand delays the build-up
dards or codes of practice (Shukla 2017). of pore water pressure and the development of a liquefied state
When fibers are present they help carry the externally applied (e.g., Maheshwari et al. 2012; Wang and Brennan 2013, 2015). Static
load and alter the stresses imposed on the soil skeleton, usually in a liquefaction has also been studied with Ibraim et al. (2010) and Liu
beneficial way. The conventional definition of the pore pressure et al. (2011) showing that fibers decrease static liquefaction potential
and prevent a strain softening behavior even when the sand is loose.
1
Research Student, Centre for Infrastructure Engineering and Safety, Many others involve drained loading (e.g., Gray and Ohashi 1983;
School of Civil and Environmental Engineering, Univ. of New South Gray and Al-Refeai 1986; Jewell and Wroth 1987; Michalowski and
Wales, Sydney, NSW 2052, Australia. Zhao 1996; Michalowski 1997; Consoli et al. 1998; Kaniraj and
2
Associate Professor, Centre for Infrastructure Engineering and Safety, Havanagi 2001; Ahmad et al. 2010; Diambra and Ibraim 2015). Fi-
School of Civil and Environmental Engineering, Univ. of New South Wales, bers increase the shear strength while also changing the deformation
Sydney, NSW 2052, Australia (corresponding author). Email: a.russell@
characteristics. The effect of fiber inclusions on a reinforced soil’s
unsw.edu.au
Note. This manuscript was submitted on March 7, 2019; approved on dilation may depend on how the volumes of solids and fibers are
August 27, 2019; published online on November 12, 2019. Discussion per- combined in measures of volumetric strain. In research where the
iod open until April 12, 2020; separate discussions must be submitted for volume of the soil skeleton is constant the presence of fibers in-
individual papers. This paper is part of the Journal of Geotechnical and creases the soil’s dilation (e.g., Diambra et al. 2010; Falorca and
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. Pinto 2011; Wood et al. 2016). Other research where total volume

© ASCE 04019125-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Particle size distribution curve of Sydney sand.


Fig. 2. (a) Agglomerate Loksand crimped fibers; and (b) four single
Loksand fibers.
Table 1. Index properties of Sydney sand
Property Value
Mean grain size, D50 0.28 mm Table 2. Properties of the Loksand fibers
Coefficient of uniformity, Cu 1.94 Property Value
Coefficient of curvature, Cc 0.86
Fines content 0.36% Length 35 mm
Maximum void ratio, emax 0.97 Diameter 0.1 mm
Minimum void ratio, emin 0.6 Tensile strength 225 MPa
Specific gravity, Gs 2.65 Specific gravity, Gf 0.91
Surface properties Crimped

of solids (i.e., the volume of fibers plus the volume of the soil
skeleton) is constant shows fiber inclusions decrease dilation wetted to 10% moisture content prior to adding the fibers. A sand or
(e.g., Michalowski and Zhao 1996; Michalowski and Čermák sand–fiber mixture is tamped into a mold in four layers with equal
2003). In the former case the added fibers replace some of the height, targeting a void ratio of 1.09. A tamped sample in a mold is
voids, producing a denser sample and therefore more dilation oc- then put into a freezer for 12 h. A sample is then taken out of the
curs (Wood et al. 2016). mold and positioned in the triaxial apparatus. The freezing did not
The motivation of this study is to demonstrate how the applied cause any significant change to a sample’s height indicating that
loads are distributed through and shared by each component of the sample disturbance may be assumed negligible. Enlarged and lu-
sand–fiber composite. The method for quantifying the load sharing bricated end platens are used in all tests to ensure deformation
will be detailed. Examples are given on how the effective stress on remains uniform throughout as shearing proceeds and strain locali-
the soil skeleton is determined. It is shown that fibers prevent lique- zation is avoided. This also facilitates the ability of fibers to con-
faction and that the conventional pore water pressure ratio incor- tinually strengthen the sand as shearing proceeds.
rectly indicates otherwise. New direction-dependant pore water Stress-controlled triaxial tests are conducted on both unrein-
pressure ratios are defined to enable more reasonable predictions forced and fiber-reinforced sand samples in both drained and un-
of liquefaction development. drained conditions. Isotropic confining stresses of 50, 100, and
200 kPa are applied to the samples. Liquefaction is most likely to
occur when low confining stresses prevail (e.g., Seed et al. 1983;
Test Materials and Test Conditions Tokimatsu and Yoshimi 1983; Robertson and Campanella 1985).
The tests performed are listed in Table 3. The sample void ratios
Materials at the end of consolidation are also included, accounting for den-
sification during saturation and consolidation. The fibers are treated
Sydney sand, a poorly graded sand, has been used in all tests. Its as being part of the solid in void ratio calculations. Skempton’s
particle size distribution curve is shown in Fig. 1 and index proper- coefficient (B-value) was at least 0.99 for each sample. In Table 3,
ties are listed in Table 1. pc represents the mean confining stress acting on samples (in ex-
Loksand fibers are used, which are crimped polypropylene fi- cess of back-pressure) at the end of consolidation.
bers (Fig. 2 and Table 2), as they are easily separable and mixed in
to sand. Mixing is done using a variety of fiber contents (FCs),
0.25%, 0.5%, and 0.75%, representing the fiber mass (M f ) as a Test Results
percentage of the dry sand mass (Ms ). A second fiber is also con-
sidered later in the paper to demonstrate that not all fibers are
equally effective at suppressing liquefaction. Stress and Strain Notations
In this study triaxial notations are adopted where p and q denote the
Test Conditions and Test Program total mean stress (in excess of back-pressure) and deviator stress on
the composite
Unreinforced and fiber-reinforced samples are prepared using
moist tamping to produce very loose states. The samples are cylin- σa þ 2σr
p¼ ; q ¼ σa − σr ð1Þ
drical in shape with a diameter and length of 50 mm. The sand is 3

© ASCE 04019125-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


where σa and σr = axial and radial stresses, respectively, corre- In this case, p 0 and q 0 denote the effective mean stress and de-
sponding to major and minor principal stresses; and p and q = viator stress carried by the sand skeleton, and pf and qf denote the
effective mean stress and deviator stress of the composite, where mean stress and deviator stress carried by the fibers.
p ¼ p − u with u denoting the pore water pressure (in excess The volumetric and deviator strains on the composite are
of back-pressure) and q ¼ q. In a drained condition p ¼ p.
2
εv ¼ εa þ 2εr ; εq ¼ ðεa − εr Þ ð2Þ
Table 3. Tests performed in the study 3
Test FC (%) pc (kPa) Void ratio after consolidation
where εa and εr = axial and radial strains, respectively.
CD-MC 0 50 1.0446
0.25 50 1.0151
0.5 50 1.0094 Drained Triaxial Compression Tests
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

0 100 1.0339
0.25 100 1.0124 Drained triaxial tests results for FC ¼ 0%, 0.25%, and 0.5% and
0.5 100 0.9874 two confining stresses (100 and 200 kPa) are shown in Figs. 3
0 200 1.0414 and 4. Both deviatoric and volumetric behaviors are affected by
0.25 200 1.0007 the addition of fibers. A greater fiber content induces a greater abil-
0.5 200 0.9966 ity to carry deviatoric stress at a given shear strain. For example, for
CU-MC 0 50 1.0480 a confining stress of 200 kPa, the deviator stress in the sample with
0.25 50 1.0337
FC ¼ 0.5% at εq ¼ 40% is 2.5 times that for an unreinforced sam-
0.5 50 0.9983
0.75 50 0.9871
ple. Fibers remain effective and enhance the deviator stress even as
0 100 1.0446 large shear strains have been reached, as illustrated by the continu-
0.25 100 1.0028 ous ascent of stress-strain curves.
0.5 100 0.9876 All samples are very loose and their volumes continuously con-
0.75 100 0.9850 tract under drained shearing. The amount of volumetric contraction
0 200 1.0371 is fiber content dependent. Adding more fibers into sand makes it
0.25 200 1.0004 more contractive. The tensile stress in the fibers provides an added
0.5 200 0.9870 confinement to the sand skeleton and may be one possible cause of
0.75 200 0.9770 the enhanced volumetric contraction. The fiber-reinforced samples
Note: CD = consolidated drained; CU = consolidated undrained; and continually contract at large shear strains, which demonstrates the
MC = monotonic compression. continuing contribution by the fibers.

Fig. 3. Drained triaxial compression tests under confining stress of 100 kPa: (a) deviator stress-shear strain; and (b) volumetric strain-shear strain.

Fig. 4. Drained triaxial compression tests under confining stress of 200 kPa: (a) deviator stress-shear strain; and (b) volumetric strain-shear strain.

© ASCE 04019125-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Undrained triaxial compression tests under confining stress of 50 kPa: (a) deviator stress-shear strain; (b) conventional pore pressure ratio–
shear strain; and (c) stress path in q∶p plane.

Undrained Triaxial Compression Tests stress ratio q=p 0 . When the effective stress path during undrained
loading intercepts the FLL instability is triggered. The FLL that best
Undrained triaxial tests results for FC ¼ 0%, 0.25%, 0.5%, and
fits the data in this paper for unreinforced samples is shown in Fig. 8.
0.75% and three confining stresses (50, 100, and 200 kPa) on loose
The stress ratio is 0.74. ru ranges from 0.50 to 0.53 as the FLL is
samples are shown in Figs. 5–7. In unreinforced sand samples static
intercepted.
liquefaction occurs for each confining stress in the usual way, with
The effective mean stress of the composite (p ) decreases during
the deviator stress peaking at an early stage of shearing (at εq ¼
the early stages of loading due to the build-up of pore water pres-
1%–2%), followed by the initiation of instability and a dramatic
sure [Figs. 5(c), 6(c), and 7(c)]. The changes to p then reverse. The
decrease of deviator stress and development of a flow deformation.
later stages of a test see an increase of p with shearing. For a given
The responses for the fiber-reinforced samples differed signifi-
confining stress, for all fiber contents, the stress paths appear to
cantly in that the deviator stress did not ever decrease with increas-
move along (or approach) unique lines in the q∶p plane, with
ing shear strain.
slopes depending on the confining stress.
The evolution of the conventionally defined pore water pressure
The p reversal point has some fundamental differences to the
ratio, ru ¼ u=pc , is also plotted for each test. Variable pc denotes phase transformation point observed for dense and unreinforced sand
the total mean stress on a sample (in excess of back-pressure) at the under undrained loading (Ishihara et al. 1993). In the reinforced sand
end of consolidation and the beginning of shear such that pc is the p reversal occurs while the pore water pressure continues to rise,
equal to the pc values in Table 3. For a standard drained or un- while in a dense and unreinforced sand it is accompanied by a di-
drained triaxial test pc is also equal to the total radial stress on rection change of the incremental pore water pressure.
a sample. The ru builds up rapidly during the initial stages of load- What should not be ignored is that the reinforced samples with
ing, reaching around 0.6 when at shear strains of less than 5% in- FC ¼ 0.25% experience a short-term instability at the initial stage
creasing to 1 thereafter. In an unreinforced sand this signifies of loading for all confining stresses. Shear deformations rapidly
liquefaction as the effective radial confining stress becomes zero. increase under almost constant deviator stresses during the periods
However, in a reinforced sand ru may reach 1 without liquefaction of temporary instability. Larger FCs were not associated with a
occurring. Reinforced sands are able to carry load, and flow defor- temporary instability. A FC of 0.25% makes only a limited contri-
mation is absent even though ru ¼ 1. An alternate pore pressure bution to the confinement of the sand skeleton, too small to com-
ratio is needed to signify liquefaction development in reinforced pletely prevent the temporary instability.
sands. This will be dealt with in the following sections.
In unreinforced sands a characteristic used to predict the initiation
of instability and strain softening prior to static liquefaction is re- Determination of Fiber Contribution
ferred to as a flow liquefaction line (FLL). It is a straight line in
the q∶p 0 plane, radiating from the origin, representing the peak q
Model for Fiber Stress Calculation
values (and associated p 0 values) observed in undrained loading
(e.g., Vaid and Sivathayalan 2000; Yang 2002; Sivathayalan and To determine how the load applied to a composite is distributed
Ha 2011; Baki et al. 2012). All points on the FLL have a constant to the sand skeleton, the fibers, and the pore water pressure, the

© ASCE 04019125-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Undrained triaxial compression test under confining stress of 100 kPa: (a) deviator stress-shear strain; (b) conventional pore pressure
ratio–shear strain; and (c) stress path in q∶p plane.

Fig. 7. Undrained triaxial compression test under confining stress of 200 kPa: (a) deviator stress-shear strain; (b) conventional pore pressure
ratio–shear strain; and (c) stress path in q∶p plane.

© ASCE 04019125-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Elastic Modulus of Fibers, Ef
Tension tests on single Loksand fibers were conducted by Diambra
et al. (2010) to determine Ef ¼ 900 MPa.

Sliding Factor, f b
The physical interaction between fibers and sand particles controls
the stress transfer between the two components. An assumption
underlying Eq. (5) is that strains in the fibers are not equal to
the strains of the sand skeleton (which are also the strains of the
composite) due to sliding (Diambra et al. 2010). An imperfect bond
exists between fibers and the sand skeleton. The fibers are able to
slide through the skeleton as the skeleton deforms. Some resistance
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

is provided, due to friction, interlocking, and the particles clamping


the fibers, so full sliding does not occur and the fibers provide some
Fig. 8. Initiation of instability in unreinforced sand. benefit. A dimensionless sliding factor fb is used to link the strains
in the fibers (ε̇f ) to the strains in the composite (ε̇)

constitutive model of Diambra et al. (2010) is adopted. In that ε̇f ¼ fb ε̇ ð7Þ


model fibers are assumed to behave elastically in tension and their
contribution to the composite is scaled by their volumetric fraction, Many features affect fb including fiber length and diameter,
as per the rule of mixtures. The sand skeleton has its own constit- sand particle sizes and size distributions, confining stresses, and
utive law and also contributes to the composite by its volumetric drainage conditions. The value of fb must lie between 0 and 1.
fraction. The effective stresses in the composite are then In this case, fb ¼ 0 represents complete sliding while fb ¼ 1 rep-
   0   resents a perfect bond. In this study fb for Loksand fiber is defined
p p pf by a power function in terms of p =pref

¼ μm 0 þ μf ð3Þ
q q qf
f b ¼ 0.034ðp =pref Þ0.43 ð8Þ
where μm ¼ ðV s þ V v Þ=V ¼ ðV − V f Þ=V is the volumetric frac-
tion of the soil skeleton; μf ¼ V f =V is the volumetric fraction of
In Eq. (8), p is the mean effective stress of composite while
the fibers; μm þ μf ¼ 1; and V s , V v , V f , and V are the volumes of
pref is a reference pressure with value of 1 kPa. With p =pref being
the sand skeleton, voids, fibers, and the overall composite, respec-
dimensionless, fb values stay within the permissible range
tively. The approximation μm ¼ 1 always applies in both drained
(0 < f b < 1) for all p values in the tests conducted. The procedure
and undrained tests and is adopted in this paper. The μf changes
used to determine this function is detailed subsequently when the
during drained tests as V changes. It may be computed using
 Gs   Gs  data is synthesized.
Gf FC Gf FC In the paper the stress in a single fiber and the relative slippage
μf ¼ V f þV v0 þΔV
¼ ð4Þ between a fiber and the adjacent sand have been averaged, assum-
1þ 1 þ e0 þ ΔVVs
Vs ing the mobilized stress and the slippage are uniform everywhere.
where Gs and Gf denote the specific gravities of sand and fibers, In Eq. (7) fb is introduced as a factor which quantifies the average
respectively; V v0 and ΔV are the initial volume of voids in a sample degree of slippage along the fiber surface. Note that this is an
and the volumetric change due to shearing; and e0 = initial void ratio approximation but convenient to formulate a continuum based
of the sample by treating fibres as being part of voids. model. For a single fiber embedded in the soil the tensile stress
The modification of Diambra et al. (2013), in which the voids in induced may not be constant along its length. They usually reduce
a composite are associated partly with the fibers and partly with the to zero at a fiber’s ends (Michalowski and Zhao 1996; Diambra
sand skeleton, is not adopted in this paper for simplicity. et al. 2010, 2013). The degree of relative slippage between fibers
The fibers’ contributions to the composite stresses in incremen- and sand skeleton may also vary along the fiber length.
tal form are as follows:
       Fiber Orientation Distribution Function, ρθ
ṗf ε̇v M 11 M 12 ε̇v The fibers and sand grains are assumed to distribute homogeneously
μf ¼ Ef fb ½M f  ¼ Ef f b ð5Þ
q̇f ε̇q M 21 M 22 ε̇q throughout a composite. However, the moist tamping method gives
rise to a transversely isotropic fiber orientation distribution. The fi-
The M 11 , : : : components are defined as bers tend to have a subhorizontal orientation (Diambra et al. 2007;
2 3 Ibraim et al. 2012), symmetrical along a cylindrical sample’s axis.
1 1 2 2 2Rθ 3
4 The fiber orientation distribution can be described by Eq. (9), which
2 3 69 9 9 7
9 76 0
0
ρðθÞ cosðθÞsin ðθÞdθ
6 7 was originally proposed by Michalowski and Čermák (2002) and
M 11 6 17 6Rθ 7
6 7 61 1 2 76 0 0 ρðθÞcos3 ðθÞsin2 ðθÞdθ 7 has been adopted by Diambra et al. (2007)
6 M 12 7 6 − − 76 Z 7
6 7 63 6 3 3 76 1 θ 0 7
6 7¼6 76 ρðθÞcos 3
ðθÞsin 2
ðθÞdθ 7
6 M 21 7 6 1 1 1 1 76 2 7 ρðθÞ ¼ μf ðA þ Bjcosn θjÞ ð9Þ
4 5 6 − − 76 0 7
63 3 3 3 76 Z θ 7
M 22 6 74 1 0 5
4 1 1 5 ρðθÞcos5 ðθÞdθ where ρðθÞ = volumetric concentration of fibers distributed in an
1 − −1 2 0
2 2 infinitesimal volume lying with an angle θ off the horizontal plane.
In Eq. (9) there are three parameters which are linked by
ð6Þ
1−A
The parameters Ef , fb , and θ0 and the function ρðθÞ are now B ¼ R π=2 ð10Þ
detailed. 0 cosnþ1 ðθÞdθ

© ASCE 04019125-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Determination of integration limits: (a) Mohr circle of strain; and (b) domains of tensile strain in triaxial compression.

so only two of them are independent. To determine the parameters is also negative and represents the amount by which p 0 shifts right-
four samples are made, frozen, and sectioned. As per the procedure ward from p. The μf qf is positive and represents the amount by
of Diambra et al. (2007) the numbers of fibers intersecting vertical which q 0 shifts downward from q. The tension in the fibers acts like
and horizontal sections through finite areas of 20 × 20 mm are an extra (anisotropic) confinement on the sand skeleton that gradu-
counted. The ratios between the numbers of fibers intersecting areas ally evolves as the composite is deformed. A general stress path for
on vertical and horizontal sections are 2.10, 2.00, 2.06, and 2.04, the sand skeleton is illustrated in Fig. 10. Notice that, at large de-
enabling the best-fit parameters A ¼ 0, B ¼ 2.04, and n ¼ 5 to be formations, the stress state of the sand skeleton is assumed to ap-
determined. The same parameters happened to be arrived at by proach the sand’s critical state line in the q 0 ∶p 0 plane.
Diambra et al. (2007), again using Loksand fibers and moist tamp-
ing, for a different sand. When A ¼ 0, B ¼ 2.04, and n ¼ 5, ρðθÞ Fiber Contribution in Undrained Triaxial Compression
represents a distribution where 97% of fibers are orientated π=4 The total stress path followed by the composite for undrained tri-
from the horizontal. Note that Eq. (9) imposes some difficulties axial compression obeys δq=δp ¼ 3. The effective stress path fol-
during the integration of Eq. (3). To avoid this, a slightly modified lowed by the composite shifts leftward due to the rise of the pore
function ρðθÞ ¼ 2μf ab2 j cos θj=ðcos2 ðθÞðb2 − a2 Þ þ a2 Þ may be water pressure as deformation proceeds (Fig. 11). The effective
used instead, with a ¼ 1.02 and b ¼ 0.46, which produces a distri- stress path of the sand skeleton is shifted from that of the composite
bution that is virtually identical to Eq. (9). by amounts controlled by μf σrf and μf σaf , as for the drained case.
In general, for the same reasons as outlined for the drained case,
Limit of Integral, θ0
μf pf is negative and represents the amount by which p 0 shifts
Only the fibers with act in tension carry a share of the stresses im-
rightward from p , while μf qf is positive and represents the
posed on the composite. The integration limit θ0 in Eq. (6) ensures
that only values of θ are considered for which δεθ < 0. The Mohr amount by which q 0 shifts downward from q . A general stress
circle of strain increments is shown in Fig. 9 with the domain of path for the sand skeleton is illustrated in Fig. 11. Again, notice
tensile strains shown. In this case, θ0 bound two directions of zero that at large deformations the stress state of the sand skeleton ap-
incremental strain, between which tensile strains prevail. The θ0 for proaches the sand’s critical state line. A detailed discussion on
triaxial compression tests is defined as stress equilibrium in a fiber-reinforced sand in an undrained con-
rffiffiffiffiffiffiffiffiffi dition will be given in the following sections.
ε̇
θ0 ¼ arctan − r ð11Þ
ε̇a

General Description

Fiber Contribution in Drained Triaxial Compression


The stress path followed by the composite during drained triaxial
compression obeys δq=δp ¼ δq =δp ¼ 3. The interaction be-
tween the fibers and the sand skeleton complicates the stress path
followed by the sand skeleton such that δq 0 =δp 0 ≠ 3. To gain a
general understanding of the fiber influences recall that (1) fibers
only carry tensile loads mobilized by tensile strains on the
composite, (2) the maximum tensile strains on the composite are
horizontal, and (3) most fibers have a near horizontal orientation.
It follows that the fiber stresses are more pronounced in the hori-
zontal (σr ) direction than the vertical (σa ) direction. The μf σrf is a
Fig. 10. Progression of the sand skeleton’s stress path under drained
significant tensile (negative) stress. The μf σaf is also a tensile
triaxial compression.
stress but with a lesser magnitude than μf σrf . Therefore, μf pf

© ASCE 04019125-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


p =pref in Fig. 12. Note that f b increases with the rise of p
and a power function [Eq. (8)] fits the data well.
Note that the sand skeleton may not have completely reached a
critical state at εq ¼ 40%. Indeed the volumetric strain is continu-
ally decreasing at εq ¼ 40% [Figs. 3(b) and 4(b)]. Therefore, the
stress state may lie slightly above the critical state causing a slight
underestimation of f b values. This results in a conservative esti-
mate of the fiber contribution when used in practice.
Using this relationship for f b and the constitutive model for the
fibers described previously, the fiber stresses under drained and un-
drained triaxial compression may be calculated at any stage of
loading.
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

The stress paths of the composite and sand skeleton for drained
triaxial compression tests are shown in Fig. 13. A greater confining
Fig. 11. Progression of the sand skeleton’s stress path under undrained
stress leads to a greater contribution by the fibers.
triaxial compression.
The stress paths of the composite and sand skeleton for undrained
triaxial compression tests are shown in Fig. 14. The FLL is also
shown. At the early stages of loading the contribution of the fibers
to the composite is minor, evidenced by the negligible difference
Application between the two sets of effective stress paths shown. It is around
the reversal point of the composite’s effective stress p that the fibers
To determine the contributions by the fibers it is necessary to know begin to have a major contribution. At large shear strains the effective
f b . The variable fb is found by trial-and-error such that the stress stress path of the sand skeleton approaches the critical state line. A
states of the sand skeleton coincide with the critical state line in the significant fiber contribution remains and continually increases with
q 0 ∶p 0 plane at εq ¼ 40%. Only the test data for samples with FCs of further loading even though the sand skeleton state has approached a
0.25% and 0.5% are used, although generality to other FCs is also critical state. The fibers’ contributions generally increase with the
confirmed, as explained subsequently. The fb is plotted against increasing of the FC. However, there is no obvious relationship be-
tween the fibers’ contributions and the confining stresses applied.
As seen in Fig. 14 the effective stress paths of the sand skeleton
always cross the FLL without initiation of instability or static lique-
faction. Although a temporary instability is observed for FC ¼
0.25% its initiation did not coincide with the FLL being crossed.
The interaction between the fibers and sand skeleton is influ-
enced by many different factors. Generally, for the most beneficial
interaction, the length of fibers must be at least one order of mag-
nitude larger than the sand particle sizes (Michalowski 1997;
Michalowski and Čermák 2003). The evolutions of fb during tests
are showed in Fig. 15. In drained tests f b gradually increases with
the development of shear strain, indicating a progressively increas-
ing bond between fibers and the sand skeleton. In undrained tests
fb decreases initially because of the rapid building up of pore water
pressure and reduction of p , then gradually increases with con-
tinuous loading. In Fig. 15(a), for the sample with FC ¼ 0.25%,
Fig. 12. Relationship between sliding factor and effective mean stress
fb decreases most significantly upon initial loading corresponding
acting on the composite.
to the temporary instability that developed. Fig. 15(a) shows that a

Fig. 13. Stress contribution of fibers under drained triaxial compression when (a) FC ¼ 0.25%; and (b) FC ¼ 0.5%.

© ASCE 04019125-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Stress contribution of fibers under undrained triaxial compression when (a) FC ¼ 0.25%; (b) FC ¼ 0.5%; and (c) FC ¼ 0.75%.

Fig. 15. Variation of the sliding factor estimated by Eq. (8) when (a) pc ¼ 100 kPa; and (b) FC ¼ 0.5%.

better bond develops in drained conditions compared to undrained skeleton becomes zero. Liquefaction occurs and the sand becomes
conditions. Fig. 15(b) shows that a greater confining stress is as- fluidized. However, in fiber-reinforced sand, when this convention-
sociated with a greater bond between fibers and the sand matrix, ally defined ru becomes equal to 1, the effective stress on the sand
which was also found by Ibraim et al. (2010). skeleton may be far from zero. A fluidized state is absent. The sand
remains stable and able to carry significant loads. Therefore, a new
pore water pressure ratio is defined in this paper (denoted ru ) such
New Pore Water Pressure Ratio to Identify that ru ¼ 1 signifies when liquefaction occurs in a fiber-reinforced
Liquefaction Development sand. Only when ru ¼ 1 does the effective stress carried by the sand
skeleton become zero and does the reinforced sand loose its load
It is well known that when the pore water pressure builds up during carrying capacity.
shearing to equal the confining stress in loose unreinforced sand, The following equilibrium of mean stresses applies at any stage
i.e., when ru ¼ u=pc ¼ 1, the effective stress acting on the sand of a triaxial test on a fiber-reinforced sample:

© ASCE 04019125-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


p − u ¼ μm p 0 þ μf pf ð12Þ where μf σaf and μf σrf = fiber stresses in axial and radial direc-
tions, respectively; and σa0 and σr0 become zero and when ru;a ¼ 1
Assuming μm ¼ 1, a suitable definition for ru becomes and ru;r ¼ 1.
u þ μf pf Fig. 17 plots the three pore water pressure ratios in fiber-
ru ¼ ð13Þ reinforced samples with FC = 0.5% and 0.75% under a confining
p
stress of 100 kPa. The other tests show similar results. Each pore
The effective mean stress of the sand skeleton (p 0 ) becomes zero water pressure ratio increases to a peak and then drops, always
and thus liquefaction occurs when ru ¼ 1. being significantly below unity. The ru;r shows the greatest peak,
Plots of ru and ru for the fiber-reinforced samples are showed in the quickest decrease, and the attainment of the lowest value at
Fig. 16. In all fiber-reinforced samples ru increases and approaches large shear strains. The predominant near horizontal orientation
1 as loading progresses. Stress-strain behaviors (Figs. 5–7) show of the fibers, and the radial strains being tensile, cause the radial
that all fiber-reinforced samples remain stable and conventional stresses to be most heavily influenced by the fiber reinforcement.
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

liquefaction is absent even though ru has approached 1. Conversely, Conversely, ru;a shows the smallest peak, the slowest decrease, and
ru remains below 1 as loading progresses. It initially increases, the attainment of the largest value at large shear strains.
reaches a peak, and then decreases. It continues to decrease as shear A practical benefit to having different pore pressure ratios in
strains become very large, indicating the samples move further and different directions is that liquefaction susceptibility can be ad-
further away from a zero effective stress condition as deformation dressed, focusing on the most vulnerable direction. What becomes
continues. the most vulnerable direction will depend on the fiber orientation
Observing the development of liquefaction is complicated by and loading condition. A liquefied state, if it is possible at all,
the fact that u acts isotropically while fiber stresses are direction- would develop first in the most vulnerable direction. It would then
ally dependent and influenced by the orientation of fibers and the most likely ruin the interaction between fibers and sand skeleton.
strains developed in the soil. Two additional pore water pressure Further research is needed demonstrate this, although it would not
ratios may be defined for the axial direction (ru;a ) and radial di- be straightforward to conduct as some fiber types prevent liquefac-
rection (ru;r ) tion altogether.
u þ μf σaf The loading conditions of triaxial compression are unable to
ru;a ¼ ð14Þ
σa cause liquefaction in the sand when reinforced with Loksand fibers.
Although the pore water pressure ratios for different directions
u þ μf σrf
ru;r ¼ ð15Þ vary, they are always well below unity. Other fiber types may not
σr work as well to suppress liquefaction, as demonstrated next.

Fig. 16. Pore water pressure ratios of fiber-reinforced sand under undrained triaxial compression when (a) pc ¼ 50 kPa; (b) pc ¼ 100 kPa; and
(c) pc ¼ 200 kPa.

© ASCE 04019125-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 17. Pore water pressure ratios of fiber-reinforced sand in different directions under undrained triaxial compression when (a) pc ¼ 100 kPa,
FC ¼ 0.5%; and (b) pc ¼ 100 kPa, FC ¼ 0.75%.

Consideration of Another Fiber Type The void ratios after consolidation for the two samples are 1.0344
and 1.0235, respectively.
Not all fibers may work as well as Loksand fibers to suppress The alternate fiber provides a lesser contribution to the sand
the tendency for liquefaction. A prototype fiber, developed as a po- skeleton than Loksand. A significant drop of deviator stress occurs
tential alternative to Loksand, is considered in this paper to dem- in a loose sample even when the reinforcement is present. Although
onstrate this. It is a synthetic fiber with a length of 30 mm, diameter complete liquefaction does not occur the prevailing load carrying
of 0.3 mm, tensile strength of 650 MPa, elastic modulus of capacity is very low. Also there is an obvious instability, accom-
9,500 MPa (manufacturer provided), and has a rough surface. Four panied by a rapid development of deformation (flow deformation).
samples reinforced by this fiber were prepared using moist tamping In this case, ru and ru;r increase to near unity [Fig. 18(b)], indicat-
to determine the distribution of fiber orientation. The ratios between ing a completely liquefied state was nearly reached.
the numbers of fibers intersecting areas (20 × 20 mm) on vertical The different behaviors of loose sand reinforced by two different
and horizontal sections were 1.96, 1.69, 1.82, and 1.54. The param- fiber types indicate the importance of fiber properties. When length
eters that best-fit the orientation distribution [Eq. (9)] are A ¼ 0, and concentration are the same the fiber aspect ratio (length/diam-
B ¼ 1.70, and n ¼ 3 for this fiber, meaning that around 93% of
eter) strongly affects the reinforcing effect. A larger aspect ratio
fibers orientate within π=4 of the horizontal. The corresponding
produces more reinforcement, captured indirectly through a larger
parameters in the modified distribution function are a ¼ 0.85 and
sliding factor f b (Michalowsk and Čermák 2003; Consoli et al.
b ¼ 0.51. The sliding factor for this fiber, obtained by trial-and-
2009). In this study the aspect ratio of a Loksand fiber is 350, being
error, is defined in Eq. (16). The sliding factor for this fiber type,
about 3.5 times larger than that of the alternate fiber used (i.e., 100).
for a given p , is much lower than that for Loksand fibers, indicat-
For a given concentration the number of Loksand fibers is much
ing that Loksand fibers have a more effective interaction with the
sand skeleton larger than for the alternate fiber, providing more fiber–sand sur-
face interaction. A crimped surface combined with negligible bend-
fb ¼ 0.0012ðp =pref Þ0.39 ð16Þ ing stiffness enables sand particles to easily lock on to the fibers and
mobilize more friction.
Typical undrained triaxial tests on sand reinforced with this al- It is not just fb that is important. The magnitude of the
ternate fiber are presented in Fig. 18(a) for FC ¼ 0.25% and 0.5%. fiber stresses also depends heavily on the fiber’s elastic modulus.

Fig. 18. Undrained triaxial compression test results on sand reinforced with alternate fiber: (a) stress paths in the q 0 -p 0 (q-p ) plane; and (b) pore
water pressure ratios.

© ASCE 04019125-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


fixed fiber orientation distribution (equal to the initial condition)
and an evolving fiber orientation distribution. It can be seen that
fibers provide a greater mean stress contribution if the induced
anisotropy is taken into account, which is expected because more
fibers act in tension. The differences in the mean stress contribu-
tions are minor at small strains, being less than 5% at shear strains
less than 10%. The differences increase with the development of
shear strain. At a shear strain of 40% assuming a fixed fiber ori-
entation distribution underestimates the fiber’s mean stress by 17%
[Fig. 20(a)]. Also, the difference becomes more significant as the
FC increases. When the confining stress is 100 kPa, the difference
is only 5 kPa at the shear strain of 40% for a FC of 0.25%, while it
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

climbs to 26 kPa for a FC of 0.75%. The difference is not minor at


large shear strains, especially when the fiber content is large. How-
Fig. 19. Development of reinforcing effectiveness with increase of
ever, assuming a fixed fiber orientation distribution as we have
effective mean stress of composite.
done in this paper, leads to more conservative calculations of the
fiber contributions. It would also be conservative to assume that the
fiber orientation distribution is fixed in many practical applications.
The product Ef f b is a useful measure of the contribution by fibers
(Diambra et al. 2013). Although the alternate fiber has greater elas-
Conclusions
tic modulus than Loksand it has a much smaller f b. Fig. 19 illus-
trates the relationship between Ef f b and p =pref . The Ef f b is A series of drained and undrained triaxial compression tests have
always smaller for the alternate fiber, meaning it is less effective been performed on very loose sand reinforced with synthetic fibers
as a reinforcement. to investigate the mechanical behavior, especially whether or not
liquefaction can occur. Several conclusions have been made as
follows:
Discussion on Distortion of Samples • Significant improvement of the drained strength of very loose
sand is produced by fiber reinforcement using Loksand. The
In this paper the fiber stress is calculated based on the original ori- strength improvement is fiber content dependent, being greater
entation distribution of fibers. In other words, it is assumed that the for larger fiber contents. A volumetric contraction prevails in
orientation of fibers remains unchanged during the deformation fiber-reinforced samples even at large shear strains (up to
process. However, the fibers in samples may rotate approaching 40%). The drained stress path of sand skeleton shifts rightward
a more horizontal orientation under triaxial compression, especially and downward from that of the composite in the q 0 ∶p 0 plane,
as large strains are imposed. The evolution of fiber orientation approaching the critical state strength at large strains.
during the deformation process was noted by Michalowski and • In undrained tests static liquefaction occurs in unreinforced
Čermák (2002) and was referred to as an induced anisotropy. The sands regardless of the initial confining stresses acting on the
orientation distribution changes of Loksand fibers have been quan- samples, corresponding to the conventionally defined pore pres-
tified in this paper, building on Michalowski and Čermák’s (2002) sure ratio ru ¼ u=pc of 1. Instability initiated in the unreinforced
proposition. The results reveal that at shear strain of 40% there are samples around a stress ratio of q=p 0 ¼ 0.74, coinciding with
97% of Loksand fibers orientated within 28.5° of the horizontal. the flow liquefaction line. Static liquefaction is prevented by fi-
This is quite different to the initial condition where 97% of fibers ber reinforcement using Loksand although a temporary instabil-
are orientated within 45° of horizontal. ity, where deformation proceeds without an obvious change to
To account for the varying orientation distribution of fibers with the deviator stress, may occur when the fiber content is low
shear strain the parameters A, B, and n within Eq. (9) can be up- (0.25%). When fiber reinforcement is present the deviator stress
dated. Fig. 20 shows the mean stress contribution of fibers using a tends not to reduce at all. The fibers cause the deviatoric stress to

Fig. 20. Difference between the mean stress contributions of fibers calculated assuming fixed (FP) and varied orientation distribution parameters (VP)
in Eq. (9): (a) FC ¼ 0.5%; and (b) pc ¼ 100 kPa.

© ASCE 04019125-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


always increase as loading continues. However, the effective Diambra, A., E. Ibraim, D. M. Wood, and A. R. Russell. 2010. “Fiber re-
mean stress on fiber-sand composite (p ) may decrease before inforced sands: Experiments and modelling.” Geotext. Geomembr.
increasing. The undrained behavior of a fiber-reinforced sand is 28 (3): 238–250. https://doi.org/10.1016/j.geotexmem.2009.09.010.
more like that of an unreinforced dense sand. Diambra, A., A. R. Russell, E. Ibraim, and D. Muir Wood. 2007. “Deter-
• The conventionally defined pore pressure ratio ru ¼ u=pc ap- mination of fiber orientation distribution in reinforced sands.” Géotech-
nique 57 (7): 623–628. https://doi.org/10.1680/geot.2007.57.7.623.
proaches unity in reinforced samples even though liquefaction
Eskisar, T., E. Karakan, and S. Altun. 2016. “Effects of fiber reinforcement
and the development of a fluidized state did not occur. This in-
on liquefaction behaviour of poorly graded sands.” Procedia Eng.
dicates ru ¼ 1 is not suited to signify liquefaction when reinfor- 161 (Jan): 538–542. https://doi.org/10.1016/j.proeng.2016.08.688.
cement is present. This is due to there being an intricate stress Falorca, I. M. C. F. G., and M. I. M. Pinto. 2011. “Effect of short, randomly
equilibrium in a fiber-reinforced sand, with the externally ap- distributed polypropylene microfibers on shear strength behaviour of
plied load being distributed across and shared by the sand ske- soils.” Geosynth. Int. 18 (1): 2–11. https://doi.org/10.1680/gein.2011
leton, pore water pressure, and fibers. The effective stress of the .18.1.2.
sand skeleton may remain well above zero when ru ¼ 1. The
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

Gray, D. H., and T. Al-Refeai. 1986. “Behavior of fabric-versus fiber-


distribution of the externally applied stresses across the fibers reinforced sand.” J. Geotech. Eng. 112 (8): 804–820. https://doi.org/10
and the sand skeleton may be quantified, invoking a suitable .1061/(ASCE)0733-9410(1986)112:8(804).
constitutive modeling framework based on the rule of mixtures. Gray, D. H., and H. Ohashi. 1983. “Mechanics of fiber reinforcement in
The fiber stresses in reinforced samples under undrained com- sand.” J. Geotech. Eng. 109 (3): 335–353. https://doi.org/10.1061
pression may be calculated. /(ASCE)0733-9410(1983)109:3(335).
• Alternate pore water pressure ratios ru , ru;r , and ru;a may be Hazirbaba, K., and H. Gullu. 2010. “California bearing ratio improvement
and freeze–thaw performance of fine-grained soils treated with geofiber
defined, taking into account the stress distributions and their di-
and synthetic fluid.” Cold Reg. Sci. Technol. 63 (1–2): 50–60. https://
rection dependence. Liquefaction occurs when any of ru , ru;r , or doi.org/10.1016/j.coldregions.2010.05.006.
ru;a become equal to unity, representing a sand skeleton’s mean, Ibraim, E., A. Diambra, A. R. Russell, and D. M. Wood. 2012. “Assessment
radial, or axial effective stress becoming zero, respectively. In all of laboratory sample preparation for fiber reinforced sands.” Geotext.
undrained triaxial compression tests the maximum values of ru , Geomembr. 34 (Oct): 69–79. https://doi.org/10.1016/j.geotexmem
ru;r , and ru;a are always below unity. In fact ru , ru;r , and ru;a .2012.03.002.
move further and further below unity as deformations become Ibraim, E., A. Diambra, D. M. Wood, and A. R. Russell. 2010. “Static lique-
large, indicating the samples move further and further away faction of fiber reinforced sand under monotonic loading.” Geotext. Geo-
from a zero effective stress condition. membr. 28 (4): 374–385. https://doi.org/10.1016/j.geotexmem.2009
• Not all fibers are equally effective at suppressing liquefaction. .12.001.
An alternate fiber influenced the undrained response of loose Ishihara, K. 1993. “Liquefaction and flow failure during earthquakes.”
sand only slightly. Thus, ru and ru;r approached almost unity Géotechnique 43 (3): 351–451. https://doi.org/10.1680/geot.1993.43
and the load carrying capacity diminished at large shear strains. .3.351.
Jewell, R. A., and C. P. Wroth. 1987. “Direct shear tests on reinforced
sand.” Géotechnique 37 (1): 53–68. https://doi.org/10.1680/geot.1987
.37.1.53.
Acknowledgments
Kaniraj, S. R., and V. G. Havanagi. 2001. “Behavior of cement-stabilized
fiber-reinforced fly ash-soil mixtures.” J. Geotech. Geoenviron. Eng.
Financial support of this research was provided by the University of
127 (7): 574–584. https://doi.org/10.1061/(ASCE)1090-0241(2001)
New South Wales in Australia. This study was also supported by
127:7(574).
China Scholarship Council. Krishnaswamy, N. R., and N. T. Isaac. 1994. “Liquefaction potential of
reinforced sand.” Geotext. Geomembr. 13 (1): 23–41. https://doi.org/10
.1016/0266-1144(94)90055-8.
References Krishnaswamy, N. R., and N. T. Isaac. 1995. “Liquefaction analysis of sa-
turated reinforced granular soils.” J. Geotech. Eng. 121 (9): 645–651.
Ahmad, F., F. Bateni, and M. Azmi. 2010. “Performance evaluation of silty https://doi.org/10.1061/(ASCE)0733-9410(1995)121:9(645).
sand reinforced with fibers.” Geotext. Geomembr. 28 (1): 93–99. https://
Liu, J., G. Wang, T. Kamai, F. Zhang, J. Yang, and B. Shi. 2011. “Static
doi.org/10.1016/j.geotexmem.2009.09.017.
liquefaction behavior of saturated fiber-reinforced sand in undrained
Baki, M. A., M. M. Rahman, S. R. Lo, and C. T. Gnanendran. 2012. “Link-
ring-shear tests.” Geotext. Geomembr. 29 (5): 462–471. https://doi
age between static and cyclic liquefaction of loose sand with a range of
.org/10.1016/j.geotexmem.2011.03.002.
fines contents.” Can. Geotech. J. 49 (8): 891–906. https://doi.org/10
Maheshwari, B. K., H. P. Singh, and S. Saran. 2012. “Effects of reinforce-
.1139/t2012-045.
ment on liquefaction resistance of Solani sand.” J. Geotech. Geoen-
Boominathan, A., and S. Hari. 2002. “Liquefaction strength of fly ash rein-
forced with randomly distributed fibers.” Soil Dyn. Earthquake Eng. viron. Eng. 138 (7): 831–840. https://doi.org/10.1061/(ASCE)GT
22 (9–12): 1027–1033. https://doi.org/10.1016/S0267-7261(02)00127-6. .1943-5606.0000645.
Consoli, N. C., L. Festugato, and K. S. Heineck. 2009. “Strain-hardening Michalowski, R. L. 1997. “Limit stress for granular composites reinforced
behaviour of fiber-reinforced sand in view of filament geometry.” Geo- with continuous filaments.” J. Eng. Mech. 123 (8): 852–859. https://doi
synth. Int. 16 (2): 109–115. https://doi.org/10.1680/gein.2009.16.2.109. .org/10.1061/(ASCE)0733-9399(1997)123:8(852).
Consoli, N. C., P. D. Prietto, and L. A. Ulbrich. 1998. “Influence of fiber Michalowski, R. L., and J. Čermák. 2002. “Strength anisotropy of fiber-
and cement addition on behavior of sandy soil.” J. Geotech. Geoen- reinforced sand.” Comput. Geotech. 29 (4): 279–299. https://doi.org/10
viron. Eng. 124 (12): 1211–1214. https://doi.org/10.1061/(ASCE)1090 .1016/S0266-352X(01)00032-5.
-0241(1998)124:12(1211). Michalowski, R. L., and J. Čermák. 2003. “Triaxial compression of sand
Diambra, A., and E. Ibraim. 2015. “Fiber-reinforced sand: Interaction at the reinforced with fibers.” J. Geotech. Geoenviron. Eng. 129 (2): 125–136.
fiber and grain scale.” Géotechnique 65 (4): 296–308. https://doi.org/10 https://doi.org/10.1061/(ASCE)1090-0241(2003)129:2(125).
.1680/geot.14.P.206. Michalowski, R. L., and A. Zhao. 1996. “Failure of fiber-reinforced granu-
Diambra, A., E. Ibraim, A. R. Russell, and D. M. Wood. 2013. “Fiber lar soils.” J. Geotech. Eng. 122 (3): 226–234. https://doi.org/10.1061
reinforced sands: From experiments to modelling and beyond.” Int. /(ASCE)0733-9410(1996)122:3(226).
J. Numer. Anal. Methods Geomech. 37 (15): 2427–2455. https://doi Noorzad, R., and P. F. Amini. 2014. “Liquefaction resistance of Babolsar
.org/10.1002/nag.2142. sand reinforced with randomly distributed fibers under cyclic loading.”

© ASCE 04019125-13 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125


Soil Dyn. Earthquake Eng. 66 (Nov): 281–292. https://doi.org/10.1016 Tokimatsu, K., and Y. Yoshimi. 1983. “Empirical correlation of soil lique-
/j.soildyn.2014.07.011. faction based on SPT N-value and fines content.” Soils Found. 23 (4):
Park, T., and S. A. Tan. 2005. “Enhanced performance of reinforced 56–74. https://doi.org/10.3208/sandf1972.23.4_56.
soil walls by the inclusion of short fiber.” Geotext. Geomembr. 23 (4): Unnikrishnan, N., K. Rajagopal, and N. R. Krishnaswamy. 2002. “Behav-
348–361. https://doi.org/10.1016/j.geotexmem.2004.12.002. iour of reinforced clay under monotonic and cyclic loading.” Geotext.
Robertson, P. K., and R. G. Campanella. 1985. “Liquefaction potential of Geomembr. 20 (2): 117–133. https://doi.org/10.1016/S0266-1144(02)
sands using the CPT.” J. Geotech. Eng. 111 (3): 384–403. https://doi 00003-1.
.org/10.1061/(ASCE)0733-9410(1985)111:3(384). Vaid, Y. P., and S. Sivathayalan. 2000. “Fundamental factors affecting lique-
Russell, A. R., M. Chapman, S. H. Teh, and T. Wiedmann. 2017. “Cost and faction susceptibility of sands.” Can. Geotech. J. 37 (3): 592–606. https://
embodied carbon reductions in cutter soil mix walls through fiber doi.org/10.1139/t00-040.
reinforcement.” Geosynth. Int. 24 (3): 280–292. https://doi.org/10 Wang, K., and A. J. Brennan. 2013. “Dynamic response of saturated fiber-
.1680/jgein.17.00001. reinforced sand.” In Proc., 2013 SECED Young Eng Conf., 113–118.
Santoni, R. L., and S. L. Webster. 2001. “Airfields and roads construction Dundee, Scotland: Univ. of Dundee.
using fiber stabilization of sands.” J. Transp. Eng. 127 (2): 96–104. Wang, K., and A. J. Brennan. 2015. “Centrifuge modelling of fiber-
Downloaded from ascelibrary.org by Delhi Technological University on 02/27/20. Copyright ASCE. For personal use only; all rights reserved.

https://doi.org/10.1061/(ASCE)0733-947X(2001)127:2(96). reinforcement using as a liquefaction countermeasure of quay wall


Seed, H. B., I. M. Idriss, and I. Arango. 1983. “Evaluation of liquefaction backfill.” In Proc., 6th Int. Conf. on Earthquake Geotechnical Engi-
potential using field performance data.” J. Geotech. Eng. 109 (3): 458– neering. Dundee, Scotland: Univ. of Dundee.
482. https://doi.org/10.1061/(ASCE)0733-9410(1983)109:3(458). Wood, D. M., A. Diambra, and E. Ibraim. 2016. “Fibers and soils: A route
Shukla, S. K. 2017. Fundamentals of fiber-reinforced soil engineering: towards modelling of root-soil systems.” Soils. Found. 56 (5): 765–778.
Application of fiber-reinfoced soil. New York: Springer. https://doi.org/10.1016/j.sandf.2016.08.003.
Sivathayalan, S., and D. Ha. 2011. “Effect of static shear stress on the cyclic Yang, J. 2002. “Non-uniqueness of flow liquefaction line for loose sand.”
resistance of sands in simple shear loading.” Can. Geotech. J. 48 (10): Géotechnique 52 (10): 757–760. https://doi.org/10.1680/geot.2002.52
1471–1484. https://doi.org/10.1139/t11-056. .10.757.
Tingle, J. S., R. L. Santoni, and S. L. Webster. 2002. “Full-scale field tests Zornberg, J. G., and E. Kavazanjian. 2002. “Prediction of the performance
of discrete fiber-reinforced sand.” J. Transp. Eng. 128 (1): 9–16. https:// of a geogrid-reinforced slope founded on solid waste.” Soils Found.
doi.org/10.1061/(ASCE)0733-947X(2002)128:1(9). 41 (6): 1–16.

© ASCE 04019125-14 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2020, 146(1): 04019125

You might also like