You are on page 1of 12

Case Studies in Construction Materials 11 (2019) e00304

Contents lists available at ScienceDirect

Case Studies in Construction Materials


journal homepage: www.elsevier.com/locate/cscm

Case study

Properties of Concrete Pedestrian Blocks Containing Crumb


Rubber from Recycle Waste Tyres Reinforced with Steel Fibres
Chalermphol Chaikaewa , Piti Sukontasukkulb,*, Udomvit Chaisakulkieta ,
Vanchai Satac , Prinya Chindaprasirtc
a
Department of Civil Engineering, Rajamangala University of Technology Rattanakosin, Nakhon Pathom, Thailand
b
Construction and Building Materials Research Center, Department of Civil Engineering, Faculty of Engineering, King Mongkut’s University
of Technology North Bangkok, Bangkok, 10800, Thailand
c
Sustainable Infrastructure Research and Development Center, Department of Civil Engineering, Faculty of Engineering, Khon Kaen
University, Khon Kaen, 40002, Thailand

A R T I C L E I N F O A B S T R A C T

Article history: Crumb rubber is a direct product from recycling abandoned tyres. It is considered one of the
Received 30 August 2019 solutions to reduce immense problems from discarded tyres. Rubberized pedestrian
Received in revised form 26 October 2019 concrete blocks are known to have good energy absorption but poor load resistance. In this
Accepted 30 October 2019
study, short steel fibres were introduced in the manufacturing of rubberized concrete
pedestrian blocks to enhance its mechanical properties. Two hooked end steel fibres with
Keywords: the lengths of 35 and 65 mm were mixed at 0.5 and 1.0% by volume fractions. The blocks
Crumb rubber
were produced by pressing and were subjected to density, water absorption, slip resistance,
Rubberized concrete
Pedestrian blocks
flexural load resistance, and field abrasion resistance tests. Results indicated that the
Steel fibre addition of steel fibre caused the density and absorption of blocks to increase slightly. The
Slip resistance flexural strength, toughness, and abrasion resistance also increased markedly with
Mechanical Properties increasing fibre content. The addition of fibre, however, did not affect the slip resistance.
© 2019 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC
BY license (http://creativecommons.org/licenses/by/4.0/).

1. Introduction

Concrete pedestrian blocks have been produced and used worldwide. They are usually manufactured by a pressing
technique in which the concrete mixtures are prepared drier than normal concrete, compressed by a pressing machine,
removed immediately, and cured in open air. The process is fast, simple, and also cost effective.
In practices, the overall qualities of concrete blocks are determined through slip resistance, abrasion resistance, and
strengths (compression and flexure). Plain pedestrian concrete blocks generally possess similar properties to plain concrete.
They exhibit excellent compressive strength but inferior flexural and tensile strength. The energy absorption is also low due
to the brittle nature of concrete. In order to enhance its energy absorption, highly elastic materials such as polymer or rubber
can be incorporated into concrete.
Rubber is one of the biggest industries in Thailand. Thailand is the world’s largest producer and exporter of natural rubber,
accounting for about one-third of the world’s supply. In 2015, Thailand’s rubber production accounted for about
4,314,975 tons [1], which was the highest among the top ten rubber producing countries [2]. In terms of consuming, about

* Corresponding author
E-mail addresses: chalermphol.c@rmutr.ac.th (C. Chaikaew), piti.s@eng.kmutnb.ac.th, piti@kmutnb.ac.th (P. Sukontasukkul), udomvit.k@rmutr.ac.th
(U. Chaisakulkiet), vancsa@kku.ac.th (V. Sata).

https://doi.org/10.1016/j.cscm.2019.e00304
2214-5095/© 2019 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/
4.0/).
2 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

85-90% are exported and about 10-15% are locally used. About half of the local usage are utilized in vehicle tyre industry. This
is no surprise as Thailand ranked no.69 in the world on the number of vehicle per capita (206 vehicles per 1000 people) [3]. In
2016, the Department of Land Transport reported the accumulative of about 37 million registered land vehicles [4].
Assuming that each vehicle changes their tyres every two years, this roughly yields number of wasted or abandoned tyres of
about 50 million tyres per year. With the increasing number of abandoned tyres, waste management has recently become a
challenging problem in Thailand.
According to UNEP, most wasted tyres at the end of service life have three possible solutions [5]: energy recover, material
recycling or recovery, and landfill waste. In the case of material recycling, tyres can be recovered into ground or crumb rubber
or recycled into reclaimed rubber. Crumb rubber can be used in applications like rubber-modified asphaltic road, sport fields/
tracks overlay, playground rubber tiles, etc.
Among several applications, crumb rubber found its way to concrete applications. Sukontasukkul et al. [6] used crumb
rubber particles from recycled waste tyres in concrete pedestrian blocks in order to enhance energy absorption and reduce
impact to the users. The crumb rubber particles successfully enhanced energy absorption. However, due to their flexibility
and low strength, both compressive strength and flexural strength were found to drop significantly. Similar findings related
to mechanical properties of rubberized concrete were also reported by several researchers [7–13]
In order to enhance the mechanical properties and further improve energy absorption of crumb rubber concrete
pedestrian blocks, several researchers proposed on pre-treating the surface of crumb rubber particles [14]. In this study, the
fibre reinforcement technology was adopted to enhance the mechanical properties of rubberized concrete. Fibre
reinforcement technology refers to a uniform mixing of short fibres into cementitious composite. Upon being subjected to
loads, cracks are initiated but the existence of fibres help intercept and slow down the crack propagation. This process of fibre
bridging across the crack leads to an increase in load carrying capacities and toughness and also permits materials to retain
load resistant ability after the first cracking [17–22].
In this study, the pedestrian blocks were prepared in forms of plain and rubberized concrete made from recycled crumb
rubber particles from waste tyres. The blocks were produced by the pressing method and cured in water for 28 days. The
effect of fibre was investigated by adding hooked end steel fibres with 2 different lengths (35 and 65 mm) at two different
volume fractions (0.5 and 1.0%) into rubberized pedestrian blocks. Four experiments were carried out include specific gravity
and absorption, slip resistance, flexural strength, and field abrasion tests.

2. Experimental Procedure

2.1. Materials

Materials consisted of Portland Cement Type I (ASTM C150) with specific gravity of 3.15, river sand with specific gravity of
2.63, and chipped rock with maximum size of 10 mm and specific gravity of 2.68. The crumb rubber was a commercial grade
type manufactured from recycled waste tyres with properties given in Table 1. The steel fibres were hooked end steel fibres
with two different lengths (35 and 65 mm) and properties as shown in Table 2.

2.2. Mix proportion and specimen preparation

The mix proportion of plain concrete blocks was set at 1 : 1.5 : 1.5 : 0.35 (Cement : Fine aggregate : Coarse aggregate :
Water) by volume for plain pedestrian concrete blocks (PC). For rubberized concrete blocks (RC), the crumb rubber was used
to replace fine aggregates at the rate of 10 and 20% by volume. For steel fibre reinforced rubberized concrete blocks (SFRRC),
two volume fractions of steel fibres were used at 0.5 and 1.0%. Detail mix proportions are given in Table 3.
To prepare test samples, all raw materials (except fibres) were dry mixed for 1-2 minutes in a pan mixer. After achieving
uniformity, water was added and the mixing continued for another 1-2 minutes. In case of SFRRC, fibre was added into the fresh
mix and the mixing continued further for 3 more minutes to obtain uniform distribution of fibres. The fresh mix was placed in a
mould and pressed into a rectangular block with dimension of 70  100 x 200 mm (Fig. 1) using a leverage pressing apparatus.
After pressing, the blocks were covered with plastic sheets for 24 hours and then moved to cure in water for 28 days.

Table 1
Properties of Crumb Rubber

Properties
Average Bulk Specific Gravity (Dry) 0.97
Average Bulk Specific Gravity (SSD) 0.98
Average Apparent Specific Gravity 0.98
Average Absorption (%) 1.01
Fineness Modulus 4.93
Average size (mm) (passing sieve no.6) 3.36
C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304 3

Table 2
Properties of Steel Fibre

No Materials Specific gravity Shape (mm.) Length (mm.) Cross-Section Aspect ratio Tensile Strength
diameter (mm) (l/d) (MPa)
1 Steel 7.8 Hooked end 35 0.55 64 1,050
2 Steel 7.8 Hooked end 65 0.90 67 1,050

Table 3
Mix Proportion

Type Cement Water Aggregate Crumb rubber Steel fibre


(kg) (kg)
Coarse (kg) Fine (kg) Vf (%) Weight Length Vf (%) Weight
(kg) (mm) (kg)
PC 498 174 839 889 - - - - -
RC10 498 174 756 799 10 63 - - -
0.5F35RC10 498 174 756 799 10 63 35 0.5 39
1.0F35RC10 498 174 756 799 10 63 35 1.0 78
0.5F65RC10 498 174 756 799 10 63 65 0.5 39
1.0F65RC10 498 174 756 799 10 63 65 1.0 78
RC20 498 174 672 711 20 126 - - -
0.5F35RC20 498 174 672 711 20 126 35 0.5 39
1.0F35RC20 498 174 672 711 20 126 35 1.0 78
0.5F65RC20 498 174 672 711 20 126 65 0.5 39
1.0F65RC20 498 174 672 711 20 126 65 1.0 78

3. Experiments

The experiments consisted of specific gravity and absorption test (ASTM C642) [23], slip test (BS7976) [24], flexural test
(ASTM C78) [25], and field abrasion test. For each test, at least three specimens were tested.
The slip resistance test was carried out based on BS7976 pendulum slip testing method specification (Fig. 2). The
apparatus consisted of a pendulum arm with an attached mechanical heel at the end. Prior to testing, the specimen was
secured rigidly to the floor, test surface was wetted with water, and the pendulum arm was moved up 90 degrees to align
parallel to the floor then locked. To begin the test, the pendulum arm was released and swung down freely. At the bottom or
point of contact, the rubber heel swept across the test surface and the maximum drag, in which the pendulum arm swings
up, called a ‘Pendulum Test Value (PTV)’ is measured. High PTV indicated low slip potential while low PTV indicated high slip
potential.
For the flexural test, the test was carried out using a 5-kN UTM. The test began by placing a sample on a support with
150 mm clear span length, then placing an LVDT underneath at 1/4 of the span (to avoid the location of crack), and slowly
applying a single point load at the centre location until failure. The rate of loading was controlled at 3.25 kN/min (equivalent
to 1 MPa/min of stress rate at the extreme surface). The mid-span deflection was extrapolated from the 1/4 deflection by
assuming zero deflection at the supports and linear deflection distribution from one end of the support to the centre point.
Field abrasion test was carried out at King Mongkut’s University of Technology-North Bangkok. The location of the test
was at the narrow walkway to the main cafeteria. Each type of block was numbered and laid in an area as shown in Fig. 3. The

Fig. 1. Block Dimension.


4 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

Fig. 2. Pendulum Slip Apparatus.

Fig. 3. Field Test Setup.

number of pedestrians was recorded every 30 minutes from 8:30 to 18:00 for three days per week (Table 4), the average of
number of pedestrians passing the test area daily was approximately 6,058. The test was carried on for 3 months, then the
blocks were removed, cleaned with brushes, and measured for weight loss.

4. Results and Discussions

4.1. Specific gravity and Absorption

Comparing between plain concrete (PC) and rubberized concrete blocks (RC), the specific gravity (SG) was found to
decrease from 2.49 to 2.41 and 2.38 with the increasing crumb rubber percentage from 10 and 20% (Table 5). This is because
SG of crumb rubber is lower than that of sand, the consequence of replacing sand with crumb rubber led to reduction in SG.
The absorption, on the other hand, was found to increase from 7.06 to 9.03 and 10.57% with the increasing crumb rubber
content from 0 to 10 and 20%, respectively (Table 4). The increase in absorption of rubberized concrete is believed to be
caused by the increase in porosity due to the non-polar effect and ability to trap of air bubbles around rubber particles [13].
In the case of SFRRC, regardless of fibre length, the addition of steel fibre increased the SG of rubberized concrete blocks
slightly. The increase in SG ranged between 1.3-3.7% for both FRRC10 and FRRC20. The increase in specific gravity is because
SG of fibre is larger than that of concrete. Although the amount of fibre is small, it still provides effect to the SG of FRRC.
Despite the increase in SG, the absorption percentage was found to increase with the increase in fibre content. For plain
rubberized concrete, the absorption of about 9.0 and 10.6% were observed in RC10 and RC20, respectively. With the addition
of steel fibre at 0.5 and 1.0% by volume, the absorption increased to about 9.7 to 11.6%. The increase in absorption is the direct
C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304 5

Table 4
Specific Gravity and Absorption

Type Specific gravity %Absorption


PC 2.49 7.06
RC10 2.41 9.03
0.5F35RC10 2.47 9.94
1.0F35RC10 2.49 10.40
0.5F65RC10 2.46 9.68
1.0F65RC10 2.50 10.02
RC20 2.38 10.57
0.5F35RC20 2.39 10.97
1.0F35RC20 2.40 11.61
0.5F65RC20 2.39 10.79
1.0F65RC20 2.41 11.35

Table 5
Slip Resistance and Friction Coefficient

Type Slip Resistance Dynamic Friction coefficient


(PTV)
PC 75 0.75
RC10 61 0.61
0.5F35RC10 60 0.60
1.0F35RC10 62 0.62
0.5F65RC10 61 0.61
1.0F65RC10 59 0.59
RC20 57 0.57
0.5F35RC20 57 0.57
1.0F35RC20 56 0.56
0.5F65RC20 55 0.55
1.0F65RC20 57 0.57

result of increasing interfacial transition zone (ITZ) between fibres and matrix. Since the ITZ is known be rich in calcium
hydroxide (CH) deposits and high porous [26], the addition of fibres caused the porosity to increase.

4.2. Slip resistance

The slip resistance (wet stage) and dynamic friction coefficient of all concrete blocks measured by Pendulum slip test
based on BS7976 test specification for Pendulum slip test are given in Fig. 4 and Table 5.
According to the specification guideline, the PTV value is categorized into three ranges: PTV < 24 = High risk,
25 < PTV < 35 = Moderate risk, 36 < PTV < 74 = Low Risk, and PTV > 75 = extremely low risk. Based on the obtained results,
all the pedestrian blocks exhibited PTV in between 55 and 75 thus categorizing them in the group of low to extremely low
risk of slipping.

Fig. 4. Slip Resistance.


6 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

Comparing between PC and RC, the PTV decreased with the addition of crumb rubber. The PTV of PC was averaged around 75
and can be classified as extremely low risk of slipping. With 10 and 20% of crumb rubber replacement, the PTV decreased to about
61 and 57, respectively. The decrease of PTV implied that the risk of slipping increases with the increasing crumb rubber content.
This is perhaps due to the smooth surface of rubber crumb particles at the surface of the specimen. Under wet conditions, the
surface of the crumb rubber may become more slippery allowing the pendulum to swing pass easier and higher.
In the case of SFRRC blocks, the addition of fibre apparently did not affect the PTV values. They were influenced mainly by
the rubber content rather than the fibre content. Regardless of volume fraction and length, the PTV were found in a very
similar range of 61 to 62 and 55 to 57 for SFRRC10 and SFRRC20, respectively. In general, metal surfaces should be more
slippery than the concrete surface especially under wet conditions. However, the results showed no effect of fibres on PTV.
This is perhaps because of the small volume fractions of fibre used in this study (0.5 and 1.0%), which allowed very small
number of fibres to be exposed at the contact surface.

4.3. Flexural performance

4.3.1. PC vs RC blocks
The flexural responses of PC and RC blocks are shown in Fig. 5. The PC block was found to behave in brittle manner. The
load increased in proportion with the deflection up to the peak, followed by an occurrence of cracks at the bottom face. The
crack propagated at fast rate which led to a sharp drop of load and a sudden failure. For the RC blocks, the flexural responses
were slightly more ductile than the PC block. A sign of ductile mode was observed in form of gradually declining post peak
responses. Although a sharp drop of load was observed right after the peak load (similar to PC blocks), the load did not drop
down to zero (complete fractured) right away. The existence of crumb rubber helped bridge the crack and support the
specimen. This allowed the RC blocks to maintain a small load carrying capacity and extend the final deflection (deflection at
complete fracture) from 0.3 mm to about 1.7-2.0 mm.
Results in terms of flexural load and strength are given in Table 6. The peak flexural load was found to decrease with the
addition of crumb rubber. The highest peak load was observed in PC block at about 8.9 kN. With the addition of crumb rubber,
the peak load dropped to about 3.96 and 2.51 kN in RC10 and RC20, respectively. Since crumb rubber particles are weaker in
strength and less stiff than sand (fine aggregates), the replacement of sand with crumb rubber particles led to the reduction
in load carrying capacity.

Fig. 5. Flexural Response of PC and RC Blocks.


C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304 7

4.3.2. RC vs FRRC blocks


The flexural responses of RC compared with SFRRC blocks are shown in Figs. 6–8. For SFRRC blocks, the load responses
were more ductile than that of RC blocks. At the beginning of loading, a linear relationship between load and deflection was
observed and then, followed by a strain hardening at near peak load region. Right after the peak load (first cracking), a sharp
drop of load occurred. However, with the effect of fibres bridging and intercepting cracks, a sudden failure was not stopped
and the load was picked up again. The load carrying response after the peak load is called the ‘post peak response’ and it is
majorly controlled by fibre type, content, geometry etc.
Figs. 6 and 7 shows the comparison between the failure mode of RC vs SFRRC blocks. The failure mode of an RC block
showed a single crack running from bottom to top which was sufficient to cause a complete fracture of the block. In case of
SFRRC block, multiple cracks were observed. The effect of fibre bridging across the crack created resistance and caused the
crack to be divert and arrested which led to the formation of new cracks.
The flexural responses of SFRRC blocks not only showed higher strength and stiffness, they also exhibited better post-peak
responses than RC blocks did. The stiffness and strength were found to increase with increasing fibre content. The effect of fibre
on post-peak responses can better be illustrated by the flexural toughness and residual strength. The flexural toughness (T) is
defined as an area under a load-deflection curve up to a certain deflection point and the residual strength (fR) is the remaining
flexural strength after first cracking up to a certain deflection (2 mm). Both can be calculated using Eqs. (1) and (2).
Z d
T¼ Pdd ð1Þ
0

3PL
fR ¼ 2
ð2Þ
2bd
where T is flexural toughness (N-m), P is bending load (kN), d is deflection (mm), fR is residual strength, L is clear span length
(mm), b is specimen width (mm) and d is specimen height (mm).
Comparing between SFRRC10 and SFRRC20, the addition of steel fibre increased the flexural strength and flexural
toughness of rubberized concrete blocks at different degrees.
For SFRRC10, the flexural strength increased with fibre content. At 0.5% volume fractions, the flexural strength increased
in the range of 33-49%. As the fibre volume fraction increased to 1.0%, the flexural strength increased by 111-147%. The
increase in flexural strength was due to the effect of fibre bridging and intercepting cracks, which helped arrest cracks and
slowed down rate of crack propagation. The increase in fibre volume fraction also doubled the number of fibres in concrete.
This doubled the chances of fibres intercepting the moving cracks. In order to fracture the specimen or overcome the bond
between fibres and matrix, more loads were required to exert into the specimen.
In the case of SFRRC20 (20% crumb rubber content), large amount of crumb rubber weakened the cement matrix. The
flexural strength of plain RC20 was observed at 0.56 MPa. The addition of steel fiber at 0.5 and 1.0% enhanced the flexural
strength by about 24-40% and 90-125%, respectively. The effect of fibre in enhancing flexural strength was more pronounced
in RC10 than in RC20. This was because the matrix of RC20 was weaker than that of RC10 by about 60%. The weaker matrix
resulted in poorer bond strength between fibres and matrix in RC20 as compared to RC10 blocks.
As for the effect of fibre length, the results showed that the blocks with long fibre performed better in both strength and
flexural toughness than those with short fibre at the same fibre volume fraction and crumb rubber content. The blocks with
65 mm fibres outperformed those with 35 mm fibres by 53-66% in terms of peak load and strength.

4.3.3. Residual strength and Toughness


Results on residual strength and toughness are shown in Fig. 9. For plain blocks without fibres (PC, RC10 and RC20), the
maximum toughness of 1.8 N-m was observed in PC block. Toughness decreased to 1.5 and 1.4 N-m in RC10 and RC20,

Table 6
Peak Load and Flexural Strength

Type Peak Load (kN) Strength (MPa)

Average S.D. Average S.D.

PC 8.90 0.26 2.00 0.06


RC10 3.96 0.39 0.89 0.09
0.5F35RC10 5.26 0.75 1.18 0.17
1.0F35RC10 8.37 0.51 1.88 0.11
0.5F65RC10 5.91 0.53 1.33 0.12
1.0F65RC10 9.79 0.51 2.20 0.11
RC20 2.51 0.35 0.56 0.08
0.5F35RC20 3.12 0.50 0.70 0.11
1.0F35RC20 4.77 0.66 1.07 0.15
0.5F65RC20 3.51 0.56 0.79 0.13
1.0F65RC20 5.63 0.61 1.27 0.14
8 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

Fig. 6. Failure mode of (a) RC Block and (b) SFRRC Block under Bending.

Fig. 7. Flexural Response of FRRC10 with Different Fibre Lengths: (A) 35 mm and (B) 65 mm.

respectively. Although RC blocks exhibited slightly more ductile responses than PC block did, their toughness was smaller
than the PC block by 22%. This is the direct result of the much lower peak load (or strength) by 55-72% in RC blocks as
compared to PC block.
In the case of SFRRC blocks, regardless of the crumb rubber content, both residual strength and toughness were found to
increase with the increasing fibre volume fraction. The increasing fibre volume fraction from 0.5 to 1.0% caused the
C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304 9

Fig. 8. Flexural Response of FRRC20 with Different Fibre Lengths: (A) 35 mm and (B) 65 mm.

Fig. 9. Residual Strength and Toughness.

toughness to increase by about 81-116% in SFRRC10 and 55-108% in SFRRC20, respectively. The residual strength, which
represents the ability to carry load beyond first crack of fibre reinforced rubberized concrete, was also found to increase with
increasing fibre content.
Comparing between SFRRC10 and SFRRC20, the SFRRC10 blocks exhibited higher residual strength and toughness than
the SFRRC20 blocks did in all fibre types and volume fractions (Fig. 8). This contributed mainly by the stronger matrix of
SFRRC10 as compared to SFRRC20.
10 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

Besides from the effect of fibre bridging across the cracks, the improvement in flexural strength and flexural toughness
was partially due to the pressing technique that was used in manufacturing the pedestrian blocks. The pressing method
caused fibres to align more in a horizontal direction (parallel to the tensile stress direction), which in turn allowed fibres to
perform in their preferable mode and provided better crack resistance.

4.4. Field Abrasion Test

Results on field abrasion resistance test are shown Fig. 10. The PC blocks exhibited about 0.55% weight loss while the RC10
and RC20 blocks exhibited higher weight loss at 1.31% and 2.57%, respectively. Higher weight loss percentage implies that RC
blocks were poorer in abrasion or wearing resistance than PC blocks. This is due to the smooth surface and the hydrophobic
nature of crumb rubber particles which caused poor bonding to the paste. The rubber particle is also low in strength. The
combination of poor bond and strength caused the rubber particles to pop out when subjected to abrasive forces and leave
behind air voids that led to surface wearing. Similar findings were reported by Gupta et al. [27] and Ozbay et al. [28].
For SFRRC blocks, the abrasion resistance was found to improve slightly as seen by the decrease in weight loss percentage
with the increasing fibre content. With fibre volume fractions of 0.5% and 1.0%, the weight loss decreased by 12-42% for
SFRRC10 and 10-40% for SFRRC20 comparing to RC blocks. This is because steel fibres are capable of providing bond of
concrete mortar at the surface and is known to improve abrasion resistance as reported by Atis et al.[29] and Felekog lu et al.
[30]. The increasing fibre length also provided positive effect on the abrasion resistance. With fibre length increase from 35 to
65 mm, the abrasion resistance also increased by 7-30% for both types of crumb rubber blocks.

5. Conclusion

The addition of steel fibre effected the properties of rubberized concrete pedestrian blocks differently.

 In terms of physical properties, the addition of fibre increased the absorption of the blocks due to the increasing ITZ
between fibre and matrix. The specific gravity also increased slightly with the increasing fibre content due to heavier
density of steel fibre.
 The slip resistance of all PC, RC, and SFRRC blocks can be categorized into low to extreme low risk of slipping group. The
replacement of sand with crumb rubber particles lowered the PTV values indicating the increase in slipping risk. The
volume fraction of steel fibre used in this study (1%) did not affect the slipping risk of the RC blocks because a small
number of fibres were exposed on the surface.
 Similar to previous studies, the addition of crumb rubber into concrete caused the mechanical properties to decrease
gradually depending on the rubber content.
 The addition fibre enhanced the flexural load and abrasion resistance of RC blocks significantly as seen by the increase in
flexural strength, residual strength and flexural toughness, and the decrease in weight loss percentage. The flexural
strength, residual strength and toughness was found to increase with fibre content. The mechanical enhancement was
more pronounced in RC10 than in RC20 due to the stronger matrix which created a stronger bond between fibres and

Fig. 10. Weight Loss Percentage from Field Abrasion Test.


C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304 11

matrix. Long fibres appeared to perform better than short fibres in the case of pressed pedestrian blocks, this was partly
due to the manufacturing process which allowed fibres to align more in horizontal directions.
 The results of this research prove that the application of crumb rubber in concrete continues to be an essential part in the
process of reducing waste from abandoned tyres. Although the mechanical properties of concrete become inferior after
incorporating crumb rubber into the mixture, there are several ways to improve the mechanical properties up to the
applicable level. Use of fibre reinforcement seems to be one method to improve the mechanical properties of rubberized
concrete. Also, there are many applications in the construction industry that require low strength and highly flexible
concrete, thus rubberized concrete seems to be a suitable choice.

Declaration of Competing Interest

The authors declare no conflict of interest.

Acknowledgements

This project is funded by King Mongkut’s University of Technology North Bangkok under contract no. KMUTNB-63-
KNOW-024. The last author would like to acknowledge the support from Thailand Research Fund under Distinguished
Research Professor grant no. DPG6180002. The first and third authors would like to acknowledge the support from
Rajamangala University of Technology Rattanakosin (RMUTR). The authors would like to thank our undergraduate students,
Mr. Jakkarin Pongpinitana, Mr. Danai Ukachot, and Mr. Pongsthorn Au-nupongpichart for their helps carry out this project.
Special thank is also to SR.Fibre Co., Ltd and Bekaert (Thailand) Co., Ltd. for providing steel fibres.

References

[1] Research and Development Centre for Thai Rubber Industry, Thai Rubber Statistic (in Thai) Retrieved May 2017, http://www.rubbercenter.org/index.
php/stat.
[2] Top 5 Leading Rubber Producing Countries, Retrieved May 2017, http://www.liveind.com/top-5-leading-rubber-producing-countries/11824.
[3] Nation Master, Motor vehicles per 1000 people Retrieved May 2017, http://www.nationmaster.com/country-info/stats/Transport/Road/Motor-
vehicles-per-1000-people.
[4] The Department of Land Transport, Number of Vehicles registered as of 31 December 2016 Retrieved May 2017, http://apps.dlt.go.th/statistics_web/
brochure/cumcar16.pdf.
[5] United Nations: Environment Programme, Basel Convention Series: Technical Guidelines on the Identification and Management of Used Tyres/
Secretariat of the Basel Convention Retrieved March 2013, from, Secretariat of the Basel Convention, 2002. http://archive.basel.int/meetings/sbc/
workdoc/old%20docs/tech-usedtyres.pdf.
[6] P. Sukontasukkul, C. Chaikaew, Properties of concrete pedestrian block mixed with crumb rubber, Construction and Building Materials 20 (7) (2006)
450–457, doi:http://dx.doi.org/10.1016/j.conbuildmat.2005.01.040.
[7] I.B. Topçu, The properties of rubberized concretes, Cement and Concrete Research 25 (2) (1995) pp. 340-310.
[8] N.N. Eldin, A.B. Senouci, Measurement and prediction of the strength of rubberized concrete, Cement and Concrete Composites 16 (4) (1994)
287–298.
[9] Z.K. Khatib, F.M. Bayomy, Rubberized Portland Cement Concrete, Journal of Materials in Civil Engineering 11 (3) (1999).
[10] C.A. Issa, G. Salem, Utilization of recycled crumb rubber as fine aggregates in concrete mix design, Construction and Building Materials 42 (2013) 48–
52, doi:http://dx.doi.org/10.1016/j.conbuildmat.2012.12.054.
[11] A. Sofi, Effect of waste tyre rubber on mechanical and durability properties of concrete – A review, Ain Shams Engineering Journal 9 (4) (2018) 2691–
2700, doi:http://dx.doi.org/10.1016/j.asej.2017.08.007.
[12] K. Bisht, P.V. Ramana, Evaluation of mechanical and durability properties of crumb rubber concrete, Construction and Building Materials 155 (2017)
811–817, doi:http://dx.doi.org/10.1016/j.conbuildmat.2017.08.131.
[13] P. Sukontasukkul, K. Tiamlom, Expansion under water and drying shrinkage of rubberized concrete mixed with crumb rubber with different size,
Construction and Building Materials 29 (2012) 520–526, doi:http://dx.doi.org/10.1016/j.conbuildmat.2011.07.032.
[14] O. Onuaguluchi, D.K. Panesar, Hardened properties of concrete mixtures containing pre-coated crumb rubber and silica fume, Journal of Cleaner
Production 82 (2014) 125–131.
[17] N. Banthia, M. Sappakittipakorn, Toughness enhancement in steel fiber reinforced concrete through fiber hybridization, Cement and Concrete Research
37 (9) (2007) 1366–1372.
[18] G. Plizzari, S. Mindess, 11 - Fiber-reinforced concrete, in: Sidney Mindess (Ed.), Woodhead Publishing Series in Civil and Structural Engineering,
Developments in the Formulation and Reinforcement of Concrete, Second Edition, Woodhead Publishing, 2019, pp. 257–287 ISBN
9780081026168.
[19] Eva O.L. Lantsoght, How do steel fibers improve the shear capacity of reinforced concrete beams without stirrups? Composites Part B: Engineering 175
(2019), doi:http://dx.doi.org/10.1016/j.compositesb.2019.107079.
[20] M.G. Alberti, A. Enfedaque, J.C. Gálvez, C. Álvarez, Using Polyolefin Fibers with Moderate-Strength Concrete Matrix to Improve Ductility, Journal of
Material in Civil Engineering 31 (2019).
[21] P. Sukontasukkul, P. Pongsopha, P. Chindaprasirt, S. Songpiriyakij, Flexural performance and toughness of hybrid steel and polypropylene fibre
reinforced geopolymer, Construction and Building Materials 161 (2018) 37–44.
[22] S. Iqbal, I. Ali, S. Room, S.A. Khan, A. Ali, Enhanced mechanical properties of fiber reinforced concrete using closed steel fiber, Materials and Structure 52
(2019) 56–66, doi:http://dx.doi.org/10.1617/s11527-019-1357-6.
[23] ASTM C642-13, Standard Test Method for Density, Absorption, and Voids in Hardened Concrete, ASTM International, West Conshohocken, PA, 2013.
[24] BS 7976, Test specification for Pendulum slip test forms the basis of our on-site slip risk assessment, British Standard Institute, UK.
[25] ASTM C78 / C78M – 18, Standard Test Method for Flexural Strength of Concrete (Using Simple Beam with Third Point Loading), ASTM International,
West Conshohocken, PA, 2013.
[26] A. Bentur, S. Mindess, N. Banthia, The interfacial transition zone in fibre reinforced cement and concrete, Engineering and Transport Properties of
the Interfacial Transition Zone in Cementitious Composites - State-of-the-Art Report of RILEM TC 159-ETC and 163-TPZ, RILEM Publications SARL,
1999, pp. 89–109.
12 C. Chaikaew et al. / Case Studies in Construction Materials 11 (2019) e00304

[27] T. Gupta, S. Chaudhary, R.K. Sharma, Assessment of mechanical and durability properties of concrete containing waste rubber tire as fine aggregate,
Construction and Building Materials 73 (2014) 562–574, doi:http://dx.doi.org/10.1016/j.conbuildmat.2014.09.102.
[28] E. Ozbay, M. Lachemi, U.K. Sevim, Compressive strength, abrasion resistance and energy absorption capacity of rubberized concretes with and without
slag, Materials and Structures 44 (2011) 1297–1307, doi:http://dx.doi.org/10.1617/s11527-010-9701-x.
[29] C.D. Atis, O. Karahan, K. Ari, Ö.C. Sola, C. Bilim, Relation between Strength Properties (Flexural and Compressive) and Abrasion Resistance of Fiber (Steel
and Polypropylene)-Reinforced Fly Ash Concrete, ASCE Journal of Materials in Civil Engineering 21 (8) (2009).
[30] B. Felekog lu, S. Türkel, Y. Altuntaş, Effects of steel fiber reinforcement on surface wear resistance of self-compacting repair mortars, Cement and
Concrete Composites 29 (5) (2007) 391–396, doi:http://dx.doi.org/10.1016/j.cemconcomp.2006.12.010.

You might also like