You are on page 1of 31

Review

2018, Vol. 47(8) 2153–2183


! The Author(s) 2016

Hydrophobic treatment Reprints and permissions:


sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1528083716654468
of natural fibers and their journals.sagepub.com/home/jit

composites—A review

Azam Ali1, Khubab Shaker2, Yasir Nawab2, Madeha Jabbar2,


Tanveer Hussain2, Jiri Militky1 and Vijay Baheti1

Abstract
There is a growing interest in the development of natural fiber-reinforced composites,
most likely due to their wide availability, low cost, environment friendliness, and sus-
tainability. The market size for natural fiber-reinforced composites is projected to reach
$5.83 billion by 2019, with a compound annual growth rate of 12.3%. The composite
materials reinforced with wood, cotton, jute, flax or other natural fibers fall under this
category. Meanwhile, some major factors limiting the large scale production of natural
fiber composites include the tendency of natural fiber to absorb water, degradation by
microorganisms and sunlight and ultimately low strength and service life. This paper has
focused to review the different natural fiber treatments used to reduce the moisture
absorption and fiber degradation. The effect of these treatments on the mechanical
properties of these composites has also been summarized.

Keywords
Natural fibers, chemical treatment, moisture regain, composite material, mechanical
properties

Introduction
The composite material is a combined assembly of two or more constituents, giving
a synergizing effect in the resulting material. The matrix and reinforcement are two
major components of the composite material. The matrix glues the reinforcement

1
Department of Material Engineering, Faculty of Textile Engineering, Technical University of Liberec, Liberec,
Czech Republic
2
Textile Composite Materials Research Group, Faculty of Engineering and Technology, National Textile
University, Faisalabad, Pakistan

Corresponding author:
Khubab Shaker, Textile Composite Materials Research Group, Faculty of Engineering and Technology, National
Textile University, Faisalabad, Pakistan.
Email: shaker.khubab@gmail.com
2154 Journal of Industrial Textiles 47(8)

and provides desired shape and properties and may be either thermoplastic or
thermoset in nature [1]. The thermoset polymeric resins harden or cure irreversibly
when subjected to a specific curing temperature [2]. The reaction is exothermic and
heat is produced as a result of polymer hardening [3]. Reinforcement is the bunch
of fibers that carry the stress applied on composite and are largely responsible for
mechanical behavior of the composite.
The glass fiber reinforced polymer composites (GFRPC) are the largest category
of fiber-reinforced composites (FRP), as 87% of the 8.7 million ton global FRP
market is based on E-glass composites [4]. Glass fiber has high density (2.50 g/cm3).
It is obtained from non-renewable source [5] and is not bio-degradable. This has led
to many problems including waste disposal and health concerns as glass fibers are
suspected to cause skin irritations and lung cancer [6]. Every year heaps of waste
are being added on the planet which is difficult to dump. There is also need to cut
short the use of petroleum based materials and use environmental friendly sustain-
able materials.
All these things have impelled the researchers to focus on eco-friendly composite
materials. There are different types of eco-composite materials, like biodegradable
polymer based materials, natural fiber reinforced materials, etc. The natural fibers
based eco-composite materials are more suitable for use [7] as compared to bio-
degradable polymeric materials due to two main reasons: cost and availability. The
advantages of natural fibers over synthetic fibers include their low cost, low density,
biodegradability, and recyclability [8, 9] as shown in Table 1, while Figure 1 pre-
sents the cost per unit weight of natural fibers [10].
Natural fibers are abundant, based on renewable source, low cost, and have less
density [11]. These properties reveal new horizon for researchers to replace the
synthetic fibers reinforcements in composites [12]. Research in the field of natural
fiber reinforced plastics has revealed that their mechanical properties are compar-
able to those of glass fiber-reinforced composites. Therefore the natural fiber rein-
forced plastics are mostly used to produce structural components of automotive
industry [13] such as panels, doors, roofs and covers [7], furniture articles like office
chairs, door panels, safety helmets, etc. [14]. The car interior and exterior parts
have been manufactured by integrating natural fiber composites, serving a two-fold
goal; to lower the overall weight of vehicle, thus increasing fuel efficiency and to
increase sustainability of manufacturing process. Companies like Mercedes Benz,

Table 1. Comparison between natural fibers and synthetic fibers [92].

Fibers Advantages Disadvantages

Natural fibers Biodegradable Inhomogeneous quality


Low density/price Dimensional instability
Synthetic fibers Moisture resistance Difficult in recycle
Good mechanical properties Relatively high price
Ali et al. 2155

2.5

2
Price/Kg ($)

1.5

0.5

0
Sisal Ramie Kenaf Jute Hemp Flax Cotton Coir Bamboo Abaca

Figure 1. Cost per weight comparison of natural fibers [10].

Toyota, and Daimler Chrysler are already using natural fiber composites [5]. The
major advantages of using natural fibers in composites are the cost of materials,
their sustainability, and low density as compared to glass fiber [14].

Classification of natural fibers


Natural fibers have good environmental impact as they solve the disposal problem,
associated to polymeric composites. In short, the natural fibers are green substitute
of synthetic fibers in composite applications. The natural fibers are classified on the
basis of origin into plant, animal, and mineral. Plant based fibers are used exten-
sively because of their renewability and availability [7]. Further subdivisions of
plant fibers are: bast, leaf, seed, fruit, and wood, depending on the part of plant
from which fibers are obtained [15]. Mostly used plant fibers are cotton, hemp, flax,
kenaf, jute, bamboo, sisal, and coconut. Animal fibers are made up of proteins; for
example silk, wool, hair, etc. Mineral fibers are not preferred in technical applica-
tions due to their carcinogenic properties [16].
All natural fibers have an uneven and rough surface that provides good
adhesion to the resin in a composite material. The bast fibers (flax, hemp, jute)
have good mechanical properties (Table 2), as they help to provide the mechanical
support to stem. Therefore bast fibers are preferred for structural application in
composites [1]. The mechanical properties of natural fibers determine the usefulness
of a composite material. It also helps to establish the service life of composite
that can be expected from a particular fiber reinforcement. The most com-
mon properties considered are tensile strength, modulus, tenacity, density, and
2156 Journal of Industrial Textiles 47(8)

Table 2. Comparison between the properties of different reinforcements [7].

Tensile Young Specific


strength modulus modulus Tenacity Density Moisture
Plant fibers (MPa) (GPa) (GPa) (MN/m2) (g/cm3) regain (%)

Cotton 400–700 6–10 4–6.5 – 1.55 8.5


Kapok 93.2 4 12.9 – 0.45 10.9
Bamboo 571 27 18 – 1.52 –
Flax 510–910 50–70 34–48 – 1.45 12
Hemp 300–760 30–60 20–41 – 1.43 12
Jute 200–460 20–55 14–39 440–553 1.34 12
Kenaf 300–1200 22–60 – – 1.30 17
Ramie 915 23 15 – 1.55 8.5
Abaca 14 41 – – 1.52 14
Banana 530 27–32 20–24 529–754 1.35 –
Pine apple 414 60–82 42–57 413–162 1.44 –
Sisal 100–800 9–22 6–15 568–640 1.45 11
Coir 100–200 6 5.2 131–175 1.15 13

moisture regain. The properties of key natural fibers, in comparison to glass fiber
are given in Table 2.

Structure of plant fiber


In order to develop composites from natural resources, it is important to under-
stand the microstructure and chemical composition of natural fibers [15]. Cellulose,
pectin, lignin, and hemi cellulose are the main components in plant fibers, and their
percentage varies from fiber to fiber (Table 3) [17–19]. Cellulose, responsible for the
hydrophilic nature of the plant fiber, is semi crystalline; while hemicellulose is
amorphous polysaccharide, partially soluble in alkaline solution and water.
These polysaccharides are a combination of five and six carbon ring. The hemicel-
lulose acts as a cementing matrix between the cellulose micro fibrils forming the
cellulose/hemicellulose network. The lignin is hydrophobic and also acts as a
cementing agent to increase the stiffness [20, 21].
The plant fiber cell wall is divided into two sections as shown in Figure 2. First is
the primary cell wall which contains irregular network of cellulose micro fibrils that
are closely packed. The second wall is known as secondary wall, which is composed
of three separated layers S1 (outer layer), S2 (middle layer), and S3 (inner layer).
The S2 layer is responsible for mechanical properties of the fiber [22], which are
dependent on the degree of polymerization, cellulose content, and microfibrillar
angle. Fibers with higher degree of polymerization, higher cellulosic content, and a
lower microfibrillar angle exhibit higher tensile strength and modulus [23].
Ali et al. 2157

Table 3. The chemical compositions of some natural fibers [17–19].

Fibers Cellulose (%) Hemicellulose (%) Lignin (%) Waxes (%)

Bamboo 26–43 30 28 –
Banana 64 20 5 –
Coir 31–42 10–20 40–46 4
Cotton 95 2 1 0.4
Flax 70–73 18–20 2 6
Hemp 70 22 6 2
Jute 71 14 13 2
Kenaf 45–57 8–13 21.5 0.8
Oil palm 65 – 29 –
Pineapple 80 – 12.8 –
Ramie 68–76 13–15 1 6
Sisal 73 14 11 2

Figure 2. Structure of the fiber cell wall.

Pectin is also polysaccharide which holds the fibers together, while lignin is a
complex hydrocarbon polymer with both aromatic and aliphatic constituents. It
absorbs less water, insoluble in most of the solvents and cannot be broken down
into monomeric units. The glass transition temperature (Tg) of lignin is around
900 C and melting temperature around 1700 C. It is considered to be thermoplastic
polymer having amorphous nature. The lignin is not hydrolyzed by acids, but
soluble in hot alkali, readily oxidized, and easily condensable with phenol [24,5].
The chemical structure of cellulose, hemicellulose, pectin, and lignin is shown in
Figure 3 [25].
2158 Journal of Industrial Textiles 47(8)

(a) (b) OH
OH
HO O O
OH OH
O OH
O
O O
O O
OH OH
O OH
OH

(c) (d) O
HO OH

HO
O OH HO OH
OH O
OH O

O OH
OH O
OH
O O O O
O HO OH
O
O
HO OH
O O
HO
O OH O

Figure 3. Structural representations of (a) cellulose, (b) hemicellulose, (c) pectin, and (d) lignin.

Extraction of natural fibers


Various methods are used for the extraction of natural fibers from the plant stems.
Retting is the process that allows fibers to be separated from the woody core by
controlled degradation of plant stem [26]. The retting of the straw is carried out
with time by exposing it to moisture; sometimes mechanical decorticator is also
used. The available techniques of retting are given in Figure 4. Most of the avail-
able methods of retting are based on the biological activity. In this process micro-
organism fungi and bacteria from the environment degrade the polysaccharides
and hence separate the fiber bundles. The fiber extraction process has a major
impact on final fiber quality and yield. To extract good quality cellulosic fibers
from plants, microbial retting is one of the widely used techniques [27].
Sain and Panthapulakkal used fungal retting for extraction of wheat straw
fibers; the fungus was obtained from the bark of an elm tree. The enzymes pro-
duced by the fungus removed the pectinic glue holding the fiber bundles together
and released cellulosic fibers from the stem. To demonstrate the effect of extraction
process on the fiber quality, they also defibrillated the wheat straw mechanically
(using laboratory-scale mechanical refiner) before and after fungal retting. They
finally confirmed that extraction technique influences the chemical composition,
structure, and properties of the fibers. The four broad categories of retting involve
the chemical, biological, mechanical, and physical fiber separation process [28].

Biological retting: Biological retting is of two types: natural and artificial. The natural
retting comprises field or dew retting and cold water retting. Field retting is the
Ali et al. 2159

Figure 4. Retting techniques for fiber extraction.

most commonly applied retting process in regions that have specific temperature
and moisture ranges. The crops after being mowed are left on the fields until the
microorganisms separate the fibers from the xylem and cortex. To prevent over-
retting, the process needs to be stopped at the right time. In cold water retting,
anaerobic bacteria is used that breakdowns the pectin of plant straw bundles
submerged in huge water tanks, ponds or rivers, and vats. The process takes
about 6–14 days and depends on the type of bacteria and temperature of the retting
water. Environmental pollution is high due to unacceptable organic fermentation
waste waters, even though the process produces high quality fibers. Artificial
retting includes canal or warm-water retting to produce fibers of high quality in
4–5 days. The bast fibers are soaked in warm water, and separated after sufficient
retting [29].

Mechanical extraction: A cost effective and simple method is to separate bast


fiber from plant straw by mechanical means. The fibers produced by decortication
are much coarser as compared to dew or water retting. It is also termed as
the decortication technique, and is based on a simple mechanical process using
toothed breaking rollers for the decortication of bast fibers conventionally. This
technique is most suitable for retted fibers. It frees the fiber from stem in the form
of strands by removing the bark and woody parts of the stem. When the stem
comes in proximity of breaking roller, the rotating blades strike the stem and
remove the green ribbon by breaking it into pieces [30]. The fiber produced
by the decorticator requires additional cleaning to remove the shives. There are
2160 Journal of Industrial Textiles 47(8)

20–25% impurities left with fiber after decortication, which are reduced to 4–5%
after cleaning.
In advanced decortication technique, bast fibers are decorticated by impact
stress and is suitable for both unretted and retted plants [31]. The impact stress
is generated by thin hammers, which impact the surface of the stem with a very
high speed of approximately 60 m/s. This thorough stress makes it possible to
process the bast fibers without any retting and ensures a complete mechanical
separation effect of fibers and shives. It differs from the conventional technology,
where slow running crushing rollers are used, which require an intensive retting of
straw stems before processing.

Physical retting: The physical retting involves ultrasound and steam explosion meth-
ods for the separation of fibers from plant straw. The stems are harvested and
crushed; then immersed in hot water bath that contains small amounts of alkali and
surfactants and subsequently exposed to high-intense ultrasound. This process
separates the hurds (coarse parts of flax or hemp that adhere to fiber after it is
separated) from the fiber. The steam explosion method shows another suitable
alternative to the traditional field-retting procedure. Under high temperature and
pressure, steam penetrates the spaces between fiber bundles in the bast. This sudden
relaxation of the steam leads to an effective breaking up of binders, which subse-
quently results in an extensive decomposition into fine fibers [32].

Chemical and surfactant retting: This type of retting refers to the retting process in
which the straws of fiber are submerged in heated tanks. These tanks contain
aqueous solutions of sulfuric acid, sodium or potassium hydroxide, chlorinated
lime, and soda ash to dissolve the pectin. The unwanted non-cellulosic components
adhering to the fibers are removed by the use of surface active agents. High quality
fibers are produced by chemical retting, but this adds costs to the final product [33].

Natural fiber composites


The natural fibers are considered as environment friendly reinforcement for com-
posite materials, having potential of use in various applications. Different natural
fibers (e.g. sisal, flax, jute, kenaf, coir, etc.), offer a number of advantages over
synthetic fibers due to their renewable nature. Some of the unique properties of
natural fibers include: low cost, biodegradability, recyclability, low density, good
thermal properties, reduced tool wear, non-irritation to the skin, and enhanced
energy recovery [34]. The acceptable mechanical properties of natural fiber com-
posites such as elongation, ultimate breaking force, flexural properties, impact
strength, acoustic absorption, suitability for processing, and crash behavior
increases its demand for automobile components [35].
These natural fiber composites reinforce either thermoset or thermoplastic
resins. The natural fiber-reinforced thermoplastic composites have been studied
by a number of researchers for automotive applications. Among various
Ali et al. 2161

thermoplastic polymers, polypropylene (PP) is perhaps one of the most widely used
because of its distinct properties such as dimensional stability, high heat distortion
temperature, flame resistance, and transparency. The thermoset resins reinforced
with natural fiber have been prepared for structural applications. Phenolics, epoxy,
unsaturated polyester (UP), and vinyl ester are the most widely used thermoset
matrices for this purpose [19]. Most thermosetting composites are processed by
simple processing techniques such as hand lay-up, spraying, resin transfer, com-
pression, injection, and vacuum bag molding.
Flax, hemp, jute, sisal, and bamboo are the most popular reinforcement mater-
ials in bio-composites. Rice husk is used as filler for the composite materials [36,37].
The textile industry cotton waste is also used as reinforcing material for the com-
posites, having tensile strength of 27 MPa [38]. The tensile strength and modulus of
jute reinforced polyester composite are 60 MPa and 7 GPa, respectively [39], while
reinforced polyester composites have tensile strength and modulus 249  25 MPa
and 23.3  3.3 GPa, respectively [40]. Polymer matrix reinforced by woven fabric
are most commonly in structural applications because it allows the control of fiber
orientation and has good reproducibility and high productivity [41]. Hybrid com-
posites are used to obtain the tailored properties, which is difficult to achieve with
single fiber. The researchers reported a combination that natural/natural and
natural/synthetic fiber-based hybrid composites have certain advantages over
individual fiber-reinforced [42].
The incorporation of hydrophilic natural fibers in the polymer matrix results in
a heterogeneous system, having inferior properties due to lack of adhesion between
fibers and the matrix. Therefore, it is important to improve the fiber–matrix adhe-
sion for the development of natural fiber-reinforced composites. The issues in the
development of natural fiber-reinforced composites can be grouped into three
categories namely thermal degradation, moisture absorption, and biodegradation.

Thermal degradation
The thermal degradation of natural fibers is a bad aspect in the development of
natural fiber composites. The chances of thermal degradation increases with
increase in the amount of natural fibers (fiber volume fraction) in polymer
matrix composites [43]. It has been observed that in natural fibers, the thermal
decomposition of about 60% mass fraction occurred within a temperature range of
around 100 C [44]. Flax fibers are thermally degraded at temperature around
200 C. It was investigated that initially degradation of flax fibers was not signifi-
cant at 170 C for 120 min, but the strength decreased at 210 C by approximately
50% [12].

Moisture absorption
All the cellulosic fibers being hydrophilic in nature absorb moisture from environ-
ment until equilibrium is established. The moisture regain of natural fibers vary
2162 Journal of Industrial Textiles 47(8)

between 5% and 12%, as shown in Table 2. The moisture absorption causes dimen-
sional variations in the fiber as well as composite material, thus affecting the inter-
face and the mechanical properties of composite material. During composite
fabrication, the moisture may lead to poor fiber–matrix interface. There will not
be effective transfer of load due to poor interface and the composite material will
deteriorate. Treatment of natural fibers with different chemicals [45] or grafting of
vinyl monomers may help to reduce its moisture regain [46, 47].

Biodegradation and photo degradation


As discussed in the previous section, natural fibers are mostly composed of high
molecular weight carbohydrate polymer, cellulose, or proteins. The microorgan-
isms degrade lignocellulose present in the natural fibers. They can attack carbohy-
drate polymers in the cell wall, resulting in loss of strength. Lignocellulose also
undergoes photochemical degradation when exposed to ultraviolet light.
Resistance against biodegradation and UV radiation can be improved by bonding
chemicals to cell wall polymers or by adding polymer to cell matrix.
Shaker et al. [48] investigated the bio-functionality of flax fiber-reinforced com-
posites by using ZnO nanoparticles. The bioactivity was tested in terms of anti-
bacterial activity (zone of inhibition). The ZnO nanoparticles have enhanced
bioactivity against microorganisms [49]. The resulting composite material was bio-
active even with lowest amount (0.02%) of ZnO nanoparticles. Galashina et al. [50]
immobilized silver nanoparticles in flax fiber composites to impart protection
against microorganisms. The silver nanoparticles are also antimicrobial [51] and
the as-produced bio protected composites are supposed to have lower risk of fiber
degradation and enhanced service life, by restricting the growth of bacteria.

Modification of natural fibers


Natural fibers being hydrophilic are incompatible with the hydrophobic polymer
matrix. Fiber–matrix interface is affected by hydrophilic nature of cellulosic fibers
[52]. In addition to this, waxy and pectin substances cap the reactive functional
groups of the fiber and act as a hindrance to interlock with the matrix. Fiber
surface needs to be modified with different chemical treatments to enhance the
effectiveness of interfacial bonding [53]. The treatments performed for this purpose
include grafting of monomers, bleaching, acetylation, etc. Grafting of fibers with
monomers also helps to improve the thermal stability of the natural fibers. The
compatibilizer or coupling agent can also be used for the effective stress transfer
across the interface. Compatibilizer is a polymer with functional groups grafted
onto the chain of the polymer. Coupling agents are tetra functional organometallic
compounds based on titanium, silicon or zirconium and are commonly known as
silane, titanate, and zirconate coupling agents [52].
More reactive groups are exposed on the fiber surface by chemical treatment,
which in turn gives better interface and enhanced mechanical properties of
Ali et al. 2163

the composite [54]. Some chemical surface modification techniques include


treatment with sodium chlorite [55], metha acrylate [56], isocyanate [57], silane
treatment [58], acetylation [59, 60], mercerization [61–64], etherification [65, 66],
enzymatic treatment [67–71], peroxide treatments [62], benzoylation [61], dicumyl
peroxide treatment [57], plasma treatment [72–75], ozone treatments [76,77], and
grafting [78–80]. The oxidation of polyolefin [81,82] has also been reported to
improve the incompatibility between the surfaces of natural fiber and polymer
matrix.
The pre-treatment of natural fibers helps to chemically modify or clean the fiber
surface [83]. The natural fibers have a hydroxyl group due to cellulose and lignin;
therefore they are amenable to modification. This hydroxyl group is involved in
hydrogen bonding within the cellulose, thereby reducing the activity towards the
matrix. Therefore a number of treatments are performed to improve the natural
fiber composite strength, ageing and fiber matrix adhesion [79]. A well-engineered
interface not only improves the strength of composites significantly but also gives
the structural stability [84–86]. The properties and performance of composite can
be controlled by the fiber–matrix interfaces chemistry and character. Good inter-
facial properties are vital to confirm efficient load transmission from matrix
to reinforcement which aids to reduce stress absorption and improves overall
mechanical properties.
The strength of reinforced composite materials depends not only on the
substrate but also on the interface strength. The composites having good
interfacial properties can be exposed to diverse range of environmental condition-
ings. On the other hand, if composites have poor interface then environ-
ments degrade fiber/matrix interfacial bond and leads in the loss of
microstructural integrity. Consequently, the creep rates of polymers are highly
sensitive to temperature and moisture diffusion along the fiber direction than
that transverse to the fiber direction. As a result the net age of composite is reduced
[87,88]. Following are some important treatments performed for different types of
natural fibers.

Sodium chlorite treatment


To improve the mechanical and thermal properties and to enhance the hydropho-
bicity, Arifuzzaman Khan et al. [55] chemically modified the okra bast fiber by
NaClO2 treatment. The extent of modification reaction was evaluated by the
Fourier transform infra-red (FTIR) spectroscopy. Scanning electron microscopy
and the wide angle X-ray diffractometry were done to observe the morphology and
crystalline index of jute fibers. They observed a noticeable variation due to
chemical treatment in the fiber surface and improvement in its tensile properties
(i.e. tensile strength, young’s modulus and extension at break) [89].
Zahran et al. developed a novel chemical formulation for bleaching flax fibers in
one-step, based on activation of sodium chlorite by hexamethylene tetramine
(HMTA) in the presence of a nonionic wetting agent. They suggested that at
2164 Journal of Industrial Textiles 47(8)

optimum formulation, HMTA activates decomposition of NaClO2 to mainly lib-


erate nascent oxygen rather than chlorine dioxide [90].

Treatment with methacrylate


Canter et al. [56] performed the esterification of flax fibers to make them hydro-
phobic. They used 10 wt% methacrylate (MA) for 25 h at 50 C and made flax/PP
composite. They noted that the composites made by flax fibers showed good flex-
ural and tensile strength. Kaith and Kalia [91] uses 20 wt% methyl methacrylate
(MMA) and treated the flax fibers for 20 min then made the flax composite by using
phenolic resin. They noticed less absorption of moisture in the treated flax fiber
composites [59]. The treatment reaction is shown below [92]:
O O
R
OH + O O R + RCOOH
R
O

Where R is CH2 C CH3

Silane treatment
Silanes act as coupling agents and to let glass fibers adhere to a polymeric matrix
also stabilize the composite materials. Silane coupling agents reduce the hydroxyl
groups of fibers and give better interface. Hydrolysable alkoxy group leads to the
formation of silanols in the presence of moisture. Hydroxyl groups of the fibers
react with silanol, forming stable covalent bonds to the cell wall that are chemi-
sorbed on to the fiber surface. As a result silane restrains the swelling of the fibers
by creating a cross-linked network due to covalent bonding. The reaction is as
follows [93–95].

CH2 CHSiðOC2 H5 Þ3 ! CH2 CHSiðOHÞ3 þ3C2 H5 OH


CH2 CHSiðOHÞ3 þFiber  OH ! CH2 CHSiðOHÞ2 O  fiber þ H2 O

The flax fibers were made hydrophobic by silane treatment [58] and composites
were fabricated using these treated and untreated flax fibers. UP was used as matrix
and the interface of composite was studied.
The degree of crosslinking in the interface region can be improved by using the
silane coupling agents. For the purpose of modifying the natural fiber matrix inter-
face, silane coupling agents play a perfect role. Improvements in the interfacial
adhesion and for getting better mechanical properties of fiber/polymer composites,
proper treatment with silane is best by following the sol gel process. Actually silane
is absorbed and condensed on the fiber surface. The hydrogen bonding is formed
between the hydroxyl group of natural fibers and silane coupling agent. This link
Ali et al. 2165

may be then converted to the covalent bonds by heating the treated fibers at
high temperatures [96, 97]. Kalia et al. [23] investigated the effect of aminopropyl
tri-ethoxy-silane (APS) to impart hydrophobicity in cellulosic fibers. Nevell et al.
[24] observed improved mechanical properties and fungal resistance for UP and
epoxy composites, after silane treatment.

Acetylation
A well-known esterification method is acetylation. Acetylation of natural
fibers introduces plasticization to natural cellulosic fibers, mostly applied to
stabilize the cell wall of wood cellulous against the moisture, environmental deg-
radation and for improving the dimensional stability. When acetic anhydride is
applied on lignocellulose material, it reacts with hydroxyl groups of cellulose and
also prevents the diffusion of the reagents, as noticed by Yao et al. [44] who made
flax/PP composite after treating the flax fiber. They have reported 18% increase in
degree of acetylation along with significant increase in tensile and flexural
strengths.
Acetylation reaction introduces an acetyl functional group (CH3COO–) into the
fiber structure. The reaction involves the generation of acetic acid (CH3COOH) as
by-product which must be removed from the lignocellulosic material before the
fiber is used. Chemical modification with acetic anhydride (CH3–C(¼O)–O–
C(¼O)–CH3) substitutes the polymer hydroxyl groups of the cell wall with acetyl
groups, modifying the properties of these polymers so that they become hydropho-
bic [98]. The reaction of acetic anhydride with fiber is shown as:

Fiber  OH þ CH3  Cð¼ OÞ  O  Cð¼ OÞ  CH3 ! Fiber


 OCOCH3 þ CH3 COOH

O O
C
H3C C
OH + O O CH3 + CH3COOH
H3C C
O

PP or polyethylene (PE) in a toluene or xylene solution is usually used


for anhydride treatment. Fibers are impregnated in it to carry out the reaction
with hydroxyl groups on fiber surface. Hughes et al. [99] noticed that propionic
anhydride (PA) and methacrylic treatment on fiber surface showed better
yield properties. Cantero et al. [56] reported that modification mechanisms are
of two types: (a) introducing a reactive vinylic group by methacrylic anhydride
(MA) and (b) coating of hydrocarbons on surface by PA. They made flax
fiber composite with methacrylic polypropylene (MAPP) and checked the strength.
The strength of unmodified flax fiber composite was less than MA treated
composite.
2166 Journal of Industrial Textiles 47(8)

Mercerization
It is a process in which natural fibers are treated with alkali which leads to fibril-
lation and causes the breakdown of bundles of fibers into smaller fibers. Thereby, it
leads to develop a rough surface topography, that results in better fiber matrix
interface adhesion and an increase in mechanical properties [61]. Moreover, fiber
active sites increase by mercerization process and allow better fiber wetting.
Chemical composition, degree of polymerization, and molecular orientation are
also affected and give long lasting effect on mechanical properties. The alkali treat-
ment also helps to improve the properties of cotton fiber composites [100].
By this treatment there is a removal of hydrogen bonding in the network struc-
ture. Reaction which takes place during this treatment is shown below.

Fiber  OH þ NaOH ! Fiber  O  Na þ þH2 O

Sreekala et al. [62] showed that a 10–30% solution of sodium hydroxide gave the
best results when they treated flax fiber with 2.5%, 5%, 10%, 15%, 18%, 20%,
25%, and 30% NaOH. They observed that 5%, 18%, and 10% were the best
concentration for mercerization. Ray et al. [101] also treated the jute fibers with
5% NaOH solution for 0, 2, 4, 6, and 8 h at 300 C. Then dried at room temperature
for 48 h followed by oven drying at 100 C for 6 h. Garcia-Jaldon et al. [102] found
that 2% alkali solution at 200 C for 90 s at 1.5 MPa is enough for defibrillation and
degumming to individual fibers. Several workers performed mercerization and
reported that mercerization leads to increase in the amount of amorphous cellulose
and also removal of hydrogen bonding [63].
Jute fibers were washed using detergent (2 vol% in aqueous solution)
then immersed in 5 wt% NaOH solution for 24 h at room temperature. Then
again washed with distilled water to remove the excess NaOH and then dried at
70 C for 24 h by using vacuum. Likewise banana fibers were also cleaned and
refluxed in 0.25% solution of sodium hydroxide for 1 h and then washed in very
dilute acid for removing non-reacting alkali. The fiber washing was carried out
until they were alkali free, after that fibers were dried in an oven at 70 C for 3 h
[103,104].
The chemical composition of natural fiber is affected by mercerization process,
as well as the degree of polymerization and molecular orientation of cellulose
crystallites, while removing the lignin and hemicellulose [64]. This type of treatment
also converts the crystalline form of cellulose I to cellulose II [105]. This transform-
ation could be seen by FT Raman spectroscope by the intensity ratio of the
stretching modes of symmetric (C–O–C) and asymmetric proposed by Bledzki
et al. [64]. Wang et al. [106] checked the chemical properties of mercerized flax
fiber composite. The treated flax fibers were used as reinforcement and polystyrene
was used as matrix. They observed that mercerization of flax fiber improved the
mechanical properties of polystyrene composites. In the same way Bledzki et al.
[107] investigated the effect of acetylation on the flax/PP composites and noticed
the increase in mechanical properties, as well as increase in the thermal stability of
Ali et al. 2167

treated flax fiber is due to the compositional change by removal of lignin and
hemicellulose.
Jabbar et al. studied the creep behavior of alkali treated woven jute/green epoxy
composites incorporated with different loadings of chemically treated pulverized
jute fibers (PJF) was presented at various environment temperatures [108]. Yan
et al. [41] improved the mechanical properties of natural fiber composites by using
the alkali treatment of flax fiber. They used 5% NaOH for 30 min, and then made
the flax epoxy composite. They noticed a significant improvement in tensile
strength (21.9%) and flexural strength (16%) along with some improvement in
transverse strength. In the same way Van de Weyenberg et al. [105] used the
same alkaline treatment technique for flax/epoxy composite but they used 4 wt%
NaOH, 45 s treatment time and noticed favorable enhancement in mechanical
properties. Manikandan Nair et al. [109] immersed dewatered sisal fiber in 5%
and 10% NaOH solution for 1 h at 30 C. These alkaline treated fibers were then
soaked in glacial acetic acid for 1 h at 30 C followed by soaking in acetic anhydride
containing one drop of concentrated H2SO4 for 5 min. There was significant
improvement in fiber surface properties after the treatment [110].
Oladele et al. [111]. investigated the effect of chemical treatment on the constitu-
ents and tensile properties of sisal fiber. The leaves of sisal were cut and buried
underground, then wetted with water regularly close to the stream in order to
subject proper fermentation for about 15 days. These fermented leaves were com-
pletely washed and sun dried. The chemical and mechanical treatment was done
over dried sisal fibers then fraction of their constituents and tensile properties were
characterized. The results showed the removal of lignin and hemicelluloses, which
is responsible for the bonding strength of composite produced from natural fibers.
The results of the tensile test revealed that sample treated sequentially with KOH
have best tensile properties. Figures 5–7 show the SEM images of untreated and
8% NaOH treated sisal, jute, and hemp fibers [112].
The key issue in terms of overall performance is the interface between the
reinforcing agent and the matrix. Coir fiber reinforced epoxy composites’ perform-
ance depends on alkali treatment and the length of fiber. Coir fibers were treated
with sodium hydroxide for 10 days with 2%, 4%, 6%, 8%, and 10% concentration.
The length of the fibers was about 10, 20, and 30 mm. It is observed that the
mechanical properties of coir-based green composites, modulus of rupture and
internal bond, increase as a result of chemical composition modification and sur-
face modification [113]. The pretreated natural fiber based composite performed
better in terms of mechanical properties as compared to the untreated natural fiber
based composites [114, 115].

Etherification
Etherification of natural fiber composites makes it more useful and enhances cer-
tain properties [12]. Sodium hydroxide plays an important role in forming a
charged intermediate species with the fiber, which allows the faster nucleophilic
2168 Journal of Industrial Textiles 47(8)

Figure 5. SEM images of sisal fiber (a) untreated and (b) 8% NaOH treated.

Figure 6. SEM images of jute fiber (a) untreated and (b) 8% NaOH treated.

addition of epoxides, alkyl halides, benzyl chloride, acrylonitrile, and formaldehyde


[12,65,66].

Cell  OH þ NaOH ! Cell  O Naþ þ H2 O


Cell  O Naþ þ Cl  R ! Cell  O  R þ NaCl
Ali et al. 2169

Figure 7. SEM images of hemp fiber (a) untreated and (b) 8% NaOH treated.

The cellulosic materials are reacted with acrylonitrile using 4% NaOH aqueous
solution saturated by NaSCN as a swelling agent and catalyst at 40 C. The reaction
of benzyl chloride with hydroxyl groups can produce etherified natural fiber [116].
Baiardo et al. [117] applied esterification and etherification reactions to flax fiber to
modify its surface properties. The treated fiber showed unchanged structure and
morphology, with desired chemical groups on its surface for reinforcing applica-
tions in polymer composites.

Enzymatic treatment
Enzymatic treatment is very useful and interesting step when enzymes are used in
combination with chemical and mechanical methods for modification of materials.
Enzymes are efficient catalysts and are highly specific in their work under mild and
energy saving conditions. To further functionalize the lignocelluloses, oxidative
enzymes such as peroxidases can be used [67]. Laccase do the oxidation of phenolic
hydroxyls to phenolic radicals in the presence of oxygen [118]. It was further
noticed that lignin contents of single cellulosic fibers were decreased from 35%
to 24% by laccase treatment [68, 69]. Laccase when used with natural phenols such
as acetosyringore, P-coumaric acid, and syringe-aldehyde offered antimicrobial
properties to natural fiber composite (flax composite) [70, 71].

Treatment with isocyanate


An isocyanate compound contains the isocyanate functional group (–N ¼ C ¼ O),
which is highly susceptible to reaction with the hydroxyl groups of cellulose and
lignin in fibers. Isocyanate is reported to work as a coupling agent used in
2170 Journal of Industrial Textiles 47(8)

fiber-reinforced composites. The reaction between fiber and isocyanate coupling


agent is shown below:

H O
j jj
R  N ¼ C ¼ O þ HO  Fiber ! R  N  C O  Fiber

Where, R may be any chemical groups (such as alkyl).


Dried fibers treated with alkali were placed in a round bottomed flask and soaked
in an appropriate volume of carbon tetrachloride (CC14) and a little (1 ml) dibutyl tin
dilaurate catalyst. After that, round bottomed flask was fitted with a pressure equal-
izing funnel containing the urethane derivative, then urethane derivative was added
dropwise into the flask with constant stirring. This reaction was allowed to continue
for one more hour after the addition of urethane. These urethane treated fibers were
purified with acetone by refluxing for 8 h in a Soxhlet apparatus and repeated wash-
ing with distilled water, followed by drying in oven at 80 C [57].

Peroxide treatments
Peroxide treatments have attracted the attractions of most of the researchers for
treatment of cellulosic fibers, because peroxide treatment is very easy and also gives
good mechanical properties. Organic peroxides are easily decomposed to free rad-
icals, which further go to react with cellulose of fiber and also hydrogen group of
matrix as proposed by Sreekala et al. [62]. The schematic of peroxide treatment is
shown below:

_ þ Cellulose  H ! R  OH þ Cellulose
RO

During peroxide treatment, fibers are treated with 6% dicummyl peroxide or


benzoyl peroxide in acetone solution for 30 min after alkali pretreatment of fibers.
Flax fibers were treated with dicumyl peroxide from acetone solution. Fibers were
soaked in solution at 70 C for 30 min. Then fibers were washed with distilled water,
and placed for 24 h in oven at 80 C, to improve the hydrophobic properties.

Benzoylation
Benzoylation is an important transformation in organic synthesis, and benzoyl
chloride is most often used in fiber treatment. Benzoyl chloride includes benzoyl
(C6H5C ¼ O) which is attributed to decrease the hydrophilic nature of the treated
fiber and improve interaction with the hydrophobic matrix. The reaction between
the cellulosic hydroxyl group of the fiber and benzoyl chloride is shown below.

O O
Fiber O Na + Cl C Fiber O C + NaCl
Ali et al. 2171

Joseph et al. [61] used sodium hydroxide and benzoyl chlorite (C6H5COCl)
solution for treatment of surface of sisal fibers. The fibers were alkaline pre-treated
to activate the hydroxyl group of lignin and cellulose of fibers. Then fibers were
immersed in 10% NaOH and benzoyl chloride solution for 15 min. After that fibers
were soaked in ethanol for 1 h to remove benzoyl chloride and fibers were then
washed and dried at 80 C for 24 h Surface modification and improvement in hydro-
phobicity was observed after the treatment.

Dicumyl peroxide treatment


Fibers pretreated with alkali (30 g) were soaked with 6% solution of dicumyl per-
oxide (DCP) in acetone for 30 min, followed by decantation of the solution and
drying of the fibers. There was significant improvement in mechanical properties
and decrease in moisture regain of the composites [57]. The SEM images of the
untreated and treated fiber surfaces are given in Figure 8. The treated fiber surfaces
underwent benzoylation, silane treatment, and dicumyl peroxide treatment [12].
Sisal–PE composites, prepared with DCP, show reduced water absorption,
although the mass% of the absorption has the same dependence on time as the
untreated composites. Significantly lower mass% change values of DCP treated
composites may be attributed to the improved interfacial bonding. Grafting reac-
tions reduce the number of voids between sisal and the PE matrix, and partially
prevent fibers getting into contact with water [119].

Plasma treatment
This is very effective treatment method to modify the surface of natural polymers
without any change in their bulk properties. The plasma discharge can be generated
by either cold plasma treatment or corona treatment. Both are the plasma treat-
ment methods when ionized gas has an equivalent number of charged (positive and
negative) molecules that react with the surface of the material present. Frequency
of the electric discharge is the main distinguishing feature between the two cate-
gories of plasmas. Cold plasma of high-frequency can be produced by microwave
energy, whereas alternating current of a lower frequency discharge at atmospheric
pressure produces corona plasma. The type of ionized gas influenced the modifi-
cation of the wood and synthetic polymer surfaces [7]. Maldas et al. [75] reported
an avenue to activate a wood surface for getting better adhesion with polyolefin by
exposure to plasmas. The plasma discharges were generated by either corona treat-
ment or cold plasma treatment. Further, Podgorski and Roux [74] have recently
studied the number of the polar component of surface energy of pine wood for
plasma modification that include power, distance of samples to plasma source,
treatment time, stability of plasma treatment, and type of gas. Corona discharge
was used to treat the pulp sheets having moisture contents of up to 85%. The
process was carried out in the presence of air and nitrogen atmospheres, and chem-
ical modification of the sheets were evaluated by dye [73]. There were no evidence
2172 Journal of Industrial Textiles 47(8)

Figure 8. SEM photographs of fiber surfaces after chemical treatment: (a) untreated, (b) silane
treated, (c) benzoylated, and (d) dicumyl peroxide treated.

of an increase in carboxylic groups found on the treated paper surface but it was
noticed that the quantity of aldehyde groups increased with increasing corona
treatment. The authors were unable to know that either the increase in bonding
strength was due to a reaction of hydroxyl groups with aldehyde groups or the
degradation of hemicelluloses that adsorb between the surfaces of the paper sheets.
Uehara and Sakata [72] noticed that air corona treatment caused a reduction in the
molecular weight of cellulose and improvement in hydrophobic properties of cel-
lulosic fibers.

Ozone treatments
The surface of cellulosic materials can be modified by exposure to ozone or oxygen-
fluorine gas. Hedenberg and Gatenholm [76] exposed cellulose fibers and PE
membranes and films to ozone gas. The adhesion properties of the composites
were improved when low density poly-ethylene (LDPE) was treated with
Ali et al. 2173

ozone treatment. These two mechanisms were proposed for the increase in bonding
strength, which involved hydrogen bonding of LDPE that contained carbonyl
groups as revealed by spectroscopy and covalent bonding initiated by the decom-
position of hydro peroxides.
The ozone treatment affects the contact angles for different liquids with the
surfaces and increases the total surface energies. More importantly, the oxidation
procedure increases the polar component of the surface energy of LDPE. This
increased polar character of LDPE will provide a significant better basis for inter-
actions with the numerous hydroxyl groups on cellulose, which in turn reduces the
hydrophilic character of the material.
Chtourou et al. [77] increased the strength properties of PE and pulp composites
by exposing the PE to oxygen-fluorine gas. The authors found that by exposure, the
specific (acid–base) interaction parameter increased, and the improvement of prop-
erties were attributed to hydrogen bonding between both fluorine and hydroxyl
groups, carbonyl, and hydroxyl groups.

Grafting
Graft copolymerization of vinyl monomers onto wood and cellulose based mater-
ials was first reported in 1953. Since then grafting has received much attention for
modification of cellulose. Meyer [79] studied the applications for wood graft
copolymers. Hon [78] reviewed four challenges of cellulosic modification by graft
copolymerization: limited cellulose involvement, homo polymerization of mono-
mer, limited control of molecular weight, and poor replication of grafting yields.
The efficiency of grafting monomers onto natural surfaces increases when the sur-
faces of fibers, wood and plant based materials have been activated. Hill et al. [120]
noticed that wood modified by methyl acrylic anhydride allowed for the grafting of
MMA and styrene. An adequate study by Hill and Cetin was made for using the
method to protect wood against UV irradiation [121]. Grelier et al. [122] grafted
with isocyanate, that have an UV absorbing chromophore to medium density fiber
boards by microwave activation.
Oligo esterification method was used to activate pine wood blocks with maleic
anhydride along with allylglycidyl ether, in order to graft hydride-terminated poly-
dimethylsiloxane and bis (trimethylsiloxy) methylsilane [80]. Firstly, the absorb-
ency of cellulose fibers was improved by activating the surface with an ethoxysilane
and then by doing copolymerization of acrylic acid onto the surface, initiated by
potassium persulfate [57, 123–134]. In comparison with the number of studies for
graft copolymerization onto cellulose, research on the copolymerization onto more
complex substances like wood and plant fiber is scarce. Presence of lignin and other
extractives found in wood act as antioxidants through resonance stabilization of
free radicals, thus inhibiting polymerization and grafting. Actually lignin inhibits
the copolymerization process [124].
The type of catalyst and the nature of lignin has inhibiting ability in graft
copolymerization [125]. However, lignin may become more modified than the
2174 Journal of Industrial Textiles 47(8)

holocellulose depending upon the reaction conditions [126] or may reduce homo-
polymerization of monomer due to a chain transfer mechanism. Zheng et al. have
described the participation of lignin in the graft copolymerization of polymethyl
methacrylate onto bagasse pith initiated by Fe2þ and hydrogen peroxide [127] and
potassium permanganate (KMNO4) [128]. Lignin was more susceptible to grafting
and specifically, hydroxyl cyclohexadienyl radicals formed by lignin participated in
the copolymerization.
Similar to the concept of bagasse treatment, same initiators were used in the
graft copolymerization of stone ground wood pulp. It was noticed that phenolic
hydroxy radicals produced by reaction with ferric ions initiated copolymerization
[129]. The initial stage of delignification, in which loss of lignin occurred by 0.6%
by the sodium chlorite method of pulping, increased the grafting efficiency, but
delignification did not have the same effect. The authors also noticed, depending
upon reaction conditions that the phenyl ring formed phenoxy radicals that either
initiated copolymerization or decomposed the phenyl rings of lignin. Marchetti
et al. reported the deposition of manganese dioxide (MnO2) on the wood cell
wall when KMnO4 was used as an initiator, which contributed to the radical
chain transfer onto wood [130]. Ghosh et al. noticed the effect of lignin on the
copolymerization in partially delignified jute fibers. Only slightly improved initial
delignification by sodium chlorite grafting, but in the presence of only 0.8% lignin,
the grafting efficiency was doubled [131]. Bleached and unbleached sugarcane fibers
were grafted with hydroxy phenyl benzotriazole UV absorber and nitroxide radical
of piperidinyloxy followed by treatment with acetic anhydride [135].

Fluorocarbon treatment
Ali et al. [136] used fluorocarbons, hydrocarbons, and hybrid fluorocarbons for the
treatment of jute fibers. These chemicals have a lower surface free energy. It is a
known fact that lower the surface free energy of a material, lower will be the
moisture regain. A significant difference in moisture regain values of treated and
untreated reinforcement samples was observed at a concentration of 40 g/l. The
composite made from treated reinforcement regained very low moisture content as
well as exhibited improved mechanical properties (tensile and flexural strength).
Being dual nature (hydrophilic and hydrophobic groups) of hybrid fluorocarbon,
the treated jute fibers and corresponding composites showed better properties as
compared to other two chemicals.
The effect of different pre-treatments on the properties of different natural fibers
is summarized in Table 4. One of the practical methods of chemical modification is
alkali treatment or mercerization of natural fibers. Alkaline treatment leads to the
disruption of hydrogen bonding in the network structure, increase in surface
roughness of fibers, removal of lignin, wax and oils from surface, etc. It is beneficial
in composite applications of fibers due to (a) better mechanical interlocking by
increase in surface roughness and (b) increase in number of possible reaction sites
due to increased cellulose percentage. The alkali treatment is advantageous over
Ali et al. 2175

Table 4. Comparison of results of the pre-treatments on natural fiber composites.

Type of pre-treatment Material Effect

Sodium chlorite Jute fibers Significant improvement in tensile strength,


treatment young’s modulus and extension at break [55]
Treatment with Jute fiber Improvement in flexural and tensile strength [56]
metha acrylate
Silane treatment Flax fibers Improvement in hydrophobic and mechanical
properties [58]
Acetylation Flax fibers Increase in tensile strength and flexural strength
[26]
Mercerization Jute, flax fibers Reduction in moisture regain due to better
interface and improvement in mechanical
properties [62]
Treatment with Flax fibers Surface modification [57]
isocyanate
Enzyme Flax fibers Decrease in lignin contents from 35% to 24% [67]
Per-oxide treatment Natural fibers Decrease in moisture regain [62]
Benzoylation Sisal fibers Surface modification and improvement in
hydrophobicity [61]
Plasma treatment Natural fibers Surface modification and improvement in
hydrophobicity [75]
Ozone treatments Natural fibers, Increased the strength properties of polyethylene
pulp and pulp composites [77]
Grafting Natural fibers, Surface modification and improvement in ultra-
wood violet protection properties, hydrophobicity,
and mechanical properties [63–74]

the other techniques because it is simple, inexpensive, quick, relatively eco-friendly,


reasonably less deterioration of fiber properties, widely applicable for different
natural fibers, good performance with higher repeatability, etc. However, the dis-
advantage is reduction in mechanical strength of fibers.

Conclusions
This article focused on the literature review about natural fiber-reinforced com-
posites and the limitations evolving in the development and serviceability of these
composites. These natural fiber-reinforced composites are a potential substitute of
glass fiber-reinforced composites. In order to exploit the full potential of natural
fibers, these limitations must be overcome. Firstly, an appropriate fiber surface
treatment should be carried out. The paper also presents literature on the natural
fiber treatments used by researchers to reduce the moisture regain of natural
fiber composites and improve the compatibility between the fiber and the
2176 Journal of Industrial Textiles 47(8)

polymer matrix. A wide range of pretreatments are available, such as sodium


chlorite, methacrylate, silane, peroxide, enzyme, plasma, ozone treatment, etc.
Some researchers have also worked to impart functionality against microorganisms
in these composite materials. Such composites not only help to maintain hygiene
but also have enhanced service life. Hence natural fibers can be used effectively to
reinforce the composite materials.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, author-
ship, and/or publication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication
of this article.

References
[1] Gay D, Hoa SV and Tsai SW. Composite materials: design and applications, 1st edn.
Boca Raton, FL: CRC Press, 2003.
[2] Blest DC, Duffy BR, McKee S, et al. Curing simulation of thermoset composites.
Compos Part A Appl Sci Manuf 1999; 30: 1289–1309.
[3] Pascault J, Sautereau H, Verdu J, et al. Thermosetting polymers. New York: CRC Press,
2002.
[4] Shah DU. Developing plant fibre composites for structural applications by optimising
composite parameters: A critical review. J Mater Sci 2013; 48: 6083–6107.
[5] Westman MP, Laddha SG, Fifield LS, et al. Natural fiber composites: a review.
Washington: Pacific Northwest National Laboratory, 2010.
[6] Bos HL. The potential of flax fibres as reinforcement for composite materials. Eindhoven
University of Technology, 2004.
[7] Polona Dobnik Dubrovski (ed.) Woven fabric engineering. Rijeka: Sciyo, 2010.
[8] Joshi SV, Drzal LT, Mohanty AK, et al. Are natural fiber composites environmentally
superior to glass fiber reinforced composites? Compos Part A Appl Sci Manuf 2004; 35:
371–376.
[9] Van Voorn B, Smit H, Sinke R, et al. Natural fibre reinforced sheet moulding com-
pound. Compos Part A Appl Sci Manuf 2001; 32: 1271–1279.
[10] Dittenber DB and Gangarao HVS. Critical review of recent publications on use of
natural composites in infrastructure. Compos Part A Appl Sci Manuf 2012; 43:
1419–1429.
[11] Celino A, Freour S, Jacquemin F, et al. Characterization and modeling of the moisture
diffusion behavior of natural fibers. J Appl Polym Sci 2013; 130: 297–306.
[12] Kalia S, Kaith BSBS and Kaur I. Pretreatments of natural fibers and their application
as reinforcing material in polymer composites—a review. Polym Eng Sci 2009; 49:
1253–1272.
[13] Huda MS, Drzal LT and Misra M. Natural fiber reinforced biodegradable polymer
composites for automotive applications. In: 7th annual automotive composites confer-
ence and exhibition ACCE 2007, Troy, 2007, pp.1–31.
Ali et al. 2177

[14] Riedel U and Nickel J. Applications of natural fiber composites for constructive parts
in aerospace, automobiles and other areas. Biopolym Online, Epub ahead of print 2005.
[15] Masuelli MA (ed) Fiber reinforced polymers—the technology applied for concrete repair.
Rijeka: InTech, 2013.
[16] Barrett JC, Lamb PW and Wiseman RW. Multiple mechanisms for the carcinogenic
effects of asbestos and other mineral fibers asbestos and multistage. Environ Health
Perspect 1989; 81: 81–89.
[17] Khalil HA and Suraya NL. Anhydride modification of cultivated kenaf bast
fibers: Morphological, spectroscopic and thermal studies. BioResources 2011; 6:
1122–1135.
[18] Hoareau W, Trindade WG, Siegmund B, et al. Sugar cane bagasse and curaua lignins
oxidized by chlorine dioxide and reacted with furfuryl alcohol: Characterization and
stability. Polym Degrad Stab 2004; 86: 567–576.
[19] Thakur VK and Thakur MK. Processing and characterization of natural cellulose
fibers/thermoset polymer composites. Carbohydr Polym 2014; 109: 102–117.
[20] Abdelmouleh M, Boufi S, Belgacem MN, et al. Short natural-fibre reinforced poly-
ethylene and natural rubber composites: Effect of silane coupling agents and fibres
loading. Compos Sci Technol 2007; 67: 1627–1639.
[21] Mwaikambo LY and Ansell MP. Chemical modification of hemp, sisal, jute and kapok
fibres by alkalization. Appl Polym Sci 2002; 84: 2222–2234.
[22] Rong MZ, Zhang MQ, Liu Y, et al. Effect of fibre treatment on the mechanical proper-
ties of unidirectional sisal-reinforced epoxy composites. Compos Sci Technol 2001; 61:
1437–1447.
[23] Kalia S, Kaith BS, Kaur I (eds) Cellulose fibers: bio- and nano-polymer composites.
Berlin: Springer, 2011.
[24] Nevell TP, Zeronian SH (eds) Cellulose chemistry and its applications. Hemel Hemstead:
Ellis Horwood Ltd., 1985.
[25] Chen H. Biotechnology of lignocellulose. Beijing: Chemical Industry Press, 2005.
Available at: http://link.springer.com/10.1007/978-94-007-6898-7
[26] Rao KMM and Rao KM. Extraction and tensile properties of natural fibres: Vakka,
date and bamboo. Compos Struct 2007; 77: 288–295.
[27] Yu HYC. Study on microbe retting of kenaf fibre. Enzyme Microb Technol 2007; 40:
1806–1809.
[28] Paridah MT, Basher AB, SaifulAzry S, et al. Retting process of some bast plant fibres
and its effect on fibre quality: A review. BioResources 2011; 6: 5260–5281.
[29] Gañán P, Zuluaga R, Velez JM, et al. Biological natural retting for determining the
hierarchical structuration of banana fibers. Macromol Biosci 2004; 4: 978–983.
[30] Das PK, Nag D, Debnath S, et al. Machinery for extraction and traditional spinning of
plant fibres. Indian J Tradit Knowl 2010; 9: 386–393.
[31] Kozlowski R, Muzyczek M and Mieleniak M. Advanced decortication technology for
unretted bast fibres. J Nat Fibers 2004; 1: 85–95.
[32] Evans JD, Akin DE and Foulk JA. Flax-retting by polygalacturonase-containing
enzyme mixtures and effects on fiber properties. J Biotechnol 2002; 97: 223–231.
[33] Kimmel L and Akin DE. Chemical/physical retting of flax using detergent and oxalic
acid at high pH. Text Res J 1998; 68: 942–947.
[34] Thakur VK, Thakur MK and Gupta RK. Review: Raw natural fiber–based polymer
composites. Int J Polym Anal Charact 2014; 19: 256–271.
2178 Journal of Industrial Textiles 47(8)

[35] Ahmad F, Choi HS and Park MK. A review: Natural fiber composites selection in view
of mechanical, light weight, and economic properties. Macromol Mater Eng 2015; 300:
10–24.
[36] Arjmandi R, Hassan A, Majeed K, et al. Rice husk filled polymer composites. Int J
Polym Sci 2015; 2015. DOI 10.1155/2015/501471.
[37] Azeez TO. Effect of rice husk filler on mechanical properties of polyethelene matrix
composite. Int J Curr Res Rev 2013; 05: 111–118.
[38] Nazir MU, Shaker K, Ahmad S, et al. Investigating the mechanical behavior of
composites made from textile industry waste. J Text Inst 2016; 108: 835–839.
[39] Munikenche Gowda T, Naidu ACB and Chhaya R. Some mechanical properties of
untreated jute fabric-reinforced polyester composites. Compos Part A Appl Sci Manuf
1999; 30: 277–284.
[40] Shah DU, Schubel PJ and Clifford MJ. Can flax replace E-glass in structural com-
posites? A small wind turbine blade case study. Compos Part B Eng 2013; 52:
172–181.
[41] Yan L, Chouw N and Yuan X. Improving the mechanical properties of natural fibre
fabric reinforced epoxy composites by alkali treatment. J Reinf Plast Compos 2012; 31:
425–437.
[42] Jawaid M, Alothman OY, Paridah MT, et al. Effect of oil palm and jute fiber treatment
on mechanical performance of epoxy hybrid composites. Int J Polym Anal Charact
2014; 19: 62–69.
[43] Fakhrul T and Islam MA. Degradation behavior of natural fiber reinforced polymer
matrix composites. Procedia Eng 2013; 56: 795–800.
[44] Yao F, Wu Q, Lei Y, et al. Thermal decomposition kinetics of natural fibers:
Activation energy with dynamic thermogravimetric analysis. Polym Degrad Stab
2008; 93: 90–98.
[45] Taj S, Munawar MA and Khan S. Natural fiber reinforced polymer composites. Proc
Pak Acad Sci 2007; 44: 129–144.
[46] Zakaria S and Kok Poh L. Polystyrene-benzoylated EFB reinforced composites. Polym
Plast Technol Eng 2002; 41: 951–962.
[47] Wang B, Panigrahi S, Tabil L, et al. Pre-treatment of flax fibres for use in rotationally
molded biocomposites. J Reinf Plast Compos 2007; 26: 447–463.
[48] Shaker K, Ashraf M, Jabbar M, et al. Bioactive woven flax-based composites:
Development and characterisation. J Ind Text 2016; 46: 549–561.
[49] Padmavathy N and Vijayaraghavan R. Enhanced bioactivity of ZnO nanoparticles—an
antimicrobial study. Sci Technol Adv Mater 2008; 9: 035004.
[50] Galashina VN, Moryganov PA and Dymnikova NS. Bio-protected flax nanocompo-
sites as basis for manufacturing of high-tech eco-products. Russ J Gen Chem 2012; 82:
2270–2278.
[51] Song HY, Ko KK, Oh IH, et al. Fabrication of silver nanoparticles and their anti-
microbial mechanisms. Eur Cell Mater 2006; 11(Suppl. 1): 58.
[52] Saheb DN and Jog JP. Natural fiber polymer composites: A review. Adv Polym Technol
1999; 18: 351–363.
[53] Militky J and Jabbar A. Comparative evaluation of fiber treatments on the
creep behavior of jute/green epoxy composites. Compos Part B Eng 2015; 80:
361–368.
Ali et al. 2179

[54] Dash B, Rana A, Mishra S, et al. Novel low-cost jute-polyester composite. II. SEM
observation of the fractured surfaces. Polym Plast Technol Eng 2000; 39: 333–350.
[55] Arifuzzaman Khan GM, Shaheruzzaman M, Rahman MH, et al. Surface modification
of okra bast fiber and its physico-chemical characteristics. Fibers Polym 2009; 10: 65–70.
[56] Cantero G, Arbelaiz A, Llano-Ponte R, et al. Effects of fiber treatment on wettability
and mechanical behaviour of flax/polypropylene composites. Compos Sci Technol 2003;
63: 1247–1254.
[57] Joseph K and Thomast S. Effect of chemical treatment on the tensile properties of short
sisal fibre-reinforced polyethylene composites. Polymer (Guildf) 1996; 37: 5139–5149.
[58] Alix S, Lebrun L, Morvan C, et al. Study of water behaviour of chemically treated flax
fibres-based composites: A way to approach the hydric interface. Compos Sci Technol
2011; 71: 893–899.
[59] John MJ and Anandjiwala RD. Recent developments in chemical modification and
characterization of natural fiber-reinforced composites. Polym Compos 2008; 29:
187–207.
[60] Velde V, Kathleen de and Kiekens P. Thermoplastic polymers: Overview of several
properties and their consequences in flax fibre reinforced composites. Polym Test 2001;
20: 885–893.
[61] Joseph K, Mattoso LHC, Toledo RD, et al. Natural fiber reinforced thermoplastic
composites. Nat Polym Agrofibers Compos 2000; 159: 159–201.
[62] Sreekala M, Kumaran M, Joseph S, et al. Oil palm fibre reinforced phenol formalde-
hyde composites: Influence of fibre surface modifications on the mechanical perform-
ance. Appl Compos Mater 2000; 7: 295–329.
[63] Mishra S, Misra M, Tripathy SS, et al. Influence of chemical surface modification on
the performance of sisal-polyster biocomposites. Polym Compos 2002; 23: 164.
[64] Bledzki AK, Fink H and Specht K. Unidirectional hemp and flax EP- and PP-compo-
sites: Influence of defined fiber treatments. J Appl Polym Sci 2004; 93: 2150–2156.
[65] Matsuda H. Chemical modification of solid wood, 1st edn. New York, NY: Marcel
Dekker Inc., 1996.
[66] Kumar S. Chemical modification of wood. Wood Fiber Sci 1994; 26: 270–280.
[67] Grönqvist S, Buchert J, Rantanen K, et al. Activity of laccase on unbleached and
bleached thermomechanical pulp. Enzyme Microb Technol 2003; 32: 439–445.
[68] Kudanga T, Nyanhongo GS, Guebitz GM, et al. Potential applications of laccase-
mediated coupling and grafting reactions: A review. Enzyme Microb Technol 2011;
48: 195–208.
[69] Kim S and Cavaco-Paulo A. Laccase-catalysed protein-flavonoid conjugates for flax
fibre modification. J Appl Microbiol 2012; 93: 585–600.
[70] Aracri E, Fillat A, Colom JF, et al. Enzymatic grafting of simple phenols on flax and
sisal pulp fibres using laccases. Bioresour Technol 2010; 101: 8211–8216.
[71] Fillat A, Gallardo O, Vidal T, et al. Enzymatic grafting of natural phenols to flax fibres:
Development of antimicrobial properties. Carbohydr Polym 2012; 87: 146–152.
[72] Uehara T and Sakata I. Effect of corona discharge treatment on cellulose prepared
from beech wood. J Appl Polym Sci 1990; 41: 1695–1706.
[73] Sakata I, Morita M, Furuichi H, et al. Improvement of plybond strength of paperboard
by corona treatment. J Appl Polym 1991; 42: 2099–2104.
[74] Podgorski I and Roux M. Wood modification to improve the durability of coatings.
Surf Coat Int 1999; 82: 590–596.
2180 Journal of Industrial Textiles 47(8)

[75] Maldas D, Kokta B and Daneault C. Influence of coupling agents and treatments on
the mechanical properties of cellulose fiber–polystyrene composites. J Appl Polym Sci
1989; 37: 751–775.
[76] Hedenberg P and Gatenholm P. Conversion of plastic/cellulose waste into composites.
II. Improving adhesion between polyethylene and cellulose using ozone. J Appl Polym
Sci 1996; 60: 2377–2385.
[77] Chtourou H, Riedl B and Kokta BV. Strength properties of wood-PE composites:
Influence of pulp ratio and pretreatment of PE fibers. TAPPI J 1997; 80: 141–151.
[78] Hon DNS. Graft copolymerization of lignocellulosic fibers. Washington, DC: American
Chemical Society, 1982.
[79] Meyer J. Wood-polymer composites and their industrial applications, in wood technology:
Chemical aspects. Washington, DC: American Chemical Society, 1977.
[80] Sebe G and Brook MA. Hydrophobization of wood surfaces: Covalent grafting of
silicone polymers. Wood Sci Technol 2001; 35: 269–282.
[81] Gugumus F. Thermooxidative degradation of polyolefins in the solid state-6. Kinetics
of thermal oxidation of polypropylene. Polym Degrad Stab 1998; 62: 235–243.
[82] Kotoyori T. Activation energy for the oxidative thermal degradation of plastics.
Thermochim Acta 1972; 5: 51–58.
[83] Edeerozey AM, Akil HM, Azhar A, et al. Chemical modification of kenaf fibers. Mater
Lett 2007; 61: 2023–2025.
[84] Drzal LT. The interphase in epoxy composites. In: Epoxy resin and composites II.
Berlin: Springer, 1986, pp.1–32.
[85] Houshyar S, Shanks R and Hodzic A. The effect of fiber concentration on mechanical
and thermal properties of fiber-reinforced polypropylene composites. J Polym Sci 2005;
96: 2260–2272.
[86] Lee S-H, Wang S, Pharr GM, et al. Evaluation of interphase properties in a cellulose
fiber-reinforced polypropylene composite by nanoindentation and finite element ana-
lysis. Compos Part A Appl Sci Manuf 2007; 38: 1517–1524.
[87] Ray BC and Rathore D. Durability and integrity studies of environmentally condi-
tioned interfaces in fibrous polymeric composites: Critical concepts and comments. Adv
Colloid Interface Sci 2014; 209: 68–83.
[88] Ray BC. Temperature effect during humid ageing on interfaces of glass and carbon
fibres reinforced epoxy composites. J Colloid Interface Sci 2006; 298: 111–117.
[89] Mustata A. Factors influencing fiber-fiber friction in the case of bleached flax. Cellul
Chem Technol 1997; 31: 405–413.
[90] Zahran MK, Rehan MF and El-Rafie MH. Single bath full bleaching of flax fibers
using an activated sodium chlorite/hexamethylene tetramine system. J Nat Fibers 2005;
2: 49–67.
[91] Kaith BS and Kalia S. Grafting of flax fiber (Linum usitatissimum) with vinyl mono-
mers for enhancement of properties of flax-phenolic composites. Polym J 2007; 39:
1319–1327.
[92] Zhu J, Zhu H, Njuguna J, et al. Recent development of flax fibres and their reinforced
composites based on different polymeric matrices. Materials (Basel) 2013; 6:
5171–5198.
[93] Mohanty A, Drzal L and Misra M. Novel hybrid coupling agent as an adhesion pro-
moter in natural fiber reinforced powder polypropylene composites. J Mater Sci Lett
2002; 21: 1885–1888.
Ali et al. 2181

[94] Kim J-K, Sham M-L and Wu J. Nanoscale characterisation of interphase


in silane treated glass fibre composites. Compos Part A Appl Sci Manuf 2001; 32:
607–618.
[95] Gupta A and Kumar A. Chemical properties of natural fiber composites and mech-
anisms of chemical modifications. Asian J Chem 2012; 24: 1831–1836.
[96] Singha A and Rana AK. Effect of aminopropyltriethoxysilane (APS) treatment on
properties of mercerized lignocellulosic grewia optiva fiber. J Polym Environ 2013; 21:
141–150.
[97] Xiea Y, Hillb CAS, Xiaoa Z, et al. Silane coupling agents used for natural fiber/
polymer composites: A review. Compos Part A Appl Sci Manuf 2010; 41: 806–819.
[98] Li X, Tabil LG and Panigrahi S. Chemical treatments of natural fiber for use in
natural fiber-reinforced composites: A review. J Polym Environ 2007; 15: 25–33.
[99] Hughes M, Carpenter J and Hill C. Deformation and fracture behaviour of flax fibre
reinforced thermosetting polymer matrix composites. J Mater Sci 2007; 42:
2499–2511.
[100] Koyuncu M, Karahan M, Karahan N, et al. Static and dynamic mechanical
properties of cotton/epoxy green composites. Fibres Text East Eur 2016; 24: 118.
[101] Ray D, Sarkar BK, Rana AK, et al. Effect of alkali treated jute fibers on composite
properties. Bull Mater Sci 2001; 24: 129.
[102] Garcia-Jaldon C, Dupeyre D and Vignon M. Fibres from semi-retted hemp bundles
by steam explosion treatment. Biomass Bioenergy 1998; 14: 251–260.
[103] Rodriguez ES, Stefani PM and Vazquez A. Effects of fibers’ alkali treatment on the
resin transfer molding processing and mechanical properties of jute–vinylester com-
posites. J Compos Mater 2007; 41: 1729–1741.
[104] Pothan LA, George CN, Jacob M, et al. Effect of chemical modification on the
mechanical and electrical properties of banana fiber polyester composites. J Compos
Mater 2007; 41: 2371–2386.
[105] Weyenberg I Van de, Truong TC, Vangrimde B, et al. Improving the properties of UD
flax fibre reinforced composites by applying an alkaline fibre treatment. Compos Part
A Appl Sci Manuf 2006; 37: 1368–1376.
[106] Wang B, Panigrahi S, Tabil L, et al. Pre-treatment of flax fibers for use in rotationally
molded biocomposites. J Reinf Plast Compos 2007; 26: 447–463.
[107] Bledzki A, Mamun A, Lucka-Gabor M, et al. The effects of acetylation on
properties of flax fibre and its polypropylene composites. Express Polym Lett 2008;
2: 413–422.
[108] Jabbar A, Militký J, Madhukar Kale B, et al. Modeling and analysis of the creep
behavior of jute/green epoxy composites incorporated with chemically treated pulver-
ized nano/micro jute fibers. Ind Crops Prod 2016; 84: 230–240.
[109] Manikandan Nair K, Thomas S and Groeninckx G. Thermal and dynamic mechanical
analysis of short sisal fibre reinforced polystyrene composites. Compos Sci Technol
2001; 61: 2519.
[110] Mishra S, Mohanty AKA, Drzal LTL, et al. Studies on mechanical performance of
biofibre/glass reinforced polyester hybrid composites. Compos Sci Technol 2003; 63:
1377–1385.
[111] Oladele IO, Omotoyinbo JA and Adewara JOT. Investigating the effect of chem-
ical treatment on the constituents and tensile properties of sisal fibre. Journal of
Minerals and Materials Characterization and Engineering 2010; 9: 569–582.
2182 Journal of Industrial Textiles 47(8)

[112] Mwaikambo LY and Ansell MP. Chemical modification of hemp, sisal, jute, and
kapok fibers by alkalization. J Appl Polym Sci 2002; 84: 2222–2234.
[113] Chanakan A, Charoenvaisarocha and Hirunlabh Jongjit KJ. Materials and mechan-
ical properties of pretreated coir-based green composites. Compos Part B Eng 2009;
40: 633–637.
[114] Athijayamani A, Thiruchitrambalam M, Natarajan U, et al. Effect of moisture
absorption on the mechanical properties of randomly oriented natural fibers/polyester
hybrid composite. Mater Sci Eng A 2009; 517: 344–353.
[115] Patel B, Acharya S and Mishra D. Environmental effect of water absorption and
flexural strength of red mud filled jute fiber/polymer composite. Int J Eng Sci
Technol 2013; 4: 49–59.
[116] Mohanty AK, Misra M and Drzal LT. Surface modifications of natural fibers and
performance of the resulting biocomposites: An overview. Compos Interfaces 2001; 8:
313–343.
[117] Baiardo M, Frisoni G, Scandola M, et al. Surface chemical modification of natural
cellulose fibers. J Appl Polym Sci 2002; 83: 38–45.
[118] Gronqvist S, Viikari L, Niku-Paavola M-L, et al. Oxidation of milled wood lignin with
laccase, tyrosinase and horseradish peroxidase. J Appl Microbiol Biotechnol 2005; 67:
489–494.
[119] Mokoena MA, Djoković V and Luyt AS. Composites of linear low density polyethyl-
ene and short sisal fibres: The effects of peroxide treatment. J Mater Sci 2004; 39:
3403–3412.
[120] Hill C and Cetin N. Surface activation of wood for graft polymerisation. Int J Adhes
Adhes 2000; 20: 71–76.
[121] Hill CA, Cetin NS, Quinney RF, et al. An investigation of the potential for chemical
modification and subsequent polymeric grafting as a means of protecting wood
against photodegradation. Polym Degrad Stab 2001; 72: 133–139.
[122] Grelier S, Castellan A, Desrousseaux S, et al. Attempt to protect wood colour against
UV/visible light by using antioxidants bearing isocyanate groups and grafted to the
material with microwave. Holzforschung Int J Biol Chem Phys Technol Wood 1997; 51:
511–518.
[123] Loria-Bastarrachea M, Carrillo-Escalante H and Aguilar-Vega M. Grafting of poly
(acrylic acid) onto cellulosic microfibers and continous cellulose filaments and char-
acterization. J Appl Polym Sci 2002; 83: 386–393.
[124] Hornof V, Kokta B and Valade J. The xanthate method of grafting. IV. Grafting of
acrylonitrile onto high-yield pulp. J Appl Polym Sci 1976; 20: 1543–1554.
[125] Da Cunha C, Deffieux A and Fontanille M. Synthesis and polymerization of lignin
macromonomers. II. Effects of lignin fragments on methyl methacrylate. J Appl Polym
Sci 1992; 44: 1212.
[126] Hornof V, Kokta B and Valade J. The xanthate method of grafting. III. Effect of
lignin content on the graftability of wood pulp. J Appl Polym Sci 1975; 19: 1573–1584.
[127] Zheng G, He S, Qi Q, et al. Fe2þ-H2O2 initiated grafting of lignocellulose with methyl
methacrylate and its mechanism. Pure Appl Chem 1995; 32: 287–299.
[128] Zheng G-Z, Zhao B-A, He S-J, et al. Initiation of graft copolymerization by direct
oxidation of lignocellulose with KMnO4 and its mechanism. Pure Appl Chem 1995; 32:
1281–1292.
Ali et al. 2183

[129] He L, Gao J and He S. Graft copolymerization of acrylonitrile onto SGW using


KMnO4 as an initiator. Cellul Chem Technol 1992; 26: 497–503.
[130] Marchetti V, Ghanbaja J, Gerardin P, et al. Localisation and characterization by
TEM and EELS of manganese species during graft copolymerization of acrylic acid
onto sawdust using KMnO4 as initiator. Holzforshung 2000; 54: 553–556.
[131] Ghosh P, Ganguly PK and Bhaduri SK. Graft copolymerization of acrylonitrile on
jute using aqueous Cu2þ IO4 combination as initiator: Some key points and mech-
anism. Eur Polym J 1994; 30: 749–756.
[132] Gtridhar J, Kishore and Rao R. Moisture absorption characteristics of natural fibre
composites. J Reinf Plast Compos 1986; 5: 141–150.
[133] Saegusa T and Oda R. Copolymerization of allylidene diacetate and styrene, and the
chemical reactions of the saponified copolymer. Bull Inst Chem Res Kyoto Univ 1956;
34: 56–64.
[134] Kaith BS, Singha AS and Kalia S. Grafting MMA onto flax under the influence of
microwave radiation and the use of flax-g-poly (MMA) in preparing PF composites.
Autex Res J 2007; 7: 119–129.
[135] Ruggiero R, Machado AE, Hoareau W, et al. Photodegradation of sugarcane bagasse
fibers. influence of acetylation or grafting UV-absorber and/or hindered nitroxide
radical on their photostability. J Braz Chem Soc 2006; 17: 763–770.
[136] Ali A, Shaker K, Nawab Y, et al. Impact of hydrophobic treatment of jute on mois-
ture regain and mechanical properties of composite material. J Reinf Plast Compos
2015; 34: 2059–2068.

You might also like