Sie sind auf Seite 1von 11

KSCE Journal of Civil Engineering (2013) 17(6):1383-1393

DOI 10.1007/s12205-013-1115-1

Structural Engineering

www.springer.com/12205

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams


Youcef Fritih*, Thierry Vidal**, Anaclet Turatsinze***, and Grard Pons****
Received March 2, 2012/Revised July 26, 2012/Accepted November 25, 2012

Abstract
This paper deals with the effect of steel fiber reinforcement on the behavior of Self-Compacting Concrete (SCC) beams. Bending
tests were carried out to examine the effect of low fiber content (0.25% by volume) on the flexural behavior of beams with different
amounts of steel rebar reinforcement. The study compares the behavior of reinforced concrete beams cast either with control SCC
and the one of Fiber-Reinforced Self-Compacting Concrete (FRSCC). Fibers used were made of stainless amorphous metal. Their
influence was studied through the global and local mechanical responses of the beams. The results show that fiber reinforcement
allows the control of cracking to be improved. Yielding, ductility and load bearing capacity are not modified by the fiber
reinforcement; its effects are limited to the kinetics and distribution of cracks. If it was observed that the used fiber content reduced
stresses in the stirrups, they could not be considered as a solution to replace stirrups. However, their ability to transfer tensile stress
through a crack provided greater beam stiffness, notably with a low steel bar reinforcement ratio. It was concluded that stainless steel
fiber reinforcement is a suitable solution to control crack width in reinforced concrete elements in aggressive environments with
respect to the limitations imposed by design codes such as the European code Eurocode 2.
Keywords: beam, fiber-reinforcement, self-compacting concrete, cracking control, flexural behavior, shear behavior

1. Introduction
The brittle nature of concrete causes it to collapse shortly after
the formation of the first crack. When steel fibers are added to a
concrete mix, they are randomly distributed and act as crack
arrestors (Beaudoin, 1990; Yun et al., 2007). The addition of
steel fibers helps to convert brittle concretes into more ductile
ones. The main role of fibers is to transfer stress across the crack
and thus to restrain crack opening and propagation. Recent
research (Banthia and Trottier, 1994; Cucchiara et al., 2004;
Edginton et al., 1978; Ezeldin and Balaguru, 1989; Furlan and de
Hanai, 1997; Khuntia and Stojadinovic, 2001) has indicated that
fiber reinforcement improves the mechanical properties of
concrete, mainly the post-cracking behavior.
Much research has investigated the influence of fibers on
behavior of reinforced concrete structural elements, as beams.
Before flexural beam cracking, the effect of fibers on global
mechanical response is contradictory. Maximum deflection is
increased in presence of fibers for Barragn (2002) whereas it is
not affected for other studies (Lim and Oh, 1999), (Narayanan et
Darwish, 1987), (Furlan and de Hanai, 1997). After flexural

cracking, (Narayanan and Darwish, 1987) and (Cucchiara et al.,


2004) have observed a more dense crack network for fiber
reinforced concrete beams. Oh (1992) noticed a better cracking
control thanks to fibers, characterized by cracks width reductions.
Furthermore, the fibers generate an increase in stiffness, load
bearing capacity, ultimate load, shear strength, resistance to
diagonal cracking (Frosch, 2000; Mirsayah and Banthia, 2002;
Shin et al., 1994). More recently, You et al. (2010) have
demonstrated the capacity of hooked steel fibers to partially
replace stirrups of self-compacting concrete beams. However,
from the synthesis of the overall studies, it is rather difficult to
precisely assess the effect of fibers since it depends on several
parameters linked with beams design (geometry, dimensions,
loading, longitudinal and shear steel reinforcements, etc.) and on
the various type and amount of fibers.
This study focuses on the effect on crack opening and on
tensile stress in the stirrups, with the prospect of a possible
reduction in the amount of shear reinforcement. However, the
usual fiber content (0.5 to 2% by volume) employed to reinforce
concrete induces a detrimental effect on workability and an
additional cost, which tend to limit industrial applications and to

*Scientific Researcher, Universit de Toulouse, UPS, INSA, LMDC (Laboratoire Matriaux et Durabilit des Constructions), F-31077 Toulouse Cedex
04, France (E-mail: youcef.fritih@insa-toulouse.fr)
**Assistant Professor, Universit de Toulouse, UPS, INSA, LMDC (Laboratoire Matriaux et Durabilit des Constructions), F-31077 Toulouse Cedex 04,
France (Corresponding Author, E-mail: thierry.vidal@insa-toulouse.fr)
***Professor, Universit de Toulouse, UPS, INSA, LMDC (Laboratoire Matriaux et Durabilit des Constructions), F-31077 Toulouse Cedex 04, France
(E-mail: anaclet.turatsinze@insa-toulouse.fr)
****Professor, Universit de Toulouse, UPS, INSA, LMDC (Laboratoire Matriaux et Durabilit des Constructions), F-31077 Toulouse Cedex 04, France
(E-mail: gerard.pons@insa-toulouse.fr)
1383

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams

compromise the material properties of the hardened concrete


(Bayasi and Soroushian, 1992).
In order to promote fibrous-concrete-based applications, the
effect of low fiber content on reinforced concrete elements needs
to be investigated. Fibers and SCC could be considered as a
judicious association since this type of material limits the
problem of workability due to friction between fibers and
aggregates. The choice was confined to the study of a series of
beams based on SCC and on FRSCC, assuming that the SCC
and plain concrete have similar shear behavior (Redon and
Chermant, 1999). This paper presents the investigations on the
influence of low fiber content (20 kg/m3, i.e., 0.25% by volume)
on the behavior of six 3-meter-long beams.
Stainless high-bond steel fibers with a high modulus of elasticity
were used. Bending tests were carried out and the global
mechanical behavior was studied through the midspan deflections
and failure mode. The cracking patterns and crack widths were also
monitored and compared. The evolution of longitudinal steel
reinforcement strain was monitored and analyzed. Finally, the
behavior of transverse bars was finely studied after the onset of
shear cracks in order to consider their total or partial substitution by
a fiber reinforcement. Reducing the number of shear bars in
reinforced concrete elements may improve the production process.
However, the relevance of this alternative is intimately linked to its
capacity to avoid brittle failure under shear forces.

2. Materials
2.1 Concrete Mixes
The choice of concretes focused on self-compacting concretes
for their many advantages: the application of SCC in construction
provides benefits from the perspective of materials technology
and environmental protection and presents exciting opportunities
to engineers and architects alike. These mixes were chosen
according to industrial and local criteria. The SCC studied here
was designed using a CEM I 52.5R cement (average particle size
d50 of 14 m) complying with European standard ENV 197-1/
A3: 2009, limestone filler (d50 = 15 m, MgO content 0.2% by
weight), river aggregates (0/4 mm sand, 4/10 mm gravel) and
acrylic copolymer-based superplasticizer.
2.2 Fiber-reinforcement
One type of macro-metallic fiber, called FIBRAFLEX, was
used (Fig. 1). The ribbon-shaped fibers are 30 mm long, 1.6 mm
wide and 0.03 mm thick, and belong to the family of metallic
glass materials. They are made of amorphous metal (Fe, Cr) 80%
and (P, C, Si) 20% by mass. No corrosion was observed when
such fibers were immersed in HCl (0.1 N) and in FeCl3 (0.4 N)
for 24 h at 35C following ASTM G48-76-A (Choulli et al.,
2008). These fibers are straight, flexible and stainless. Their
characteristics are given in Table 1. They can be used in severe
conditions, maintaining their mechanical efficiency without
generating the unsightly rust stains that appear on the concrete
surface when traditional metal fibers are used. These fibers have

Fig. 1. Amorphous Metallic Fibers Used


Table 1. Main Characteristics of Fibers Used
Fiber type
Length (mm)
Cross section (mm2)
Density
Tensile strength (MPa)
Elastic modulus (GPa)
Specific properties

Amorphous metal
30
Rectangular 1.6 0.03
7.20
2000
140
Stainless

a high bond with the concrete matrix due to their rough surface
and high specific surface. So, at the time of crack initiation, the
fibers do not slide from the matrix and are immediately
tensioned, thus acting to restrain the crack opening (Turatsinze et
al., 2005). When the stress is too high and exceeds the tensile
strength of the fibers, they break instead of pulling out.
The effect of this amorphous fiber reinforcement on the tensile
behavior of concrete has been widely investigated and is well
established thanks to direct tensile test according to Rilem
recommendations (TC 162-TDF Rilem, 2001) that specify the
experimental procedure. Uni-axial tension tests have been
extensively performed on mono fibered concretes (Turatsinze et
al., 2003) and on hybrid fibered concretes (Pons et al., 2007)
(Hameed et al., 2010a) reinforced by different types of fiber. All
results show that, due to the high tensile strength and high
modulus of elasticity of amorphous fiber used and because of
their large specific surface which provides high bond with the
matrix, the concrete reinforced with this type of fibers exhibits
high residual post peak strength. The corresponding post peak
plateau is at a high level but short, given these fibers, unlike the
conventional steel fibers, do not slide but are broken when
critical crack opening estimated to 200 microns is reached
(Turatsinze et al., 2005). Hameed et al. (2010b) showed that the
type of amorphous fibers used in this research allows the tension
stiffening of a reinforced concrete to be increased as soon as a
crack is initiated under uniaxial tensile loading.
2.3 Mix Proportions
The studied Self-Compacting Concrete (SCC) and Steel Fiber
Reinforced Self-Compacting Concrete (FRSCC) were developed
by Pons et al. (2007), following Rossis recommendations (1991).
The mix proportions of these concretes are detailed in Table 2.
They use the same materials. Compared to the SCC mix, the

1384

KSCE Journal of Civil Engineering

Youcef Fritih, Thierry Vidal, Anaclet Turatsinze, and Grard Pons

Table 2. Concrete Mixes


Mix components [kg/m3]

SCC (control)

FRSCC

311.2

353.0

Cement CEM I 52.5R


Limestone Filler (Batite 12)

171.4

220.4

Sand 0/4 mm

824.5

919.3

Gravel 4/10 mm

864.3

574.4

20

Metal fibers
Superplasticizer SIKA3030

3.56

4.30

Water

195.9

232.0

Table 3. Fresh State Characteristics of Concretes Used


SCC

FRSCC

Slump flow (mm)


Flow time t50 (s)

720
2.1

680
2.3

Average J-RING diameter (mm)


hint(cm)
hext(cm)

680
4.3
1.0

650
4.5
1.1

Specific gravity (kg/m3)

2388

2342

1.9

1.5

Air content (%)

Fig. 2. Residual Post-peak Strength versus Crack Opening in a


Direct Tensile Test on Notched Specimen

Table 4. Mechanical Properties of Concretes Used


SCC

FRSCC

28-day compressive strength (MPa)

45.7

42.3

28-day splitting tensile strength (MPa)

3.9

4.0

28020

27330

28-day modulus of elasticity (MPa)

FRSCC is characterized by slightly higher filler, sand and cement


content, and a lower gravel content to compensate for the loss of
workability due to the addition of fiber.
The compositions were optimized to satisfy both the rheological behavior of fresh concrete, highlighted by dedicated tests of
workability according to the usual recommendations, and the
mechanical properties of concrete in the hardened state. The
fibers were used at 20 kg/m3 (0.25% by volume).
2.4 Properties of the Concrete
Fibers affect both the rheological behavior of fresh concrete
and the mechanical behavior of the concrete in the hardened
state. The use of fibers is generally detrimental to the workability
of the fresh material. The properties of both types of concrete in
the fresh state are given in Table 3. These results are within the
range recommended by EN 206-9 (2010) with regard to the fresh
properties of self-compacting concrete.
The compressive strength, modulus of elasticity and splitting
tensile strength of the two concretes at 28 days (the date of the
beam bending tests) are presented in Table 4. Each value
represents the average for 3 tests carried out on cylindrical
specimens 236 mm high and 118 mm in diameter.
2.5 Direct Tensile Tests
In order to quantify the contribution of fibers to the residual
post-cracking strength, direct tensile tests were carried out on
notched prismatic specimens (100 100 200 mm3) using closed
Vol. 17, No. 6 / September 2013

loop CMOD-controlled loading according to RILEM recommendations (TC 162-TDF Rilem, 2001) with the aim of determining
the post-peak residual strength-crack opening relationship. The
results illustrated in Fig. 2 allow a comparison between SCC and
FRSCC sample behavior. For SCC, the diagrams demonstrate
the well-known brittle behavior of the material: a sudden
decrease in the residual strength with increase of the crack
opening after the peak load. When the macrocrack is localized,
its propagation requires little energy.
For FRSCC, the diagram has a different shape. Up to peak load,
the curve is the same as the one for SCC. The zone just before the
peak corresponds to the microcrack initiation and propagation in
the concrete. After the peak load, coalescence phenomenon leads to
damage localization. The post-peak behavior of FRSCC can be
schematized as a three-phase law. The first part corresponds to a
stress decrease from peak load to a residual strength plateau. The
action of fibers allows a high level of residual strength to be
maintained for a crack opening at which the residual strength is
close to zero in the case of control SCC. During the development of
the crack opening in a softening material such as the one studied
here, the drop in the residual strength from the peak load is
unavoidable. The second phase implies a plateau of residual postpeak strength. This plateau is also largely influenced by fiber
properties (modulus of elasticity, bond with the matrix) and fiber
content. The amplitude of the residual strength is high, but the
plateau is short due to the stress concentration in the bridging part of
the fiber, which causes its failure.
The third part coincides with the failure of the specimen. It is
associated with the successive fracture of fibers. According to
Turatsinze et al. (2005), the fibers used in this study do not resist
for crack openings wider than 0.2 mm.

3. Experimental Setup
The objectives of this experimental program were to assess the
ability of a steel fiber reinforcement to control crack opening and
the resulting consequences on the mechanical flexural postcracking response of beams. Their mechanical contribution to
shear strength was also investigated by comparing the responses

1385

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams

Table 5. Geometrical and Mechanical Characteristics of Tested Beams


Identification
of beams

Series A
Series B
Series C

SCC
FRSCC
SCC
FRSCC
SCC
FRSCC

Transverse
reinforcement
spacing
Asw
(cm2)
(mm)
0.50
150
0.50
150

Longitudinal
reinforcement
Asl
(cm2)
3.08
3.08
8.29
8.29
8.29
8.29

of SCC- and FRSCC-based beams.


3.1 Specimen Details
Three series of reinforced concrete beams having identical
rectangular cross sections of 150 280 mm were tested. They
were 3000 mm long with a span length of 2800 mm. For each
series, one beam was cast with SCC and one with FRSCC. A
total of six beams were tested to investigate the influence of the
fiber reinforcement on the mechanical behavior of reinforced
concrete beams. The major test variables were the presence of
fibers (0 or 0.2% by volume), the use of transverse reinforcement
and the longitudinal steel reinforcement ratio. The concrete
cover was maintained constant and equal to 20 mm for all
specimens. Ribbed steel longitudinal reinforcing bars of 14, 16
and 20 mm diameter were used.

Volume
of fibers

Type
of loading

w (%)

Vf (%)

flexure

0.81
0.81
2.28
2.28
2.28
2.28

0
0.25
0
0.25
0
0.25

3 points
3 points
4 points
4 points
4 points
4 points

Expected
failure mode

flexure
flexure
shear
shear
shear
shear

The first group, series A, was made up of beams without


stirrups and with little longitudinal reinforcement, i.e. two 14mm-diameter ribbed re-bars. They were subjected to three-point
flexure tests in order to obtain flexural failure. The second group,
series B, consisted of specimens without stirrups, with a high
longitudinal reinforcement, i.e. two 20-mm and one-16 mm
diameter ribbed re-bars, subjected to four-point flexure tests in
order to obtain shear mode failure. Beams of the third series C
were cast with 6-mm-diameter ribbed stirrups and with strong
longitudinal reinforcement. The beams were designed to fail in
diagonal tension. These details and the experimental setup are
summarized in Table 5 and in Fig. 3.
The parameter rw corresponding to the reinforcement ratio was
calculated using the following formula: rw = Asl/(b.d), where Asl
is the cross-section of the longitudinal reinforcement, b is the
beam web width and d is the effective depth.
3.2 Testing Procedure and Measurements
Two different conditions of loading were adopted for the tests
depending on the targeted mode of failure. The beams were tested

Fig. 3. Geometric Characteristics of Beams and Experimental


Setup (Dimensions in mm)

Fig. 4. Experimental Setup: (a) Experimental Setup for All Beams


(LVDT Sensor at Mid-span and a Strain Gauge on Longitudinal Reinforcement) and for Series C (Strain Gauges
on Stirrups), (b) Location of Strain Gauges on Stirrups for
Series C

1386

KSCE Journal of Civil Engineering

Youcef Fritih, Thierry Vidal, Anaclet Turatsinze, and Grard Pons

under three-point loading for series A, and four-point loading for


series B and C. At midspan of all the beams (Fig. 4(a)), a strain
gauge was placed on the longitudinal reinforcement and a LVDT
sensor was used to measure the deflection. For series C, twelve
additional strain gauges were pasted on the shear reinforcement,
symmetrically located in the 5th, 6th and 7th positions (the 1rst
position corresponding to the closest one of the support). For each
stirrup, two strain gauges were spaced 5 cm apart to maximize the
possibility of these gauges to be crossed by a shear crack. The
exact locations of all strain gauges are shown on Fig. 4(b).The
crack widths were measured at the most tensioned fiber of the
transverse cross-section of the beams. The system used was a
video microscope at a magnification of X175, with an accuracy of
0.02 mm.

4. Results and Discussion


4.1 Cracking Patterns
The crack pattern at regular load levels for the SCC and
FRSCC can be seen in Fig. 5.
Table 6 presents a global synthesis of the cracking state at
failure including the number of cracks, the length of the cracked
zone (distance between the two extreme cracks), the maximum
crack opening width (wmax) and the average spacing between
consecutive cracks.
In all specimens, the first crack was a flexural crack located in the

maximal moment area for a load value between 10 and 15 kN.


Other flexural cracks appeared for higher levels of loading. The
shear cracks were initiated near the support at loads between
35% and 50% of the ultimate load and propagated towards the
loading points.
For beams with fiber reinforcement, it can be seen that the
crack spacing is smaller while the crack network is denser than in
the case of the control-concrete-based beam. For each series, an
overall view also shows that the crack propagation was delayed
for fiber-reinforced beams in terms of opening and height.
This phenomenon can be explained by the ability of fibers to
transfer stresses to the concrete through a crack. The crack
distribution is slightly more regular than the one observed on
SCC beams. Thus, the contribution of the concrete area between
two existing flexural cracks to the tensile strength, i.e. the
concrete tension stiffening, seems to be improved. This mechanism induces the development of built-in tensile stresses
generating new cracks when the concrete tensile strength is
reached. In such conditions, the crack density is higher, a result
which can be compared to the one observed for a high
longitudinal reinforcement ratio (Abrishami and Mitchell, 1997)
which also improves the tensile stress transfer to concrete
through the steel-concrete interface.
In consequence, it can be noted that, at a given load level,
the propagation of flexure and of shear cracks is limited in
FRSCC specimens. However, with the type and content of
fibers used, no significant effect was observed on the extent
of the cracked area at failure.
4.2 Global Mechanical Behavior (Load deflection)
Representative curves of load-deflection for all beams tested

Fig. 5. Crack Pattern in Beams vs Loading Levels

Fig. 6. Load-deflection

Table 6. Effects of Fiber Reinforcement on Cracking Observed at the Failure Level


Beams
A-SCC
A-FRSCC
B-SCC
B-FRSCC
C-SCC
C-FRSCC

Number
of cracks
22
19
26
31
28
29

Vol. 17, No. 6 / September 2013

Length of cracked
zone (cm)
208
175
220
228
242
220

wmax
(m)
270
180
213
191
361
300
1387

wmax
reduction (%)
33.3
10.3
16.9

Average crack spacing


(cm)
10.5
9.7
8.8
7.6
9.0
7.8

Spacing
reduction (%)
7.6
13.6
13.3

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams

are plotted in Fig. 6. Using these curves, the following aspects


were analyzed :
Behavior before first crack
First cracking load
Stiffness effect after longitudinal cracking
Behavior after yielding of steel bar
Before the first crack initiation, the SCC control beams and the
FRSCC beams behaved similarly in terms of load versus
deflection response. At this stage, the evolution of stiffness did
not depend on the presence of fibers.
After the first crack appeared, all the beams showed a nonlinear
response. During the stabilized cracking phase under service
moment (load), a slight increase of stiffness was observed for
beams of series A with fiber reinforcement (A-FRSCC) only.
This behavior is thought to result from the greater tension
stiffening effect induced by a better tensile stress transfer by
fibers through cracks (Bischoff, 2000; Bischoff, 2003; Mitchell
et al., 1996). However, as in the case of this experimental study,
when a low fiber content is associated with a strong longitudinal
reinforcement (B and C series), the latter prevails and the
fiber-reinforcement effect becomes negligible. The bending
stiffness of the beams based on SCC or on FRSCC is thus
similar. Yielding, ductility and load bearing capacity were not
significantly modified by the fiber-reinforcement. This can be
explained by two opposing effects of fibers on the global
mechanical behavior of beams: On the one hand, fibers restrain
the crack opening and thus help to improve both stress transfer
through the crack and tension stiffening but, on the other hand,
fiber reinforcement generates more cracks which counterbalance
the gain of stiffness of the beam.
4.3 Failure mode
Table 7 summarizes the theoretical and experimental values of
ultimate strength for all the elements. The values were calculated
according to the Eurocode 2 model (EN1992-1-1, 2005), based
on the regulatory method of shear design in Ultimate Limit
States. This method uses partial factors equal to 1. The Eurocode
2 model specifies that the shear strength should be calculated as
follows (1):
VRd = VRd,c + VRd,s

(1)

In the case of elements for which no shear reinforcement is


required, the shear resistance of a section in the cracked zone is
calculated according to relation (2).

1/3

0.18 [.k. ( 100.w .fcm ) ].bw d


VRd, c = max
[ 0.035k3/2.fcm1/2].bw d

(2)

The reinforcement ratio rw is calculated using the following


Asl
formula: w = --------- 0, 02 , where : Asl is the cross-section of the
bw .d
longitudinal reinforcement, bw is the beam web width and d is the
200
effective depth. k = 1 + --------- 2 and fcm is the 28-day compressive
d
strength of concrete.
In the case of elements for which shear reinforcement is
required, the shear resistance of a section in the cracked zone is
calculated as follows (3).
Asw
VRd, s = -------.z.f
ywd . cot
s

(3)

is the angle of inclination of the diagonal compressive stress,


z is the effective shear depth, taken as the distance between the
resultants of tensile force and compressive one due to flexure,
Asw is the cross-section of the transverse reinforcement, and fywd is
the yield strength of the shear reinforcement.
First, beams of series A, without shear reinforcement and
with a low percentage of longitudinal bar (0.81%) failed by
flexure as predicted (Fig. 7(a)). The failure mode for both SCC
and FRSCC beams was compressive concrete crushing after the
tension steel yielding. The ultimate load remained unchanged
despite the addition of fibers. The beams of series B failed by
shear as expected following the occurrence of diagonal tension
cracks (Fig. 7(b)). In this test configuration, the fiber reinforcement used did not modify the failure mode or increase the
ultimate shear strength which remained identical to that
predicted for the beam without fiber reinforcement. This result
clearly demonstrates that the type of fibers used, at a rate of
0.25% by volume, cannot efficiently replace the traditional shear
reinforcement. These conclusions are in agreement with the
Eurocode 2 model.
Beams of series C, designed to fail by shear at 130.7 kN,
failed by flexure at 170.5 kN and 170.7 kN respectively for SCC
and FRSCC beams (Fig. 7(c)). The effective ultimate shear
strength appeared to be underestimated by the Eurocode 2
model. These findings can be attributed to the mechanical
contribution of concrete shear strength, which is not taken into
account in the model prediction. Results also showed that the
fiber reinforcement helped to prevent the spalling of concrete

Table 7. Ultimate Loads and Failure Mode of Beams: Model Predictions and Experimental Results
Beams
A-SCC
A-FRSCC
B-SCC
B-FRSCC
C-SCC
C-FRSCC

Calculated ultimate load (kN)


63.8
63.8
111.0
111.0
130.7
130.7

Experimental ultimate load (kN)


63.5
63.1
116.9
111.3
170.5
170.7
1388

Failure mode
Flexure
Flexure
Shear
Shear
Flexure
Flexure
KSCE Journal of Civil Engineering

Youcef Fritih, Thierry Vidal, Anaclet Turatsinze, and Grard Pons

Fig. 7. Mode of Failure of all Beams Tested: (a) Series A Beams, (b) Series B Beams, (c) Series C Beams

Fig. 8. Cracking Distribution in Various Beams Tested


Vol. 17, No. 6 / September 2013

1389

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams

observed before failure.


4.4 Crack Opening
The development of crack opening of the tested beams is
described in Fig. 8.
The crack openings were measured at the most tensioned fiber
of the transverse cross-section of the beams, i.e. at the bottom
fiber. The system used was a video microscope at a magnification of X175 and with an accuracy of 0.02 mm. Crack
evolution is expressed in terms of normalized bending moment.
This is the bending moment M(x) at abscissa x of the crack
considered, divided by the maximum value M(x)max reached,
for the same abscissa, at beam failure. This parameter allows
different type of cracks to be identified and their distribution
reflects the development of the beam damage. Cracks that are
initiated for high M(x)/M(x)max ratios and present low crack
opening correspond to cracks located near a support. Cracks
occurring for lower normalized bending moment and having
large openings are located in the central area of the beam. This
kind of analysis shows that fiber reinforcement reduces the
shape of the diagram. This result is essentially due to the effect
of the fibers on the distribution of the cracks and their
openings.
Figure 9 shows the variations of the average crack openings (a)
and of the cumulated crack openings (b).
Results in Fig. 9 show that for a given normalized bending
moment, the presence of fibers limits the crack opening

significantly. In fact, steel fibers in the tensile zone of the beams


allow the strain in the rebars to be reduced and it is well known
that crack opening is strongly dependent on the strain of the
reinforcement (Oh, 1992, Meda et al., 2005). This effect is less
pronounced in the case of high longitudinal reinforcement ratios.
Increasing the amount of steel reinforcement has the obvious
consequence of reducing reinforcing steel strain and, as
explained above, it results in reduced crack openings, as
evidenced on the corresponding diagrams.
It can therefore be concluded that fiber addition can be a
relevant solution to respect the crack opening limitations at the
Serviceability Limit State for concrete subjected to aggressive
environments in accordance with design codes such as Eurocode
2 (EN 1992-1-1, 2005). In such situations, the use of stainless
fibers that avoid the unsightly rust stains generally observed on
the concrete surface is of obvious interest if the aesthetic
character of the application is a factor to take into account.
4.5 Local Strain Distribution in Steel
4.5.1. Behavior of Longitudinal Reinforcement: Series A
and B
The strain evolution of the longitudinal steel reinforcement
measured at midspan for series A and B is presented in Fig. 10.
The curves highlight the similarity of the longitudinal
reinforcement behavior before crack initiation whether the
concrete is fiber-reinforced or not. However, the type of response
is strongly influenced by the crack path location relative to the
location of the strain gauge. According to the crack pattern
presented in Fig. 5, it appears that the central crack is located on
the midspan gauge for all beams tested. When the first flexure
crack reaches the position of the longitudinal reinforcement, a
change in the steel bar response is recorded. For beams of type A
(Fig. 10), the SCC-based element presents quasi-instantaneous
increase of the strain, 620 m/m, while the strain in the FRSCC
based beam evolves progressively. Fiber reinforcement prevents
this sharp, sudden increase of the strain. This means that the
ability of the fibers to transfer tensile stress through the crack
helps to decrease the tensile stress in the longitudinal reinforcement.

Fig. 9. Crack Openings: (a) Average of Crack Openings, (b) Cumulated Crack Openings
1390

Fig. 10. Midspan Strains of Longitudinal Bars


KSCE Journal of Civil Engineering

Youcef Fritih, Thierry Vidal, Anaclet Turatsinze, and Grard Pons

However, this phenomenon is not observed on beams of type B


(Fig. 10). In this case, when a crack appears, the stress in the steel
bar for high longitudinal reinforcement ratios is lower and, as
explained before, the effect of the fiber reinforcement with low
fiber content becomes of second order.
Once the crack has occurred, the slope of the curve of
normalized moment versus steel bar strain for the type A beam is
steeper when the material is fiber-reinforced. Once again, this is
a corollary of the ability of the fibers to transfer stress across the
crack, which helps to limit the stress in the steel rod and thereby
reduces its deformation. The same causes produce the same
effects; the contribution of the fiber reinforcement is not visible
in the case of beams of series B.
In the case of beams of series A (lightly reinforced, Asl = 3.08
cm2), for a given bending moment in a range between 15 and 35
kN.m, one should observe a decrease of 500 m/m in the strain
due to the contribution of fibers. However, no effect of fiberreinforcement can be seen in the case of beams of series B
(highly reinforced Asl = 8.29 cm2), reported in Fig. 10. This result
can be explained by smaller crack openings that spread more
slowly when the amount of tensile reinforcement is large.
Consequently, the crack opening is small, the bridging effect of
the fibers is not fully activated and the number of fibers bridging
the crack is lower. In such conditions, the normal stress carried
by fibers is limited.
4.5.2. Behavior of Shear Reinforcement: Series C
Figure 11 shows the evolution of local strain of the transverse
reinforcement according to the applied force.

Fig. 11. Strains in the Stirrups: (a) SCC, (b) FRSCC


Vol. 17, No. 6 / September 2013

Fig. 12. Strains at Different Locations (Distances from the Support) of Stirrups Just before Failure of the Beams

Before cracking initiation, the stress level is low and the


subsequent strains are small. The material can be considered as
homogenous, and the transfer of stress is ensured by the steelconcrete bond. As fibers have no significant effect before the
cracking of the material, the deformation is similar whether the
concrete is fiber-reinforced or not. However, after the formation
of a crack, stirrups are tensioned (Barragan, 2002; Furlan and de
Hanai, 1997; Furlan and de Hanai, 1999; Jurez et al., 2007) and
the stress in them increases. It must be noted that the use of
metallic fibers does not delay the contribution of transverse
reinforcement. This phenomenon is relative to the strength of the
concrete, which is almost identical for both formulations whether
they are fiber-reinforced or not. However, the deformations of
transverse reinforcement in the fiber-reinforced beam are
significantly lower than those in the beam without fibers, thanks
to a better stress transfer through the crack, which results in a
higher contribution of concrete to the shear resistance.
It is also worth noting that the values of local deformations of
beams with fibers are less scattered because the control the fibers
exert on the crack openings makes the concrete-steel-fiber
composite more homogenous.
Fiber-reinforced-concrete based beams are characterized by a
higher number of cracks with significantly reduced crack
openings. Additionally, fiber-reinforcement appears to be very
effective in restraining the shear crack.
Figure 12 shows the effect of fiber reinforcement on the
reduction (in %) of shear reinforcement strain at failure level,
where it appears that fibers help to reduce the stirrup deformation.
Because of the failure mode obtained (flexural) and the
formation of few shear cracks crossing stirrups, it was difficult to
evaluate the part of shear stress carried by fibers. According to
the cracking patterns, two cases of steel bar behavior were
distinguished depending on the type of crack.
Case of flexural crack crossing a stirrup
Mostly, flexural cracks occurred at the same location as the
stirrups, where the concrete cover is pre-tensioned before
external loading due to restrained shrinkage. Fig. 13(a) describes
the behavior of the stirrups of two beams having a similar
configuration with regards to flexural cracking (load and crack

1391

Flexural and Shear Behavior of Steel Fiber Reinforced SCC Beams

Fig. 13. Strains in the Stirrups Crossed by a: (a) Flexural Crack, (b)
Shear Crack

path). These flexural cracks were initiated at the same load


(about 30 kN) for both SCC- and FRSCC-based beams and
reached the same level of spread at identical load. The diagram
illustrates the influence of fibers on the bending cracks. Fiber
reinforcement reduces the local deformation of the stirrup under
the effect of the crack, the latter being restrained: at a given load
level, the fibers modify the stress state in a section of concrete
cracked in flexure and inhibit cracking, thus causing a decrease
in shear stresses.
Case of shear crack crossing a stirrup
Figure 13(b) shows the evolution of local deformations of
stirrups n 6 (the 6th position from the support). The same
configuration of the diagonal crack is noted. The inclined crack
crosses these bars at a load level of 100 kN, at equal distance
from the strain gauges. The deformation resulting from this crack
in the FRSCC-based beam is significantly lower than the one
measured on the stirrup of the SCC-based beam. This reduction
is about 45% at failure. During the propagation of inclined cracks,
the fibers are progressively tensioned and their contribution leads
to a significant reduction in the shear force carried by the
transverse reinforcement.

5. Conclusions
This paper has investigated the effect of steel-fiber reinforcement on the behavior of six reinforced Self-Compacting Concrete

beams with a fiber content of 20 kg/m3 (0.25% by volume) under


flexure bending. Low and high amounts of longitudinal steel bars
were used with or without shear reinforcement. The results
reported here lead to the following conclusions:
1. Global mechanical behavior is not affected by the fiber content used. Just a slight enhancement of stiffness after crack
initiation is observed in the case of beams with a low longitudinal reinforcement ratio. This can be explained by the relatively greater tension stiffening effect induced by a better
tensile stress transfer by fibers to the concrete through the
crack opening Yielding, ductility and load bearing capacity
are not improved with the addition of fibers. The fibers do
not modify the failure mode. Fiber contents such as that used
here cannot efficiently replace traditional shear reinforcement.
2. Due to the effect of fibers described above dense crack network
for the fiber-reinforced beam is noticed. It involves built-in tensile stress in the concrete zone located between existing cracks,
generating the occurrence of a new crack when the concrete
tensile strength is reached. Fibers restrain the crack opening,
which is of great interest with regard to design codes such as
Eurocode 2 in the case of aggressive environments. In this case,
the use of stainless fibers is a suitable solution to meet requirements with respect to limitation of crack width and durability.
This type of fiber provides an additional advantage in the sense
that it avoids unsightly rust stains on the concrete surface.
3. The local evaluation of the strain in the longitudinal reinforcement shows that, by controlling the crack opening,
fibers make a tensile mechanical contribution that induces
relaxation in steel bars. This effect appears to be more
marked when the longitudinal reinforcement ratio is low.
4. The mechanical action of the fibers has also been identified
in transverse reinforcement when a bending crack grows
along the stirrup or a shear crack intercepts it. Finally, the
analysis of concrete deformations shows that, for a given
level of loading, the fiber reinforcement changes the stress
state of the concrete in front of the crack tip.
This contribution shows a positive synergetic effect from a
traditional steel-reinforcement and a fiber-reinforcement with a
significant impact on cracking kinetics. However, the effect on shear
behavior is limited because of, firstly, the low fiber content used and,
secondly, the type of fibers, which have a high modulus of elasticity
and a strong bond with the matrix, and are brittle. They do not resist
for crack openings greater than 200 m. These fibers can be added
into the concrete as an additional reinforcement but cannot be used
as a complete substitute for shear bars.
In our ongoing research program, hybrid fiber reinforcements
with various fiber contents are being tested to investigate the
positive synergetic effect between stirrups and fibers of different
natures and properties. The ultimate goal of the program is to
create conditions for reducing the amount of shear reinforcement
without compromising strength and ductility requirements,
particularly with a view to precast applications.

1392

KSCE Journal of Civil Engineering

Youcef Fritih, Thierry Vidal, Anaclet Turatsinze, and Grard Pons

References
Abrishami, H. H. and Mitchell, D. (1997). Influence of steel fibers on
tension stiffening. ACI Structural Journal, Vol. 94, No. 6, pp. 769-776.
Banthia, N. and Trottier, J. (1994). Concrete reinforced with deformed
steel fibres, part I: Bond-slip mechanisms. ACI Materials Journal,
Vol. 91, No. 5, pp. 435-444.
Barragn, B. (2002). Failure and toughness of steel fiber reinforced
concrete under tension and shear, PhD Thesis, Universitat Politecnica de Catalunya, Spain.
Bayasi, M. Z. and Soroushian, P. (1992). Effect of steel fiber reinforcement on fresh mix properties of concrete. ACI Materials
Journal, Vol. 89, No. 4, pp. 369-374.
Beaudoin, J. J. (1990). Handbook of fiber reinforced concrete, principles,
properties, Developments and Applications, Noyes Publications,
Park Ridge, NJ, USA.
Bischoff, P. H. (2000). Comparison of tension stiffening of plain and
steel fiber reinforced concrete. Fiber-Reinforced Concrete (FRC)
BEFIP 2000, 5th RILEM Symposium, Lyon, Septembre 2000, RILEM
publication S.A.R.L Cachan-cedex, In: Rossi P. and Chanvillard G.
Eds., pp. 633-641.
Bischoff, P. H. (2003). Tension stiffening and cracking of steel fiberreinforced concrete. Journal of Materials in Civil Engineering, Vol.
15, No. 2, pp. 174-182.
Choulli, Y., Mari, A. R., and Cladera, A. (2008). Shear behavior of fullscale prestressed i-beams made with self-compacting concrete.
Materials and Structures, Vol. 41, No. 1, pp. 131-141.
Cucchiara, C., La Mendola, L., and Papia, M. (2004). Effectiveness of
stirrups and steel fibres as shear reinforcement. Cement and
Concrete Composites, Vol. 26, No. 7, pp. 777-786.
Edgington, J., Hannant, D. J., and Williams, R. I. T. (1978). Steel fibre
reinforced concrete, fibre reinforced materials. Practical Studies
from the Building Research Establishment, The Construction Press,
Lancaster, pp. 112-128.
EN 1992-1-1. (2005). Eurocode 2: design of concrete structures, Part 1-1:
General Rules and Rules for Buildings.
EN 206-9. (2010). Concrete part 9: Additional rules for self-compacting concrete, AFNOR.
Ezeldin, A. and Balaguru, P. (1989). Bond behavior of normal and
high-strength fibre reinforced concrete. ACI Materials Journal,
Vol. 86, No. 5, pp. 515-524.
Frosch, R. J. (2000). Behavior of large-scale reinforced concrete beams
with minimum shear reinforcement. ACI Structural Journal, Vol.
97, No. 6, pp. 814-820.
Furlan, S. and de Hanai, J. B. (1997). Shear behavior of fibre reinforced
concrete beams. Cement and Concrete Composites, Vol. 19, No. 4,
pp. 359-366.
Furlan, S. and de Hanai, J. B. (1999). Prestressed fiber reinforced
concrete beams with reduced ratios of shear reinforcement. Cement
and Concrete Composites, Vol. 21, No. 3, pp. 213-221.
Hameed, R., Turatsinze, A., Duprat, F., and Sellier, A. (2010a). Study on
the flexural properties of metallic-hybrid-fibre reinforced concrete.
Maejo International Journal of Science and Technology, Vol. 4, No.
2, pp. 169-184.
Hameed, R., Turatsinze, A., Duprat, F., and Sellier, A. (2010b). A study on
the reinforced fibrous concrete elements subjected to uniaxial tensile
loading. KSCE Journal of Civil Engineering, KSCE, Vol. 14, No. 4, pp.
547-556.

Vol. 17, No. 6 / September 2013

Jurez, C., Valdez, P., Durn, A., and Sobolev, K. (2007). The diagonal
tension behavior of fiber reinforced concrete beams. Cement and
Concrete Composites, Vol. 29, No. 5, pp. 402-408.
Khuntia, M. and Stojadinovic, B. (2001). Shear strength of reinforced
concrete beams without transverse reinforcement. ACI Structural
Journal, Vol. 98, No. 5, pp. 648-56.
Lim, D. H. and Oh, B. H. (1999). Experimental and theoretical
investigation on the shear of steel fibre reinforced concrete beams.
Engineering Structures, Vol. 21, No. 10, pp. 937-944.
Meda, A., Minelli, F., Plizzari, G., and Riva, P. (2005). Shear behavior
of steel fibre reinforced concrete beams. Materials and Structures,
Vol. 38, No. 227, pp. 343-351.
Mirsayah, A. and Banthia, N. (2002). Shear strength of steel fiberreinforced concrete. ACI Materials Journal, Vol. 99, No. 5, pp.
473-479.
Mitchell, D., Abrishami, H. H., and Mindess, S. (1996). The effect of
steel fibers and Epoxy-Coated reinforcement on tension stiffening
and cracking of reinforced concrete. ACI Materials Journal, Vol.
93, No. 1, pp. 61-68.
Narayanan, R. and Darwish, I. Y. S. (1987). Use of steel fibers as shear
reinforcement. ACI Structural Journal, Vol. 84, No. 3, pp. 216226.
Oh, B. H. (1992). Flexural analysis of reinforced concrete beams
containing steel fibres. Journal of Structural Engineering, Vol. 118,
No. 10, pp. 2821-2836.
Pons, G., Mouret, M., Alcantara, M., and Granju, J. L. (2007). Mechanical
behavior self-compacting concrete with hybrid fibre reinforcement. Materials and Structures, Vol. 40, No. 2, pp. 201-210.
Redon, C. and Chermant, J. L. (1999). Damage mechanics applied to
concrete reinforced with amorphous cast iron fibers, concrete
subjected to compression. Cement and Concrete Research, Vol. 21,
No. 3, pp. 197-204.
Rossi, P. (1991). Formulation et comportement mcanique des btons
de fibres mtalliques. Annales de l'ITBTP, N492, Srie Bton, Vol.
279, pp. 90-107 (in French).
Shin, S. W., Oh, J. G., and Ghosh, S. K. (1994). Shear behavior of
laboratory-sized high strength concrete beams reinforced with bars
and steel fibers. American Concrete Institute, Special Publication,
Vol. 142, No. 10, pp. 181-200.
TC 162-TDF Rilem. (2001). Test and design methods for steel fibre
reinforced concrete: Uniaxial tension test for steel fibre reinforced
concrete. Materials and Structures, Vol. 34, No. 235, pp. 3-6.
Turatsinze, A., Farhat, H., and Granju J. L. (2003). Influence of autogenous
cracking on the durability of repairs by cement-based overlays
reinforced with metal fibres. Materials and Structures, Vol. 36, No.
10, pp. 673-677.
Turatsinze, A., Granju, J. L, Sabathier, V., and Farhat, H. (2005).
Durability of bonded cement-based overlays: Effect of metal fibre
reinforcement. Materials and Structures, Vol. 38, No. 3, pp. 321327.
You, Z., Ding, Y., and Niederegger, C. (2010). Replacing stirrups of
self-compacting concrete beams with steel fibers. Transactions of
Tianjin University, Vol. 16, No. 6, pp. 411-416.
Yun, H. D., Yang, I. S., Kim, S. W., Jeon, E., Choi, C. S., and Fukuyama,
H. (2007). Mechanical properties of high-performance hybridfibre-reinforced cementitious composites (HPHFRCCs). Magazine
of Concrete Research, Vol. 59, No. 4, pp. 257-271.

1393

Das könnte Ihnen auch gefallen