Sie sind auf Seite 1von 14

European Journal of Environmental and Civil Engineering

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

Ductility behaviours of oil palm shell steel fibre-


reinforced concrete beams under flexural loading

Soon Poh Yap, U. Johnson Alengaram, Kim Hung Mo & Mohd Zamin Jumaat

To cite this article: Soon Poh Yap, U. Johnson Alengaram, Kim Hung Mo & Mohd Zamin
Jumaat (2017): Ductility behaviours of oil palm shell steel fibre-reinforced concrete beams
under flexural loading, European Journal of Environmental and Civil Engineering, DOI:
10.1080/19648189.2017.1320234

To link to this article: http://dx.doi.org/10.1080/19648189.2017.1320234

Published online: 24 Apr 2017.

Submit your article to this journal

Article views: 19

View related articles

View Crossmark data

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


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

Download by: [University of Malaya] Date: 08 May 2017, At: 23:16


European Journal of Environmental and Civil Engineering, 2017
https://doi.org/10.1080/19648189.2017.1320234

Ductility behaviours of oil palm shell steel fibre-reinforced


concrete beams under flexural loading
Soon Poh Yap, U. Johnson Alengaram, Kim Hung Mo and Mohd Zamin Jumaat
Faculty of Engineering, Department of Civil Engineering, University of Malaya, Kuala Lumpur, Malaysia

ABSTRACT ARTICLE HISTORY


In the effort of developing sustainable concrete, studies on the replacement Received 9 September 2016
of conventional coarse aggregates with local waste material are increasing. Accepted 30 March 2017
The new oil palm shell fibre-reinforced concrete (OPSFRC) utilises Malaysia’s
KEYWORDS
waste material and is emerging as a promising sustainable concrete Deflection; fibre-reinforced
attributed to its lightweight and enhanced concrete properties. In this concrete; lightweight
study, the flexural behaviours of OPSC and OPSFRC beams (with steel fibres concrete; oil palm shell; steel
.25, .50, .75 and 1.00% by volume) were investigated. The results show that fibre
the addition of steel fibres improved both the mechanical properties and
moment capacities of OPSFRC specimens. However, the OPSFRC exhibited
reduced deflection at failure of about 30–40% relative to the OPSC beams.
Meanwhile, the steel fibres significantly improved the failure load to ultimate
load ratios of OPSFRC to .65–.80 compared to the ratio of .19 as in OPSC. In
addition, different approaches to compute the ductility ratios of OPSFRC
beams are also conducted and compared.

1. Introduction
The newly developed concrete known as the oil palm shell fibre-reinforced concrete (OPSFRC) has
gained substantial research interests for the past five years, attributed to its enhanced mechanical
properties compared to the normal weight concrete (NWC) and oil palm shell concrete (OPSC) (Shafigh,
Mahmud, & Jumaat, 2011; Yap et al., 2015a; Yap, Bu, Alengaram, Mo, & Jumaat, 2014; Yap, Alengaram,
Jumaat, & Khaw, 2015b; Yew, Mahmud, Ang, & Yew, 2015). The combined advantages of oil palm shell
(OPS) and fibre (Figure 1) enabled the production of OPSFRC as high-strength sustainable lightweight
concrete.
Reports by Teo, Mannan, and Kurian (2006) and Alengaram, Jumaat, and Mahmud (2008) revealed
that the flexural behaviours of reinforced OPSC beams were comparable to that of NWC and other LWCs
containing pumice and expanded clay. However, like other LWCs, the low tensile strength of OPSC has
limited its structural applications (Yap, Alengaram, & Jumaat, 2013; Yap, Khaw, Alengaram, & Jumaat,
2015c). Hence, the incorporation of fibres in OPSC has been proven to enhance its mechanical properties,
torsional behaviours, impact resistance and toughness (Mo, Yap, Alengaram, Jumaat, & Bu, 2014; Yap et
al., 2015a; Yew et al., 2015). Despite that, the development of fibre-reinforced LWCs including OPSFRC
remains as a new and challenging area of research, attributed to the types of lightweight aggregates
and fibres used (Gribniak, Kaklauskas, Hung Kwan, Bacinskas, & Ulbinas, 2012; Hassanpour, Shafigh, &
Mahmud, 2012; Khelifa, Leklou, Bellal, Hebert, & Ledesert, 2016; Yap et al., 2013). The great diversity in

CONTACT  Soon Poh Yap  spyap@um.edu.my


© 2017 Informa UK Limited, trading as Taylor & Francis Group
2    S. P. YAP ET AL.

Figure 1. (a) Oil palm shell and (b) steel fibre.

the type and geometry of fibres yields varying bond characteristics, thus limiting the application of
fibre-reinforced concrete in structural members (Gribniak et al., 2012; Marthong & Sarma, 2016; Toraldo,
Mariani, & Crispino, 2016). Furthermore, the flexural behaviours of fibre-reinforced LWCs become more
complicated with different types of lightweight aggregate used.
From the literature, improvement in flexural performances of fibre-reinforced concrete beams has
been observed, such as reduced brittleness and improved crack resistance of fibre-reinforced concrete
beam attributed to the enhanced crack resistance (Altun & Aktaş, 2013; Altun, Haktanir, & Ari, 2007;
Meda, Minelli, & Plizzari, 2012; Qian & Indubhushan, 1999; Wang & Belarbi, 2011; Yang, Min, Shin, &
Yoon, 2012). This is supplemented by the delayed initiation of flexural cracks, decreased crack widths
and doubled first crack load in the flexural beam test conducted by Yang et al. (2012). Moreover, the
addition of steel fibre into the reinforced concrete system improved flexural rigidity, yield load, moment
capacity and toughness of reinforced concrete beam (Altun et al., 2007; Mahmud, Yang, & Hassan, 2013;
Qian & Indubhushan, 1999; Wang & Belarbi, 2011; You, Chen, & Dong, 2011). Fibres significantly enhance
the behaviour at service conditions by increasing the stiffness in the cracked stage and, therefore, by
limiting the crack openings and deformations (Meda et al., 2012).
However, the main challenge in this study lies on the contradictory observations on the flexural
ductility. In the plain concrete without reinforcement, steel fibres improved the tensile strength, post
cracking behaviours and ductility (Ng, Foster, Htet, & Htut, 2013; Yap et al., 2014). Meda et al. (2012)
and You et al. (2011) had demonstrated that the fibre reinforcement can lead to a reduction in the
rotation capacity and sectional ductility of reinforced concrete members. On the other hand, Altun
and Aktaş (2013) shows the addition of steel fibre increased the flexural ductility by 9–18%. Qian and
Indubhushan (1999) and Wang and Belarbi (2011) also reported that steel fibres can increase displace-
ment of beams at failure.
To date, there are no reports available on the flexural behaviours of OPSFRC beams. Therefore,
the main objective of this study is to investigate the effects of steel fibre in the flexural behaviours of
OPSFRC reinforced beams. This study compares the flexural ductility characteristics (i) between OPSC
and OPSFRC beams and (ii) OPSFRC with different fibre volumes (.25, .50, .75 and 1.00%).

2.  Experimental program


2.1. Materials
Type 1 Ordinary Portland cement with Blaine-specific surface area and specific gravity of the cement
are 3450 cm2/g and 3.13, respectively, was used. Silica fume (SF) of 10% cement weight was added to
improve the mechanical properties of all mixes. Meanwhile for the aggregates, lightweight OPS was col-
lected from a local oil palm factory. OPS were used as coarse aggregates to replace conventional granite
aggregates. Figure 1(a) shows the OPS with different sizes and concave/convex surfaces. The distinct
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   3

Table 1. Mix proportions.

Mix designa- Cement Mining sand Water Silica fume Superplasti- Steel fibre
tions (kg/m3) OPS (kg/m3 (kg/m3) (kg/m3) (kg/m3) zer (kg/m3) (%vol.)
OPSC 530 320 970 170 53 3.5 0
OPSFRC-25 530 320 970 170 53 3.5 .25
OPSFRC-50 530 320 970 170 53 3.5 .50
OPSFRC-75 530 320 970 170 53 3.5 .75
OPSFRC-100 530 320 970 170 53 3.5 1.00

properties of OPS which enable OPS as a potential material to produce lightweight green concrete
are its low bulk density and good impact/abrasion resistance. The compacted bulk density, aggregate
impact value and Los Angeles abrasion value of OPS are 635 kg/m3, 2.11 and 5%, respectively. Before
the mixing process, the OPS below 2.36 mm was sieved and removed, while the OPS with sizes between
2.36–15 mm was used in all the mixing proportions. In addition, all OPS were soaked in water for 24 h
before the mixing process to keep the OPS in saturated surface dry condition and thus preventing the
mixing water to be absorbed by the OPS. Meanwhile, mining sand with specific gravity and fineness
modulus of 2.65 and 2.7, respectively was used as fine aggregates. Potable water (pH = 6.5) was used
in the pre-treatment of OPS, mixing and curing processes. A constant water to binder ratio of .30 was
used in all the mixing proportions. To improve the workability of the fresh OPSC and OPSFRC mixes, a
polycarboxylate-based superplasticizer of .65% of cement weight was added.
In order to produce OPSFRC mixes, hooked-end steel fibres of aspect ratio 65 (Figure 1(b)) were
added into OPSC. The length and diameter of steel fibre are 35 and .55 mm, respectively. Specific gravity,
tensile strength and modulus of elasticity of the steel fibres are 7.9, 1100 MPa, and 205 GPa, respectively.

2.2.  Specimen preparation and testing


The mix proportions of all mixes were shown in Table 1. The parameter studied is the fibre volume
while the amount of other materials was kept constant. For each mix proportion, mechanical properties
and flexural beam testing were conducted. For the testing of the mechanical properties of each mix
proportion,100 mm cubes, 100ϕ × 200 mm cylinders, 150ϕ × 300 mm cylinders, 100 × 100 × 500 mm
prisms and 100 × 100 × 300 mm prisms were prepared for compressive strength (BS EN 12390: Part 3,
2009), splitting tensile strength (ASTM C496/ C496M, 2011), modulus of elasticity/Poisson’s ratio (ASTM
C469, 2010), flexural strength (ASTM C78/ C78M, 2010) and flexural ductility/toughness (ASTM C1018,
1997), respectively. All the specimens are cured in water until the age of testing. The cube specimens
were tested at the age of 1, 3, 7, and 28-day, while other specimens including the beams were tested
at the age of 28 days.
For the reinforced concrete beam test, a total of 10 beams were designed and prepared as under-re-
inforced concrete beams in accordance to the BS 8110. Two beams were tested for each mix design as
given in Table 1. The reinforcement arrangement for all beams is shown in Figure 2. Ribbed steel bars
with diameter 12 mm and plain steel bars with diameter 6 mm were used as compression/tension
and shear reinforcement, respectively. A clear cover of 30 mm was used. In the pure bending region
(middle-third of the beam), the compression and shear reinforcements were not used, while the shear
reinforcement was used only in the shear span (.7 m from each support) at a spacing of about 75 mm
centre to centre to ensure yielding of tension steel occurs before the concrete crushing. The flexural
testing of the beam specimens was conducted using an Instron universal testing machine with a built-in
load cell capacity of 600 kN as illustrated in Figure 3.

3.  Results and discussions


3.1.  Fresh and hardened properties
The fresh properties of OPSC and OPSFRC were measured by slump cone test and the slump values of
all mixes are reported in Table 2. The slump of OPSC was 80 mm and the addition of .25–1.00% steel
4    S. P. YAP ET AL.

Figure 2. Reinforcement details of flexure beams (all dimensions are in mm).

Figure 3. Flexural test set-up (all dimensions are in mm).

Table 2. Fresh and hardened properties of OPSC and OPSFRC mixes.

Flexural Modulus of Poisson’s


Compressive strength (MPa) strength (MPa) Brittleness* elasticity (GPa) ratio
Slump
Mix (mm) 1-day 3-day 7-day 28-day 28-day 28-day 28-day 28-day
OPSC 80 18.7 25.6 33.0 33.9 3.26 10.40 13.87 .244
OPSFRC-25 60 21.3 30.4 38.7 39.3 4.28 9.17 15.67 .286
OPSFRC-50 50 22.1 31.7 39.7 41.2 5.38 7.66 16.11 .294
OPSFRC-75 40 22.7 33.6 44.6 45.9 6.70 6.85 16.29 .275
OPSFRC-100 20 25.7 35.0 45.1 47.3 8.16 5.80 15.50 .292
*
Brittleness = compressive to flexural strength ratio.

fibres decreased the slump by about 30–70%. As the fibre content increases, lower slump was reported
in the OPSFRC mix. The slump reduction in the OPSFRC mixes can be explained by the fibre–cement
matrix network in the fresh concrete. The network increases the concrete viscosity and substantially
restricts the flow of the fresh OPSFRC (Domagała, 2011; Yap et al., 2015a; Yap, Alengaram, Jumaat, &
Khaw, 2016). All OPSFRC mixes can be compacted with good finishing. However, the fibre content
cannot be further increased beyond 1% due to the poor compaction.
Previous studies on OPSFRC with steel fibres proved that the addition of fibres of different aspect
ratios enhanced the mechanical properties of OPSFRC (Yap et al., 2015a; Yap et al., 2015c). In this work,
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   5

the effect of fibre volume on the mechanical properties of OPSFRC is discussed. The mechanical prop-
erties of all mixes are tabulated in Table 2. Firstly, the development of the compressive strength of
OPSC and OPSFRC showed comparable trend. The 1, 3 and 7-day compressive strength of all mixes
were 50–55, 73–77 and 95–98%, respectively, compared to the 28-day compressive strength. The high
7 to 28-day compressive strength ratios indicate that the incorporation of silica fume contributed to
the high early strengths in all the mixes.
The results show that the addition of steel fibres up to 1% by volume resulted in improved mechan-
ical properties of OPSFRC, compared to the OPSC. In addition, the mechanical properties of OPSFRC
show increasing trends with increasing fibre volume. Therefore, the highest compressive and flexural
strengths are found in the OPSFRC-100 mix with 1% steel fibres at 47.3 and 8.2 MPa, respectively and
both these values are 40 and 150% higher than the control OPSC. The addition of steel fibres below
.50% only produced improvements in compressive and flexural strength within the range of 16–21 and
31–65%. However, the incorporation of fibre volume beyond .50% yields significant improvement of
compressive and flexural strength up to 105–150%. This indicates that a minimum fibre volume of .5%
is required to produce notable improvement on the strengths of OPSFRC specimens. Meanwhile, the
benefits of steel fibres are also evident in the modulus of elasticity (MOE) and Poisson’s ratio as shown in
Table 2. MOE was measured as stress to longitudinal strain ratio, while Poisson’s ratio was measured as
lateral to longitudinal strain ratios under compression loading. Previous studies reported that the MOE
values of OPSC are about 35–50% lower than that of NWC (Alengaram, Mahmud, & Jumaat, 2011; Yap et
al., 2016). The reason is that OPS has a lower stiffness and restraining effect. Hence, under compressive
loading, the OPS undergoes higher strain, which produces a lower MOE than the NWC.
Both the MOE and Poisson’s ratio of OPSFRC specimens surpassed the OPSC mix by 11–18%. This
indicates that the addition of steel fibres allowed the OPSFRC to sustain higher strains in both lon-
gitudinal and lateral directions. However, the improvements are independent on the fibre content
and this can be explained by the equal increase in both upper stress and strain which resulted in the
minimal change in the modulus of elasticity when fibre volume increases (Yap et al., 2016). In addition,
the porous lightweight aggregates leads to high brittleness of LWC and can result in a complete and
immediate loss of load carrying capacity once the tensile stress capacity is reached (Domagała, 2011;
Yap et al., 2015a). The brittleness of LWC including OPSC needs to be measured, and a direct indication
will be the brittleness ratios (compressive to flexural strength ratios) (Sun & Xu, 2009; Yap et al., 2015a;
Yap et al., 2016). In this study, the brittleness ratios for both OPSC and OPSFRC are reported in Table 2.
The brittleness ratio of OPSC was reported to be 10.40. The brittleness of OPSC was reduced by 12–45%
when fibre is added to the OPSC mix.
The enhancement mechanism of the steel fibres in the mechanical properties of OPSFRC can be
explained by the crack bridging effect. Figure 4(c) and (d) shows images of fibre bridging across the
beam cracks under flexural loadings taken using hand-held digital microscope. Additional energy is
required to overcome the fibre-cement matrix interfacial bond by both fibre debonding and fibre
pullout processes. The crack propagation in the OPSFRC is diverted, blunted or even stopped, thus
increasing the energy capacity of the concrete (Hamoush, Abu-Lebdeh, & Cummins, 2010; Makita &
Brühwiler, 2013; Şahin & Köksal, 2011; Singh, Shukla, & Brown, 2004; Sun & Xu, 2009; Yap et al., 2014).

3.2.  Moment capacities


The moment–deflection curves for all the specimens are presented in Figure 5. Based on Figure 5,
the plain OPSC beam produced an ultimate moment of 27.7 kNm. As expected, the addition of steel
fibres improved the ultimate moment capacity of the OPSFRC beams by 4–18%. The increase in the
fibre content tends to increase the moment capacity in all OPSFRC beams as reflected in Figure 5. The
highest ultimate moment of 32.6 kNm was obtained for the OPSFRC-100 beam with 1% steel fibres
and it was 18% (or about 5 kNm) higher than the ultimate moment of control OPSC beam. In Section
3.1, the crack bridging effect of steel fibres improved the mechanical properties of OPSFRC specimens
and similar explanation is applicable for the enhanced moment capacity of OPSFRC as well.
6    S. P. YAP ET AL.

Figure 4. Comparison of cracks between (a) OPSC and (b) OPSFRC-100 beams; (c) & (d) crack bridging effects of steel fibres.

35
OPSC
30 OPSFRC-25
OPSFRC-50
25 OPSFRC-75
Moment (kNm)

OPSFRC-100
20

15

10

0
0 10 20 30 40 50 60 70 80 90 100
Deflection (mm)

Figure 5. Bending moment–deflection curves for flexure beams.

In addition, steel fibre reinforcement of volume fraction lower than .50% produced slight improve-
ment on the moment capacity of OPSFRC specimen (increment within the range of 4–8%), whereas
steel fibre addition of volume higher than .50% produced significant increment of 16–18% compared
to the plain OPSC beam. This shows that a minimum fibre volume of .50% is required to provide notable
improvement on the moment capacity of OPSFRC. This can be further supplemented by the discussions
in the Section 3.1, where a fibre content of more than .50% is essential to enhance the tensile strength
of OPSFRC specimens. In the case of OPSFRC beams, the fibres were randomly dispersed in both the
tension and compression zones. The fibre reinforcement enhanced the compression and tension tough-
ness, hence substantially improved the moment capacity of the OPSFRC beams.
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   7

3.3.  Ductility characteristics


3.3.1. Strains
Figure 6 shows the concrete (compressive) and steel (tensile) strains of all beam specimens. The max-
imum tensile strains are within the range of 16000–18000 × 10−6. Meanwhile, the compressive strains
within the service moments lie within the range of 550–900 × 10−6. The comparison between the OPSC
and OPSFRC shows that the steel fibre increased the moment capacity of OPSFRC specimens at a given
strain (in both compressive and tensile strains). This can be explained by the crack bridging effect. The
strong fibre–matrix interfacial bond takes up a portion of the applied loading and hence improving
the load capacity of concrete. In addition, the strong fibre-matrix interfacial bond also caused the
OPSFRC beams to reach maximum moment at a lower strain value compared to the OPSC beams. The
OPSC specimen achieved maximum moment capacity at the steel strain value of about 4500 × 10−6,
but the OPSFRC attained the respective maximum moment capacities at lower steel strains of about
2500–3500 × 10−6. Further, lower steel strains are observed in OPSFRC beams with higher steel fibres
contents. This scenario could be explained by the confinement effect caused by the steel fibres. The pres-
ence of the strong fibre–matrix interfacial bond (as shown in Figure 4(c) and (d)) holds the s­ urrounding
concrete together and this causes a confinement effect which substantially leads to stronger the rebar-
to-concrete bond (Peled, Zaguri, & Marom, 2008). The improved rebar-to-concrete bond does not allow
the yielded rebar to develop the plastic deformation that happens in the reinforced concrete beams
without fibre. This leads to strain localisation effect, at which, maximum moments in OPSFRC beams
occurs with increasing fibre content at lower strains compared to OPSC beams. Similar finding was
reported for NWC beams with fibre (du béton, 2000; Meda et al., 2012).

3.3.2.  Mid span deflection


The deflections of all beam specimens at different loadings are shown in Table 3. Based on the results
on deflections as shown in Table 3, it can be seen that the OPSC produced a deflection of about 93 mm
prior to failure. This result is very close to the deflection reported in the previous study (Alengaram
et al., 2008). However, the addition of steel fibres up to 1% by volume reduced the deflection at failure
of OPSFRC beams by about 30–40%. The deflections prior to failure of the OPSFRC beams fall within
the range of 55–66 mm. Furthermore, the increase in volume fraction of steel fibres resulted in lower
deflection. In the case of OPSFRC, the OPSC produced strong bond properties and the combined effects
of strong aggregate–matrix and fibre–matrix bonds contributed to a reduced ductility in OPSFRC beams.
The confinement effect of steel fibres can be further supported by Figure 4. As observed in Figure 4,
the OPSC shows wide and deep flexural cracks with the formation of large concrete crushing wedge.
Meanwhile, the confinement effect of steel fibres causes the OPSFRC-100 beam to fail at a smaller
deflection before the cracks and concrete crushing can fully develop, as in the OPSC beam.

Figure 6. (a) Concrete and (b) steel strains.


8    S. P. YAP ET AL.

Table 3. Deflections of OPSC and OPSFRC beams at different loading stages.

Ultimate Failure
  Service load stage stage stage
Deflection Deflection Deflection Deflection Deflection at Deflection Deflection
at low ser- at yield at ultimate at moder- Deflection service load at moment prior to
vice load load, δyld load, δult ate service at service predicted by capacity failure, δf
Beam (mm) (mm) (mm) load (mm) load (mm) BS 8110 (mm) (mm) (mm)
OPSC 2.70 8.12 12.10 4.90 7.65 7.94 12.10 92.80
OPSFRC-25 1.07 7.14 11.16 3.17 5.97 8.55 11.16 66.06
OPSFRC-50 .85 9.98 15.28 2.92 6.08 8.88 15.28 62.97
OPSFRC-75 .57 7.71 10.56 2.90 5.80 9.91 10.56 60.54
OPSFRC-100 .91 12.04 18.92 5.05 8.81 10.53 18.92 55.13

Table 4. Span to deflection ratio and load at 50 mm deflection of OPSC and OPSFRC beams.

Span to service load deflection Load at 50 mm deflection,


Beam ratio P50 mm (kN) P50 mm/Pult Toughness (kNm)
OPSC 22.6 55.64 .70 4.291
OPSFRC-25 31.8 73.28 .89 4.589
OPSFRC-50 33.3 78.29 .92 4.435
OPSFRC-75 34.7 83.12 .91 4.287
OPSFRC-100 38.1 84.57 .91 4.116

Similar observations could be drawn at other stages including yield (when rebar attains yield strain),
low service load (20% of ultimate load), moderate service load (40% of ultimate load), service load (BS
8110) and at ultimate load. The service load stages are selected according to the report of Teo et al.
(2006). In order to fulfil the serviceability limited state as outlined in both BS 8110 and Eurocode 2, the
span to service load deflection should not exceed 250. In addition, ASTM C1609 stated that the max-
imum span to service load deflection ratio is 150. Based on the results shown in Table 4, the span to
service load deflection ratios for OPSC and OPSFRC beams were found to be 23 and 30–40, respectively.
These values are well below the deflection limit stipulated in BS 8110, Eurocode 2 and ASTM C1609.
The deflections at service load for all beam specimens were also calculated based on curvature of
the beams as proposed in BS 8110 using Equation (1).
𝛿 = Kl 2 (1∕r) (1)
Where δ = deflection in mm, K = a constant depends on the distribution of bending moments of a
member, l = effective span of beam and 1/r = curvature. All the predicted deflections at service load
are shown in Table 3. It was observed that the experimental deflection of OPSC closely agreed to the
predicted deflection. However, the calculated deflections using the Equation (1) for the OPSFRC beams
are 20–80% higher than the values obtained experimentally. This is because BS 8110 code provisions do
not include the effect of steel fibres. The difference between calculated and experimental deflections
increased with the increasing steel fibre volume.
Despite the decrease in the flexural deflection, the role of fibres in post-yielding flexural behaviour
of OPSFRC beams was evident after the yielding of steel bars. According to the deflection data in Table
3, the difference between the deflection at the yield stage and the ultimate stage is found to be 4 mm
in the OPSC beam. However, in the OPSFRC beams, this margin is slightly extended to about 6–7 mm
irrespective of the volume fraction of steel fibres. After the steel bar yields, the load is not fully sustained
by the reinforcement but transferred from the bars to the concrete to achieve a new equilibrium (Hsu,
1968). The fibre–matrix interfacial bond in OPSFRC beams increased the load capacity of the concrete
thus improved the toughness within the yielding and ultimate stages.
For the post-ultimate load stage, the addition of steel fibres improved the ultimate compressive strain
which resulted in softened post-peak behaviours (Bencardino, Rizzuti, Spadea, & Swamy, 2008; Yang et
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   9

al., 2012). It can be observed in Figure 4 that the addition of steel fibres produced a flattened curve for
the post-peak behaviours of the OPSFRC beams, while the OPSC exhibited a steep descending post-
peak curve. One of the indicative parameters to reveal the effect of steel fibre in post-peak behaviours
of concrete is the failure load to ultimate load ratio. The failure load to ultimate load ratios for OPSC,
OPSFRC-25, OPSFRC-50, OPSFRC-75 and OPSFRC-100 were .19, .69, .71, .66, .80 and .72, respectively.
When the OPSC beam failed, the beam achieved only 19% of its maximum load. Whereas for the OPSFRC
beams, the values of 66–80% of their ultimate load capacity show that steel fibre reinforcement signif-
icantly enhanced the post-ultimate load behaviours of OPSFRC reinforced members.
The post-peak behaviours of both OPSC and OPSFRC could also be compared using the P50/Pult ratio
for the load at constant deflection of 50 mm, as shown in Table 4. The deflection of 50 mm was chosen
as it was closest to the deflection prior to failure in OPSFRC-100 beam. The ratio of P50/Pult for OPSC
beam was found to be about 70% while higher value of about 90% was obtained for the OPSFRC beams.
This could be attributed to the crack bridging effect of steel fibre that requires additional energy to
pull the fibre out from the fractured cement matrix. This post-peak tensile stiffening mechanism sub-
stantially increases the post-cracking strength and ductility (Abu-Lebdeh, Hamoush, Heard, & Zornig,
2011; Hamoush et al., 2010; Okay & Engin, 2012; Singh et al., 2004).

3.3.3.  Ductility ratio


Flexural ductility can be defined based on a reference state, which can be obtained from the yield
point of internal reinforcements. The unique yield plateau in the stress–strain curve of steel makes it
possible for the structural member to be able to resist load while undergoing large deformations (Kim &
Shin, 2011). Ductility of reinforced concrete is generally measured by ductility ratio. High ductility ratio
indicates that a structural member is capable of undergoing large deflections prior to failure (Teo et al.,
2006). There are two main approaches to define the ductility: energy-based approach and deforma-
tion-based approach (Wang & Belarbi, 2011). The frequently utilised approach is the deformation-based
approach which is reflected by the deformation margin between the ultimate stage and service stage
(Alengaram et al., 2008; Teo et al., 2006; Wang & Belarbi, 2011). Equation (2) shows the calculation of
ductility ratio based on the deformation approach.
𝜇D = 𝛿f ∕𝛿yld (2)
where μD = deformation-based ductility ratio, δf = final mid-span deflection and δyld = mid-span deflec-
tion when tensile steel yields.
Jaeger, Tadros, and Mufti (1997) introduced Equations 3 to 5 to consider both the strength and
deflection on the ductility characteristics.
𝜇D2 = Cs × Cd (3)
where μD2 = ductility ratio by Jaeger et al. (1997), Cs = strength factor calculated from Equation (4) and
Cd = deflection factor calculated from Equation (5).
Cs = Mu ∕ M𝜀 .001 (4)

Cd = 𝛿u ∕ 𝛿𝜀 .001 (5)
where Mu  =  moment capacity, Mε.001  =  moment when tensile steel strain is .001, δu  =  deflection at
moment capacity and δε.001 = deflection when tensile steel strain is .001.
Meanwhile, in energy-based approach, ductility can be defined as a capacity of energy absorption
(Kim & Shin, 2011; Wang & Belarbi, 2011). Equations 6 and 7 show two examples to calculate the ener-
gy-based ductility ratio reported by Kim and Shin (2011).
𝜇E1 = AMu ∕AMyld (6)
where μE = energy-based ductility ratio up to ultimate moment, AMu = area of load–displacement curve
up to moment capacity and AMyld = area of load–displacement curve up tensile steel yields.
10    S. P. YAP ET AL.

Table 5. Comparison of ductility ratios of OPSC and OPSFRC beams.

Ratio to the con- Ratio to the con- Ratio to the con- Ratio to the con-
Beam μD trol specimen μD2 trol specimen μE1 trol specimen μE2 trol specimen
OPSC 11.43 – 8.51 – 8.96 – 4.71 –
OPSFRC-25 9.25 .81 11.66 1.37 9.30 1.04 5.70 1.21
OPSFRC-50 6.31 .55 12.11 1.42 6.34 .71 5.32 1.13
OPSFRC-75 7.85 .69 8.74 1.03 8.84 .99 4.18 .89
OPSFRC-100 4.58 .40 9.75 1.15 4.45 .50 4.72 1.00

𝜇E2 = AM.75f ∕AMyld (7)


where μE2 = energy-based ductility ratio up to 75% of post-peak load, AMu = area of load–displacement
curve up to 75% of post-peak load and AMyld = area of load–displacement curve up tensile steel yields.
The comparison between different ductility ratios of the OPSC and OPSFRC beams is reported in
Table 5. The ductility ratio of OPSC (μD) obtained in this work is about 11.4 and is higher than the previ-
ously published data (Alengaram et al., 2008; Teo et al., 2006). OPSC is a ductile material due to its low
modulus of elasticity. The low modulus of elasticity in OPSC allows the concrete to undergo high strain
under increasing load (Alengaram et al., 2011; Yap et al., 2014). For the deformation-based ductility ratio,
μD, the addition of fibres significantly reduced the ductility of OPSFRC beams, and the increase in fibre
content further reduced the ductility of OPSFRC beams. It can be seen that the OPSFRC-25 recorded
a 20% reduction in the ductility, while the ductility value further declined by 40% in the OPSFRC-100
mix. According to Ashour (2000), members with ductility ratio in the range of 3–5 has adequate duc-
tility and can be considered for structural members subjected to large displacements, such as sudden
forces caused by earthquake. The ductility ratios in OPSFRC specimens fell in the range of 4.5–9.3 and
the requirement for the structural ductility is fulfilled.
Previous discussions mentioned that the strong fibre-matrix interfacial bond holds the concrete
in place and this substantially increases the beam’s load capacity and crack resistance. However, this
phenomenon is accompanied by early fracture in the reinforcements. The ductility ratio μD considers
only the deflection of OPSFRC beams, but not taking into account the enhancement effect of steel
fibres on the load capacity. Therefore, ductility ratio, μD2 is more appropriate for OPSFRC specimens as
it takes both strength and deflection aspects into account. The μD2 values of the OPSFRC beams indicate
that the steel fibre reinforcement provides positive effect on the ductility of the OPSFRC with higher
increment in the moment capacity of the OPSFRC and smaller decrease in the deflection. Both the
OPSFRC-25 and OPSFRC-50 beams produced about 40% higher μD2 value than the OPSC mix. However,
the μD2 values of OPSFRC-75 and OPSFRC-100 specimens decreased notably by 20–30% compared to
the OPSFRC-50. It was reported that a minimum amount of steel fibre (.3–1.5%) is essential to produce
significant increment in the concrete’s mechanical properties and structural behaviour (Gao, Sun, &
Morino, 1997; Okay & Engin, 2012). For OPSFRC, the minimum volume of steel fibres is .50% (explained
in Section 3.1). Figure 7 compares the confinement effect of steel fibres with different volume. The
confinement effect is represented by the effective stress zone, which can be defined as the concrete
area held together by the fibre-matrix interfacial bond. When the volume fraction of steel fibre in the
cement matrix is lower (less than .50%), the confinement effect of steel fibre is weaker because the
effective stress zone of the steel fibres is not interconnected (as illustrated in Figure 7). For higher
volume fraction of steel fibre, although the strong confinement effect of the steel fibres improved the
load capacity of OPSFRC specimens, but at the same time strain localisation takes place and eventually
leads to early bar fracture (Meda et al., 2012).
Referring to the values of μE1 in Table 5, the variations of μE1 values in the OPSFRC specimens were
inconsistent as μE1 only takes into account the energy from the post-yielding stage up to ultimate stage.
Meanwhile, the observation of μD2 is also applicable to μE2. The OPSFRC with volume fraction .25–.50%
produced 13–21% increment in the energy-based ductility ratios while OPSFRC with .75–1.00% steel
fibres produced similar ductility ratios compared to the control OPSC beam. The observation can further
be supplemented by the results of the toughness values as shown in Table 4. The toughness values
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   11

Figure 7. Confinement effect of steel fibres when (a) low-volume fraction and (b) high-volume fraction.

were calculated from the area under the load–displacement curves (Dinh, Choi, & Kim, 2016). In this
study, the area under the curve was calculated using the trapezoidal area approximation method at
which the curve is broken up into multiple small trapeziums. The toughness is taken as the total area of
all trapeziums. The results show that the toughness of OPSFRC-25 and OPSFRC-50 beams improved by
3–6%; while the other two OPSFRC specimens show comparable toughness to that of OPSC beam. In
the case of OPSFRC, the effects of steel fibres enhance the toughness of both pre-peak and post-peak
behaviours. Hence, the ductility ratio, μE2 is more appropriate to indicate the energy-based ductility
of OPSFRC.

4. Conclusions
This study investigated the effects of steel fibres addition (.25% to 1.00% by volume) in the flexural
behaviour of OPSC. Based on the experimental results and discussions, the following conclusions are
drawn:

(1) The addition of steel fibres up to 1% by volume resulted in improved mechanical properties in


the OPSFRC, compared to the OPSC. The highest compressive and flexural strengths are 47.3
and 8.2 MPa, respectively.
(2) OPSFRC mixes show 4–18% higher moment capacities relative to the control OPSC mix which
could be attributed to the crack bridging effect of steel fibres.
(3) The OPSC specimen achieved maximum moment capacity at the steel strain value of about
4500 × 10−6. On the other hand, the strong fibre-matrix interfacial bond results in confinement
effect which enables OPSFRC to attain the respective maximum moment capacities at lower
steel strains of about 2500–3500 × 10−6.
(4) The confinement effect of steel fibres up to 1% by volume reduced the mid span deflection at
failure of OPSFRC beams by about 30–40%. Despite the reduction in deflection, the service load
deflection of OPSFRC beams satisfied the deflection limit as stipulated in BS 8110, Eurocode 2
and ASTM C1609.
(5) The benefits of steel fibres in ductility behaviours of OPSFRC are reflected in the post-yielding
stage, at which the OPSFRC beams exhibited flattened post-peak curve compared to the steep
post-peak curve of OPSC. In addition, the failure load to ultimate load ratios of OPSFRC are .65–.80
while the ratio is only .19 in OPSC beams.
12    S. P. YAP ET AL.

(6) Different ductility ratios are compared in this study. For the case of OPSFRC, the ductility ratio
calculation has to consider both the reduced deflection as well as the enhanced load capacity
of OPSFRC beams.

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
This work was supported by Bantuan Kecil Penyelidikan (BKP) titled ‘Development of waterproofing cement mortar using
plastic filling’ [Project Number: BK004-2016].

References
Abu-Lebdeh, T., Hamoush, S., Heard, W., & Zornig, B. (2011). Effect of matrix strength on pullout behavior of steel fiber
reinforced very-high strength concrete composites. Construction and Building Materials, 25, 39–46. doi:10.1016/j.
conbuildmat.2010.06.059
Alengaram, U. J., Jumaat, M. Z., & Mahmud, H. (2008). Ductility behaviour of reinforced palm kernel shell concrete beams.
European Journal of Scientific Research, 23, 406–420.
Alengaram, U. J., Mahmud, H., & Jumaat, M. Z. (2011). Enhancement and prediction of modulus of elasticity of palm kernel
shell concrete. Materials & Design, 32, 2143–2148. doi:10.1016/j.matdes.2010.11.035
Altun, F., & Aktaş, B. (2013). Investigation of reinforced concrete beams behavior of steel fiber added lightweight concrete.
Construction and Building Materials, 38, 575–581. doi:10.1016/j.conbuildmat.2012.09.022
Altun, F., Haktanir, T., & Ari, K. (2007). Effects of steel fiber addition on mechanical properties of concrete and RC beams.
Construction and Building Materials, 21, 654–661. doi:10.1016/j.conbuildmat.2005.12.006
Ashour, S. A. (2000). Effect of compressive strength and tensile reinforcement ratio on flexural behavior of high-strength
concrete beams. Engineering Structures, 22, 413–423. doi:10.1016/S0141-0296(98)00135-7
ASTM C1018. (1997). Standard test method for flexural toughness and first-crack strength of fiber-reinforced concrete (using
beam with third-point loading). West Conshohocken, PA: American Society for Testing and Materials (ASTM).
ASTM C469. (2010). Standard test method for static modulus of elasticity and Poisson’s ratio of concrete in compression. West
Conshohocken, PA: American Society for Testing and Materials (ASTM).
ASTM C496/ C496M. (2011). Standard test method for splitting tensile strength of cylindrical concrete specimens. West
Conshohocken, PA: American Society for Testing and Materials (ASTM).
ASTM C78/ C78M. (2010). Standard test method for flexural strength of concrete (using simple beam with third-point loading).
West Conshohocken, PA: American Society for Testing and Materials (ASTM).
Bencardino, F., Rizzuti, L., Spadea, G., & Swamy, R. (2008). Stress–strain behavior of steel fiber-reinforced concrete in
compression. Journal of Materials in Civil Engineering, 20, 255–263. doi:10.1061/(ASCE)0899-1561(2008)20:3(255)
BS EN 12390: Part 3. (2009). Testing hardened concrete- compressive strength of test specimens. London: BSI.
Dinh, N.-H., Choi, K.-K., & Kim, H.-S. (2016). Mechanical properties and modeling of amorphous metallic fiber-reinforced
concrete in compression. International Journal of Concrete Structures and Materials, 10, 221–236. doi:10.1007/s40069-
016-0144-9
Domagała, L. (2011). Modification of properties of structural lightweight concrete with steel fibres. Journal of Civil Engineering
and Management, 17, 36–44. doi:10.3846/13923730.2011.553923
du béton, F. (2000). Bond of reinforcement in concrete: State-of-the-art report. Lausanne: International Federation for Structural
Concrete.
Gao, J., Sun, W., & Morino, K. (1997). Mechanical properties of steel fiber-reinforced, high-strength, lightweight concrete.
Cement & Concrete Composites, 19, 307–313.
Gribniak, V., Kaklauskas, G., Hung Kwan, A. K., Bacinskas, D., & Ulbinas, D. (2012). Deriving stress–strain relationships for
steel fibre concrete in tension from tests of beams with ordinary reinforcement. Engineering Structures, 42, 387–395.
doi:10.1016/j.engstruct.2012.04.032
Hamoush, S., Abu-Lebdeh, T., & Cummins, T. (2010). Deflection behavior of concrete beams reinforced with PVA micro-fibers.
Construction and Building Materials, 24, 2285–2293. doi:10.1016/j.conbuildmat.2010.04.027
Hassanpour, M., Shafigh, P., & Mahmud, H. B. (2012). Lightweight aggregate concrete fiber reinforcement – A review.
Construction and Building Materials, 37, 452–461. doi:10.1016/j.conbuildmat.2012.07.071
Hsu, T. T. C. (1968). Torsion of structural concrete-behavior of reinforced concrete rectangular members. ACI Special
Publication, 18, 261–306.
Jaeger, G. L., Tadros, G., & Mufti, A. A. (1997). The concept of the overall performance factor in rectangular-section reinforced
concrete beams. In Proceeding of 3rd International Symposium on Non-metallic (FRP) reinforcement for concrete structures
(Vol. 2, pp. 551–558). Sapporo.
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   13

Khelifa, M. R., Leklou, N., Bellal, T., Hebert, R. L., & Ledesert, B. (2016). Is alfa a vegetal fiber suitable for making green reinforced
structure concrete? European Journal of Environmental and Civil Engineering, 1–21. doi: 10.1080/19648189.2016.1217792
Kim, H. S., & Shin, Y. S. (2011). Flexural behavior of reinforced concrete (RC) beams retrofitted with hybrid fiber reinforced
polymers (FRPs) under sustaining loads. Composite Structures, 93, 802–811. doi:10.1016/j.compstruct.2010.07.013
Mahmud, G. H., Yang, Z., & Hassan, A. M. T. (2013). Experimental and numerical studies of size effects of ultra high
performance steel fibre reinforced concrete (UHPFRC) beams. Construction and Building Materials, 48, 1027–1034.
doi:10.1016/j.conbuildmat.2013.07.061
Makita, T., & Brühwiler, E. (2013). Tensile fatigue behaviour of ultra-high performance fibre reinforced concrete (UHPFRC).
Materials and Structures, 47, 475–491. doi:10.1617/s11527-013-0073-x
Marthong, C., & Sarma, D. K. (2016). Influence of PET fiber geometry on the mechanical properties of concrete: an
experimental investigation. European Journal of Environmental and Civil Engineering, 20, 771–784. doi:10.1080/1964
8189.2015.1072112
Meda, A., Minelli, F., & Plizzari, G. A. (2012). Flexural behaviour of RC beams in fibre reinforced concrete. Composites Part B:
Engineering, 43, 2930–2937. doi:10.1016/j.compositesb.2012.06.003
Mo, K. H., Yap, S. P., Alengaram, U. J., Jumaat, M. Z., & Bu, C. H. (2014). Impact resistance of hybrid fibre-reinforced oil palm
shell concrete. Construction and Building Materials, 50, 499–507. doi:10.1016/j.conbuildmat.2013.10.016
Ng, T. S., Foster, S. J., Htet, M. L., & Htut, T. N. S. (2013). Mixed mode fracture behaviour of steel fibre reinforced concrete.
Materials and Structures, 47, 67–76. doi:10.1617/s11527-013-0045-1
Okay, F., & Engin, S. (2012). Torsional behavior of steel fiber reinforced concrete beams. Construction and Building Materials,
28, 269–275. doi:10.1016/j.conbuildmat.2011.08.062
Peled, A., Zaguri, E., & Marom, G. (2008). Bonding characteristics of multifilament polymer yarns and cement matrices.
Composites Part A: Applied Science and Manufacturing, 39, 930–939. doi:10.1016/j.compositesa.2008.03.012
Qian, C., & Indubhushan, P. (1999). Properties of high-strength steel fiber-reinforced concrete beams in bending. Cement
and Concrete Composites, 21, 73–81. doi:10.1016/S0958-9465(98)00040-7
Şahin, Y., & Köksal, F. (2011). The influences of matrix and steel fibre tensile strengths on the fracture energy of high-strength
concrete. Construction and Building Materials, 25, 1801–1806. doi:10.1016/j.conbuildmat.2010.11.084
Shafigh, P., Mahmud, H., & Jumaat, M. Z. (2011). Effect of steel fiber on the mechanical properties of oil palm shell lightweight
concrete. Materials & Design, 32, 3926–3932. doi:10.1016/j.matdes.2011.02.055
Singh, S., Shukla, A., & Brown, R. (2004). Pullout behavior of polypropylene fibers from cementitious matrix. Cement and
Concrete Research, 34, 1919–1925. doi:10.1016/j.cemconres.2004.02.014
Sun, Z. Z., & Xu, Q. W. (2009). Microscopic, physical and mechanical analysis of polypropylene fiber reinforced concrete.
Materials Science and Engineering: A, 527, 198–204. doi:10.1016/j.msea.2009.07.056
Teo, D. C. L., Mannan, M. A., & Kurian, V. J. (2006). Flexural behaviour of reinforced lightweight concrete beams made with
oil palm shell (OPS). Journal of Advanced Concrete Technology, 4, 459–468.
Toraldo, E., Mariani, E., & Crispino, M. (2016). Laboratory investigation into the effects of fibres and bituminous emulsion
on cement-treated recycled materials for road pavements. European Journal of Environmental and Civil Engineering, 20,
725–736. doi:10.1080/19648189.2015.1061460
Wang, H., & Belarbi, A. (2011). Ductility characteristics of fiber-reinforced-concrete beams reinforced with FRP rebars.
Construction and Building Materials, 25, 2391–2401. doi:10.1016/j.conbuildmat.2010.11.040
Yang, J.-M., Min, K.-H., Shin, H.-O., & Yoon, Y.-S. (2012). Effect of steel and synthetic fibers on flexural behavior of high-
strength concrete beams reinforced with FRP bars. Composites Part B: Engineering, 43, 1077–1086. doi:10.1016/j.
compositesb.2012.01.044
Yap, S. P., Alengaram, U. J., & Jumaat, M. Z. (2013). Enhancement of mechanical properties in polypropylene– and nylon–fibre
reinforced oil palm shell concrete. Materials & Design, 49, 1034–1041. doi:10.1016/j.matdes.2013.02.070
Yap, S. P., Bu, C. H., Alengaram, U. J., Mo, K. H., & Jumaat, M. Z. (2014). Flexural toughness characteristics of steel–polypropylene
hybrid fibre-reinforced oil palm shell concrete. Materials & Design, 57, 652–659. doi:10.1016/j.matdes.2014.01.004
Yap, S. P., Alengaram, U. J., & Jumaat, M. Z. (2015b). The effect of aspect ratio and volume fraction on mechanical properties
of steel fibre-reinforced oil palm shell concrete. Journal of Civil Engineering and Management, 22, 168–177. doi:10.384
6/13923730.2014.897970
Yap, S. P., Alengaram, U. J., Jumaat, M. Z., & Khaw, K. R. (2015). Torsional behaviour of steel fibre-reinforced oil palm shell
concrete beams. Materials & Design, 87, 854–862. doi:10.1016/j.matdes.2015.08.078.
Yap, S. P., Khaw, K. R., Alengaram, U. J., & Jumaat, M. Z. (2015). Effect of fibre aspect ratio on the torsional behaviour of steel
fibre-reinforced normal weight concrete and lightweight concrete. Engineering Structures, 101, 24–33. doi:10.1016/j.
engstruct.2015.07.007
Yap, S. P., Alengaram, U. J., Jumaat, M. Z., & Khaw, K. R. (2016). Torsional and cracking characteristics of steel fiber-reinforced
oil palm shell lightweight concrete. Journal of Composite Materials, 50, 115–128. doi:10.1177/0021998315571431
Yew, M. K., Mahmud, H. B., Ang, B. C., & Yew, M. C. (2015). Influence of different types of polypropylene fibre on the
mechanical properties of high-strength oil palm shell lightweight concrete. Construction and Building Materials, 90,
36–43. doi:10.1016/j.conbuildmat.2015.04.024
You, Z., Chen, X., & Dong, S. (2011). Ductility and strength of hybrid fiber reinforced self-consolidating concrete beam with
low reinforcement ratios. Systems Engineering Procedia, 1, 28–34. doi:10.1016/j.sepro.2011.08.006

Das könnte Ihnen auch gefallen