ArticlePDF Available

Towards widespread properties of cellulosic fibers composites: A comprehensive review

Authors:

Abstract

To meet the growing demand for the use of environmentally and friendly materials in various applications, many efforts have been done in order to provide lighweighting, biodegradable, and renewable composites leading to a diminution of greenhouse gas emissions. However, they are confronting some challenges that may limit their utilization in industries such as poor interfacial adherence, high moisture absorption, low processing temperature, and impact strength. Several research works have been carried out regarding the mechanical, thermal, and hydric properties of natural fibers reinforced composites to deal with above underlined drawbacks. Actually, based on our analysis, there are two relevant concerns that prevent their integration in other industries such as marine and renewable energy sectors. In this context, the current critical review offers an up to date overview of the factors influencing the properties of the different components of natural fibers composites (fiber, matrix, fiber/matrix interface) as well as their physicochemical, mechanical, thermal and hydric behaviors. Recent works on the mechanical properties of nanobiocomposites based on cellulosic fibers, crystals, and nano-clay fillers are reviewed. The future directions are also discussed to overcome the challenges confronted by these materials, that is, fire resistance and moisture absorption, in order to widespread their utilization especially for wet and fire retardant purposes.
Original article
Journal of Reinforced Plastics and
Composites
2022, Vol. 0(0) 142
© The Author(s) 2022
Article reuse guidelines:
sagepub.com/journals-permissions
DOI: 10.1177/07316844221112974
journals.sagepub.com/home/jrp
Towards widespread properties of cellulosic
bers composites: A comprehensive review
Mouad Chakkour
1
, Mohamed Ould Moussa
1
, Ismail Khay
1
,
Mohamed Balli
1
and Tarak Ben Zineb
2
Abstract
To meet the growing demand for the use of environmentally and friendly materials in various applications, many efforts have
been done in order to provide lighweighting, biodegradable, and renewable composites leading to a diminution of
greenhouse gas emissions. However, they are confronting some challenges that may limit their utilization in industries such
as poor interfacial adherence, high moisture absorption, low processing temperature, and impact strength. Several re-
search works have been carried out regarding the mechanical, thermal, and hydric properties of natural bers reinforced
composites to deal with above underlined drawbacks. Actually, based on our analysis, there are two relevant concerns that
prevent their integration in other industries such as marine and renewable energy sectors. In this context, the current
critical review offers an up to date overview of the factors inuencing the properties of the different components of natural
bers composites (ber, matrix, ber/matrix interface) as well as their physicochemical, mechanical, thermal and hydric
behaviors. Recent works on the mechanical properties of nanobiocomposites based on cellulosic bers, crystals, and nano-
clay llers are reviewed. The future directions are also discussed to overcome the challenges confronted by these materials,
that is, re resistance and moisture absorption, in order to widespread their utilization especially for wet and re retardant
purposes.
Keywords
Cellulosic bers, natural ber composites, environmental aging, modeling
Introduction
Nowadays, the environmental consciousness has increased
and the use of ecofriendly materials that combine several
functional properties such as lightness and strength may be
encouraged.
1,2
Actually, the utilization of renewable and
lightweighting materials orient research activities toward
natural bers that can reinforce polymers, bio-resins as well
as ceramics.
1,3
Indeed, natural bers are characterized by
their important mechanical properties, low density, low cost,
biodegradability, renewability, and low environmental
impact.
1,2,4
These make them among the best candidates to
replace non-renewable and expensive synthetic bers such
as glass, carbon, and kevlar bers.
4,5
However, the low
temperature processing and the incompatibility with poly-
mers limit the utilization of natural bers in industrial ap-
plications.
1,6
Currently, they are subdivided into three
categories based on their chemical compositions: Vegetal,
animal, and mineral bers, where the latter are the most used
due to their abundance,
7
low cost
6
as well as attractive
mechanical properties.
6
Besides, the properties of renewable natural bers de-
pend on several parameters such as: local climate (moisture,
temperature, and ultraviolet light), morphology, chemical
composition and extraction processes. Indeed, they are
highly inuenced by the amount of absorbed water and sun
exposition during the plants growth in addition to the age of
plants.
8,9
These factors determine the polysaccharides
contents within plants especially the cellulose content and
its crystallinity.
6,8
Several characterizations could be used to
better understand the thermal, mechanical and physico-
chemical behaviors.
10,11
In fact, the morphology is one of the factors that in-
uence the properties of natural bers. The decrease of
mechanical properties is related to the increase of the mi-
crobrilar angle (MFA) of bers.
6
Bourmand et al.
6
reported
1
AMEEC team, Laboratory of Renewable Energies and Advanced Materials
LERMA, College of Engineering and Architecture, International
University of Rabat, Morocco
2
Laboratory for the Study of Microstructures and Mechanics of Materials
LEM3, University of Lorraine, France
Corresponding author:
Mohamed Ould Moussa, AMEEC team, Laboratory of Renewable Energies
and Advanced Materials LERMA, College of Engineering and
Architecture, International University of Rabat, Universit´
e Internationale
de Rabat Techn, sala al jadida 1100 AQ10, Morocco.
Email: mouad.chakkour@uir.ac.ma
that the ax and hemp bers that exhibit high mechanical
properties have a small MFA comprised between and 11°,
while sisal bers displaying a low stiffness and strength
properties have a large MFA of about 20°. Likewise, the
local environment of natural bers can also inuence their
mechanical properties because of their high hydrophility.
Effectively, Rowell
12
studied the moisture adsorption be-
havior of some lignocellulosic bers under 21°C and rel-
ative humidity RH of about 65%. They found that their
moisture adsorption ranges from 7% to 15%. Regarding
thermal resistance, the processing temperature of natural
bers is limited due to the partial and complete degradation
of their chemical components at a certain temperature.
Actually, several studies
13,14
reported that the degradation
temperature of the most known natural bers usually varies
between 215°C and 240°C. Concerning extraction pro-
cesses, the production of ne and strong bers is technically
difcult
15
and dramatically depends on the used extraction
methods. For this occasion, Parnia zakikhani et al.
5
reported
the effect of different extraction methods on the tensile
properties of lignocellulosic bamboo bers. They found that
the highest value is obtained for bers extracted by com-
bining chemical and compression methods, while the lowest
one is exhibited by bers extracted by using mechanical
rolling method. However, the roughness of the pure ex-
tracted bers is small which make them incompatible with
considered matrices. Accordingly, different treatments ei-
ther chemical (NaOH, stearic acid, silane, and peroxide) or
physical (plasma, ozone, and corona) can be used in order to
enhance the compatibility and the interfacial adherence
between bers and matrix by removing impurities and some
amorphous components present at the bers surface.
1,5,16,17
In this case, Chin SC et al.
18
investigated the tensile strength
of lignocellulosic bers, and, concluded that the tensile
strength of bers treated with high NaOH concentrations
decreases when compared to pure ones, which is due to the
degradation of crystalline cellulose caused by the solution.
This review will provide a map for selecting the type of
bers based on their physicochemical properties, cost, and
applications needs.
Besides, natural bers are particularly used to reinforce
petrochemical-based resins including thermosets, thermo-
plastics and employed bio-resins matrices in bio-
composites.
8,9
Usually, petrochemical matrices are favored
to bio-resins due to their good mechanical properties,
1
abundance,
19
and low cost (approximately 10% lower
than bio-resins).
19
Regarding thermoplastic polymers, the
macromolecular chains are linked together with weaker
intermolecular forces such as van der Waals and hydrogen
bonds.
20
These polymers can change reversibly from vis-
cous to solid states under cooling and heating, respectively.
The advantages of utilizing thermoplastic polymers are their
important fracture resistance, high strain under failure, and
impact strength.
20
Also, their recyclability enables us to
reduce the waste amount. Polypropylene is one of the most
used thermoplastics because of its good mechanical prop-
erties (tensile and exural strengths), small porosity, and
high dimensional stability which result in the enhancement
of the water absorption property.
21
The use of polyethylene
and polycarbonate are recommended because of their good
thermal and mechanical properties. Due to its excellent re
and chemical resistance as well as low cost and good du-
rability, the polyvinyl chloride becomes the most appro-
priate material in building sector.
21
However,
macromolecular chains of thermosets matrices are linked by
hydrogen elements which are removed irreversibly upon
heating.
20
They are suitable for applications where the high
thermal stability and chemical reactions are needed.
20
In
addition, they exhibit a small viscosity at the initial stage of
processing because of their small molecular weight re-
sulting in a good wettability of bers. Epoxy and polyester
are the most used thermosets because of their excellent
thermal, mechanical and chemical resistances in addition to
their high dimensional stability. Indeed, their hydrophobic
structure makes them the best candidate for composite
laminates and advanced composite surfaces.
21
Many studies have been carried out to investigate the
multiphysical properties of natural bers reinforced com-
posites. Several researchers studied the effect of different
chemical treatments, bers fraction, orientation and length
on the mechanical, hydric, and thermal properties. Most of
studies reported that the augmentation of bers content
increases the void content in composites. The optimal
mechanical properties are generally obtained for 30 wt%
longitudinal bers loading composites.
22,23,24,25
Regarding
hydric properties, the equilibrium moisture content and
absorption rate was found to increase while increasing the
bers content. These results are explained by the important
void content within composites. The mechanical properties
are highly inuenced by the moisture absorption which
induces swelling of bers, leads to the development of shear
stresses at the interface and nally to the debonding of bers
and matrix.
26
Some solutions presented in the literature such
as chemical treatments enhance the mechanical properties,
reduce partially the hydrophility, and moisture absorption of
biocomposites. However, they are still not sufcient to
address their use for wet applications. Other researchers
focused on the thermal properties of natural bers com-
posites. Most of them reported that the thermal conductivity
rises when the ber content increases due to the increase of
the porosity within composites that affects the capability of
heat transfer.
27,28
From modeling point of view, several works used nite
elements method FEM and analytical models to predict the
mechanical behavior of natural bers reinforced composites
(NFRCs). In this way, Palombini et al.
29
used FEM to
investigate the effect of the different structural components
of the bamboo. Cao et al.
30
used some analytical models
such as Halpin-Tsai and the rule of mixture (ROM) to study
the mechanical properties of wood/ax reinforced
2Journal of Reinforced Plastics and Composites 0(0)
polypropylene composites. Potluri et al.
31
studied also the
mechanical properties of pineapple bers reinforced epoxy
composites using analytical models.
As early reported, natural bers are incompatible with
hydrophobic polymers due to their hydrophilicity. To en-
hance the bers-matrix interface, the incorporation of
nanollers such as nanoclays, SiO
2
,TiO
2
, and graphene
remains an efcient solution to improve the mechanical,
physicochemical, optical, and thermal properties of bio-
composites.
20
Nanoclays llers have attracted the attention
of researchers due to their great aspect ratio, low cost, and
wide availability
32
in addition to their excellent mechanical
and re retardancy behaviors. Preliminary investigations
reveal that the addition of nano-additives reduces the water
absorption while improving the stress-transfer between
matrix and bers, until a critical ller content where the
mechanical properties drop due to the agglomeration of
llers. Indeed, Haq et al.
33
have studied the effect of 1.5
wt.% of Cloisite 30B nano-additives on the mechanical
properties of hemp reinforced polyester hybrid composites.
They particularly reported that the tensile modulus and
strength increase by 6% and 20%, respectively. Arulmur-
ugan et al.
34
investigated the inuence of montmorillonite
nano-clays contents on the mechanical properties of jute
bers reinforced polyester nanocomposites and reported
that the optimal mechanical and vibration properties are
obtained for 5 wt.% nano-clays contents. Upon 5 wt.%
nano-clays content, the agglomeration of llers prevents
a better load transfer between the composite components
weakening mechanical properties. However, few studies
have addressed the inuence of nanollers on the water
absorption and re retardancy behaviors of biocomposites.
In general, the adoption of these renewable bers in
industries would be of great interest due to their lighweight,
low cost, and non-toxicity.
35,36
Currently, automotive in-
dustry is leader in the use of natural bers reinforced
composites (NFRCs) especially German companies
including Mercedes, Volkswagen, Audi, BMW as well as
Opel that encourage the utilization of natural bers in
several interior applications such as decking, window
frames, oor trays, and the headliners.
1,37
In addition,
natural bers are used in building, medical, and packaging
applications.
38
This current review paper aims to provide in a critical
way, the whole picture of natural bers reinforced polymer
composites NFRCs and help to understand the issues for
their widespread integration into more applications mainly
for wet and re-retardant targets. It is organized into dif-
ferent sections, as shown in the Figure 1. A detailed pre-
sentation of the factors inuencing the different components
of NFRCs (ber, matrix, ber/matrix interface) as well as
the most used characterization techniques are provided in
the rst part, then, the mechanical, hydric, and thermal
properties of biocomposites are reviewed while discussing
the encountered challenges. Afterward, progress works
regarding the mechanical properties of nanobiocomposites
are reported. Finally, applications and the future trends and
directions will be reported and discussed.
Composition of natural bers reinforced
composites NFRCs
Natural bers
In recent years, natural bers have been widely used to
reinforce composite materials
1
with various origins such as
vegetal, animal, or mineral. Considered as renewable re-
sources, they exhibit attractive lightweight, mechanical, and
economic properties (see Table 1), which make them serious
alternatives to the expensive and non-renewable synthetic
bers (glass, carbon, and kevlar bers) at least for soft
structure applications.
39,40
Regarding vegetal bers, they are namely composed of
sugar-based polymer components (cellulose and hemicellulose)
Figure 1. Graphical summary of the review paper.
Chakkour et al. 3
while protein is the main component of the animal ones.
8
The former can be categorized into eight groups based on
the nature of plants and plant part, they are extracted from
bast bers (jute, ax, hemp, banana) from the skin of the
plants stems; leaf bers (banana, pineapple) from leaves
plants; seed bers (cotton, coir) from seeds; grass bers
(bamboo, rice) and core bers (corn, wheat stalk) from
stalks; wood bers; foot bers (cassava); and fruit bers
(banana, coir).
8,19,41
Hence, according to above-mentioned
categories (wood and non-wood bers)
1
they can be dis-
tinguished. The non-wood bers are more lightweight and
show indeed high specic strength, high specic modulus,
exibility during processing, fatigue and corrosion re-
sistance, world-wide abundance,
42
and higher cellulose
concentration compared to wood bers.
9
In addition, there
are three types of animal bers: Animal hairs bers, avian
bers, and silk bers (for instance, spider silks). Bast bers
(especially Flax) show rapid absorption and desorption of
water, low thermal conductivity, and the high acoustic in-
sulation.
1
The leaf bers (like abaca bers) present high
mechanical strength. Seed bers are characterized by their
water absorption, strength, and concentration of lignin in
comparison to other natural bers. Bamboo bers (grass
bers) are known by their excellent durability and good
properties such as stability, tenacity, exibility, ultraviolet
radiation resistance, and small density.
1
Furthermore, sheep
animal bers are known by their good thermal and acoustic
insulation, high deformability, and durability similarly as
the lightweight feathers bers. Regarding mineral bers,
they exhibit a good re resistance.
19,41,43,44
Figure 2 summarizes the natural bers classication il-
lustrated by some images.
However, the widespread utilization of the natural bers
remains confronted to certain limitations
9,45
such as the high
moisture absorption, low compatibility with conventional
resins and low homogeneity compared to synthetic bers
(glass, carbon, and kevlar bers).
Nowadays, the vegetal lignocellulosic bers are usually
encountered as reinforcement for composites in several
applications such as automotive, aeronautic, military,
building, and sport equipments.
9,47
This is due to their
mechanical performances compared to animal bers. Cur-
rently, known as sugar-based plant bers, lignocellulosic
bers include mainly polysaccharides which are carbohy-
drates and associate several monosaccharides made of
carbon, oxygen, and hydrogen.
9
They are bonded by hy-
drogen links via the relevant hydroxyl groups in the
structure.
48
In fact, starch, cellulose, hemicellulose, and
proteins are the most widely usual polysaccharides. Sugar-
based plant bers are composed mainly of cellulose,
hemicellulose, and lignin with some amount of pectines,
waxes, and impurities. Cellulose is considered as a rigid
armature surrounded by amorphous areas consisting namely
of lignin and hemicellulose with the other minor con-
stituents. It is reported that the properties of the vegetal
bers are affected by the amount and the crystallinity of
cellulose, the polysaccharide composition of cell walls and
the microbril angle which corresponds to the orientation of
cellulosic microbrils with respect to the ber axis.
6
These
key parameters are highly sensitive to the age, the plants
type, and the climate conditions (amount of sun radiation
and absorbed water during the plant growth).
6,9
According
to this particular microstructure of the vegetal bers, it is
worth in the following sections to introduce more details
about the mentioned chemical components:
Cellulose
It belongs to the family of periodic homopolysaccharides and
then appears as a sequence of linear macromolecules in which
the monomer is the glucose elements including repetitive
motif called cellubiose that is composed of two glucose
Table 1. Advantages and disadvantages of natural bers.
Advantages of natural
bers
Disadvantages of natural
bers References
Low cost High moisture absorption
1
Low density Variation in bers quality
19
Renewable Low maximum processing
temperature
19
Biodegradable Hydrophilic structure
1,6
Non-toxic Discontinued bers
1,6
Figure 2. Classication of natural bers with respect of their origin
1,46
(Google images reviewed on 2021).
4Journal of Reinforced Plastics and Composites 0(0)
molecules and six hydroxyl groups per monomer
5
(see
Figure 3). These hydroxyl groups ensure intra and in-
termolecular hydrogen bonds with other macromolecules
leading to crystalline arrangement of the cellulose macro-
molecules into microstructures called microrbils.
49
The latter
are embedded in amorphous cellulose matrix showing small
packing density.
50
Moreover, the polymerization degree of
cellulose is 10100 higher than the hemicellulose one.
5
Hemicellulose
Hemicellulose is the second most widely organic material
found within the bers in terms of quantities. It comprises
heterogeneous polysaccharides of xylopyranose, glucopyr-
anose, and mannopyranose chains.
51
Regarding the moisture
absorption, it is reported that hemicellulose components are
approximately 2.6 higher than the lignin ones.
9
Lignin
Lignin is a 3D polymer composed of aromatic units strongly
intramolecular bonded and lling the space between cel-
lulose, hemicellulose, and pectins.
52
In addition, lignin acts
as a coupling agent between cellulose and hemicellulose
increasing the stiffness of bers.
53
It is reported that the
specic tensile strength of plant bers ranges between 1600
and 2950 MPa while Youngs modulus is comprised be-
tween 10 and 100 GPa.
53
The Youngs modulus of cellulose
is 140 GPa, and that for hemicellulose is equal to 8 GPa. The
evolution of these parameters is strongly inuenced by the
temperature and the humidity. In fact, the presence of an
important quantity of lignin improves stiffness and rigidity
of the lignocellulosic bers.
1,6
Figure 4 gives a sketch of the
morphology of a lignocellulosic bamboo ber.
5
Nowadays, the majority of NF is produced in Asian
countries especially China, India, and Vietnam while small
quantities can be produced in South America (Brazil) and
Africa (Nigeria and South Africa)
1
(see Table 2). Abaca
bers are species of banana that are cultivated as a com-
mercial crop in Asia, especially in the Philippines. They
exhibit an important tensile strength lying between 600 MPa
and 900 MPa and Youngs modulus ranging from 30 GPa to
50 GPa,
57
making them very promising for industrial ap-
plications. Sugarcane (Bagasse) attracts the attention of
researchers and scientists because of its biodegradability
and abundance. It is widely used for making furniture,
insulating boards, and lter muds.
55
Bamboo bers are
abundant in humid equatorial zones, especially in Africa,
Asia, and South America. They are mainly used in
construction, agriculture, and wood-based panels.
55,58
Ba-
nana is obtained from the bast of the plant Musa sapientum
Linn. The high cellulose content and low microbrillar
angle MFA indicate that banana bers could be a good
reinforcement for hydrophobic polymer matrices.
55,59
Co-
conut palm is a member of the palm family that grows in the
Southeast Asia (Malaysia, Indonesia, and Philippines) and
tropical islands such as Seychelles, Mauritius, Polynesia,
Comoro, and Maldives. Date palm bers are found in the
Middle East, Northern Africa, India, United States (Cal-
ifornia), and Pakistan. Flax is one of the most used natural
bers because of its important specic mechanical prop-
erties which are comparable to those of glass bers. Canada
has been the world largest producer and exporter of ax
since 1994. The latter are widely used as reinforcements for
composites in several industrial applications, including
automotive interiors.
55,60
Hemp (Cannabis Sativa) is a plant
native to Philippines and China. The corresponding bers
are extensively used in textiles and automotive compo-
nents.
55,61
Jute is one of the cheapest natural bers that
grows in China, India, and Bangladesh. It is known as the
golden berbecause of its biodegradability, excellent
thermal and electric insulations.
8
Pineapple is a short
tropical plant where related bers are inexpensive and
exhibit a high specic strength and stiffness.
55,62
Sisal is one
of the widely used natural bers because of its good me-
chanical and acoustic properties as well as easy cultivation.
The main producers of sisal bers are Tanzania and
Figure 3. Monomer unity of cellulose polymer.
1,6
Figure 4. (a) Vascular bundle of bamboo, (b) elementary ber
1020 μm, (c) nano bril 110 μm involves lignin, and
hemicellulose.
5,54
Reprinted from Materials and Design 63, by
Parnia Zakikhani, R. Zahari, M.T.H. Sultan, D.L. Majid, Extraction
and Preparation of bamboo ber-reinforced composites, pp 820
828, Copyright 2021, with permission from Elsevier, (License
number: 5,118,180,687,934).
Chakkour et al. 5
Brazilia.
55,63
Aliakbar et al.
1
report that the three bers most
produced cellulosic are bagasse, bamboo, and cotton bers.
The annual production reaches 75, 30, and 25 millions of
tons, respectively.
1
However, the effective cost of bamboo
and alfa bers are the lowest among all lignocellulosic -
bers.
6
Actually, the bamboo price ranges between 0.3 and
0.5 $/kg which is 2.89.3 times cheaper than glass bers.
Some of natural bers such as sisal and hemp are more
expensive than others due to their quality, neness, and
mechanical performance.
6
In this way, it is worth to
highlight that uctuations of both supply and demand
dramatically affect the bers price.
6
In addition to economic aspects, the mechanical prop-
erties of natural bers are vital in their selection as
Table 2. Commercial prices
6
and world production
7,8,19,55,56
of the most used cellulosic bers.
Fiber nature Producer Cost ($/Kg)
World production
(millions of tons) References
Abaca Philippines, Ecuador 0.451.3 0.07
6,19
Sisal Brazil, Kenya, Tanzania, China, Cuba, Haiti, Madagascar, India 0.378
6,8
Kenaf India, China, Malaysia, USA, Thailand, Vietnam 0.40.7 0.97
6,7
Jute India, Nepal, Bangladesh 0.51.15 3.45
6,8
Hemp China, France, Germany, UK 0.71.75 0.214
6,19
Bamboo China, Japan, India, Chile, Ecuador, Nigeria, Philippines, Pakistan 0.30.5 30
6,56
Cotton China, Brazil, India, Pakistan, Turkey 0.62.25 25
6,56
Elephant grass ——Abundant
55
Broom ——Abundant
55
Oil palm fruit ——Abundant
55
Wood 0.10.65
6
Ramie Tropical Asia 0.51.2 0.28
1,6
Flax China, Brazil, Philippines, India 1.752.7 0.83
1,6
Coir India, Thailand, Vietnam, Philippines, Indonesia 0.40.75 1.2
1,6
Alfa 0.20.4
6
Table 3. Averaged mechanical and physical properties of the most used natural bers.
55,46,64,65,66,67
Fiber
Diameter
(μm)
Tensile
strength
(MPa)
Tensile
modulus
(GPa)
Elongation
(%)
Density
(g/cm3)
Specic tensile
strength
(MPa.cm3.g1)
Specic tensile
modulus
(GPa.cm3.g1)
Jute 393860 1360 1.51.8 1.441.52 258.5597.2 8.541.6
Flax 3451500 27.690 2.73.2 1.421.52 226.91056.3 18.263.4
Hemp 550920 5570 241.471.52 361.8625.8 36.247.6
Bamboo 140800 1132 2.53.7 0.6 1.1 127.31333.3 1053.3
Bagasse 200400 220290 1727.1 1.1 1.25 177.6232 13.621.7
Feather 100203 310 6.9 0.9 111.1225.5 3.3311.1
Sisal 50300 468790 9.425 2 7 1.41.45 322.75564.3 6.5 17.2
Silk 1001500 5 25 15 60 1.3 76.91153.8 3.819.2
Wool 50315 2.35 13.235 1.3 38.46242.3 1.773.85
Abaca leaf
ber
114130 418486 1213.8 - 0.83 503.61585.54 14.4516.62
Alfa 35 22 5.8 0.89 39.32 24.71
Banana 80250 529759 8.20 13.5 1.35 391.85562.22 6.07
Cotton 287597 5.512.6 310 1.6 179.37373.12 3.437.87
Coir 100460 108252 461540 1.15 93.91219.13 3.47 5.21
Curaua 170 158729 5 1.4 112.85 520.71
Kenaf 81 250 4.3 1.4 178.57 3.07
Oil palm 150500 80248 0.53.2 17 25 0.71.55 51.62354.28 0.32 2.06
Piassava 134143 1.074.59 7.8 21.9 1.4 95.71102.14 1.213.27
Pineapple
leaf
2080 4131627 34.582.5 0.81 1.3 317.69
1251.53
26.5363.46
Ramie 2080 4001000 24.5128 1.2 411.55 2581000 15.8128
6Journal of Reinforced Plastics and Composites 0(0)
reinforcement for composites. Several works have reported
such issues and analyzed the parameters that inuence the
quality of the selected bers.
5
The Table 3 shows the av-
eraged mechanical properties of the most widely used
natural bers.
Usually, the specic mechanical properties of natural
bers are generally important due to their small density. In
addition, one notes that hemp, jute, ax, and bamboo exhibit
tensile strength and Youngs modulus similar to those of
E-glass bers that exhibit about 480600 MPa.g/cm
3
and
28 GPa, respectively
68
(see Table 3). Sometimes, bamboo
bers are considered as glass bers due to their morphologic
longitudinal alignment.
Parameters inuencing the mechanical properties of natural
bers. The reported literature issues
6,9
have revealed re-
cently that the multiphysical properties of natural bers
depend on several parameters such as extraction methods,
chemical composition, processing temperature, and con-
ditions of plants growth. Figure 5 shows the general factors
that inuence the properties of bers going from the plant
growth to the supplying stage. More details are highlighted
in following sections.
Chemical composition. Plant-based bers are composed
mainly of cellulose, hemicellulose, lignin, and pectines as
mentioned early, in addition to waxes, fats and other in-
organic components which represent a small percentage of
bers mass. Actually, the chemical composition varies with
the plant type, the growth conditions and the amount of
absorbed water and sun exposition. These factors are re-
sponsible especially for the change of cellulose amount and
crystallinity.
8
The latter represents the capacity of the cel-
lulose OH groups to form high-density hydrogen bonds
leading to compacted cellulosic chains. Hence, it is reported
that the crystalline cellulose shows higher Youngs modulus
that reaches 136 GPa and only 75 GPa for synthetic glass
bers.
6
Table 4 shows chemical compositions of the most
used plant-based bers.
Usually, more natural bers contain cellulosic elements
more their mechanical properties increase except for the
cotton bers.
6
Gurunathan et al.
9
showed that the specic
tensile strength of the plant-based bers falls between
1600 MPa and 2950 MPa, while the specic tensile
modulus is comprised between 10 GPa and 130 GPa.
Besides, Cousins et al.
53
found that the Youngs modulus of
cellulose is almost 140 GPa whereas that of hemicellulose is
estimated to be only 8 GPa, provided, these values may
change with respect of the relative humidity conditions.
53
Further, Bledzki et al.
74
reported that the Youngs modulus
of crystalline cellulose lies between 134 and 136 GPa, while
that of the hemicellulose is about 8 GPa in dry atmosphere
(about 12% of relative humidity) and only 0.2 GPa at
ambient atmosphere.
75
Moreover, lignin that is considered
a compatibilizer or coupling agent between cellulose and
hemicellulose contributes to increase stiffness properties of
plant-based bers.
8
According to above underlined works, it seems clearly
that crystalline cellulose and lignin dramatically contribute
mortally to enhance the mechanical behavior of vegetal
bers.
6
Morphology. Morphological aspects including micro-
brillar angle (MFA) and dimensions of bers strongly affect
the mechanical properties of lignocellulosic bers.
76
In fact,
the microbrillar angle MFA characterizes the orientation of
cellulosic microbrils in the secondary plants wall with
respect of the longitudinal direction.
6
In literature, several
authors used X-Ray diffraction method to measure mi-
crobrillar angle which may change during tensile tests.
Wang et al.
77
showed that during tensile tests, cellulose
microbrils are realigned allowing a reduction in micro-
brillar angle. Trivaudey and Placet
78
investigate the effect of
tensile loading on the evolution of microbrillar angle of
hemp bers. They found that MFA decreased from 11° to
7.2° under a loading amplitude of about 0.5 N. Eder et al.
79
investigated the evolution of tensile stiffness of wood
samples and different extracted bers with respect of mi-
crobrillar angle MFA. They found that tensile stiffness of
wood samples varies between 8 GPa and 56 GPa, while their
correspondent MFA ranges between 10° and 40°. However,
stiffness does not exceed 30 GPa for MFA between and
15° in the case of 65% of dried wood bers obtained by
chemical extraction, while the lower Youngs modulus is
obtained for bers extracted by other processes exhibiting
higher microbrillar angles.
Regarding the bers dimensions, Humphrey et al.
80
investigated the effect of aspect ratio of coconut, ba-
gasse, and oil palm bers on the compressive and tensile
strengths of soil blocks. They opted for the following ex-
perimental protocol of block elaborating that exists to
maintain the maximum moisture content at 18%, to keep the
bers immerged in water for 48 h before being beaten, use
a compound light microscope to determine the ber di-
ameter, adapt the ber length for considering 5 values of the
Figure 5. Factors inuencing the bers quality.
55
Chakkour et al. 7
aspect ratio between 25 and 125, mix bers with soil
provided a volume fraction of bers of about 1%, and
elaborate blocks with uniform section by imposing a con-
stant pressure of about 100 bars. They obtained that the
increase in the aspect ratio from 50 to 125 leaded to improve
the compressive strength by 26%, while the aspect ratio of
100 is the optimum value for blocks reinforced by oil palm
and bagasse bers showing an increase of 14% and 32%,
respectively. At these maximum ratios, the tensile strength
increased by 61%, 24%, and 20% for blocks reinforced by
coconut, oil palm, and bagasse bers, respectively. A.S.
Singha et al.
81
also investigated the effect of ber length on
the tensile and compressive properties of agave bers re-
inforced polystyrene composites. They used raw agave -
bers with 90 μm, short bers with 3 mm and 8 mm,
respectively, of lengths were used as continuous bers for
reinforcement. They concluded that reinforced composites
with particles exhibits excellent mechanical properties
where compared to long and short reinforcements.
Moisture adsorption. Hydrophilic features of cellulosic
bers result in high water and moisture absorption.
82
In fact,
the resulting swelling alters their mechanical properties and
leads to poor interfacial adherence and to the interface
debonding phenomenon accordingly within natural bers
reinforced composites when associated with hydrophobic
matrices.
Rowell
12
studied the equilibrium moisture content of the
widely used natural bers at operating temperature of 21°C
and relative humidity RH of 65%. He showed that Abaca
and Pina bers adsorbed almost 15% and 13%, re-
spectively. This corresponds to the highest moisture
contents whereas ax and bamboo bers adsorb 7% and
8.9%, respectively.
Such behavior can be explained by the fact that cellulosic
bers have a hydrophilic structure being responsible for the
high water and moisture adsorptions. Further, it is reported
that the hydrophilic structure of natural bers lead to
a swelling after water or moisture absorption and causes
poor interfacial adherence with hydrophobic matrices. This
results in poor mechanical properties of natural bers re-
inforced composites (NFRCs).
The equilibrium moisture content of the most widely
used natural bers at a temperature of 21°C and relative
humidity RH of 65% are reported by Rowell and sum-
marized in Table 5
12
(see Table 5). Accordingly, Abaca,
Pina bers adsorbed almost 15% and 13%, respectively,
which are the higher humidity contents whereas ax and
bamboo bers have adsorbed 7% and 8.9%, respectively.
Chung KF et al.
83
have studied the effect of water ab-
sorption of two types of bamboo specimens. Their length
was about 1200 mm and the external diameter was esti-
mated to be 25 mm. Specimens are immersed in water inside
an oven at 105°C for various time periods: 124812 h
and 1237 days. They found that the compressive and
bending strengths of bamboo have decreased by half and one-
third after 7 days of immersion. During the same experiment,
the Youngs modulus has decreased by one-third. Godbole
et al.
84
have investigated the effect of water absorption on the
mechanical properties of bamboo samples. In their works,
specimens are immersed in distilled water at an ambient
temperature of 20°C for 144 h. They absorb 81.2% of their
weight. In addition, their Youngs modulus and tensile
strength have decreased by 48% and 37%, respectively.
To deal with these drawbacks, several studies devoted to
investigate the effect of chemical treatments on the moisture
adsorption and the mechanical behavior of natural bers and
biocomposites. However, few researchers are focused on
the study of dimensional changes (swelling) of natural bers
or natural ber composites upon exposure to relative hu-
midity. It is very important to conduct research on this
aspect to constitute key data for the predictive modeling of
the hygromechanical behavior of biocomposites.
Table 4. Chemical composition of the most widely used lignocellulosic bers.
55,69,70,71
Fiber Cellulose (wt%) Hemicellulose (wt%) Lignin (wt%) Waxes (wt%) Pectins (wt%) References
Abaca 56 63 20 25 7 93 1
8,72,73
Coir 32 43 0.15 0.25 40 45 34
37,49
Cotton 85 90 5.7 0.7 1.6 0.6 0.1
37,49,70
Flax 71 18.6 20.6 2.2 1.5 2.3
37,49,72
Jute 61 71 14 20 12 13 0.5 0.2
37,49,70
Bamboo 26 43 30 21 31 ——
8,49
Hemp 68 15 10 0.8 0.9
37,49,69,72
Ramie 68.6 76.2 13 16 0.6 0.7 0.3 1.9
37,71,72
Sisal 67 78 10 14 8 11 2 10
37,49
Bagasse 55.2 16.8 25.3 ——
8,49
Agave 68.42 4.85 4.85 0.26
55
Banana 60 65 6 8510 ——
55
Borassus 53.40 29.6 17 ——
55
Kenaf 31 72 20.3 21.5 8 19 ——
55
Pineapple leaf 70 83 512.7 ——
55
8Journal of Reinforced Plastics and Composites 0(0)
Thermal resistance. The ammability of cellulosic bers
leads to the loss of stiffness and strength which limits their
utilization under high temperatures.
85
It is reported that the
thermal decomposition of bers depends on their chemical
composition.
85
Kozlowski et al.
85
reported that the increase
of crystalline cellulose within bers structures decreases the
rate of the thermal decomposition of lignocellulosic bers,
increasing then the re resistance of natural bers reinforced
composites.
Several authors observed that cellulose degrades in
a temperature range comprised between 300°C and 420°C,
while pectins and hemicelluloses between 250°C and
320°C
8688
Due to the complex chemical structure of lignin,
there is a permanent discussion about its thermal de-
composition. Some works showed that lignin breaks down at
200°C according to Ref.
89
and 400°C according to Ref.
90,91
while Yang et al.
88
demonstrated that lignin can be decom-
posed in a large temperature range ranging from 160°C up to
900°C. It can be concluded that the high content of amor-
phous structures as hemicelluloses and pectins within bers
decrease their re resistance. However, chemical treatments
usually utilized for natural bers can enhance their thermal
stability by removing amorphous parts. Ta ble 6 shows the
averaged decomposition temperature of the most used lig-
nocellulosic bers. Figure 6 illustrates a typical TGA curve of
non-treated Agave Americana bers.
Based on several studies, it seems that chemical treat-
ments enhance the thermal stability of bers by eliminating
a signicant amount of lignin and non-amorphous com-
ponents that degrade at low temperatures.
Extraction processes. Nowadays, it is well known
15
that
the production of ne and strength bers is now technically
difcult because of limited and expensive involved pro-
cesses. Currently, the common extraction methods are me-
chanical such as steam explosion, retting, crushing, rolling,
and/or chemical such as those using alkaline treatments.
Mechanical procedures
Steam explosion process
This process consists of cutting plant culms into different
specimens which are then placed in autoclave housing su-
perheated steam at a temperature of 175°C and a pressure
between 0.70.8MPafor60min.
5
This is cycled until cells
fracture. At the end of the process, the ash resulted is removed
by washing bers using water at 90°C, then, the latter are
dried in an oven at a temperature of 105°C for 24 h. As
a consequence, lignin is condensed on the bers surface
increasing their roughness. This negatively impacts the in-
terfacial adhesion between bers and matrix leading to the
decrease of mechanical properties of the resulting composite.
Retting method
There are two types of retting such as dew retting and water
retting.
11
The rst one consists in breaking down the cellular
tissues and the substances that surround the bers in the
presence of bacteria, sunlight and dew. This method is
widely used in the industrial production of blast bers.
However, in water retting, water molecules penetrate the
stems and cause the internal cells to swell resulting in
a bursting of the plant outer layer. In fact, the process in-
cludes maintaining the stem specimens in water for 3 days at
room temperature in order for it to be soft, and then, it is
beaten and separated using a knife or a razor. This operation
is repeated until the bers are separate from each other.
96
Crushing method
This method consists to cut stems into several samples
which are crushed using a roll in order to extract bers.
Then, the extracted bers are washed with boiled water at
90°C for 10 h in order to remove fat traces.
97
Rolling mill method
This method is used to cut samples into several strips of
small thickness which are immersed in water for 1 hour in
order to be softened. Then, strips are passed through
a rolling mill at small rotation speed, soaked another time in
water for 30 minutes to ease their extraction and are sep-
arated using a knife or a razor.
98
Chemical proceeding
Chemical extraction can affect the bers microstructure by
modifying its composition. One distinguishes namely the
alkali procedure which consists to immerse plant strips into
container of NaOH with a concentration of 1.5 N (Nor-
mality) at 70°C for 5 hours. Then, alkali-treated strips are
pressed. The resulting bers are washed in water and dried
in an oven.
99
It is reported that the main advantage of this
method is that bers exhibit less damage.
99
Combination of chemical and mechanical methods
The compression and roller mill techniques are usually used
after chemical treatments in order to extract bers. How-
ever, the choice of the corresponding extraction method
may depend on the expected quality of extracted bers.
Table 5. Some reported moisture content of some selected natural bers at 21°C and RH 65%.
12,55
Fiber Moisture content (%) Fiber Moisture content (%) Fiber Moisture content (%)
Abaca 15 Hemp 9 Bamboo 8.9
Coir 10 Pina 13 Agave 7.69
Flax 7 Ramie 9 Bagasse 8.8
Jute 12 Sisal 11 Pineapple leaf 11.8
Chakkour et al. 9
Among all natural bers, some studies have investigated the
effect of extraction processes on the mechanical properties
of bamboo bers due to their high growth rate and their
ability to x atmospheric carbon dioxide.
5,100
. Indeed,
Zakikhani et al.
5
investigated the mechanical behavior of
bamboo bers using different extraction methods (see
Table 7). The combination of chemical and mechanical
methods enables to extract bers with high tensile strength
ranging between 645 and 1000 MPa.
5
Furthermore, the
obtained bers by steam explosion exhibits moderate
strength that change from 308 MPa to 862 MPa but remain
better than those extracted using the rolling mill method
where the strength did not exceed 270 MPa.
5
The morphology of bers depends on the extraction
methods. Mechanical methods lead to a small global
roughness due to the presence of the cellulose surrounded
with lignin, hemicelluloses, and pectins. The local roughness
at the microbril level remains important because of the
presence of some impurities and inorganic components at the
microbril surface. The combination of chemical and me-
chanical methods shows small local roughness because of the
suppression of above-mentioned impurities, while the global
roughness increases as some of hemicelluloses and lignin are
removed.
1,5,102,101
Modication of natural bers. The main drawbacks that would
restrict the utilization of natural bers as reinforcements of
composites are their high moisture absorption and low thermal
stability.
103
In fact, the high hydrophility of the former makes
them incompatible with polymer matrices which lead to in-
terfacial debonding phenomena within composites.
8,19
Therefore, the surface of either natural bers or some ma-
trices like bio-based resins needs to be treated in order to
improve the stiffness and the strength of composites.
8,104
One
can distinguish between physical and chemical treatments.
Physical modications improve the mechanical adhesion
between matrix and natural bers without changing the
chemical properties of the latter. Corona, plasma, and ultra-
violet (UV) are the mainly used physical treatments to change
the surface properties of natural bers.
8
As in corona, plasma
treatments change the surface energy of cellulosic bers under
high voltage at low temperature. Regarding the UV treatment,
it increases the polarity of bers which allows a good in-
terfacial adherence with matrices.
105,106
In addition, chemical
treatments improve the adhesion between bers and matrices
via chemical reactions.
107,108
The most used techniques are
described in the following sections.
Chemical modication
Alkaline treatment
This method attacks the molecular structure of bers by
decreasing the hydroxyl groups.
109
It involves a decrease in
Table 6. Decomposition temperature of some selected lignocellulosic bers.
1,13,14,93,94,95
Fiber Decomposition temperature (°C) References
Bagasse 232
1,14
Bamboo 224
13,14
Cotton 232
1,13,14
Hemp 215
1,13,14
Jute 215
1,14
Kenaf 229
1,13,14
Pina 240
1,13,14
Chicken feather 350
93
Raw muntingia calabura 213
94
Silane-treated muntingia calabura 268
94
Alkali-treated muntingia calabura 254
94
Raw water hyacinth 289.22
95
Alkali-treated water hyacinth 294.67
95
Silane-treated water hyacinth 329.71
95
Citrifolia 248.44
92
Silane-treated citrifolia 274.71
92
Figure 6. TGA curves of raw and treated Agave Americana
natural ber adapted to Madhu et al.
136
10 Journal of Reinforced Plastics and Composites 0(0)
the amount of non-cellulosic components such as hemi-
cellulose, lignin, pectins, and waxes.
19
The resulting ef-
fective surface area of bers becomes cleaner allowing good
stresses transfer through cells. Further, given the optimum
concentrations, alkali treatment is the most widely used
treatment because of its simplicity and efciency.
Silane treatment
Silane forms a chemical link between ber surface and
matrix via siloxane molecules.
8
First, silanol is created
using existing moisture and hydrolyzable alkoxy groups and
the free chemical bridges reacts with cellulosic OH groups.
This ensures molecular continuity along the interface and
the formed hydrocarbon chains that contribute to reduce
swelling.
19
Acetylation treatment
In this treatment, bers are immersed into acetic acid and
acetic anhydride, respectively, in order to accelerate re-
actions.
8
An esterication reaction occurs between OH
groups and carboxy/anhyride groups of natural bers.
110
Accordingly, the ber surface becomes consequently
smoother and the mechanical stress transfer toward cells
interfaces is improved.
1
Benzoyl treatment
Benzoyl chloride is used to decrease the hydrophilic surface
property improving then the interfacial adherence and the
thermal stability of bers. This treatment consists of two
process steps in which waxes and impurities are removed,
then, bers are immersed in benzoyl chloride solution where
benzoyl groups are substituted to the OH groups.
1
Potassium permanganate treatment
This method consists of immersing bers in 0.5 wt.%
permanganate potassium and acetone for 30 min
11,112
The
resulting bers are washed and air dried for 24 h at
a temperature of 105°C. Mohammed et al.
111
have treated
sugar palm bers with 0.66 wt.% potassium permanganate
leading to an improvement of the tensile properties of
treated bers and their composites. Sanjay et al.
112
have
reported studies dealing with the effect of 0.05 wt.% of
KMnO
4
treatment for 2 min at room temperature on the
morphological, tensile, exural, and impact properties of
ramie reinforced polymer composites. The obtained results
show that there is an increase of the roughness of bers
leading to better adhesion with matrix in addition to an
improvement of the tensile, exural, and impact strengths.
Stearic acid treatment
Regarding stearic acid treatment, a solution of 1% stearic
acid in ethyl alcohol is slowly poured into bers which are
placed in a glass vessel.
11,112
Afterward, the treated bers
are dried at 80°C for 45 min. Bagasse bers were treated
with 3 wt.%, 5 wt.%, 7 wt.%, and 9 wt.% of stearic acid, and
then, dried at a temperature of 105°C for 45 min.
113
The
rheological studies revealed that stearic acid acts as a good
lubricant that minimizes shear viscosity. Kenaf bers were
treated with 0.4 wt.% stearic acid, washed with distilled
water, and dried in an air oven at a temperature of 80°C.
Salem et al.
114
have reported that such a treatment reduces
the hydrophilicity of bers by decreasing their water
absorption.
Seawater treatment
Seawater treatment is the simplest and most economical
method for modifying natural bers. After checking the pH
and the salinity of seawater, the bers are immersed for
30 days, and then, washed with water and dried at ambient
temperature.
11,112
Bleaching
The bers are treated using calcium hypochlorite Ca(ClO)
2
for 45 min, then washed with deionized water, and oven
dried at 80°C for 24 h.
11
Graft copolymerization
Regarding the graft copolymerization reaction, a monomer
(MMA) and an initiator are used once the bers are im-
mersed into water. The grafting parameters such as time,
temperature, volume of solvent, and the initiator and
monomer concentrations should be controlled in order to
obtain the optimal grafting percentage.
11
Actually, it is reported that a pretreatment using hot water
can be utilized in order to dissolve undesirable hemi-
celluloses, waxes, and impurities from the surface leading to
good matrix/bers compatibility.
1
In addition, many studies
have focused on effects of chemical treatments on the
mechanical properties of natural bers. Doan, Thi Thu
Loan
87
studied the effect of alkali treatment with different
weight concentrations and immersion time on the tensile
strength of jute bers.
87
The considered samples are
Table 7. Mechanical properties of extracted bamboo bers by using different extraction methods.
5
Extraction method Tensile strength (MPa) Youngs modulus (GPa)
Steam explosion 308862 1736
Retting 503 36
Crushing 420 38
Rolling 270
Combined chemical+compression 6451000
Combined chemical+roller mill 370480
Chakkour et al. 11
composed of several elementary bers. Tensile strengths of
jute bers treated by 1% NaOH for 4 h, 5% NaOH for
10 min, and 5% NaOH for 12°h are compared with that of
pure jute bers. It was found that the weight loss of bers
increases upon large treatment periods and concentrations
of alkaline. The weight loss of treated bers by 5% NaOH is
found to be higher than those treated with 1% NaOH at the
same immersion time. Doan Thi Thu Loan found that alkali
treatment with a weight percentage of 1% and immersion
time of 4 h was optimum for jute bers. This can be ex-
plained that by the fact that the treatment removes a part of
amorphous components such hemicellulose and lignin
which causes a rearrangement of cellulosic brils in the
loading direction leading to high tensile strength due to the
load distribution. The 5% alkali-treated bers for an im-
mersion time of 12 h show relatively small tensile strength
due to cellulosic brils damage and attacks by NaOH.
Furthermore, Chin SC et al.
18
studied the effect of alkali
treatment on the tensile strength of bamboo bers. The
soaking time is maintained to 48 h, while the NaOH con-
centration was of 5%, 10%, and 15%, respectively. They
showed that untreated bamboo bers exhibit the smallest
tensile strength and modulus, of about 140.1 MPa and
10.89, respectively. Using the above-mentioned treatments,
the tensile strength of bers was found to increase up to
255.67 MPa, 319.52 MPa, and 212.96 MPa for 5%, 10%,
and 15% NaOH concentrations, respectively. The tensile
modulus increases until reaching 16.63, 27.93, and
25.99 GPa, respectively. The lowest tensile properties of 5%
NaOH treated bers compared to 10% are attributed to the
incomplete delignication of bers. However, it is worthy to
note that the tensile properties of bers were improved by
the increase of NaOH concentrations until a critical value,
the solution is expected to damage the cellulose chains
reducing then the crystallinity of bers
115
Moreover, the
10% alkali-treated bers with soaking periods of 24 h and
72 h exhibit a tensile strength of about 267.44 and
176.90 MPa, respectively, while the tensile modulus is
found to be 18.78 and 24.29 GPa, respectively. The small
tensile strength is obtained for bamboo bers immersed
within 10% NaOH solution for 72 h that could be attributed
to the elimination of lignin which links the cellulosic chains
together.
18
Table 8 shows a brief summary of the treatments
effects on the mechanical properties of some lignocellulosic
bers and composites.
Osorio L. et al.
116
have also studied the tensile strength
and the tensile modulus of alkali-treated bamboo bers with
the effect of 1%, 3%, and 5%, respectively, soaked for
20 min. They showed that the tensile strength of bers
increases by 1%, 6%, and 4% for 1%, 3%, and 5% treated
bers, respectively. The tensile modulus increases by 7%,
21%, and 19%, respectively, after chemical treatment. This
was explained by the removal of the weak non-cellulosic
components.
116
The effect of silane treatment has been
Table 8. Effects of treatment on the mechanical properties of some selected bers and composites.
1,11,112,123
Fiber/composite Treatment Results References
Jute 1%, 5% NaOH Increase in tensile strength of 1% treated bers compared to 5%
treated ones
87
Bamboo 5%, 10%, and 15% alkali
treatment
Increase in tensile properties compared to untreated bers
18
Borassus 5% NaOH The tensile properties such strength and Youngs modulus were
improved
124
New cane 6% NaOH Improvement of tensile and exural strengths
125
Bamboo 1%, 3%, and 5% NaOH The tensile strength and Youngs modulus increased after
treatments
116
Sisal bers 5% benzoyl chloride The crystallinity index and thermal stability are improved
126
Areca bers 6% NaOH Alkali treatment enhances the thermal stability of bers
127
Coir-polyester composite 5% NaOH Better impact and exural strengths
128
Hemp-polypropylene
composite
2%, 4%, and 6% silane The silane does not affect the tensile and exural strengths
129
Hemp-polypropylene
composite
4% and 6% NaOH The treated bers exhibited the highest tensile and exural
strengths
129
Flax-polypropylene
composite
2.5% silane 6% and 3% enhancement in the tensile and exural strengths
130
Kenaf-polypropylene
composite
2% NaOH Removal of hemicellulose
131
Bagasse-epoxy composite 5% KMnO
4
Improvement of tensile and thermal properties
132
Alfa- polypropylene 0.25 M NaOH The removal of non-cellulosic impurities and the increase of
microbrillar angle MFA
112
Abaca-phenolic composite 50% acetylation The tensile and impact strengths were improved
14
Banana epoxy composite 1% NaOH Improvement of the tensile and exural properties
133
12 Journal of Reinforced Plastics and Composites 0(0)
investigated by Tung Nguyen et al.
117
They found that the
tensile strength of raw steam exploded bamboo bers is
higher than that of silane-treated bamboo.
Other investigations have reported on hygrothermal
treatments of natural bers. Cuiyin Ye et al.
119
have studied
the effect of hygrothermal treatment at temperatures 160°C,
180°C, 200°C, and 220°C and relative humidity of 100% on
the porous structure of Moso bamboo and its mechanical
properties. They found that 220°C hygrothermal bamboo
sample appeared to be more brittle than untreated sam-
ples.
119
This was attributed to the loss of hemicellulose
leading to an increase of pores numbers after hygrothermal
treatment at 220°C. The total volume of pores and the
average diameter maximize at 160°C and is increased by
46% and 16%, respectively, compared to untreated ones.
This is due to the hemicellulose degradation and the sup-
pression of volatile components at high temperatures. At
160°C, the average pore size slightly decreases because the
hemicellulose degradation is still processing and the free
pores can be relled by the chemical components owing
under high temperatures. In addition, the number of hy-
drogen bonds between cellulose chains increases which
results in the reduction of the number of pores and their
sizes. Further, they showed that the relative crystallinity
reaches the maximum (36.92%) at a temperature of 180°C
being 11.1% higher when compared to untreated samples.
They explained this growth of crystallinity by the high
reactivity of the amorphous celluloses. The latter areas lost
water by condensation and produce hydrogen bonds en-
hancing consequently the crystallinity. Moreover, the re-
lease of hemicellulose parts led to the rearrangement of
amorphous cellulosic chains producing low energy bonds. It
is also reported that after overpassing a temperature of
180°C (200°C220°C), the degradation of crystalline cel-
lulose begins leading to a decrease of the relative crystal-
linity. Yun et al.
118
and Cuiyin et al.
119
independently found
that relative crystallinity of Moso bamboo increased at
180°C by 5.67% in comparison to untreated samples.
Cuiyin et al.
119
showed that the microbrillar angle
(MFA) decreases slightly during the hygrothermal treatment
and explained such a result by the softening of lignin and
hemicellulose around cellulosic brils. The amorphous
chains of cellulose are moved and reoriented when some
polysaccharides are degraded, resulting in a drop of the
microbrillar angle. The reduction of the number of voids
and the microbrillar angle as well as the increase of
crystallinity at 180°C was found to enhance both the elastic
modulus and hardness by almost 21%. At temperatures
higher than 180°C, they found that the crystalline cellulose
began to degrade on account to crystallinity decrease
leading to small elastic modulus.
Besides, there are various physical treatments that could
be used to change the surface energy of bers and enhance
the bers-matrix interlocking. Some reported treatments are
cited below.
Physical modication
Plasma treatment
Plasma treatment is used to remove impurities and change
the surface energy in cellulosic bers, using semi-industrial
prototype machines at ambient conditions.
122
Sanjay
et al.
112
have reported on jute bers which are treated with
low pressured O
2
plasma environment for a duration of
15 min at several power ranges. The obtained results reveal
that plasma treatment enhances the bers-matrix adhesion,
resulting in an improvement of the mechanical properties. It
is worthy to highlight that the increase of mechanical
properties of composites is proportional to plasma power.
Vacuum ultraviolet irradiation treatment
Vacuum ultraviolet irradiation is a new method consisting of
removing impurities from the surface of plant bers. Sanjay
et al.
112
have reported an interesting work revealing the
surface oxidation of bers by vacuum ultraviolet irradiation.
Initially, the bers are placed in a chamber of stainless steel
and high energy below 200 nm were used for irradiation.
Ozone treatment
This process allows to generate surface oxidation on the
bers surfaces using ozone gas at a temperature of 25°C and
aow rate of 70 L/h, while the time of exposure ranges
between 5 min and 9 h.
11,112
Then, the treated bers are
washed with distilled water and dried for 24 h at 60°C.
Sanjay et al.
112
have found that the degree of crystallinity
and crystallites sizes increase after ozone treatment. Avinc
et al.
20
have studied the effect of ozone treatment duration
on the burst strength of soybean bers reinforced compo-
sites showing that the optimal strength is obtained for high
treatment duration.
Corona treatment
Sanjay et al.
11
described the experiment set up of some
previous works devoted to studying the effect of corona
treatment on the cellulose ber surface. For this purpose,
about 1 g of bers is placed in the corona cell and treated for
1 min with an applied potential of 15 kV and frequency of
60 Hz at a temperature of 25°C and relative humidity of 50
RH%.
Y-Ray treatment
The bers are treated with NaOH solution (16 mol.dm
3
),
followed by tetramethylammonium hydroxide (1
3 mol.dm
3
). Afterward, the treated bers are neutralized,
dried, and post-processed with 5 kGy/h dose rate. Some
ndings
112,121
show that the 25 kGy γ-ray treatment im-
proves the bers-matrix adhesion and enhances the physical
properties of sisal bers composites.
Characterization of natural bers. This section is devoted to
present the most widely useful characterizations techniques
for natural bers.
Chakkour et al. 13
Spectroscopic characterization of Fourier transform infrared
spectroscopy. Fourier transform infrared spectroscopy (FTIR)
is used to determine the chemical composition of natural
bers, composites reinforced natural bers, and their func-
tional groups that interact with lignocellulosic bers.
1,11,123
This method consists of imposing an infrared radiation to the
sample and measuring the wavelengths, thresholds, and the
intensity of the radiation absorption (see Table 9).
Mofokeng et al.
134
have investigated the spectroscopic
characteristics of randomly oriented sisal bers reinforced
PLA composite. They found that O-H bonds became more
pronounced when increasing the bers content, which is
explained by the presence of free hydroxyl groups at bers
surfaces.
Marianne
135
reported the most common functions in the
study of natural bers and polysaccharides.
Doan Thi Vi
87
studied the effect of 1% NaOH (1 im-
mersion day) treatment on bamboo bers by infrared
spectroscopy FTIR. According to Ref.
87
the peaks corre-
sponding to carbonyl groups C=O (1750 cm
1
) and the
lignin characteristic group OCH
3
(1240 cm
1
) have
disappeared. Doan Thi concluded that alkali treatment
contributes to reduce the content of amorphous components
especially lignin and hemicellulose.
87
Chin SC et al.
18
reported a comparison between raw and
alkali-treated bamboo bers using infrared spectroscopy
method.
18
They showed that the peak of the band at
895 cm
1
is due to the C-O stretching vibrations of
amorphous cellulose that increases with the NaOH treat-
ment. They showed that the intensity of hydroxyl group
peak of the band at 3446 cm
1
increases after alkali
treatment allowing a good ber-matrix interaction. In ad-
dition, the intensity of bands peaks attributed to C-O, C-C
stretching (1036 cm
1
) decreases after chemical treatments
showing a reduction of hemicellulose and lignin con-
stituents of treated bamboo bers. Table 9 gives an overview
of the infrared bands peaks and the associated chemical
bands of bamboo bers.
135
P. Madhu et al.
136
have studied the chemical composition
of raw and chemically treated Agave Americana bers
(AAF) with NaOH, stearic acid (SA), benzoyl peroxide
(BP), and potassium permanganate (PP).
136
They observed
that the band peak appearing at 2357 cm
1
is attributed to
waxes components that become more pronounced for un-
treated bers compared to chemically treated ones. This was
explained by the partial removal of waxes within treated
bers. The two peaks found between 3100 cm
1
,
3427 cm
1
and 2917 cm
1
, respectively, are expected to
characterize the cellulose respective hydroxyl and C-H
stretching groups.
X-ray diffraction. Vegetable bers are mainly constituted
of sugar-based polysaccharides especially cellulose,
hemicellulose, and lignin. The cellulose which is the main
component of lignocellulosic bers appears into two
chemical states which are crystalline (cellulose 1) and
amorphous (cellulose 2) celluloses. The content of cellulose
1 in natural bers determines their mechanical
properties.
6,11,123
The crystallinity index (%) corresponds to
the amount of crystalline cellulose compared to the global
quantity of amorphous components.
6,135
Marianne Le-
troedec
135
have studied the crystallinity of hemp bers
under NaOH and Ca(OH)
2
treatments. The obtained
crystalline peak (2θ) referring to the angle between the
incoming and the scattered X-Ray beam is comprised be-
tween 22° and 23° and corresponds to the crystallographic
plane family (002) of cellulose 1. Some depicted small
peaks corresponds to the plane (110). According to Ma-
rianne Letroedec,
135
if the content of crystalline cellulose is
high, the peaks remain distinct and if bers contain a high
content of amorphous components, the peaks merge and form
a unique one. On the other hand, the obtained results showed
that NaOH treatment increases the crystallinity of hemp bers
from 55% to 80%. This was explained by the removal of
a part of amorphous areas as hemicellulose, pectins, and
lignin in addition to other inorganic components. Further, the
hot water and calcium chloride treatments were found to have
Table 9. Overview of the principal infrared bands of bamboo bers and their corresponding chemical bonds.
135
Wavelength (cm1) Chemical bonds Associated components
3300 O-H Polysaccharides
2885 C-H symmetric bending Polysaccharides
2850 CH2 symmetric bending Cires
1732 C=O Xylanes (hemicellulose)
16501630 OH Water
1505 C=C Lignin
1425 Deformation at the group plane C-H Pectines, lignin, hemicelluloses
1373 CH2 Polysaccharides
1335 C-O vibrations Cellulose
1240 Acetyl groups deformation (xylanes) Lignin, polysaccharides
1162 C-O-C Cellulose, hemicellulose
895 Glycosidic bonds vibrations Polysaccharides
670 C-OH vibration Cellulose
14 Journal of Reinforced Plastics and Composites 0(0)
a negligible inuence on either cellulose or amorphous
parts.
135
Chin SC et al.
18
performed X-Ray diffraction analysis of
bamboo bers with different NaOH concentrations and
soaking durations. The amorphous peak is characterized by
a low intensity at a 2θof 16°, whereas the peak corre-
sponding to the crystalline part is located at 2θ= 20°. Both
peaks are associated with the crystallographic plane (002).
Further, the results showed that the crystallinity index of
pure bamboo bers is small (49.92 %) which is explained by
the high content of amorphous components in such bers. In
fact, the crystallinity of specimens soaked within NaOH
solutions for 48 h has increased to 52.86, 58.73, and 53.68%
for 5%, 10%, 15% alkali-treated samples, respectively. The
increase of crystallinity index after chemical treatment is
due to the removal of waxes and a part of hemicellulose,
lignin, and pectines components.
Thermogravimetric analysis. Thermogravimetric analysis
(TGA) allows to determine the evolution of mass loss of
materials as a function of temperature.
11,123
Several studies
have addressed this point. For example, P. Madhu
136
have
explored thermogravimetric curves of Agave Americana
bers. They divided the degradation of bers into three
stages: The rst stage is comprised between 50°C and
185°C corresponding to the moisture evaporation and the
removal of waxes components. The weight loss at this
degradation stage did not exceed 15%. The second stage
was found to be between 220°C and 340°C where the mass
loss was attributed to the degradation of hemicellulose and
cellulose components. The mass loss at this stage remains
under 60%. Further, they found that the third stage occurs
above 340°C and is related to the degradation of lignin and
other non-cellulosic components. Furthermore, Chin SC
et al.
18
have performed thermogravimetric analysis of pure
bamboo bers and 5%, 10%, 15% NaOH alkali-treated
bamboo bers for 24, 48, and 72 h.
18
They also observed
the above mentioned three degradation steps in the ther-
mogravimetric curve of non-treated and treated bers. The
rst stage between 26°C and 155°C is related to the
evaporation of moisture. The amount of adsorbed moisture
by the non-treated bers (11%) is higher than the treated
bers (7%10%). This was explained by the effects of
alkali treatment on the hydrophilicity of cellulosic bers
leading to the removal of amorphous zones. The second
decomposition stage ranges from 199°C to 379°C and is
essentially caused by the removal of holocellulose (cel-
lulose and hemicellulose) components. At this stage, the
mass loss of untreated bamboo bers is much higher (52%)
due to the high content of hemicelluloses around cellulosic
brils. The nal stage occurs between 364°C and 499°C
where the decomposition is due to the degradation of the
hardest components. Chin SC et al.
18
reported that alkali
treatment increases the content of lignin in bers.
18
In
Refs.
18,136139
all reported tests showed that alkali treat-
ment increases the thermal stability of bers while three
degradation stages have been identied. Dehydration,
degradation of cellulose components, and nal de-
composition of the bers lead to the formation of char.
Figure 6 illustrates a typical TGA curve of a non-treated
Agave Americana ber.
Density measurement. To measure the density, the bers
are dried for 48 h in a desiccator that contains calcium
chloride, immersed in toluene for 2 h to eliminate micro-
bubbles, and then, kept in a pycnometer. The density of
bers is calculated by the following formula
11,112,123
ρ¼m2m1
ðm3m1Þðm4m2ÞρT(1)
where, m
1
,m
2
,m
3
,m
4
, and ρ
T
are the mass of the empty
pycnometer, pycnometer containing bers, pycnometer
lled with toluene, pycnometer lled with bers and tol-
uene, and the density of toluene, respectively. Deeksha
et al.
123
reported several works in which bers density is
measured using Archimedesprinciple. In fact, several
solutions such as water and canoila oil are used to measure
the apparent loss in weight of natural bers. Raja et al.
140
and Khan et al.
141
have estimated the density of Baobab and
Eleusine Indica bers using Archimedesprinciple with
ethanol as solvent and reported values of 1.1041 g/cm
3
and
1.1430 g/cm
3
, respectively.
Scanning electron microscopy and OM analyses. Scanning
electron microscopy (SEM) is usually used to study the
morphology of natural bers reinforced composites. Several
studies have been carried out in order to analyze either the
morphology of bers or the quality of ber-matrix
bonding.
11,112,123
Madhu et al.
136
analyzed the surface
morphology of untreated and treated Agave Americana
natural ber with sodium hydroxide, stearic acid, benzoyl
peroxide, and potassium permanganate. They noticed that
raw bers are constituted of parallel microbrils surrounded
by waxes, oils, and others impurities which make the ber
surface less rough. However, they highlighted that chemical
treatments induce an increase in the number of pores on the
ber surface while impurities such waxes are eliminated.
Chin SC et al.
18
investigated the effect of NaOH treatment
on the bamboo ber surface. They observed that the ber
diameter decreased by 25% after alkali treatment which is
explained by the dissolution of hemicellulose and the re-
moval of lignin and other impurities. Regarding the ber-
matrix bonding, Pothan et al.
142
have investigated the effect
of ber volume content on the morphology of banana-
polyester composites, and reported that the composite
with 40% ber content has a better ber-matrix bonding.
The adhesion between short bamboo bers and poly-
propylene has been studied by Thwe and Liao
23
using
Scanning Electron Microscopy (SEM). The SEM images
showed that the ber/matrix adhesion is relatively poor and
the bers are completely debonded from the matrix.
Regarding the diameter of natural bers, it is measured
by optical microscopy OM or a scanning electron
Chakkour et al. 15
microscope SEM.
11,123
These techniques allow measure-
ments with an accuracy of 10
6
μm. Several studies have
pointed out difculties in measuring the diameter of bers
because of their irregular shape and variable thickness.
Indeed, the presence of amorphous and inorganic com-
ponents makes the bers cross-section uncircular. Hence,
the diameter is evaluated in 5 different locations along the
ber length and the average value is considered. It is re-
ported that the diameter of pineapple leaf and ramie bers
ranges between 20 μmand80μm.
55
The averaged bers
diameters of the most used natural bers are presented in
Tab le 3.
Moreover, the hierarchical structure of plant-based bers
such as vascular bundles, ber bundles, elementary bers,
cellulosic microbrils, cells, primary and secondary cell
walls could be investigated using optical microscope OM or
by SEM analysis.
11,112,123
The EDX analysis allows to determine the chemical
compositions of bers surfaces.
11,123
Indeed, the reactions
of carbon C and oxygen O with sodium Na, aluminum Al,
silicium Si, and magnesium Mg could be detected, except
hydrogen H that represents the main constituent of natural
bers. Najeeb et al.
143
have performed an elementary
composition analysis of Yankee pineapple bers and have
reported values of 43.84 wt.%, 46.80 wt.%, and 9.36 wt.%
of carbon, oxygen, and potassium, respectively.
Chemical analysis. The chemical analysis permits to de-
termine the content of each natural bers component. The
content of waxes and fats, pectins, lignin, holocellulose,
cellulose, and ash are determined according to the ASTM D
110756, ASTM 111056, ASTM D 110656, ASTM D
110456, ASTM D 110360 and ASTM D 110284, re-
spectively.
87
Sanjay et al.
11
have reported that the cellulose
is generated in an insoluble solution after treating the bers
with 95% nitric acid solution and ethanol. The hemi-
cellulose content is estimated by exposing the bers to
10 mL of cold neutral detergent solution and sodium sulte
for 1 h, ltering the mixture through glass crucible and
washing the residue by hot distilled water. It is reported that
APPITA, TAPPI, and Conrad methods may be used to
determine the lignin, ash and waxes contents, respectively.
Najeeb et al.
143,123
have identied the chemical composition
of pineapple leaf bers by TAPPI standards T20305 OS-74
and T222 OS-83 to determine the content of cellulose/
hemicellulose and lignin, respectively. The obtained com-
positions are 47.74% of cellulose, 15.98% of hemicellulose,
and 2.44% of lignin. Vijay et al.
144
have investigated the
chemical composition of Leucasaspera bers by following
the APPITA method. The corresponding contents are
58.3%, 8.9%, and 4.5% of cellulose, hemicellulose, and
lignin, respectively.
Differential scanning calorimetry analysis. This analysis is
performed by DSC machine and permits the determination
of melting and glass transition (T
g
) temperatures of natural
bers
11,123
The tests are performed with a heating rate of
10°C/min starting from 0°C until reaching the melting peak,
in nitrogen atmosphere. Palai et al.
145
reported that four
temperature peaks (76.3 °C, 154.75 °C, 242.75 °C, and
330.8 °C) are observed in the DSC curve of Crassipes bers
and corresponds to the water removal, glass transition
temperature, hemicellulose, and cellulose degradations,
respectively.
Single ber tensile test. As early mentioned, the tensile
properties of natural bers are inuenced by several factors
such as dimensions (length, width and thickness), envi-
ronmental conditions and plant type.
11,123
The test should
be performed by following the ASTM C1557-03 standard,
with a cell load of 5 kN or less and speed rate not exceeding
0.5 mm/min. It is worthy to note that each ber edge needs
to be xed with epoxy resin to avoid bers damage. Vijay
et al.
144
and Najeeb et al.
143
have investigated the tensile
strength of Vachellia and pineapple leaf bers by using
universal testing machine and recorded values of
33.075 MPa and 420.3 MPa, respectively.
Activation energy. This technique allows to determine the
minimum required energy to decompose the bers.
123
It is
reported that a high activation energy corresponds to
a better thermal stability. Kumar et al.
146
and Manimaran
et al.
147
have investigated the activation energy Acacia
nilotica and Furcracea Foetida bers and recorded values
of 69.73 kJ/mol and 66.64 kJ/mol, respectively. In
Refs.
145,148,149
the activation energy of Albizia Lebbeck
bark, Sida Cordifolia, and Eichhornia Crapssipes bers is
89 kJ/mol, 73.1 kJ/mol, and 66.32 kJ/mol, respectively.
Ganapathy et al.
150
have reported that alkali treatment
increases the activation energy of banyan tree bers from
72.45 kJ/mol to 72.65 kJ/mol. Similarly, alkali treatment
improves the thermal stability and activation energy of
treated Saharan Aloevera bers when compared to un-
treated ones.
151
Matrices
In industrial applications, natural bers are used as re-
inforcement of different matrices constituting the well-
known biocomposites. The main role of matrices is to
Figure 7. Classication of the used matrices in natural bers
reinforced composites.
1
16 Journal of Reinforced Plastics and Composites 0(0)
ensure bers protection and solid bonds with them allowing
mechanical resistance of the composite.
19,152
They can be
classied into two parts, namely, petrochemical-based
resins such thermoplastics, thermosets and bio-based res-
ins (see Figure 7).
Petrochemical-based matrix is a chemical product de-
rived from petroleum as coal and natural gas. The most used
one is the thermoplastic resin which is a polymer that can
change reversibly from the viscous to solid states under
cooling and heating respectively. Indeed, macromolecular
chains of thermosets matrices are linked by hydrogen el-
ements which are removed irreversibly upon heating.
However, each type of thermoplastic resins is characterized
by its own properties. The polyethylene (PE) exhibits high
ductility, impact strength, and good fatigue resistance, while
the polypropylene (PP) shows high temperature and di-
electric resistance.
153
Regarding thermosets resins, they are infusible materials
that can be cured by heating or by the use of a catalyst. In fact,
macromolecular chains of thermosets are linked by covalent
and non-covalent bonds making them sensitive to the heating
based reshaping when compared to thermoplastic ones.
154
In
addition, thermosets show higher modulus, good thermal
stability, and high chemical resistance. Among thermosets
resins, epoxy presents high thermal and mechanical prop-
erties, high water resistance and long working times.
155
Moreover, bio-resins are polymers based on renewable
resources and obtained from the cellulose and starch or by
the polymerization of plant-based sugars and oils
1,8,157
such
as polylactic acid (PLA) and polyethylene terephthalate
(PEET). Polylactic acid is a homopolymer of starch, while
the monomer unity of the PEET is a bio-ethylene glycol.
The latter is obtained from glucose fermentation and the
oxidation of the resulting bio-ethylene.
156
Furthermore, bio
polyethylene terephthalate, bio polypropylene, and bio
polyethylene are bio-based resins that are partially bio-
degradable.
8
The PLA exhibits a tensile strength and failure
elongation of about 410, respectively and 517 times
higher than the starch ones in addition to an important
exural strength (see Table 10).
However, the development and the production of bio-
based resins still suffer from limitations. For example
a production of 3.4 million tons has been reported in 2011
and 235 million tons of petrochemical based ones in 2019,
19
in addition to the high moisture absorption and low thermal
stability of bio-resins.
8,158
Fiber-matrix interface
Actually, the mechanical properties of composite materials
mainly depend on the ber content and the ber-matrix
interface.
161163
Indeed, the matrix transfers the load to the
bers thanks to shear stress in order to ensure their reinforcing
role. However, to understand the adhesion phenomena, it is
important to consider the different types of bonding.
Mechanical bonding: This type results from the mechanical
interlocking and depends on the ber roughness. The
principle is to introduce the matrix within the microvoids of
reinforcements in order to provide a good ber-matrix load
transfer by means of the shear stress
163
(see Figure 8(a)).
Chemical bonding: The chemical bonds may be formed
thanks to the reactive sites of lignocellulosic bers which
are the hydroxyl groups of cellulose.
164
Besides, several
coupling agents are utilized to form chemical bonding
between the hydrophobic matrices and vegetable bers
19,165
(see Figure 8(b)).
Table 10. Relevant mechanical properties of the most used matrices including thermoplastics/thermosets and bio-resins
67,159,160
.
Type Resin
Tensile
strength
(MPa)
Youngs
modulus
(GPa)
Elongation
(%)
Flexural
strength
(MPa)
Flexural
modulus
(MPa)
Density
(Kg/m
3
)
Thermoplastic Polypropylene 2641.4 0.951.77 15700 40 1.5 890910
Polystyrene 2569 4512.5 70 2.5 9601040
Low-density
polyethylene
4078 0.0550.38 90800 9 0.2 910925
High-density
polyethylene
14.538 0.41.5 2130 32 1.2 940960
Polycarbonate 5570 ——200 1.2
Polyvinyl alcohol 1600 ——6 1.191.31
Thermoset Polyester 41.489.6 2.074.41 22.6 70110 24 10401400
Epoxy 55130 36210 110150 34 11101400
Polyurethane 18 - 200800 ——1250
Bioresin Starch 56 0.1250.85 3144 52 2.4 10001390
PLA 2160 0.353.5 2.565170 4.2 12101250
PHA 1824 0.71.8 3.25 ———
Chakkour et al. 17
Electrostatic bonding: The electrostatic bonding results from
arrangement of charged particles with opposite polarities on
both matrix and bers
163
(see Figure 8(c)).
Several studies were focused on the optimization of
lignocellulosic bers-polymer matrix interface in order to
enhance the mechanical properties of the well-known bi-
ocomposites. In this way, chemical and physical treatments
were previously introduced in the section 1.1.2 to deal with
the ber-matrix incompatibility.
9
Several works used chemical treatments such as NaOH
to increase the roughness of bers for a better mechanical
interlocking with matrix. Others researchers used coupling
agents to further optimize the interface. Despite these
chemical treatments, a signicant void content is still no-
ticed. Besides, the developments are directed toward the use
of nanoparticles, specically, cellulose, clay, graphene,
silicone dioxyde SiO
2
, titane dioxyde TiO
2
, and calcium
carbonate CaCO
3
, to reduce the void content and enhance
the properties of biocomsposites. Indeed, the low cost and
ecofriendly character of clays and cellulose make them the
best candidate to reinforce composites.
Mechanical properties and theoretical
modeling of natural bers reinforced
composites NFRCs
Mechanical and tribological properties:
Experiments side
Many research works investigated the mechanical proper-
ties of biocomposites depending on several parameters.
166
Indeed, Rivalani et al.
167
investigated the effect of alkali
treatment (20% by weight) for 1 h on the tensile, exural,
hardness properties, and the water absorption behavior of
polyester composite reinforced by randomly dispersed sisal
and glass bers. They tested the composites following ve
congurations. The obtained composites by alternating
three layers of treated sisal bers and four layers of glass
bers have the highest tensile and exural strength of
128 MPa and 184 MPa, respectively. Similar tests were
carried out on composites with untreated sisal bers. In this
case, the tensile and exural strength reach 112 MPa and
156 MPa, respectively, decreasing by 12.5% and 15.2%
compared to treated bers reinforced composites . In ad-
dition, Rivalani et al. showed the water absorption behavior
of different composites and concluded that the increase in
sisal bers layers increases the water absorption ca-
pacity. Moreover, the alkali treatment was found to
decrease the hardness of composites from 78 to 55.2
HRB, where the highest value is obtained for layered
composites by six untreated mats of sisal bers and four
mats of glass bers.
The mechanical properties of a polyvinyl butyral (PVB)
composite reinforced by kenaf bers were studied by
Salman et al.
168
They found that the composite with 45/
45° oriented bers exhibit the highest impact resistance
compared to the case of 0/90° orientation.
168
Yan et al.
169
investigated the mechanical properties of alkali-treated coir
bers reinforced epoxy composites. The tensile, compres-
sive, and exural strengths are 23.8 MPa, 2.98 MPa, and
40.4 MPa, respectively. The tensile, impact, and exural
properties of sisal bers reinforced epoxy composites were
studied by Yan et al.
170
using injection molding processing
technique. Their values reach 180.45 MPa, 46.5 MPa, and
191.37 MPa for the tensile, compressive, and exural
strengths, respectively. Ramesha et al.
171
carried out the
tensile test of hemp bers reinforced epoxy composites.
They showed that composites exhibited 55 MPa and
4.5 GPa for the tensile strength and Youngs modulus,
respectively. Codispoti et al.
172
studied the effect of the
matrix on the tensile properties of Hemp bers reinforced
epoxy and polyester composites. The obtained tensile
strength reaches 63.12 MPa and 58.20 MPa, respectively,
for epoxy and polyester-based composites. Biswas et al.
173
explored the effect of bers types (bamboo and jute) on the
mechanical properties of epoxy-based composites. They
showed that bamboo bers reinforced composites exhibit
a high tensile strength while the jute bers based ones
display the highest Youngs modulus. Mir et al.
174
high-
lighted the effect of temperature on the tensile properties of
jute bers reinforced epoxy biocomposites, and showed
that their mechanical properties dropped by 50% at about
180°C. Ramesha et al.
175
investigated the effect of volume
fraction on the tensile and exural properties of banana
bers reinforced epoxy composites. They reported that
composites with 50% bers volume fraction exhibited the
highest tensile strength while the optimal exural strength
is obtained for composites with a bers volume fraction of
almost 60%. Punyamurthy et al.
176
studied the effect of
ber content on the impact strength of treated abaca bers
reinforced epoxy composites, and concluded that a ber
content of about 40 wt.% within composites shows the
Figure 8. Different types of bonding (a) Mechanical, (b) Chemical, and (c) Electrostatic bonding.
19,165
18 Journal of Reinforced Plastics and Composites 0(0)
maximum impact strength of 7.68 mJ/mm
2
. Sanjay MR
et al.
177
have assessed the effect of stacking sequences on
the tensile and exural properties of jute/kenaf/glass woven
fabrics. Their results reveal that the laminates made only
from Glass exhibited a tensile strength of 332 MPa. The
laminates including only pure jute and pure kenaf fabrics
display low strength values, not exceeding 35 MPa and
45 MPa, respectively. However, their ndings show that the
jute/kenaf hybridization slightly increases the tensile
strength of the laminates to reach 47 MPa. They have found
that the incorporation of glass fabrics as the outer layers
increases gradually the tensile strength of the laminates to
129 MPa. Similarly, they found that glass/kenaf/jute hybrid
laminates exhibit the highest exural strength reaching
239 MPa. The authors highlighted that the exural prop-
erties depend on the laminate stacking sequences. They
mentioned that having glass and kenaf as skin plies im-
prove the exural strengths of hybrid laminates. Arpitha
et al.
178
have investigated the hybridization of sisal and
glass on the tensile and exural properties of hybrid
laminates. The reported results show that the highest tensile
and exural strengths are exhibited by glass laminates and
reach 346 MPa and 318 MPa, respectively. However, pure
sisal laminates display the lowest tensile and exural
strengths remain under 33 MPa and 124 MPa, respectively.
The authors have pointed out that the glass/sisal hybrid-
ization increases the tensile and exural strengths of
laminates to reach 168 MPa and 241 MPa, respectively. In
addition, Arpitha et al.
178
studied the effect of 0.72 wt.% of
silicon carbide on the mechanical properties of sisal lam-
inates. They revealed that the ller decreases the tensile
strength of laminates from 33 MPa to 31 MPa and increases
the exural strength from 124 MPa up to 168 MPa. Athith
et al.
179
have studied the effect of tungsten carbide llers
with 23μm size on the physical, mechanical, and tri-
bological properties of jute/sisal/E-glass reinforced natural
rubber/epoxy composites. Their ndings reveal that the
incorporation of tungsten carbide llers enhances the
density and the mechanical properties of composites. For
10 wt.% llers content, the corresponding tensile strengths
are found to be 109.7 MPa, 71.38 MPa and 222.7 MPa for
glass/sisal-epoxy, jute/rubber-epoxy, and glass/jute-epoxy
laminates, respectively. The exural strengths are 149 MPa,
35.38 MPa, and 100.3 MPa for glass/rubber-epoxy, glass/
sisal-epoxy, and jute/rubber-epoxy laminates, respectively.
Regarding tribological properties, the wear rate decreases
with the llers content. Vinod et al.
180
have investigated the
inuence of stacking sequences on the mechanical prop-
erties of jute/hemp reinforced bio-epoxy hybrid compo-
sites. They reported that the laminates with jute core layer
H/J/H exhibit the highest tensile strength of 65.44 MPa
when compared to other laminates. The optimal Youngs
modulus is obtained for pure hemp laminates H/H/H and
reaches 1.87 GPa. The maximum exural strength and
modulus are observed for jute core layer laminates H/J/H
and pure hemp composites H/H/H reaching 121.2 MPa and
3.48 MPa, respectively. The high exural and tensile
strengths for H/J/H laminates can be explained by the
presence of jute as a core that acts as a plasticizer due to its
high amorphous components which prevents the fracture
growth in laminates. Jothibasu et al.
181
have studied the
mechanical behavior of areca sheath/jute/glass bers re-
inforced hybrid composites. The reported results show that
the jute/areca/glass bers hybrid composites exhibit the
highest mechanical strengths compared to jute/glass and
pure areca/glass reinforced hybrid laminates. The tensile,
exural, compression, and shear strengths reach
46.99 MPa, 5.4 MPa, 44.5 MPa, and 32.58 MPa, re-
spectively. For jute/areca/glass bers composites, the areca
layer is sandwiched between high strength jute and glass
layers which explain their important mechanical properties
compared to other laminates. Krittirash et al.
182
have in-
vestigated the thermal, tensile, and impact properties of
kenaf/sisal bers reinforced bio-epoxy hybrid composites
and reported that the incorporation of bers decreases the
mechanical properties of composites which suggest that the
applied load is not effectively transferred from matrix to the
reinforcements. However, the hybrid composites show im-
portant resistance to weathering conditions exhibiting
moderate mechanical strengths when compared to neat bio-
epoxy. The optimal tensile and impact strengths are obtained
for pure kenaf bers and kenaf/sisal/kenaf bers reinforced
laminates and reach 23.56 MPa and 7.2 kJ/mm
2
,respectively.
The Tabl e 11 gives a summary of the mechanical properties
of the most relevant natural bers reinforced polymer
composites. The Figures 9 and 10 show the tensile and
exural strengths of some NFRCs, respectively.
On the other hand, during the relative motion of surfaces,
the wear and friction coefcients remain the main investigated
tribological properties.
183
The friction coefcient denes the
frictional behavior of composite materials while the wear one
corresponds to the progressive loss of the material during
mechanical and/or chemical processes.
183
Wei et al.
184
have
studied the effect of temperature on the tribological properties
of sisal bers/phenol formaldehyde composites and concluded
that the wear rate increases with temperature. Mutlu et al.
185
have studied the tribological properties of rice straw and rice
husk dust reinforced pads and reported that the incorporation
of these bers increases the tribological properties of the
corresponding composites. Similarly, Ojha et al.
186
showed
that wood apple shell particles reinforced composites exhibit
the lowest erosion wear compared to coconut composites. El-
Tay eb
187
have investigated and reported that the wear rate of
chopped sugarcane bers composites is better than that of glass
bers reinforced composites.
Comparative conclusion
Figures 9 and 10 show that the pure bamboo, sisal, and ax
bers composites exhibit an important tensile and exural
Chakkour et al. 19
Table 11. Mechanical properties of the relevant natural bers reinforced composites.
Nature of composite Observation Reference
Randomly oriented sisal/glass bers
reinforced polyester composites
Obtained composites by alternating three layers of treated sisal bers and
four layers of glass bers exhibit the highest tensile and exural strengths
167
Kenaf bers reinforced polyvinyl butyral
(PVB) composites
Composites with 45/45° oriented bers exhibit the highest impact
resistance compared to the case of 0/90°
168
Poplar wood PW reinforced high-density
polyethylene HDPE
Decrease of exural properties from 64 MPa for longitudinal bers reinforced
composites to 34 MPa for composites with transversal bers
188
Radiata pine RP reinforced high-density
polyethylene HDPE
Decrease of exural properties from 48 MPa for longitudinal bers reinforced
composites to 33 MPa for composites with transversal bers
188
Rice husk RH reinforced high-density
polyethylene HDPE
Decrease of exural properties from 39 MPa for longitudinal bers reinforced
composites to 28 MPa for composites with transversal bers
188
Randomly oriented kenaf bers with silica
reinforced epoxy composites
The highest compressive and exural strengths are obtained for composites
with 2 vol% of silica
189
Undirectional hemp bers reinforced epoxy
composites
The obtained exural strength at a mass fraction of 35% equal to 128.3 MPa
22
Alkali-treated coir bers reinforced epoxy
composites
The tensile, compressive and exural strengths were respectively 23.8 MPa,
2.98 MPa, and 40.4 MPa
169
Sisal bers reinforced epoxy composites The obtained values are equal to 180.45 MPa, 46.5 MPa, and 191.37 MPa for
the tensile, compressive, and exural strengths respectively
170
Hemp bers reinforced epoxy composites Specimens exhibit a tensile strength of about 55 MPa
171
Hemp bers reinforced epoxy composites Specimens exhibit a tensile strength of about 63.12 MPa
172
Hemp bers reinforced polyester
composites
Specimens exhibit a tensile strength of about 58.20 MPa
172
Flax bers reinforced epoxy composites The obtained tensile strength is almost 117.37 MPa
172
Flax bers reinforced polyester composites The obtained tensile strength is almost 109.80 MPa
172
Bidirectional jute bers laminates reinforced
epoxy composites
The incorporation of jute bers increases the tensile, impact, and exural
strengths of composites compared to neat matrices
191
Jute bers reinforced composites The optimum tensile strength is obtained for unidirectional composites
compared to 45° and 90° oriented bers within composites
115
Ramie and coir bers reinforced epoxy
composites
The highest tensile, impact, and exural strengths are obtained for ramie-
based epoxy composites
190
Banana bers reinforced synthesized bio-
resins
Banana bers increases the tensile strength, Youngs modulus, and exural
modulus by 15%, 12%, and 25% compared to neat matrix respectively
192
Coir bers reinforced epoxy composites The obtained maximum exural strength for composites with a ber length of
about 30 mm and a volume fraction of bers of 40%
193
Jute and bamboo reinforced epoxy
composites
Bamboo bers reinforced composites exhibit the highest tensile strength,
whereas the highest Youngs modulus is obtained by jute bers reinforced
epoxy
173
Jute bers reinforced epoxy biocomposites The tensile properties decrease by 50% at 180°C
174
Banana bers reinforced epoxy composites 50% banana and 50% epoxy-based epoxy composites exhibit the highest
tensile strength, and the optimum exural one is obtained for composites
with 60% of banana bers
175
Treated abaca bers reinforced epoxy
composites
The maximum impact strength is obtained for composites with a ber loading
of about 40%
176
Jute/kenaf/glass woven fabrics The glass/kenaf/jute hybrid laminates exhibit the highest tensile and exural
strengths reaching 47 MPa and 239 MPa, respectively
177
Sisal/glass hybrid laminates The addition of glass lamina increases the tensile and exural strengths of
composites
178
Jute/sisal/E-glass reinforced composites The incorporation of tungsten carbide additives increases the density and the
mechanical properties of composites
179
Hemp/jute hybrid composites The laminates with jute core layer H/J/H exhibit the highest tensile and
exural strengths
180
Jute/areca/glass bers hybrid composites The jute/areca/glass bers hybrid composites exhibit the highest mechanical
strengths compared to jute/glass and pure areca/glass laminates
181
20 Journal of Reinforced Plastics and Composites 0(0)
strengths ranging from 117 MPa to 180 MPa and between
191.4 MPa and 161.6 MPa, respectively. The lower values
are particularly obtained for radiata pine rice husk, coir, and
ramie bers composites, and could be attributed to the small
cellulose content and microbrillar angle of the reinforce-
ments. It can be clearly stated that the incorporation of glass
layers enhances the composites properties. Indeed, the
tensile strength reaches 168 MPa and 129 MPa for sisal/
glass and jute/kenaf/glass bers reinforced hybrid compo-
sites, respectively. The exural strength is 239 MPa and
184 MPa for jute/kenaf/glass and treated sisal/glass lami-
nates, respectively.
Theoretical modeling and prediction of the NFRCs
response to thermomechanical stimulus
Different theoretical and experimental approaches have
been proposed to understand the multiscale and multi-
physical deformation mechanism in NFRCs and partic-
ularly the debonding phenomenon of their matrix/ber
Figure 9. Summary of the tensile strength of the relevant NFRCs.
Figure 10. Summary of the exural strength of the relevant NFRCs.
Chakkour et al. 21
interface. In fact, some useful details of such issues are
addressed in the following sections.
Empirical models. Curve tting with polynomial, expo-
nential, and trigonometric functions are the main used
tools to model the thermomechanical behavior of NFRCs.
Gupta et al.
194
predicted the effect of Young modulus and
tensile strength of NFRPC using multiple non-linear re-
gressions techniques based on available experimental data.
In the case where coupling phenomena exist in the
composite, the model is extended by adding a new as-
sociated parameter. Epaarachichi et al.
195
used two sim-
plied empirical models of the mechanical behavior of
randomly distributed short bers in the matrix. In this
study, different short bers are embedded in an elastic matrix
and the whole composite is then subjected to mechanical tests.
The obtained results show that predictions using empirical
models overestimate the evolution of elastic modulus as
a function of the ber volume content. Furthermore, we
recommend to enrich these models by considering interfacial,
thermal, and moisture effects.
Finite element based models. One of the most commonly
used methods to simulate the multiphysical behavior of
materials and structures is the nite elements approach. It
predicts the response of NFRCs subjected to thermo-
mechnical stimulus on the basis of well-known parameters
such as Youngs modulus, Poisson ratio, density, thermal
conductivity, and specic heat capacity.
196
Palombini
et al.
29
used the nite element method to investigate effects
of structural components of the bamboo, such as paren-
chyma, sclerenchyma, and the large lumina of the xylem
and the phloem on the ber behavior. Sliseris et al.
197
implemented a nonlinear plasticity model to simulate me-
chanical fracture in ax short ber/polypropylene and ax
fabric/epoxy composites. Then, damage mechanisms such
as ber breakage initiated at the defects neighborhood,
damage nucleation in polymer matrix and the ber de-
bonding at ber/matrix interface are addressed using so-
phisticated numerical tools. Furthermore, results show good
agreement with experimental ones.
Analytical models. This type of models
198
involves rule of
mixtures (ROM), inverse ROM (IROM), Cox model,
Hirsch model Halpin-Tsai model, Kelly-Tyson model, and
Bowyer-Bader model and leads in general to mathematical
expressions of the NFRCs properties with respect of those
of the matrix and the bers separately. Cao et al.
199
used the
mentioned models to t the mechanical response of the
wood/ax/PP composite. IROM is the most accurate
method to successfully mimic the experimental mechanical
behavior of composites. Potluri et al.
31
used analytical
models of the mechanical behavior in case of pineapple
bers reinforced epoxy composite. On the basis of Chamiis
model and MROM, the predicted values of longitudinal,
transverse, and shear modulus are in agreement with ex-
perimental ones. The above underlined models are reviewed
in the following parts.
Rule of mixtures. Rule of mixtures are the simplest, ef-
fective methods that can be used. In fact, they are based on
matrices and bers contents in order to predict the behavior
of the resulting composites. One can distinguish between
parallel model called ROM and series model called IROM.
The latter results from the Voigts assumption where both of
matrices and bers exhibit the same strain. It is based on
Reussassumption in which the applied stress is equivalent
for both matrices and bers
30
In fact, the ROM is usually used
to predict the mechanical properties of composites such as the
tensile and exural strengths. Equations ((2), (3)) and ((4), (5))
show the expression of the strength and the modulus of the
composite following ROM and IROM models, respectively.
σcV¼ασ1V1þσ2V2(2)
EcV¼αE1V1þE2V2(3)
σc
V¼σ1σ2
σ1V2þσ2V1
(4)
Ec
V¼E1E2
E1V2þE2V1
(5)
where V is the volume fraction of ber and matrix (V=V
1
+
V
2
), σis the strength, and E the modulus. The subscripts 1,
2, c refer to the ber, matrix, and composite, respectively.
The parameter αis a harmonic coefcient which depends on
the orientation of bers within matrix.
30,200
The Hirsh model. The Hirsh model is considered as
a combination of parallel and series rule of mixtures ROM.
The elastic moduli and tensile strength of the resulting
composite can be dened using the equations below.
30,133
σcV¼αðσ1V1þσ2V2Þþð1αÞσ1σ2
σ1V2þσ2V1(6)
EcV¼αðE1V1þE2V2Þþð1αÞE1E2
E1V2þE2V1(7)
Halpin-tsai model. Regarding this model, it considers the
matrix to ber aspect ratio instead of the combination of
components content. It is usually used to determine the
mechanical behavior of continuous and discontinuous
oriented ber-reinforced composites. The following equa-
tions (8) and (9) show the strength and modulus of the
composite following Halpin-Tsai model.
30,133
σc¼σ21þBηV1
1ηV1(8)
Ec¼E21þBηV1
1ηV1(9)
22 Journal of Reinforced Plastics and Composites 0(0)
where
η¼
σ1
σ21
σ1
σ2þB¼
E1
E21
E1
E2þB(10)
B¼2L
D(11)
where L is the ber length and D its diameter.
Bowyer and Baders model. This model is developed to
analyze the tensile strength and elastic moduli of compo-
sites.
133,201
The following equations (12) and (13) are given
in terms of K1 which corresponds to the ber orientation
(ranging from 0 to 1) and K
2
to the ber length factor given
by equation. (14)
σc
V¼σ1K1K2V1þσ2V2(12)
Ec
V¼E1K1K2V1þE2V2(13)
K2¼lc
2lfor l <lc(14)
where lis the ber length, l
c
the critical length, and σand E
are the tensile strength and elastic modulus, respectively.
Self-consistent and Mori-Tanaka models. These ap-
proaches are based on the Eschelby inclusion problem.
202
Let e be the deformation eld in the volume V and E the
macroscopic deformation. The localization tensor A dened
by e(x)=A(x)Efor x2V, veries < A>V¼I, where Iis the
identity tensor. The expression of the homogenized stiffness
tensor C
hom
is given by
Chom ¼<CA >V(15)
If the matrix (Subscript 0) reinforced by several in-
clusions (subscript i ranging from 1 to n), the expression of
the homogenized stiffness tensor C
hom
is given by
Chom ¼X
n
i¼1
fiCiAi¼C0þX
n
i¼1
fiCiC0Ai(16)
where f
i
is the volume fraction of inclusion i (ratio of its
volume to the volume of the VER), C
i
the stiffness
tensor, and A
i
the deformation localization tensor of
inclusion idened by A
i
=< A>V
i
,whereV
i
is the
inclusion volume.
Self-consistent model. In this model, all reinforcements
are assumed to be immersed in the equivalent homogeneous
volume with an effective property described by C
SC
.
The deformation localization tensor for the re-
inforcement i is written using the Eschelby tensor Si
Esh
203
Ai¼hIþSi
EshCSC 1CiCSC i1
(17)
By replacing in equation (16), we obtain the expression
of the stiffness tensor C
SC
CSC ¼C0þX
n
i¼1
fiCiC0hIþSi
EshCSC 1CiCSC i1
(18)
Mori-Tanaka model. In this model, it is assumed that each
inclusion is embedded in the matrix of the composite. When
a stress is applied to the composite, an average stress σ
0
is
developed within it, resulting in an average strain ϵ
0
.
204
In this model, the localization tensor A
i
is written as
follow
Ai¼Ti X
n
j¼0
fjTj!1
(19)
Where
Ti¼hIþSi
EshC01CiC0i1
(20)
By replacing in equation (16), we obtain the expression
of the stiffness tensor C
MT
CMT ¼C0þX
n
i¼1
fiCiC0Ti X
n
j¼0
fjTj!1
(21)
Fracture based models. According to these models, crack
nucleation and propagation through the composite bulk or
along the interfacial area are predicted with regard to the
loading characteristics and the material fracture toughness.
Pupur et al.
205
used a shear lag model of wood bers
combined with PLA or Tencel matrices to predict the effect
of ber/matrix weight content on the elastic modulus and
the mechanical strength of the composite. The results were
compared with experimental data and reveal that in wood
bers reinforced composites, the governing mechanism
changes in material stiffness and strength, is related to the
interaction between the bers on one side and the interaction
between the bers, and the matrix including porosities
effects on the other side. Keck and Fulland
206
investigate
the crack path through ax reinforced composites and
fracture behavior of several compact tensile specimens
corresponding to different bers orientations and volume
fraction. The obtained results show that the latter strongly
induce changes in the crack path. Indeed, a large ber
contents results in a fast crack path growth along the ber.
Cohesive models. In fracture mechanics, cohesive zone
models allow to predict the crack nucleation and propa-
gation via a constitutive law giving the tensile normal stress
as a function of the relative displacement between the crack
lips.
207
Beakou and Charlet
208
used a bilinear cohesive zone
law to model the bundle strength in a ax ber-reinforced
Chakkour et al. 23
composite. Results show good agreement with the exper-
imental data pointing out the efciency of the cohesive
models of the matrix/ber interface to model the fracture
properties of the bundles. Hbib et al.
209
investigated the
interfacial damage in a hemp reinforced composite using 1D
interface cohesive elements. A linear correlation is evoked
between the damage and the interfacial stiffness criteria and
evidences the great inuence of the interface properties on
the damage evolution.
Hydric and thermal properties of natural
bers reinforced composites (NFRCs)
Hydric properties
The hydrophilic structure of natural bers is responsible for
the high moisture absorption of biocomposites. Several
studies revealed the water diffusion kinetic in some selected
biocomposites.
Chandrakanta Mishra et al.
210
studied the moisture
sorption behavior of kenaf/glass polyester hybrid composite
with different stacking sequences. They identied the
distinctive parameters such as: diffusion coefcient, (D
x
),
moisture content at equilibrium (M
m
), and the mechanism of
water transport. They found that composites made with
glass bers only revealed low moisture gain (2.314%),
whereas composites made with kenaf bers only shows high
moisture gain of about (10.560%). It is observed that the
water absorption of all specimens follows the mechanism of
Fickians diffusion.
210
They also noticed that there is a 51%
reduction of tensile strength for specimens made of kenaf
bers only (from 65 MPa to 30 MPa). However, the in-
corporation of glass bers was found to restrict the re-
duction to 37%. After drying samples, the obtained results
showed that 97% of tensile strength is recovered in case of
composites made with glass bers only, whereas in case of
kenaf/glass hybrid composite, it is limited to 87.22%.
Venkatesha B.K et al.
211
investigated the moisture ab-
sorption behavior of woven bamboo/glass ber-reinforced
epoxy hybrid composites with different laminate orienta-
tion. The volume fraction of bers within composite
specimens ranges between 31.71% and 32.52%. The
specimens are immersed in different environmental con-
ditions: Distilled and saline waters. They found that the
moisture absorption rate in case of saline water is less than
that of distilled one. This was explained by the accumu-
lation of NaCl ions in the ber surface immersed in saline
water which prevents further diffusion of moisture. The
obtained results revealed that the high values belonged to
composites with 60° laminate and showed 7.4% and 5.5%
for specimens immersed in distilled water and seawater
conditions, respectively.
211
They also noticed that the
tensile and exural strengths decreased after being im-
mersed in the solutions. The maximum reduction in these
properties is observed in case of saline water environment.
This is explained by the physical damage or chemical de-
terioration that specimens exhibited.
211
Chittaranjan Deo
and Acharya
212
investigated the moisture adsorption be-
havior of lantana camara bers reinforced epoxy composites
at different ber contents and environmental conditions
(steam, saline water and sub-zero temperature). They found
that the moisture adsorption increases by increasing the ber
content within composites. The moisture uptake is high for
composites immersed in steam environment. It is noticed
that moisture diffusion within all composites followed
Ficks law.
212
The Figure 11 shows the typical water ab-
sorption of cellulosic banana bers reinforced compo-
sites
213
at different bers contents and reveals that the
equilibrium moisture content of composites increases with
respect of bers content. The Table 12 gives a summary of
the hydric properties of the most relevant natural bers
reinforced polymer composites. Figure 12 shows the typical
evolution of moisture absorption of some NFRCs as
a function of time. One can observe that the moisture
diffusion curves of cellulosic bers composites are char-
acterized by a rapid diffusion rate during the rst 60 h before
reaching the saturation stage after 200 h of water immersion.
The current research studies revealed that the chemical
treatments enhance the dimensional stability of bers,
improve the interfacial adherence and decrease partially the
moisture absorption. However, these improvements are not
sufcient to integrate them in wet applications. The future
developments may focus on the integration of nanoparticles,
namely Mmntmorillonite nano-clays, nano silicone SiO
2
,
titane TiO
2
, and carbon nanotubes. Indeed, the preliminary
studies showed that the integration of nanoparticles de-
creased signicantly the moisture absorption of
Figure 11. Typical moisture absorption curves of cellulosic
bers composites.
24 Journal of Reinforced Plastics and Composites 0(0)
nanocomposites, leading to an increase in the mechanical
performance.
214221
Moreover, many studies investigated the effects of water
on the mechanical behavior of biocomposites, but still now,
few researchers are interested on the effect of relative hu-
midity. It is very important to conduct research studies to
further understand the hygromechanical behavior of ad-
vanced biocomposites.
Recent works revealed a lack on analytical models ad-
dressing the prediction of hydric properties. It will be ef-
fective to implement micromechanical models such as
ROM, IROM, and Hirsh to predict the equilibrium moisture
Table 12. Hydric properties of the relevant natural bers reinforced composites.
Nature of composite Observation Reference
Kenaf/glass reinforced polyester hybrid
composite
The water absorption of all specimens follows the mechanism of Fickians
diffusion. The composites made from kenaf bers only exhibited high
moisture absorption
210
Woven bamboo/glass ber-reinforced epoxy
hybrid composites
The moisture absorption rate in case of saline water is less than that of
distilled water
211
Lantana camara bers reinforced epoxy
composites
The moisture adsorption increases by increasing the ber content within
composites. It is higher for composites immersed in steam environment
212
Non-woven kenaf bers reinforced polyester
composites
Alkali treatment enhances the water absorption resistance of composite
specimens
24
Bamboo bers reinforced polyester composites At high temperatures, the moisture gain of specimens decreases
25
Randomly oriented oil palm empty fruit bunch
bers reinforced epoxy composites
The moisture gain increases by increasing the ber content in composites
222
Kenaf bers reinforced unsaturated polyester
composites
The temperature increases the diffusion kinetic
223
Bagasse bers reinforced epoxy composites The moisture uptake decreases from 12% to 8% after sodium hydroxide
treatment
224
Jute/sisal reinforced epoxy hybrid composites The water absorption decreased after chemical treatment. The optimal
moisture gain properties are related to 50sisal/50jute composites
225
Bamboo/sisal reinforced polyester hybrid
composites
The composite that possess higher content of bamboo bers display lower
water absorption
226
Figure 12. Summary of the moisture absorption of the relevant NFRCs.
Chakkour et al. 25
content and the moisture diffusivity in the composite based
on the constituentsproperties.
Thermal properties
This section aims to develop the reported works of the
thermal behavior of the relevant natural bers reinforced
composites (NFRCs). Indeed, both their thermal conduc-
tivity and stability are addressed in the following sections.
Thermal conductivity. Nowadays, there is a signicant de-
mand for new and improved housing, which may lead to
increased greenhouse gas emissions. To address these is-
sues, many studies have been conducted to improve the
global sustainability by incorporating natural bers. Ac-
tually, Idicula et al.
227
studied the thermal properties of
banana/sisal reinforced polyester hybrid composites. An
amount of bers are immersed in 10% NaOH and 5%
PSMA in toluene for 1 h and half an hour respectively. The
obtained results revealed that the incorporation of bers
enhances the thermal conductivity of composites speci-
mens. It decreases from 0.153 W.m
1
.k
1
to
0.140 W.m
1
.k
1
for 20% and 40% (by volume) ber
content composites. In contrast, the thermal diffusivity and
specic heat did not have a signicant variation. In case of
40 vol.% ber content, the thermal conductivity increased
and reached 0.201 W.m
1
.k
1
and 0.213 W.m
1
.k
1
for
NaOH and PSMA treated composites. This was explained
by the formation of pores due to the partial removal of non-
cellulosic components.
227
Takagi et al.
228
investigated the
thermal conductivity of unidirectional abaca and bamboo
bers reinforced epoxy composites. For abaca composites,
the conductivity values equal to 0.298 W.m
1
.k
1
,
0.292 W.m
1
.k
1
and 0.273 W.m
1
.k
1
for neat epoxy,
13.4% and 24.9% ber content composite (by volume)
respectively. In case of bamboo composites, it increased
from 0.298 W.m
1
.k
1
to 0.317 W.m
1
.k
1
for neat epoxy
and 22.42% of bers volume fraction composites.
228
Bhaskar
et al.
229
investigated the potential of using natural bers in
building thermal insulation. They studied the thermal con-
ductivity in longitudinal and transverse direction for banana/
palmyra bers reinforced epoxy hybrid composites. They
showed that the incorporation of bers enhances the thermal
insulation of the material. It is found that the thermal con-
ductivity in the longitudinal direction decreased from
0.363 W.m
1
.k
1
to 0.251 W.m
1
.k
1
for neat epoxy and 20%
palmyra/20%banana reinforced hybrid composite, re-
spectively. In the transversal direction, it increased from
0.363 W.m
1
.k
1
to 0.247 W.m
1
.k
1
respectively. They
found that the optimal value is related to 30%palmyra/10%
banana epoxy composites. In this case, the conductivity
reached 0.246 W.m
1
.k
1
and 0.240 W.m
1
.k
1
in the lon-
gitudinal and transversal direction, respectively.
229
.Figure 13
shows the thermal conductivity of some NFRCs.
However, few researchers are nowadays focused on the
development of re resistance performances of NFRCs. It is
necessary to orient research studies in this area to expand
their utilization in aeronautic and marine industries. As
discussed in the previous sections, the integration of nano-
clays may be effective to further enhance the thermal
properties of biocomposites, due to their excellent re re-
sistance.
230232
It may also be important to address research
Figure 13. Summary of the thermal properties of some relevant NFRCs.
26 Journal of Reinforced Plastics and Composites 0(0)
on the thermo-hygromechanical behavior of biocomposites
due to the lack of data on this aspect.
Thermal stability. Krittirash et al.
182
have investigated the
thermal properties of kenaf/sisal bers reinforced bio-epoxy
hybrid composites and reported that the neat epoxy and pure
kenaf laminates are degraded at 313.33°C while the max-
imum degradation temperature is extended to 315.83°C for
pure sisal bers reinforced laminates. Vinod et al.
92
have
studied the effect of various chemical treatments on the
thermal properties of Morinda citrifolia bers and their
corresponding composites. The obtained results show that
silane treatment increases the onset degradation temperature
of bers from 248.44°C to 274.71°C which is explained by
the high ratio of cellulose weight related to the removal of
sensitive-hemicellulose components by the chemical
treatments. Besides, they reported that the highest thermal
expansion coefcient is obtained for raw bers reinforced
composites due to the presence of high amorphous content
in the bers, while the lowest value is related to the silane-
treated bers composites and reached 90.82 ppm°C-1 and
64.62 ppm°C-1, respectively. Sanjay et al.
93
have studied
the thermal properties of chicken feather/Ceiba Pentandra
llers reinforced bio-epoxy hybrid nanocomposites and
showed that these materials are suitable for applications up
to 310°C. Vinod et al.
94
reported that the chemical treat-
ments enhance the thermal stability of Muntingia Calabura
bark bers. Indeed, the corresponding onset degradation
temperatures are about 213°C, 268 °C, and 254 °C for raw,
silane and alkali-treated bers, respectively. Similarly,
Sumrith et al.
95
studied the effect of various chemical
treatments on the thermal stability of water hyacinth bers.
The corresponding degradation temperatures are 289.22 °C,
294.67
°
C, and 329.71 °C for raw, NaOH and silane-treated
bers, respectively.
George et al
159
have studied the effect of hybridization
on the thermal behavior of sisal/ramie/curaua bers hybrid
composites and reported that the addition of the ramie/
curaua bers enhances the thermal stability of the corre-
sponding composites compared to pure sisal laminates.
Radzi et al.
236
have investigated the thermal stability of
roselle/sugar palm bers reinforced polyurethane compo-
sites. The reported results show that the incorporation of 75
wt.% of sugar palm bers improves the thermal stability of
the corresponding laminates. Teixeira et al.
237
have reported
that the addition of alkali and silane-treated curaua bers
enhances the degradation temperatures of composites. The
corresponding yield temperatures are 248°C, 273°C, and
293
o
C for pure polyester, alkali-treated and silane-treated
bers reinforced polyester composites, respectively.
237
They reported that the rst stage (at 200°C) corresponds
to the breakage of cross-links and the degradation of
polyester while the second stage occurs at 330°C and
corresponds to the decomposition of cellulose. Shanmu-
gasundaram et al.
238
have highlighted that the alkali
treatment improves the initial degradation temperature from
220°C to 250°C for untreated and treated mulberry bers
reinforced polyester composites. They found that the
temperature related to the decomposition of cellulose,
hemicellulose and lignin increases when increasing the -
bers contents. It reaches 348°C, 353°C, and 357°C for 5
wt.%, 10 wt.%, and 15 wt.% bers composites, respectively.
Jawaid et al.
239
have reported that the incorporation of jute
bers improves the thermal stability of jute/oil palm hybrid
composites. Indeed, the rst degradation temperature is
260°C, 283°C, and 288°C for pure oil palm composites, 1:1
Figure 14. Summary of the yield temperatures of the thermal stability of some natural bers reinforced composites NFRCs.
Chakkour et al. 27
jute/oil palm hybrid composites and pure jute bers com-
posites, respectively. Maou et al.
240
have studied the thermal
properties of data palm bers reinforced high-density
polyethylene HDPE and reported that the acid hydrolysis
treatment enhances the onset and maximum degradation
temperatures of composites when compared to NaOH and
silane treatments. The onset temperatures reach 230.33°C,
188.76°C, and 194.72°C for acid hydrolysis, alkali and
silane-treated bers composites. Fang et al.
241
noted that the
addition of 2 layers of jute bers to 5 layers of polylactic
acid PLA showed an increase of the onset degradation
temperature of jute bers reinforced polylactic acid PLA
laminates. The Jandas et al.
242
ndings showed that the
presence of banana bers enhances the stability yield
temperature of banana bers reinfoced PLA composites to
reach 113.75°C. The high decomposition temperature of
jute bers (almost 230°C) allows to improve the thermal
stability of PLA (30°C60°C). The thermal stability of
some NFRCs is reported in Figure 14.
Comparative conclusion. Figure 13 illustrates the thermal
conductivity of some NFRCs. It can be clearly seen that
bamboo bers epoxy composites show the highest value,
followed by abaca and abaca/palmyra epoxy composites.
The addition of carbon bers to Agave Americana bers
enhances the thermal insulation of the composites. In ad-
dition, hemp-polyurethane composites seem to be a good
insulator showing a thermal conductivity not exceeding
0.0407 W.m
1
.k
1
.Figure 14 shows that the chicken
feather/Ceiba Pentandra composites have the highest deg-
radation temperature of about 310°C. Indeed, the thermal
stability of these animal bers exceeds 350°C and permits to
enhance the resistance of composites to re. However, the
lowest degradation temperature is obtained for banana be AQ2rs
reinforced polylactic acid PLA composites which is due to
the poor thermal resistance of bioresin matrix. (Table 13)
Progress in mechanical properties
of nanocomposites
Biocomposites still suffer from the poor bers-matrix
bonding due to the hydrophilic nature of the reinforce-
ments.
32
As early mentioned, these limitations can be
overcome by using chemical or physical treatments, leading
to a better bers-matrix bonding.
123
However, several
works shed light on the effect of some nano-additives; for
instance, eggshell powder, titanium dioxide TiO
2
, nano-
clay, graphene, and silicon dioxide SiO
2
on the physical and
mechanical properties of nanocomposites.
243
Nanoclay llers/polymer composites
Clay nano-additives attract the attention of researchers due
to their great aspect ratio, low cost, and availability.
32
Montmorillonite (MMT), kaolinite, smectite, chlorite,
kenyaite, and ilerite are the most known types of clays.
32
Recent research works regarding natural bers/clays re-
inforced nanocomposites are summarized in the following
section. (Table 14)
Ramakrishnan et al.
244
have studied the effect of Cloisite
20A nano-clays on the morphological properties of jute
bers reinforced epoxy nanocomposites and reported that
bers-matrix interfaces bonding is improved after the in-
corporation of nano-additives. Hossain et al.
245
have in-
vestigated the effect of MMT K10 nano-clay on the water
Table 13. Thermal properties of the relevant natural bers reinforced composites.
Nature of composite Observation Reference
Banana/sisal reinforced polyester hybrid
composites
Decrease of the thermal conductivity by increasing the ber volume fraction
227
Unidirectional abaca bers reinforced
epoxy composites
The incorporation of bers decreased the thermal conductivity from 0.298 to
0.273 W.m1. k1 for neat epoxy and 24.9% ber loading composite (by
volume) respectively
228
Unidirectional bamboo bers reinforced
epoxy composites
The thermal conductivity increased from 0.298 W.m1. k1to
0.317 W.m1.k1 for neat epoxy and 22.42% ber loading composites
228
Banana/palmyra bers reinforced epoxy
hybrid composites
The incorporation of bers enhances the thermal insulation of the material
229
Miscanthus natural ber-reinforced
geopolymer composites
The thermal conductivity increased by increasing the ber size and foaming
agent content
233
Bamboo bers reinforced polypropylene
composites
The thermal conductivity increases with respect of ber content, and reaches
0.1073 W.m1.k1 at 60 wt% ber loading composites
27
Agave americana/carbon ber hybrid
reinforced epoxy composites
The thermal conductivity of 10wt% treated composites is higher than that of
7.5wt% treated ones
28
Esculentus cyperus natura ber-reinforced
polypropylene composites
The thermal conductivity decreases as the ber volume content increases
234
Hemp bers reinforced polyurethane
composites
The thermal conductivity increases with respect of the ber loading. It is found
to be affected by water
235
28 Journal of Reinforced Plastics and Composites 0(0)
absorption and mechanical properties of woven jute bers
biopol nanocomposites. The ndings reveal that adding
clays decreases the drop in exural properties after water
absorption. Arulmurugan et al.
34
have reviewed the me-
chanical properties of jute bers reinforced polyester
nanocomposites at different MMT nano-clays contents and
found that the optimal mechanical and vibration properties
are obtained for 5 wt.% nano-clays contents. Deepak
et al.
246
have reported that the garamite nano-clays enhance
stresses transfer at the interface of coir bers reinforced
polyester composites. Similarly, Borba et al.
247
showed that
the incorporation of 5 wt.% MMT nano-clays increases the
mechanical properties of curaua bers reinforced styrene-
butadiene-styrene composites and reported that the ag-
glomeration of llers leads to the decrease of the mechanical
properties of composites for high ller contents ( > 5 wt.%).
In a similar vein, Borba et al.
247
have reported that the
incorporation of 10 wt.% of MMT nano-clays in curaua
bers reinforced polyester composites leads to a cluster
formation preventing effective load distribution. Saba
et al.
248
have studied the effect of organo-modied mont-
morillonite OMMT on the physical and mechanical prop-
erties of kenaf bers reinforced epoxy and showed that the
corresponding llers increase the density and the hardness
of composites. Venkatram et al.
249
have noted that the
addition of 3 wt.% of garamite nano-clays increases the
tensile, exural and impact strengths of sisal-epoxy nano-
composites. Haq et al.
33
have studied the effect of 1.5 wt.%
of Cloisite 30B nano-additives on the mechanical properties
of hemp reinforced polyester hybrid composites. The ob-
tained results reveal that the tensile modulus and strength
increase by 6% and 20%, respectively. Prasad et al.
250
have
reported that the tensile strength and modulus of wild cane
grass reinforced polyester increases by 6.3% and 18.3%
after the addition of 4 wt.% MMT nano-clays, respectively.
Similarly, Wang et al.
251
showed that the tensile strength,
tensile modulus, exural strength, and exural modulus of
ax bers reinforced epoxy composites increase by 14.3%,
7%, 20.7%, and 13.6% after the incorporation of 1.3 wt.%
of organo-modied OMMT nano-additives, respectively.
Biswal et al.
252
have investigated the effect of Cloisite 20A
nano-clays on the mechanical properties of pineapple bers
reinforced polypropylene composites, manufactured using
compression molding. The obtained data reveal that the 3
wt.% ller reinforced nanocomposites exhibit the optimal
mechanical properties. The tensile strength, exural
strength, tensile modulus and exural modulus are im-
proved by 20%, 24.3%, 45.6%, and 38.57%, respectively.
Samariha et al.
253
have studied the inuence of Cloisite 30B
addition on the mechanical properties of bagasse bers
reinforced high-density polyethylene HDPE composites
manufactured using extrusion injection molding method.
Table 14. Yield temperatures of the thermal stability of the relevant natural bers reinforced composites.
Nature of composite Observation Reference
Kenaf/sisal reinforced bio-epoxy The degradation temperature of sisal laminates is higher when compared to neat
epoxy and pure kenaf bers laminates
182
Morinda citrifolia bers reinforced
bio-epoxy
Silane treatment increases the thermal stability of the bers and the corresponding
composites compared to alkali treatment
92
Chicken feather/ceiba pentandra
reinforced bio-epoxy
The incorporation of chicken feather llers increases the thermal stability of the
composite reaching 310°C
93
Muntingia calabura bark bers
reinforced green epoxy
Silane and alkali treatments enhances the thermal stability of the bers and
composites when compared to raw bers
94
Sisal/ramie/curaua bers composites The addition of ramie/curaua bers increases the thermal stability of the hybrid
composites
159
Sugar palm/roselle bers reinforced
polyurethane
The maximum degradation temperature is obtained for 75.% sugar palm bers
loadings
236
Curaua bers reinforced polyester
composites
Silane treatment improves the degradation temperature of composites from 273°C
to 293°C
237
Mulberry bers reinforced polyester
composites
The alkali treatment improves the thermal stability of composites from 220°C to
250°C
238
Jute/oil palm hybrid composites The addition of jute bers increases the thermal stability of bers of the hybrid
composites from 260°C to 283°C
239
Data palm bers reinforced
polyethylene HDPE
The acid hydrolysis treatment enhances efciently the onset degradation
temperature of composites when compared to alkali and silane-treated bers
composites
240
Jute bers reinforced polylactic acid
PLA
The addition of 2 layers of jute to 5 layers of PLA increases the thermal stability of
composites compared to other congurations
241
Banana bers PLA composites The presence of 30 wt.% banana bers enhances the decomposition temperature of
PLA composites
242
Chakkour et al. 29
Samariha et al. ndings
253
show that the 2 wt.% ller re-
inforced composites exhibit the highest tensile and exural
strengths while the optimal tensile and exural modulus are
obtained for 4 wt.% loaded composites. Shahroze et al.
254
investigated the tensile, exural, and impact properties of
sugar palm bers/organo-modied montmorillonite OMMT
reinforced polyester composites processed through hot
pressing method. The composites reinforced with 4 wt.%
OMMT exhibited the optimal tensile properties while the
high exural and impact strengths are obtained at 2 wt.%
OMMT content. Table 15 shows the mechanical properties
of various nano-clays reinforced biocomposites.
Cellulose bers cellulose nanobers and crystals
cellulose nanocrystals composites
polymer nanocomposites
There is a growing interest in the production of cellulose
nanocrystals composites (CNC) due to their attractive
mechanical characteristics, small density, biodegradability
and renewable nature.
255,56
CNC can be isolated from
different agricultural sources such as wood, cotton, bamboo,
jute, and sisal bers using various chemical pretreatments to
fully/partially remove the amorphous hemicellulose and
lignin components from the lignocellulosic bers.
256
Table 15. Mechanical properties of the relevant natural bers reinforced nanocomposites.
Nature of composite Additives
Tensile
strength (MPa)
Tensile
modulus (GPa)
Flexural
strength (MPa)
Flexural
modulus (GPa) References
Basalt bers epoxy 2 wt.% MMT 325 15.9 306 18.49
265
Curaua bers-polyester 2.5 wt.% Organophilic
clay
36 32.55
266
Banana bers epoxy 3 wt.% Cloisite 173 10 88 8.1
267
Coccinia indica bers epoxy 3 wt.% Cloisite 30B 38.29 92.77
268
Jute bers epoxy 5 wt.% Cloisite 20A 103.05 1.29 162.8 2.8
244
Jute bers-polyester 1.5 wt.%
(MMT+Eggshell)
29.5 39.52 ——
269
Sugar palm bers-
polyester
4 wt.% OMMT 24.56 3.68 68.12 3.78
254
Jute bers-polyester 5 wt.% MMT 40.38 234.98
34
Sisal bers polypropylene 5 wt.% Cloisite 30B 55.95 1.7 ——
270
Curaua bers-polyester 2.5 wt.% organophilic
clay
36 32.55
266
Curaua bers-polyester 2.5 wt.% organophilic
clay
36 32.55
266
Jute/coir bers-polyester 3 wt.% garamite 43 ——
246
Curaua bers-styrene-
butadiene-styrene
2 wt.% Cloisite 10A 7.8 4.8 ——
247
Hemp bers-polyester 1.5 wt.% Cloisite 30B 24 6 ——
33
Wild cane grass bers-
polyester
4 wt.% MMT 99.57 2.26 221.61 4.19
250
Flax bers epoxy 1.3 wt.% OMMT 87.5 7.55 140 6.2
251
Pineapple bers
polypropylene
3 wt.% Cloisite 20A 45.14 6.45 65.01 4.46
252
Poplar wood bers
polypropylene
3 wt.% MMT 41.7 57.5
271
Sisal bers epoxy 5 wt.% Cloisite 30B 57 25 ——
272
DCNC-epoxy 3.5 wt.% DCNC 72 2.25 ——
257
CNF-epoxy 15 vol.% CNF 109 5.9 ——
258
CNF-epoxy 0.1 wt.% CNF 6.2 0.42 ——
259
CNF-polyester 20 vol.% CNF 150 1.15 125
261
CNF-phenol formaldehyde 12 wt.% CNF 17 38
262
CNF-phenol formaldehyde 10 wt.% CNF 18 36
262
CNF-phenol formaldehyde 8 wt.% CNF 16 33
262
CNF-phenol formaldehyde 6 wt.% CNF 15 28
262
CNF-phenol formaldehyde 4 wt.% CNF 13 25
262
CMF-melamine
formaldehyde
5 wt.% CMF 142 16.6 ——
264
30 Journal of Reinforced Plastics and Composites 0(0)
Actually, they are extracted from the pretreated bers using
strong acid hydrolysis leading to the dissolution of amor-
phous regions.
255
During this process, the hydronium ions
penetrate the amorphous regions and cause the cleavage of
glycosidic bonds resulting in the separation of individual
cellulose crystallites.
However, cellulose nanobers CNFs are usually ex-
tracted from the pretreated bers through mechanical
processes such as high-pressure homogenization, micro
uidization, ball milling, and grinding.
255
Afterward, it is
reported to dry the obtained CNFs and CNCs by using
several methods such as evaporation of solvent, drying by
iyophilization or using supercritical state uids.
255,256
Wang et al.
257
have studied the addition effect of
chemically modied CNC with dodecenyl succinic anhy-
dride to obtain hydrophobic CNC (DCNC) on the me-
chanical properties of cellulose nanocrystalls reinforced
epoxy composites. The reported results reveal that the
presence of 3.5 wt.% of DCNC enhances the tensile
strength, Youngs modulus, and strain at break by 82%,
21%, and 198% compared to the neat epoxy . This is ex-
plained by the homogeneous dispersion of nanollers
DCNC in the matrix and the strong interactions between the
composite components. The values reach 72 MPa,
2.25 GPa, and 7.2% for tensile strength, Youngs modulus,
and strain at break, respectively. Ansari et al.
258
have in-
vestigated the inuence of the incorporation of cellulose
nanobers CNF mixed with acetone/epoxy/amine on the
mechanical performance of composites and revealed that
the nanocomposites exhibit a good tensile strength and
Youngs modulus at 15 vol.% bers contents, reaching
109 MPa and 5.9 GPa, respectively. AL-Turaif
259
have
found that the addition of 0.1 wt.% of cellulose nanobers
(CNF) improves the tensile strength, strain at break,
Youngs modulus and toughness by 121%, 73%, 64%, and
300% which are larger than the improvements by any other
reported nanocomposites reported in the literature.
260
Gao et al.
261
found that the addition of 20 vol.%
silane-treated cellulose nanobers (CNF) increases the
tensile strength, exural strength, shear strength, and
Youngs modulus of polyester nanocomposites by
117.7%, 38.4%, 38.7%, and 27.6%, respectively. They
are foun to be 150 MPa, 125 MPa, 123 MPa, and
1.15 GPa, respectively. It is reported that 12 wt.% of
cellulose nanobers (CNF) induces a signicant increase
of 142% in the tensile strength, 280% in exural strength,
and 133% in impact strength for the Phenol Formalde-
hyde nanocomposites.
262
Theyexhibit17MPa,38MPa,
and28kJ/m
2
in tensile strength, exural and impact
strengths, respectively. Nakagaito et al.
263
reported that
the exural strength and Youngs modulus of 25 layers of
cellulose microbers reinforced Phenol Formaldehyde
compositesare370MPaand16GPa,respectively.The
obtained results by Marielle et al.
264
show that the tensile
strength and Youngs modulus of cellulose microbers
(CMF) reinforced melamine formaldehyde (MF) com-
posites increase by 50% and 30% at 5 wt.% bers content,
when compared to neat matrix, respectively.
Applications and future trends of NFRCs
Because of their good mechanical properties, lightweight,
and sound attenuation, natural bers based composites have
been utilized in the automotive, building, sport equipment,
and biomedical industries
37,65,273
(see Table 16). The
adoption of those bers in several industries is encouraged
by their low cost and lightweight nature. They are usually
used to reinforce polyester and polypropylene matrices.
35
Table 16. Applications of natural bers in different elds.
Fiber Applications References
Coir Containers and boxes, car seat covers, car door mats, helmet, car roof, sports applications
276,281,284
Flax Racing bicycle, packaging, car headliners, car side and back walls, car armrests, car trunk trim, car inner door
panel, car door trim panels, tissue engineering application
35,273,280,285
Hemp Cases for musical instruments, car side and back walls, car headliners, car armrests, car trunk trim, car inner door
panel, femur bone application
273,281,285
Kenaf Mobile phone casing, car door inner panel, car side and back walls, car headliners, car armrests, car trunk trim,
underbody shield applications
280,285
Wood Door and covered seatback panels in cars
286
Cotton Soundproong, insulation
286
Abaca Car under door panels
286
Jute Packaging, containers, cladding applications
281
Bamboo Construction, interior design (roof, stairs, and window outlines of houses) and furnitures
274,282
Sisal Car inner door panel, car door trim panels, scaffolds in medical eld
35,273
Coconut Seats of Mercedes a-class model
274
Bagasse Sustainable and thermal insulation materials
112
Banana Computer elements, door panels
112
Ramie Military applications
112
Chakkour et al. 31
However, these natural bers could not replace glass bers
in wet applications such as boats and kayaks because of
their high sensitivity to water.
37
The bast bers reinforced
composites are found to exhibit high tensile strength and
a small impact one in contrast to cotton bers. Sanjay et al.
showed that the combination of those two types of bers
could be benec for the high impact stressed applications
such as door panels of cars and safety helmets, combining
the high tensile properties generated by plants bers and the
important impact strength of cotton bers.
35
Nowadays, automotive industry is leader in the use of
natural bers composites especially German companies
such as Mercedes, Volkswagen, Audi, and BMW that took
the initiative to use natural bers in exterior as well as
interior applications such decking, window frames, and
headliners.
1,37
The rst attempt to use natural bers in
automotive industry was in 1941 by Henry Ford using ax
and hemp bers as reinforcement for petrochemical poly-
mers.
1
The rst commercial example of NFRC in the
automotive industry was the inner door panel of the 1999
S-Class Mercedes-Benz.
273
Indeed, it was made of 35% of
polyurethane PUR elastomer reinforced by 65% of hemp,
ax and sisal bers. On the other hand, Mercedes-Benz
developed jute bers reinforced epoxy composites in the
door panels of its E-Class vehicles in 1996. The company
used also coconut bers as reinforcement for rubber latex
composites for the seats of the A-class model.
274
Similarly,
Audi has launched the A2 car in 2000 where door trim
panels are made of mixed ax/sisal reinforced polyurethane
composite.
35
Ford used kenaf bers reinforced Poly-
propylene in the door panels of their Mondeomodel,
whereas, ax is used in oor trays.
274
For Opel, the mixture
of ax and kenaf bers is inserted in door panels for Vectras
model.
274
Volkswagen used cellulose to make the door
panels, seatbacks, boot-lid nish panels, and boot-liners of
all Passat, Golf, and Bora models. Holbery et al.
275
reported
that BMW group use a lot of natural bers reinforced
composites in its automobiles. The group used about 10,000
tons of natural bers in 2004. The BMW 7 series model is
composed of 24 kg of raw natural bers including ax and
sisal in the interior door lining panels, cotton in the
soundproong in addition to the wool and wood bers in the
seatback parts.
274,275
On the other hand, Toyota used Bagasse bers to re-
inforce used plastic polymers in the interior of cars.
273
Furthermore, the coir/polyester composites have been
also used to produce mirror casing, projector cover, helmet,
and roof.
276
It is worthy to note that 4000 tons of natural
bers reinforced composites have been used in 2008 as
interior pieces for automotive industries.
277
In 2009, all Peugeots, 207 came out with an external
mirror fabricated using 30% hemp bers.
277
Djamelhas
277
showed that the density of hemp and ax bers reinforced
composites is 20%30% much lower than glass bers
reinforced ones. In the works of Yang et al.,
278
it was
reported that a reduction of 25% of automotive weight
would enable us to avoid 100 billion kilograms of CO
2
emission annually.
In addition, the Cambridge industry inserted ax bers
polypropylene composites in its Chevrolet Impala model.
Moreover, plant-based bers have attracted the attention
of researchers for the insulation in building.
277
The mixture
cement/lignocellulosic bers could replace cement matrix/
Asbestos bers in several applications as wall coverings and
piping.
277
Because of their thermo-acoustic properties,
hemp wool may be a good substituent to the glass one.
277
In
addition, Van de Weyenberg et al.
279
have shown that sisal
reinforced composites can be used in several building ap-
plications such as structural building, facades and span
roong elements. Also, bamboo bers are used as re-
inforcement in the construction of low cost houses.
274
In addition to the construction and automotive industries,
natural bers composites can be implemented in other
applications including medical, pharmaceuticals, packag-
ing, bioenergy, and biofuels elds.
38
Currently, ax and
hemp bers are used in racing bicycle and musical in-
struments,
280
coir and jute in packaging and containers
281
and kenaf in mobile phone casing.
274,280
Besides, bamboo is
also used in interior design (roof, stairs and window outlines
of houses), furniture, and spring chairs. One can note that
Barajas Madrid international airport was built using bamboo
strips.
282
Table 16 shows the different applications of
natural bers.
Pending future developments, the application of natural
ber composites remains limited, mainly in the automotive
and construction sectors. Further efforts should be made to
disseminate their use in sectors such as marine, aeronautic,
and renewable energy. Figure 15 shows the different ap-
plications of natural bers reinforced composites.
Figure 15. Applications of natural bers reinforced
composites.
283
32 Journal of Reinforced Plastics and Composites 0(0)
Conclusion and perspectives
The good thermo-acoustic properties, mechanical proper-
ties, ecofriendly character, low cost, biodegradability, re-
newability make natural bers the best candidates to replace
non-renewable and expensive synthetic bers such as glass,
carbon, and kevlar ones. They are nowadays used as re-
inforcements for petrochemical matrices including ther-
moplastics and thermosets especially in automotive
industries, building as well as in medical and pharmaceu-
tical elds. However, the utilization of those natural bers
remains limited because of their sensitivity to different
parameters such as moisture adsorption, temperature in
addition to their incompatibility with conventional resins.
Several solutions have been presented in the literature but,
still now, biocomposites still suffer from the above limi-
tations. This paper has presented an up to date of different
parameters inuencing the properties of natural bers, used
matrices types to coat natural bers, ber-matrix interface,
mechanical, thermal, and hydric behaviors of natural bers
composites. The current critical review presents the recent
developments and applications of natural bers reinforced
composites and help to understand the issues for widespread
integration into more applications mainly for wet and re-
retardant targets. The possible solutions and future ori-
entations are discussed as well. A brief update of analytical
and FEM-based models has been presented to predict the
mechanical response of natural bers composites. On basis
of the current literature review and our own analysis, two
main trends that will address the future development of
NFRCs are discussed.
- Hybrid composites: One of the future trends that will help
NFRCs to overcome the lack of interfacial adhesion is to
combine different reinforcements in order to enhance the
properties of the nal composite. Using more than one
reinforcement permit to counter any limitations by adding
other relevant constituents. For instance, to counter the
hydrophilic behavior of natural bers, hybrid composites
are made of synthetic and natural bers. Instead of the
chemical treatment of bers, synthetic bers are hydro-
phobic and therefore would give the best matrix/bers
compatibility.
- Nanobiocomposites: Another new research attempt in the
NFRCs eld is the use of nano reinforcement particles. The
latter possess a large surface to volume ratio that allows to
enhance the interfacial adhesion. Besides, nanoparticles such
as silicium oxide SiO
2
,titaneoxideTiO
2
, calcium carbonate
CaCO
3
, zeolite, and graphene may be the best candidates to
reinforce biocomposites. The preliminary studies reveal that
their incorporation decreases the void content and enhances the
stress transfer between bers and matrix. However, few studies
have addressed their corresponding water absorption and
ame retardancy properties. Many efforts should be done to
understand the hydromechanical behavior of these nano-
biocomposites to extend their applications in wet conditions. In
fact, the hybridization of nanollers would be an attractive
solution to improve the ammability and the water absorption
behavior of these bio-sourced materials. In addition, pre-
liminary investigations reveal that cellulose bers and crystals
nanocomposites exhibit higher mechanical properties when
compared to macro and micro-cellulose composites; however,
they still suffer from limitations such as high moisture ab-
sorption, poor wettability, and incompatibility with most
polymeric matrices. For these reasons, scientists should ad-
dress research studies to the issues related to the modication
of nanocellulose to produce high-performance nano-
composites with a better hydromechanical and re properties.
Furthermore, the main issue of the NFRCs behavior
modeling remains undoubtedly 3D multiscale predictions
which consider the mechanical properties of the matrix, the
bers and their interfaces respectively. This allows to size
up the whole composite material and ensures a good in-
terfacial adhesion by selecting compatible matrix and bers.
In addition, cohesive models enhance the knowledge re-
garding the prediction of crack initiation and propagation
under complex loadings as well as fatigue conditions. In
critical applications such as aerospace and aeronautics, it is
necessary to develop models for NFRCs that integrate their
thermomechanical dynamic properties. However, it is sure
that a big progress must be done in modeling the behavior of
NFRCs in order to nely characterize interfacial properties
and damage pronostic but also to understand the complex
geometrical structure of the natural bers and relate it to the
mechanical, hydric, and thermal coupling effects on the
state of health of the composite material.
Acknowledgements
Mouad Chakkour, M. O, I.K and M.B acknowledge the In-
ternational University of Rabat. T.B acknowledges the Lorraine
University.
Declaration of conicting interests
The author(s) declared no potential conicts of interest with re-
spect to the research, authorship, and/or publication of this article.
Funding
The author(s) received no nancial support for the research, au-
thorship, and/or publication of this article.
ORCID iDs
Mouad Chakkour https://orcid.org/0000-0002-4148-369X
Mohamed Ould Moussa https://orcid.org/0000-0001-8944-
1467
References
1. Gholampour A and Ozbakkaloglu T. A review of natural
ber composites: properties, modication and processing
techniques, characterization, applications. J Mater Sci 2019;
■■■:■■■.
Chakkour et al. 33
2. Ramesh M, Deepa C, Rajesh Kumar L, et al. Life-cycle and
environmental impact assessments on processing of plant
bres and its bio-composites: a critical review. J Industrial
Textiles 2020; 125: ■■■.
3. Tarannum N, Pooja K and Khan R. Preparation and appli-
cations of hydrophobic multicomponent based redispersible
polymer powder: a review. Construc Build Mater 2020; 247:
118579.
4. Rangappa SM and Siengchin S. Moving towards biober-
based composites: Knowledge gaps and insights. Express
Polym Lett 2022; 16: 451452.
5. Zakikhani P, Zahari R, Sultan MTH, et al. Extraction and
preparation of bamboo ber-reinforced composites. Mater
Design 2014; 63: 820828.
6. Bourmaud A, Beaugrand J, Shah DU, et al. Towards the
design of high-performance plant bre composites. Prog
Materials Science 2018; 97: 347408.
7. Balaji A, Karthikeyan B and Raj CS. Bagasse berthe
future biocomposite material: a review. Int J Cemtech Res
2014; 7: 223233.
8. Faruk O, Bledzki AK, Fink H-P, et al. Biocomposites re-
inforced with natural bers: 20002010. Prog Poly Sci 2012;
37(11): 15521596.
9. Gurunathan T, Mohanty S, Nayak SK, et al. A review of the
recent developments in biocomposites based on natural bres
and their application perspectives. Compos Part A Appl Sci
Manuf 2015; 77: 125.
10. Madhu P, Sanjay MR, Senthamaraikannan P, et al. A review
on synthesis and characterization of commercially available
natural bers: part-1. J Nat Fibers 2018; 16(4): 113.
11. Sanjay MR, Siengchina S, Parameswaranpillai J, et al.
A comprehensive review of techniques for natural bers
as reinforcement in composites: preparation, processing
and characterization. Carbohydr Polym 2019; 207:
108121.
12. Rowell RM. Natural bres: types and properties. In: Pick-
ering K (ed). Properties and performance of natural bre
composites. Amsterdam: Elsevier, 2008, pp. 366.
13. Yao F, Wu Q, Lei Y, et al. Thermal decomposition kinetics
of natural bers: activation energy with dynamic ther-
mogravimetric analysis. Polym Degrad Stab 2008; 93(1):
9098.
14. Joseph S, Koshy P, Thomas S, et al. The role of interfacial
interactions on the mechanical properties of banana bre
reinforced phenol formaldehyde composites. Compos Inter
2005; 12: 581600.
15. London OX. A decade of observations of a Guadua an-
gustifolia plantation in Colombia. J Am Bamboo Soc 1998;
12: 3742.
16. Yashas Gowda TG, Rangappa SM, Parameswaranpillai J,
et al. Natural bers as sustainable and renewable resource for
development of eco-friendly composites: a comprehensive
review. Front Mat 2019; 6: ■■■.
17. Sanjay MR and Siengchin S. Exploring the applicability of
natural bers for the development of biocomposites. eX-
PRESS Polymer Letters 2021; 15(5): 193.
18. Chin SC, Tee KF, Tong FS, et al. Thermal and mechanical
properties of bamboo ber reinforced composites. Mater
Today Commun 2019; 23: 100876.
19. Campilho RDSG. Natural ber Composites. Boca Raton:
CRC Press, 2015.
20. Hemath M, Rangappa SM, Kushvaha V, et al. A compre-
hensive review on mechanical, electromagnetic radiation
shielding, and thermal conductivity of bers/inorganic llers
reinforced hybrid polymer composites. Polym Composites
2020; 41: 126.
21. Yashas Gowda TG, Sanjay MR, Subrahmanya Bhat K, et al.
Polymer matrix-natural ber composites: an overview. Co-
gent Eng 2018; 5: ■■■.
22. Boccarusso L, Carrino L, Durante M, et al. Hemp
fabric/epoxy composites manufactured by infusion
process: improvement of re properties promoted by
ammonium Polyphosphate. Composites Part B 2016; 89:
117126.
23. Thwe MM and Liao K. Effects of environmental aging on the
mechanical properties of bamboo glass ber reinforced
polymer matrix hybrid composites, composites part A. Appl
Science Manufacturing 2002; 33: 4352.
24. Mohd Nazarudin Z, Mohd Ariff J, Masitah Abu K, et al. The
effect of alkaline treatment on water absorption and tensile
properties of non-woven kenaf polyester composite. Adv
Mater Res 2013; 812: 258262.
25. Sugiman S, Dwi Setyawan P and Anshari B. Water ab-
sorption and impact strength of alkali-treated bamboo ber/
polystyrene-modied unsaturated polyester composites.
J Appl Sci Eng 2020; ■■■:920.
26. Sekar S, Shanmugam V, Kumar S, et al. Effects of water
absorption on the mechanical properties of hybrid natural
bre/phenol formaldehyde composites. Sci Rep 2021; ■■■:
■■■.
27. Wanga C, Qi Z, Lina T, et al. Predicting thermal conductivity
and mechanical property of bamboo bers/polypropylene
nonwovens reinforced composites based on regression
analysis. Int Commun Heat Mass Transfer 2020; 118:
104895.
28. Jani SP, Sajith S, Rajaganapathy C, et al. Mechanical and
thermal insulation properties of surface-modied agave
Americana/carbon bre hybrid reinforced epoxy composites.
Mater Today Proc 2020; 37: 16481653.
29. Palombini FL, Kindlein W, de Oliveira BF, et al. Bionics and
design: 3D microstructural characterization and numerical
analysis of bamboo based on X-ray microtomography. Mater
Characterization 2016; 120: 357368.
30. Cao Y, Wang W, Wang Q, et al. Application of mechanical
model for natural bre reinforced polymer composites.
Mater Res Innov 2014; 18: 354357. sup2.
31. Potluri R, Diwakar V, Venkatesh K, et al. Analytical model
application for prediction of mechanical properties of natural
ber reinforced composites. Mater Today Proc 2018; 5:
58095818.
32. Rajeshkumar G, Arvindh Seshadri S, Ramakrishnan S, et al.
A comprehensive review on natural ber/nano-clay re-
inforced hybrid polymeric composites: materials and tech-
nologies. Polym Composites 2021; 42: 115.
33. Haq M, Burgueño R, Mohanty AK, et al. Hybrid bio-based
composites from blends of unsaturated polyester and soy-
bean oil reinforced with nanoclay and natural bers. Com-
posites Sci Technol 2008; 68: 1516.
34 Journal of Reinforced Plastics and Composites 0(0)
34. Arulmurugan S and Venkateshwaran N. Vibration analysis of
nanoclay lled natural ber composites. Polym Polym
Composites 2016; 24: 507516.
35. Sanjay MR, Arpitha GR, Laxmana Naik L, et al. Applica-
tions of natural bers and its composites: an overview. Earth
environ sci, 2016; 7: ■■■.
36. AL-Oqla FM and Salit MS. Natural ber composites. Mater
Selection Nat Fiber Compo 2017; ■■■:2348.
37. Dicker MPM, Duckworth PF, Baker AB, et al. Green
composites: a review of material attributes and comple-
mentary applications. Compos Part A Appl Sci Manuf 2014;
56: 280289.
38. Peças P, Carvalho H, Salman Haz, et al. Natural bre
composites and their applications: a review. J Compos Sci
2018; 2: 66.
39. Mokhtar M. Characterization and treatments of pineapple
leaf bre thermoplastic composite for construction appli-
cation. Johor Bahru: Universiti Teknologi Malaysia, 2007.
40. Campilho RDSG. Recent innovations in biocomposite
products. In: Biocomposites for high-performance applica-
tions, 2017, pp. 275306.
41. Khalil HPSA, Bhat IUH, Jawaid M, et al. Bamboo bre
reinforced biocomposites: a review. Mater Des 2012; 42:
353368.
42. Rana AK, Mandal A, Bandyopadhyay S, et al. Short jute
ber reinforced polypropylene composites: effect of com-
patibiliser, impact modier and ber loading. Compos Sci
Technol 2003; 63(6): 801806.
43. Chandramohan D and Marimuthu K. A review on natural
bers. Int J Res Rev Appl Sci 2011; 8(2): 194206.
44. Ramamoorthy SK, Skrifvars M, Persson A, et al. A review of
natural bers used in biocomposites: plant, animal and re-
generated cellulose bers. Polym Rev 2015; 55(1): 107162.
45. Jawaid M and Khalil HPSA. Cellulosic/synthetic bre re-
inforced polymer hybrid composites: a review. Carbohydr
Polym 2011; 86(1): 118.
46. Rajeshkumar G, Arvindh Seshadri S, Devnani GL, et al.
Environment friendly, renewable and sustainable poly lactic
acid (PLA) based natural ber reinforced composites
a comprehensive review. J Clean Prod 2021; 310: 127483.
47. Karimah A, Ridho MR, Munawar SS, et al. A review on
natural bers for development of eco-friendly bio-composite:
characteristics, and utilizations. J Mater Res Technol 2021;
13: 24422458.
48. Shang S, Zhu L, Fan J, et al. Intermolecular interactions
between natural polysaccharides and silk broin protein.
Carbohydr Polym 2013; 93: 561573.
49. Satyanarayana KG, Arizaga GGC, Wypych F, et al. Bio-
degradable composites based on lignocellulosic bers an
overview. Prog Polym Sci 2009; 34: 9821021.
50. Summerscales J, Dissanayake NPJ, Virk AS, et al. A review
of bast bres and their composites. Part 1 bres as re-
inforcements. Compos Part A: Appl Sci Manuf 2010; 41:
13291335.
51. Azwa ZN, Yousif BF, Manalo AC, et al. A review on the
degradability of polymeric composites based on natural -
bres. Mater Des 2013; 47: 424428.
52. John MJ and Thomas S. Biobres and biocomposites.
Carbohydr Polym 2008; 71: 343364.
53. Cousins WJ. Youngs modulus of hemicellulose as related to
moisture content. Wood Sci Technol 1978; 12: 161167.
54. Fuentes C, Tran LQN, Dupont-Gillain C, et al. Wetting
behaviour and surface properties of technical bamboo bres.
Colloids Surf A: Physicochem Eng Asp 2011; 380: 8999.
55. Rangappa SM, Siengchin S, Parameswaranpillai J, et al.
Lignocellulosic ber reinforced composites: progress, per-
formance, properties, applications, and future perspectives.
polymer Composites 2022; 43: 645691.
56. Mohanty AK, Misra M, Drzal LT, et al. Surface modications
of natural bers and performance of the resulting J Mater Sci
biocomposites: an overview. Compos Inter 2001; 8: 313343.
57. Shibata M, Ozawa K, Teramoto N, et al. Biocomposites made
from short abaca ber and biodegradable polyesters. Mac-
romolecular Mater Eng 2003; 288(1): 3543.
58. Khalil HA, Bhat IUH, Jawaid M, et al. Bamboo bre re-
inforced biocomposites: a review. Mater Des 2012; 42:
353368.
59. Thyavihalli Girijappa YG, Mavinkere Rangappa S, Para-
meswaranpillai J, et al. Natural bers as sustainable and
renewable resource for development of eco-friendly com-
posites: a comprehensive review. Front Mater 2019; 6: 226.
60. Yan L, Chouw N, Jayaraman K, et al. Flax bre and its
composites areview.Composites Part B 2014; 56: 296317.
61. Shahzad A. Hemp ber and its composites a review.
J Compos Mater 2012; 46(8): 973986.
62. Nadirah WW, Jawaid M, Al Masri AA, et al. Cell wall
morphology, chemical and thermal analysis of cultivated
pineapple leaf bres for industrial applications. J Environ
Polymer Degra 2012; 20(2): 404411.
63. Di Bella G, Fiore V, Galtieri G, et al. Effects of natural bres
reinforcement in lime plasters (kenaf and sisal vs. poly-
propylene). Constr Build Mater 2014; 58: 159165.
64. Sanjay MR, Madhu P, Jawaid M, et al. Characterization and
properties of natural ber polymer composites: a compre-
hensive review. J Clean Prod 2017; 172: 566581.
65. Dittenber DB and GangaRao HVS. Critical review of recent
publications on use of natural composites in infrastructure.
Compos Part A Appl Sci Manuf 2012; 4343(8): 14191429.
66. Pickering KL, Efendy MGA, Le TM, et al. A review of recent
developments in natural bre composites and their me-
chanical performance. Compos Part A Appl Sci Manuf 2016;
83: 98112.
67. Ashby MF. Materials and the environment: eco-informed
material choice. Amsterdam: Elsevier, 2012.
68. Kumar V, Kumari M and Kumar R. Review: raw natural
bersbased polymer composites. Int Journal Polymer Anal
2014; 19(3): 256271.
69. Hattalli S, Benaboura A, Ham-Pichavant F, et al. Adding
value to alfa grass (stipa tenacissima L.) soda lignin as
phenolic resins 1. Lignin characterization. Polym Degrad
Stab 2002; 76: 259264.
70. 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: 567576.
71. Malkapuram R, Kumar V, Negi YS, et al. Recent de-
velopment in natural ber reinforced polypropylene com-
posites. J Reinf Plast Compos 2009; 28: 11691189.
Chakkour et al. 35
72. Fuqua MA, Huo S, Ulven CA, et al. Natural ber reinforced
composites. Polym Rev 2012; 52: 259320.
73. Li X, Tabil LG, Panigrahi S, et al. Chemical treatment of
natural bre for use in natural bre-reinforced composites:
a review. Polym Environ 2007; 15: 2533.
74. Bledzki AK and Jaszkiewicz A. Mechanical performance of
biocomposites based on PLA and PHBV reinforced with
natural bres a comparative study to pp. Compos Sci
Technol 2010; 70: 16871696.
75. Kozlowski R, Wladyka-Przybylak M and Kicinska-Jaku-
bowska A. A state of art in the research on natural bres and
their properties used in composites. In: Bledzki AK and
Sperber VE (eds) Proc 7th Glob. WPC Nat Fibre Compos,
2008.
76. Alves Fidelis ME, Pereira TVC, Gomes O, et al. The effect of
ber morphology on the tensile strength of natural bers.
J Mater Res Technol 2013; 2(2): 149157.
77. Wang H, Zhou S, Li X, et al. The inuence of climate change
and human activities on ecosystem service value. Ecol Eng
2016; 87: 224239.
78. Trivaudey F and Placet V. Inuence de langle des micro
brilles (AMF) sur le module d´
elasticit´
e longitudinal des
bres v´
eg´
etales approche exp´
erimentale et num´
erique.
Journ´
ees Sci Tech 2010; ■■■:■■■.
79. Eder M, Arnould O, Dunlop JWC, et al. Experimental mi-
cromechanical characterisation of wood cell walls. Wood
Sci Technol, 2012.
80. Humphrey Danso, Martinson D. Brett, Ali Muhammad,
et al. Effect of ber aspect ratio on mechanical properties of
soil building blocks, Construction and building materials,
2015.
81. Singha A.S. and Rana Raj K. Natural ber reinforced
polystyrene composites: Effect of ber loading, ber di-
mensions and surface modication on mechanical proper-
ties, Materials and design, 2012; 47: 163182.
82. Vinod A, Sanjay MR, Suchart S, et al. Renewable and
sustainable biobased materials: an assessment on biobers,
biolms, biopolymers and biocomposites. J Clean Prod
2020; 258: 120978.
83. Chung KF and Yu WK. Mechanical properties of structural
bamboo for bamboo scaffoldings. Eng Struct 2002; 24(4):
429442.
84. Godbole VS and Lakkad SC. Effect of water absorption on
the mechanical properties of bamboo. J Mater Sci Lett 1986;
5: 303304.
85. Kozowski R and Wadyka-Przybylak M. Flammability and
re resistance of composites reinforced by natural bers.
Polym Adv Technol 2008; 19: 446453.
86. Martins MA and Joekes I. Effect of Mercerization and
Acetylation on Reinforcement. J Appl Polym Sci 2003; 89:
25072515.
87. Doan Thi Thu Loan. Investigation on jute bers and their
composites based on polypropylene and epoxy matrices.
Doctoral thesis. Dresden University, 2006
88. Yang H, Yan R, Chen H, et al. Characteristics of hemi-
cellulose, cellulose and lignin pyrolysis. Fuel 2007; 86:
17811788.
89. Joseph PV, Joseph K, Thomas S, et al. The thermal and
crystallization studies of short sisal bre reinforced
polypropylene composites. Composites: Part A 2003; 34:
253266.
90. De Rosa IM, Maria Kenny J, Puglia D, et al. Morphological,
thermal and mechanical characterization of okra (Abelmo-
schux exculentus) bers as potential reinforcement in
polymer composites. Composites Science Technology 2010;
70: 116122.
91. Tomczak Fabio, Helena T, Sydenstricker D, et al. Studies on
lignocellulosic bres of Brazil. Part II: morphology and
properties of Brazilian coconut bres. Composites: Part A
2007; 38: 17101721.
92. Vinod A, Sanjay MR, Siengchin S, et al. Fatigue and thermo-
mechanical properties of chemically treated Morinda cit-
rifolia ber-reinforced bio-epoxy composite: a sustainable
green material for cleaner production. J Clean Prod 2021;
326: 129411.
93. Rangappa SM, Parameswaranpillai J, Siengchin S, et al.
Bioepoxy based hybrid composites from nanollers of
chicken feather and lignocellulose Ceiba Pentandra. Sci Rep
2022; 12(1): ■■■.
94. Vinod A, Yashas Gowda TG, Vijay R, et al. Novel muntingia
calabura bark ber reinforced green-epoxy composite:
a sustainable and green material for cleaner production.
J Clean Prod 2021; 294: 126337.
95. Sumrith N, Techawinyutham L, Sanjay MR, et al. Char-
acterization of alkaline and silane treated bers of water
hyacinth plantsand reinforcement of water hyacinth -
berswith bioepoxy to develop fully biobased sustainable
ecofriendly composites. J Polym Environ 2020; 28:
27492760.
96. Rao KM and Rao Mohana K. Extraction and tensile prop-
erties of natural bers: vakka, date and bamboo. Compos
Struct 2007; 77: 288295.
97. Tien PN, Fujii T, Bui C, et al. Study on how to effectively
extract bamboo bers from raw bamboo and wastewater
treatment. J Mater Sci Res 2011; 1: 144155.
98. Jindal UC. Development and testing of bamboo-bres re-
inforced plastic composites. J Compos Mater 1986; 20:
1929.
99. Kim H, Okubo K, Fujii T, et al. Inuence of ber extraction
and surface modication on mechanical properties of green
composites with bamboo ber. J Adhes Sci Technol 2013; 27:
13481358.
100. Ray AK, Das SK and Mondal S. Microstructural charac-
terization of bamboo. J Mater Sci 2004; 39: 10551060.
101. Vijay R, Daniel J, Dhilip J, et al. Characterization of
natural cellulose ber from the barks of vachellia far-
nesiana characterization of natural cellulose ber from the
barks of vachellia farnesiana. J Nat Fibers 2020; ■■■:
110.
102. Vigneshwaran S, Sundarakannan R, John KM, et al. Recent
advancement in the natural ber polymer composites:
a comprehensive review. J Clean Prod 2020; 277: 124.
103. Rajak DK, Pagar DD, Kumar R, et al. Recent progress of
reinforcement materials: a comprehensive overview of
composite materials. J Mater Res Technol 2019; 8:
63546374.
104. Hamidon MH, Sultan MTH, Arifn AH, et al. Effects of bre
treatment on mechanical properties of kenaf bre reinforced
36 Journal of Reinforced Plastics and Composites 0(0)
composites: a review. J Mater Res Technol 2019; 8:
33273337.
105. Marais S, Gouanve F, Bonnesoeur A, et al. Unsaturated
polyester composites reinforced with ax bers: effect of
cold plasma and autoclave treatments on mechanical and
permeation properties. Compos Part A Appl Sci Manuf 2005;
36: 975986.
106. Martin AR, Manolache S, Mattoso LHC, et al. Plasma
modication of sisal and high-density polyethylene compo-
sites: effect on mechanical properties, Third International
symposium on natural polymers and composites, 2000,
pp. 431436.
107. Sathish S, Karthi N, Prabhu L, et al. A review of natural ber
composites: extraction methods, chemical treatments and
applications. Mater Today Proc 2020; ■■■:■■■.
108. El-Shekeil YA, Sapuan SM, Khalina A, et al. Inuence of
chemical treatment on the tensile properties of kenaf ber
reinforced thermoplastic polyurethane composite. eXPRESS
Polym Lett 2012; 6: 10321040.
109. Khan A, Rangappa SM, Siengchin S, et al. Asiri, biobers
and biopolymers for biocomposites: synthesis, character-
ization and properties.■■■: book, 2020.
AQ3
110. Li X, Tabil LG, Panigrahi S, et al. Chemical treatments of
natural ber for use in natural ber-reinforced composites:
a review. J Polym Environ 2007; 1515(1): 2533.
111. Mohammed AA, Bachtiar D, Rejab MRM, et al. Effect of
potassium permanganate on tensile properties of sugar palm
bre reinforced thermoplastic polyurethane. Indian J Sci
Technol 2017; 10(7): 15.
112. Jagadeesh P, Puttegowda M, Rangappa SM, et al. A review
on extraction, chemical treatment, characterization of natural
bers and its composites for potential applications. Polymer
Composites 2021; 2: 62396264.
113. Kiattipanich N, Kreua-Ongarjnukool N, Pongpayoon T, et al.
Properties of polyropylene composites reinforced with
stearic ACID treated sugarcane FIBER. J Polym Eng 2007;
27: 67.
114. Salem IAS, Rozyanty AR, Betar BO, et al. Study of the effect
of surface treatment of kenaf ber on chemical structure and
water absorption of kenaf lled unsaturated polyester
composite. J Phys Conf Ser 2017; 908(1): 012001.
115. Hossain SI, Hasan M, Hasan N, et al. Effect of chemical
treatment on physical , mechanical and thermal properties of
ladies nger natural ber. Adva mate Sci Engineering 2013;
2013(6): ■■■.
116. Osorio L, Trujillo E, Van Vuure AW, et al. Morphological
aspects and mechanical properties of single bamboo bers
and exural characterization of bamboo/epoxy compo-
sites. J Reinforced Plastics Composites 2011; 44:
433438.
117. Nguyen TH, Yamamoto H, Takashi M, et al. Effect of surface
treatment on interfacial strength between bamboo ber and
PP resin. JSME Int J Ser A 2004; 47: 561565.
118. Yun H, Li K, Tu D, et al. Effect of heat treatment on bamboo
ber morphology crystallinity and mechanical properties.
Wood Res.-Slovakia 2016; 61: 227234.
119. Ye C, Huang Y, Feng Q, et al. Effect of hygrothermal
treatment on the porous structure and nanomechanics of
moso Bamboo. Nature 2020; ■■■:■■■.
120. Avinc O, Eren HA, Uysal P, et al. The effects of Ozone
treatment on soybean bers. Ozone: Sci Eng 2012; 34:
143150.
121. de Castro BD, Silva KMMN, Maziero R, et al. Inuence of
gamma radiation treatment on the mechanical properties of
sisal bers to use into composite materials. Fibers Polym;
21(8): 18161823.
122. Radoor S, Karayi J, Rangappa SM, et al. A review on the
extraction of pineapple, sisal and abaca bers and their use as
reinforcement in polymer matrix. eXPRESS Polym Lett 2020;
14: 309335.
123. Jaiswa D, Devnani GL, Rajeshkumar G, et al. Suchart
siengchin, review on extraction, characterization, surface
treatment and thermal degradation analysis of new cellulosic
bers as sustainable reinforcement in polymer composites.
Curr Res Green Sustain Chem 2022; 5: 100271.
124. Obi RK, Uma Maheswari C, Shukla M, et al. Studies on
borassus fruit ber and its composites with polypropylene.
Comp Part B 2013; 44: 433438.
125. Chikouche MD, Merrouche A and Azizi A. Inuence of
alkali treatment on the mechanical properties of new cane
bre/polyester composites. J Reinf Plast Comp 2015; 34:
13291339.
126. Kalia S, Kaushik VK, Sharma RK, et al. Effect of Ben-
zoylation and graft copolymerization on morphology, ther-
mal stability, and crystallinity ofssisal bers. J Nat Fibers
2011; 8: 2738.
127. Sampathkumar D, Punyamurthy R, Bennehalli B, et al.
Physical characterization of natural lignocellulosic single
areca ber, ci ˆ
encia.■■■: Tecnologia dos Materiais, 2015,
pp. 121135.
128. Prasad SV, Pavithran C, Rohatgi PK, et al. Alkali treatment of
coir for coir-polyester composites. J Mater Sci 1983; 18:
14431454.
129. Suardana NPG, Piao Y and Lim JK. Mechanical properties of
hemp bers and hemp/pp composites: effects of chemical
surface treatment. Mater Phys Mech 2011; 11: 18.
130. Cantero G, Arbelaiz A, Llano-Ponte R, et al. Effects of bre
treatment on wettability and mechanical behaviour of ax/
polypropylene composites. Compos Sci Technol 2003; 63:
12471254.
131. Han YH, Han SO, Cho D, et al. Kenaf/polypropylene bio-
composites: effects of electron beam irradiation and alkali
treatment on kenaf natural bers. Compososite Inter 2007;
14: 559578.
132. Vidyashri V, Lewis H, Narayanasamy P, et al. Preparation of
chemically treated sugarcane bagasse ber reinforced epoxy
composites and their characterization. Cogent Engineering,
2019, 1708644.
133. Venkateshwaran N, Perumal AE, Arunsundaranayagam D,
et al. Fiber surface treatment and its effect on mechanical and
visco-elastic behaviour of banana/epoxy composite. Mater
Des 2013; 47: 151159.
134. Mofokeng JP, Luyt AS, Tabi T, et al. Comparison of injection
moulded, natural bre-reinforced composites with PP and
PLA as matrices. J Thermoplast Compos Mater 2012; 25(8):
927948.
135. Letroedec Marianne. Caract´
erisation des interactions
physico-chimiques dans un mat´
eriau composite `
a base de
Chakkour et al. 37
phyllosilicates, de chaux et de bres cellulosiques. Thèse de
Doctorat, 2009.
136. Madhu P, Sanjay MR, Jawaid M, et al. A new study on effect
of various chemical treatments on Agave Americana ber for
composite reinforcement: physico-chemical, thermal, me-
chanical and morphological properties. Polymer Testing
2020; ■■■:■■■.
137. Xu M, Cui Z, Chen Z, et al. Experimental study on com-
pressive and tensile properties of a bamboo scrimber at ele-
vated temperatures. Construc build mat 2017; 151: 732741.
138. Li Z, Zhang X, Fa C, et al. Investigation on characteristics
and properties of bagasse bers: performances of asphalt
mixtures with bagasse bers. Construc Build Mat 2020; 150:
118648.
139. Parre A, Karthikeyana B, Balaji A, et al. Investigation of
chemical, thermal and morphological properties of untreated
and NaOH treated banana bers. Mater Today: Proc 2019;
22: 347352.
140. Raja K, Prabu B, Ganeshan P, et al. Characterization studies
of natural cellulosic bers extracted from shwetark stem
characterization studies of natural cellulosic bers extracted
from shwetark stem. J Nat Fibers 2020; ■■■:112.
141. Khan A, Vijay R, Singaravelu DL, et al. Extraction and
characterization of natural ber from Eleusine Indica grass as
reinforcement of sustainable ber-reinforced polymer com-
posites extraction and characterization of natural ber from
Eleusine Indica grass as reinforcement of sustainable b.
J Nat Fibers 2019; ■■■:19.
142. Pothan LA, Oommen Z, Thomas S, et al. Dynamic me-
chanical analysis of banana ber reinforced polyester
composites. Compos Sci Technol 2003; 63(2): 283293.
143. Najeeb MI, Sultan MTH, Andou Y, et al. Characterization of
lignocellulosic biomass from Malaysians Yankee pineapple
AC6 toward composite application characterization of lig-
nocellulosic biomass from Malaysians Yankee pineapple
AC6 toward composite application. J Nat Fibers 2020; ■■■:
113.
144. Vijay R, Manoharan S, Arjun S, et al. Characterization of
silane-treated and untreated natural bers from stem of leucas
aspera characterization of silane-treated and untreated natural
bers. J Nat Fibers 2020; ■■■:117.
145. Palai BK, Sarangi SK, Mohapatra SS, et al. Investigation of
physiochemical and thermal properties of Eichhornia Cras-
sipes bers investigation of physiochemical and thermal
properties of Eichhornia Crassipes bers. J Nat Fibers 2019;
18: 13202133.
146. Kumar R, Sivaganesan S, Senthamaraikannan P, et al.
Characterization of new cellulosic ber from the bark of
Acacia nilotica L. Plant characterization of new cellulosic
ber from the bark of Acacia. J Nat Fibers 2020; ■■■:110.
147. Manimaran P, Sanjay MR, Senthamaraikannan P, et al. A
new study on characterization of pithecellobium dulce ber
as composite reinforcement for light- weight applications
composite reinforcement for light-weight applications. J Nat
Fibers 2018; ■■■:112.
148. Manimaran P, Kumar KSS and Prithiviraj M. Investigation of
physico chemical, mechanical and thermal properties of the
Albizia Lebbeck bark bers investigation of physico chemical,
mechanical and thermal. JNatFibers2019; ■■■:112.
149. Manimaran P, Prithiviraj M, Saravanakumar SS, et al.
Physicochemical, tensile, and thermal characterization of
new natural cellulosic bers from the stems of sida cordifolia.
J Nat Fibers 2017; 15: 110.
150. Ganapathy T, Sathiskumar R, Senthamaraikannan P, et al.
International journal of biological macromolecules charac-
terization of raw and alkali treated new natural cellulosic
bres extracted from the aerial roots of banyan tree. Int J Biol
Macromolecules 2019; 138: 573581.
151. Balaji AN and Nagarajan KJ. Characterization of alkali
treated and untreated new cellulosic ber from Saharan aloe
vera cactus leaves. Carbohydr Polym 2017; 174: 200208.
152. Mahmud S, Faridul Hasan KMF, Anwar Jahid Md, et al.
Comprehensive review on plant ber reinforced polymeric
biocomposites. J Mater Sci 2021; 4: 72317264.
153. Lokensgard E. Industrial plastics: theory and applications.
6th ed. ■■■: Cengage Learning, 2016, p. 544.
154. Mohammad NAB. Synthesis, characterization, and prop-
erties of the new unsaturated polyester resins for composite
applications.■■■: MARA University of Technology, 2007,
pp. 4556.
155. Mukherjee RN, Pal SK, Sanyal SK, et al. Role of interface in
bre reinforced polymer composites with special reference to
natural bres. J Polym Mater 1984; 1: 6981.
156. Siracusa V and Blanco I. Bio-Polyethylene (Bio-PE), Bio-
Polypropylene Bio-PP) and Bio-Poly(ethylene terephthalate
(Bio-PET): recent developments in Bio-Based polymers
analogous to petroleum-derived ones for packaging and
engineering applications, review. Polymers 2020; ■■■:■■■.
157. Luo S and Netravali AN. Interfacial and mechanical prop-
erties of environment-friendly greencomposites made
from pineapple bers and poly(hydroxybutyrate-co-valerate)
resin. Journal Mat Sci 1999; 34: 37093719.
158. Kuruppalil Z. Green plastics: an emerging alternative for
petroleum-based plastics. Int J Eng Res Innov 2011; 3:
5964.
159. George A, Sanjay MR, Srisuk R, et al. A comprehensive
review on chemical properties and applications of bio-
polymers and their composites. Int J Biol Macromolecules
2020; 154: 329338.
160. Clarinval AM and Halleux J. Classication of biodegradable
polymers. In: Smith R (ed) Biodegradable polymers for in-
dustrial applications. Amsterdam: Elsevier, 2005, pp. 331.
161. Zafeiropoulos NE. Engineering the bre matrix interface in
natural-bre composites. Properties Perform Natural-Fibre
Comp 2008; ■■■: 127162.
162. Huang S, Fu Q, Yan L, et al. Characterization of interfacial
properties between bre and polymer matrix in composite
materials A critical review. J Mater Res Technol 2021; 13:
14411484.
163. Scarponi C and Pizzinelli CS. Interface and mechanical
properties of natural bers reinforced composites: a review.
Int Journal Materials Product Technol 2009; ■■■:■■■.
164. JHerrera-Franco P and Valadez-Gonz´
alez A. Properties of
ber-matrix interfaces of natural ber composites. Reference
module in materials science and materials engineering, 2021.
165. Shalwan A and Yousif BF. State of art: mechanical and
tribological behaviour of polymeric composites based on
natural bres. J Mater Des 2012; 48: 1424.
38 Journal of Reinforced Plastics and Composites 0(0)
166. Meht G, Mohanty AK, Misra M, et al. Effect of novel sizing
on the mechanical and morphological characteristics of
natural ber reinforced unsaturated polyester resin based bio-
composites. J Mat Sci 2004; 39: 29612964.
167. Baloyi RB, Ncube S, Moyo M, et al. Analysis of the
properties of a glass/sisal/polyester composite. Nature 2021;
■■■:■■■.
168. Salman SD, Leman Z, Sultan MTH, et al. The effects of
orientation on the mechanical and morphological properties
of woven kenaf reinforced poly vinyl butyral lm. BioRes
2016; 11: 11761188.
169. Yan L, Chouw N, Huang L, et al. Effect of alkali treatment on
microstructure and mechanical properties of coir bres, coir
bre reinforced-polymer composites and reinforced-
cementitious composites. Construction Building Mater
2016; 112: 168182.
170. Yan L, Hao M, Yiou S, et al. Effects of resin inside ber
lumen on the mechanical properties of sisal ber reinforced
composites. Compos Sci Technol 2015; 108: 3240.
171. Ramesha M, Palanikumar K, Reddy KH, et al. Comparative
evaluation on properties of hybrid glass ber- sisal/jute re-
inforced epoxy composites. Proced Eng 2013; 51: 745750.
172. Codispoti R, Oliveira VD, Olivito RS, et al. Mechanical
performance of natural ber-reinforced composites for the
strengthening of masonry. Composites Part B 2015; 77:
7483.
173. Biswas S, Shahinur S, Hasan M, et al. Physical, mechanical
and thermal properties of jute and bamboo ber reinforced
unidirectional epoxy composites. Proced Eng 2015; 10:
933939.
174. Mir A, Zitoune R, Collombet F, et al. Study of mechanical
and thermomechanical properties of jute/epoxy composite
laminate. J Reinforc Plast Compos 2010; 29: 16691680.
175. Ramesha M, Atreyaa TSA, Aswina US., et al. Processing and
mechanical property evaluation of banana ber reinforced
polymer composites. Proced Eng 2014; 97: 563572.
176. Punyamurthy R, Sampathkumar D, Bennehalli B, et al. Abaca
ber reinforced epoxy composites: evaluation of impact
strength. Int J Sci Basic Appl Res 2011; 18: 16481656.
177. Sanjay MR and Yogesha B. Studies on hybridization effect of
jute/kenaf/E-glass woven fabric epoxy composites for po-
tential applications: effect of laminate stacking sequences.
J Industrial Textiles 2017; 47(00): 119.
178. Arpitha GR, Sanjay MR, Senthamaraikannan P, et al. Hy-
bridization effect of Sisal/Glass/epoxy/ller based woven
fabric reinforced composites. Experi techn 2017; ■■■:■■■.
179. Athith D, Sanjay MR, Yashas Gowda TG, et al. Effect of
tungsten carbide on mechanical and tribological properties of
jute/sisal/E-glass fabrics reinforced natural rubber/epoxy
composites. J Industrial Textiles 2017; 48: 125.
180. Vinod A, Tengsuthiwat J, Gowda Y, et al. Suchart siengchin,
hom nath dhakal, jute/hemp bio-epoxy hybrid bio-
composites: inuence of stacking sequence on adhesion of
ber-matrix. Int J Adhesion Adhesives 2022; ■■■:■■■.
181. Jothibasu S, Mohanamurugan S, Vijay R, et al. In-
vestigation on the mechanical behavior of areca sheath
bers jute bers/glass fabrics reinforced hybrid composite
for light weight applications. Int J Adhesion Adhesives
2018; 00: 125.
182. Yorseng K, Rangappa SM, Pulikkalparambil H, et al. Ac-
celerated weathering studies of kenaf/sisal ber fabric re-
inforced fully biobased hybrid bioepoxy composites for
semi-structural applications: morphology, thermo-
mechanical, water absorption behavior and surface hydro-
phobicity. Constr Build Mat 2020; 235: 117464.
183. Sanjay MR, Madhu P, Jawaid Mohammad, et al. Charac-
terization and properties of natural ber polymer composites:
a comprehensive review. J Clean Prod 2017; 172: 566581.
184. Wei C, Zeng M, Xiong X, et al. Friction properties of
sisalber/nano-silica reinforced phenol formaldehyde com-
posites. Polym Comp 2015; 36: 433438.
185. Mutlu I. Investigation of tribological properties of brake pads by
usingricestrawandricehuskdust.JApplSci2009; 9: 377381.
186. Ojha S, Raghavendra G, Acharya S, et al. A comparative in-
vestigation of bio waste ller (wood apple-coconut) reinforced
polymer composites. Polym Comp 2014; 35: 180185.
187. El-Tayeb N. A study on the potential of sugarcane bers/
polyester composite for tribological applications. Wear 2008;
265: 223235.
188. Hao X, Zhou H, Mu B, et al. Effects of ber geometry and
orientation distribution on the anisotropy of mechanical
properties, creep behavior, and thermal expansion of natural
ber/HDPE composites. Composites Part B, 2020; ■■■:
■■■.
189. Bajuri F, Mazlan N, Ishak MR, et al. Flexural and com-
pressive properties of hybrid kenaf/silica nanoparticles in
epoxy composite. Proced Chem 2016; 19: 955960.
190. Jeyapragash RSV and Sathiyamurthy S. Mechanical prop-
erties of natural ber/particulate reinforced epoxy
composites a review of the literature. Mater Today Proc
2020; 22: 12231227.
191. Mishra V. and Biswas S. Physical and mechanical properties
of Bi-directional jute ber epoxy composites. Proced Eng
2013; 51: 561566.
192. Paul V., Kanny K., Redhi G.G., et al. Mechanical, thermal
and morphological properties of a bio-based composite
derived from banana plant source. Composites Part A 2015;
68: 90100.
193. Shringi D, Solanki G and Sharma P. Mechanical properties
characterization of natural (coir based) ber polymer com-
posite by numerical methods. Int J Enhanced Res Sci Technol
Eng 2014; 3: 465470.
194. Gupta R, Sulaiman N, Gupta A, et al. An empirical approach
for prediction of natural ber reinforced polypropylene
composite properties. Appl Mech Mater 2014; 534: 6973.
195. Epaarachchi J, Ku H, Gohel K, et al. A simplied empirical
model for prediction of mechanical properties of random
short ber/vinylester composites. J Compos Mater 2010; 44:
779788.
196. Xiong X, Hua L, Miao M, et al. Multi-scale constitutive
modeling of natural ber fabric reinforced composites.
Composites Part A: Appl Sci Manufac 2018; 115: 383396.
197. Sliseris J, Yan L, Kasal B, et al. Numerical modeling of ax
short ber reinforced and ax bre fabric reinforced polymer
composites. Composites Part B Eng 2015; 89: 143154.
198. Migneault S, Koubaa A, Erchiqui F, et al. Application of
micromechanical models to tensile properties of wood-
plastic composites. Wood Sci Technol 2011; 45: 521532.
Chakkour et al. 39
199. Cao Y, Wang W, Wang Q, et al. Application of mechanical
models to ax ber/wood ber/plastic composites. Bio-
Resources 2013; ■■■:■■■.
200. Beckermann GW and Pickering KL. Engineering and
evaluation of hemp bre reinforced polypropylene compo-
sites: micro-mechanics and strength prediction modelling.
Composites A 2009; 40: 210217.
201. Mulenga TK, Ude AU and Vivekanandhan C. Techniques for
modelling and optimizing the mechanical properties of
natural ber composites: a review. Fibers 2021; ■■■:■■■.
202. Eshelby JD. The determination of the elastic eld of an
ellipsoidal inclusion, and related problems. Proc R Soc Lond
A: Math Phys Eng Sci 1957; 241: 376396.
203. Mur T. Micromechanics of defects in solids.■■■: Martinus
Nijhoff Publishers, 1987.
204. Mori T and Tanaka K. Average stress in matrix and average
elastic energy of materials with mistting inclusions. Acta
Metallurgica 1973; 21: 571574.
205. Pupure L, Varna J, Joffe R, et al. Mechanical properties of
natural ber composites produced using dynamic sheet
former. Wood Material Science & Engineering, 2018;
15(31): 111.
206. Keck S and Fulland M. Effect of bre volume fraction and
bre direction on crack paths in ax bre-reinforced com-
posites. Eng Fracture Mech 2016; 167: 201209.
207. Chaboche JL. 2.04 - Damage mechanics. In: Milne I, Ritchie
R and Karihaloo B (eds) Comprehensive structural integrity.
Pergamon, 2003, pp. 213284.
208. Beakou A and Charlet K. Mechanical properties of interfaces
within a ax bundle-Part II: numerical analysis. Int J Adhes
Adhesives 2013; 43: 5459.
209. Hbib M, Guessasma S, Bassir D, et al. Interfacial damage
in biopolymer composites reinforcedusing hemp bres:
nite element simulation and experimental investigation.
Compo Sci Technol - Compo Sci Technol.2011;71:
14191426.
210. Mishra C, Ranjan Deo C, Baskey S, et al. Inuence of
moisture absorption on mechanical properties of kenaf/glass
reinforced polyester hybrid composite. Mater Today Proc
2020; 38: 25962600.
211. Venkatesha BK, Saravanan R, Anand Babu K, et al. Effect of
moisture absorption on woven bamboo/glass ber reinforced
epoxy hybrid composites. Mater Today Proc Journal 2020;
45: 216221.
212. Deo C and Acharya SK. Effect of moisture absorption on
mechanical properties of chopped natural ber reinforced
epoxy composite. J Reinforced Plastics Compo 2010; 29:
25132521.
213. Jannah M, Mariatti M, Abu Bakar A, et al. Effect of chemical
surface modications on the properties of woven banana-
reinforced unsaturated polyester composites. J Reinforced
Plastics Composites 2009; 28: 15191532.
214. Shen X, Jia J, Chen C, et al. Enhancement of mechanical
properties of natural ber composites via carbon nanotube
addition. J Mater Sci 2014; 49: 32253233.
215. Chee SS, Jawaid M, Sultan MTH, et al. Effects of nanoclay
on physical and dimensional stability of Bamboo/Kenaf/
nanoclay reinforced epoxy hybrid nanocomposites.
J Mater Res Technol 2020; 9: 58715880.
216. Sapiai N, Jumahat A, Shaari N, et al. Mechanical properties
of nanoclay-lled kenaf and hybrid glass/kenaf ber com-
posites. Mater Today Proc 2020; 46: 17871791.
217. Mylsamy B, Kumar Palaniappanb S, Subramani SP, et al.
Impact of nanoclay on mechanical and structural properties
of treated Coccinia indica bre reinforced epoxy composites.
J Mater Res Technol 2019; 8: 60216028.
218. Prasad V, Joseph MA and Sekar K. Investigation of me-
chanical, thermal and water absorption properties of ax bre
reinforced epoxy composite with nano TiO2 addi-
tion.Composites: Part A 2018; ■■■:■■■
219. Kumar Sinha R, Sridhar K, Purohit R, et al. Effect of nano
SiO2 on properties of natural ber reinforced epoxy hybrid
composite: a review. Mater Today Proc 2020; 26:
31833186.
220. Devnania GL and Sinha S. Effect of nanollers on the
properties of natural ber reinforced polymer composites.
Mater Today 2017; ■■■:■■■.
221. Manohar Maharana S, Pradhan AK, Kumar Pandit M, et al.
Moisture absorption behaviour of nanoller reinforced
jute-kevlar hybrid polymer composite. Mater Today Proc
2020; 26: 775780.
222. Suhana Hassan CIK, Rao Chellaiah N, Barkawi S, et al.
Effect of chemical treatment on oil palm empty fruit bunch
(OPEFB) ber on water absorption and tensile properties of
OPEFB ber reinforced epoxy composite. Key Eng Mat,
2016; ■■■: 295299.
223. Osman E, Vakhguelt A, Sbarski I, et al. Water absorption
behavior and its effect on the mechanical properties og kenaf
natural FIBER unsaturated polyester composites. In: 18th
internaional conference on composite materials, 2018.
224. Mittal V and Sinha S. Effect of chemical treatment on the
mechanical and water absorption properties of bagasse bre-
reinforced epoxy composites. J Polym Eng 2015; ■■■:■■■.
225. Gupta MK and Srivastava RK. Mechanical, thermal and
water absorption properties of hybrid sisal/jute bre re-
inforced polymer composite. Indian J Eng Mater Sci 2016.
226. Prasanna Venkatesh R, Ramanathan K, Srinivasa Raman V,
et al. Tensile, exual, impact and water absorption properties
of natural bre reinforced polyester hybrid composites. FI-
BRES Textiles East Europe 2016; 117: 9094.
227. Idicula M, Boudenne A, Umadevi L, et al. Thermophysical
properties of natural bre reinforced polyester composites.
Composites Sci Technol 2006; 66: 27192725.
228. Takagi H, Nakagaito AN and Liu K. Heat transfer analyses of
natural bre composites. High Perform Optimum Des
Structures Mat 2014; 137: ■■■.AQ4
229. Vijaya Bhaskar V, Srinivas K, Devireddy SBR, et al. A novel
mathematical correlation for thermal conductivity of hybrid
composites reinforced with natural bers. Mater Today Proc
2020; 26: 22082211.
230. Yang W, Hu Y, Tai Q, et al. Fire and mechanical performance
of nanoclay reinforced glass-ber/PBT composites con-
taining aluminum hypophosphite particles. Composites: Part
A2011; 4242: 794800.
231. Dhanushka Hapuarachchi T and Peijs T. Ton peijs, multi-
walled carbon nanotubes and sepiolite nanoclays as ame
retardants for polylactide and its natural bre reinforced
composites. Composites: Part A 2010; 41: 954963.
40 Journal of Reinforced Plastics and Composites 0(0)
232. Suhartya NS, Almanar IP, Sudirman K, et al. Flammability,
biodegradability and mechanical properties of bio-
composites waste polypropylene/kenaf ber containing
nano CaCO3 with diammonium phosphate. Proced Chem
2012; 4: 282287.
233. Walbrück K, Drewler L, Witzleben S, et al. Factors inu-
encing thermal conductivity and compressive strength of
natural ber-reinforced geopolymer foams, open Ceramics
2021; ■■■:■■■.
234. Dilli Babu G, Kumar Gudapati SP, Ratna Prasad AV, et al.
Experimental investigation on mechanical and thermal
properties of Esculentus Cyperus ber reinforced poly-
propylene composites. Mater Today Proc 2020; 23:
557560.
235. Sair S, Oushabi A, Kammouni A, et al. Mechanical and
thermal conductivity properties of hemp ber reinforced
polyurethane composites. Stud Construction Mater 2018; 8:
203212.
236. Radzi AM, Sapuan SM, Jawaid M., et al. Water absorption,
thickness swelling and thermal properties of roselle/sugar
palm bre reinforced thermoplastic polyurethane hybrid
composites. J Mater Res Technol 2019; 8: 39883994.
237. Teixeira LA, Junior LVD, Luz SM, et al. Chemical treatment
of curaua bres and its effect on the mechanical performance
of bre/polyester composites. Plastics, Rubber and Compo
2020; 50: 189199.
238. Shanmugasundaram N, Rajendran I, Ramkumar T, et al.
Static, dynamic mechanical and thermal properties of un-
treated and alkali treated mulberry ber reinforced polyester
composites. Polym Composites 2018; 39: 19081919.
239. Jawaid M, Khalil HA, Bakar AA, et al. Effect of jute bre
loading on the mechanical and thermal properties of oil
palmepoxy composites. J Compos Mater 2013; 47:
16331641.
240. Maou S, Meghezzi A, Grohens Y, et al. Effect of various
chemical modications of date palm bers (DPFs) on the
thermo-physical properties of polyvinyl chloride (PVC)
high-density polyethylene (HDPE) composites. Ind Crops
Prod 2021; 171: 113974.
241. Fang C-C, Zhang Y, Qi S-Y, et al. Inuence of structural
design on mechanical and thermal properties of jute re-
inforced polylactic acid (PLA) laminated composites. Cel-
lulose 2020; 27: 93979407.
242. Jandas PJ, Mohanty S, Nayak SK, et al. Effect of surface
treatments of banana ber on mechanical, thermal, and
biodegradability properties of PLA/banana ber bio-
composites. Polym Composites 2011; 32: 16891700.
243. Devnani GL and Sinha S. Effect of nanollers on the
properties of natural ber reinforced polymer composites.
Mater Today: Proc 2019; 18: 647654.
244. Ramakrishnan S, Krishnamurthy K, Rajasekar R. et al. An
experimental study on the effect of nano-clay addition on
mechanical and water absorption behaviour of jute bre
reinforced epoxy composites. J Industrial Textiles 2019; 49:
597620.
245. Hossain MK, Dewan MW, Hosur M, et al. Mechanical
performances of surface modied jute ber reinforced biopol
nanophased green composites. Composites Part B 2011; 42:
17011707.
246. Deepak K, Vattikuti SP, Venkatesh B, et al. Experimental
investigation of jute ber reinforced nano Clay Composite.
Proced Mater Sci 2015; 10: 238242.
247. Borba PM, Tedesco A, Lenz DM, et al. Effect of reinforcement
nanoparticles addition on mechanical properties of SBS/curau´
a
ber composites. Mater Res 2014; 17: 412419.
248. Saba N, Paridah MT, Abdan K, et al. Physical, structural and
thermomechanical properties of oil palm nano ller/kenaf/
epoxy hybrid nanocomposites. Mater Chem Phys 2016; 184:
6471.
249. Venkatram B, Kailasanathan C, Seenikannan P, et al. Study
on the evaluation of mechanical and thermal properties of
natural sisal ber/general polymer composites reinforced
with nanoclay. Int J Polym Anal Characterization 2016; 21:
647656.
250. Prasad AR, Rao KB, Rao KM, et al. Inuence of nanoclay on
the mechanical performance of wildild cane grass ber-
reinforced polyester nanocomposites. Int J Polym Anal
Characterization 2015; 20: 541556.
251. Wang A, Xia D, Xian G. et al. Effect of nanoclay grafting
onto ax bers on the interfacial shear strength and me-
chanical properties of ax/epoxy composites. Polym Com-
posites 2019; 40: 34823492.
252. Biswal M, Mohanty S, Nayak SK, et al. Inuence of or-
ganically modied nanoclay on the performance of pineapple
leaf ber-reinforced polypropylene nanocomposites. J Appl
Polym Sci 2009; 114: 40914103.
253. Samariha A, Hemmasi AH, Ghasemi I, et al. Effect of
nanoclay contents on properties, of bagasse our/
reprocessed high density polyethylene/nanoclay compo-
sites. Maderas: Ciencia y Tecnolog´
ıa, 2015.
254. Shahroze RM, Ishak MR, Salit MS, et al. Effect of organo-
modied nanoclay on the mechanical properties of sugar
palm ber-reinforced polyester composites. BioResources
2018; 3: 74307444.
255. Nagarajan KJ, Ramanujam NR, Sanjay MR, et al. A com-
prehensive review on cellulose nanocrystals and cellulose
nanobers: Pretreatment, preparation, and characterization.
Polym Composites 2021; 42: 15881630.
256. Shojaeiarani J, Dilpreet SB, Chanda S, et al. Cellulose
nanocrystal based composites: a review. Composites Part C
2021; 5: 100164.
257. Wang WJ, Wang WW, Shao Z-Q, et al. Surface modication
of cellulose nanowhiskers for application in thermosetting
epoxy polymers. Cellulose 2014; 2529: 2538.
258. Ansari F, Galland S, Johansson M, et al. Berglund, cellulose
nanober network for moisture stable, strong and ductile
biocomposites and increased epoxy curing rate. Composites:
Part A 2014; 63: 3544.
259. Al-Turaif HA. Relationship between tensile properties and
lm formation kinetics of epoxy resin reinforced with
nanobrillated cellulose. Prog Org Coat 2013; 76: 23.
260. Kargarzadeh H, Mariano M, Huang J, et al. Recent devel-
opments on nanocellulose reinforced polymer nano-
composites: A review. Polymer 2017; 132: 368393.
261. Gao C, Wan Y, He F, et al. Mechanical, moisture absorption,
and photodegradation behaviors of bacterial cellulose
nanober- reinforced unsaturated polyester composites. Adv
Polymer Technologies 2011; 30: 249256.
Chakkour et al. 41
262. Neelamana IK, Thomas S, Parameswaranpillai J, et al.
Characteristics of banana bers and banana ber reinforced
phenol formaldehyde composites-macroscale to nanoscale.
Appl Polym Sci 2013; 130: 12391246.
263. Nakagaito AN, Iwamoto S, Yano H, et al. Bacterial cellulose:
the ultimate nano-scalar cellulose morphology for the pro-
duction of high-strength composites. Appl Phys A 2005; 80:
9397.
264. Henriksson M and Berglund LA. Structure and properties of
cellulose nanocomposite lms containing melamine form-
aldehyde. J Applied Polymer Science 2007; 106: 28172824.
265. Bulut M, Bozkurt ¨
OY, Erkli A, et al. Mechanical and dy-
namic properties of basalt ber-reinforced composites with
nanoclay particles. Arabian J Sci Eng 2020; 45: 10171033.
266. del Pino GG, Kieling AC, Bezazi A, et al. Hybrid polyester
composites reinforced with curau´
abres and nanoclays.
Fibers Polym 2020; 21: 399406.
267. Mohan TP and Kanny K. Compressive characteristics of
unmodied and nanoclay treated banana ber reinforced
epoxy composite cylinders. Composites Part B 2019; 169:
118125.
268. Mylsamy B, Palaniappan SK, Subramani SP, et al. Impact of
nanoclay on mechanical and structural properties of treated
Coccinia indica bre reinforced epoxy composites. J Mater
Res Technol 2019; 8: 60216028.
269. Ganesan K, Kuttalam KC, Palaniappan M, et al. Molecules
2020; 25: 5579.
270. Ibrahim ID, Jamiru T, Sadiku ER, et al. Impact of surface
modication and nanoparticle on sisal ber reinforced poly-
propylene nanocomposites. J Nanotechnology 2016: 19.
271. Ashori A and Nourbakhsh A. Preparation and character-
ization of polypropylene/wood our/nanoclay composites.
Eur J Wood Wood Prod 2011; 69: 663666.
272. Kanny K. Water barrier properties of nanoclay lled sisal
bre reinforced epoxy composites. Composites Part A Appl
Sci Manufacturing 2011; 42: 385393.
273. Koronis G, Silva A, Fontul M, et al. Green composites:
a review of adequate materials for automotive applications.
Compos Part B Eng 2013; 44(1): 120127.
274. Mohammed L, Ansari MNM, Grace P, et al. A review on
natural ber reinforced polymer composite and its applica-
tions. Int Journal Polymer Science 2015; ■■■:115.
AQ5
275. Holbery J and Houston D. Natural-ber-reinforced polymer
composites in automotive applications. JOM 2006; 58: 8086.
276. Rwawiire S, Tomkova B, Militky J, et al. Development of
a biocomposite based on green epoxy polymer and natural
cellulose fabric (bark cloth) for automotive instrument panel
applications. Composites: Part B 2015; 81: 149157.
277. Djamel Eddine AIZI, Extraction, caract´
erisation morpho-
logique, physico-chimique et m´
ecanique des bres caulin-
aires de Retama onosperma L.Boiss. Doctorate thesis, 2017.
278. Yang Y, Boom R, Irion B, et al. Recycling of composite
materials. Chem Eng Process Process Intensif 2012; 51:
5368.
279. Van de W, Ivens J, De Coster A, et al. Inuence of processing
and chemical treatment of ax bres on their composites.
Composites Sci Technol 2003; 63: 12411246.
280. Lucinte. Opportunities in Natural Fiber Composites. TX,
USA: Lucintel: Las Colinas, 2011.
281. Ngo T-D. Natural bers for sustainable bio-composites. In:
Natural and articial ber-reinforced composites as re-
newable sources. London ,UK: InTech, 2018.
282. Kaur N, Saxena S, Gaur H, et al. A review on bamboo ber
composites and its applications. In: International Conference
on Infocom Technologies and Unmanned Systems, 2017.
283. Nagaraj K. Bharath und, satyappa basavarajappa, applica-
tions of biocomposite materials based on natural bers from
renewable resources: a review. Sci Eng Comp Mat 2015;
■■■:■■■.AQ6
284. Food and Agriculture Organization of the United Nations.
Common fund for commodities. In: Proceedings of the
symposium on natural bres. Rome: Italy, 1999.
285. Witayakran S, Smitthipong W, Wangpradid R, et al. Natural
ber composites: review of recent automotive trends. In:
Reference module in materials science and materials engi-
neering. Amherst, MA, USA: Elsevier Publishing, 2017.
286. Huda MS, Drzal LT, Ray D, et al. Natural-ber composites in
the automotive sector. In: Properties and performance of
natural-bre composites. Oxford, UK: Woodhead Publish-
ing, 2008.
Nomenclature
A The deformation localization tensor
C
MT
Mori-Tanaka stiffness tensor
CNC Cellulose nanocrystal
CNF Cellulose nanober
C
SC
Self-consistent stiffness tensor
E
2
Youngs Modulus of matrix
E
c
Effective Youngs Modulus of composite
E
1
Youngs Modulus of ber
FTIR Fourier transform infrared spectroscopy
IROM Inverse rule of mixture
K
1
Fiber orientation factor
K
2
Fiber length factor
MFA Microbrillar angle
NFRCs Natural bers composites
NFs Natural bers
OM Optical microscope
PLA Polylactic acid
PHA Polyhydroxyalkanoates
RH Relative humidity
SEM Scanning electron miscroscopy
Si
Esh The Eschelby tensor
TGA Thermogravimetric analysis
UV Ultraviolet treatment
V
1
Volume fraction of bers
V
2
Volume fraction of matrix
XRD X-Ray Diffraction
σ
0
Average stress
ϵ
0
Average strain
αHarmonic coefcient
σ
1
Strength of ber
σ
2
Strength of matrix
σ
c
Strength of composite
42 Journal of Reinforced Plastics and Composites 0(0)
... Environmental consciousness has increased these recent years, and therefore the new environmental regulations are somehow one of the factors that lead to think about the use of ecofriendly materials that combine the lightness and strength [1][2][3][4][5]. In particular, cellulosic fibers attract the attention of researchers and seem to be the best candidate to replace synthetic fibers (i.e., carbon and glass fibers) in several industrial applications [1,[5][6][7][8]. ...
... Environmental consciousness has increased these recent years, and therefore the new environmental regulations are somehow one of the factors that lead to think about the use of ecofriendly materials that combine the lightness and strength [1][2][3][4][5]. In particular, cellulosic fibers attract the attention of researchers and seem to be the best candidate to replace synthetic fibers (i.e., carbon and glass fibers) in several industrial applications [1,[5][6][7][8]. Indeed, they are currently used as reinforcement for composites in several applications such as automotive, aeronautic, construction, military and sporting equipment's thanks to their attractive mechanical properties, low cost, lightness, renewability and biodegradability [1,7]. ...
... In particular, cellulosic fibers attract the attention of researchers and seem to be the best candidate to replace synthetic fibers (i.e., carbon and glass fibers) in several industrial applications [1,[5][6][7][8]. Indeed, they are currently used as reinforcement for composites in several applications such as automotive, aeronautic, construction, military and sporting equipment's thanks to their attractive mechanical properties, low cost, lightness, renewability and biodegradability [1,7]. Automotive industry is leader in the use of bio composites especially in internal applications [9,10]. ...
Conference Paper
Recently, research activities are oriented towards natural fibers that seem to be the best candidate to replace synthetic fibers (i.e., carbon, glass, and Kevlar). Particularly, bamboo fiber is a promising candidate in several industries due to its low cost and important specific mechanical properties. However, a limited attention is devoted to its contribution to the mechanical properties of polymer composites. Herein, the current paper reports on the finite element and microme-chanical modeling of the tensile behavior of continuous bamboo fiber composites. More importantly, the experimental tensile test is performed for validation purposes. First, bamboo fibers are isolated from stems using combined mechanical and manual techniques. Composite specimens are manufactured using hand layup and compression techniques. Interestingly, the findings show that the addition of 30 wt.% of bamboo fibers drastically improves the tensile strength of the composite. SEM (Scanning electron microscope) observations of the fractured surface reveal adequate fiber-matrix interfaces indicating the good stress distribution between matrix and fibers. More, the rule of mixture model validates the experimental data due to the low void content of the considered composites. Attractively, the anisotropic finite element model correctly predicts the tensile strength of composites.
... Wang et al. [30] discovered that flax fiber-reinforced polymer (FFRP) reduced flexural strength by 11.2%, 14.9%, and 15.5% and flexural modulus by 21.3%, 32.3%, and 35.8% after natural aging for 60, 120, and 180 days, respectively. Chakkour et al. [31] demonstrated that exposing plant fiber composites to a wet environment significantly diminishes their mechanical properties. ...
Article
Full-text available
To promote resource reuse and the green, low-carbon transformation of the construction industry, this study uses recycled aggregate from crushed waste concrete and natural bamboo fibers to formulate bamboo fiber-reinforced recycled-aggregate concrete. This study investigates the effects of natural bamboo fiber (NBF) content, NBF length, and the water-to-cement ratio on the performance of concrete through an orthogonal experiment to determine the optimal mixing proportions of NBF-reinforced concrete. Additionally, recycled aggregate completely replaced natural aggregate. The mechanism by which NBF influences concrete was also analyzed. The results demonstrate that the NBF-reinforced specimens exhibited good integrity during compression failure, with NBFs effectively tying the concrete together. The optimized parameters for NBF-reinforced concrete were an NBF length of 20 mm, an NBF content of 0.4v%, and a water-to-cement ratio of 0.55. Almost no flaky Ca(OH)2 crystals were observed in the NBF-hardened cement–paste transition zone, indicating effective bonding at the interface.
... In approach to anticipations about the ecosystem, the well-being of living creatures, and the diminishing supply of natural resources, these guidelines were formed. Environmentally friendly fibers are suggested to replace fibers made from artificial yarns in the composites industry, which are not reusable, require greater amounts of energy while manufacturing, and raise questions about the end product's durability over time [6,7]. Fibers that are derived from plants are considered as the best substitute due to their extensive obtainability, minimum price, easiness of production, and decomposable characteristics. ...
Article
Full-text available
Utilizing readily available earth’s assets and agro-waste has grown crucial for the sustainability of business entities. In the current study, Oplismenus hirtellus grass stem fiber (OHSFs) mined from the commonly grown grass Oplismenus hirtellus that retains as herbal waste after their life expectancy. To ensure OHSF’s capabilities and feasibility in boosting the ecological footprint of the automotive industry, resemblance among OHSF and other regularly utilized organic fibers were listed and discussed. The chemical investigations expose the sufficiency of cellulose (60.25 wt%) in OHSF with least wax (0.27 wt%) contributing to its mechanical and interfacial characteristics. The observed chemical features were further endorsed by the Fourier transform infrared spectrum and crystallinity index (49%) of OHSF. The minimum density (1.01 g/cm³) and tensile strength (112–181 MPa) of OHSF experimented through mass-volume method and single fiber tensile test respectively exposes its specific mechanical characteristics. The surface features (Arithmetic mean height, Ra = 21.8 nm) and thermal stability (226 °C) of OHSF ensure the sufficiency of interfacial characteristics and ability to survive fabrication heat respectively to be castoff as stiffening material in polymer composites.
... The lignocellulosic fibers (rice husk, coir, jute, abaca, water hyacinth, etc.) are lightweight, reduce machine wear, and are readily accessible, sustainable, and affordable. Additionally, they are biodegradable and do not leave behind hazardous residues [10][11][12]. Lignocellulosic polymeric composites are made at an affordable price. As a result, these composites have received considerable interest and are playing an ever-growing part in creating a wide range of affordable, lightweight, environmentally friendly composites [13]. ...
Article
Full-text available
Hybrid fiber reinforcements can incorporate a wider array of qualities than single fiber reinforcement. Instead of synthetic fibers, applying plant and animal-based organic materials as reinforcement in polymer matrix offers certain benefits, such as low price, greater availability, and better biodegradability. In the present study, water hyacinth fiber and sheep wool fiber reinforced polypropylene hybrid composites were fabricated at three different (5, 10, and 15 wt.%) fiber loading. The effect of fiber loading on thermo-mechanical, structural, and biodegradability properties was subsequently investigated. The manufacturing process of the composite material was carried out with utmost consideration for biodegradability, since polypropylene, the primary constituent, is not inherently biodegradable. The insertion of fibers into the polypropylene matrix showed variance in properties of different aspects. The tensile strength of the composites displayed a downward trajectory (from 25 to 10 MPa) with a 15% increase in fiber loading due to voids, and fiber dispersion, while impact strength exhibited an opposite trend (from 25 to 32 J/m). Except for hardness, all the mechanical properties degraded slightly after the employment of the reinforcement. Fourier transform infrared spectroscopic analysis revealed the movement of typical peaks and the appearance of new peaks demonstrating the bonding between the fiber and the matrix. Thermogravimetric analysis showed that the thermal degradation temperature of the composites improved at maximum fiber loading. On the other hand, the goal of achieving biodegradability has been succeeded by the implementation of a combination of plant and animal-based fibers as biodegradability of the manufactured composites thrives with increasing fiber content for the presence of cellulosic bonds, as evident from the FTIR spectrum. Even though some properties of the hybrid composite declined slightly with increasing fiber loading, the other characteristics, including service temperature and biodegradability experienced a prospective advancement. Hence, the 15% fiber-loaded composite was found to be a potential candidate in terms of slightly high temperature and environment-friendly applications.
... Due to its inherent characteristics, such as low density, high mechanical strength, renewable nature, and nano-dimensions, cellulose fiber obtained from plant fibers is generating a great deal of attention in the field of materials research. [10][11][12][13] Microfiber cellulose, which has microscale structural dimensions, is derived from cellulose, a biosynthetic product from plants. Plant fiber mostly consists of cellulose, lignin, and hemicellulose. ...
Article
The Paederia foetida (PF) stem's abundant and environmentally friendly novel nanocellulose has immense potential as a reinforcing agent for bio-nanocomposites. This work aims to investigate the physical, thermal, chemical, and morphological properties of cellulose microfibrils from Paederia foetida stems (CMFPFs). Sodium chlorite, acetic acid, and 5 % sodium hydroxide were used to perform bleaching and mercerization operations to extract cellulose. Chemical composition, X-ray diffraction (XRD) analysis, Fourier transform infrared spectroscopy, density , moisture content, examination of cellulose morphology using field emission scanning electron microscopy, and thermal stability were assessed using thermogravimetric analysis. The outcomes showed that the delignification and mercerization methods, respectively, removed lignin and hemicellulose from the extracted cellulose. Higher crystallinity was produced from chemically treated fibers, according to XRD examination. The CMFPF sample had a crystallinity index of 97.9 %, which was higher than the raw PFs (81.9 %) and bleached PFs (84.38 %). The maximum crystallinity was identified for CMFPFs, which had the lowest moisture content, fiber diameter, and density of the samples. After bleaching and mercerization have been applied to the fiber, changes in functional groups take effect.
... Sustainability in industrial practices is the order of the day due to unbalance in ecosystems which reflects as global warming; pollution of air, water, and soil; and exhaustion of natural resources [1,2]. Sustainable practices were further enforced by laws in view to protect and preserve nature and its ecosystem. ...
Article
Full-text available
Bio-waste is the new source of raw materials focussed by composite industries to sort out the sustainability issues in their products and processes. This research portrays the characterization of Licuala grandis tree leaf stalk fibers (LGTLSFs) mined from leaf stalks of the Licuala grandis tree which is a floral waste. The comprehensive investigation aids in obtaining quantifiable data such as cellulose proportion (56.47 wt.%), least wax (0.27 wt.%), minimum density (1.36 g/cm³), greater crystallinity index (49%), tensile strength (312–354 MPa), and Young’s modulus (2.3–6.6 GPa) of LGTLSFs. The thermogravimetric (TGA/DTG) and differential scanning calorimetry (DSC) analysis helps in predicting the thermal behaviour of the LGTLSF and suggests thermal stability until 218 °C. Fourier transform infrared (FTIR) spectroscopy analysis aids in validating the results of chemical investigations. The exterior roughness of the LGTLSFs was analysed through a scanning electron microscope (SEM) to favour its possibility as a support material in polymer composites. Positive findings from the experiment indicate that LGTLSFs can be utilized as a supporting material in polymer composites used in structural applications. Graphical Abstract
Article
Full-text available
In the era of smart and sustainable technology driven by naturally occurring materials, various nanocellulose-based materials play a crucial role. Shape memory behaviour and self-healing capabilities of nanocelluloses are emerging as focal points in numerous research domains. Nanocellulose and its derivatives such as cellulose nanocrystals (CNC) and cellulose nanofibers (CNF), are currently in the limelight due to their excellent shape-memory and self-healing properties, making them suitable for multifunctional devices. In this regard, CNF, as a cutting-edge material, has spurred researchers to explore its potential in developing contemporary multifunctional and personalized health devices. Therefore, a timely and comprehensive review is essential to gain deep insights into the effectiveness of shape-memory and self-healing capabilities of CNF for multifunctional devices. Herein, we first provide a succinct introduction to all nanocellulose materials. This review also depicts recent advancements and breakthroughs in the large and effective synthesis of CNF-based hybrid materials. Next, focusing on their self-healing and shape-memory performance, this review sheds new light on the advanced applications of CNF materials. Finally, perspectives on the current challenges and opportunities in this field are summarized for future researchers to gain an in-depth understanding of "CNF-based smart and sustainable materials."
Article
Full-text available
Lignocellulosic-based polymer composites have gained significant interest due to their ‘’green’’ character as a response to environmental concerns. A diverse array of lignocellulosic fibers is utilized, depending on fiber dimensions, chemical composition, moisture content, and the fiber–matrix interface. The aim of this study is to establish an alternative standardized methodology, aimed at comparatively estimating the performance of polymer composites through the examination of individual plant fibers. The fibers studied are ramie, hemp, flax, and kenaf, and HDPE-based corresponding composites were analyzed for their performance across various fiber-content levels (10, 20, and 30 wt.%). It was found that kenaf showcases the largest average fiber diameter, succeeded by hemp, ramie, and flax. Additionally, ramie and kenaf exhibit elevated levels of crystallinity, suggesting increased cellulose content, with kenaf having the lowest crystallinity index among the fibers compared. Based on Thermogravimetric analysis, ramie displays the lowest moisture content among the examined fibers, followed by hemp, flax, and ultimately kenaf, which is recorded to have the highest moisture content, while, similarly, ramie exhibits the lowest mass loss at the processing temperature of the corresponding composites. Composites containing fibers with smaller diameters and higher crystallinity indexes and lower moisture absorptions, such as ramie and hemp, demonstrate superior thermal stability and exhibit increased Young’s modulus values in their respective composites. However, poor interfacial adhesion affects mechanical performance across all composites. Understanding fiber morphology, inner structure, and thermal stability is important for developing new composite materials and optimizing their selection for various applications.
Article
Full-text available
Nanoclays (layered silicates) have been applied as effective reinforcements for range of natural and synthetic polymeric matrices. Recent research has turned toward design and exploration of green nanocomposites using green polymers and nanoclay nanofillers. This state-of-the-art comprehensive overview debates design and performance prospects of green nanoclay nanocomposites. In this regard, numerous green polymers like poly(lactic acid), poly(vinyl alcohol), natural rubber, cellulose, starch, etc. have been considered. The effectiveness of green nanoclay nanocomposites has been analyzed through microscopic, electrical, mechanical, thermal, adsorption, and biomedical properties and wide span of applications such as packaging, dye removal, and biomedical sectors. Packaging based on cellulose/montmorillonite had very low water vapor transmission rate of 43 g/m².day, whereas poly(lactic acid)/cellulose/montmorillonite packaging performed better with high water vapor transmission rate of 512–1861 g/m².day. Poly(vinyl alcohol)/Cloisite Na possess optimum water vapor transmission rate of 533 g/m².day. Nanocellulose/nanoclay packagings have also been found ideal due to low water vapor permeability (6.3–13.3 g.μm/m².day.kPa) and oxygen permeability (0.07 cm³μm/m².day.kPa) values. In dye removal applications, poly(ethylene glycol)/montmorillonite revealed optimum dye adsorption capacities of 190–237 mg/g, where chitosan/montmorillonite had high dye adsorption capacity of 446.43 mg/g. Poly(lactic acid)/modified Cloisite 20 A systems also own high dye adsorption efficiency of 97%. Poly(ɛ-caprolactone) and poly(vinyl alcohol) systems with montmorillonite nanoclay have effective drug delivery, tissue engineering, and wound healing applications. Furthermore, dielectric, mechanical, non-flammability, and self-extinguishing features of cellulose/montmorillonite nanocomposite systems have been reported. Future of these nanomaterials definitely relies on innovative design, facile fabrication strategies, and overcoming related challenges.
Thesis
Ecological issues have recently encouraged researchers to focus on eco-friendly and lightweight materials. In this way, plant fiber composites offer numerous benefits, such as lower material cost and density as well as important specific mechanical properties. The objective of this thesis work is to contribute to the understanding of the physical and mechanical properties of bamboo fiber composites under moisture aging. Preliminary studies were performed to evaluate the moisture absorption and mechanical properties as well as dimensional stability of bamboo fiber composites at different fiber contents and relative humidity conditions. The effects of montmorillonite (MMT) and eggshell (ESP) particles at different weight contents on the physical, mechanical and morphological properties of bamboo fiber composites, were analyzed. As a result, composites with 3 wt.% of MMT fillers exhibit better mechanical and physical properties in both dry and moisture equilibrium states. However, the exposure to long-term water aging (up to 120 days) significantly influences their long-term mechanical properties. Furthermore, the durability of composites after exposure to 7 hygroscopic cycles (up to 7 months), was explored. Herein, the physicochemical modifications induced by the cyclic aging were analyzed to understand the changes in mechanical properties. Interestingly, neat bamboo fiber composites show perfect resilience to cyclic aging while MMT-filled composites are significantly affected. Finally, the effect of fiber mercerization with various soaking times (0.5, 4, 24 and 48 hours) on the long-term water aging mechanisms and mechanical properties of bamboo fiber composites, was analyzed after prolonged aging (120 days). Based on the obtained results, raw and short period treated fiber composites reveal good resilience to long-term moist aging.
Article
Full-text available
The objective of this investigation is to check the suitability novel cellulosic fibre extracted from the aerial roots of Banyan tree (ARBFs) as reinforcement in fibre reinforced plastics. The Fundamental properties of ARBFs such as density, tensile strength, chemical composition, crystallinity index, crystalline size, thermal stability, maximum degradation temperature and surface roughness were studied. The chemical analysis results revealed that after the alkalization cellulose content was improved while hemi-cellulose, lignin and wax content were demised. The enhancement in the crystallinity index (76.35% from 72.47%) ACCEPTED MANUSCRIPT A C C E P T E D M A N U S C R I P T and crystalline size (7.74 nm from 6.28 nm) of alkali treated ARBFs were evidenced by the X-ray diffraction analysis. Thermal analysis results confirmed that maximum degradation temperature (368°C) and kinetic activation energy (75.45 kJ / mol) of alkali treated ARBFs had increased from 358°C and 72.65 kJ / mol respectively. The results of scanning electron microscopic and atomic force microscopic analysis exhibited that impurities and wax on the outer surface of the ARBFs were removed after the alkali treatment. All the above finding concluded that ARBFs is the appropriate material to use as a reinfocement in fibre reinforced plastis.
Article
Full-text available
In this work, fillers of waste chicken feather and abundantly available lignocellulose Ceiba Pentandra bark fibers were used as reinforcement with Biopoxy matrix to produce the sustainable composites. The aim of this work was to evaluate the mechanical, thermal, dimensional stability, and morphological performance of waste chicken feather fiber/Ceiba Pentandra bark fiber filler as potential reinforcement in carbon fabric-layered bioepoxy hybrid composites intended for engineering applications. These composites were prepared by a simple, low cost and user-friendly fabrication methods. The mechanical (tensile, flexural, impact, hardness), dimensional stability, thermal stability, and morphological properties of composites were characterized. The Ceiba Pentandra bark fiber filler-reinforced carbon fabric-layered bioepoxy hybrid composites display better mechanical performance compared to chicken feather fiber/Ceiba Pentandra bark fiber reinforced carbon fabrics layered bioepoxy hybrid composites. The Scanning electron micrographs indicated that the composites exhibited good adhesion at the interface of the reinforcement material and matrix system. The thermogravimetric studies revealed that the composites possess multiple degradation steps, however, they are stable up to 300 °C. The thermos-mechanical studies showed good dimensional stability of the composites. Both studied composites display better thermal and mechanical performance compared to neat bioepoxy or non-bioepoxy thermosets and are suitable for semi-structural applications.
Article
Nowadays, as environmental awareness is key issue among researchers, scientific community is looking for natural materials as they are biodegradable, low cost, eco-friendly and also safe for health. Researchers and academicians have found many natural fibers and studied their properties for their sustainable applications in various possible sectors, and studies are also going on. So, in that context several natural fiber like jute, sisal, banana, pineapple, flax, hemp, kenaf, bamboo, cornstalk waste, coir, etc. have been successfully utilized as a reinforcing material in polymer composites by replacing man made synthetic fiber. Apart from traditional natural fibers, scientific community is also looking for locally available natural fibers across the globe in different geographical locations for successful reinforcement in polymer matrix. This will not only decrease burden on traditional fibers and but also at the same time it would be helpful to enrich the rural economy. Natural fiber based composites can be used in different areas such as auto motive industry, construction industry, sports industry and food industry. This study is related with extraction, characterization, surface treatment thermal analysis and activation energy of different uncommon natural fibers available at different geographical locations worldwide. The purpose of this study is to provide a comprehensive knowledge on extraction techniques, treatment methodologies, and properties of these uncommon natural fibers so that these novel materials can be utilized efficiently as a reinforcing material in different polymer matrix. Discussions on traditional natural fibers like Bagasse, Wheat straw, Coir, Pineapple, Banana etc. have been compiled extensively in various review papers but compilation on these new uncommon natural fibers is rare. Thermal analysis along with activation energy evaluation is another aspect which has been given emphasis in discussion because this is also a very important examination to evaluate the thermal stability of these natural fibers.
Article
Environmental awareness across the world has motivated researchers to focus their attention on the use of cellulosic fiber as reinforcement in polymer matrices. Lignocellulosic fibers are an abundantly available resource in all countries, which is cheap and easily renewable. Also, due to their properties, cellulosic plant fibers exhibit a great potential for use in polymer reinforcement. As a result, considerable research and development efforts have been directed towards the use of cellulosic fibers as a reinforcing material in composites. The use of cellulosic fiber reinforced composites has continuously increased during recent years, which benefits the cultivation of fiber plants and the economy of the country. This research area continues to be of interest to both industry and academia, the use of cellulosic fibers in composite applications being investigated throughout the world. Cellulosic fiber reinforced composites are reasonably strong, lightweight, harmless to human health and the environment, and biodegradable, hence they have the potential to be used in various applications. Recent progress in cellulosic fiber composites research has illustrated their great potential as structural components in automobiles, aerospace structures, construction, and building, and so forth. This study is an effort to create awareness about the research works that have been published in the field of natural fiber composites. This review briefly illustrates the main paths and results of major research published in the field of natural fiber reinforced polymer composites. The topics covered include the aspects of fiber treatment to improve the mechanical properties of the composites, manufacturing methods, performance of hybrid composites, effect of laminate configuration, and many different applications of natural fiber composites. By presenting a systematic view of the work performed in this area so far, this review will hopefully serve as a starting point for the development of new ideas in the research on natural fiber polymer composites.
Article
In the current research, four different sequential laminates of Jute (J) and Hemp (H) fibers reinforced hybrid bio-epoxy composites were developed by using the hand lay-up method. The effect of the layering sequence of laminates was evaluated through mechanical properties such as tensile, flexural, interlaminar shear, and impact tests, and physical properties like water absorption, percentage void, and density. The stacking sequence showed minor changes in the hardness of the composites, while the mechanical performance of the hybrid laminates was comparable to the pure hemp laminate. The hybrid composite hemp/jute/hemp showed the highest tensile strength of 65.44 MPa, and a similar trend was observed in flexural results. The mechanical test results revealed that hybridization with hemp fiber as a skin layer and jute fiber as a core layer can improve the tensile and flexural strength, but it will decrease interlaminar shear strength. Thermal stability and fracture morphology are analyzed using thermogravimetric analysis and scanning electron microscopic images.
Article
Due to the environmental awareness, researchers are focusing to replace the synthetic with natural fibers. However, the use of natural fibers has certain limitations like limited to geographical location, hydrophobicity, and compatibility. To overcome these, the current research presents a new novel resource of natural fiber from Morinda citrifolia as a cleaner material and bridges the gap between the materials and sustainability. The fibers were subjected to chemical treatments like NaOH, silane, and nitric acid to improve its compatibility and four different composites were developed, respectively using bio-epoxy to promote the bio-based concept. The physical, mechanical, thermal, viscoelastic, and morphological properties of the fibers and composites were analyzed and studied according to ASTM. The mechanical results revealed that the chemical treatments improved the tensile and flexural strength by 14.372% and 46.716%, respectively. Surface roughness plots showed improved roughness for NaOH and Nitric acid-treated fibers. The thermal stability of the silane and NaOH treated fibers was improved by 6.785% and 5.583% respectively. Lifetime of the best performed composite under dynamic loading is analyzed and studied through fatigue test. Finally, from the results, it was found that Morinda citrifolia fiber is a novel potential resource of sustainable raw material for reinforcement in polymer composites and can be used to develop lightweight structural applications.
Article
In recent days, natural fiber reinforced polymer composite is more popular due to its extensive properties suitable for various potential applications. The attention towards natural fibers is because of low cost, biodegradable, recycla-bility, nonabrasive, combustible, lightweight, and nontoxic properties. However , there is a need for furthermore fundamental knowledge for the raw materials processing and fabrication of composite structures, which is still challenging in current days. Natural fiber sources exist all over the world, which is obtained from animals, plants and minerals. The quality of the natural fibers depends on the extraction methods and different processing techniques. These natural fibers surface characteristics could be enhanced by selecting suitable surface treatment and chemical treatment. These fiber treatments reduce the water intake percentage, improve the adhesive nature, and enhance the overall performance of resulting polymer composites. Among all the chemical treatments, alkaline treatment (NaOH) is the most preferred chemical treatment because of its effectiveness and its low cost. This review article proposes the natural fibers detailed classification, composition, structure , properties, and extraction methods, chemical and surface treatments. We also summarize the previous research work findings on the fibers treatment, properties of natural/natural hybrid polymer composites and natural/synthetic hybrid polymer composites with applications.