You are on page 1of 28

Journal of Cleaner Production 338 (2022) 130676

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

A review of coir fibre and coir fibre reinforced cement-based composite


materials (2000–2021)
Bo Wang a, Libo Yan a, b, *, Bohumil Kasal a, b
a
Organic and Wood-based Construction Materials, Technische Universität Braunschweig, Hopfengarten 20, 38102, Braunschweig, Germany
b
Center for Light and Environmentally-Friendly Structures, Fraunhofer Wilhelm-Klauditz-Institut, Bienroder Weg 54E, 38108, Braunschweig, Germany

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

Handling Editor: Zhen Leng This paper provides a comprehensive review on the research of coir fibre and coir fibre reinforced cementitious
composite (CFRC) in the past 20 years. In the first part, the extraction process, morphology, density, chemical
Keywords: composition, and tensile performance of coir fibres are discussed. Then, the pull-out performance, physical
Coir fibre properties (i.e., density, thermal and acoustic insulation), short- and long-term properties (i.e., compressive,
Fibre reinforced cementitious matrix
flexural, impact, and dynamic performance) of CFRC are reviewed. Existing modification methods (i.e.,
Fibre pull-out
cementitious matrix and fibre surface modifications) to improve the bond and mechanical behaviour of CFRC
Mechanical properties
Coir fibre composites and the practical application of CFRC in construction are presented. Future perspectives of CFRC studies are
highlighted, including the validation of existing models (i.e., for the prediction of coir tensile strength as well as
bond strength and total energy of CFRC), further investigations on long-term, seismic, fire performance of CFRC,
and the use of coir fibre in geopolymer and coconut shell aggregate concrete towards practical application.

1. Introduction emission should be made primarily in the building and construction


sector. To date, using materials from agricultural and forest wastes and
The earth planet is suffering from climate change such as global by-products in construction has increasingly attracted the attention from
warming along with a steady increase of green-house gases (GHG) researchers, governments as well as industries. Using such wastes and
emission. The emission of global energy-related CO2 (i.e., the main by-products in a cement-based matrix brings about number of positive
GHG) was around 33.0 Gt in 2019, while that in 2000 was only 23.1 Gt aspects such as reduction of weight and costs, improved ductility and
(IEA, 2019). Considering the pressing climate crisis, ambitious objec­ thermal properties, and recycling of what would otherwise be consid­
tives have been set by all countries as well as international organiza­ ered a waste. For example, wood chips from the residual of logging and
tions. “Carbon neutrality” (i.e., net zero-carbon emission) is the core to sawmills were used in wood chip concrete as lightweight construction
these objectives. In the frame of Paris Agreement signed in 2015, the material (Fu et al., 2020a, 2020b; Waldmann et al., 2017; Li et al., 2017)
European Union had set a target to achieve a net domestic GHG with improved thermal and sound insulation properties. Straw (Wei
reduction of at least 55% of that in 1990 by 2030. In 2019, the European et al., 2015; Nasser et al., 2018; Bakatovich et al., 2018; Palumbo et al.,
Commission released the “European Green Deal” (European Commis­ 2015) such as rice, wheat, and oat as well as sugarcane bagasse (Mes­
sion, 2021) aiming at carbon neutrality by 2050. Most recently, China quita et al., 2018; Ramlee et al., 2019; Doost-hoseini et al., 2014) or rice
has also announced the same target of carbon neutrality by 2060 (Chen husks (Akinyemi et al., 2020) from agriculture waste were usually uti­
et al., 2021). Among all the industries, building and construction sector lized for producing insulation board. The ashes produced from com­
provided the largest contribution to the energy-related CO2 emission, bustion of biomass such as waste wood, rice husk, sugarcane bagasse,
which accounted for 39% of the global energy-related CO2 emission in and corn cob can be used as supplementary cementitious materials as
2018 (Global Alliance for Buildings and Construction and International replacement of ordinary Portland cement (Fořt et al., 2021; Jahanzaib
Energy Agency and the United Nations Enrionment Programme, 2019). Khalil et al., 2021; Raheem et al., 2010; Adesanya and Raheem, 2010;
To achieve the carbon neutrality, efforts for the mitigation of GHG Shakouri et al., 2020; Sutas et al., 2012; Chao-Lung et al., 2011; Givi

* Corresponding author. Organic and Wood-based Construction Materials, Technische Universität Braunschweig, Hopfengarten 20, 38102, Braunschweig,
Germany.
E-mail addresses: l.yan@tu-braunschweig.de, libo.yan@wki.fraunhofer.de (L. Yan).

https://doi.org/10.1016/j.jclepro.2022.130676
Received 28 June 2021; Received in revised form 12 January 2022; Accepted 23 January 2022
Available online 26 January 2022
0959-6526/© 2022 Elsevier Ltd. All rights reserved.
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

et al., 2010; Kang et al., 2019; Thomas, 2018; Vieira et al., 2020). 2. Coir fibre
Another effective application for utilization of agricultural and forest
waste in construction is to use plant-based fibres (Ahmad et al., 2019; 2.1. Morphology, physical properties and chemical composition of coir
Parveen et al., 2017; Onuaguluchi and Banthia, 2016; Pacheco-Torgal fibre
and Jalali, 2011; Ramakrishna and Sundararajan, 2005a, 2005b; Ram­
akrishna and Priyadharshini, 2018; Tolêdo Filho et al., 2003) such as Coconut (cocos nucifera) is one of the most essential crop species in
coir (Yadav and Tiwari, 2016; Pateriya et al., 2021; Tolêdo Filho et al., almost all tropical and subtropical regions (Chan and Craig, 2006; Ali,
2000; Ahmad et al., 2020; Ali et al., 2012; Andiç-Çakir et al., 2014; 2011; Kadereit et al., 2014). It is usually used in the copra production
Asasutjarit et al., 2007; Hwang et al., 2016; Khan and Ali, 2019; Khatri such as coconut oil extraction for food and cosmetic application (Prab­
et al., 2015; Lertwattanaruk and Suntijitto, 2015; Li et al., 2006, 2007a; hakaran Nair, 2010). Coir fibre is a by-product from the copra produc­
Maia Pederneiras et al., 2021; Quiñones-Bolaños et al., 2021; Ramli tion and extracted from mesocarp of the coconut (Lai et al., 2005). The
et al., 2013; Yan and Chouw, 2013a; Yan et al., 2015, 2016a), sisal extraction process of coir fibre is shown in Fig. 1. After the separation of
(Tolêdo Filho et al., 2000; Wei, 2014; Ferreira et al., 2014), Spanish the coconut husk (i.e., mesocarp and exocarp), the coconut husk is
Broom (Juradin et al., 2019), oil palm (Lertwattanaruk and Suntijitto, usually stored in water for a water retting process. The retting process
2015) and jute (Islam and Ahmed, 2018; Zakaria et al., 2017) as sub­ aims to separate the fibres by biologically degrading the pectin, which
stitutions to the man-made fibre reinforcement applied in cementitious holds the fibres together (Xia and Li, 2019). The retting process of coir
materials. Comparing with man-made fibres such as steel, poly­ fibre lasts around 4–12 months (Hasan et al., 2021). The retted husk is
propylene, carbon, and glass fibres, the plant-based fibres require then washed and dried. After drying the exocarp can be readily peeled
significantly less energy during production and are renewable. Table 1 by hand. Through pressing, hackling (e.g., combing or spinning), and
lists the tensile properties and embodied carbon footprint of conven­ screening, the coir fibre and coco peat are separated. Coir fibres have
tional fibre materials used as monofilament fibre reinforcement in relatively low modulus of elasticity compared with other plant-based
cementitious materials. Among all the fibres listed in Table 1, coir fibre, natural fibres (e.g. flax) and man-made fibres. Due to the high avail­
as a typical representative, has not only lowest global warming potential ability of coconut plants, coir fibres are widely processed and used to
(GWP, 0.2 kg equivalent CO2/kg), but also a significantly larger elon­ make different products for different applications, especially in tropical
gation at failure (15.0%–51.4%) than most of other fibres. The elonga­ and subtropical regions. For example, coir fibres are produced to be
tion at failure is a property that may positively ensure the bridging effect fertilizer, brushes or coir ropes. In construction, coir fibres can be used
of coir fibre in cement-based matrix and affect the ductility of otherwise as insulation materials and reinforcement of cementitious materials.
brittle cement-based matrix. It was also reported that amongst Other plant-based natural fibres, such as flax fibres with higher tensile
plant-based fibres, coir fibre had the highest toughness in tension (i.e. strength and modulus, are used to produce textiles for automotive (e.g.,
the area under the stress-strain curve of a material under tension), as door panel of car (RWTH AAchen University, 2022)) and construction
therefore it has been widely investigated as reinforcement of cementi­ (e.g., as decking system for footbridge (Blok et al., 2019)). The coco peat
tious composite for low-cost concrete structure (Ali et al., 2012). The is usually applied as coco peat soil for planting. In the last ten years, the
readers must be cautioned that the comparisons in Table 1 are only annual production of coir fibre was around 1.2 million tonnes (Food and
general due to the number of different sources used. Additionally, only Agriculture Organization of the United Nations, 2015). India, Vietnam,
about 15% of the produced coir fibres are recovered (Gu, 1980–2015; and Sri Lanka are the three largest coir production countries with annual
Bui, 2021). Due to its mechanical properties and ecological impact, coir yielding of 580, 382, and 161 thousand tonnes, respectively (Food and
is considered to be one of the ideal natural fibres as monofilament fibre Agriculture Organization of the United Nations, 2015). Based on the
reinforcement used in cementitious materials. maturity of coconut, there are two types of coir fibres, i.e., brown fibre
and white fibre (Fig. 2). The fibre extracted from matured coconut is the
brown fibre while that from immature coconut is the white fibre. White
coir fibre usually has a colour of white or light brown. Since the white
coir fibre is fine and smooth, it is usually spun into yarn for coir mat and
rope (Ashby, 2013). In comparison with white coir fibre, brown coir
fibre is usually thicker, stronger and has higher abrasion resistance due
to its longer growth time. Therefore, brown fibre is suitable to be used as
Table 1
monofilament reinforcement in CFRC.
Tensile properties and embodied carbon footprint of conventional fibre as
Coir fibre typically has a cylindrical structure (i.e., around 10–460
reinforcement in cementitious matrix (Wang et al., 2021; Zoghi, 2013; Richaud
et al., 2018; Jones and Hammond; Das, 2011; Grasselly, Hamm, Quaranta,
μm in diameter (Dittenber and GangaRao, 2012)) with a middle hollow
Vitrou; Dittenber and GangaRao, 2012; Thienel, 2020). area (i.e., the lacuna) surrounded by 200–300 elementary fibres (Tran
et al., 2015) as presented in Fig. 3 (a). An elementary fibre is a hollow
Fibre type Tensile Elastic Elongation at GWPa
strength Modulus failure
cylindrical structure, which is comprised of individual fibre cells with
the length of around 1 mm and the diameter of 10–20 μm (Yan, 2016).
(MPa) (GPa) (%) (kg equivalent
The scanning electron microscope (SEM) image analysis and
CO2/kg)
SEM-computer tomography (CT) scans from Tran et al. (2015) revealed
Steel 300–2500 200–210 3–4 2.0–3.1
that the lacuna area of one coir fibre was 4000 to 20,000 μm2. The total
Polypropylene 50–600 0.5–3.0 50–600 5.0
E-glass 1800–3500 70–75 2.0–3.5 1.4 (in
fibre porosity was measured to be around 21%–31%. Such high porosity
general) provides the floating ability of coconuts and ensures the inhabitability
Carbon (PAN 3500–5000 200–260 1.2–1.8 1.3–1.5 and widespread of coconuts in almost all tropical and subtropical re­
HT)b gions. The coir fibre surface along fibre direction is presented in Fig. 3
Coir 95–230 2.8–6 15.0–51.4 0.2
(b). Small round openings can be found on the surface which are known
Sisal 363–700 9–38 2.0–7.0 –
Flax 343–2000 27.6–103 1.2–3.3 – as pit. Since the lignification of the cell wall makes the cell less
Hemp 270–900 23.5–90 1.0–3.5 0.6–1.0 permeable, the water or general nutrition exchange are achieved by the
Kenaf 223–930 14.5–53 1.5–2.7 0.6–0.9 pits (Kadereit et al., 2014). Globular protrusions are usually found in the
a
GWP (global warming potential) was calculated for the entire life of material pit, which are known as tylosis (Madueke et al., 2021). Under stress or
(from cradle to grave). during invasion of pathogens, the protoplast of the adjacent parenchy­
b
Carbon fibre from polyacrylonitrile (PAN); HT for high tensile strength. matous cells is overgrown to generated a new wall material, i.e., the

2
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 1. Coir fibre extraction process and the conventional application of coir fibre and coco peat.

A plant-based fibre cell typically consists of three parts, i.e., primary


cell wall, secondary cell wall, and the hollow lumen (Fig. 5). The pri­
mary cell wall is an easily swellable, isotropic gelatine, which pre­
dominated by amorphous structures such as pectin substances,
hemicellulose, and proteins. Around 15%–40% cellulose (Cosgrove and
Jarvis, 2012) (i.e., crystalline structure) can be found in the primary cell
wall as disorderly arranged scaffold fibrils. The secondary cell wall (S) of
fibre consists of several layers, for example S1, S2, and S3 as illustrated
in the Fig. 5. Cellulose is the main component in the secondary cell wall,
which can be more than 90% of the dry weight of the wall layer (Wang
et al., 2021). In each S layer, the cellulose fibrils are densely packed and
helically arranged. The microfibril angle (MFA) refers to the angle be­
tween cellulose fibrils and the cell long axis, which varies from different
S layers and plant species. Since cellulose fibrils provide most of the
tensile load carrying capacity of a fibre, a fibre with high cellulose
content and low MFA usually exhibits high tensile strength and initial
modulus (Djafari Petroudy, 2017).
Throughout the reviewed investigations, the MFA of coir fibre
ranged from 30 to 49◦ (Gassan et al., 2001; Saw et al., 2012; Sumesh and
Kanthavel, 2020), which was much higher than that of the conventional
plant-based fibres such as flax (6–10◦ ), jute (around 8◦ ), sisal (10–25◦ ),
Fig. 2. The (a) white and (b) brown coir fibre (White and Brown Coir and hemp (around 6◦ ) (Djafari Petroudy, 2017). Table 2 presents the
Fibr, 2021). cellulose content in percentage by weight from various studies (Pache­
co-Torgal and Jalali, 2011; Ramakrishna and Sundararajan, 2005b;
tylosis (AGRIOS, 2005). The tylosis usually consists of cellulose, hemi­ Gassan et al., 2001; Saw et al., 2012; Sumesh and Kanthavel, 2020;
cellulose, pectin, suberin as well as lignin (Micco et al., 2016). During Bayer et al., 2017; Mathura and Cree, 2016; Akinyemi and Adesina,
the formation of tylosis, the pit membrane is stretched and pushed into 2021; JAYABAL et al., 2012; Mohammed et al., 2015) in comparison of
the vessel. Simultaneously, the tylosis protrude into xylem vessels other conventional natural fibres (Djafari Petroudy, 2017). The cellulose
through pit. Tylosis can completely clog the plant vessel and block the content of coir varies from 22% to 44% of the fibre weight. In com­
pathogens from further invasion in the plant. In the case of coir fibre, the parison with flax (64%–71%), jute (61%–72%), sisal (66%–78%), and
tylosis prevents the water exchange by filling the pits on the coir surface. hemp (70%–74%), coir has lower cellulose content. Due to the high MFA
Therefore, this may contribute to the reduction of permeability of coir and low cellulose content, coir exhibits inferior tensile strength but su­
fibre and in turn improve the stability of coir fibre in severe environment perior elongation at failure to other plant-based fibres as shown in
such as alkaline environment. Considering the overall morphology, a Table 1. More details of the tensile behaviour of coir fibre can be found
coir fibre is not usually a single straight fibre but with branches. A in section 2.2.2. Apart from cellulose, hemicellulose and lignin are the
typical branch observed along fibre direction and on the cross-section other two main chemical components of coir fibre. Compared with other
are presented in Fig. 4. The branches are usually thinner than the natural fibres such as flax and sisal, coir fibre usually has a high lignin
main fibre and cannot be avoided when applied as construction mate­ content (40%–45%), which could be even more than cellulose content.
rial. This increases the variation of the mechanical performance of the With the hydrophobic nature of lignin, a high lignin content makes coir
coir fibre itself or the construction materials using coir fibre. fibre more resistant against water and fungi compared to other

3
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 3. (a) Cross-section of coir fibre adapted from Tran et al. (Tran et al., 2015) (permission granted) and (b) coir surface along fibre direction under scanning
electron microscope adapted from Madueke et al. (Madueke et al., 2021).

plant-based fibres. This implies the good potential durability of coir fibre elementary fibres. As a result, the measured fibre volume increased and
as construction materials especially in tropical countries such as Brazil thereby the calculated fibre density decreased. In this case, it is strongly
(John et al., 2005) and Colombia (Quiñones-Bolaños et al., 2021). The recommended that the length of coir fibre should be given for density
hemicellulose in coir fibre only accounts for a small portion, which is measurement with pycnometer or other similar methods based on
usually less than 15%. According to the Bayer et al. (2017), Jayabal et al. Archimedes’ principle (i.e., the buoyancy force of an object immersed in
(JAYABAL et al., 2012), Saw et al. (2012), Sumesh et al. (Sumesh and a fluid is equal to the weight of displaced fluid due to the immersion).
Kanthavel, 2020), and Mohammed et al. (2015), the hemicellulose Different densities should be used in different situations. If the object of
contents of coir were found even less than 1% of the fibre weight. The the investigation is the whole coir fibre (i.e. with lumen and lacuna),
typical thermogravimetric analysis (TGA) with nitrogen atmosphere and coir fibre should not be grinded for density measurement. In some cases
Fourier-transform infrared spectroscopy (FT-IR) analysis of coir fibre are such as tensile strength calculation (section 2.2.1), the density of solid
presented in Fig. 6. Conventionally, three peaks can be found in the material without lumen and lacuna is recommended for the compara­
derivative weight curve (red line in Fig. 6 (a)) of natural coir fibre under bility of all the coir fibre tensile strength. Therefore, the coir fibre should
TGA. The first peak is at around 60 ◦ C–100 ◦ C as water in the coir fibre be cut or grinded for the density measurement.
evaporates. According to Yang et al. (2007), the degradation tempera­
ture of hemicellulose and cellulose is at 250 ◦ C–300 ◦ C and
350 ◦ C–400 ◦ C, respectively. These are also reflected in the second and 2.2. Tensile behaviour of coir fibre
third peak of the derivative weight curve of natural coir fibre in the
Fig. 6 (a). The typical peak location ranges for specific function groups The tensile behaviour is one of the most significant behaviour of coir
were summarized by Bledzki et al. (2010), which are presented in Fig. 6 fibre, especially when coir fibres are used as reinforcement in composite
(b). With a wavelength between 3460 and 3400 cm− 1, a peak is detected materials (e.g., fibre reinforced polymer or CFRC). In this section, the
due to the stretching of O–H bonds. This is a typical characteristic of methodology of tensile test and tensile properties of coir fibres, as well
plant-based fibres (Hasan et al., 2021). When the fibre was treated with as the models to predict the tensile properties of coir/natural fibres are
alkaline such as sodium hydroxide, the hydroxyl groups could be discussed in detail.
removed resulting in the descending of intensity of the O–H peak
(Abraham et al., 2013). 2.2.1. Methodology on tensile test of coir fibre
Since all the plant-based fibres have similar chemical compositions, The most feasible tensile testing setup for coir fibre is the direct
the density of coir (1.15–1.46 g/cm3) is similar to that of other plant- tensile test as illustrated in Fig. 7 (a). Before tension, both sides of a fibre
based fibres such as flax (1.4–1.5 g/cm3), hemp (1.4–1.5 g/cm3), and are gripped by the clamps. Since the plant-based fibres are not homo­
jute (1.3–1.49 g/cm3) (Andiç-Çakir et al., 2014; Dittenber and Gang­ geneous along the fibre length, they cannot be always aligned straight
aRao, 2012; Yan et al., 2014a; Mwaikambo). It is worth noticing that between the clamps of the test machine. Therefore, to adjust the fibre in
some densities of coir measured by the previous study lay far lower than the middle of the clamps to ensure the vertical tension is the difficult
1.15 g/cm3. For example, the natural coir fibre had a density between point for this setup. The authors’ own experience on different natural
0.38 g/cm3 and 0.98 g/cm3 - Mohammad et al. (Hosseini Fouladi et al., fibres showed that fixing a small weight (e.g., a clamp) to the bottom of
2011). Despite of the species variation of coconuts, the measured low the fibre as a plumb bob can solve this problem. However, the additional
density depends on the fibre length during the density measurement. weight provides pre-loading on the fibre. Assuming coir fibre is cylin­
Tran et al. (2015) investigated the relationship between coir fibre length drical, the diameter of coir varies from different fibres, i.e., less than 100
and its density measured by pycnometer. When the coir fibres were μm to more than 400 μm. This leads to a great difference in the pre-
grinded into powder with around 0.05 mm length, the measured density tensile stress in coir fibre. A large additional weight may cause plastic
was around 1.3 g/cm3. The measured density decreased as the fibre deformation of thin fibres and a small additional weight may have
length increased that the measured density of coir fibre was around 0.9 hardly influence on the alignment of thick coir fibres. Additionally,
g/cm3 with the fibre length of 4.0 mm. As explain by Tran et al. (2015), small fibre could be sensitive to clamping, so that the direct tensile test
the increase of coir fibre length raised the enclosed porosity in may not be suitable for the fibre. To solve the above-mentioned prob­
lems, a special sample preparation can be conducted, i.e. gluing the fibre

4
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Nahazanan; Bui et al., 2020). Another tensile test setup is the sample
holder with deflection (Fig. 7 (c)). Basically, this sample holder is
designed for tensile test of yarns. During the test, a yarn is winded on
two cylinders and clamped on both ends. During tension, additional
friction provided by the cylinders can reduce the tensile force in yarn at
the near-clamp area. This helps to avoid the yarn failure at clamps. Since
the diameter of coir can be over 10 times of that of other conventional
fibres (e.g., carbon fibre: 5–7 μm (Acatay, 2017), glass fibre: 5–25 μm
(Middleton, 2016), and flax fibre: 12–16 μm (Yan et al., 2014a)) and the
length of coir fibre (can be more than 200 mm) is sufficient for the setup
of deflection sample holder, this setup is also suitable for coir fibre from
our perspective. It is worth noticing that the plant-based fibres (e.g. coir
and flax) are not cylinder and their shape varies along fibre length. The
diameter of the plant-based fibre is an equivalent diameter of a cylinder
that gives the same volume as the fibre.
The tensile behaviour of coir fibre can be comprehensively described
by the tensile stress-strain curve. The establishment of stress-strain re­
quires the calculation of tensile stress σ and strain ε during the tensile
test. The tensile stress is calculated as follows:
F
σ= (1)
Ae

where F is the load measured by the testing machine and Ae is the


equivalent area of the solid material without lumen and lacuna in coir
fibre. The equivalent area is calculated in the following expression:
m
V = Ae × lf = (2)
ρ

where V is the total volume of the solid material. The lf , m, and ρ


represent the fibre length, mass, and density of solid material without
lumen and lacuna, respectively.
Due to the tensile test setup of coir fibre, extensometer can be hardly
applied to measure the fibre elongation. A calibration process was
needed to generate the fibre stress-strain curve based on the displace­
ment measurement by the testing machine, as reported by Defoirdt et al.
(2010). In the calibration, the total displacement of fibre (the travel of
the crosshead) Δltotal is the sum of fibre elongation Δlfibre , and the
displacement by slippage and test setup compliance Δlnon fibre . At the
elastic stage of the fibre, the fibre obeys the following relationship:
Δltotal Δlnon fibre Δlfibre σ
− = =ε= (3)
ltest ltest ltest E

where E is the fibre elastic modulus, σ is the stress applied on the fibre,
ltest is the initial gauge length for the testing, and Δlfibre is the elongation
of fibre during tension. For infinite long ltest (i.e., 1/ltest = 0), Δlnon fibre
can be ignored in Eq. (3), and the elastic modulus can be directly
calculated by the travel of crosshead Δltotal and the stress σ . This can be
achieved by extrapolating the elastic modulus at 1/ltest = 0 from an
elastic modulus vs. 1/ltest curve.
Assuming that Δlnon fibre and the load on fibre F have a linear relation,
Fig. 4. Coir fibre with branches (a) overview and (b) a close up (red circle area a factor α with a unit of mm/N can be set. For each fibre i, its factor αi can
from the overview of (a)) along fibre direction of a coir fibre, (c) cross-section at be calculated with the following expression:
the branching location of one coir fibre. For the cross-section, a coir fibre was
penetrated with epoxy and polished. αi =
Δlnon fibre, i Δltotal,i − Δlfibre,i
= (4)
Fi Fi
on a paper frame (Fig. 7 (b)). This method is in accordance with DIN EN
where Fi , Δlnon fibre, i , Δltotal,i , and Δlfibre,i are the corresponding load,
ISO 5079 (5079:2021–02, 2021). During the preparation, a paper is
displacement by slippage and test setup compliance, total displacement,
hollowed in the middle to be a paper frame and a fibre is glued on both
and elongation of fibre i, respectively. At linear elastic stage of the fibre
sides of the paper frame. Both glued sides will be then clamped in the
under tensile, Δlfibre,i can be calculated by the extrapolated elastic
testing machine. Shortly before the test, the paper frame will be cut
modulus from Eq. (3). In turn, the factor αi can be determined. Since the
along the cutting line. With this preparation, the fibre at the clamping
factor α should be constant if the fibres are from same species in the ideal
area is strengthened by the glue and the middle adjustment in the clamps
case, it can be estimated by a linear regression or middle value from
can be readily achieved. In most cases of tensile tests for coir fibres, the
factor αi . In this case, the fibre elongation at failure can be determined.
paper frame preparation was frequently applied (Tran et al., 2015;
Instead of the calibration, optical strain mapping system was used in
Defoirdt et al., 2010; Yusoff et al., 2016; Anggraini, Huat, Asadi,

5
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 5. (a) Schematic illustration of plant-based fibre cell structure adapted from Wang et al. (Wang et al., 2021) (b) and (c) the details of coir elementary fibres
under scanning electron microscope from Valášek et al. (Valášek et al., 2018). Permission granted.

the previous research to track the fibre elongation with the help of high- relationships should be further experimentally proved, and the cali­
speed cameras. As reported by Tran et al. (2015), the strain distribution bration method requires validation with more experimental results for
map of coir fibre can be easily obtained by the calculation of differences various fibre species.
between speckle patterns within a series of images. Additionally, Tran
et al. (2015) also used aforementioned calibration method (i.e. Eq. (3) 2.2.2. Tensile properties of coir fibre
and Eq. (4)) for the tensile behaviour of coir fibre. Even though, the coir Fig. 8 illustrates the typical stress-strain curve of a coir fibre in ten­
test length for optical strain mapping (i.e., 5 mm) and calibration sion. The curve can be divided into three stages, i.e. OA, AB, and BC in
method (i.e., 10, 15, 20, 25, and 30 mm) was different, the elastic Fig. 8. At stage OA, coir fibre exhibits a linear elastic behaviour while it
modulus from the mapping (i.e., 4.6 GPa in average) and from extrap­ shows plastic behaviour at stage BC. Between OA and BC, there is a
olation in calibration method (i.e., 4.54 GPa) was similar. transition stage where the strain of coir fibre is usually around 2%–3%
It is worth mentioning that the implementation of calibration (Tran et al., 2015; Teli et al., 2016; Martinschitz et al., 2008). Despite of
methods has several prerequisites, i.e. (1) Δlnon fibre and the load on fibre both linear elastic and plastic behaviour, the MFA of coir fibre was
F, and (2) elastic modulus of the fibre E and the 1/ltest all have linear observed to have a constant decrease as the applied strain on coir fibre
relationship. Linear regression can be then conducted to determine the increased (Martinschitz et al., 2008). This indicates that the change of
factor α and extrapolated fibre elastic modulus. Therefore, these linear MFA due to tension is not the reason for the initiation of the

6
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Table 2 elastic-to-plastic transition. In comparison with an idealized theoretical


Cellulose content of coir fibre in comparison with flax, jute, sisal, and hemp. relationship between MFA and strain (derived from the behaviour of
Reference Year Cellulose Content (%) idealized helical structure), the idealized MFA was found smaller than
the experimental MFA with the same applied strain.
Coir Other fibres
It has been well reported that lignin in a plant acted as a coupling
Gassan et al. (Gassan et al., 2001) 2001 43 agent filling the voids between hemicellulose and cellulose, and mainly
Paramakrishna et al. (Ramakrishna and 2005 33.2
Sundararajan, 2005b)
provided stiffness for plant development and lodging resistance (Liu and
Pacheco-Torgal et al. (Pacheco-Torgal and 2010 21.5 Le LuoZheng, 2018; Yang et al., 2019). The possible explanation of the
Jalali, 2011) stress-strain behaviour with long elongation may lie in the fibre lignin
Jayabal et al. (JAYABAL et al., 2012) 2012 43.4 content as well as morphology and cell wall structure of coir fibre. Under
Saw et al. (Saw et al., 2012) 2012 32.1–34.5
tension, the elementary fibre tends to be stretched along the direction of
Mohammed et al. (Mohammed et al., 2015) 2015 32–43
Mathura et al. (Mathura and Cree, 2016) 2016 33–54 the elementary fibre and contracted in the hoop direction of the
Bayer et al. (Bayer et al., 2017) 2017 32–43 elementary fibre. During the stretching, the pitch of the spring-like
Sumesh et al. (Sumesh and Kanthavel, 2020) 2020 44 microfibril tends to increase and MFA tends to decrease. However,
Dittenber et al. (Dittenber and GangaRao, 2011 64-71 (flax) these increasing and decreasing tendency can be impeded by the lignin
2012) 61-72 (jute)
66-78
existing between two adjacent coils of the microfibril spring. During the
(sisal) contraction, the microfibril is impeded by the cell wall matrix mainly
70-74 through shearing and bonding between them. In the first linear stage
(Hemp) (OA in Fig. 8), the microfibril and the cell wall matrix are well-bonded
and the matrix provides the shear resistance. In the transition stage

Fig. 6. Typical (a) pyrolysis curve under thermogra­


vimetric analysis (TGA) and (b) spectra of Fourier
transform infrared spectroscopy (FT-IR) of natural
coir fibre. The TGA was conducted by the authors at
Department of Organic and Wood-based Construction
Materials, Technische Universitaet Braunschweig.
The TGA atmosphere was nitrogen with a rate of 20
ml/min, and the heating speed was 20 K/min. The
FT-IR was conducted by the authors at Fraunhofer
Wilhelm-Klauditz-Institut (WKI). The description for
FT-IR was adapted from Bledzki et al. (Bledzki et al.,
2010).

7
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 7. Schematic illustration of single fibre tensile setup. (a) direct clamping on fibre, (b) fibre preparation with paper frame before tensile test, and (c) sample
holder with deflection. Summarized from references (Tran et al., 2015; 5079:2021–02, 2021; Defoirdt et al., 2010; Yusoff et al., 2016; Anggraini, Huat, Asadi,
Nahazanan; Bui et al., 2020).

tensile strength, tensile modulus, and the large strain at failure of coir
fibre might be mainly caused by the low cellulose content and large MFA
of the coir fibre. These properties also make coir fibre a proper substi­
tution of steel fibre to promote the ductility of cementitious materials.
More details on the mechanical properties of coir fibre reinforced
cementitious materials are discussed in section 3. Additionally, the
tensile properties of coir fibre can be significantly affected by the gauge
length during the tensile test. Tomczak et al. (2007) investigated the
effect of gauge length on the tensile properties of the Brazilian coir fibre.
As the gauge length increased from 5 mm to 25 mm, the tensile strength
and strain at failure reduced 17% and 58%, respectively. This could be
explained that larger gauge length increased the number of defects or
weak point along the fibre. In the contrary, due to the inhomogeneity
and the multicellular structure of the fibre, the elastic modulus increased
from 1.3 GPa to 2.7 GPa as the gauge length increased from 5 mm to 25
mm.

2.2.3. Models of coir fibre tensile properties


Fig. 8. Typical stress-strain response of coir fibre under tensile. OA: elastic
Since 1900s (Kulkarni et al., 1981; Satyanarayana et al., 1981;
behaviour, AB: transition zone, and BC: plastic behaviour. The transition zone
Bledzki, 1999), the models to describe the tensile properties of
AB can be usually found at around 2%–3% of the ultimate strain. Summarized
from references (Tran et al., 2015; Teli et al., 2016; Martinschitz et al., 2008). plant-based fibres have been widely discussed. A typical model derived
from the fibre structure can be found in Fig. 9. The crystalline fibrils (i.e.,
cellulose in the case of coir) are helically arranged and embedded in
(AB in Fig. 8), cracks initiate in the interface between the microfibril and
non-crystalline region (i.e., lignin in the case of coir). This model de­
the cell wall matrix as well as inside the cell wall matrix. As a result, the
scribes the viscoelastic nature of plant-based fibres and considers only
shear resistance to prevent the contraction of microfibril spring reduces.
two parameters, i.e. the fibre cellulose content and MFA (Bledzki, 1999).
In the second linear stage (BC in Fig. 8), the contraction of cellulose
The overall modulus of a plant-based fibre could be considered as a
spring stops as the hollow area in the elementary fibre (i.e., lumen) is
combined modulus of two materials (springs) that are parallel arranged.
filled or fibre ruptures. Such explanation is however, speculative in
Material I (spring I) describes the tensile behaviour for fibrils with low
nature since no experimental evidence exists. To prove the explanation,
MFA (i.e., less than 45◦ ) while material II (spring II) is for large MFA (i.
the changing of fibre morphology and cell wall structure of coir fibre
e., greater than 45◦ ). Material I (spring I) is based on an isochoric
should be observed during tensile test, which might be achieved by
deformation of both crystalline and non-crystalline region. Material II
micro or nano computed tomography with or without delignification
(spring II) describes the spring-like deformation of the fibrils. The
methods (e.g., using NaOH (Yang et al., 2019)).
expression to calculate the overall modulus Ef of a plant-based fibre can
Table 3 summarizes the tensile strength, tensile modulus, and elon­
be found as follows (Tomczak et al., 2007; Satyanarayana et al., 1981;
gation at failure under tensile of coir fibre. They are compared with the
Bledzki, 1999):
corresponding tensile properties with flax, jute, sisal, and hemp fibres,
which are adapted from Dittenber et al. (Dittenber and GangaRao, 2012) E1 E2
Ef = (5)
and Yan et al. (2014a). The tensile strength of coir fibre lies between 80 E1 + E2
MPa and 304 MPa and the tensile modulus is between 0.9 GPa and 6.2
GPa, respectively. These values are smaller than those of other for
plant-based fibres as shown in Table 3. Conversely, the strain at failure
E1 = [pc Ec + (1 − pc )Enc ]cos 2 θ (6)
of coir fibre can be found between 12.5% and 83.7%, which is higher
than that of flax (i.e., 1.2%–3.3%), jute (i.e., 1%–1.8%), sisal (i.e., 2.0%–
7.0%), and hemp (i.e., 1%–3.5%). As explained in section 2.1, the low

8
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Table 3
The tensile strength, modulus, and elongation at failure of coir fibre in com­
parison with flax, jute, sisal, and hemp.
Reference Year Tensile Tensile Elongation at
strength Modulus failure (%)
(MPa) (GPa)

Tolêdo Filho et al. ( 2000 108–251 2.5–4.5 13.7–41


Tolêdo Filho et al.,
2000)
Mwaikambo et al. ( 2006 175 4–6 30
Mwaikambo)
Li et al. (Li et al., 2007 175 4–6 30
2007b)
Tomczak et al. ( 2007 82.4–195.9 0.9–3.6 12.5–83.7
Tomczak et al., 2007)
Cheung et al. (Cheung 2009 106–175 4–6 14.2–49
et al., 2009)
Defoidt et al. (Defoirdt 2010 120–184 4–6 –
et al., 2010)
Pacheco-Torgal et al. ( 2010 95–118 2.8 –
Pacheco-Torgal and Fig. 9. A typical model to describe the plant-based fibre tensile properties
Jalali, 2011) redrawn from Bledzki et al. (Bledzki, 1999). Permission granted.
Jayabal et al. ( 2012 – 4.5 30
JAYABAL et al.,
2012) where ε is the strain at break of coir fibre and θ is the spiral angle (i.e.,
Saw et al. (Saw et al., 2012 144.6 3.1 32–32.4 MFA). This equation was well agreed to their testing results with a
2012) correlation coefficient of 0.974, as well as the later investigations
Dittenber et al. ( 2012 95–230 2.8–6 15–51.4
Dittenber and
(Tomczak et al., 2007).
GangaRao, 2012) As discussed in section 2.1, coir fibre has a porous structure, with
Al-Maadeed et al. ( 2014 175 4–6 30 large volume of lumens in elementary fibres and lacuna in the middle.
Al-Maadeed and This porous structure eases the contraction of the cell wall of elementary
Labidi, 2014)
fibres towards lumen and the outer elementary fibres towards lacuna.
Yan et al. (Yan et al., 2014 95–230 2.8–6 15–51.4
2014a) This may strongly influence the tensile behaviour of coir fibre, which
Tran et al. (Tran et al., 2014 150–250 3.5–5.7 18–36.7 was not considered in the classical model as shown in Fig. 9. The
2015) regression-based model in Eq. (8) considered the MFA as the only
Lecompte et al. ( 2015 80–152 3.6–6.2 – parameter that affected the strain at break. However, the strain at break
Lecompte et al.,
2015)
should be also affected by the bond between lignin and micro-fibrils or
Mohammed et al. ( 2015 220 6 15–25 the porous structure of coir fibre. The major drawback of the
Mohammed et al., morphology-based models are the difficulties of measuring individual
2015) properties, such as E-moduli and MFA, which by themselves are random
Pickering et al. ( 2016 131–220 4–6 15–30
variables. In our opinion, such measurements are often impossible to
Pickering et al.,
2016) conduct with reasonable amount of reliability and accuracy. The authors
Gupta et al. (Gupta and 2016 175 4–6 30 question the applicability and practicality of such models that are based
Srivastava, 2016) on significant simplifications and approximations. To sum up, despite of
Sumesh et al. (Sumesh 2020 94 4.1 – the aforementioned classical model for plant-based fibres before 21st
and Kanthavel, 2020)
Akinyemi et al. ( 2021 108–251 2.5–4.5 13.7–41
century, the model specifically for coir fibre should be further developed
Akinyemi and towards the understanding and safety design of coir fibre for construc­
Adesina, 2021) tion applications.
Dittenber et al. ( 2011 343-2000 27.6–103 1.2–3.3 (flax) Unlike fibre modulus or strain at break, strength could not be simply
Dittenber and (flax) (flax) predicted based on the cellulose content and spiral angle, since defects in
GangaRao, 2012) 320-800 30 (jute) 1–1.8 (jute) fibre could significantly affect the fibre strength as discussed in section
(jute) 2.2.2. Throughout the literature review of the authors, the prediction of
363-700 9-38 (sisal) 2-7 (sisal)
(sisal)
coir tensile strength has not been well discussed, but the prediction of
270-900 23.5–90 1–3.5 (Hemp) tensile strength from other plant-based fibres has been reported. For
(Hemp) (Hemp) example, Weibull analysis has been widely used for plant-based fibres
(such as jute (Defoirdt et al., 2010; Xia et al., 2009; Roy et al., 2012;
Alves Fidelis et al., 2013), bamboo (Defoirdt et al., 2010; Wang and
Knc ( )2
E2 = 1 − 2cot 2 θ (7) Shao, 2014; Trujillo et al., 2014), and sisal (Alves Fidelis et al., 2013;
1 − pc
Amico et al., 2005)) with linear elastic tensile behaviour, assuming that
the total failure was caused by the failure at the most critical flaw.
where E1 , and E2 are the Young’s modulus of material I and II. pc is the
Defoirdt et al. (2010) used Weibull distribution to assess the tensile
content of crystalline part. Knc is the bulk modulus of non-crystalline
strength of coir fibre and compared it with bamboo and jute fibres. They
part. θ is the spiral angle (i.e., MFA).
pointed out that due to the noticeable plastic deformation under tensile,
The strain at break of coir fibre could be predicted by Kulkami et al.
coir was not justified to be used with Weibull distribution. The mean and
(Kulkarni et al., 1981) in 1981. This regression-based model is presented
standard deviation for normal and Weibull distribution were found
in Eq. (8).
almost equal. The Weibull parameters were different for different test
θ2 lengths, which was not justified as well. Therefore, the strength pre­
ε = 1.24 + (8)
60 diction based on statistic distribution should be further investigated. A
normal distribution could be sufficient and suitable for the strength

9
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

prediction of coir fibre since the possibility of defect along the testing The incorporation of dispersed short fibres in concrete to form the
fibre could obey normal distribution. Hence, it is strongly recommended fibre reinforced concrete (FRC) was firstly introduced by Romualdi et al.
to investigate the applicability of various statistics for the coir fibre (Romualdi and Mandel, 1964) in early 1960s. The main role of those
strength. fibres is to control the crack opening and propagation. The mechanism
of fibre function in concrete is illustrated in Fig. 11. Under tension (i.e.,
3. Coir fibre reinforced cement-based composite materials load direction under tensile test, hoop direction under compression test,
or bottom part of the beam under bending test), tensile strength of
There are several configurations of coir fibres that were reported to concrete is reached easily, and cracks are initiated. A crack process
be able to reinforce the cement-based composite materials, e.g. coir fibre (Fig. 11 (a)) usually consists of three parts, i.e. macro-cracking, aggre­
rope (Ali and Chouw, 2013; Ali, 2018), coir mesh (Li et al., 2007a), and gate interlocking, and micro-cracking. In the macro-cracking area, no
monofilament fibre reinforcement. To produce the coir rope and mesh, interaction can be found between both sides of the crack so that no
coir yarn is used, which is twisted by a batch of coir fibres. The me­ tensile stress occurs. In the aggregate interlocking area, small crack
chanical performance of coir rope and mesh is highly dependent on the width enables the contact and interlocking between aggregates. Hence,
friction between the twisted fibres, determined by the fibre surface a small amount of tensile stress can be carried at this area. In the
roughness and the contact area. In general, the longer the fibre is, the micro-cracking area, not only the tensile stress can be carried by the
larger contact area can be found between fibres. Coir fibre is a seed fibre interlocking between aggregates but also the path flow of tensile stress
that is usually short and thick, i.e. around 3.5–15 mm in length and can go around the crack tips. Therefore, higher tensile stress can be
150–400 μm in diameter according to Ashby (2013). The stem fibres are recorded in this area compared to the rest of the crack. With the addition
more suitable to be used as yarn as they are long and thin. For example, of short fibres, both sides of the crack can be bridged by the fibres so that
the length of flax fibre could be up to 900 mm and its average diameter the tensile stress is readily transported along the fibre. This results in an
was 12–16 μm. In comparison with other stem fibres such as flax, coir overall increase in tensile stress along the crack in comparison with the
fibre is more suitable to be used as monofilament fibre reinforcement in plain concrete (Fig. 11 (b)). This effect of the fibre in the concrete is
the cement-based composite materials and this was the most commonly referred to as “fibre bridge effect”.
used configuration that has been investigated in the past 20 years. In The typical failure modes of fibre in cementitious matrix are illus­
addition, coir fibre had the highest toughness in tension among trated in Fig. 12. The le and w represent the effective embedded length
plant-based fibres, resulting in a wide investigation on coir fibre as and the crack width, respectively. If the embedded length of a straight
monofilament reinforcement for concrete structure (Ali et al., 2012). fibre in concrete (at least one side) is smaller than le , the fibre will be
Therefore, in this review paper, the focus is on the coir fibre as mono­ pulled out (Fig. 11 (a)). If the fibre embedded length is equal to or larger
filament fibre reinforcement in cement-based composite materials. than le , fibre rupture occurs (Fig. 12 (b)). In some cases, especially for
steel fibre, fibre rupture still occurs even though the embedded length of
fibre is smaller than le (Fig. 12 (c)). This is due to the modification of the
3.1. Mechanism of fibre in cement-based composite materials fibre (e.g., crimped, hooked fibre, or fibre with button end (Brandt,
2008)). Generally, the effective embedded length le depends on the
Cement-based composite materials, for example concrete, usually interfacial bonding and interlocking between fibre and cementitious
exhibit high compressive strength but low tensile strength and brittle matrix. The better the interfacial bonding and interlocking are, the
failure mode. Adding dispersed short fibres in concrete is regarded as larger is the le . Yan et al. (2015) have investigated the surface
one of the efficient methods to increase the tensile strength and change morphology of coir fibre and the fractured surface of CFRC under SEM.
the brittle failure of concrete into ductile. Typical tensile stress-strain Fig. 13 (a) and (b) shows the coir fibre surface before and after mixing
curve of fibre reinforced concrete in comparison with plain concrete is with concrete, respectively. It was found that a thin layer of hydration
illustrated in Fig. 10 (a). The concrete tensile strength (i.e., the product, i.e. portlandite (calcium hydroxide), was generated on the coir
maximum stress) and fracture energy (i.e., the area under the stress- fibre surface. This can be observed in the close-up SEM figure of coir
strain curve) of fibre reinforced concrete are higher than the plain fibre surface after concrete mixing in Fig. 13 (c). The majority of the coir
concrete. According to Marcalikova et al. (2020), the type of the FRC surface was covered with the hydration product that coir fibre surface
was related to the fibre volume fraction, concrete recipe and mixing and pits with tylosis could be still detected. After the bending test, both
process. In the case of coir fibre, the splitting tensile load-time behaviour fibre pull-out failure and fibre rupture (Fig. 13 (d)) were found at the
of coir fibre reinforced concrete showed the similar behaviour of FRC fractured surface of coir-concrete. These failures could result from the
type 1 in Fig. 10 (a), according to the previous limited research (Fig. 10 difference of fibre surface condition, fibre/matrix adhesion, or the
(b)) (Ali et al., 2012).

Fig. 10. (a) typical tensile stress-strain curve of fibre reinforced concrete, redrawn from Marcalikova et al. (Marcalikova et al., 2020). PC stands for plain concrete,
FRC stands for fibre reinforced concrete. FRC type 1 and 2 refer to low and high-volume fibre fraction respectively. (b) Splitting tensile load-time curve of coir fibre
reinforced concrete from Ali et al. (Ali et al., 2012) (permission granted). CFRC stands for coir fibre reinforced concrete and PC stands for plain concrete.

10
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 11. Schematic illustration on crack in (a) plain concrete and (b) fibre reinforced concrete with the corresponding tensile stress in the crack along the crack
length redrawn from Alfes et al. (Alfes and Wiens, 2010).

3.2.1. Methodology of pull-out test of fibre from cementitious matrix


The interfacial properties of fibre in cementitious matrix are
conventionally detected by fibre pull-out test. Generally, these methods
can be classified into two categories, i.e. matrix-fibre method (Ram­
akrishna and Priyadharshini, 2018; Lecompte et al., 2015) (Fig. 14 (a))
and matrix-fibre-matrix method (Bhutta et al., 2017; Farooq and Ban­
thia, 2019) (Fig. 14 (b)). The matrix-fibre method is a single fibre
pull-out method that the fibre is directly pulled out from the matrix. The
matrix will be either pulled from the bottom or pushed from top. The
matrix-fibre-matrix method tests the overall bonding strength of one
fibre or several fibres. During the test, two cementitious blocks will be
connected only with one fibre or several fibres. One side of the
embedded length of the fibres should be larger than the other side so that
the fibres are ensured to be pulled out from one side (bottom side in the
case of Fig. 14 (b)). Comparing both methods with each other, the
matrix-fibre method is easier to be conducted. However, controlling the
fibre parallel to the tensile direction to avoid additional moment is
Fig. 12. Typical failure mode of fibre in cementitious matrix: (a) fibre pull-out
impossible especially in the case of plant-based fibres with inevitable
failure with insufficient embedded length, (b) fibre rupture with embedded
natural curvature. Additionally, the clamping from the test machine may
length equal to or larger than effective embedded length, and (c) fibre rupture
with fibre modification redrawn from Thienel et al. (Thienel, 2020). The le and
crush the fibre. The fibres in the matrix-fibre-matrix method are pro­
w represent the effective embedded length and the crack width, respectively. tected by the matrix against clamping. No additional moment is needed
to be considered for the overall setup since the component forces from
the fibres in the plane of A-A in Fig. 14 (b) are in all directions and the
position of crack initiation.
horizontal component of force can be cancelled out with each other. The
overall pull-out performance of the setup is a cumulative performance of
3.2. Interfacial properties of fibre in cementitious matrix individual fibres that no specific testing machine with sensitive load cell
is needed. However, the setup of the matrix-fibre-matrix is complex that
According to the failure modes of fibre in cementitious matrix, a during the matrix casting, the cross-section A-A in Fig. 14 (b) should be
conclusion can be drawn that the mechanical performance of fibre free of matrix. Since plant-based fibres usually have defects along the
reinforced cementitious matrix depends on the fibre tensile properties fibre, fibre rupture may occur during the pull-out testing even though
and interfacial properties of the fibre in the matrix. The fibre tensile small embedded length is applied. Hence, the pull-out performance ac­
properties have been discussed in section 2.2. In this section, the inter­ quired from the matrix-fibre-matrix method may not reflect the real
facial properties of fibre in matrix will be extensively described.

11
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 13. The SEM micrograph of (a) coir fibre before concrete mixing, (b) coir fibre after concrete mixing, (c) a close-up of coir fibre after concrete mixing, and (d)
fractured surface of coir-concrete after bending test adapted from Yan et al. (Yan et al., 2015). Permission granted.

Fig. 14. Setup for fibre pull-out test: (a) matrix-fibre method and (b) matrix-fibre-matrix method.

pull-out performance of a fibre from cementitious matrix. the results from this modified method showed the same trend that the
Considering that plant-based fibres can be easily failed before pull- fibre with higher average peak pull-out load in the standard methods
out, Naik et al. (2019) proposed a modified method for single fibre had higher average peak pull-out load in the modified method as well.
pull-out test. Instead of consolidated matrix, fresh cementitious matrix Despite of such trend, Naik et al. (2019). also suggested to have more
was used in their investigation. The pull-out test was conducted directly investigations to validate the feasibility of their modified method for
after the casting of the matrix, i.e. the fibres were directly pulled out single fibre pull-out test.
from fresh cement mortar. Comparing with the standard pull-out test,

12
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

3.2.2. Pull-out behaviour of fibre from cementitious matrix follows:


The typical pull-out load-displacement curve of fibre from cementi­
Fmax
tious matrix is schematically illustrated in Fig. 15 (a). With an increase τmax = (9)
π dl
in the applied force, the fibre and the matrix will go through well-
bonded stage (segment OA), debonding stage (segment AB and BC), where d is the fibre equivalent diameter considering lumen and lacuna,
and pull-out stage (segment CD). In the well-bonded stage (segment and l is the embedded length of fibre in the matrix. The equation above
OA), the fibre free length that is not embedded in the matrix stretches assumes perfect contact between the matrix and fibre and models the
elastically. From our opinion, the interfacial performance between fibre fibre as an ideal cylinder. This means that one has to have a reliable
and matrix at this stage is dominated by the chemical bond (adhesion). method to convert the contact surface into an equivalent cylinder with
The mechanical bond (static friction) is equal to zero until point A where the same contact area. Even if this were possible, the true contact area is
an interfacial crack is initiated. In the debonding stage (segment AB and hard to estimate due to the non-prismatic fibre shape and varying sur­
BC), interfacial debonding occurs resulting the reduction of adhesion face properties (such as roughness or even chemical properties) along
and increase of static friction. At point B, the fibre intact embedded the fibre.
length is equal to the critical fibre embedded length. At point C, fibre The total energy Wt is the area under the load-displacement curve
and matrix are fully debonded that no interfacial adhesion remains. In for the entire pull-out test (i.e., area OBCD in Fig. 15). It is the sum of
the pull-out stage (segment CD), fibre is pulled out with the effect of debonding energy Wd (i.e., area OBCE in Fig. 15) and the pull-out en­
dynamic friction. During this stage, three mechanisms (i.e., slip- ergy Wp (i.e., area ECD in Fig. 15).
hardening, constant friction, and slip-softening) can take place. A fibre The pull-out behaviour of coir fibre was not frequently but
surface with high roughness (part ① in Fig. 15 (a)), and similar hardness comprehensively discussed and investigated in the previous studies.
or shear modulus to as the matrix surface (part ② in Fig. 15 (a)) can Ramakrishna et al. (Ramakrishna and Priyadharshini, 2018) has inves­
again be blocked by parts of the surrounding matrix during pull-out, tigated the bond strength of coir in correlation with curing days, water
resulting in a slip-hardening mechanism. If the fibre surface has hard­ cement ratio (w/c ratio), and the mixing design (i.e., cement to fine
ness and shear modulus much significantly different from the sur­ aggregate ratio). The results are summarized and illustrated in Fig. 16
rounding matrix, fibre (part ① in Fig. 15 (a)) or the surround matrix (a), (b), and (c), respectively. The bond strength at one day curing, with
(part ② in Fig. 15 (a)) can be easily damaged. This results in less contact w/c ratio of 0.3, and with mixing design of 1:3 are set as the benchmarks
between fibre and matrix and in turn a slip-softening occurs. If the in each figure, respectively. The rest of the results are compared with the
reduction of dynamic friction due to pull-out is constantly reduced, the corresponding benchmarks. With the curing days of mortar increased
constant friction mechanism occurs (Naik et al., 2019). For the soft and from one day to 28 days, the bond strength of coir had a reduction of
ductile fibres (e.g. polypropylene fibre), alternate slip-hardening and around 22%. This reduction of bond strength was explained due to the
slip-softening could be observed in the pull-out stage (Fig. 15 (b)) (Naik degradation of plant-based fibre in the pore solution, which is exten­
et al., 2019). The abrasion of soft fibre surface against hard matrix led to sively explained in the section 3.3.3. The bond strength also decreased if
the peeling of small fibrils. With the increasing of slip between fibre and more fine aggregate and less cement were in the mortar mixture. A bond
matrix, the quantity of peeled fibres increased. These peeled fibres strength reduction of around 23% was found in mixing design of cement
jammed together and improve the resistance of the pull-out load. This to fine aggregate ratio at 1:5 in comparison with that at 1:3. The reason
resistance would reduce after the failure of the jamming part or the was not mentioned in the paper. From our perspective, this could be
protruding part of surrounding matrix (part ③ in Fig. 15 (b)). explained that the interlocking between aggregate and fibre may have
The maximum load, total energy, and the interfacial shear strength less influence on the bond strength than the adhesion bonding due to
are the three most important properties to describe the fibre pull-out cement hydration process. However, further validation should be con­
behaviour. The maximum load is the load at point B. The interfacial ducted to support this statement. The bond strength did not have
shear strength τmax can be calculated by the maximum load Fmax as

Fig. 15. (a) Typical pull-out load-displacement curve of fibre from matrix redrawn from Naik et al. (Naik et al., 2019). Permission granted. (b) Abrasion of soft fibre
against hard matrix.

13
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 16. The relationship of coir fibre bond strength


in percentage and (a) curing days, (b) water cement
ratio (c) mixing design (i.e., cement to fine aggregate
ratio), from Ramakrishna et al. al. (Ramakrishna and
Priyadharshini, 2018), and (d) embedded length from
Ramakrishna et al. (Ramakrishna and Priyadharshini,
2018) and Ali et al. (Ali et al., 2013) The bond
strength with day 1, w/c ratio of 0.3, mixing design of
1:3, and embedded length of 40 mm are set as the
benchmark in (a), (b), (c), and (d), respectively. The
rest results are compared with the corresponding
benchmark.

constant reduction along with the increase of w/c ratio. The highest diameter is an equivalent diameter of a cylinder with the same volume
bond strength appeared when w/c ratio was equal to 0.4, i.e. around as the one of the coir fibre. The calculated contact area of the coir and
66% higher than that at w/c ratio of 0.3. As explained by Ramakrishna matrix interface was different from the true contact area. This also had
et al. (Ramakrishna and Priyadharshini, 2018), this highest bond influence on the calculated bond strength of coir fibre. Additional to
strength was due to the reach of normal mortar consistency at w/c ratio aforementioned parameters, Ali et al. (2013) found out that the fibre
of 0.4. Since this statement was not proved by the authors, it is recom­ pre-treatment significantly improved the bond strength of coir fibre in a
mended that the relationship of coir fibre bond strength and the con­ concrete matrix. The coir fibre bond strength was improved by about
sistency of mortar/concrete should be investigated. It is worth to 180% when the fibre was boiled for 2 h and washed. During the
mention that the bond strength in the study of Ramakrishna et al. pre-treatment process, the extractives could be washed away. The
(Ramakrishna and Priyadharshini, 2018) was calculated by the equivalent diameter could reduce as a results of the loss of extractives.
maximum load measured during the pull-out test. In some cases from Hence the calculated bond strength increased after Eq. (10).
their study, the fibre rupture failure occurred rather than pull-out fail­ Lecompte et al. (2015) investigated coir fibre pull-out shear strength
ure, e.g. the samples with 50 mm embedded length at w/c ratio of 0.5 with additional consolidation pressure (i.e., a pressure applied to the
with mixing design of 1:4, or the samples with 40 mm embedded length mortar during the mortar consolidation with 0.5 MPa, 0.75 MPa, and 1
with mixing design of 1:3. In such cases, the actual bond MPa) during the casting of fresh mortar (i.e., a mixture with cement to
strength/pull-out shear strength should be higher than the tested results. kaolin ratio of 1:1 and w/c ratio of 0.4). With the increase of consoli­
Hence, the readers must be cautioned that the presented results are dation pressure, the pull-out bond strength increased by around 50%.
specific in character and may not present the general bonding behaviour The addition of consolidation pressure provided a mechanical adhesion
between coir fibre and cementitious matrix. Despite of that, fracture between fibre and matrix through the residual radial stress along the
failure occurred more frequently (i.e., pull-out shear strength reduced) fibre after consolidation and the fracture coefficient of the interface.
when curing days increased, less fine aggregate was added, or the With such results, Lecompte et al. (2015) suggested to use an extrusion
embedded length (from 25 mm to 50 mm) increased in the case of w/c process rather than casting for coir fibre reinforced cementitious con­
ratio of 0.5. struction materials to optimise the interfacial bonding between fibre and
The effect of coir embedded length in mortar/concrete on pull-out matrix.
strength was investigated by Ramakrishna et al. (Ramakrishna and The bond strength τmax and the total energy Wt of coir pull-out from
Priyadharshini, 2018) and Ali et al. (2013), which is presented in Fig. 16 concrete have been provided by Ali et al. (2013) through the following
(d). In Fig. 16 (d), the bond strength at embedded length of 40 mm is empirical equations.
considered as benchmark since this embedded length was studies in both [ ]
τmax = αβ al3 + bl2 + c(l − 1) (10)
references. Based on the figure, a conclusion can be drawn that the
optimal embedded length providing the highest bond strength was be­ [ ]
tween 20 mm and 40 mm, i.e., the bond strength of embedded lengths of Wt = γλ pl3 + ql2 + rl + s (11)
25 mm and 30 mm was about 19% and 11% higher to that of 40 mm,
where α and γ are factor for coir fibres, while β and λ are matrix factors
respectively. With the increase of fibre embedded length, more fibre
with:
rupture was observed, which was not considered in the bond strength

calculation. The bond strength of the ruptured fibres should be higher ⎨ 1 medium and soaked fibre
than those fibres which were pulled out. Therefore, the bond strength α = 2.85 thick and soaked fibre

with large embedded length could be underestimated. Another reason 7 thick and boiled fibre
could lie on the diameter of the coir fibre. As already discussed, the

14
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676


⎨ 1 medium and soaked fibre relatively good thermal and sound insulation properties. Bui et al.
γ = 2.65 thick and soaked fibre (2020) investigated the thermal conductivity of coir fibre mat using heat

7 thick and boiled fibre flow meters. The measured thermal conductivity was between 0.024
{ W/mK and 0.052 W/mK, which were comparable to the design value of
0.43 cement : fine aggregate : coarse aggregate = 1 : 4 : 4 conventional thermal insulation materials. Additionally, Bui et al.
β=
1 cement : fine aggregate : coarse aggregate = 1 : 3 : 3 (2020) found that the pre-treatment (i.e., fibre boiling for 2 h or im­
{ mersion in 5% NaOH for 30 min) had no significant effect on the coir
λ=
0.35 cement : fine aggregate : coarse aggregate = 1 : 4 : 4 fibre mat (density from 30 to 120 kg/m3) thermal conductivity. Taban
1 cement : fine aggregate : coarse aggregate = 1 : 3 : 3 et al. (2019) investigated the acoustic absorption ability of coir fibre mat
a, b, c, p, q, r, and s are constants with the value of − 0.0166, 0.112, with a density of 130 kg/m3. The acoustic absorption coefficient
− 0.188, − 10.4, 81, − 150, and 88, respectively. For the fibres, medium measured by impedance tube was considered to describe the acoustic
and thick refer to the fibre with the diameter range of 0.15 mm–0.20 mm absorption ability of the material. The higher the acoustic absorption
and 0.3 mm–0.35 mm, respectively. Soaked and boiled fibres indicate coefficient was (close to 1), the material could have better sound ab­
the pre-treatment of fibre with soaking in tap water, or firstly soaked and sorption ability. It was found that coir fibre mat had a relative low ab­
then in boiling water for 2 h, respectively. Compared with the experi­ sorption coefficient at low frequency, e.g. the acoustic absorption
mental results, the results from these empirical expressions showed a coefficient was 0.34 for the coir fibre sample with 35 mm thickness at
great agreement that the difference between experimental and calcu­ 1000 Hz. Significant increase in acoustic absorption coefficient could be
lated results were less than 3%. However, from our perspectives, such found for higher frequencies, e.g., the acoustic absorption coefficient
was more than 0.9 for the coir fibre sample with 25 mm thickness at
good agreement is because that the empirical equations were generated
based on the experimental results. Therefore, the feasibility of such 2000 Hz. In addition, the thickness of coir fibre mat had significant in­
fluence on the sound absorption coefficient especially for low frequency.
equations should be validated with further experiments on coir fibre
pull-out tests. The coir fibre mat with 25 mm, 35 mm, and 45 mm thickness had a
sound absorption coefficient of 0.11, 0.34, and 0.97 at 1000 Hz,
respectively. Similar results could be found in the work by Hosseini
3.3. Physical and mechanical properties of coir fibre reinforced cement- Fouladi et al. (Hosseini Fouladi et al., 2011). An average acoustic ab­
based composite materials sorption coefficient of 0.8 could be achieved for the coir fibre mat
thickness of 20 mm (mat density: 130 kg/m3), 30 mm (mat density 153
3.3.1. Physical properties of coir fibre reinforced cement-based composite kg/m3), and 45 mm (mat density: 159 kg/m3) at sound frequency of
materials higher than 1360 Hz, 940 Hz, and 578 Hz, respectively.
CFRC also shows good insulation ability from coir fibre. With 5%,
3.3.1.1. Light-weight construction materials. Comparing with the 10%, and 15% coir fibre to binder in weight, Lertwattanaruk et al.
cement-based composite materials, coir fibre has lower density. As dis­ (Lertwattanaruk and Suntijitto, 2015) discovered a reduction of 40%,
cussed in Section 2.1, the density of coir ranges from 1.15 to 1.46 g/cm3 44%, and 46% in the thermal conductivity of cementitious matrix,
(for solid material in coir). The density of normal strength concrete is respectively. With high content of coir fibre (e.g., the ratio of cement to
around 2.40 g/cm3. The addition of coir fibre in cementitious materials coir was 1:2 in the case of Asasutjarit et al. (2007)), the thermal con­
can reduce the density of the concrete. Existing investigations showed ductivity could be around 0.0547–0.1205 W/mK. These values were
that with a coir fibre addition less than 5% of the cement weight (i.e., the close to 0.1 W/mK, which was regarded as upper threshold for the
conventional coir fibre content for structural application), the reduction identification of insulation materials according to Bui et al. (2020). The
of density of the resulting concrete was up to 12% (Ali et al., 2012; sound insulation ability of CFRC was not thoroughly investigated in the
Lertwattanaruk and Suntijitto, 2015; Li et al., 2006; Maia Pederneiras previous studies. Based on Olukunle et al. (2018), the sound decibel
et al., 2021). Asasutjarit et al. (2007) measured the density of coir fibre reduced around 5% in normal mortar compared with the sound input,
board in which the cement to coir fibre ratio in weight was 1:2. With while that for 0.25%, 0.5%, and 0.75% of addition coir had a reduction
such high content of coir fibre, the density of the board could be as low of 12%, 18%, and 19%, respectively. It was also found that the sound
as 0.25 g/cm3 due to the porous structure of coir fibres. However, adding absorption ability had no significant difference with different curing
fibres could reduce the workability of cementitious matrix and increase days (i.e., 7–28 days).
the porosity of the coir fibre reinforced cementitious matrix. Coir fibres Even CFRC has good thermal and sound insulation ability, it was
or other natural fibres are hydrophilic, which absorbs moisture from the found out that most of the studies only concentrated in its thermal
cementitious matrix (Onuaguluchi and Banthia, 2016). The amount of insulation ability. Its sound insulation ability was not comprehensively
free water will decrease resulting in lower workability. The addition of investigated. Since coir fibre mat is good at sound absorption especially
fibre will increase the total surface area (including surface of aggregate for high frequency, more investigations should be done on sound insu­
and fibre) which should be covered by the cementitious mortar. As a lation ability of CFRC. If the CFRC is comparable to other sound insu­
result, the cementitious mortar on aggregates and fibres become thinner. lation materials, the application of CFRC can be further extended. The
Gaps will be generated between the mortar layers, which will be filled increasing utilization of coir fibre will push us a step closer towards the
with air. In order to have the same workability (i.e., flow value of the global sustainability and circularity.
mixture was 16 ± 1 cm), 80.3%, 96.7%, and 221.3% more super­
plasticizer (i.e., modified polycarboxylates as chemical base) were 3.3.2. Short-term mechanical properties of coir fibre reinforced cement-
added by Andiç-Çakir et al. (2014) in the mixture with 0.4%, 0.6%, and based composite materials
0.75% of coir fibres by weight of the total mixture, respectively. Addi­ Throughout the literature, most of the short-term mechanical in­
tionally, they discovered that soaking coir fibre in 5% NaOH solution for vestigations on CFRC were for compressive and bending. The
2 h could improve the workability for the fibre-mortar board. Only compressive and flexural strength are respectively summarized in
19.7%, 41.0%, and 141.0% more superplasticizers were respectively Fig. 17 (a) and (b). In the figures, the corresponding property of
needed in the mixture with 0.4%, 0.6%, and 0.75% of coir fibres by cementitious mortar without coir fibre is regarded as control value. The
weight of the total mixture to achieve the flow value of 16 ± 1 cm. property in percentage is the quotient of test value to control value. With
the increase of coir fibre content, the compressive strength increases at
3.3.1.2. Thermal and sound insulation. A porous structure with a first and then decreases. Similar tendency was often discovered by
porosity around 21%–31% (Tran et al., 2015) gives the coir fibres cementitious matrix reinforced with other sorts of fibres, such as steel

15
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 17. The relationship of CFRC (a) compressive strength and (b) flexural strength in percentage with coir fibre content to binder in weight (Ali et al., 2012;
Andiç-Çakir et al., 2014; Lertwattanaruk and Suntijitto, 2015; Li et al., 2006; Maia Pederneiras et al., 2021; Yan et al., 2016a). The compressive and flexural strength
of cementitious matrix without fibre with the corresponding condition is considered as the benchmark. A horizontal line (red line) is set at 100% to clearly show the
improvement or decline of the strength by coir fibre addition.

fibre (Saidani et al., 2016), basalt fibre (Katkhuda and Shatarat, 2017; the compressive strength of cementitious mortar. Different from
Dong et al., 2017), or polypropylene fibre (Saidani et al., 2016). This is compressive strength, the addition of coir fibre can improve the mortar
due to the change of workability through fibre addition that additional flexural strength even with low coir fibre content (coir fibre content of
voids can reduce the compressive strength. With a coir fibre content to 0.26% in Fig. 17 (b)). Since pure mortar usually has low tensile strength,
binder in weight of around 1%–5%, the addition of coir fibre improves cracks can be easily generated in the tension area of beam under

16
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

bending. As a result, the tensile loading is transferred to fibre in the NaOH solution for 2 h) on coir fibre could improve the fracture tough­
mortar due to the fibre bridge effect. With sufficient bonding between ness by around 8%–16% of the CFRC with untreated fibre. It should be
fibre and matrix, the flexural strength of fibre reinforced mortar can still pointed out that the fracture toughness is not a material property – the
be improved even if more voids exist. However, if too many fibres are toughness measurements are equipment and geometry dependent
added in the mixture (e.g., 5%, 10%, 15% in Fig. 17 (b)), additional (Polocoșer et al., 2017) but can be used to compare materials.
surface area due to the fibre is too large to be covered by the mortar The impact resistance of cementitious materials can be significantly
paste. Therefore, the flexural strength reduces due to the insufficient improved by the addition of monofilament short fibres due to the fibre
bonding. In addition to fibre content, fibre pre-treatment and fibre “bridge effect”. Ramakrishna et al. (Ramakrishna and Sundararajan,
length also affect the compressive and flexural strength. After immersing 2005a) investigated the effect of different natural fibres with various of
coir fibre (length of 20 mm) in 5% NaOH solution for 2 h, the results fibre content and length in cement mortar. In the study, sisal, coir, jute,
from Andiç-Çakir et al. (2014) showed that the mortar with treated fibre and Hibiscus Cannubinus fibres with fibre length of 20 mm, 30 mm, and
had slightly higher compressive strength and flexural strength than that 40 mm and fibre content of 0.5%, 1%, 1.5%, and 2% by cement weight
with untreated coir fibre. However, with fibre length of 40 mm and were considered. The mixing design of mortar was 0.47:1:3 by water:
pre-treatment with 1% NaOH solution for 48 h, no significant difference cement: sand. A metal ball weighing 475 g was set at 200 mm fall height
could be found between cementitious mortar reinforced by coir fibre for the impact test. The impact energy absorbed by the samples until
with/without pre-treatment - Li et al. (2006). The average flexural ultimate failure was one of the indicators for the impact resistance. The
strength of mortar with treated coir fibre was 7.47 MPa while that with higher the impact energy absorption, the greater was the impact resis­
untreated fibre was 7.53 MPa. Due to the limited literature, it is still tance of the sample. The impact energy absorptions of all the fibre
unknown how NaOH or other pre-treatment on coir fibre can affect the reinforced cement mortar are presented in Fig. 18. Among all the fibre
mechanical properties of CFRC. A proper pre-treatment should be reinforced cement mortar, the lowest impact energy absorption was
carefully selected to improve the mechanical properties of the CFRC and found by the sample of Hibiscus Cannebinus with 20 mm and 0.5% fibre
in turn promote it structural application. Large coir fibre length usually content, i.e., 25.2 J. Despite of that, this impact energy absorption was
had negative effect on the mechanical performance of CFRC. Ali et al. still around double of that without fibre reinforcement (i.e., 14.0 J),
(2012) investigated the compressive strength of CFRC with fibre length indicating that fibre reinforcement could strongly improve the impact
of 25, 50, and 75 mm. For fibre length of 25 and 50 mm, no significant resistance of cement mortar. Additionally, the impact energy absorption
difference can be found in the compressive strength of CFRC. With both increased as fibre length or fibre content increased. This was mainly due
lengths, the compressive strength of the matrix was improved by around to the higher interlocking possibilities as longer or more fibres were
22%, 16%, and 12% in average for fibre content of 1%, 2%, and 3%, added in the matrix. Among all four plant-based fibres, coir fibre
respectively. For fibre length of 75 mm, the compressive strength was exhibited significant improvement on the impact resistance, especially
only improved by fibre content of 1% with an increase of 9%. With fibre with the fibre contents of 1.5% and 2%. For fibre length of 40 mm and
content of 2% and 3%, the compressive strength of matrix was reduced fibre content of 2%, the impact energy absorption of coir fibre-cement
by around 6% and 10%. Large fibre length can be easily intertwined mortar was 253.5 J, which was around double of that with sisal fibre
during the mortar mixing, and in turn reduce the workability and in­ (i.e., 121.2 J) and four times of that with jute (i.e., 68.0 J) and Hibiscus
crease the porosity of CFRC. Since short fibre length may lead to Cannebinus (64.3 J). According to the authors, the higher impact
insufficient embedded length for fibre bridge effect, a proper coir fibre resistance of coir fibre-cement mortar was mainly due to the higher
length should be identified in the future studies for practical application. ductility of coir fibre than other plant-based fibres, i.e., the plastic
In addition to compressive and flexural strength, other short-term deformation under tensile discussed in section 2.2.2. Similar results
mechanical properties of CFRC such as fracture toughness, splitting could be found in the study of Hwang et al. (2016), in which coir fibre
tensile strength, dynamic properties and impact resistance were also with 0%, 1%, 2.5%, and 4% of the mortar volume was added in the
investigated in the previous studies. cementitious matrix. The cementitious matrix was modified with fly ash
Fracture toughness is the energy absorption capacity of a beam under and ground blast furnace slag replacing 15% of sand and 25% of cement
bending, i.e., the total area under load-displacement curve of the beam. by weight, respectively. The coir fibre reinforced cementitious matrix
Ali et al. (2012) used toughness index to describe the CFRC toughness, was casted as thin slabs with a dimension of 40 x 30 × 0.12 cm and
which was the ratio of total energy absorption to the area under tested under falling ball impact test. The failing height was set as 180 cm
load-displacement curve until generation of fibre crack. Compared with and the ball was made of steel with a weight of 530g. After the impact
the samples without fibre, the samples with 1%, 2% and 3% coir fibre to test, the samples without coir fibre were broken into pieces while the
cement by weight had a higher toughness index, i.e. around 357% 448% samples with coir fibre were intact. The samples with 1% coir fibre had a
and 496% of the unreinforced samples in average, respectively. Addi­ recessed diameter of 0.9 cm with small cracks. In the contrary, samples
tionally, Ali et al. (2012) found out that CFRC length of 50 mm had with 2.5% and 4% coir fibre showed no cracks and the recessed diameter
highest toughness index to counterparts with 25 mm and 75 mm coir was 1 cm. However, the impact resistance did not always increase as the
fibre. Explanation was given by the authors that 25 mm coir fibre had coir fibre in cementitious matrix was longer. Wang et al. (Wang and
insufficient embedded length for fibre bridge effect. The 75 mm fibre Chouw, 2017a) cast coir fibre reinforced concrete with fibre length of
reduced the fibre number in the matrix leading to the reduction of 25 mm, 50 mm, and 75 mm. The mixing design of cement: water: sand:
toughness index. Maia Pederneiras et al. (Maia Pederneiras et al., 2021) gravel was 1:0.52:2:2 and the coir fibre content was 1.5% of cement
investigated the fracture toughness of CFRC with low coir fibre content, weight. During the test, coir fibre reinforced concrete with 25 mm and
i.e., coir fibre to cement ratio in weight was up to 1.1%. The increase of 50 mm survived with four blows of the drop weight, i.e., impact from
fracture toughness could be up to 102%. Exception could be found for mass of 40 kg once at each 50 cm (blow 1), 100 cm (blow 2), 150 cm
the sample with 0.16% coir fibre. A fracture toughness of 192 N mm was (blow 3), and 200 cm (blow 4) impact height. However, concrete with
measured by the sample, which was slightly lower than that of the fibre length of 75 mm survived with only three blows. The reason was
sample without fibre addition (i.e., 195 N mm). The explanation was not not provided by the authors. From our authors’ perspectives, longer
provided by the authors. From our authors’ perspective, this could be fibre increased the possibility of formation of fibre clumps, resulting in
due to the low coir fibre content that insignificant improvement on the inhomogeneous distribution of coir fibre in the concrete.
fracture toughness could be discovered. The improvement of coir fibre in Splitting tensile and dynamic performance of coir fibre reinforced
fracture toughness was also found in the research of Andiç-Çakir et al. concrete were also investigated by Ali et al. (2012). After the splitting
(2014) that the fracture toughness increase was up to 62% as up to 3.4% tensile test, plain concrete samples were broken into two half pieces
of untreated coir fibre was added. Additionally, pre-treatment (5% while concrete with coir fibre was still held by fibres. The splitting

17
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 18. Impact energy absorption of coir fibre reinforced cement mortar with various plant-based fibres (i.e., Hibiscus Cannebinus, jute, sisal, and coir fibre), fibre
lengths (i.e., 20 mm, 30 mm, and 40 mm), and fibre contents (i.e., 0.5%, 1%, 1.5%, and 2% by cement weight) according to Ramakrishna et al. (Ramakrishna and
Sundararajan, 2005a).

tensile strength increased generally as fibre content increased up to coir indicated that the coir fibre reinforced soil-cement composite had good
content of 5%. Additionally, middle length (i.e., 50 mm) of coir fibre in erosion resistance and could be applicable for the external wall
most of the cases could achieve higher splitting tensile strength than the application.
corresponding samples with small (i.e., 25 mm) and large (i.e., 75 mm)
fibre length. Exception could be found in sample with coir fibre of 75 3.3.3. Long-term mechanical properties of coir fibre reinforced cement-
mm length and content of 1%. This was not explained by the authors and based composite materials
hence more investigations should be conducted for the explanation. As one of the plant-based fibres, coir fibre also faces challenges of
Damping ratio and fundamental frequency are two main parameters that degradation in cementitious matrix like all the others. Generally, the
present the dynamic performance of a structure. Damping ratio is a degradation of plant-based fibres in cementitious matrix is due to the
measure (a nonnegative value) to describe the ease of decay of oscilla­ presence of water (e.g., pore solution) and is classified into two cate­
tions to the rest or equilibrium position. No decay takes place in a system gories, i.e. alkaline degradation and mineralisation. In the case of
if its damping ratio is 0. Fundamental frequency is the lowest frequency alkaline degradation, lignin and hemicellulose undergo the alkaline
of oscillation. In the fibre reinforced cementitious matrix beams, the hydrolysis easily since they are amorphous. According to Wei (2014),
main damping mechanism was the friction when cracks slides over the lignin and part of hemicellulose are firstly degraded, following by the
fibre as high shear stress exists (Chi et al., 2019). Ali et al. (2012) applied degradation of the rest of hemicellulose. As a result, the stability and
dynamic loading on coir fibre reinforced concrete beams and four integrity of fibre cell wall will be decreased and cellulose will be exposed
damage stages (i.e., without cracking, before crack initiation, cracking, in the pore solution. The amorphous regions in the cellulose will be then
and 2–3 cycles of static loading after crack initiation) were considered. hydrolysed due to alkaline pore solution. This degradation process is
During the testing, the damping ratio increased, and fundamental fre­ diagrammatically illustrated in Fig. 19. Additionally, the existence of
quency decreased along with the four stages. The fibre content and fibre calcium ion in the pore solution also leads to the degradation of cellu­
length also influenced the dynamic performance of the concrete but only lose. Based on the investigation from Van Loon et al. (van Loon and
after cracking initiation. At the third cycle of static loading after Glaus, 1998), calcium ion could catalyse the benzylic acid rearrange­
cracking, the damping ratio of samples with 3% coir fibre lay over 0.15 ment to generate isosaccharinic acid (ISA). With the presence of ISA, the
while that of 2% was less than 0.1. Samples with 50 mm coir fibre calcium ion can have a complexation reaction with one carboxylic and
usually exhibited the highest damping ratio and lowest fundamental one hydroxylic group closest to the carboxylic group with the pH value
frequency to other fibre length (i.e., 25 mm and 75 mm). All these results of the pore solution over 10. This leads to a concentration reduction of
indicated that coir fibre could change the material damping. However, free calcium ion and further dissolution of solid calcium hydroxide to
due to the low amplitude (maximum 4–5 kN during testing), the dy­ rebuild the calcium ion solubility equilibrium. The corresponding re­
namic testing could be different from seismic condition. Therefore, a action is as follows:
suggestion was provided by the authors that the performance of coir
Ca2+ + H4 ISA− ↔ CaH3 ISA + H + (12)
fibre reinforced concrete under earthquake loading should be further
investigated. Within the alkaline degradation, the final degradation of cellulose
Soil-cement, a composite consisting of cement and soil was investi­ leads to the reduction of tensile strength as cellulose is the main load
gated as an environmentally friendly material. This composite was carrying component in the plant-based fibres. With the degradation of
widely used as reinforcement of soft soil foundation, slope support, as lignin and hemicellulose, the hydration products (e.g., soluble calcium
well as channel lining for roads and bridges (Duan and Zhang, 2019). To silicate hydrate or calcium hydroxide) can be infiltrated into the cell
improve the stabilisation of soil-cement mortar, coir fibre and lime were wall, causing the mineralisation of plant-based fibres. The crystalliza­
added in the composite by Danso et al. (Danso and Manu, 2020). The tion process (e.g., the nucleation and growth process) of the hydration
results showed that the samples with coir fibre of 0.2% and lime of 5% products destroys the bond between different components and corrodes
by total weight were found to have largest compressive strength, tensile cellulose micro-fibrils. As a result, fibre becomes brittle and its
strength, and density to other samples with other coir fibre content (i.e., deformability significantly reduces, in turn the fibre tensile strength
0.4%, 0.6%, and 0.8%) and lime ratio (i.e., 0%, 10%, and 15%). In decreases. Based on the aforementioned degradation mechanisms, lignin
addition, no pit or indent was found on the surface of the samples after and hemicellulose are considered as protection of cellulose and help to
erosion test (i.e., 100 ml water dripping with a height of 400 mm). This delay the decomposition of plant-based fibres. As discussed in section

18
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Fig. 19. Degradation process of plant-based fibre adapted from Wei et al. (Wei and Meyer, 2015). Permission granted.

2.1, coir fibre has lower cellulose content but higher lignin content than flexural strength than that with coir fibre. This was mainly due to the
other conventional plant-based fibres such as flax, hemp, sisal, and jute. higher tensile strength of sisal fibre than that of coir fibre used in the
This makes coir fibre more suitable as monofilament fibre in the study. The average tensile strength of sisal fibre was 578 MPa while that
cementitious matrix as it is more stable than the other plant-based fibres. of coir fibre was only 174 MPa. However, the reduction of the post-crack
Tolêdo Filho et al. (Tolêdo Filho et al., 2000) compared sisal and coir flexural strength of beam with sisal fibre exhibited significantly larger
fibres on the reduction of tensile strength in calcium hydroxide solution. than that with coir fibre. This again proved that coir fibre was more
It was found that at 210th day, only 33.7% of the original strength stable in cementitious matrix due to its higher content of lignin than
retained by sisal fibre while that of coir fibre was 58.7%. Ramakrishna other plant-based fibres.
et al. (Ramakrishna and Sundararajan, 2005b) conducted similar in­ John et al. (2005) investigated the coir fibre obtained from a
vestigations but for four types of plant-based fibres, i.e. coir, sisal, jute, 12-year-old CFRC wall. Instead of cement, the low alkaline binder was
and Hibiscus Cannabinus. Around 20%–40% of the original strength used for the wall which mainly consisted of blast-furnace slag activated
could be retained at 60 days by coir fibre, which was higher than jute by 2% of lime and 10% of gypsum. The time period was from year
fibre with 10%–20% of the original strength in saturated calcium hy­ 1961–1990 with an average temperature range of 11.7 ◦ C–27.9 ◦ C,
droxide solution. Sisal and Hibiscus Cannabinus fibres were completely average rainfall of 1477 mm/year, and average relative humidity range
destroyed after 60 days. of 74%–80%. From the scanning electron microscope, it was found that
Throughout our literature review on the research in the recent 20 calcium and small amount of silicon were impregnated in the fibre cell
years, limited investigations can be found on the long-term mechanical wall. A fully impregnation was hardly found in the lumens of the
performance of CFRC. This is critical for the coir-fibre reinforced elementary fibres. Additionally, the obtained fibres had an average
cement-based composites applications in structures. Tolêdo Filho et al. lignin content of around 35%, while that of new fibres was reported to
(Tolêdo Filho et al., 2000) compared sisal and coir fibre on the long-term be 48%. Therefore, a conclusion was drawn by the authors that lignin
performance of fibre reinforced cementitious mortar beams. The con­ was still decomposed even low alkaline binder was used. Despite of that,
ditions simulated the effects of ageing were set to be in water, exterior the decomposition of lignin showed no effect on the use performance of
environment (outdoors), and with wet/dry cycles. The samples at 28 the wall.
days were considered as control group. For samples in water and out­ In the consideration of the CFRC application for marine structure,
doors, the exposure time of 180 and 322 days were considered. For Ramli et al. (Maia Pederneiras et al., 2021) investigated the CFRC in
wet/dry cycles, 25 and 46 cycles were used. The post-crack flexural outdoor condition (i.e., tropical climate with radiation and raining),
strength under the corresponding condition of the beams with sisal and combined outdoor-seawater condition (i.e., 4 days in seawater and 10
coir fibres were compared with the control group, shown in Fig. 20. In days outdoor), and seawater condition. In the mixture, 12% by weight of
most cases, the samples with sisal fibre retained higher post-crack type I Portland cement was replaced by condensed silica fume. The

Fig. 20. The relationship of coir and sisal fibre reinforced cement-based composite flexural strength in percentage with different (a) ageing time in water, (b) ageing
time for outdoors, and (c) numbers of wet/dry cycles from Tolêdo Filho et al. (Tolêdo Filho et al., 2000) The 28-days flexural strength of cementitious matrix with
corresponding fibres (i.e., sisal and coir fibre) is considered as the benchmark.

19
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

samples with coir fibre contents of 0%, 0.6%, 1.2%, 1.8%, and 2.4% by compressive strength was stronger when the samples were longer
binder volume were studied under compression and bending with exposed. For example, with 1.2% coir fibre, the compressive strength
exposure period of 3, 6, 12, and 18 months. In the outdoor condition, increase was around 2% after 28 days, while that was around 11% and
both compressive strength and flexural strength of all the samples 6% after 18 months exposure under outdoor-seawater and seawater
increased consistently along the exposure time. This was explained by condition, respectively (Fig. 21 (a) and (b)). However, large coir fibre
the further hydration of cement and pozzolanic reaction (i.e., silica from content (e.g., 2.4%) had negative effect on the compressive strength
silica fume reacting with calcium hydroxide) with the presence of water after 18 months exposure. Similar to the compressive strength, the coir
from tropical climate condition. The effects of exposure time and fibre fibre content of 0.6% and 1.2% improved the long-term flexural strength
content on compressive and flexural strength increase under of the cementitious matrix. The flexural strength of the samples with
outdoor-seawater and seawater are presented in Fig. 21. The corre­ 1.2% coir fibre were improved by around 2% and 9% under
sponding strength of the samples without coir fibre in the corresponding outdoor-seawater and seawater condition. Even high coir content (e.g.,
month is considered as the benchmark (i.e., Surface 0). The blue surface 2.4%) improved the flexural strength (i.e., around 15%) before expo­
(i.e., Data) indicates the increase of the strength to the benchmark at the sure, the improvement turned to decline after 18-month exposure. All
same month. Under both outdoor-seawater and seawater condition, the these results can be explained by the mechanism of fibre in cementitious
0.6% and 1.2% coir fibre in most cases improved the compressive matrix. The fibre bridging effect limited the crack propagation,
strength of the cementitious matrix. The improvement effect on impeding the penetration of the seawater. High fibre content, however,

Fig. 21. Compressive and flexural strength increase under outdoor-seawater and seawater condition with various exposure periods (i.e., 3, 6, 12, and 18 months) and
coir fibre content (0%, 0.6%, 1.2%, and 2.4%) (Ramli et al., 2013). The fibre content is the percentage by binder volume. The blue surface (Data) is the increase of
tested value to the corresponding benchmark in percentage and the pink surface (Surface 0) is the flat surface where strength increase is equal to zero.

20
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

reduced the workability of the matrix, increasing the porosity of the time points in Gutiérrez et al. (2005) and Khan et al. (Khan and Ali,
CFRC. Therefore, more seawater was able to penetrate into the samples 2019)) to prove whether pozzolana can retard the reduction of me­
through pores, reducing the strength of the CFRC. It is worth noticing chanical properties of CFRC along time. Therefore, further in­
that all the data shown in Fig. 21 are average values of the test results vestigations are recommended to prove and quantitatively analyse the
from Ramli et al. (2013). Since most of the increases were within ±5%, retardation effect of pozzolana on coir degradation in cementitious
they could be statistically insignificant. The above discussions are all matrix.
based on the average values and readers should be cautious with them. Accelerated carbonation is an alternative matrix modification
Additionally, the degradation mechanisms of coir with seawater should method for plant-based fibre reinforced cement-based composites. The
be further investigated to validate the results from this study. existing calcium hydroxide in the matrix reacts with carbon dioxide to
form calcium carbonate (Pacheco-Torgal and Jalali, 2011; Yan et al.,
3.4. Modification methods for coir in cementitious materials 2016a; Ardanuy et al., 2015). Tolêdo Filho et al. (Tolêdo Filho et al.,
2003) compared the effect of accelerated carbonation on the long-term
The modification methods for CFRC focused on the improvement of bending performance of fibre reinforced cement-based composites.
interfacial bonding between fibre and matrix as well as the retardation Hybrid fibre reinforcement was used in the experiments, i.e. 2 vol%
of fibre degradation. These methods can be classified as two categories, randomly distributed coir fibre with 25 mm and 1 vol% aligned long
i.e. matrix modification and fibre modification. Additionally, the overall sisal fibre with 375 mm. The accelerated carbonation was conducted by
mechanical performance of CFRC can be also improved through indirect CO2 incubator with 9.8% of CO2 for 109 days. Water immersing, out­
methods such as using external confinement. door, and wet/dry cycle were selected as the aging condition in the
study. Generally, the aged and carbonated samples had lower
3.4.1. Matrix modification post-cracking flexural strength than that of the carbonated samples
The matrix modification mainly targets on the retardation of fibre without ageing. However, the post-cracking flexural strength for out­
degradation through reduction of pH value in the cementitious matrix. door condition of 250 days and wet/dry cycle of 36 was higher than that
Using pozzolana (e.g., husk ash, blast furnace slag, silica fume, or fly for outdoor condition of 180 days and wet/dry cycle of 25. Similar
ash) to replacing Portland cement is the most commonly modification tendency was found also in the Japanese toughness index, i.e., the
methods for plant-based fibre reinforced cement-based composites required energy of a beam to have a mid-span deflection of 1/150 of the
(Pacheco-Torgal and Jalali, 2011; Tolêdo Filho et al., 2003; Yan et al., span. The authors suggested to have further research for the confirma­
2016a; Agopyan et al., 2005; Gutiérrez et al., 2005; Ardanuy et al., tion of the obtained data.
2015). Pozzolana in the cementitious matrix can have the pozzolanic
reaction with calcium hydroxide and silica in the matrix, producing the 3.4.2. Fibre modification
calcium silicate hydrates or calcium aluminate hydrates. The final The fibre modification focuses on the improvement of fibre-matrix
products are the main components that provide load carrying capacity to interfacial bonding or fibre coating against fibre degradation. There
the matrix. With the reaction of calcium hydrates, the pH value of the were various of modification methods on plant-based fibre in cementi­
pore solution in the matrix can be reduced. Agopyan et al. (2005) used tious matrix (Ahmad et al., 2019). Alkali treatment was found to have
granulated blast furnace slag as binder of roofing tiles (with 10% gyp­ improvement on the mechanical performance of plant-based fibre
sum and 2%–4% lime). The proportion of coir fibres was 3.7% to the reinforced cement-based composites. At 90 days of curing, the
binder in weight and the water binder ratio was 0.45. After exposure in compressive and flexural strength of alkali-treated jute fibre reinforced
outdoor condition for 16 and 60 months, the maximum bending load cement-based composites was around 5% and 41% higher than that of
were 48.7% and 51.5% of that with one month exposure respectively. untreated samples (Jo et al., 2016). As reported by Sedan et al. (2008),
Despite that the maximum bending load reduced significantly from the flexural strength of hemp fibre reinforced cement-based composites
month 1 to month 16, a slight improvement in the load could be found was improved by 55% after hemp was treated with NaOH. In the case of
from month 16 to month 60. Even if no statistical analysis on signifi­ coir fibre, alkali treatment can also improve mechanical performance of
cance was conducted for month 16 and month 60, it was obvious that the CFRC. This has been presented in Fig. 17 from section 3.3.2 (the research
weakening of the bending load carrying capacity slowed down along of Andiç-Çakir et al. (2014)). After 2 h NaOH treatment on coir fibre, the
time. This could be also proved by the reduction of the specific energy (i. fracture toughness of CFRC was improved by 8%–16%. Yan et al.
e., the area under the load-displacement curve) of the samples. The (2016a) treated coir fibre with 5 wt% NaOH for 30 min. Their results
specific energy of the samples with 16- and 60-months outdoor exposure showed that the average compressive strength, compressive strain,
was around 35.6% and 28.7% of that with one-month outdoor exposure, flexural strength, and fracture energy of the treated coir-concrete were
respectively. Gutiérrez et al. (2005) compared the performance of CFRC 1%, 17%, 6%, and 5% higher than the counterpart without treatment. As
with four different pozzolana types, i.e. silica fume, metakaolin, fly ash, reported by Siakeng et al. (2018), alkali treatment removed the surface
and ground granulated blast furnace (GGBS). The pozzolana content was lignin and impurities such as wax, resulting in a rougher and cleaner
15% of the binder in weight except for GGBS with 70%. Coir fibre surface than the untreated fibre. This improved the fibre-matrix inter­
content was 2.5% of the cement weight. Compared with the control facial bond. Wetting agent was reported to have improvement on the
group (cementitious matrix without fibres), the compressive strength of fibre-matrix bonding as proved in coir mesh in the cementitious matrix
CFRC was 92%, 91%, 95%, 102%, and 85% of the control group for the by Li et al. (2007a). The wetting agent used in the study was 2-ethylhex­
matrix with no pozzolana, silica fume, metakaolin, fly ash, and GGBS at anol. It was found that the overall flexural performance (including
90 days, respectively. This result showed that metakaolin and fly ash flexural strength and toughness) had an improvement of around 20% of
may retard the coir fibre degradation. However, the number of repli­ the treated mesh to the untreated counterpart. Gaps were found in the
cation and standard deviation of the compressive strength were not interface between untreated fibre mesh and matrix, indicating that the
given in this study. The compressive strength of CFRC with different wetting agent could promote the interfacial bonding (Yan et al., 2016a).
pozzolana may have no statistically significant difference. Khan et al. Fibre coating with epoxy paint and varnish were introduced by Het­
(Khan and Ali, 2019) investigated CFRC with fly ash and silica fume. tiarachchi et al. (Hettiarachchi and Thamarajah, 2020). The purpose of
With the addition of both fly ash and silica fume, the improvement of using coating was to reduce the moisture absorption of coir fibre. With
compressive strength, flexural strength, and splitting tensile strength 1% coir fibre by cement weight, around 4.6% and 4.4% less moisture
were 5%–35%, 1%–30%, 9%–27%, respectively. However, all the was absorbed by the matrix with coir fibre coated with epoxy and var­
aforementioned studies provided only limited data (i.e., lack of com­ nish than that without coating, respectively. They were larger than that
parison with unmodified matrix in Agopyan et al. (2005) and different with NaOH modification (i.e., 3.4% less moisture absorption of

21
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

untreated CFRC). Comparing with NaOH modification and varnish average maximum deflection of ten samples by 683% and 1300% for
coating, the epoxy coating provided the greatest improvement on 2-layer and 4-layer, respectively (Yan and Chouw, 2013a). The damping
compressive, and splitting tensile strength of CFRC, i.e., 3.4% and 8.9% ratio obtained from the dynamic testing for coir-concrete was 380%–
of that of untreated CFRC, respectively. Epoxy coating provided com­ 462% of that of plain concrete. With 2-layer and 4-layer flax FRP (FFRP),
parable improvement on flexural strength to the NaOH modification the damping ratio was further increased by 34%–68% and 76%–83%,
that the improvements were 6.5% and 5.3% by NaOH modification and respectively (Yan and Chouw, 2014b). Additionally, Wang et al. (Wang
epoxy coating to the untreated CFRC, respectively. and Chouw, 2017b) investigated the impact resistance of CFRC rein­
forced with FFRP. In the study, coir-concrete beams were externally
3.4.3. Other modification strengthened with 2-, 4-, and 6-layer FFRP, similar to the configuration
Despite CFRC has advantage in sustainability, its mechanical per­ in Fig. 22 (b). The impact was conducted through a drop of 40 kg to the
formance is usually inferior in comparison with cementitious matrix mid-span of the beams. The impact height started from 200 mm. For
with other fibres such as steel fibre. The steel fibre reinforced cement- each repeat of impact, an increment of 100 mm for the impact height
based composite has been widely used as structural elements (e.g., was considered. This means that at second repeat, the impact height was
prestressed concrete bridge beams (Brandt, 2008)), while CFRC was 300 mm while it was 400 mm at the third repeat. The samples with 2-,
only found as light-weight façade board or coating layer for masonry 4-, and 6-layer FFRP survived 2, 4, and 5 repeats before failure,
wall (details will be presented in section 3.5). In order to broaden the respectively. The energy absorption increased as more layers were
application of CFRC, especially the structural application, comprehen­ applied on the coir-concrete. The average total energy absorption of
sive investigations on coir fibre reinforced concrete externally con­ eight samples with 6-layer was 140.9J, which was more than 700% of
fined/strengthened with fibre reinforced polymer (FRP) composites that of the coir-concrete without FRP strengthening (i.e., 19.4 J). All
have been conducted by the second author of this study. The configu­ these investigations showed that flax-FRP could significantly improve
rations of FRP for coir-concrete column and beam are illustrated in both static and dynamic mechanic performance of coir fibre reinforced
Fig. 22 (a) and (b), respectively. The FRP could partially or fully replace concrete. It can be therefore concluded that the using environmentally
the steel rebar in the concrete and provide tensile carrying capacity and friendly plant-based fibre reinforced polymer composite (e.g., flax-FRP)
shear resistance. The coir fibre was considered in the concrete to prevent can increase the possibility of using coir fibre reinforced concrete in
the brittle failure of concrete (Yan and Chouw, 2013a). Additionally, the structural applications.
FRP tube shown in Fig. 22 (a) could be also used as concrete mould,
realizing mould-free concrete casting in the practical application. 3.5. Practical application of coir fibres and coir fibre reinforced cement-
Since flax fibres have comparable tensile properties to those of glass based composite materials
fibres, they are considered as reinforcement in the FRP. The flax-FRP
tube encased coir-concrete with/without steel rebar under compres­ The practical application of coir fibres can be ascribed to their light
sion (Yan and Chouw, 1980-2015, 2013a, 2014a; Yan et al., 2014b), weight and excellent insulation ability. It can be fabricated into ply
bending (Yan and Chouw, 1980-2015, 2013a, 2013b), and dynamic board for interiors and furniture (TNKKSS, 2021; Kenstar. Coir Ply
loading (Yan and Chouw, 2014b), as well as the flax-FRP strengthened Board., 2021), or acoustic and thermal insulation board for roof, floor,
coir-concrete beam under bending (Yan et al., 2015) were experimen­ wall, and façade (Coir Insulation B, 2021; Enkev, 2021; Biodegradable
tally and theoretically investigated. For example, with 2-layer and and Compost, 2021). Coir fibre can be also used as filler in masonry
4-layer flax-FRP confinement the axial compressive strength of walls to promote the thermal insulation performance of the walls. Iwaro
coir-concrete was improved by 38% and 99%, respectively. Addition­ et al. (Iwaro and Mwasha, 2019) filled hollow concrete block (150 x 200
ally, the average axial strain of 2-layer and 4-layer confined × 400 mm) and clay masonry block (100 x 200 × 300 mm) with coir
coir-concrete was respectively 1.89% and 2.70%. The average axial fibre as illustrated in Fig. 23. The insulation performance of coir fibre
strain of coir-concrete was only 0.54%, although it was higher than that filled block as masonry walls was investigated by establishing small
of plain concrete (i.e., 0.20%) (Yan and Chouw, 2013a). The average demonstrations (around 1.5 m long, 1.2 m wide, and 1.5 m high with
peak load of cylindrical coir-concrete under three-point bending test was 26G corrugated, galvanized sheeting as roof). It was found that in the
10.1 N, which was higher than that of plain concrete (i.e., 7.4 N). hot climate environment, coir fibre filled block had good thermal
Confined with 2-layer and 4-layer flax-FRP, the increase of 267% and insulation performance and could reduce the energy consumption of air
946% in peak load was detected. The flax-FRP tube also improved the conditioning. For example, without air conditioning, the average indoor
temperature was 27.8 ◦ C for coir fibre filled concrete block while that of
uninsulated wall was 28.5 ◦ C. Almost no difference was detected when
glass fibre was used as filler in the block. With air conditioning, the

Fig. 22. Configuration of fibre reinforced polymer (FRP) composite as external


(a) confinement for coir-concrete column and (b) reinforcement of coir- Fig. 23. Illustration of coir fibre filled hollow concrete or clay masonry block
concrete beam. based on Iwaro et al. (Iwaro and Mwasha, 2019).

22
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

indoor temperature could be kept constant (i.e., 24 ◦ C in the study). The distributions on coir fibre strength prediction has not been verified
energy consumption for coir fibre filled concrete block was found to be and should be further investigated (Section 2.2.3).
82% and 97% of the uninsulated and glass fibre filled concrete block, • The regression-based predictions on bond strength and total energy
respectively. (Eq. (11) and Eq. (12) in section 3.2.2) were generated by Ali et al.
Compared with a direct application of coir fibre, the CFRC was rarely (2013) based on their experimental results. However, these pre­
practically applied as construction elements. A coir fibre cement board dictions have not been validated with other coir fibre pull-out test
was developed by the PCA-Zamboanga Research Center in Philippines results, which should be conducted in the future.
(Ali, 2011; Business DiaryCoconu, 2021). This coir fibre cement board • Most of the studies focused on the thermal insulation ability of CFRC.
(CFB) consisted of 30%–40% coir fibre and 60%–70% cement by weight. Since coir fibre mat is good at sound absorbing, CFRC is expected to
The CFB had a low thermal conductivity of 0.9 W/mK and performed have good acoustic insulation ability. Comprehensive investigations
well in flame test with slow burning rate and low smoke emission. In the are recommended to broaden the practical application of CFRC as
early 2000, a two-story house was built by this research center with CFB. acoustic insulation (Section 3.3.1).
The CFB was used as walls, partitions, and ceiling in the house. More­ • Using pozzolana (e.g., fly ash or silica fume) to modify the cemen­
over, CFRC was also reported to be applicable as façade coating layer of titious matrix can improve the compressive, flexural, and splitting
housing. Quiñones-Bolaños et al. (2021) have investigated a house (42 tensile strength of CFRC. The pozzolana is also reported to be able to
m2 of constructed area) in Cartagena de Indias, Colombia (ambient retard the loss of these mechanical properties. However, limited data
temperature of 27 ◦ C) for 39 days, which was coated with a 3 cm-layer of (i.e., lack of comparison with matrix without pozzolana or lack of
coir fibre reinforced mortar. In this mortar, the ratio of cement, water, comparison at same time points) was provided by the previous in­
sand was 1:0.5:3. With an addition of 15% coir fibre by weight, the vestigations (Khan and Ali, 2019; Agopyan et al., 2005; Gutiérrez
indoor temperature reduced by 0.5 ◦ C–1.5 ◦ C during the maximum et al., 2005). To promote the application of pozzolana in CFRC for
sunlight radiation period and the energy costs for cooling could reduce better long-term performance, the retardation effect of the pozzolana
by 16% through calculation. Based on the results, it was pointed out by should be quantitatively analysed and proved.
the authors that CFRC as façade coating had its potential in tropical
areas for the low-income housing structures. The following points should be investigated:

4. Future perspectives • The coir fibre in cementitious matrix is in most cases randomly ori­
ented. When the fibre bridge effect takes place due to the crack
Adding the agricultural waste, i.e., coir fibre, in cementitious matrix initiation, the surface of the crack cannot always be perpendicular to
can reduce the environmental impact of concrete. The aforementioned the fibre direction. With the crack propagation, the soft coir fibre can
studies showed that coir fibre could improve the mechanical perfor­ be readily bent and later cut off at the bending point by the crack
mance (e.g., compressive, flexural, and tensile), dynamic and impact surface. The failure pattern of the coir fibre with embedded angle
performance, and physical properties (e.g., reduction of weight, thermal could be different from steel fibre as steel fibre still has bending
and acoustic insulation ability) of cementitious material. However, the resistance. Therefore, the orientation of coir fibre in cementitious
investigation on coir in cementitious matrix is still at the early stage and matrix may significantly affect the mechanical performance of the
a number of questions still need to be considered and answered. Several CFRC. In this case, the investigations on for example the embedded
future perspectives or recommendations that have been discussed above angle of coir fibre in cementitious matrix on fibre pull-out perfor­
are provided below: mance should be conducted.
• Alkaline treatment is one of the most common treatment of coir fibre
• Considering slippage and test setup compliance, a calibration as well as other plant-based fibres. Previous results showed that
method on coir tensile test was reported by Defoirdt et al. (2010) (Eq. using alkaline (e.g., NaOH solution) could improve the mechanical
(3) and Eq. (4) in section 2.2.1). In the calibration, the displacement properties (e.g., compressive strength, strain, flexural strength, or
by slippage and test setup compliance should be removed from the fracture energy). However, alkaline treatment dissolves the lignin in
measured displacement from the testing machine. However, since the coir fibre, which is the main component protecting coir fibre
the factor α and extrapolated fibre elastic modulus are determined by from degradation in the cementitious environment. The long-term
linear regression, the validation of the calibration should be con­ performance of cementitious mortar reinforced with alkaline
ducted with more experimental results. The validation could be treated coir fibre should be investigated. The result from the inves­
achieved by comparing the results from calibration method and the tigation may also contribute to the optimization of alkaline treat­
tensile test with optical strain mapping system. ment of coir in cementitious matrix. Additionally, the effect of low
• The stress-strain curve of coir fibre in tension is bilinear. This is alkaline cement/binder on the coir fibre degradation can be also
different from other plant-based fibres such as flax or sisal, which studied. All these studies will promote the design of CFRC in the
exhibit always linear elastic behaviour. The fibre morphology, cell practical application.
wall structure, and the high lignin content are expected to be the • Besides alkaline treatment, there are many other fibre modification
main aspects contributing to the bilinear stress-strain curve in ten­ methods that have been reported for other plant-based fibres, e.g.
sion of coir. This expectation should be further verified by using hornification. Hornification method is to use dry and wet cycles on
micro or nano computed topography during the tensile test of coir the plant-based fibres to promote their dimensional stability and
fibre with or without delignification treatment. (Section 2.2.2). reduce the water retention values (Yan et al., 2016b). For example,
• Existing literature showed that the tensile strength of plant-based 40% and 50% improvement could be found on the adhesion and
fibres can be predicted by Weibull analysis. Based on Defoirdt frictional bond strength of sisal fibre after hornification (Ferreira
et al. (2010), coir fibre was not justified by using the Weibull dis­ et al., 2014). As one type of plant-based fibres, coir fibre could be
tribution for the prediction of its tensile strength, since coir has also effectively modified by the methods to improve the interfacial
noticeable plastic deformation under tension. Additionally, the mean bonding between fibre and matrix. Further investigations are
value and standard deviation for normal and Weibull distribution of required for this verification.
coir fibre tensile strength were found almost equal. Therefore, • One of the objectives of using coir fibre in cementitious matrix is that
normal distribution could be sufficient to predict the tensile strength the fibre can change the brittle failure of cementitious matrix into
of coir that the coir fibre strength is expected to obey normal dis­ ductile. The brittle failure is critical in the case of collapse of a
tribution. However, the applicability of normal distribution or other structure especially during earthquake. Despite dynamic loading was

23
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

conducted on the CFRC (Ali et al., 2012; Yan and Chouw, 2014b), it steel fibre, glass fibre, and carbon fibre) to produce monofilament fibre
is still different from seismic condition. As also suggested by Ali et al. reinforced cement-based composite. Under tension, coir fibre exhibits
(2012), the seismic performance of CFRC should be further bilinear stress-strain respond. Previous studies showed that coir fibre
investigated. had lower tensile strength, tensile modulus, but higher elongation at
• Previous studies showed that polypropylene fibre in concrete could failure in comparison with other plant-based fibres such as flax, sisal,
promote the fire resistance of concrete (Brandt, 2008). As tempera­ hemp, and jute. The fibre-matrix interfacial bonding can be detected by
ture increases, polypropylene fibre can be melted by absorbing the fibre pull-out test. Existing investigations found out that the bond
heat and simultaneously creating free channels to take the thermal strength of coir in the cementitious matrix decreased along with the
expansion of the mortar. As plant-based fibre, coir fibre is inflam­ curing days of matrix. However, the bond strength was not propor­
mable in the open air. However, in CFRC, coir is covered by tionally related to other parameters such as water cement ratio, mixing
cementitious matrix and has rarely contact with the oxygen in the design (i.e., cement aggregate ratio), or embedded length.
air. Instead of being ignited, carbonization may occur in the coir CFRC can be regarded as a lightweight construction material with
fibre. During the carbonization process, the hydrogen and oxygen well thermal and acoustic insulation ability. Its mechanical properties
atom in the fibre can be released from organic components of the such as compressive, flexural, splitting tensile, and dynamic perfor­
fibre as water and simultaneously absorb the heat. Additionally, the mance as well as fracture toughness under bending have been discussed
carbonized coir fibre becomes thinner and in turn creates the free in this study. It was found that the mechanical performance of CFRC
room for mortar thermal expansion. Hence, it is expected that coir could be significantly affected by the fibre content. With more coir fibre
fibre may improve the fire resistance of cementitious matrix. This in the cement-based matrix, the workability of the matrix reduces. This
should be further validated in the future experiments. results in the increase of matrix porosity and in turn weakens the load
• Geopolymer is considered as an environmentally friendly sub­ carrying capacity of the CFRC. This study has also reviewed the in­
stitutions to the ordinary Portland cement (OPC) as it consists of a vestigations on the long-term performance of CFRC. It was reported
large content of industrial by-products and waste products (e.g., fly from the previous investigations that CFRC could retain higher per­
ash and blast-furnace slag) with less CO2 emission than OPC (Yan centage of the original mechanical properties than that with other plant-
et al., 2016b; Camargo et al., 2020). Existing studies showed that coir based fibres, e.g., sisal, jute, Hibiscus Cannabinus. The current modifi­
fibre could significantly improve the mechanical performance of cation methods focused on improving the coir fibre-matrix interfacial
geopolymer. Wongsa et al. (2020) compared coir fibre with glass bonding, retarding the coir fibre degradation, or improving the overall
fibre for high calcium fly ash geopolymer flexural strength, and mechanical performance through indirect methods. These methods
splitting tensile strength. Coir fibre provided higher flexural strength could be classified into matrix modification, fibre modification, and
(i.e., 5.3–6.6 MPa) than that with glass fibre (i.e., 3.1–3.7 MPa). The other modification. Matrix modification includes addition of pozzolana
sample with coir fibre showed comparable splitting tensile strength and accelerated carbonation, while fibre modification consists of alkali
(i.e., 2.0–2.2 MPa) to that with glass fibre. These were all greater treatment, wetting agent, and coating. Externally strengthening the
than the corresponding flexural strength (i.e., 3.1 MPa) and splitting CFRC with fibre reinforced polymer composite was considered as other
tensile strength (i.e., 1.9 MPa) of the unreinforced geopolymer. In modification method, which significantly improved the compressive,
addition, due to the pozzolanic reaction through high content of fly flexural, and dynamic performance of CFRC.
ash, blast-furnace slag, or metakaolin, the durability of coir fibre can So far, coir fibre was practically applied into ply board or as filler in
be improved as explained in section 3.4.1. Therefore, coir fibre the masonry walls due to its lightweight and insulation ability. Such
reinforced geopolymer is expected to be one of the ideal environ­ advantages of coir fibre also contributed to the application of coir fibre
mentally friendly substitutions for steel fibre reinforced concrete. cement board. In addition, CFRC was found practically available as
However, based on our literature review, no practical application façade coating to improve the living comfort and reduce the energy
was established for the coir fibre reinforced geopolymer. With the consumption for example in tropical area.
concern of sustainability and the promotion of further application, Several recommendations for the future investigation have been
the next step investigation on coir fibre or other plant-based fibre listed in this study. For example, the existing calibration and prediction
reinforced geopolymer could focus on the full-scale demonstration in models should be further validated. More investigations for influence of
the relevant environment such as hygrothermal environment, or parameters on coir fibre performance in cementitious matrix are sug­
freezing-thawing condition. gested, including mortar consistency on fibre bond strength, long-term
• Statistical methods (e.g., analysis of variance, t-test, F-test, Shapiro- investigations on modification methods, the investigation on seismic
Wilk test, and Kolmogorov-Smirnov test) have been widely applied and fire performance of CFRC, and demonstration investigation of coir
for natural fibre reinforced polymer composites. For example, Le fibre reinforced geopolymer towards practical application.
Moigne et al. (Le Moigne et al., 2011) analysed the length and aspect
ratio distributions of flax, sisal and wheat straw reinforced poly­ Credit roles
propylene with Kolmogorov-Smirnov test. It was found that
two-parameter Weibull distribution had generally good correlation Wang B: Original draft; Formal analysis; Methodology. Yan LB:
to the length and aspect ratio distribution. Based on our literature Funding acquisition; Conceptualization; Methodology; Project admin­
review, large variation can be always found in plant-based materials istration; Supervision; Review & Editing. Kasal B. Supervision; Review &
such as coir fibre material properties (e.g., density, chemical Editing.
composition) and mechanical behaviours (e.g., tensile and pull-out
behaviour). It is therefore recommended that the statistical Declaration of competing interest
methods should be considered in the future investigation on coir
fibre as well as other plant-based fibres as reinforcement in The authors declare that they have no known competing financial
cement-based composites to ensure the reliability of the test results. interests or personal relationships that could have appeared to influence
the work reported in this paper.
5. Conclusions
Acknowledgement
The agricultural waste from the copra industry, coir fibre, has low
embodied carbon footprint with large elongation during tensile. This The research was financially supported by Bundesministerium für
makes coir fibre an ideal fibre for the replacement of other fibres (e.g., Bildung und Forschung (BMBF, Federal Ministry of Education and

24
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Research of Germany) (Grant No.: 031B0914A) and Fachagentur Asasutjarit, C., Hirunlabh, J., Khedari, J., Charoenvai, S., Zeghmati, B., Shin, U.C., 2007.
Development of coconut coir-based lightweight cement board. Construct. Build.
Nachwachsende Rohstoffe e.V. (FNR, Agency for Renewable Resources)
Mater. 21 (2), 277–288. https://doi.org/10.1016/j.conbuildmat.2005.08.028.
founded by Bundesministerium für Ernährung und Landwirtschaft Ashby, M.F., 2013. Material profiles. In: Materials and the Environment. Elsevier,
(BMEL, The Federal Ministry of Food and Agriculture of Germany), pp. 459–595.
under the Grant Award: 22011617. The first author would like to thank Bakatovich, A., Davydenko, N., Gaspar, F., 2018. Thermal insulating plates produced on
the basis of vegetable agricultural waste. Energy Build. 180, 72–82. https://doi.org/
the China Scholarship Council to support his PhD study. The authors 10.1016/j.enbuild.2018.09.032.
would also like to thank Ms. Dagmar Brammertz for her support of the Bayer, J., Granda, L.A., Méndez, J.A., Pèlach, M.A., Vilaseca, F., Mutjé, P., 2017.
TGA measurement on coir fibre. The gratitude would be extended to Cellulose polymer composites (WPC). In: Advanced High Strength Natural Fibre
Composites in Construction. Elsevier, pp. 115–139.
anonymous reviewers for their constructive comments during the re­ Bhutta, A., Farooq, M., Zanotti, C., Banthia, N., 2017. Pull-out behavior of different fibers
view process. in geopolymer mortars: effects of alkaline solution concentration and curing. Mater.
Construcción 50 (1), 467. https://doi.org/10.1617/s11527-016-0889-2.
Biodegradable and compostable pots [25.Apr.2021]; Available from: https://biodegrada
References ble-pots.com/insulation-coconut-fibre/.
Bledzki, A., 1999. Composites reinforced with cellulose based fibres. Prog. Polym. Sci. 24
5079:2021-02, 2021DIN EN ISO 5079:2021-02, Textilfasern_- Bestimmung der (2), 221–274. https://doi.org/10.1016/S0079-6700(98)00018-5.
Höchstzugkraft und Höchstzugkraftdehnung an einzelnen Fasern (ISO_5079:2020); Bledzki, A.K., Mamun, A.A., Volk, J., 2010. Barley husk and coconut shell reinforced
Deutsche Fassung EN_ISO_5079:2020. Berlin: Beuth Verlag GmbH. https://doi.org/10.3 polypropylene composites: the effect of fibre physical, chemical and surface
1030/3181427. properties. Compos. Sci. Technol. 70 (5), 840–846. https://doi.org/10.1016/j.
Abraham, E., Deepa, B., Pothen, L.A., Cintil, J., Thomas, S., John, M.J., et al., 2013. compscitech.2010.01.022.
Environmental friendly method for the extraction of coir fibre and isolation of Blok, R., Smits, J., Gkaidatzis, R., Teuffel, P., 2019. Bio-based composite footbridge:
nanofibre. Carbohydr. Polym. 92 (2), 1477–1483. https://doi.org/10.1016/j. design, production and in situ monitoring. Struct. Eng. Int. 29 (3), 453–465. https://
carbpol.2012.10.056. doi.org/10.1080/10168664.2019.1608137.
Acatay, K., 2017. Carbon fibers. In: Fiber Technology for Fiber-Reinforced Composites. Brandt, A.M., 2008. Fibre reinforced cement-based (FRC) composites after over 40 years
Elsevier, pp. 123–151. of development in building and civil engineering. Compos. Struct. 86 (1–3), 3–9.
Adesanya, D.A., Raheem, A.A., 2010. A study of the permeability and acid attack of corn https://doi.org/10.1016/j.compstruct.2008.03.006.
cob ash blended cements. Construct. Build. Mater. 24 (3), 403–409. https://doi.org/ Bui, T.T.H., 2021. Study on Performance Enhancement of Coconut Fibres Reinforced
10.1016/j.conbuildmat.2009.02.001. Cementitious Composites [PhD Dissertation]. Normandie Université.
Agopyan, V., Savastano, H., John, V.M., Cincotto, M.A., 2005. Developments on Bui, H., Sebaibi, N., Boutouil, M., Levacher, D., 2020. Determination and review of
vegetable fibre–cement based materials in São Paulo, Brazil: an overview. Cement physical and mechanical properties of raw and treated coconut fibers for their
Concr. Compos. 27 (5), 527–536. https://doi.org/10.1016/j. recycling in construction materials. Fibers 8 (6), 37. https://doi.org/10.3390/
cemconcomp.2004.09.004. fib8060037.
Agrios, G.N., 2005. How plant defend themselves against pathogens. In: Plant Pathology. Business Diary PH. Coconut fiber cement board (CFB) [25.Apr.2021]; Available from:
Elsevier, pp. 207–248. https://businessdiary.com.ph/2933/coconut-fiber-cement-board-cfb/.
Ahmad, R., Hamid, R., Osman, S.A., 2019. Physical and chemical modifications of plant Camargo, M.M., Adefrs Taye, E., Roether, J.A., Tilahun Redda, D., Boccaccini, A.R.,
fibres for reinforcement in cementitious composites. Adv. Civ. Eng. 2019 (4), 1–18. 2020. A review on natural fiber-reinforced geopolymer and cement-based
https://doi.org/10.1155/2019/5185806. composites. Materials 13 (20). https://doi.org/10.3390/ma13204603.
Ahmad, W., Farooq, S.H., Usman, M., Khan, M., Ahmad, A., Aslam, F., et al., 2020. Effect Chan, E., Craig, R.E., 2006. Cocos nucifera (coconut). Species profiles for Pacific Island
of coconut fiber length and content on properties of high strength concrete. Materials agroforestry 2, 1–27.
13 (5). https://doi.org/10.3390/ma13051075. Chao-Lung, H., Le Anh-Tuan, B., Chun-Tsun, C., 2011. Effect of rice husk ash on the
Akinyemi, B.A., Adesina, A., 2021. Utilization of polymer chemical admixtures for strength and durability characteristics of concrete. Construct. Build. Mater. 25 (9),
surface treatment and modification of cellulose fibres in cement-based composites: a 3768–3772. https://doi.org/10.1016/j.conbuildmat.2011.04.009.
review. Cellulose 28 (3), 1241–1266. https://doi.org/10.1007/s10570-020-03627- Chen, B., Lu, Y., Liu, G., Palme, T., Ming, T.K., Jacobsen, R., et al.. 中国气候路径报告
3. (Report of the China climate path in English) [Jun. 2021]; Available from: https://
Akinyemi, B., Omoniyi, T.E., Elemile, O., Arowofila, O., 2020. Innovative husk-crete web-assets.bcg.com/89/47/6543977846e090f161c79d6b2f32/bcg-climate-plan-
building materials from rice chaff and modified cement mortars. Acta Technol. for-china.pdf.
Agric. 23 (2), 67–72. https://doi.org/10.2478/ata-2020-0011. Cheung, H-y, Ho, M-p, Lau, K-t, Cardona, F., Hui, D., 2009. Natural fibre-reinforced
Al-Maadeed, M.A., Labidi, S., 2014. Recycled polymers in natural fibre-reinforced composites for bioengineering and environmental engineering applications. Compos.
polymer composites. In: Natural Fibre Composites. Elsevier, pp. 103–114. B Eng. 40 (7), 655–663. https://doi.org/10.1016/j.compositesb.2009.04.014.
Alfes, C., Wiens, U., 2010. Stahlfaserbeton nach DAfStB-Richtlinie: moeglichkeiten und Chi, L., Lu, S., Yao, Y., 2019. Damping additives used in cement-matrix composites: a
Herausforderungen (Steel fiber concrete according to DAfStB guideline: possibilities review. Compos. B Eng. 164 (1–2), 26–36. https://doi.org/10.1016/j.
and challenges i English). Beton 60 (4), 128–135. compositesb.2018.11.057.
Ali, M., 2011. Coconut fibre: a versatile material and its applications in engineering. TNKKSS. Coir insulation board [25.Apr.2021]; Available from: https://www.tnkkss.
J. Civ. Eng. Construct. Technol. 2 (9), 189–197. org/coco-peat-blocks.
Ali, M., 2018. Role of post-tensioned coconut-fibre ropes in mortar-free interlocking TNKKSS, 25.Apr.2021. Coir ply. Available from: https://www.tnkkss.org/coir-ply.
concrete construction during seismic loadings. KSCE J Civ Eng 22 (4), 1336–1343. Cosgrove, D.J., Jarvis, M.C., 2012. Comparative structure and biomechanics of plant
https://doi.org/10.1007/s12205-017-1609-3. primary and secondary cell walls. Front. Plant Sci. 3, 204. https://doi.org/10.3389/
Ali, M., Chouw, N., 2013. Experimental investigations on coconut-fibre rope tensile fpls.2012.00204.
strength and pullout from coconut fibre reinforced concrete. Construct. Build. Mater. Danso, H., Manu, D., 2020. Influence of coconut fibres and lime on the properties of soil-
41 (1), 681–690. https://doi.org/10.1016/j.conbuildmat.2012.12.052. cement mortar. Case Studies in Construction Materials 12 (2), e00316. https://doi.
Ali, M., Liu, A., Sou, H., Chouw, N., 2012. Mechanical and dynamic properties of coconut org/10.1016/j.cscm.2019.e00316.
fibre reinforced concrete. Construct. Build. Mater. 30 (1), 814–825. https://doi.org/ Das, S., 2011. Life cycle assessment of carbon fiber-reinforced polymer composites. Int. J.
10.1016/j.conbuildmat.2011.12.068. Life Cycle Assess. 16 (3), 268–282. https://doi.org/10.1007/s11367-011-0264-z.
Ali, M., Li, X., Chouw, N., 2013. Experimental investigations on bond strength between Defoirdt, N., Biswas, S., Vriese, L.D., Le Tran, Q.N., van Acker, J., Ahsan, Q., et al., 2010.
coconut fibre and concrete. Mater. Des. 44, 596–605. https://doi.org/10.1016/j. Assessment of the tensile properties of coir, bamboo and jute fibre. Compos. Appl.
matdes.2012.08.038. Sci. Manuf. 41 (5), 588–595. https://doi.org/10.1016/j.compositesa.2010.01.005.
Alves Fidelis, M.E., Pereira, T.V.C., Gomes, OdFM., Andrade Silva, F de, Toledo Filho, R. Dittenber, D.B., GangaRao, H.V.S., 2012. Critical review of recent publications on use of
D., 2013. The effect of fiber morphology on the tensile strength of natural fibers. natural composites in infrastructure. Compos. Appl. Sci. Manuf. 43 (8), 1419–1429.
J. Mater. Res. Technol. 2 (2), 149–157. https://doi.org/10.1016/j.jmrt.2013.02.003. https://doi.org/10.1016/j.compositesa.2011.11.019.
Amico, S.C., Sydenstricker, T.H.D., da Silva, P.S.C.P., 2005. Evaluation of the influence of Djafari Petroudy, S.R., 2017. Physical and mechanical properties of natural fibers. In:
chemical treatment on the tensile strength of sisal fibres by a Weibull distribution Advanced High Strength Natural Fibre Composites in Construction. Elsevier,
analysis. Met. Mater. Process. 17 (3/4), 233. pp. 59–83.
Andiç-Çakir, Ö., Sarikanat, M., Tüfekçi, H.B., Demirci, C., Erdoğan, Ü.H., 2014. Physical Dong, J.F., Wang, Q.Y., Guan, Z.W., 2017. Material properties of basalt fibre reinforced
and mechanical properties of randomly oriented coir fiber–cementitious composites. concrete made with recycled earthquake waste. Construct. Build. Mater. 130 (8),
Compos. B Eng. 61 (11), 49–54. https://doi.org/10.1016/j. 241–251. https://doi.org/10.1016/j.conbuildmat.2016.08.118.
compositesb.2014.01.029. Doost-hoseini, K., Taghiyari, H.R., Elyasi, A., 2014. Correlation between sound
Anggraini V, Huat B, Asadi A, Nahazanan H. Relationship between the compressive and absorption coefficients with physical and mechanical properties of insulation boards
tensile strengths of lime-treated clay containing coconut fibres. Acta Geotechnica made from sugar cane bagasse. Compos. B Eng. 58 (1), 10–15. https://doi.org/
Slovenica;12(1):49–52. 10.1016/j.compositesb.2013.10.011.
Ardanuy, M., Claramunt, J., Toledo Filho, R.D., 2015. Cellulosic fiber reinforced cement- Duan, X-l, Zhang, J-s, 2019. Mechanical properties, failure mode, and microstructure of
based composites: a review of recent research. Construct. Build. Mater. 79 (102), soil-cement modified with fly ash and polypropylene fiber. Adv. Mater. Sci. Eng.
115–128. https://doi.org/10.1016/j.conbuildmat.2015.01.035. 2019 (2), 1–13. https://doi.org/10.1155/2019/9561794.
European Commission. A European green deal [02.Jun.2021]; Available from: https://e
c.europa.eu/info/strategy/priorities-2019-2024/european-green-deal_en.

25
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Farooq, M., Banthia, N., 2019. FRP fibre-cementitious matrix interfacial bond under Kadereit, J.W., Körner, C., Kost, B., Sonnewald, U., 2014. Strasburger − Lehrbuch der
time-dependent loading. Mater. Construcción 52 (6), 406. https://doi.org/10.1617/ Pflanzenwissenschaften (Strasburger - Textbook of Plant Sciences in English), 37th
s11527-019-1409-y. ed. Springer Berlin Heidelberg, Berlin, Heidelberg. Imprint; Springer Spektrum.
Ferreira, S.R., Lima, P.R.L., Silva, F.A., Toledo Filho, R.D., 2014. Effect of sisal fiber Kang, S.-H., Hong, S.-G., Moon, J., 2019. The use of rice husk ash as reactive filler in
hornification on the fiber-matrix bonding characteristics and bending behavior of ultra-high performance concrete. Cement Concr. Res. 115, 389–400. https://doi.org/
cement based composites. KEM 600, 421–432. https://doi.org/10.4028/www.sci 10.1016/j.cemconres.2018.09.004.
entific.net/KEM.600.421. Katkhuda, H., Shatarat, N., 2017. Improving the mechanical properties of recycled
Food and Agriculture Organization of the United Nations. FAOSTAT-Crops [25. concrete aggregate using chopped basalt fibers and acid treatment. Construct. Build.
Apr.2015]; Available from: http://www.fao.org/faostat/en/?#data/QC. Mater. 140 (1), 328–335. https://doi.org/10.1016/j.conbuildmat.2017.02.128.
Fořt, J., Šál, J., Ševčík, R., Doleželová, M., Keppert, M., Jerman, M., et al., 2021. Biomass Kenstar. Coir ply board [25.Apr.2021]; Available from: http://kenstarexim.com/coir-p
fly ash as an alternative to coal fly ash in blended cements: functional aspects. ly-board/.
Construct. Build. Mater. 271 (7), 121544. https://doi.org/10.1016/j. Khan, M., Ali, M., 2019. Improvement in concrete behavior with fly ash, silica-fume and
conbuildmat.2020.121544. coconut fibres. Construct. Build. Mater. 203 (4), 174–187. https://doi.org/10.1016/
Fu, Q., Yan, L., Ning, T., Wang, B., Kasal, B., 2020a. Interfacial bond behavior between j.conbuildmat.2019.01.103.
wood chip concrete and engineered timber glued by various adhesives. Construct. Khatri, V.N., Dutta, R.K., Venkataraman, G., Shrivastava, R., 2015. Shear strength
Build. Mater. 238 (4), 117743. https://doi.org/10.1016/j. behaviour of clay reinforced with treated coir fibres. Period. Polytech. Civ. Eng.
conbuildmat.2019.117743. https://doi.org/10.3311/PPci.7917.
Fu, Q., Yan, L., Ning, T., Wang, B., Kasal, B., 2020b. Behavior of adhesively bonded Kulkarni, A.G., Satyanarayana, K.G., Sukumaran, K., Rohatgi, P.K., 1981. Mechanical
engineered wood – wood chip concrete composite decks: experimental and behaviour of coir fibres under tensile load. J. Mater. Sci. 16 (4), 905–914. https://
analytical studies. Construct. Build. Mater. 247 (10), 118578. https://doi.org/ doi.org/10.1007/BF00542734.
10.1016/j.conbuildmat.2020.118578. Lai, C.Y., Sapuan, S.M., Ahmad, M., Yahya, N., Dahlan, K.Z.H.M., 2005. Mechanical and
Gassan, J., Chate, A., Bledzki, A.K., 2001. Calculation of elastic properties of natural electrical properties of coconut coir fiber-reinforced polypropylene composites.
fibers. J. Mater. Sci. 36 (15), 3715–3720. https://doi.org/10.1023/A: Polym. Plast. Technol. Eng. 44 (4), 619–632. https://doi.org/10.1081/PTE-
1017969615925. 200057787.
Givi, A.N., Rashid, S.A., Aziz, F.N.A., Salleh, Mohamand Amran Mohd, 2010. Le Moigne, N., van den Oever, M., Budtova, T., 2011. A statistical analysis of fibre size
Contribution of rice husk ash to the properties of mortar and concrete: a review. and shape distribution after compounding in composites reinforced by natural fibres.
J. Am. Sci. 6 (3), 157–165. Compos. Appl. Sci. Manuf. 42 (10), 1542–1550. https://doi.org/10.1016/j.
Global Alliance for Buildings and Construction, International Energy Agency and the compositesa.2011.07.012.
United Nations Enrionment Programme, 2019. Global Status Report for Building and Lecompte, T., Perrot, A., Subrianto, A., Le Duigou, A., Ausias, G., 2015. A novel pull-out
Construction: towards a zero-emission, efficient and resilient buldings and device used to study the influence of pressure during processing of cement-based
construction sector [25.Apr.2021]; Available from: https://www.worldgbc.org/s material reinforced with coir. Construct. Build. Mater. 78, 224–233. https://doi.org/
ites/default/files/2019%20Global%20Status%20Report%20for%20Buildings%20 10.1016/j.conbuildmat.2014.12.119.
and%20Construction.pdf. Lertwattanaruk, P., Suntijitto, A., 2015. Properties of natural fiber cement materials
Grasselly D, Hamm F, Quaranta G, Vitrou J. Carbon footprint of coconut fibre (coir) containing coconut coir and oil palm fibers for residential building applications.
substrates. Infos-Ctifl;249:55–59. Construct. Build. Mater. 94 (5), 664–669. https://doi.org/10.1016/j.
Gu, H., 1980-2015. Tensile behaviours of the coir fibre and related composites after conbuildmat.2015.07.154.
NaOH treatment. Mater. Des. 30 (9), 3931–3934. https://doi.org/10.1016/j. Li, Z., Wang, L., Wang, X., 2006. Flexural characteristics of coir fiber reinforced
matdes.2009.01.035, 2009. cementitious composites. Fibers Polym. 7 (3), 286–294. https://doi.org/10.1007/
Gupta, M.K., Srivastava, R.K., 2016. Mechanical properties of hybrid fibers-reinforced BF02875686.
polymer composite: a review. Polym. Plast. Technol. Eng. 55 (6), 626–642. https:// Li, Z., Wang, L., Ai Wang, X., 2007a. Cement composites reinforced with surface
doi.org/10.1080/03602559.2015.1098694. modified coir fibers. J. Compos. Mater. 41 (12), 1445–1457. https://doi.org/
Gutiérrez, RM de, Díaz, L.N., Delvasto, S., 2005. Effect of pozzolans on the performance 10.1177/0021998306068083.
of fiber-reinforced mortars. Cement Concr. Compos. 27 (5), 593–598. https://doi. Li, X., Tabil, L.G., Panigrahi, S., 2007b. Chemical treatments of natural fiber for use in
org/10.1016/j.cemconcomp.2004.09.010. natural fiber-reinforced composites: a review. J. Polym. Environ. 15 (1), 25–33.
Hasan, K.M.F., Horváth, P.G., Bak, M., Alpár, T., 2021. A state-of-the-art review on coir https://doi.org/10.1007/s10924-006-0042-3.
fiber-reinforced biocomposites. RSC Adv. 11 (18), 10548–10571. https://doi.org/ Li, M., Khelifa, M., El Ganaoui, M., 2017. Mechanical characterization of concrete
10.1039/d1ra00231g. containing wood shavings as aggregates. Int. J. Sustain. Built Environ. 6 (2),
Hettiarachchi, C., Thamarajah, G., 2020. Effect of surface modification and fibre content 587–596. https://doi.org/10.1016/j.ijsbe.2017.12.005.
on the mechanical properties of coconut fibre reinforced concrete. AMR (Adv. Magn. Liu, Q., Le Luo, Zheng, L., 2018. Lignins: biosynthesis and biological functions in plants.
Reson.) 1159, 78–99. https://doi.org/10.4028/www.scientific.net/amr.1159.78. Int. J. Mol. Sci. 19 (2) https://doi.org/10.3390/ijms19020335.
Hosseini Fouladi, M., Ayub, M., Jailani Mohd Nor, M., 2011. Analysis of coir fiber Madueke, C.I., Kolawole, F., Tile, J., 2021. Property evaluations of coir fibres for use as
acoustical characteristics. Appl. Acoust. 72 (1), 35–42. https://doi.org/10.1016/j. reinforcement in composites. SN Appl. Sci. 3 (2), 100. https://doi.org/10.1007/
apacoust.2010.09.007. s42452-021-04283-3.
Hwang, C.-L., Tran, V.-A., Hong, J.-W., Hsieh, Y.-C., 2016. Effects of short coconut fiber Maia Pederneiras, C., Veiga, R., Brito, J de, 2021. Physical and mechanical performance
on the mechanical properties, plastic cracking behavior, and impact resistance of of coir fiber-reinforced Rendering mortars. Materials 14 (4). https://doi.org/
cementitious composites. Construct. Build. Mater. 127 (11), 984–992. https://doi. 10.3390/ma14040823.
org/10.1016/j.conbuildmat.2016.09.118. Marcalikova, Z., Cajka, R., Bilek, V., Bujdos, D., Sucharda, O., 2020. Determination of
IEA. Global CO2 emissions in 2019 [25.Apr.2021]; Available from: https://www.iea.org/ mechanical characteristics for fiber-reinforced concrete with straight and hooked
articles/global-co2-emissions-in-2019. fibers. Crystals 10 (6), 545. https://doi.org/10.3390/cryst10060545.
Islam, M.S., Ahmed, S.J.U., 2018. Influence of jute fiber on concrete properties. Martinschitz, K.J., Boesecke, P., Garvey, C.J., Gindl, W., Keckes, J., 2008. Changes in
Construct. Build. Mater. 189, 768–776. https://doi.org/10.1016/j. microfibril angle in cyclically deformed dry coir fibers studied by in-situ synchrotron
conbuildmat.2018.09.048. X-ray diffraction. J. Mater. Sci. 43 (1), 350–356. https://doi.org/10.1007/s10853-
Iwaro, J., Mwasha, A., 2019. Effects of using coconut fiber–insulated masonry walls to 006-1237-7.
achieve energy efficiency and thermal comfort in residential dwellings. J. Architect. Mathura, N., Cree, D., 2016. Characterization and mechanical property of Trinidad coir
Eng. 25 (1), 4019001. https://doi.org/10.1061/(asce)ae.1943-5568.0000341. fibers. J. Appl. Polym. Sci. 133 (29), 10. https://doi.org/10.1002/app.43692.
Jahanzaib Khalil, M., Aslam, M., Ahmad, S., 2021. Utilization of sugarcane bagasse ash Mesquita, RGdA., Mendes, L.M., Sanadi, A.R., Sena Neto, AR de, Claro, P.I.C., Corrêa, A.
as cement replacement for the production of sustainable concrete – a review. C., et al., 2018. Urea formaldehyde and cellulose nanocrystals adhesive: studies
Construct. Build. Mater. 270 (4), 121371. https://doi.org/10.1016/j. applied to sugarcane bagasse particleboards. J. Polym. Environ. 26 (7), 3040–3050.
conbuildmat.2020.121371. https://doi.org/10.1007/s10924-018-1189-4.
Jayabal, S., Sathiyamurthy, S., Loganathan, K.T., Kalyanasundaram, S., 2012. Effect of Micco, V de, Balzano, A., Wheeler, E.A., Baas, P., 2016. TYLOSES and gums: a review OF
soaking time and concentration of NaOH solution on mechanical properties of structure, function and occurrence OF vessel occlusions. IAWA J. 37 (2), 186–205.
coir–polyester composites. Bull. Mater. Sci. 35 (4), 567–574. https://doi.org/ https://doi.org/10.1163/22941932-20160130.
10.1007/s12034-012-0334-2. Middleton, B., 2016. Composites: manufacture and application. In: Design and
Jo, B.W., Chakraborty, S., Kim, H., 2016. Efficacy of alkali-treated jute as fibre Manufacture of Plastic Components for Multifunctionality. Elsevier, pp. 53–101.
reinforcement in enhancing the mechanical properties of cement mortar. Mater. Mohammed, L., Ansari, M.N.M., Pua, G., Jawaid, M., Islam, M.S., 2015. A review on
Construcción 49 (3), 1093–1104. https://doi.org/10.1617/s11527-015-0560-3. natural fiber reinforced polymer composite and its applications. Int. J. Polym. Sci.
John, V.M., Cincotto, M.A., Sjöström, C., Agopyan, V., Oliveira, C.T.A., 2005. Durability 2015 (2), 1–15. https://doi.org/10.1155/2015/243947.
of slag mortar reinforced with coconut fibre. Cement Concr. Compos. 27 (5), Mwaikambo LY. Review of the history, properties and application of plant fibres. Afr. J.
565–574. https://doi.org/10.1016/j.cemconcomp.2004.09.007. Sci. Technol.;7:120–133.
Jones, Craig, Hammond, Geoffrey. Embodied carbon-the ICE database (V3.0 Beta). Naik, D.L., Sharma, A., Chada, R.R., Kiran, R., Sirotiak, T., 2019. Modified pullout test for
Available from: https://circularecology.com/embodied-carbon-footprint-database. indirect characterization of natural fiber and cementitious matrix interface
html. properties. Construct. Build. Mater. 208, 381–393. https://doi.org/10.1016/j.
Juradin, S., Boko, I., Netinger Grubeša, I., Jozić, D., Mrakovčić, S., 2019. Influence of conbuildmat.2019.03.021.
harvesting time and maceration method of Spanish Broom (Spartium junceum L.) Nasser, R., Radwan M, A., Sadek M, A., Elazab H, A., 2018. Preparation of insulating
fibers on mechanical properties of reinforced cement mortar. Construct. Build. material based on rice straw and inexpensive polymers for different roofs. Int. J. Eng.
Mater. 225 (8), 243–255. https://doi.org/10.1016/j.conbuildmat.2019.07.207. Technol. 7 (4), 1989. https://doi.org/10.14419/ijet.v7i4.14082.

26
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Olukunle, B.G., Uche, N.B., Efomo, A.O., Gideon, A., Joshua, J.K., 2018. Data on acoustic Sutas, J., Mana, A., Pitak, L., 2012. Effect of rice husk and rice husk ash to properties of
behaviour of coconut fibre-reinforced concrete. Data Brief 21, 1004–1007. https:// bricks. Procedia Eng. 32, 1061–1067. https://doi.org/10.1016/j.
doi.org/10.1016/j.dib.2018.10.133. proeng.2012.02.055.
Onuaguluchi, O., Banthia, N., 2016. Plant-based natural fibre reinforced cement Taban, E., Tajpoor, A., Faridan, M., Samaei, S.E., Beheshti, M.H., 2019. Acoustic
composites: a review. Cement Concr. Compos. 68 (6), 96–108. https://doi.org/ absorption characterization and prediction of natural coir fibers. Acoust Aust. 47 (1),
10.1016/j.cemconcomp.2016.02.014. 67–77. https://doi.org/10.1007/s40857-019-00151-8.
Pacheco-Torgal, F., Jalali, S., 2011. Cementitious building materials reinforced with Teli, M.D., Valia, S.P., Mifta, J., 2016. Application of functionalized coir fibre as eco-
vegetable fibres: a review. Construct. Build. Mater. 25 (2), 575–581. https://doi.org/ friendly oil sorbent. J. Textil. Inst. 16, 1–6. https://doi.org/10.1080/
10.1016/j.conbuildmat.2010.07.024. 00405000.2016.1220048.
Palumbo, M., Avellaneda, J., Lacasta, A.M., 2015. Availability of crop by-products in Thienel, K.-C.. Faserbeton (Fibre reinforced concrete in English) [15.Apr.2020];
Spain: new raw materials for natural thermal insulation. Resour. Conserv. Recycl. 99 Available from: https://www.unibw.de/werkstoffe/lehre/masterstudium/skripte-a
(5), 1–6. https://doi.org/10.1016/j.resconrec.2015.03.012. norganische-bindemittel/faserbeton.pdf.
Parveen, S., Rana, S., Fangueiro, R., 2017. Macro- and nanodimensional plant fiber Thomas, B.S., 2018. Green concrete partially comprised of rice husk ash as a
reinforcements for cementitious composites. In: Sustainable and Nonconventional supplementary cementitious material – a comprehensive review. Renew. Sustain.
Construction Materials Using Inorganic Bonded Fiber Composites. Elsevier, Energy Rev. 82, 3913–3923. https://doi.org/10.1016/j.rser.2017.10.081.
pp. 343–382. Tolêdo Filho, R.D., Scrivener, K., England, G.L., Ghavami, K., 2000. Durability of alkali-
Pateriya, A.S., Dharavath, K., Robert, D.J., 2021. Enhancing the strength characteristics sensitive sisal and coconut fibres in cement mortar composites. Cement Concr.
of No-fine concrete using wastes and nano materials. Construct. Build. Mater. 276 Compos. 22 (2), 127–143. https://doi.org/10.1016/S0958-9465(99)00039-6.
(3), 122222. https://doi.org/10.1016/j.conbuildmat.2020.122222. Tolêdo Filho, R.D., Ghavami, K., England, G.L., Scrivener, K., 2003. Development of
Pickering, K.L., Efendy, M.A., Le, T.M., 2016. A review of recent developments in natural vegetable fibre–mortar composites of improved durability. Cement Concr. Compos.
fibre composites and their mechanical performance. Compos. Appl. Sci. Manuf. 83, 25 (2), 185–196. https://doi.org/10.1016/S0958-9465(02)00018-5.
98–112. https://doi.org/10.1016/j.compositesa.2015.08.038. Tomczak, F., Sydenstricker, T.H.D., Satyanarayana, K.G., 2007. Studies on lignocellulosic
Polocoșer, T., Kasal, B., Stöckel, F., 2017. State-of-the-art: intermediate and high strain fibers of Brazil. Part II: morphology and properties of Brazilian coconut fibers.
rate testing of solid wood. Wood Sci. Technol. 51 (6), 1479–1534. https://doi.org/ Compos. Appl. Sci. Manuf. 38 (7), 1710–1721. https://doi.org/10.1016/j.
10.1007/s00226-017-0925-6. compositesa.2007.02.004.
Prabhakaran Nair, K.P., 2010. The Agronomy and Economy of Important Tree Crops of Tran, L.Q.N., Minh, T.N., Fuentes, C.A., Chi, T.T., van Vuure, A.W., Verpoest, I., 2015.
the Developing World. Elsevier, Oxford. Investigation of microstructure and tensile properties of porous natural coir fibre for
Quiñones-Bolaños, E., Gómez-Oviedo, M., Mouthon-Bello, J., Sierra-Vitola, L., use in composite materials. Ind. Crop. Prod. 65, 437–445. https://doi.org/10.1016/j.
Berardi, U., Bustillo-Lecompte, C., 2021. Potential use of coconut fibre modified indcrop.2014.10.064.
mortars to enhance thermal comfort in low-income housing. J. Environ. Manag. 277, Trujillo, E., Moesen, M., Osorio, L., van Vuure, A.W., Ivens, J., Verpoest, I., 2014.
111503. https://doi.org/10.1016/j.jenvman.2020.111503. Bamboo fibres for reinforcement in composite materials: strength Weibull analysis.
Raheem, A.A., Oyebisi, S.O., Akintayo, S.O., Oyeniran, M.I., 2010. Effects of admixtures Compos. Appl. Sci. Manuf. 61 (2), 115–125. https://doi.org/10.1016/j.
on the properties of corn cob ash cement concrete. Leonardo Electron. J. Pract. compositesa.2014.02.003.
Technol. 16, 13–20. Enkev, 25.Apr.2021. Available from: https://www.enkev.com/en/.
Ramakrishna, G., Priyadharshini, S., 2018. Effect of embedment length of untreated Valášek, P., D’Amato, R., Müller, M., Ruggiero, A., 2018. Mechanical properties and
natural fibres on the bond behaviour in cement mortar. Front. Struct. Civ. Eng. 12 abrasive wear of white/brown coir epoxy composites. Compos. B Eng. 146 (1),
(4), 454–460. https://doi.org/10.1007/s11709-017-0454-2. 88–97. https://doi.org/10.1016/j.compositesb.2018.04.003.
Ramakrishna, G., Sundararajan, T., 2005a. Impact strength of a few natural fibre van Loon, L.R., Glaus, M.A., 1998. Experimental and Theoretical Studies on Alkaline
reinforced cement mortar slabs: a comparative study. Cement Concr. Compos. 27 Degradation of Cellulose and its Impact on the Sorption of Radionuclides.
(5), 547–553. https://doi.org/10.1016/j.cemconcomp.2004.09.006. Vieira, A.P., Toledo Filho, R.D., Tavares, L.M., Cordeiro, G.C., 2020. Effect of particle
Ramakrishna, G., Sundararajan, T., 2005b. Studies on the durability of natural fibres and size, porous structure and content of rice husk ash on the hydration process and
the effect of corroded fibres on the strength of mortar. Cement Concr. Compos. 27 compressive strength evolution of concrete. Construct. Build. Mater. 236, 117553.
(5), 575–582. https://doi.org/10.1016/j.cemconcomp.2004.09.008. https://doi.org/10.1016/j.conbuildmat.2019.117553.
Ramlee, N.A., Jawaid, M., Zainudin, E.S., Yamani, S.A.K., 2019. Modification of oil palm Waldmann, D., May, A., Thapa, V.B., 2017. Influence of the sheet profile design on the
empty fruit bunch and sugarcane bagasse biomass as potential reinforcement for composite action of slabs made of lightweight woodchip concrete. Construct. Build.
composites panel and thermal insulation materials. J. Bionic. Eng. 16 (1), 175–188. Mater. 148 (5), 887–899. https://doi.org/10.1016/j.conbuildmat.2017.04.193.
https://doi.org/10.1007/s42235-019-0016-5. Wang, W., Chouw, N., 2017a. The behaviour of coconut fibre reinforced concrete (CFRC)
Ramli, M., Kwan, W.H., Abas, N.F., 2013. Strength and durability of coconut-fiber- under impact loading. Construct. Build. Mater. 134 (5), 452–461. https://doi.org/
reinforced concrete in aggressive environments. Construct. Build. Mater. 38 (19), 10.1016/j.conbuildmat.2016.12.092.
554–566. https://doi.org/10.1016/j.conbuildmat.2012.09.002. Wang, W., Chouw, N., 2017b. Behaviour of CFRC beams strengthened by FFRP laminates
Richaud, E., Fayolle, B., Davies, P., 2018. Tensile properties of polypropylene fibers. In: under static and impact loadings. Construct. Build. Mater. 155 (3), 956–964. https://
Handbook of Properties of Textile and Technical Fibres. Elsevier, pp. 515–543. doi.org/10.1016/j.conbuildmat.2017.08.031.
Romualdi, James P., Mandel, James A., 1964. Tensile strength of concrete affected by Wang, F., Shao, J., 2014. Modified Weibull distribution for analyzing the tensile strength
uniformly distributed and closely spaced short lengths of wire reinforcement. of bamboo fibers. Polymers 6 (12), 3005–3018. https://doi.org/10.3390/
J. Proceedings 61 (6), 657–672. polym6123005.
Roy, A., Chakraborty, S., Kundu, S.P., Basak, R.K., Majumder, S.B., Adhikari, B., 2012. Wang, B., Huang, S., Yan, L., 2021. Natural/synthetic fiber-reinforced bioepoxy
Improvement in mechanical properties of jute fibres through mild alkali treatment as composites. In: Parameswaranpillai, J., Rangappa, S., Siengchin, S., Jose, S. (Eds.),
demonstrated by utilisation of the Weibull distribution model. Bioresour. Technol. Bio-Based Epoxy Polymers, Blends and Composites. Wiley, pp. 73–116.
107, 222–228. https://doi.org/10.1016/j.biortech.2011.11.073. Wei, J., 2014. Durability of Cement Composites Reinforced with Sisal Fiber [Ph.D.
Rwth AAchen University. ITA at composites Europe in stuttgart, Germany [11.Jan.2022]; thesis]. Columbia University.
Available from: https://www.ita.rwth-aachen.de/go/id/okxs?lidx=1. Wei, J., Meyer, C., 2015. Degradation mechanisms of natural fiber in the matrix of
Saidani, M., Saraireh, D., Gerges, M., 2016. Behaviour of different types of fibre cement composites. Cement Concr. Res. 73, 1–16. https://doi.org/10.1016/j.
reinforced concrete without admixture. Eng. Struct. 113, 328–334. https://doi.org/ cemconres.2015.02.019.
10.1016/j.engstruct.2016.01.041. Wei, K., Lv, C., Chen, M., Zhou, X., Dai, Z., Shen, D., 2015. Development and
Satyanarayana, K.G., Kulkarni, A.G., Rohatgi, P.K., 1981. Structure and properties of coir performance evaluation of a new thermal insulation material from rice straw using
fibres. Proc. Indian Acad. Sci. 4, 419–436. https://doi.org/10.1007/BF02896344. high frequency hot-pressing. Energy Build. 87 (9), 116–122. https://doi.org/
Saw, S.K., Sarkhel, G., Choudhury, A., 2012. Preparation and characterization of 10.1016/j.enbuild.2014.11.026.
chemically modified Jute-Coir hybrid fiber reinforced epoxy novolac composites. White and Brown coir fibre [Jun. 2021]; Available from: https://ar.pinterest.com/pi
J. Appl. Polym. Sci. 125 (4), 3038–3049. https://doi.org/10.1002/app.36610. n/304626362271819039/visual-search/.
Sedan, D., Pagnoux, C., Smith, A., Chotard, T., 2008. Mechanical properties of hemp fibre Wongsa, A., Kunthawatwong, R., Naenudon, S., Sata, V., Chindaprasirt, P., 2020. Natural
reinforced cement: influence of the fibre/matrix interaction. J. Eur. Ceram. Soc. 28 fiber reinforced high calcium fly ash geopolymer mortar. Construct. Build. Mater.
(1), 183–192. https://doi.org/10.1016/j.jeurceramsoc.2007.05.019. 241 (4), 118143. https://doi.org/10.1016/j.conbuildmat.2020.118143.
Shakouri, M., Exstrom, C.L., Ramanathan, S., Suraneni, P., 2020. Hydration, strength, Xia, J-l, Li, P-j, 2019. Pectic enzymes. In: Encyclopedia of Food Chemistry. Elsevier,
and durability of cementitious materials incorporating untreated corn cob ash. pp. 270–276.
Construct. Build. Mater. 243 (11), 118171. https://doi.org/10.1016/j. Xia, Z.P., Yu, J.Y., Cheng, L.D., Liu, L.F., Wang, W.M., 2009. Study on the breaking
conbuildmat.2020.118171. strength of jute fibres using modified Weibull distribution. Compos. Appl. Sci.
Siakeng, R., Jawaid, M., Ariffin, H., Salit, M.S., 2018. Effects of surface treatments on Manuf. 40 (1), 54–59. https://doi.org/10.1016/j.compositesa.2008.10.001.
tensile, thermal and fibre-matrix bond strength of coir and pineapple leaf fibres with Yadav, J.S., Tiwari, S.K., 2016. Behaviour of cement stabilized treated coir fibre-
poly lactic acid. J. Bionic. Eng. 15 (6), 1035–1046. https://doi.org/10.1007/s42235- reinforced clay-pond ash mixtures. J. Build. Eng. 8, 131–140. https://doi.org/
018-0091-z. 10.1016/J.JOBE.2016.10.006.
Sumesh, K.R., Kanthavel, K., 2020. The influence of reinforcement, alkali treatment, Yan, Y., 2016. Developments in fibers for technical nonwovens. In: Advances in
compression pressure and temperature in fabrication of sisal/coir/epoxy composites: Technical Nonwovens. Elsevier, pp. 19–96.
GRA and ANN prediction. Polym. Bull. 77 (9), 4609–4629. https://doi.org/10.1007/ Yan, L., Chouw, N., 1980-2015. Compressive and flexural behaviour and theoretical
s00289-019-02988-5. analysis of flax fibre reinforced polymer tube encased coir fibre reinforced concrete
composite. Mater. Des. 52 (5), 801–811. https://doi.org/10.1016/j.
matdes.2013.06.018, 2013.

27
B. Wang et al. Journal of Cleaner Production 338 (2022) 130676

Yan, L., Chouw, N., 2013a. Experimental study of flax FRP tube encased coir fibre Yan, L., Chouw, N., Huang, L., Kasal, B., 2016a. Effect of alkali treatment on
reinforced concrete composite column. Construct. Build. Mater. 40 (400–42), microstructure and mechanical properties of coir fibres, coir fibre reinforced-
1118–1127. https://doi.org/10.1016/j.conbuildmat.2012.11.116. polymer composites and reinforced-cementitious composites. Construct. Build.
Yan, L., Chouw, N., 2013b. A comparative study of steel reinforced concrete and flax Mater. 112 (4), 168–182. https://doi.org/10.1016/j.conbuildmat.2016.02.182.
fibre reinforced polymer tube confined coconut fibre reinforced concrete beams. Yan, L., Kasal, B., Huang, L., 2016b. A review of recent research on the use of cellulosic
J. Reinforc. Plast. Compos. 32 (16), 1155–1164. https://doi.org/10.1177/ fibres, their fibre fabric reinforced cementitious, geo-polymer and polymer
0731684413487092. composites in civil engineering. Compos. B Eng. 92 (3), 94–132. https://doi.org/
Yan, L., Chouw, N., 2014a. Natural FRP tube confined fibre reinforced concrete under 10.1016/j.compositesb.2016.02.002.
pure axial compression: a comparison with glass/carbon FRP. Thin-Walled Struct. 82 Yang, H., Yan, R., Chen, H., Lee, D.H., Zheng, C., 2007. Characteristics of hemicellulose,
(121), 159–169. https://doi.org/10.1016/j.tws.2014.04.013. cellulose and lignin pyrolysis. Fuel 86 (12–13), 1781–1788. https://doi.org/
Yan, L., Chouw, N., 2014b. Dynamic and static properties of flax fibre reinforced polymer 10.1016/j.fuel.2006.12.013.
tube confined coir fibre reinforced concrete. J. Compos. Mater. 48 (13), 1595–1610. Yang, J., Ching, Y.C., Chuah, C.H., 2019. Applications of lignocellulosic fibers and lignin
https://doi.org/10.1177/0021998313488154. in bioplastics: a review. Polymers 11 (5). https://doi.org/10.3390/polym11050751.
Yan, L., Chouw, N., Jayaraman, K., 2014a. Flax fibre and its composites–A review. Yusoff, R.B., Takagi, H., Nakagaito, A.N., 2016. Tensile and flexural properties of
Compos. B Eng. 56, 296–317. polylactic acid-based hybrid green composites reinforced by kenaf, bamboo and coir
Yan, L., Chouw, N., Jayaraman, K., 2014b. Effect of column parameters on flax FRP fibers. Ind. Crop. Prod. 94, 562–573. https://doi.org/10.1016/j.
confined coir fibre reinforced concrete. Construct. Build. Mater. 55 (5), 299–312. indcrop.2016.09.017.
https://doi.org/10.1016/j.conbuildmat.2014.01.061. Zakaria, M., Ahmed, M., Hoque, M.M., Islam, S., 2017. Scope of using jute fiber for the
Yan, L., Su, S., Chouw, N., 2015. Microstructure, flexural properties and durability of coir reinforcement of concrete material. Text Cloth Sustain 2 (1), 123. https://doi.org/
fibre reinforced concrete beams externally strengthened with flax FRP composites. 10.1186/s40689-016-0022-5.
Compos. B Eng. 80 (17), 343–354. https://doi.org/10.1016/j. Zoghi, M., 2013. The International Handbook of FRP Composites in Civil Engineering.
compositesb.2015.06.011. CRC Press.

28

You might also like