Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Arab J Sci Eng (2016) 41:1183–1198

DOI 10.1007/s13369-016-2099-1

REVIEW ARTICLE – CIVIL ENGINEERING

Fiber–Matrix Interactions in Fiber-Reinforced Concrete:


A Review
Yassir M. Abbas1 · M. Iqbal Khan1

Received: 1 September 2015 / Accepted: 1 March 2016 / Published online: 30 March 2016
© King Fahd University of Petroleum & Minerals 2016

Abstract A significant breakthrough in concrete technol- List of symbols


ogy was achieved using fibers to reinforce concrete. Vari- Am , Af , Ab The cross-sectional area of the matrix,
ous researchers have reported that fiber reinforcement can the fiber and a concrete member, respectively
alter the brittleness of concrete. The efficiency of fiber re- Ec, Ef , Em Moduli of elasticity for the composite,
inforcement is based on the fiber–matrix interactions. The fiber and matrix, respectively
understanding of these interactions is a challenging engi- K IC Plane strain fracture toughness
neering problem, where the frictional bond governs and the Lc Critical length of the fiber
physical/chemical bond plays a minor role. This problem is Lp The maximum length of the fiber at
extremely sophisticated because of the following nonlinear which instantaneous pullout of fiber occurs
interactions: interfacial debonding, plastic material deforma- Lt The minimum length of the fiber at
tions, mechanical bond deformations, and frictional sliding. which a tensile failure occurs without debond-
This paper reports a comprehensive and up-to-date litera- ing
ture review on the fiber–matrix interactions, and physical P Shear force along the fiber–matrix interface
and theoretical modeling of the fiber–matrix interactions is Pt the applied pullout load
reported in detail. In addition, the most important conclu- k A constant correlates the relation between
sions of the parametric studies of the fiber–matrix interfacial shear stress on the fiber–matrix interface
bond are summarized. The information of the pullout test and the corresponding relative slippage
standardization to assess the fiber–matrix behavior of a fiber- le The embedded length of fibers
reinforced concrete is reviewed. The current research in the n, df Number and diameter of the fibers, respec-
area of fiber–matrix interactions of fiber-reinforced concrete tively
are discussed. s Slippage of the fiber due to pullout force
u Length of the debonded zone
Keywords Fiber-reinforced concrete · Fiber–matrix inter- x Variable distance parallel to the fiber length
actions · Pullout tests · Physical modeling · Standardization · εu The ultimate strain of the matrix.
Theoretical modeling εe The strain of the matrix
corresponding to the first crack stress.
τo The limiting interfacial bond strength
neglecting the Poisson’s effect on fibers
τfd the decreasing frictional bond stress
corresponding to the end slip after full
B Yassir M. Abbas debonding
yabbas@ksu.edu.sa
η Coefficient based of the shape of the load–slip
1 Department of Civil Engineering, College of Engineering, curve. It can be assumed to be equal to 0.2 for
King Saud University, P. O. Box 800, 11421 Riyadh, straight fibers
Kingdom of Saudi Arabia σf Tensile strength of fiber

123
1184 Arab J Sci Eng (2016) 41:1183–1198

γm The work of fracture of the matrix understood [2,3]. Naaman [4] reported that the strength of the
τ Frictional shear stress at fiber–matrix interface matrix controls the cracking strength and the fiber–matrix in-
τ Shear strength for fiber–matrix interface terfacial bond dominates the post-cracking strength. It is well
ψ The spacing between fibers agreed that the inclusion of fibers in a brittle matrix consider-
υf , υm The Poisson’s ratio ably improves its post-cracking performance; however, Gray
for the fiber and the matrix, respectively and Johnson [5] reported that the modulus of rupture (tensile
δ Fiber orientation factor strength and first crack) and the associated absorbed energy
, o Fiber displacement response of after of a fiber-reinforced concrete can be dramatically improved
complete pullout and at the end of complete by increasing the fiber–matrix bond strength. Moreover, these
pullout researchers observed that the ultimate tensile strength, strain
ζ Damage coefficient taking a value between 0 capacity and the energy absorbed of a FRC are improved by
and 1 increasing the shear strength of the fiber–matrix interface.
Generally, the fiber–matrix interactions (Fig. 2) govern
the mechanical properties of FRC [5–9]. These properties
include (1) compressive, bending and tensile strengths, (2)
1 Introduction
modulus of elasticity, (3) ultimate strain fracture toughness,
Concrete cracking arises at notably low tensile strains be- (4) impact and seismic resistance, (5) ductility and (6) dura-
cause it has a notably low tensile strength (approximately 7– bility [5,7,9–11]. It has been proved that FRC with superior
11 % of its compressive strength) [1]. Thus, concrete is clas- mechanical properties could not be developed unless the
sified as “brittle or fragile material”. The typical stress–strain fiber–matrix interfacial bond was at least equal to the tensile
relationship for a stressed FRC element (Fig. 1) demonstrates strength of the matrix [4]. However, notably strong fiber–
that it has two different behaviors pre- and post-initiation of matrix bond does not guarantee the production of ductile
cracks. As shown in Fig. 1, the cracking and post-cracking FRC, which indicates the importance of the properties of the
strength of a FRC are characterized based on these behav- matrix and fiber. The tensile strength of the fiber, microstruc-
iors. It is often thought that this unfavorable behavior of ture of the matrix and the length/alignment with respect to
concrete in the post-crack stage can be altered by incor- the applied stress/content of the fibers are important to the
porating discontinuous fibrous reinforcement (fibers) in the fiber capability to mitigate cracking [7].
concrete to formulate fiber-reinforced concrete (FRC). How- In the past, direct and indirect experimental methods were
ever, the benefits of fibers in the pre-crack stage are not well developed to quantify the fiber–matrix bond of FRC [12,13].

Fig. 1 Mechanics and


Micromechanics of FRC beam,
a Four-point flexure test, b
stressed element, c stress–strain
of stressed element

(a)

Stress (b)

Cracking

Post-cracking

Strain
Crack opening
(c)

123
Arab J Sci Eng (2016) 41:1183–1198 1185

2 Fiber–Matrix Interfacial Bond


Debonding
The mechanics of the fiber–matrix interfacial bond are ex-
Bridging
tremely sophisticated as a result of different linear and non-
Pullout linear modes of interaction: physical/chemical bond, inter-
facial debonding, plastic material deformations, mechanical
Failure bond deformations, and frictional sliding. The fiber bridging–
debonding–pullout (Fig. 2) is generally considered the most
Fig. 2 Mechanisms of fiber–matrix interactions [13] possible failure mode of FRC [17].
The pullout behavior of straight fibers is classified into
two modes of fracture (Fig. 4): (1) elastic physical/chemical
adhesion between the fiber and the matrix; (2) frictional
Load sliding [18,19]. However, the pullout behavior of deformed
(hooked-end) fibers is characterized into five modes of frac-
Matrix ture (Fig. 4): (1) elastic bond; (2) fiber–matrix debonding;
FRC
(3) plastic deformation of the hook; (4) Coulomb friction on
the hook; (5) fiber–matrix frictional bond [20]. As observed
in Fig. 4, the hooked-end fiber significantly enhances the
fiber–matrix pullout toughness because of the slip-hardening
Fiber behavior caused by the mechanical bond of the hook. How-
ever, Wille and Naaman [21] reported that the fiber–matrix
behavior of brass-coated straight fibers embedded in ultra-
Fig. 3 Physical modeling of debonding and pullout of fibers by fiber high-performance concrete (UHPC) also exhibited the slip-
pullout test [13] hardening behavior as hooked fiber embedded in normal
concrete. This behavior was attributed to the scratching of
the fiber surface using micro-sand, which increases the fric-
tional bond of the fiber–matrix interface in the post-cracking
In direct methods, the fiber–matrix bond behavior is assessed stage.
by measuring the uniaxial tensile pullout load and the corre-
sponding slippage during fiber pullout from a cementitious
2.1 Parameters Influencing the Fiber–Matrix Bond
matrix [5,7,13–15]. This test can provide important infor-
mation on the fiber–matrix bond of FRC because it simulates
The bonding strength of the fiber–matrix interface of a FRC is
the debonding and pullout of fibers (Fig. 3). Because of the
influenced by the characteristics of the fiber, matrix and ITZ.
nonlinearities associated with the fiber–matrix shear stress
Table 1 gives the general parameters influencing the fiber–
distribution along the fiber length, particularly for deformed
matrix interfacial bond. So far, numerous parametric studies
fibers, many researchers adopted the fiber pullout test to as-
sess the efficiency and surface treatment of fibers. The “crack
mouth opening width” is defined as the total slip of the fiber
including elongation and straightening effects. In the indirect
Hooked fiber
methods, the fiber–matrix bond strength is assessed from the Mode of fracture
mechanical property (mostly the flexural strength) of FRC I Elastic
[16]. Straight fiber II Mechanical
Pullout force

The subject of the fiber–matrix bond strength of FRC re- III Frictional
A
ceives great interest from the researchers of modern concrete I
technology. However, the reported study on this topic is scat-
tered in the literature. The objectives of this study are (1)
to comprehensively review the literature on the physical and
III
theoretical modeling of the fiber–matrix interactions of FRC,
(2) to provide information on the potential of standardiz- II
ing the pullout test to assess the fiber–matrix behavior of
FRC and (3) to address the gaps in the literature of fiber– End slip
matrix interactions of FRC for possible future research pro-
grams. Fig. 4 Pullout modes of fracture for steel fibers [53]

123
1186 Arab J Sci Eng (2016) 41:1183–1198

Table 1 Properties affecting the fiber–matrix bond [22] of deformed fibers (crimped and rounded-end fibers) de-
Fiber ITZ Matrix creases with increasing the alignment angle of fiber. In ad-
dition, these researchers concluded that, unless the orien-
Strength Microstructure Strength tation causes a tensile failure of the fiber, the increase in
Geometry Porosity Composition the fiber orientation increases the pullout toughness with
Stiffness Packing density Stiffness maximum toughness occurred at 10◦ –20◦ fiber orientation
Poisson’s ratio Local stiffness Poisson’s ratio angles. Up to 50 % increase in the ultimate pullout load
Embedded length Local strength Age of straight fibers can be achieved by optimization of the
Orientation Processing Curing condition fiber orientation. Grunewald [20] demonstrated that the ul-
timate pullout load is independent of the fiber orientation
for hooked fibers. Due to the combined effect of snubbing
on the fiber–matrix pullout behavior have been reported. In “the increase in bridging effect of the fiber due to its orien-
these studies, the most investigated parameters are: (1) the tation” and spalling of the matrix, Lee et al. [10] considered
type and geometry of fibers, (2) alignment (orientation) of the optimum range of fiber orientation is 30◦ –45◦ , which
the fiber, (3) surface treatment of fibers, (4) strength and achieve the maximum bridging performance. Most of the re-
composite of the matrix, and (5) loading rate. However, few searchers agree that the probability of fiber tensile failure or
studies have taken the embedment length and size of the fiber matrix spalling increases with an increase in fiber orientation
in consideration. Based on the reported research, it seems [20].
that there is no available information on the pullout behavior
of hooked fibers with embedment length less than 10 mm
[18]. The most important conclusions of these studies are 2.1.3 Strength and Composition of the Matrix
summarized in following sections.
Generally, increasing the strength of the matrix increases the
ultimate pullout load. In addition, the overall behavior of the
2.1.1 Type of the Fiber fiber–matrix interactions is enhanced as the strength of the
matrix increases, but not to the extent that makes the fiber
The most comprehensive study on the influence of fiber tensile failure possible. For hooked-end steel fibers aligned
type on the fiber–matrix bond was conducted by Chin and in the direction of loading, Robins et al. [18] found that an
Xiao [17]. In their study, the investigated fibers were: (1) increase in the strength of the matrix, also, increases the pull-
the flat-crimped, (2) the flat-end, (3) the hooked-end, (4) out toughness and ultimate response. However, the rate of the
the round-crimped, and (5) double-anchored of steel fibers. increase decreases as the fiber orientation angle increases.
These researchers reported that the flat-crimped fiber had Grunewald [20] has pointed out that increasing strength of
exhibited the highest pullout load (about 1.2 kN for fiber the matrix increases the probability of fiber tensile failure
having 50 mm length and 1.14 mm equivalent diameter). and/or matrix spalling. Abu-Lebdeh et al. [25,26] found that
However, the highest pullout toughness (10.8 Nm for fiber for only deformed fibers, increasing the strength of the matrix
having 50 mm length and 1.00 mm equivalent diameter) was increases the ultimate pullout load and toughness; however,
obtained for flat-end steel fiber. Chin and Xiao [17] concluded the pullout behavior of some of the straight fibers is not af-
that making the fiber deformations greater in size, amount, fected by the strength of the matrix. In addition, they found
intensity, or degree increases the ultimate pullout load and that the increase in the pullout toughness in very high strength
decreases the pullout toughness, and vice versa. Richardson concrete is more pronounced than the increase in the ultimate
and Heather [23] had compared the fiber–matrix bond of load.
hooked steel fibers and 3D steel fibers. It was concluded Tuyan and Yazici [27] investigated the influence of mix
that the bond of fiber matrix interface to fiber strength of 3D proportions of Slurry Infiltrated Fiber Concrete (SIFCON)
fibers is more than three times that for hooked steel fibers. on the fiber–matrix bond utilizing the single-fiber pullout
However, Richardson and Heather [23] had used hooked test. They reported that increasing the strength of SIFCON
fiber with tensile strength more than that for 3D fibers. matrix increases fiber–matrix bond. Correlation between the
fiber–matrix bond and strength of the matrix was observed
2.1.2 Alignment of the Fiber for SIFCON under standard water and steam curing; how-
ever, this was not evident in autoclave curing. The effect
Naaman [24] demonstrated that increasing the alignment of age of standard water curing on the interfacial bond of
of the fiber with respect to the loading direction increases the fiber to the matrix was studied by Singh et al. [28]. It
both of the fiber pullout load and response at the ultimate was found that the maximum fiber–matrix interfacial bond
load. However, Robins et al. [18] found that the toughness is reached after 2-day curing (Fig. 5). Although water curing

123
Arab J Sci Eng (2016) 41:1183–1198 1187

30

Embedded length of the fiber


25

Frictional fiber-matrix bond


Fiber-matrix bond (MPa)

20

Model-1
15

10

5
0 1 2 3 4 5 6 7
Age of curing (days) End slip
Fig. 5 Effect of the age of curing of the fiber–matrix bond [37] Fig. 6 The fiber–matrix frictional bond models [21]

more than 2 days enhances the strength of the matrix, Singh increase in the fiber pullout load. However, Shannag et al.
et al. [28] reported that it does not improve the interfacial [8] reported that increasing the embedment depth of straight
bond of the fiber to the matrix. Recently, Wille and Naaman steel fibers could increase the ultimate pullout load and pull-
[21] reported that increasing the strength of the matrix is im- out toughness by 200–300 %. Singh et al. [28] observed that
portant for increasing the fiber–matrix bond; however, it does the performance of fiber–matrix interface during fiber pullout
not guarantee the development of very high fiber to matrix changes with increasing the fiber embedment length. How-
bond. ever, for polypropylene fibers, constant average fiber–matrix
Tuyan and Yazici [27] investigated the influence of dif- bond for different embedment lengths was observed. Tuyan
ferent curing conditions (standard, steam and autoclave) of and Yazici [27] had investigated the fiber–matrix of SIFCON
SIFCON on the fiber–matrix bond. It was reported that, gen- with different embedment lengths. It was reported that in-
erally, the fiber–matrix interfacial bond enhances as the cur- creasing the embedment length of a fiber from 10 to 30 mm
ing conditions improve. The highest fiber to matrix bond in SIFCON matrix increases the ultimate pullout load. Also,
was achieved for autoclave curing, which changes the fiber Ali et al. [31] found that increasing the embedded length
mode to tensile failure. However, 8 h of steam curing achieved of a coconut fibers increases the ultimate pullout load and
fiber–matrix bond as that obtained for 28-day standard water toughness.
curing. Barluenga [29] conducted experimental and a nu- However, Boshoff et al. [32] found that fiber–matrix fric-
merical study on the fiber–matrix interactions at early age of tional bond decreases as the fiber embedment length increa-
FRC made by AR-glass fibers. He concluded that the stress ses, not as expected. Thus, it was concluded that the micro-
distribution along the fiber depends on the development of mechanical model assuming a uniform interfacial bond along
the strength of concrete and size, type and material of the the length (Model-1 in Fig. 6) of the fiber is not perfect. This
fiber. was attributed to the Poisson’s effects on the fiber. In addition,
Kim et al. [30] studied the influence of the fine aggre- these researchers have revealed that increasing the fiber em-
gate to coarse aggregate ratio on the fiber–matrix bond for bedment length increases the probability of the tensile model
straight, hooked and twisted fibers. It was found that only of failure for a fiber.
twisted fibers, increasing the S/a ratio from 0.444 to 0.615 in-
creases fiber–matrix interfacial bond. However, the S/a ratio 2.1.5 Rate of Loading
has no influence on fiber–matrix bond for straight and hooked
fibers. This was attributed to the enhancement of microstruc- Boshoff et al. [33] conducted a comparison study on the fiber–
ture of the matrix and fiber–matrix mechanical interactions matrix interfacial bond under rates of utilizing the pullout
for twisted fibers. It is worth mentioning that the increase test. As expected, they found that the fiber–matrix inter-
in fine aggregate to course aggregate ratio was achieved by facial bond depends on the rate of the loading, and it in-
decreasing the content of the coarse aggregate. creased as the rate of the loading increases. However, at high
rate of loading the possibility of fiber tensile failure is high.
2.1.4 Length of the Fiber Also in this regard Abu-Lebdeh et al. [25] have confirmed
that the ultimate pullout load and toughness are sensitive
Robins et al. [18] demonstrated that increasing the length of to the rate of loading. However, due to geometry of fibers,
hook steel fiber from 12 to 25 mm caused a limited range this sensitivity for deformed fibers is different from that for

123
1188 Arab J Sci Eng (2016) 41:1183–1198

straight fibers. For deformed fibers, the maximum pullout 2.2.1 Macroscopic Altering of Fiber Surface
load and toughness increase with an increase in the loading
rate. Naaman [4] developed a new generation of twisted steel
fibers, known as “Torex,” with an enhanced fiber–matrix in-
terfacial bond. Because of its slip-hardening behavior, these
2.1.6 Other Parameters fibers are appropriate for the development of FRC with supe-
rior ductility properties. To enhance the fiber–matrix interfa-
Chan and Chu [34] investigated the effect of the content cial bond, Singh et al. [28] proposed a mechanical indentation
of the silica fume on the fiber–matrix interfacial bond of treatment for polypropylene. This method involves applying
UHPC. It was reported that up to 30 % content of silica different pressures on the surface of the fiber by two rough
fume, increasing the content of silica fume does not affect steel plates. These researchers revealed that the 700 kPa pres-
the interfacial bond of the fiber to the matrix. However, due sure is the optimum since it creates up to 300 % increase
to the decrease in the interfacial-toughening property, the in the bond of fiber–matrix without tensile failure of the
fiber–matrix bond decreases when the silica fume content is fiber.
more than 30 %. These researchers concluded that the 20–
30 % is the optimal content of silica fume for producing
UHPC.
Singh et al. [28] studied the influence of seawater and
salt environments on the fiber–matrix interfacial bond. It 2.2.2 Microscopic Altering of Fiber Surface
was found that a concrete kept in salt water and seawater
for 6 months has more the fiber–matrix bond than that kept Brass coating on the surface of steel fibers significantly en-
in plain water. In addition, the maximum pullout load oc- hances the interfacial bond of the fiber to the matrix and thus
curs for the concrete kept in the seawater. The improvement enhances the overall performance of steel FRC. However,
in the concrete exposed to the seawater was attributed to its it was reported that the use of oxidizing or degreasing en-
improved microstructural behavior as compared to salt envi- hances the interfacial bond of the fiber to the matrix, but it
ronment exposure. has no influence on the mechanical properties of FRC. Most
Regarding the content of fiber in a FRC, Shannag et al. of the mechanical treatments of fibers (hooked or enlarged
[8] demonstrated that increasing the volume content from 3 end, twisting) have positive effect on the fiber matrix interfa-
to 6 % enhances the pullout behavior of steel fiber, namely cial bond [5]. Wu and Li [35] showed that up to 600–700 %
the ultimate pullout load and toughness increased by about enhancement in the fiber–matrix interfacial bond strength
20 %. can be attained by plasma treating for polypropylene fibers.
However, it was reported that the inclusion of PVA pow-
der or high alumina cement in the cementitious has a little
2.2 Enhancement of Fiber–Matrix Interfacial Bond influence on the fiber–matrix interfacial bond. So far little
information on the combined effect of the combined enhance-
Various techniques for enhancing the fiber–matrix interfa- ment techniques is available. Sedan et al. [37] investigated
cial bond have been proposed [4,28,31,35,36]. These tech- the mechanical properties of a FRC made with hemp fibers.
niques can be classified into: (1) macroscopic altering of An alkaline treatment of this fibers increases the elastic (ad-
the mechanical properties and fibers (fibrillation, twisting, hesion) fiber–matrix interfacial bond, which increases the
crimping, hooked of enlarged end), (2) microscopic alter- flexural strength of the FRC by 94 % compared to the same
ing of the fiber surface properties (mechanical or chemical made with un-treated fibers.
treatment) and (3) microscopic enhancement of the matrix Sebaibi et al. [36] studied the influence of treating fibers
by: (a) using microfibers (hybridization) or (b) densification 0.5, 0.75 and 1 % of vinyl trimethoxy silane (VTMO) on
of the transition zone. All of these techniques can result in the fiber–matrix bond of UHPC. They found that the VTMO
significant enhancement in the fiber–matrix bond. However, treatment of fiber significantly enhances the fiber–matrix in-
the transition zone densification method is appropriate for terfacial bond with an optimal content of 0.5 %. Up to 20 MPa
metallic fibers. So far, most of the enhancement methods fiber matrix bond of straight fiber embedded in UHPC was
have been made at the microscale levels of the constituent developed by Wille and Naaman [21] by using zircon sand
materials of FRC. However, Li and Stang [9] reported that with high hardness, which enhances the dispersion of mi-
the macroscopic notching of fibers had significantly enhance croparticles in the matrix. Ali et al. [31] had demonstrated
the fiber–matrix interfacial bond, while the microscopic abra- that boiling coconut fibers increases the fiber–matrix bond by
sion texturing of the fiber surface has no influence on the 184 %; however, the chemical treatment of the same fibers
fiber–matrix interfacial bond. decreases the fiber–matrix bond by 25 %.

123
Arab J Sci Eng (2016) 41:1183–1198 1189

Table 2 Types of fiber matrix


Effects Significance
interfacial bond [2]
Tensile Provide resistance to stresses Has great influence on the tensile
perpendicular to fiber–matrix strength of FRC
interface
Shear
Elastic Facilitate two-way fiber–matrix Almost all the mechanical
bridge for transferring stresses, if properties of FRC are influence
the applied shearing stresses on by the shearing bond strength of
the fiber matrix interface being the fiber–matrix interface
more than the matrix capability
to resist tensile stresses.
Frictional Resists the fiber slip when the
applied stresses being more than
the elastic bond strength of the
fiber–matrix interface

2.2.3 Microscopic Enhancement of the Matrix the frictional shear bond strength are (1) coefficient of fric-
tion of the fiber–matrix interface, (2) normal stresses in the
A microstructural investigation on fiber–matrix fracture of matrix and (3) Poisson’s effects on the fiber–matrix inter-
steel and polypropylene fibers revealed that the failure oc- face [2,16]. In the following sections, the conducted efforts
curred at the fiber–matrix interface; however, for brass fibers, in the last 50 years to theoretically and physically model the
the failure occurred at the transition zone (which has a width mechanics of fiber–matrix interactions of FRC are summa-
of 60–80 µm for the fiber–matrix interface) [7]. Thus, the rized.
densification of transition zone is not appropriate for enhanc-
ing the fiber–matrix bond strength for steel and polypropy-
lene fibers. However, a research conducted by Shannag et al. 3.1 Theoretical Modeling
[8] concluded that up to 300 % enhancement of the fiber–
matrix interfacial bond for steel fiber can be obtained by Various researchers have proposed theoretical models to sim-
densifying the ITZ. Mu et al. [38] succeeded to increase ulate the fiber–matrix interactions [2,10,14,40,41]. Gener-
cost/toughness ratio of polypropylene FRC compared to AR- ally, the following procedure is followed: (1) assume an
glass FRC by improving its fiber–matrix bond. This was elastic fiber–matrix bond until the initiation of the first crack,
achieved by enhancing the strength of the matrix and mini- (2) assume a notably simple post-cracking model, and (3)
mizing the porosity of the ITZ. estimate the stress distribution along the fiber–matrix using
Markovich et al. [39] investigated the inclusion of 4–6 % semiempirical methods. At present, most available models
by volume of microfibers in the cementitious matrix for con- assume a constant fiber–matrix interfacial bond. In the re-
trolling the micro-cracks occurring around the macro-fiber ported research, only Beaumont and Aleszka [16] assumed a
during the fiber pullout test. It was reported found that the variable stress distribution along the fiber–matrix interface.
fiber hybridization improves up to 40 % the fiber–matrix Generally, these models differ only in the simulation of the
bond, especially for the matrix having low water–binder ra- fiber–matrix behavior at the post-cracking stage, which the
tio. researchers agree to be at the end of the linearity of the stress–
strain relation of the FRC. However, this conclusion is not
absolutely true because the fiber–matrix interface debonds
after the matrix fractures [2]. The simplest model (model 1
3 Mechanics of Fiber–Matrix Interactions in Fig. 6) for the fiber–matrix interactions assumes a uniform
frictional shear bond during the fiber pullout. This model was
Based on the fiber alignment with respect to the applied modified by Li et al. [40] to incorporate the slip-hardening
stresses, the fiber–matrix interfacial bond can be classified effects of deformed fibers (model 2 in Fig. 6). Figure 7 shows
into tensile bond and shear bond strength. The shear bond the simulated pullout–slip curves, which resulted from mod-
strength is characterized into elastic and frictional. In the lit- els 1 and 2.
erature, the information available on the tensile bond strength Assuming that the elastic properties of the fiber and matrix
of the fiber–matrix interface is limited. Table 2 provides a govern the shearing-stress distribution over the fiber surface,
general description of the effect and significance of each class Lawrence [41] modeled the shear force and stress distribution
of fiber–matrix interfacial bond. The properties that govern along the fiber using Eqs. (1)–(2) as shown in Figs. 8 and 9,

123
1190 Arab J Sci Eng (2016) 41:1183–1198

   √  
π kdf Pt cosh ax
τ = √ √  and
Embedded length of the fiber R a sinh ale
  
Simulation based π kdf Pt √ 
τmax = √ coth ale (2)
on Model-2 R a
Pullout load

where

π df k
a=
R
Simulation based on Am E m − Af E f
Model-1 R=
(Am E m ) (Af E f )

Crack opening width Based on Lawrence [41] model, it can be seen from Fig. 8
that the pullout force distribution is linear for small embed-
Fig. 7 The simulated fiber–matrix interactions [21]
ment length but nonlinear (taking the form of an exponential
curve) for long embedded fibers. In addition, it can be seen
1
from Fig. 8 that the shear stress distribution over the fiber
data1
Normalized fiber shear stress τ/τ max

surface significantly depends on its embedded length. In this


0.8 data2
model, the maximum shearing stress τmax occurs when the
data3 fiber enters the matrix. In addition, the fiber begins debonding
0.6 data4 (Fig. 2) when τmax is greater than the fiber–matrix bonding
strength. Generally, the rule of mixture (Eq. 3) governs the
0.4 elastic behavior of the FRC. Moreover, it is usually assumed
that the first crack occurs when the matrix strain is equal to
0.2 the ultimate strain (Eq. 4).

0
0 0.2 0.4 0.6 0.8 1
E c = E f Vf + E m (1 − Vf ) (3)
Normalized fiber embeddment depth x l e
 1
24τ γm E f Vf2 3
εm u = 2d V
(4)
Fig. 8 Distribution of the pullout force over the fiber surface [23] Ec Em f m

1.4 Assuming the fiber and the matrix are in full contact, the
Normalized fiber embeddment depth x l e

data1 first cracking stress σm can be estimated by Eq. (5).


1.2
data2
 1  1
1 4E m γm 2 2G m E c 4
σm = (5)
0.8 df (1 + α) ψ E f E m Vm

0.6 where
0.4
E m Vm
α=
0.2 E f Vf
π√ 1
2
0
0 0.2 0.4 0.6 0.8 1
ψ = ln 3Vf
2
Normalized fiber embeddment depth x l e
Equation (6) provides a semiempirical expression to evaluate
Fig. 9 Distribution of the shear stress over the fiber surface [23] the generated tensile stress over the fiber surface during the
pullout test. Assuming that the shearing stress on the fiber–
matrix interface is inversely proportional to the tensile stress
respectively. This model uses experimental data to establish on fiber (Eq. 6), the composite strength σc after the first crack
the fitting constant k. initiation is obtained using Eq. (7).
 √  
sinh ax  
P = Pt √  (1) τo 4k1le
sinh ale σf = 1 − exp − (6)
k1 df

123
Arab J Sci Eng (2016) 41:1183–1198 1191
 
τo df 2k1le df (a)
σc = Vf 1+ exp − − (7) Fiber
k1 2k1le df 2k1le

where

E m υf μ
k1 =
E f (1 + υm )

Assuming the tensile strength of FRC does not affect the Fiber
alignment of the fiber, it can be estimated by σt = σf × factor,
in which a factor less than one is encountered to take into
account the actual number of fibers resisting the formation
of cracks. The fibers having embedment length less than the
critical length (Eq. 8) will start pullout under decreasing load.

df σf
Lc = (8)
4τ (b)
Beaumont and Aleszka [16] developed Eq. (9) to find the Fig. 10 Nonlinear distribution of shearing stress over the fiber surface:
toughness of FRC beams γp , which represent the work done a at pre-cracking stage, b at post-cracking stage [2]
by pullout force that causes the frictional shear stress τ  at Table 3 Mode of failure based on the fiber embedment length [2,49]
fiber–matrix interface.
Case Mode of failure
Vf τ le2 le ≤ L p
γp = for le ≤ L c /2 The fiber–matrix elastic bond resists the fiber
12df pullout; however, prompt pullout occurs since
Vf σf3le very short embedment length does not allow
γp = for le > L c /2 (9) the propagation of frictional resistance along
12τ 2 the fiber–matrix interface
L p ≤ le ≤ L c Fiber pullout process begins when the pullout
To simulate the fiber–matrix interactions, it is generally stresses is more than the elastic bond on the
assumed a non-uniform elastic shear bond with frictional fiber–matrix interface, and then, the debonded
shear bond in the formulation of models. Different modes of zone with frictional resistance propagates, and
failure have been established: (1) no debonding, (2) instan- finally complete pullout occurs when the
applied pullout stresses approach the tensile
taneous debonding and (3) gradual debonding [16,42–44]. strength of fiber
However, to solve the same problem, Naaman et. al. [45] L c ≤ le ≤ L t Fiber pullout process begins when the pullout
proposed a uniform elastic shear bond with frictional shear stresses is more than the elastic bond on the
bond. The following parameters were incorporated in the the- fiber–matrix interface, and a fiber tensile failure
oretical model of fiber–matrix interactions: (1) orientation, occurs with a partial debonded when the
applied pullout stresses exceed the tensile
spacing and aspect ratio of the fibers; (2) continuous fibers strength of fiber
[43,46–48]. However, notably few models have incorporated le > L t Prompt tensile failure of the fiber occurs when
the radial deformations because of shrinkage and Poisson’s the fiber is in full contact with the matrix and
effects [16] and the effect of repetitive loading [46]. No re- the applied pullout stresses exceed the tensile
ported theoretical model appeared to consider the nonlinear strength of fiber
stress distribution along the fiber in the pre-cracking stage
(Fig. 10a).
The initiation of the first crack changes the role of the of fibers. Table 3 defines four modes of failure being char-
fiber–matrix interfacial bond. Before the first crack begins, acterized based on the critical embedment length, maximum
the ultimate shearing stress over the fiber surface is located fiber length L p . The prompt pullout of fiber occurs as the
at its edges (Fig. 10a). However, after the first crack, the embedded length of fiber is below the minimum fiber length
ultimate shearing stress is located at approximately at the L t ; however, the tensile failure occurs when the embedded
mid-length of the fiber (Fig. 10b). The pullout toughness of depth of fiber exceeds a critical value [2,49].
steel/cement systems is approximately 10–20 J/m2 , which is Wang et al. [14] demonstrated that the theoretical model
L c comparable to the tensile toughness of cement paste [9]. could be simplified by neglecting the elastic shear bond
Several modes of failure are defined in terms of the fiber– strength of the fiber–matrix interface τs . In addition, Shannag
matrix bond and the mechanical and geometrical properties et al. [8] found that the frictional bond of the fiber–matrix in-

123
1192 Arab J Sci Eng (2016) 41:1183–1198

Table 4 Theoretical mode


Type of fiber Aspect ratio τo a1 a2 Range of slip (mm)
constants for different fibers [9]
Nylon – 0.05 0.001 0.0002 50
Monofilament polypropylene – 0.1 0.0005 0.0003 50
Straight steel – 2.35 −0.35 0.014 12.7
Fibrillated polypropylene – 0.8 0 0 0.3
Straight steel 62.50 4.2 −4.0 1.0 0.3
2.9 −4.0 0 0.3
Straight steel 31.25 6.0 −4.0 1.0 0.3
Hooked steel 60.00 3.5 4.0 −10.0 0.3
3.5 4.0 −10.0 0.3
4.5 −2.0 0.0 0.3

terfacial significantly contributed to resisting shear stresses


along the interface during the pullout test. In addition, these Embedded length of the fiber
researchers showed that the frictional shear bond strength
of the fiber–matrix interface τf for metallic fibers generally

Fiber-matrix bond
decreased when the relative slippage of the fiber–matrix in-
terface increased. However, for synthetic fibers, τf increases τ max
when the relative slippage of the fiber–matrix interface in- τf
creases. Therefore, the frictional bond strength of the fiber–
matrix interface should not be considered constant. Wang et
al. [14] presented a nonlinear model of the bond–slip rela-
tions ship using Eq. (10). 1

τ = τo + a1 x + a2 x 2 (10) End slip

where the constants τo , a1 and a2 for different types of fibers Fig. 11 Fiber–matrix bond model proposed by Naaman et al. [50]
are provided in Table 4. Li and Stang [9] demonstrated that
the model of Wang et al. [14] (Eq. 10) efficiently predicted
the slip-hardening relationship behavior of two-sided fiber strength can be considered equal to the frictional bond of
pullout experiments. the fiber–matrix interface. In addition, a 3D finite-element
Naaman et al. [50] proposed the model in Fig. 11 to sim- model of the fiber–matrix interactions by Georgiadi et al. [3]
ulate the fiber–matrix bond during the pullout test. Based on revealed that the appropriate ratio of fiber–matrix frictional
this model, Lee et al. [10] derived Eqs. (12)–(14) to model the coefficient μ is 10 %, which made the best correlations with
pullout force–slip behavior for the elastic, partial debonding experimental results.
and fully debonded modes, respectively. It should be noted
that the experimental investigations of Lee et al. [10] showed   
λAm E m 1 + e−λle
that the pullout behavior of fibers embedded in UHPC did P= s (12)
Q−2 1 − e−λle
not exhibit the drastic decrease in fiber–matrix bond when
the fiber debonding began. Thus, the maximum elastic bond
where
 π df κ Am E m
λ= K Q; K = ; and Q = 1 +
Am E m Af E f
⎧ ⎡ ⎤⎫

⎨ πd τ  ⎬
f max ⎣ 1 − e−2λ(1−u) ⎦
P = π df τf u +     
⎩ λ 2 −λ(1−u) + 1 − 1 1 + e−2λ(1−u) ⎭
Q e Q
  
1−e−λ(1−u)
P (Q − 1) u − π df2τf u (Q − 2) + (P − π df τf u) 1−e
2 Q−2
−2λ(1−u) λ − π df τf ule
s= (13)
Am E m

123
Arab J Sci Eng (2016) 41:1183–1198 1193

P = π df τfd s (l − s + so ) (14)
  
−4υ
f μ(l−s+so ) 
η η
1 − exp 1+υf 1−υf
e−(s−so ) − ςe−l E f df Em + E
τfd (s) = τf η   f

1 − ςe−(l−s+so )
1 − exp −4υf μle 
1+υ 1−υ
E f df E mf + E f
f

(15)
Pmax
τ= (16)
nπ df le

3.2 Physical Modeling (a) (b) (c) (d)


Fig. 12 Different configurations for pullout test: a single-fiber single-
The fiber–matrix interfacial interaction has been investigated sided, b multiple-fiber single-sided, c single-fiber double-sided config-
using direct and indirect methods. Direct methods are based uration, d multiple-fiber double-sided [14]
on pullout tests of single fiber or group of fibers [51]. In indi-
rect methods, the fiber–matrix interfacial bond is quantified
from a mechanical property of a type of FRC (mostly the flex- Table 5 The criteria for efficient pullout test configuration [6,12,18]
ural strength). In the pullout test, the fiber pullout load against
Description of the criteria
the fiber response history (slip or crack mouth opening dis-
placement) is acquired [52]. Despite great efforts to propose a Testing configuration should be able to measure the
an ideal pullout test method, no available method satisfies all fiber–matrix interfacial response accurately
required measurements. Till date, there is no standard test in b The test configuration should facilitate pullout of the fiber
with very small free length
any code or specification to assess the fiber–matrix interac-
c The test configuration should facilitate pullout of the fibers
tions for provided fiber and matrix.
with up to 90◦ orientation with respect to the loading
d Testing configuration should facilitate the visual inspection
3.2.1 Direct Methods of the fiber–matrix interface
e Testing specimen should be easily fabricated
The pullout test is an experimental method to assess the dis- f Testing specimen should simulate an element of a FRC
placement response of fiber or group of fibers embedded in g Testing specimen should facilitate an accurate placement of
a cementitious matrix to a uniaxial tensile quasi-static or dy- fibers with different geometries during the fabrication of
namic load, which increases with a uniform rate to cause the testing specimens
fracture in the fiber–matrix interface or the fiber. The ac- h Testing specimens should be easy and safely transported
curacy of the pullout test measurement system is important i Testing specimens should facilitate high productivity and
in the reliability of the results because the fiber response is safe storage in different curing regimes
highly sensitive even to small pullout loads. This test has j Testing specimens should be appropriate for applying the
secondary effects “as the radial or axial stresses
been used to (1) optimize a mechanical property of FRC; (2)
k Testing specimens should minimize stress concentration
investigate the effect of different parameters of the fiber, ma-
and shrinkage effects
trix and interfacial transition zone (ITZ) of the fiber–matrix
interface; (3) evaluate the critically embedded fiber lengths;
and (4) quantify the properties of the fiber–matrix bond for a
provided type of FRC. However, these studies are rare in the sided test configuration. However, the single-fiber double-
literature [16,53]. Many researchers reported that there was sided test configuration was rarely used [12]. The single-fiber
no direct correlation between the fiber pullout test results single-sided test configuration is associated with technical
and the mechanical properties of the applied FRC [6,54]. problems in controlling the free tip of the fiber without break-
Nonetheless, the information obtained from the fiber pullout ing [10].
test can be used to optimize the FRC properties [51]. Many researchers suggested the criteria in Table 5 to de-
So far, several pullout test setups have been developed velop an efficient pullout test setup [6,12,18]. Based on these
to establish the pullout load–slip relationship for embedded criteria, Table 6 presents a review on the previously used test
fiber in a matrix. The test configurations can be classified configurations. Equation 16 is used by several researchers
into four types (Fig. 12): (1) single-fiber single-sided, (2) to evaluate the shearing bond of the fiber–matrix interface.
multiple-fiber single-sided, (3) single-fiber double-sided, and Based on the assumption that the pullout load is equally dis-
(4) multiple-fiber double-sided. Because of its simplicity, tributed between the fibers on both sides for the double-sided
the researchers have extensively used the single-fiber single- test configurations, the slip response of a single fiber is esti-

123
1194 Arab J Sci Eng (2016) 41:1183–1198

Table 6 Pullout test methods


Testing procedure Date Ref. Criteria as defined in Table 4 Remarks
a b c d e f g h i j k
√ √ √ √
1972, 2003, 2013 [4,21,24]    ×  × × Fiber orientation up to 75◦
was investigated

√ √ √ √ √
1976 [58]     × × A cylindrical specimen was
used to investigated the
Poisson’s effects of the
fiber–matrix bond

√ √ √ √ √ √
1978 [56]     × This test configuration was
developed for better
physical simulation for
the fracture of the fiber.
However for low
percentage volume of
fibers, it was
demonstrated that the
effect on the fiber–matrix
is marginal [56]

√ √ √ √
1991 [11] ×   ×  × ×


1997 [8] × ×   ×   × × ×

√ √
1999 [35]  × × ×    × × The test involves applying
simultaneous two
direction loading

√ √ √ √ √
2001 [39] × ×  × ×  A 65 mm dia. × 50 mm
height cylindrical
specimen drilled from
150 × 150 × 150 mm3
concrete cubes

√ √ √ √ √ √
2002 [18]  × × ×  Fiber orientation up to 60◦
was investigated

√ √
2004 [34]  × ×     × × Nine steel fibers were
embedded in the ASTM
tensile test specimen

√ √ √ √
2004 [28] × ×    ×  A 2 mm dia. × 50 mm
height cylindrical
specimen was used

123
Arab J Sci Eng (2016) 41:1183–1198 1195

Table 6 Continued
Testing procedure Date Ref. Criteria as defined in Table 4 Remarks
a b c d e f g h i j k
√ √ √ √
2009 [53] × × ×  × × ×

√ √ √ √ √ √
2010 [3]  × ×  × A 25 mm dia. × 100 mm height
cylindrical specimen was
used. A microstructural check
for the fiber location was
conducted.

√ √ √ √ √ √ √ √
2010 [10]   × 32 fibers oriented at up to 60◦
were investigated in a
dog-bone specimen

√ √ √
2011 [26] ×   × ×  × × The test configuration was used
for comparing the pullout
load and toughness for
different fibers, lading rates
and matrix strengths

√ √ √ √ √
2012 [30] ×  ×  × ×

√ √ √ √ √ √ √
2012 [17]   × × A 65 mm × 65 mm × 200 mm
double-sided prismatic
specimen was used

√ √
2012 [36] × ×    ×   × A 10 mm × 10 mm × 20 mm
prismatic specimen was used

√ √ √ √ √
2012, 2013 [27,58] × × ×   × A 50 and 70 mm cubic
specimen was used to study
the fiber–matrix bond

√ √ √ √
2013 [31] ×  ×  × ×  A coconut fiber is embedded in
beam having
100 mm × 100 mm cross
section. The fiber pullout test
does require any type of
fixing since its fiber pullout
loads are much less than the
weight of the beam

√ √ √ √ √ √
2013 [23]     × A cubic specimen was used to
compare the fiber–matrix
bond of hooked and 3D fibers


Satisfied  Partially satisfied × Not satisfied

123
1196 Arab J Sci Eng (2016) 41:1183–1198

mated by dividing the measured slip by two [10]. In addition, Pre-cracking stage
the pullout toughness (energy absorbed in the fiber–matrix
interface for a particular crack opening width) is evaluated Post-cracking stage
by numerically integrating the load–slip relationship with a
limited amount of slip (e.g., 5 mm or the embedded length of

Pullout force
the fiber) to satisfy the maximum permissible crack opening
for certain serviceability requirement [28].
It can be seen from Table 6 that none of the used test
steps satisfies the technical requirements for a perfect phys-
ical simulation of the fiber–matrix interactions in an actual
FRC structure with high accuracy, productivity and safety
(Table 5). Most used test setups satisfy the accuracy of the
fiber-slip curve condition (condition a in Table 5). However,
most of them fail to simulate the secondary effects as the Crack opening width
radial stresses (condition l in Table 5). Generally, the double-
Fig. 13 Typical result for single-fiber pullout test [38]
sided test setups succeed to simulate an element of FRC;
however, the standard or modified tensile strength specimen
(dog-bone specimen) is subjected to concentrated stresses 3.3 Indirect methods
because of the gripping systems. Most single-sided test con-
figurations are held in position during testing by applying Prudencio et al. [19] suggested an indirect method to predict
bearing stresses opposite to the loading direction; however, the pullout behavior of fiber from the results of the four-point
this case does not occur in the actual FRC. Some single-sided loading test. In this method, the load in a fiber that crosses the
test setups were glued to the bottom of the specimen, which critical section of the beam is estimated from the equilibrium
simulated an actual FRC structure better [39]. of the forces and moments if the number of the fibers (Eq. 17)
Ali et al. [31] used a single-sided specimen without re- is assumed to cross that section.
straining because its weight was much more than the ulti-
mate pullout load. However, this test setup failed to satisfy Vf Ab δ
the safety and productivity requirements (conditions e, h n= (17)
25π df2
and i in Table 5). The closest setup to achieve the physical
simulation requirements was presented by Gray and John-
ston [55], where they modified the basic single-fiber single-
sided test. For better physical modeling, this modification 4 Conclusions
involves transferring the restraining stresses to inside the
matrix instead of the specimen surface. Only the cylindri- The conclusions of this study are summarized as follows:
cal specimens introduced by Pinchin and Tabor [56] and The fiber–matrix interactions govern the mechanical prop-
Georgiadi-Stefanidi et al. [3] satisfy the requirement for ap- erties of FRC, and these interactions are governed by fric-
plying secondary (radial) effects. However, these test setups tional bond. The physical/chemical bond plays a minor role.
are subjected to stress concentrations because of the radial The frictional bond depends on the coefficient of friction of
pressure stresses on the circumferential surface. It should be the fiber–matrix interface, normal stresses in the matrix and
noted that the testing methodology should minimize the devi- Poisson’s effects on the fiber–matrix interface. The bond of
ation of results. Generally, 20–55 % coefficient of variation the fiber–matrix interface has been quantified by direct and
was reported for the results of the pullout test setups [56]. indirect methods. In direct methods, the fiber–matrix bond
As reported by Bentur and Mindess [51], this variation was behavior is assessed by measuring the uniaxial tensile pullout
attributed to (1) high variability of the measured properties, load and the corresponding slippage during the fiber pull-
(2) testing management, and (3) variations in gripping and out from a cementitious matrix, while in indirect methods,
alignments. Figure 13 shows a typical result of a single-fiber the fiber–matrix bond strength is assessed as a mechanical
pullout test. This figure demonstrates that until the ultimate property of FRC. In theoretical modeling of the fiber–matrix
pullout load (pre-cracking stage), the pullout load and slip- interactions, most researchers assume a constant fiber–matrix
page are linearly proportional. However, at the post-cracking interfacial bond in the post-cracking stage. All researchers as-
state, the slippage drastically increases with the decrease in sumed that the debonding began when the first crack began in
pullout force. Because of the frictional fiber–matrix interfa- the FRC. Increasing the fiber deformations in size, amount,
cial bond, the post-cracking stage has a random behavior for intensity or degree increases the ultimate pullout load and
the pullout force-slippage relation [24]. decreases the pullout toughness.

123
Arab J Sci Eng (2016) 41:1183–1198 1197

The pullout test is an experimental method to assess the References


displacement response of fiber or group of fibers embedded
in a cementitious matrix to a uniaxial tensile quasi-static or 1. Mehta, P.K.; Monteiro, P.J.: Concrete: Microstructure, Properties,
and Materials, vol. 3. McGraw-Hill, New York (2006)
dynamic load, which increases with a uniform rate to cause
2. Bartos, P.: Review paper: bond in fibre reinforced cements and
fracture in the fiber–matrix interface or the fiber. Despite concretes. Int. J. Cement Compos. Lightweight Concrete 3(3), 159–
great efforts to develop an ideal pullout test method, none of 177 (1981). doi:10.1016/0262-5075(81)90049-X
the used test setups satisfies the technical requirements for a 3. Georgiadi-Stefanidi, K.; Euripidis, M.; Dafni, P.; Michalis, Z.: Nu-
merical modelling of the pull-out of hooked steel fibres from
perfect physical simulation of the fiber–matrix interactions
high-strength cementitious matrix, supplemented by experimental
in an actual FRC structure with high accuracy, productivity results. Constr. Build. Mater. 24(12), 2489–2506 (2010). doi:10.
and safety. The used test configurations can be classified into: 1016/j.conbuildmat.2010.06.007
single-fiber single-sided, multiple-fiber single-sided, single- 4. Naaman, A.E.: Engineered steel fibers with optimal properties
for reinforcement of cement composites. J. Adv. Concrete Tech-
fiber double-sided, and multiple-fiber double-sided. Because
nol. 1(3), 241–252 (2003). doi:10.3151/jact.1.241
of its simplicity, the researchers have intensively used the 5. Gray, R.; Johnston, C.: The influence of fibre–matrix interfacial
single-fiber single-sided test configuration. Due to the high bond strength on the mechanical properties of steel fibre reinforced
variability of the measured properties, testing management, mortars. Int. J. Cement Compos. Lightweight Concrete 9(1), 43–
55 (1987). doi:10.1016/0262-5075(87)90036-4
gripping and alignments, the used pullout test setups exhibit 6. Maage, M.: Interaction between steel fibers and cement based
high variability (20–55 % coefficient of variation). matrixes. Mater. Constr. 10(5), 297–301 (1977). doi:10.1007/
The initiation of the first crack alters the role of the fiber– BF02478831
matrix interfacial bond. Before the first crack begins, the 7. Chan, Y.; Li, V.C.: Effects of transition zone densification on
fiber/cement paste bond strength improvement. Adv. Cement Based
ultimate shearing stress over the fiber surface is located at its Mater. 5(1), 8–17 (1997). doi:10.1016/S1065-7355(97)90010-9
edges; after the first crack begins, the ultimate shearing stress 8. Shannag, M.; Brincker, R.; Hansen, W.: Pullout behavior of
is at approximately the fiber mid-length. Experimental and steel fibers from cement-based composites. Cement Concrete
theoretical investigations are required to enhance the under- Res. 27(6), 925–936 (1997). doi:10.1016/S0008-8846(97)00061-6
9. Li, V.C.; Stang, H.: Interface property characterization and
standing of the fiber effects in the pre-crack stage. Moreover, strengthening mechanisms in fiber reinforced cement based com-
based on the fiber alignment with respect to the applied posites. Adv. Cement Based Mater. 6(1), 1–20 (1997). doi:10.1016/
stresses, the fiber–matrix interfacial bond can be classified S1065-7355(97)90001-8
into tensile, elastic shear and frictional shear. The tensile 10. Lee, Y.; Kang, S.; Kim, J.: Pullout behavior of inclined
steel fiber in an ultra-high strength cementitious matrix. Con-
bond strength is important to the ductility of FRC. str. Build. Mater. 24(10), 2030–2041 (2010). doi:10.1016/
Studies are required to rationally understand the contribu- S1065-7355(97)90001-8
tion of the interfacial tensile strength of the fiber to the matrix 11. Li, Z.; Mobasher, B.; Shah, S.P.: Characterization of interfacial
to the overall behavior of FRC. Theoretically, the orientation, properties in fiber-reinforced cementitious composites. J. Am.
Ceram. Soc. 74(9), 2156–2164 (1991). doi:10.1111/j.1151-2916.
spacing and aspect ratio of fibers are incorporated to model 1991.tb08276.x
the fiber–matrix interactions, however, to precisely simulate 12. Gray, R.: Experimental techniques for measuring fibre/matrix
the fiber–matrix interactions, a theoretical model incorpo- interfacial bond shear strength. Int. J. Adhes. Adhes. 3(4), 197–
rating all possible governing parameters must be developed. 202 (1983). doi:10.1016/0143-7496(83)90094-5
13. DiFrancia, C.; Ward, T.; Claus, R.: The single-fibre pull-
These parameters should include the radial deformations be- out test. 1: review and interpretation. Compos. Part A
cause of shrinkage and Poisson’s effects, repetitive-loading Appl. Sci. Manuf. 27(8), 597–612 (1996). doi:10.1016/
effect and nonlinear stress distribution along the fiber in the 1359-835X(95)00069-E
pre-cracking stage. 14. Wang, Y.; Li, V.C.; Backer, S.: Modelling of fibre pull-out from
a cement matrix. Int. J. Cement Compos. Lightweight Con-
Little information is available in the literature on the dura- crete 10(3), 143–149 (1988). doi:10.1016/0262-5075(88)90002-4
bility of the fiber–matrix bond under different environmental 15. Lin, Z.; Kanda, T.; Li, V.C.: On interface property characterization
conditions. Research programs must study the environmen- and performance of fiber reinforced cementitious composites. Con-
tal impact on the long-term behavior of the fiber–matrix crete Sci. Eng. 1(3), 173–184 (1999)
16. Beaumont, P.; Aleszka, J.: Cracking and toughening of concrete
bond with emphasis on the microstructural investigations and polymer-concrete dispersed with short steel wires. J. Mater.
under different exposure conditions. So far, various tech- Sci. 13(8), 1749–1760 (1978). doi:10.1007/BF00548738
niques were proposed to enhance the fiber–matrix interfacial 17. Chin, C.; Xiao, R.: Experimental and nonlinear finite element
bond. To enhance the ductility of high-performance concrete analysis of fiber-cementitious matrix bond-slip mechanism. High
Perform. Fiber Reinf. Cement Compos. 6, 145–152 (2012). doi:10.
and UHPC, innovative solutions are required to enhance the 1007/978-94-007-2436-5_18
fiber–matrix interfacial bond. 18. Robins, P.; Austin, S.; Jones, P.: Pull-out behaviour of hooked
steel fibres. Mater. Struct. 35(7), 434–442 (2002). doi:10.1007/
Acknowledgments The Author extends his appreciation to the Dean- BF02483148
ship of Scientific Research at King Saud University for funding the work 19. Prudencio, L. Jr.; Simon, A.; Peter, J.; Hugo, A.; Peter, R.: Predic-
through the research group Project No. RGP-VPP-105. tion of steel fibre reinforced concrete under flexure from an inferred

123
1198 Arab J Sci Eng (2016) 41:1183–1198

fibre pull-out response. Mater. Struct. 39(6), 601–610 (2006). interaction. J. Eur. Ceram. Soc. 28(1), 183–192 (2008). doi:10.
doi:10.1617/s11527-006-9091-2 1016/j.jeurceramsoc.2007.05.019
20. Grünewald, S.: Performance-based design of self-compacting fibre 38. Mu, B.; Meyer, C.; Shimanovich, S.: Improving the interface bond
reinforced concrete. Delft University Press, Delft, The Nether- between fiber mesh and cementitious matrix. Cement Concrete
lands (2004) Res. 32(5), 783–787 (2002). doi:10.1016/S0008-8846(02)00715-9
21. Wille, K.; Naaman, A.: Bond stress-slip behavior of steeh fibers 39. Markovich, I.; Van Mier, J.; Walraven, J.: Single fiber pullout from
embedded in ultra high performance concrete. In: ECF18, Dresden hybrid fiber reinforced concrete. HERON 46(3), 191–200 (2001)
(2013) 40. Li, V.C.; Cynthia, W.; Shuxin, W.; Atsuhisa, O.; Tadashi, S.: In-
22. Kim, J.; Kim, D.; Kang, S.; Lee, J.: Influence of sand to coarse terface tailoring for strain-hardening polyvinyl alcohol-engineered
aggregate ratio on the interfacial bond strength of steel fibers in con- cementitious composite (PVA-ECC). ACI Mater. J. Am. Concrete
crete for nuclear power plant. Nuclear Eng. Des. 252, 1–10 (2012). Inst. 99(5), 463–472 (2002)
doi:10.1016/j.nucengdes.2012.07.004 41. Lawrence, P.: Some theoretical considerations of fibre pull-out
23. Richardson, A.; Heather, M.: Improving the performance of con- from an elastic matrix. J. Mater. Sci. 7(1), 1–6 (1972). doi:10.1007/
crete using 3D fibres. Procedia Eng. 51, 101–109 (2013). doi:10. BF00549541
1016/j.proeng.2013.01.016 42. Laws, V.: The efficiency of fibrous reinforcement of brittle ma-
24. Naaman, A.E.: A statistical theory of strength for fiber reinforced trices. J. Phys. D Appl. Phys. 4(11), 1737 (1971). doi:10.1088/
concrete. Ph.D. Thesis, Massachusetts Institute of Technology 0022-3727/4/11/318
(1972) 43. Aveston, J.; Kelly, A.: Theory of multiple fracture of fibrous
25. Abu-Lebdeh, T.; Hamoush, S.; Zornig, B.: Rate effect on pullout be- composites. J. Mater. Sci. 8(3), 352–362 (1973). doi:10.1007/
havior of steel fibers embedded in very-high strength concrete. Am. BF00550155
J. Eng. Appl. Sci. 3(2), 454 (2010) 44. Swamy, R.; Mangat, P.; Rao, C.K.: The mechanics of fiber rein-
26. Abu-Lebdeh, T.; Sameer, H.; William, H.; Brian, Z.: Effect of ma- forcement of cement matrices. Spec. Publ. 44, 1–28 (1974)
trix strength on pullout behavior of steel fiber reinforced very-high 45. Naaman, A.; Argon, A.; Moavenzadeh, F.: A fracture model
strength concrete composites. Constr. Build. Mater. 25(1), 39– for fiber reinforced cementitious materials. Cement Concrete
46 (2011). doi:10.1016/j.conbuildmat.2010.06.059 Res. 3(4), 397–411 (1973). doi:10.1016/0008-8846(73)90078-1
27. Tuyan, M.; Yazıcı, H.: Pull-out behavior of single steel fiber from 46. Allen, H.: Stiffness and strength of two glass–fiber reinforced ce-
SIFCON matrix. Constr. Build. Mater. 35, 571–577 (2012). doi:10. ment laminates. J Compos. Mater. 5(2), 194–207 (1971). doi:10.
1016/j.conbuildmat.2012.04.110 1177/002199837100500205
28. Singh, S.; Shukla, A.; Brown, R.: Pullout behavior of 47. Nair, N.: Mechanics of glass fibre reinforced cement. In: Rilem
polypropylene fibers from cementitious matrix. Cement Concrete Symposium. The Construction Press Ltd Hornby (1975)
Res. 34(10), 1919–1925 (2004). doi:10.1016/j.cemconres.2004. 48. Argon, A.; Shack, W.: Opening Paper: Theories of Fibre Cement
02.014 and Fibre Concrete. DTIC Document, Fort Belvoir (1975)
29. Barluenga, G.: Fiber–matrix interaction at early ages of concrete 49. Bartoš, P.: Analysis of pull-out tests on fibres embedded in brittle
with short fibers. Cement Concrete Res. 40(5), 802–809 (2010). matrices. J. Mater. Sci. 15(12), 3122–3128 (1980). doi:10.1007/
doi:10.1016/j.cemconres.2009.11.014 BF00550385
30. Kim, J.J.; Dong, J.; Su, T.; Jang, H.: Influence of sand to coarse 50. Naaman, A.E.; George, G.; Jamil, M.; Husam, S.: Fiber pullout
aggregate ratio on the interfacial bond strength of steel fibers in con- and bond slip. I: analytical study. J. Struct. Eng. 117(9), 2769–
crete for nuclear power plant. Nuclear Eng. Des. 252, 1–10 (2012). 2790 (1991). doi:10.1061/(ASCE)0733-9445(1991)117:9(2769)
doi:10.1016/j.nucengdes.2012.07.004 51. Bentur, A.; Mindess, S.: Fibre reinforced cementitious composites,
31. Ali, M.; Li, X.; Chouw, N.: Experimental investigations on bond 2nd edn. CRC Press, Boca Raton (2006)
strength between coconut fibre and concrete. Mater. Des. 44, 596– 52. Cunha, V.; Barros, J.; Sena-Cruz, J.M.: Pullout behaviour of
605 (2013). doi:10.1016/j.matdes.2012.08.038 hooked-end steel fibres in self-compacting concrete. Civil Engi-
32. Boshoff, W.; Mechtcherine, V.; van Zijl, G.: Characterising the neering, Report 07-DC/E06.Universidade do Minho, Guimarães
time-dependant behaviour on the single fibre level of SHCC: 53. Aiello, M.; Leuzzi, F.; Centonze, G.; Maffezzoli, A.: Use of
Part 1: mechanism of fibre pull-out creep. Cement Concrete steel fibres recovered from waste tyres as reinforcement in con-
Res. 39(9), 779–786 (2009). doi:10.1016/j.cemconres.2009.06. crete: pull-out behaviour, compressive and flexural strength. Waste
007 Manag. 29(6), 1960–1970 (2009). doi:10.1016/j.wasman.2008.12.
33. Boshoff, W.; Mechtcherine, V.; van Zijl, G.: Characterising the 002
time-dependant behaviour on the single fibre level of SHCC— 54. Hughes, B.; Fattuhi, N.: Load–deflection curves for fibre-
Part 2: the rate effects on fibre pull-out tests. Cement Concrete reinforced concrete beams in flexure. Mag. Concrete
Res. 39(9), 787–797 (2009). doi:10.1016/j.cemconres.2009.06. Res. 29(101), 199–206 (1977). doi:10.1680/macr.1977.29.
006 101.199
34. Chan, Y.; Chu, S.: Effect of silica fume on steel fiber bond 55. Gray, R.; Johnston, C.: The measurement of fibre–matrix interfacial
characteristics in reactive powder concrete. Cement Concrete bond strength in steel fibre-reinforced cementitious composites. In:
Res. 34(7), 1167–1172 (2004). doi:10.1016/j.cemconres.2003.12. Proceedings of the RILEM Symposium (1978)
023 56. Pinchin, D.; Tabor, D.: Inelastic behaviour in steel wire pull-
35. Wu, H.; Li, V.C.: Fiber/cement interface tailoring with plasma treat- out from Portland cement mortar. J. Mater. Sci. 13(6), 1261–
ment. Cement Concrete Compos. 21(3), 205–212 (1999). doi:10. 1266 (1978). doi:10.1007/BF00544732
1016/S0958-9465(98)00053-5 57. Pinchin, D.: Poisson contraction effects in aligned fibre com-
36. Sebaibi, N.; Mahfoud, B.; Nor Edine, A.; Christophe, B.: Me- posites. J. Mater. Sci. 11(8), 1578–1581 (1976). doi:10.1007/
chanical properties of concrete-reinforced fibres and powders with BF00540895
crushed thermoset composites: The influence of fibre/matrix inter- 58. Zı̄le, E.; Zı̄le, O.: Effect of the fiber geometry on the pullout re-
action. Constr. Build. Mater. 29, 332–338 (2012). doi:10.1016/j. sponse of mechanically deformed steel fibers. Cement Concrete
conbuildmat.2011.10.026 Res. 44, 18–24 (2013). doi:10.1016/j.cemconres.2012.10.014
37. Sedan, D.; Cécile, P.; Agnèse, S.; Thierry, C.: Mechanical proper-
ties of hemp fibre reinforced cement: Influence of the fibre/matrix

123

You might also like