You are on page 1of 58

This article was downloaded by: [University of Boras]

On: 29 January 2015, At: 03:48


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Polymer Reviews
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lmsc20

A Review of Natural Fibers Used in


Biocomposites: Plant, Animal and
Regenerated Cellulose Fibers
a a b
Sunil Kumar Ramamoorthy , Mikael Skrifvars & Anders Persson
a
Swedish Centre for Resource Recovery, University of Borås, Sweden
b
The Swedish School of Textiles, University of Borås, Sweden
Published online: 28 Jan 2015.

Click for updates

To cite this article: Sunil Kumar Ramamoorthy, Mikael Skrifvars & Anders Persson (2015) A Review
of Natural Fibers Used in Biocomposites: Plant, Animal and Regenerated Cellulose Fibers, Polymer
Reviews, 55:1, 107-162, DOI: 10.1080/15583724.2014.971124

To link to this article: http://dx.doi.org/10.1080/15583724.2014.971124

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [University of Boras] at 03:48 29 January 2015
Polymer Reviews, 55:107–162, 2015
Copyright © Taylor & Francis Group, LLC
ISSN: 1558-3724 print / 1558-3716 online
DOI: 10.1080/15583724.2014.971124

A Review of Natural Fibers Used in Biocomposites:


Plant, Animal and Regenerated Cellulose Fibers

SUNIL KUMAR RAMAMOORTHY,1 MIKAEL SKRIFVARS,1


AND ANDERS PERSSON2
1
Swedish Centre for Resource Recovery, University of Borås, Sweden
2
The Swedish School of Textiles, University of Borås, Sweden
Downloaded by [University of Boras] at 03:48 29 January 2015

Natural fibers today are a popular choice for applications in composite manufacturing.
Based on the sustainability benefits, biofibers such as plant fibers are replacing synthetic
fibers in composites. These fibers are used to manufacture several biocomposites. The
chemical composition and properties of each of the fibers changes, which demands the
detailed comparison of these fibers. The reinforcement potential of natural fibers and
their properties have been described in numerous papers. Today, high performance
biocomposites are produced from several years of research. Plant fibers, particularly
bast and leaf, find applications in automotive industries. While most of the other fibers
are explored in lab scales they have not yet found large-scale commercial applications.
It is necessary to also consider other fibers such as ones made from seed (coir) and
animals (chicken feather) as they are secondary or made from waste products. Few plant
fibers such as bast fibers are often reviewed briefly but other plant and animal fibers are
not discussed in detail. This review paper discusses all the six types of plant fibers such
as bast, leaf, seed, straw, grass, and wood, together with animal fibers and regenerated
cellulose fibers. Additionally, the review considers developments dealing with natural
fibers and their composites. The fiber source, extraction, availability, type, composition,
and mechanical properties are discussed. The advantages and disadvantages of using
each biofiber are discussed. Three fabric architectures such as nonwoven, woven and
knitted have been briefly discussed. Finally, the paper presents the overview of the
results from the composites made from each fiber with suitable references for in-depth
studies.

Keywords natural fiber composite, biocomposite, plant fiber, animal fiber, bast fiber,
leaf fiber

1. Introduction
Fiber reinforced composites (FRC) have been used in several applications for several years
and the market is growing continuously. It is known that the addition of fibers to polymers
has several advantages, especially the mechanical properties of the composites. Synthetic
fibers such as carbon/glass are reinforced in polymers to be used in high-performance
applications such as automobiles and aircraft industries.1–3 The performance of these com-
posites was improved continuously through rigorous research, often through mixing of

Received May 14, 2014; accepted September 18, 2014.


Address correspondence to Mikael Skrifvars, Swedish Centre for Resource Recovery, University
of Borås, Allégatan 1, SE-501 90, Borås, Västra Götaland, Sweden. E-mail: Mikael.Skrifvars@hb.se
Color versions of one or more of the figures in the article can be found online at
www.tandfonline.com/lmsc.

107
108 S. K. Ramamoorthy et al.

two or more reinforcements/polymers or fillers.4–7 However, these high-performance com-


posites are difficult to recycle as the separation of the components are quite difficult.8–10
Therefore, these composites are often disposed in unsatisfactory ways such as landfills or
incineration which causes a vast environmental impact.8–10 It is imperative to note that most
of these polymers are made from petroleum-based non-renewable resources.
Ecological problems in recent decades have urged the necessity to look for new al-
ternatives which could replace the traditional FRCs with lower environment impact ma-
terials.11–14 This created a renewed interest in natural materials which could be used as
reinforcements or fillers in the composites and are thus referred to as “natural fiber re-
inforced composites” or “ecocomposites.”15 They are also termed as “biocomposites.”16
Several researchers came up with ideas of reinforcing different natural fibers in polymers
to produce eco-friendly composites for several applications which do not require excellent
mechanical properties, such as secondary/tertiary building structures, car door panels, pack-
aging, etc.13 The use of natural fibers as reinforcements in composites has been growing
Downloaded by [University of Boras] at 03:48 29 January 2015

since then and has replaced several synthetic fiber reinforced composites in many appli-
cations such as automotive, marine, aerospace, construction industries, etc.11–14,17 This is
mainly due to extensive research undertaken, its environmentally-friendly/biodegradability
character, low cost, and interesting physical and mechanical properties (low density, good
strength, processing flexibility, high specific stiffness, etc.).14,16
In chronological order, the very first attempts focused on replacing the synthetic fibers
by natural fibers in petroleum-based matrices to form composites.13 The production and
characterization of these composites were studied and the properties were improved sub-
stantially.18–24 These composites gained support due to the environment-friendly element,
considerable reduction of non-renewable resources used in composites, and the replace-
ment of mineral-inorganic materials with natural-organic ones.13 The use of natural fibers in
composite industry created a new alternative for the farmers. The possibilities of using recy-
cled plastics like polyolefin succeeded as it reduced the consumption of non-biodegradable
polymers.25–29 Furthermore, the biodegradable polymers were derived from renewable
resources, i.e., polylactic acid thermoplastic polymer from corn starch, soybean based ther-
moset matrix, etc., and the composites were made completely from renewable resources,
and were thus referred to as “green composites.”30–37
Natural fiber composites were used in airplane seats as early as 1896.38 The use of
natural fibers critically declined and came to a near halt due to the rise in the usage of
standard plastics and their low cost. Few countries like India sustained the use of natural
fibers as reinforcements for composites in applications like pipes, panels, etc.39 The Indian
government encouraged natural fiber reinforced composites for buildings such as Madras
House.40 Critical discussions about the environmentally-friendly materials and preservation
of non-renewable resources regained the interest in natural fiber composites.11 The US
market reports that the composite market was 2.7 billion pounds in 2006 and estimated to
reach 3.3 billion pounds by 2012 with 3.3% annual growth.41 The natural fiber market in
the US experienced 13% growth rate (275 million kilograms) from 1994 to 2004, and the
demand for the natural fibers continues to rise.42 The average global annual market growth
for bio-based matrix from 2003 to 2007 was 38%; Europe saw the highest annual growth
rate of 48%. Bio-based matrix market was estimated at 0.36 million metric ton in 2007
and is expected to reach 3.45 million metric ton in 2020.43 The fiber reinforced market is a
multibillion dollar business and the natural fiber composites take a fair part of it.44
Many articles including reviews, conference proceedings, and books reflect the grow-
ing interest and importance of natural fibers, bio-based matrix, and biocomposites.11–22
Bledzki and Gassan reviewed the cellulose based composites until 1999, and Omar et al.
A Review of Natural Fibers Used in Biocomposites 109

reviewed biocomposites from 2000 to 2010.11,12 Detailed review about biocomposites,


mostly from bast and leaf fibers, and their properties until 2010 have been discussed in
the above-mentioned two papers. There is a lack of detailed discussion on other plant and
animal fibers. New techniques including prediction of composite properties using different
software/methods play an important role in designing the composites.45,46 The potential of
man-made regenerated cellulose fiber is highlighted due to uniformity and superior prop-
erties; these fibers are used as reinforcements in composites.47–52 The detailed review of
the regenerated cellulose fibers and their composites have not been done. This paper aims
at studying the natural fibers such as plant and animal fibers, and regenerated fibers. All
six types of plant fibers such as bast, leaf, seed, straw, grass, and wood are discussed, and
animal fibers such as silk, wool, and chicken feather are discussed. Regenerated cellulose
fibers like Lyocell and viscose are also discussed.
Though the study covers a wide range of fibers for biocomposites, it will certainly be
incomplete due to widespread research in this field, but this paper will hopefully provide a
Downloaded by [University of Boras] at 03:48 29 January 2015

sensible overview of the fibers used for biocomposites.

2. Natural Fibers and Cellulose-Based Fibers


Fibers are broadly classified as natural or man-made fibers. Figure 1 shows the schematic
representation of fiber classification.47 Natural fibers have replaced man-made fibers in
many applications in the last few decades due to low cost, low density, and low tool
wear.11,12 Wide use of natural fibers as reinforcements due to renewability and sustainability
has brought several fibers into the composite field. Natural fibers are categorized into three
types based on their origins; plant, mineral, and animal fibers. Animal fibers such as hair
and silk and mineral fibers have not been widely used as reinforcement fibers. But several
plant fibers have been used widely in biocomposites field for applications in the areas of
automotive, marine and construction. Regenerated cellulose fibers fall between natural and
man-made fibers which have been used as reinforcement fibers in recent times. Natural and
regenerated cellulose fibers, which are key parts of biocomposites reinforcements, have
been discussed in this paper.

Figure 1. Schematic representation of fiber classification.


110 S. K. Ramamoorthy et al.

Table 1
Some plant fiber production and their producers.13,14,53–59

Fiber World production (103 ton) Largest Producers


Bast
Flax 830 Canada, France, Belgium
Hemp 214 China, France, Philippines
Jute 2300 India, China, Bangladesh
Kenaf 970 India, Bangladesh, USA
Ramie 100 China, Brazil, Philippines, India
Leaf
Abaca 70 Philippines, Ecuador, Costa Rica
Curaua >1 Brazil, Venezuela
Pineapple 74 Philippines, Thailand, Indonesia
Downloaded by [University of Boras] at 03:48 29 January 2015

Sisal 378 Tanzania, Brazil


Seed
Coir 100 India, Sri Lanka
Cotton 25000 China, India, USA
Oil Palm 40 Malaysia, Indonesia
Grass
Bagasse 75 000 Brazil, India, China
Bamboo 30 000 India, China, Indonesia

2.1 Plant Fibers


The fibers from the plants can be in the form of hairs (cotton, kapok), hard fibers (coir,
sisal), and fiber sheaves (flax, hemp, jute). The plant fibers are classified depending on
their utility such as primary and secondary. Plants to be used as fibers for primary utilities
include hemp, jute, kenaf, etc., while the byproducts of plants such as coir, pineapple, etc.,
belong to the secondary group.12
There are six types of plant fibers namely bast fibers (flax, hemp, jute, kenaf, and ramie),
leaf fibers (abaca, pineapple and sisal), seed fibers (coir, cotton, and kapok), straw fibers
(corn, rice, and wheat), grass fibers (bagasse and bamboo), and wood fibers (softwood and
hardwood), Fig. 1. Table 1 shows the quantity of few fiber produced and their geographical
distribution.13,14,53–59 The general structure of biofiber is shown in Fig. 2.
The major chemical composition of plant fibers is lignocellulose (cellulose, hemicel-
lulose, and lignin) and the quantity of these components change from plant to plant. This
is due to age, species, and could also vary in different parts of the same plant. These basic
components partially determine the physical properties of the fibers. These polymers could
be distributed unevenly throughout the plant cell wall which makes it difficult to know the
composition and properties of the fibers. Table 2 shows the average chemical composition
of some plant fibers.12,16,60–72 Cotton fiber has the highest amount of cellulose while coir
fiber has the highest amount of lignin.
The unevenness of the chemical composition of the plant is one concern of natural
fibers to be used as reinforcements in composites. Cellulose is the strongest and stiffest
component of the fiber which is linear 1,4-β-glucan polymer consisting D-anhydroglucose
(C6 H11 O5 ) repeating units containing hydroxyl groups (Fig. 3). The OH groups form
inter- and intra-molecular hydrogen bonding making it hydrophilic in nature. The cellulose
A Review of Natural Fibers Used in Biocomposites 111
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 2. Structure of biofiber.14 © Elsevier. Reproduced by permission of Elsevier. Permission to


reuse must be obtained from the rightsholder.

Table 2
Chemical composition of some natural fibers.12,16,60–72

Fiber Cellulose (wt%) Lignin (wt%) Hemicellulose (wt%) Wax (wt%)


Bast
Flax 71.0 2.2 18.6–20.6 1.7
Hemp 70.2–74.4 3.7–5.7 17.9–22.4 0.8
Jute 61.0–71.5 12.0–13.0 13.6–20.4 0.5
Kenaf 31.0–39.0 15.0–19.0 21.5 —
Ramie 68.6–76.2 0.6–0.7 13.1–16.7 0.3
Leaf
Abaca 56.0–63.0 7.0–9.0 20.0–25.0 3.0
Curaua 73.6 7.5 9.9 —
Henequen 77.6 13.1 4.0–8.0 —
Pineapple 70.0–82.0 5.0–12.0 — —
Sisal 67.0–78.0 8.0–11.0 10.0–14.2 2.0
Seed/Fruit
Coir 36.0–43.0 41.0–45.0 0.15–0.25 –
Cotton 82.7 — 5.7 0.6
Oil Palm 65.0 — 29.0 —
Grass
Bagasse 55.2 25.3 16.8 —
Bamboo 26.0–43.0 21.0–31.0 30.0 —
Straw
Rice 41.0–57.0 8.0–19.0 33.0 8.0–38.0
Wheat 39.0–45.0 13.0–20.0 15.0–31.0 —
Others
Rice Husk 35.0–45.0 20.0 19.0–25.0 14.0–17.0

© Wiley. Reproduced by permission of Wiley. Permission to reuse must be obtained from the
rightsholder.
112 S. K. Ramamoorthy et al.

Figure 3. Chemical structure of cellulose.57 © Elsevier. Reproduced by permission of Elsevier.


Permission to reuse must be obtained from the rightsholder.
Downloaded by [University of Boras] at 03:48 29 January 2015

chemical structure remains the same for all the natural fibers while the degree of poly-
merization changes, which influences the mechanical properties of the fibers. It has been
observed that bast fibers have the highest degree of polymerization compared to most of
the other plant fibers.73 Lignin is the phenolic compound which is believed to support the
structure of the plant, and it is also resistant to microbial degradation until disturbed through
physical/chemical treatment. The chemical structure of lignin is not clear until today even
though most of the functional groups and units are identified.56,74 Lignin is believed to be
a binder which links the celluloses to retain the structure of the plant. The wax content of
the fiber plays a crucial role on processing composites as it influences wettability when the
matrix is introduced, furthermore influencing interfacial fiber-matrix adhesion.
The mechanical properties of some natural fibers and made-made fibers are listed in
Table 3.11–12,16,60,62,70–71,75–78 Variation in natural fiber properties are a result of several
factors such as geographical location of the plant, maturity, size, chemical composition,
part of the plant from which fibers are extracted, etc. The variation can also be attributed to
different stages: growth, harvesting, fiber extraction, and storage, and each stage has several
influencing factors. It is demonstrated that the strength of the fiber is also related to fibrillar
angle and physical properties of the fiber.16,80 Small fibrillar angle, small fiber diameter, and
high aspect ratio give high mechanical properties.12,16,80 Owing to low density and relatively
good strength and modulus of natural fibers, these fibers are preferred reinforcements in
several applications. The high strength of kenaf and ramie fibers and the low strength of coir
fiber could be attributed to high and low cellulose content. The modulus of kenaf and ramie
fibers can be compared to that of man-made fibers. The natural fibers are hydrophilic by
nature and have substantial amount of moisture content which influences the mechanical
properties. Table 4 shows the moisture content of some natural fibers and these could
be related to fiber pores, relative humidity, chemical composition, and crystallinity of the
fibers.60–69,79 In general, high amount of cellulose, high relative humidity, high pore volume,
and low crystallinity in fibers tend to cause high moisture content.
The cost of natural fibers is substantially low when compared to synthetic fibers such as
carbon, steel, and glass.75 Performance price ratio (modulus/cost) of natural fibers is several
times higher than that of man-made fibers.16,75 Figure 4 shows the price performance of
some plant fibers. As the result of continuous decline in the availability of natural resources
to produce synthetic fibers, the cost of these fibers is expected to increase. So, renewable
raw material production and contribution to fiber market continue to rise. The data showed
that the production of jute fibers is higher than that of E-glass; it also shows that the amount
of flax and sisal produced together crosses the quantity of E-glass produced.68,81
A Review of Natural Fibers Used in Biocomposites 113

Table 3
Mechanical properties of some natural and man-made fibers.11–12,16,60,62,70–71,75–78

Density Tensile E-Modulus Elongation


Fiber (g/cm3) Strength (MPa) (GPa) at break (%)
Bast
Flax 1.5 345–1100 27.6 2.7–3.2
Hemp — 690 30–60 1.6
Jute 1.3–1.4 393–773 13.0–26.5 1.2–1.5
Kenaf — 930 53.0 1.6
Ramie 1.5 400–938 61.4–128.0 1.2–3.8
Leaf
Abaca 1.5 400 12.0 3.0–10.0
Curaua 1.4 500–1150 11.8 3.7–4.3
Downloaded by [University of Boras] at 03:48 29 January 2015

Pineapple — 413–1627 34.5–82.5 1.6


Sisal 1.4 468–640 9.4–22.0 3.0–7.0
Seed/Fruit
Coir 1.1 131–175 4.0–6.0 15.0–40
Cotton 1.5–1.6 287–800 5.5–12.6 7.0–8.0
Oil Palm 0.7–1.55 248 3.2 25.0
Grass
Bagasse 1.25 290 17 —
Bamboo 0.6–1.1 140–230 11–17 —
Man-made
Aramid 1.4 3000–3150 63.0–67.0 3.3–3.7
Carbon 1.7 4000 230–240 1.4–1.8
E-glass 2.5 2000–3500 70.0 2.5
S-glass 2.5 4570 86.0 2.8

© Wiley. Reproduced by permission of Wiley. Permission to reuse must be obtained from the
rightsholder.

2.1.1 Bast Fibers.


2.1.1.1 Flax. Flax (Linum usitatissimum) is a food and fiber crop mainly produced in
Canada, France, and Belgium. Nowadays, it is predominantly grown for fiber and linseed
oil and accounts 830 × 103 ton, Table 1. It is one of the oldest fiber crops and its fiber is
one of the first to be spun and woven into textiles. Reports claim that the flax fibers were
used for many applications well before 5000 BC in Egypt and Georgia.82,83 High grade
long fibers are usually made into yarns for textiles and the low-grade fibers are used as
reinforcements/fillers in composites. The structure of flax fibers (Fig. 5) is explained in
detail by Charlet et al.84 The average flax plant grows up to 90 cm tall and possesses strong
fibers along the stem which is composed of bark, phloem, xylem, and center void.84 Macro
and microscopic analyses have been performed to see the fiber bundles and hierarchical
organization.84,85 The chemical composition of the flax fibers varied with different authors
which underlines high dependence on the flax fibers used.24,86–88 Flax fiber is considered
for reinforcement as it has high cellulose content and a high degree of crystallinity which
makes it stronger and stiffer.
114 S. K. Ramamoorthy et al.

Table 4
Moisture content of some natural fibers.60–69,79

Fiber Moisture Content (wt%)


Bast
Flax 10.0
Hemp 10.8
Jute 12.6
Ramie 8.0
Leaf
Abaca 15
Pineapple 11.8
Sisal 11.0
Seed
Downloaded by [University of Boras] at 03:48 29 January 2015

Coir 8.0
Grass
Bagasse 8.8
Bamboo 8.9

The tensile properties of flax fiber are given in Table 3. The tensile strength of the fiber
is between 345 and 1100 MPa. The modulus of the fiber is 27.6 GPa and the percentage
elongation of the fiber is between 2.7 and 3.2%. Good quality flax fiber has the best tensile
properties compared to other bast fibers. Charlet et al. have studied the tensile deformation

Figure 4. Ternary Diagram of the Plant Fibers.17 © Elsevier. Reproduced by permission of Elsevier.
Permission to reuse must be obtained from the rightsholder.
A Review of Natural Fibers Used in Biocomposites 115

Figure 5. Multi-scale composite structure of flax.84 © Elsevier. Reproduced by permission of


Elsevier. Permission to reuse must be obtained from the rightsholder.
Downloaded by [University of Boras] at 03:48 29 January 2015

of flax fiber and divided the stress-strain curve into three parts: linear strain (0 to 0.3%),
non-linear strain (0.3 to 1.5%) and end linear strain (1.5% to fracture).89 The reasons
for each of the segments are loading of fiber, alignment of micro-fibrils along machine
direction, and response of the micro-fibrils.89 Similar tensile strain response was obtained
by other researchers.90,91 Variation of tensile properties of flax fibers were studied based
on the fiber location in a stem, plant variety and place of cultivation.84,90,92,93 It showed
that the fibers extracted from the middle of the stem had the highest strength and modulus;
fibers from Hermes variety showed better tensile strength and modulus; and that Hivernal
cultivated plants fibers showed better tensile strength and modulus (Fig. 6).84,90,92,93 The
size distributions of the fiber diameters, lumen diameters, and porosity is shown in Fig. 7.
The effect of moisture absorption and drying of flax fibers were investigated and the results
showed that both the moisture intake and drying affected the tensile properties.94 Defects
in the cell wall, diameter of the fiber, and testing condition (gauge length) are few other
reasons for the variation of tensile properties of flax fibers. A detailed study was done and
it was discovered that all the above factors affected the tensile properties.95,96 Flax fibers
are used as reinforcements in several natural fiber composite applications as it has good

Figure 6. Tensile stress-strain curves for the different varieties of oleaginous flax fibers.54 © Elsevier.
Reproduced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.
116 S. K. Ramamoorthy et al.

Figure 7. Size Distributions of the (a) fiber diameters, (b) lumen diameters, and (c) porosity.93 ©
Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.

tensile properties, availability and is studied extensively, apart from its environmental and
economic benefits. More detailed review of flax fiber and its composites was done by Yan
et al.54
2.1.1.2 Hemp. Hemp (Cannabis sativa) is also a bast fiber crop cultivated mainly in
Downloaded by [University of Boras] at 03:48 29 January 2015

China and France for fiber, oil, and seed. Cannabis family plants were indigenous to central
Asia and were believed to have reached Europe in the Iron Age.97 Today, it is also widely
grown in temperate climate countries such as Chile, North Korea, India, Japan, and many
European Union (EU) countries. EU considers promoting hemp cultivation in member
countries by subsidy and is looking forward for further developments. Nowadays, hemp is
used in several applications such as textile fiber, paper, composite fiber, seed food, oil, wax,
resin, pulp, biofuel, etc., and its usage mainly depends on the grade/quality of the hemp
plant.12,55 The hemp plant secretes small amounts of tetrahydrocannabinol (THC) which is
famous for the narcotic substance marijuana. Since the amount of THC present in hemp
is lower than 0.2%, it cannot be used as a drug like other varieties of Cannabis sativa.
Reports suggest that the oldest hemp fabric dates back to 8000 BC (The Columbia History
of the World). Hemp fiber cultivation was briefly banned between 1971 and 1993 to avoid
misuse and on lifting the ban, it has grown exponentially but still hemp fiber counts less
than 0.5% of total natural fiber production: 214 × 103 ton (Table 1). The average height of
industrial hemp is 10 feet tall and the width is normally very narrow as it is grown close
together. Hierarchical organization of hemp fiber and fiber bundle size were analyzed by
several authors.98,99 It has been reported that the diameter of the average fiber bundle is
25 μm and the length of the average fiber bundle is 25 mm.99 The chemical composition of
the hemp fiber varied with different authors; cellulose content varied from 70.2% to 74.4%
(Table 2). Even though it has several applications, owing to its high strength and stiffness
hemp fiber is also used as reinforcement in biocomposites.100
The variation in the chemical composition due to several factors101 leads to variability
in mechanical properties of the hemp fibers. The plant population affected the hemp fiber
morphology and the physical properties of the fiber which was studied over a 4 year
research by Amaducci et al.102 Similar research was done in the colder country, Sweden,
taking several growing conditions into account.103 The tensile properties of the hemp fiber
are given in Table 3. The tensile strength of the fiber is around 690 MPa and the modulus is
between 30 and 60 GPa. The percentage elongation is around 1.6%. The stress-strain curve
obtained by tensile test had a similar trend to that of jute fibers. The variation of tensile
properties is evident and common to all natural fibers, and this will continue as there are
several factors which are uncontrollable in a large-scale set up.11–18,101 The tensile properties
of the hemp fibers were studied based on the age of the plant and testing parameters.104
It was found that a fully matured plant had better tensile strength than the partly matured
plant; it was also found that the testing parameter and gauge length affected the tensile
strength of the hemp fiber.104 Renowned automotive company, Bayerische Motoren Werke
A Review of Natural Fibers Used in Biocomposites 117

AG (BMW), makes car parts out of hemp fiber reinforced composites.100 A detailed review
of hemp fiber and its composites was done by Shahzad.55

2.1.1.3 Jute. Jute (Corchorus capsularis/ Corchorus olitorius) is mainly grown for
its fiber and it is one of most important natural fibers after cotton.105 It belongs to bast
fiber and is one of the cheapest fibers grown in tropical regions. History of jute dates back
to 206 BC–221 AD; jute paper was discovered in Dunhuang, Gansu Province, China and
was believed to be produced during the reign of the Western Han Dynasty (Jute Paper
from Western Han Dynasty).106,107 Historical documents show that the jute fibers were
used predominantly in India during the era of Mughal Emperor Akbar (1542–1605).108,109
The British East India Company traded jute from India. Since then, jute has been one
of the chief fibers in terms of usage, consumption, production, and availability. The jute
fiber industry grew big during 1800s in Dundee, Scotland.108,109 The global production
of jute fibers is around 2300 × 103 ton and major portion of it is contributed by India,
Downloaded by [University of Boras] at 03:48 29 January 2015

China, and Bangladesh (Table 1). The height range of the jute plant is between 2–4 m
and the fibers are drawn by the retting process. Jute fibers were studied to understand the
fiber structure and the properties.105 The chemical composition of jute fibers was studied by
various researchers and the results show the range of cellulose content between 61%–71.5%
(Table 2). Availability of the jute fiber in large quantities makes it popular reinforcement
among German automotive manufacturers.110
The tensile strength of the jute fibers varies from 393 MPa to 773 MPa according to
different authors (Table 3). The modulus of the fiber is 13–26.5 MPa and the elongation
is 1.2–1.5%. Jute fibers absorb most water in bast fibers (Table 4). It is one of the most
explored fibers for reinforcements; epoxy composites,111 PLA composites,112 polyester
amide composites,113 phenolic composites,114,115 etc. Water uptake and morphological
analysis of the jute fibers were performed.115 Fiber length dependence upon impact strength
in jute-based composites was studied by Razera and Frollini115 Figures 8a and 8b show the
effect of humidity on jute fibers. There was more than 500% growth in use of jute fibers
for car door panels from the year 1996 to 1999 (from 4000 to 21000 tonnes.110 Detailed
review of jute fibers and its composites was done by Mohanty and Misra.56

2.1.1.4 Kenaf. Kenaf (Hisbiscus cannabinus) is cultivated in tropical regions mainly


for its fiber and seed oil. It is primarily considered as a central Africa and southern Asia
crop, and today it is grown primarily in India and Bangladesh (Table 1). It is a new crop
in the US and has shown good potential in biocomposite applications. It is believed that
kenaf has been cultivated since around 4000 BC. In ambient conditions the plant may
grow up to 10 cm/day and matures in 3 months.101,116 As it is a fast growing crop it can
annually yield up to 10000 kg/ha and the latest varieties may yield 30000 kg/hg annually.
The average plant grows 3 m tall with a woody base diameter 3–5 cm. The bast (bark)
constitutes about 40% of the plant where the fibers are extracted and the remaining part is
core wood.117 The bark has high crystalline fiber orientation while the core wood has an
amorphous pattern. The stem diameter is often 1–2 cm. It has environmental and economic
advantages; energy consumption to produce 1 kg of kenaf is 15 MJ while it takes 54 MJ
for glass fibers and the cost of the fibers is considerably low.101 The properties of the kenaf
fiber are similar to that of jute fibers and the structure of kenaf fiber is the same as other
natural fiber. The chemical composition of the kenaf fibers is given in Table 2; it has the
lowest cellulose content (31–39%) and the highest lignin content (15–19%) among bast
fibers. Several authors suggest using kenaf fiber as a reinforcement in composites due to its
superior toughness.57,118–120
118 S. K. Ramamoorthy et al.
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 8. (a) Weight gain and swelling in thickness of jute composites at different humidity levels
and immersed water conditions. (b) Effect of humid conditions on the tensile and flexural strength of
jute composites.114 © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be
obtained from the rightsholder.

Table 3 shows the tensile properties of the kenaf fibers: it possesses a tensile strength
and modulus of about 930 MPa and 53 GPa, respectively. The percentage elongation is
around 1.6%. Several composites were made by reinforcing kenaf fibers in thermoset and
thermoplastic matrix.120–124 Chin and Yousif studied the potential of kenaf fiber composites
in tribological applications125 and Serizawa et al. considered these composites for electronic
applications.126 Few other applications of kenaf fibers are discussed by Ramaswamy et al.127
Chemical analysis of the fibers were performed to see the chemical composition of the kenaf
seed oil and maturing of core/bast.128,129 Properties of the fiber based on growing conditions
A Review of Natural Fibers Used in Biocomposites 119

were studied by Ogbonnaya et al.130 Kenaf fiber and its composites are reviewed in detail
by Akil et al.57
2.1.1.5 Ramie. Ramie (Boehmeria nivea) is the bast fiber that is native to Asia and
today it is mainly produced in China and Brazil (Table 1). It has been grown in China
for many centuries and is commonly referred to as “China Grass”.131 It is an herbaceous
perennial plant belonging to the Nettle family and can be harvested 3–6 times a year. The
plant lasts for around 7–20 years and grows 1–2.5 m tall.132 The presence of gum, pectin,
and other substances in the bark makes the chemical treatment mandatory before the usage
of the fibers. The fibers are coarse and have thick walls.133 The global production of these
fibers is around 100 × 103 ton and it is lowest among bast fibers (Table 1). The ramie fiber
has the least amount of lignin content amount in bast fibers (0.6–0.7%) and has cellulose
content of about 68.6–76.2% (Table 2). The low production and impure nature of the plant
makes it difficult to consider it as reinforcement for composite in large scale.
Table 3 shows the tensile properties of the fiber: tensile strength and modulus is
Downloaded by [University of Boras] at 03:48 29 January 2015

400–938 MPa and 61.4–128 GPa, respectively. The elongation is about 1.2–3.8%. Ramie
fiber is the stiffest among bast fibers (Table 3). The fiber yield of the perennial plant
each year was studied by Hearle and Peters.134 Physical properties of the ramie fibers
were studied by Nam and Netravali135 and the composites were prepared by the same
authors.136 Ramie fibers were reinforced in many thermoset and thermoplastic resins to
produce biocomposites.136–138 Ramie fibers are comparatively less explored than any other
mentioned bast fibers due to the above-stated issues (availability and impure) and most of
the ramie fibers are consumed domestically.
Summerscales et al. and other researchers reviewed bast fibers and the use of these
fibers as reinforcements in composites.132,139–141 Few more bast fibers were also used as re-
inforcements for biocomposites (Table 5). The properties of these fibers, their reinforcement
potential, and the composite properties were not studied in detail.

2.1.2. Leaf Fibers.


2.1.2.1 Abaca. Abaca (Musa textilis) is predominantly grown in the Philippines,
Ecuador, and Costa Rica (Table 1). It belongs to the banana family (Musaceae) and is
mainly cultivated for its fiber. It is native to the Philippines and the first large-scale cul-
tivation was done in Sumatra in 1925. It was later cultivated in America with funding
from the US Department of Agriculture.142 The lifetime of the plant is around 10 years
and the fibers are harvested periodically (two or three times a year) after initial growth.
Cultivation, extraction, and processing of abaca fibers are discussed in detail by Goltenboth
and Muhlbauer.142 The average height of the plant is roughly 4 meters with leaves from
the trunk and the fibers are extracted from these leaves. The chemical composition of the
abaca fibers varies like other natural fibers: cellulose (56-63%), hemicellulose (20-25%),
and lignin (7-9%) (Table 2). Several authors have used these fibers as reinforcements in
composites as it has good mechanical properties.143–149 Other applications include usage in
marine ropes as it is resistant to salt water.
The tensile strength and modulus of the abaca fiber are around 400 MPa and 12
GPa, respectively (Table 3). The percentage elongation is between 3 and 10%. Abaca
fiber reinforced thermoset composites were produced and good tensile properties were
achieved.150,151 Similarly, thermoplastic composites reinforced with abaca fibers possessed
good tensile properties.143–149 Bledzki at al. compared the properties of abaca/PP composites
with jute/PP and flax/PP composites.148 Abaca/PP composites had comparable tensile
strength to jute/PP and flax/PP composites.148 Abaca/PP composites had higher flexural
and impact strength than jute/PP and flax/PP composites.148 The odor emission test done
120 S. K. Ramamoorthy et al.

Table 5
Some more important fibers from plants.14

Fiber Type Fiber Species


Bast China jute Abutilon theophrasti
Isora Helicteres isora
Kudzu Pueraria thunbergiana
Nettle Urtica dioica
Roselle Hibiscus sabdariffa
Sun hemp Crorolaria juncea
Cadillo/urena Urena lobata
Wood (>10,000 species)
Leaf Banana Musa indica
Cantala Agava cantala
Downloaded by [University of Boras] at 03:48 29 January 2015

Caroa Neoglaziovia variegata


Date Palm Phoenix Dactylifera
Henequen Agave fourcroydes
Istle Samuela carnerosana
Mauritius Hemp Furcraea gigantea
Piassava Attalea funifera
Phormium Phormium tenas
Sansevieria Sansevieria
Seed/Fruit/Root Kapok Ceiba pentranda
Sponge Gourd Luffa cylinderica
Broom Root Muhlenbergia macroura

© Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from


the rightsholder.

by Bledzki at al. showed that odor emission concentration is more than 2 times in abaca/PP
composites when compared to jute/PP and flax/PP fibers which makes its usage difficult in
several indoor applications.148 Abaca fibers are vulnerable to pathogens and due to this the
cultivation was affected severely.152–155 Banana fiber, which is very close to abaca fiber, is
reviewed by Jarman.156
2.1.2.2 Curaua. Curaua (Ananas erectifolius) is cultivated in Amazon regions (Brazil
and Venezuela), mainly in the semi-arid areas. Brazil is the one country to produce curaua
fibers for commercial purposes; it produced around 150 tons in the year 2003.157 The curaua
plant leaves are hard and erect with flat surfaces and required minimum of 2000 mm of
precipitation annually. The plant produces about 50–60 leaves annually and the leaves grow
to about 1.5 m in eight months period while the dry fiber content is only about 5–8%.157 The
plant lifetime is about five years and each plant produces up to seven scions a year.158 This
fiber is often replaced by sisal fiber due to low production/cultivation of curaua and a higher
demand for natural fibers. Curaua fibers along with other fibers are used in interior and
exterior automotive applications.159,160 But the commercial use is quite limited considering
the global production of only 150 tons and 5–8 wt% dry content. Marcia et al. studied the
curaua fiber, its density, diameter, tensile strength, moisture absorption, thermal behavior,
chemical composition, and morphology.161 Table 2 shows the chemical composition of the
curaua fiber: cellulose (73.6%), hemicellulose (9.9%), and lignin (7.5%). The high cellulose
A Review of Natural Fibers Used in Biocomposites 121

content fiber of curaua is one of the reasons for its high strength and this fiber is used as
reinforcement in composites.162
The tensile properties of the curaua fiber are given in Table 3: tensile strength (500–1150
MPa), modulus (11.8 GPa), and elongation (3.7–4.3%). Two articles conclude that curaua is
the best natural fiber for composite applications due to its performance.157,163 Thermoset and
thermoplastic composites reinforced with curaua fibers were prepared and the properties
of the composites are discussed by several authors.164–171 Curaua fiber production for
industrial applications with respect to characterization and micro-propagation was discussed
by Leao et al.172 Figure 9 shows the SEM image of the fiber with leaf residue adhered to
surface. Curaua fibers were reinforced in recycled polypropylene and polystyrene and good
composite properties were observed.169,173 Detailed viscoelastic analysis of the curaua
fibers reinforced composites was done by Ornaghi et al. where cole-cole plots were used
to identify the homogeneity of the system.174 Furthermore, curaua fibers are reinforced
in cement for construction applications by d’Almeida et al.175 Curaua fibers have good
Downloaded by [University of Boras] at 03:48 29 January 2015

potential to be used as reinforcements in biocomposites but the low amount of production


hinders the use of fibers in industrial scale and as a result is often replaced by other
plant fibers. A brief review of curaua fiber and its composites was done by the authors in
references.161,162,176
2.1.2.3 Pineapple. The pineapple (Ananas comosus) plant is indigenous to South
America but today it is primarily grown in Southeast Asian countries such as Philippines,
Thailand, and Indonesia. It is also grown widely in Costa Rica, Chile, Brazil, and India. The
botany, cultivation, and utilization of the pineapple were discussed by Collins.177 Pineapple
fiber is obtained from the leaves of the plant, Pineapple leaf fiber (PALF). The plant is
about 1–1.5 m tall and has leaves which are 30–100 cm long. Each plant may have 30 or
more leaves which are dark green in color, sword shaped, and bear spines.178 The fibers
are extracted by mechanical or retting process and thereafter spun to fine yarn.178 The
pineapple fibers are inexpensive and it is not the primary product; fibers are extracted from
leaves which were considered to be waste and now it can be used for industrial applications.
PALF has cellulose content of 70–82% and lignin content of 5–12% (Table 2). Extraction
of PALF is limited due to low production which makes it difficult to be used in industrial
applications.

Figure 9. SEM Image of Curaua fiber with leaf residue adhered to surface.162 © Sage. Reproduced
by permission of Sage. Permission to reuse must be obtained from the rightsholder.
122 S. K. Ramamoorthy et al.

The high cellulose content and low micro-fibrillar spiral angle (14 deg) are associated
with good tensile properties.178 Tensile properties of the pineapple fiber are given in Ta-
ble 3: tensile strength (413–1627 MPa), modulus (34.5–82.5 GPa), and elongation (1.6%).
PALF has the highest modulus among leaf fibers and good fiber modulus which can be
compared to man-made fibers such as Aramid and glass (Table 3). The tensile strength of
good quality PALF is the highest among leaf fibers (Table 3). On the other hand, PALF
has the lowest elongation at break among leaf fibers (1.6%) (Table 3). Like other plant
fibers, PALF is also reinforced in thermosets and thermoplastics.179–184 Matrix from natural
sources were reinforced with PALF to produce green composites.179,184 Thermal properties
such as thermal conductivity and thermal diffusivity of pineapple fiber reinforced phenol-
formaldehyde composites were studied using transient plane source (TPS) technique by
Mangal et al.185 Inclusion of fiber decreased the thermal conductivity and thermal diffusiv-
ity of the composite, and this is applicable to all plant fiber composites.185 The X-ray and
di-electric tests were studied to see the crystallinity of the fibers, orientation of spiral angle
Downloaded by [University of Boras] at 03:48 29 January 2015

and dielectric behavior.186,187 It was found that the PALF had a high degree of crystallinity
(0.63 to 0.68), low micro-fibrillar spiral angle (14 deg), and the dielectric properties were
comparable to that of jute fibers.186–188 The properties and structure of PALF were studied in
detail by researchers.67 The pineapple leaf fibers and its composites are reviewed by Mishra
et al.178
2.1.2.4 Sisal. Sisal (Agave sisilana) is a commercial crop produced mainly in Tanzania
and Brazil (Table 1). It is also widely cultivated in China and Kenya. It is believed that
sisal is native to Central America and its fiber was used already in pre-Columbian times.189
The sisal plant was explained in terms of botany, cultivation, and utilization by Lock.190
The plant grows to about 1 m tall and 28 mm wide with 200–250 leaves. The leaves are
dark green in color, rigid, fleshy, and lance-shaped and grown in a rosette from the stalk.189
The lifetime of the plant is about 7–10 years upon maturity the fibers are extracted from
leaves; each leaf has about 1000 fiber bundles of which only 4% is fiber.191 The fibers are
traditionally used for rope, twine, and as textile fiber. Decortication is the common process
by which the fibers are extracted. These fibers were classified into three types based on the
place of extraction, namely mechanical, ribbon, and xylem.75 The mechanical fibers have
the highest strength among the three while xylem fibers have the lowest strength, and this
is because mechanical fibers are extracted from the periphery of the leaf and have a defined
shape whereas xylem fibers are irregular in shape and have thin-walled cells.75 Ribbon
fibers are intermediate fibers which are extracted from conducting tissues in the median
line of the leaf and have considerable mechanical strength.75 The sisal fiber dimensions and
their mechanical properties were studied by Bisanda and Ansell.192 The fiber extraction
was studied by several researchers.191,193 Table 2 shows the chemical composition of sisal
fibers: cellulose (67–78%), hemicellulose (10–14%), and lignin (8–11%). These fibers are
widely used as reinforcement in composites.58,75,178,192
Several authors came up with different tensile properties of sisal fibers and this variation
was expected in plant fibers. Table 3 shows the tensile properties of the sisal fiber: tensile
strength (468–640 MPa), modulus (9.4–22 GPa), and elongation (3–7%). The sisal fibers are
reinforced in thermoset and thermoplastic matrix like other plant fibers; and the mechanical,
thermal, viscoelastic, dielectric, water absorption, and morphological properties of the
composites were published.194–202 Morphological analysis is very common among natural
fiber composites’ research to study interfacial adhesion, pores, fiber pull-out, etc. Evaluation
of fiber pull-out was studied in detail by Sydenstricker et al.203 Oksman et al. studied
scanning electron microscopy images of sisal-epoxy composites.194 The sisal fiber and its
composites are reviewed by many researchers.58,178,204
A Review of Natural Fibers Used in Biocomposites 123

Other leaf fibers (Table 5) such as banana, henequen, etc., were also used as reinforce-
ments in composites. The characterization of the fibers, properties of the fiber, reinforcement
potential, and composite properties were studied briefly.205–211

2.1.3. Seed/Fruit Fibers.


2.1.3.1 Coir. Coir (Cocos nucifera) is a natural fiber extracted from coconut husk;
mainly produced in India and Sri Lanka (Table 1). Coir is the byproduct of coconut and it
is an inexpensive material in local markets.12,16 The coconut refers to the entire fruit which
has a hard inner shell and outer fibrous material. The fibrous layers are separated either by
manual de-husking or through a crushing machine.212 The fibers are white in color when
the coconut is unripe (white fiber) and the color changes to brown when coconut ripens
(brown fiber).212 These fibers are mainly used to make ropes, twines, doormats, mattresses,
brushes, etc. The coir fibers have cellulose (36–43%), hemicellulose (0.15–0.25%), and
lignin (41–45%) (Table 2), but the chemical composition changes with maturity. These
Downloaded by [University of Boras] at 03:48 29 January 2015

fibers have the highest lignin content among natural fibers (Table 2). A coir fiber cell is
narrow, hollow, stiff, and coarse with thick walls made up of cellulose.212 The coir fiber is the
only plant fiber which is relatively waterproof and absorbs low amount of salt water.212–214
Low water absorption can be related to lower cellulose content and high amount of lignin.
Table 3 shows that the coir fiber has low tensile strength and a modulus of about
131–175 MPa and 4–6 GPa, respectively, but it has high elongation (15–40%). These fibers
can be stretched beyond its elastic limit without rupture due to a micro-fibrillar spiral angle
(45å). Water absorption of coir fiber was explored by several researchers as it possesses the
lowest moisture absorption among plant fibers.213–217 The coir dust was studied by Prasad
with respect to physical, chemical, and biological properties.218 Satyanarayana et al. studied
the coconut fibers with respect to density and electrical resistivity.219 Various matrices
were reinforced with coir fibers and the composite properties were published: namely
viscoelastic,220 physic-mechanical,212,221–223 thermal,69,223 electric,224 morphological,225,226
and sound absorption.227 Varma et al. studied coir fibers and their composites in detail.66,228
2.1.3.2 Cotton. Cotton (Gossypium) is the most famous natural fiber which is native
to tropical and subtropical regions around the world. Today, China, India, and the United
States are the top three cotton producers in the world. The cotton fibers have been used
since 5000 BC and woven cotton fabrics are known from 900 to 200 BC.229,230 The cotton
fibers are grown in a protective capsule (boll) in the cotton plant and harvested on maturity.
The cotton crops and fibers were widely produced and used between 2000 and 1000 BC.231
It is predominantly used as textile fiber and the growth and the demand for the cotton
has been increasing. The development of cotton fibers, their growth kinetics, cytology of
early fiber development, chemical changes during fiber development, and cell wall were
studied by Basra and Malik.232 During the fiber development, it undergoes striking amount
of elongations; 1000 to 3000 times longer than its diameter.232 Genetically modified (GM)
cotton crops were developed to reduce dependence on pesticides; its effect on farmers’
health233 and economic impact234 were studied. The chemical composition of the cotton
fiber is given in Table 2: cellulose (82.7%) and hemicellulose (5.7%). The cotton plant
produces one of the purest forms of cellulose known to man.232 The hydrophilicity of the
cotton fibers are high due to a large amount of cellulose (OH groups). The compatibility
between hydrophilic fibers and hydrophobic matrix is poor which makes fiber surface
modification necessary for technical applications that involve hydrophobic matrixes.
The tensile properties are given in Table 3; cotton fibers have tensile strength between
287–800 MPa, modulus between 5.5–12.6 GPa, and elongation between 7–8%. Mois-
ture absorption and release performance of knitted fabrics made from composite yarns
124 S. K. Ramamoorthy et al.

were studied by water absorption capacity, diffusion rate, and drying rate.235 The perfor-
mance of cotton based composites were studied by Mwaikambo et al.236,237 The comparison
of the mechanical properties of wood and cotton fiber composites were studied, and the
properties of cotton-based composites were improved by an increase in fiber% in the com-
posite while the trend in wood based composites were contrary.238 This was due to the better
interface between fiber and matrix in cotton-based composites than wood-based compos-
ites. Nonwoven cotton-based composites were produced for automotive applications by
Kamath et al.239 Cotton fibers are used in many value-added products and the demand for
the fibers have increased which makes the fibers expensive. Recycled cotton blend fabrics
were used as reinforcement to reduce the cost.240,241 Cotton fiber composites were studied
by Mueller and Krobjilowski 242
2.1.3.3 Oil Palm. Oil Palm (Elaeis guineensis) is the West African tree that is the
main source of palm oil. It is closely related to American oil palm, Elaeis oleifera, which
is also used for oil production. African oil palm is native to west and southwest Africa,
Downloaded by [University of Boras] at 03:48 29 January 2015

mainly between Angola and Gambia whereas the American oil palm is native to Central
and South America. The use of oil palm dates back to 3000 BC in Africa.243 Today,
Elaeis guineensis is the highest yielding edible oil crop and is mostly cultivated outside
the African continent; Malaysia and Indonesia produce most of the world supply.244 Oil
palm fiber (OPF) is the secondary product which is extracted from empty fruit bunches
which were used as reinforcement in biocomposites; it is also possible to extract fibers
from other parts of the tree but the yield is very low compared to that of fruit bunch.245,246
Empty fruit bunches create waste disposal problems, while fibers can be extracted from
this waste. The retting process is used to extract the fibers and water retting is most
common.247 The more sophisticated machine extraction was developed by Jayashree et al.248
The chemical composition, anatomy, lignin distribution, and cell wall structure was studied
in detail by Abdul Khalil et al.244,249,250 These fibers are hard and tough and their properties
similar to coir fibers.250 The OPF consists of around 65% cellulose and 29% hemicellulose
(Table 2).
OPF has tensile strength about 248 MPa, modulus 3.2 GPa and elongation 25% (Ta-
ble 3). The environmental exposure of plant fiber composites has a big impact on mechanical
properties of the composites and this was studied by Hill et al.251 The effect on the mass
loss, tensile, and flexural properties were evaluated and it was found that the mechanical
properties fell drastically. Water sorption of OPF reinforced composites was studied by
Sreekala et al.252 Viscoelastic analysis of OPF composites was studied by Jacob et al. and
storage modulus was improved.253 Figures 10a and 10b show the SEM images of OPF.
Unlike cotton fibers, OPF has only few applications such as ethanol production254, gasi-
fication,255 and biocomposites. Many researchers have prepared composites reinforcing
OPF in various thermoset and thermoplastics; a detailed review was done on OPF and its
composites by Shinoj et al.256
Other leaf/fruit fibers such as kapok236,237,257,258 and sponge gourd259–263 were used as
reinforcement in biocomposites.

2.1.4 Grass.
2.1.4.1 Bagasse. Bagasse is the fibrous residue that remains after sugarcane stalks
are crushed to extract juice. It was principally a waste material that causes disposal cost
for sugar mills. Wet bagasse consists of 30% sugarcane which means 3 tons of bagasse is
produced for each 10 ton batch of sugarcane. The bagasse is comprised of 49% moisture,
48.7% fiber, and the rest consists of soluble solids.264 Currently, bagasse is used as a raw
material in some applications such as bioethanol,265–269 combustion,270–273 cellulose source
A Review of Natural Fibers Used in Biocomposites 125
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 10. (a) SEM images of transverse sections of OPF(4x). F: fiber, L: lacuna. Permission to reuse
must be obtained from the rightsholder. (b) SEM Images of Longitudinal Section of OPF (750x).256
© Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.

for dissolving pulp, value-added products,274 and biocomposites. The low calorific value
and the low level of sucrose left in the bagasse makes energy and bioethanol productions a
low efficiency process. The world bagasse production is high (75 × 106 ton) and is mainly
produced in Brazil, India, South Africa, and China (Table 1). The cellulose content in
bagasse fiber is around 55.2% while the hemicellulose and lignin content is about 16.8%
and 25.3%, respectively (Table 2).
The bagasse fiber has inferior tensile properties when compared to bast and leaf
fibers; strength (290 MPa) and modulus (17 GPa) (Table 3). Different composite processes
like extrusion, compression molding, etc., were adopted to produced bagasse reinforced
thermoset and thermoplastic composites.275–278 The comparison between compression and
injection molding with respect to mechanical behavior and microstructural analysis of
bagasse composites was done by Luz et al.276 Preparation and properties of recycled
PE/bagasse composites were studied by Lei et al.,279 and impact properties of bagasse
126 S. K. Ramamoorthy et al.

composites was studied by Stael et al.278 and dimensional stability of esterified bagasse
composites was analyzed by Hassan et al.280 Bagasse reinforced cement composites were
prepared and analyzed by Bilba et al.281 The potential of the bagasse fiber and its composites
were studied in detail.264,282–284
2.1.4.2 Bamboo. Bamboo (Bambusa Shreb) is a perennial plant mainly grown in Asian
countries such as India, China, and Indonesia (Table 1). It has more than 1,250 species and
the high quality bamboo is stronger than steel; it is used as building material.285 It belongs to
the grass family and the fibers extracted are of two kinds: natural original bamboo fiber and
bamboo pulp fiber (regenerated bamboo viscose fiber). The world production has reached
30 × 106 ton and bamboo is used in the textiles and paper industry.286–288 Microstructural
characterization of bamboo was studied by Ray et al. where the fracture of fiber and cross-
section of the fibrils was shown.289 The chemical composition of bamboo fiber is given in
Table 2; cellulose (26–43%), hemicellulose (30%), and lignin (21–31%).
The tensile properties of the bamboo fiber are given in Table 3; strength (140–230 MPa)
Downloaded by [University of Boras] at 03:48 29 January 2015

and modulus (11–17 GPa). Green composites were prepared by reinforcing bamboo fibers
in polylactic acid (PLA) and their properties were evaluated.290 Thermal conductivity of
bamboo/PLA were evaluated by Takagi et al. and it was found out that these composites have
lower thermal conductivity compared to that of wood.291 Xiao et al. studied the design and
construction of modern bamboo bridges.292 Computer simulations of the field test were done
and it showed the trend of mid-span deflection.292 Modern bamboo structures were discussed
by several authors in conferences and a book was published based on several papers.293 The
durability of bamboo-based fibers were studied by cyclic loading and the results showed
that the composites were produced with good fatigue resistance.294 Mechanical properties,
crystallization, and interfacial morphology of bamboo/PP composites were studied by Chen
at al.295,296 A detailed review of bamboo fiber reinforced structural concrete elements was
done by Ghavami.297 Bamboo fiber and its composites were reviewed in detail by Khalil
et al.298 The review based on structure and properties of bamboo fiber and its composites
was also done.299

2.1.5 Straw. Straw is an agro-residue or a byproduct obtained from cereal plants when
grain and chaff have been removed. The straw is obtained from several cereal crops such
as rice, wheat, barley, oats, and rye. It is normally gathered and stored in a bale (straw
bale) and is used in various applications: namely animal feed,300 biofuels,301 biomass,302
construction material,303 paper,304 packaging,304 etc. The straw is also gaining importance
in biocomposite research.
2.1.5.1 Rice Straw. Rice is the most widely consumed staple food in the world, espe-
cially in Asia. The genetic evidence shows that rice originated between 6200 and 11500
BC.305 It is usually an indigenous annual plant which is native to Asia, but today it is
grown worldwide. Straw makes up about half of the yield of rice. Rice production, mor-
phology, growth, development, and harvesting were clearly discussed by Datta.306 After
harvesting the rice is separated from the straw either manually or by a machine process. The
rice straw consists of 41–57% cellulose, 33% hemicellulose, and 8–19% lignin (Table 2).
This lignocellulosic fiber could be used as reinforcement in biocomposite production. Rice
straw was mainly reinforced in thermoplastic polymers such as polyvinyl chloride, poly-
hydroxybutyrate, polyethylene, and polypropylene.307–310 Tensile, impact, bending, and
viscoelastic properties of rice straw based composites were evaluated.307–311 The tensile
strength and modulus of rice straw fibers were 74.6 MPa and 3.3 GPa, respectively.311
Although it is very low compared to other plant fibers, straw could be used in low-stress
applications. It was concluded that straw-based composites have potential to be used as a
A Review of Natural Fibers Used in Biocomposites 127

core reinforcement material for structural board products.311 Rice straw based composites
were also tested for acoustic properties and it was found that these composites could be
suitable for sound absorbing insulation material in wooden construction.312 Rice straw
can partially replace the wood particles in the composites without reducing the bending
strength.312
2.1.5.2 Wheat Straw. Wheat is the third most produced cereal after corn and rice
with an annual production of about 607 million tons.313 Wheat is grown worldwide with
USA, China, and India being the major producers.313 It is a food crop which could also be
used to produce alcoholic beverages and biofuel. The production, utilization, development,
and harvesting of wheat were discussed in detail by Gooding and Davies.314 Grains are
separated from straw on harvest and the straw is considered as a byproduct or residue. Tra-
ditionally, wheat straw is used as cattle feed and construction material.303 Table 2 shows the
chemical composition of wheat straw: cellulose (39–45%), hemicellulose (15–31%), and
lignin (13–20%). The researchers have used this lignocellulosic material to produce bio-
Downloaded by [University of Boras] at 03:48 29 January 2015

composites. These fibers were reinforced in polypropylene and their dispersion was studied
with an increase in the fiber loading.315 Mechanical properties, flowability, and fracture
surface morphology of the composites were evaluated.315 Furthermore, wheat straw/PP and
jute/PP composites were compared; superior properties were obtained by lightweight wheat
straw/PP composites.316 Morphology (Fig. 11) and thermal properties of wheat straw based
biocomposites were studied and it showed that the composites could be used in packaging
and medical science.317 A detailed study of wheat straw/PHBV composites was done by
Avella et al. where the mechanical, thermal, and biodegradation properties of the composites
were explored.318 The rate of PHBV crystallization increased on addition of wheat straw
fibers due to nucleating effect and these fibers did not reduce the biodegradability rate.318
Wheat straw was used as reinforcement material by other researchers and the composite
properties were discussed.319–320 High-performance composite panels were produced by
Han320 and the mechanism of bondability between fiber and resin was discussed.
2.1.5.3 Barley Straw. Barley is the fourth most produced cereal crop in the world and
the world’s biggest producers are Russia, Ukraine, France, and Germany. Barley production,
harvesting, processing, and utilization were discussed in detail by Shands et al.321 Ullrich
discussed about the improvement and uses of barley.322 It is a food crop which is also used as
animal fodder, in beer production, distilled beverages, and health foods. Large quantities of
barley straw were not used until recently but the potential of barley straw has been explored
for energy production and composite material preparation. Researchers used barley straw
as reinforcement in biocomposites.323 The chemical composition of the barley straw was
studied; cellulose (48.6%) and lignin (16.4%).324 Bouhicha et al. investigated the barley
reinforced composite material for construction in Algeria; the effect of fiber length and
fiber fraction on shrinkage was studied along with its mechanical properties.323 Thermal
conductivity of the composite was studied and the results showed that straw lowers the
thermal conductivity.325 The comparative study of mechanical properties between wheat
and barley straw fiber reinforced composites revealed that barley fiber composites had lower
compressive strength than wheat straw fiber composites.325
White and Ansell studied straw-reinforced polyester composites.324 Structural and
thermal characterization of various straws such as rice, wheat, rye and barley were studied
by Sun and Sun.326

2.1.6 Husk.
2.1.6.1 Rice Husk. Rice husk (or rice hull) is the hard protecting cover of the rice
grain. After rice harvesting, the grain together with husk undergoes further processing to
128 S. K. Ramamoorthy et al.
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 11. SEM images of the wheat straw: (a) cross section, (b) microfibers, and (c) tem image of
wheat straw nanofibers.317 © Elsevier. Reproduced by permission of Elsevier. Permission to reuse
must be obtained from the rightsholder.
A Review of Natural Fibers Used in Biocomposites 129

Figure 12. TGA analysis of barley husk, coconut shell and softwood.333 © Elsevier. Reproduced
Downloaded by [University of Boras] at 03:48 29 January 2015

by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.

remove the hull. The husk separation was a manual process until the late 1800s, after which
the modern rice hulling machine was invented. The rice husks could be used as filler in
construction, insulation material, fertilizer, or fuel. The husk is formed from hard materials
including silica and lignin. Amorphous reactive silica is obtained on burning rice husk
(ash) which could be used in various applications in material science.327 The chemical
composition of rice husk is given in Table 2: cellulose (35–45%), hemicellulose (19–25%),
and lignin (20%).
Rice husk based composites were prepared and their properties were evaluated with re-
spective to hygrothermal aging, mechanical, morphological, and thermal properties.328–331
The study showed that the potential of rice husk as a filler was lower than that of talc as
the processing of talc composites were easier than that of rice husk composites.329 The
talc composites showed higher tensile strength, E-modulus, flexural modulus, and impact
strength than rice husk composites.329 Similar inferior properties of rice husk composites
were obtained by Yang et al.330 The enhancement of processing of rice husk/PE compos-
ites was investigated by Panthapulakkal et al.;332 water absorption, swelling, density, and
extrusion rate were also studied. The limitation in mechanical properties makes the rice
husk undesirable as filler in several applications.
There was limited research on composites based on husk from other grains; Barley
Husk,333 Corn Husk,334,335 Wheat Husk,336 and Rye Husk.336 The moisture absorption of
barley husk was lower than that of the coconut shell which makes barley husk a possi-
ble reinforcement for moisture sensitive applications.333 The tensile strength and impact
properties of barley husk composites were better than that of coconut shell and softwood
composites.333 Figure 12 shows that the thermal stability of barley husk (235◦ C) was com-
parable to softwood fiber (245◦ C) and it was higher than that of a coconut shell (195◦ C)
which makes husk a possible alternative filler for softwood in composite applications.333
Figure 13 shows the comparison of the particle size distribution of barley husk with coconut
shell and softwood. Elemental and FTIR analyses of barley husk are shown in Figs. 14a
and 14b. Huda and Yang used corn husk as fillers in biocomposites where the tensile, flex-
ural, and impact properties were investigated and compared to that of jute composites.334
Bledzki et al. studied the thermal, physical, chemical, and surface properties of wheat and
rye husk composites.336 It was found that the rye husk (180◦ C) has lower thermal stabil-
ity than wheat husk (205◦ C).336 Surface morphology of these two husks were studied by
130 S. K. Ramamoorthy et al.

Figure 13. Particle size distribution of barley husk, coconut shell and soft wood.333 © Elsevier.
Reproduced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.
Downloaded by [University of Boras] at 03:48 29 January 2015

SEM which showed that rye husk had smoother surface than that of wheat. The chemical
composition of these husks was investigated and the results showed that wheat husk had
higher cellulose (36%) and lignin content (16%) than that of rye (cellulose-26%, lignin-
13%). The tensile and impact strength of wheat husk composites were better than the
composites with rye husk fillers.336

Figure 14. (a) Elemental analysis of barley husk. (b) FTIR analysis of barley husk and coconut
shell.333 © Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained
from the rightsholder.
A Review of Natural Fibers Used in Biocomposites 131

2.1.7 Wood Fiber. There are two types of wood reinforcements, namely, softwood (gym-
nosperm) and hardwood (angiosperm). Hardwood is from a deciduous tree which loses its
leaves annually while the softwood is from conifer ones which do not shed their leaves
throughout the year. Hardwoods are denser than softwoods with exceptions like balsa wood.
The growth rate of softwood is faster than that of hardwood. Both of these are used for
several applications but hardwood is less frequently used than softwood.
2.1.7.1 Softwood. Softwoods such as Pinus (Pines), Picea (spruces), and Larix
(larches) are relatively less complex than hardwood in terms of cells present. All soft-
wood species have two main cell types, longitudinal tracheid and ray parenchyma. It might
contain resin canals, ray tracheid, and longitudinal parenchyma depending on the species.
The softwood anatomy is straightforward in most cases, as 90% or more of softwood vol-
ume is composed of longitudinal tracheid which transports water and gives mechanical
strength to the wood.
Downloaded by [University of Boras] at 03:48 29 January 2015

2.1.7.2 Hardwood. Hardwoods (aspen, birch) are more complex when compared to
softwoods. Unlike softwood, the roles of conduction and support are carried out by different
cells in hardwood. The presence of specialized water-conducting cells, vessel elements,
is a primary feature that distinguishes hardwood from softwood. These vessel elements
are narrow tubes and open at the ends, and are called pores; therefore, hardwood is called
“porous wood.” Hardwood has ray parenchyma which varies in the structure and dimensions
in different species. It may contain longitudinal parenchyma and other cells depending upon
the species.
The lignin content varies between these two types of wood; spruce (softwood)–28% and
aspen (hardwood)–18%. Softwood is preferred for composite applications as it has higher
aspect ratio than hardwood. The properties and cell types of both softwood and hardwood
was discussed in detail.337 Morphological analysis was done on hardwood and softwood.338
The composite processing with wood reinforcement and the mechanical properties of the
composites were discussed.337,338
The wood reinforcement was mostly reinforced in thermoplastic composites, especially
PP.339-344 Softwood composites had better stiffness than hardwood composites and this
could be due to higher lignin content in softwood compared to that in hardwood.342,344
But hardwood composites showed better tensile strength, impact strength, and elongation
which could be attributed to higher cellulose content.342 A comparative study of wood
fiber reinforced PLA and PP composites showed that fibers dispersed better in PLA than
in PP.343 Fiber morphology and composite properties were also investigated.343 The wood
fibers were reinforced in HDPE and the composite properties were studied.341 Detailed
review of wood composites was done by review articles and books.338,344–346
There are other plant fibers which were used in composite production and some of
them are listed in Table 5. The most commonly used fibers are described in detail and
out of these bast and leaf fibers are widely preferred in industrial applications. These
plant fibers were well explored in terms of chemical composition, mechanical and ther-
mal properties, and reinforcement potential. There are many books which describes the
potential of the plant fibers in composites.337,347–350 The industrial applications of natural
fibers with respect to structure, properties and technical applications were discussed.351
Several fibers, especially bast and leaf fibers, and its composites were reviewed previ-
ously.54–58,66,156,161,162,176,178,204,228,242,256,264,282–284,299,345 A general review on plant fiber
and its composites was done by several researchers.13,16–19,22–26,139,140 Bledzki and Gassan
have reviewed plant fiber reinforced composites until 1999.11 Faruk et al. have reviewed
the plant fiber reinforced composites from 2000 to 2010.12
132 S. K. Ramamoorthy et al.

Figure 15. (a) Raw cocoon silks; (b) side view of silk fiber.21 © Elsevier. Reproduced by permission
of Elsevier. Permission to reuse must be obtained from the rightsholder.
Downloaded by [University of Boras] at 03:48 29 January 2015

2.2 Animal Fibers


Animal fibers such as silk, wool, hair, feathers, etc., are the second most important source
of natural fiber after plant fibers for reinforcement in composites. It is also the second most
important natural fiber in terms of availability after plant fibers. There are several sources
of each type of animal fibers, i.e., wool is obtained from sheep, alpaca, angora, bison,
cashmere, muskox, etc. Similarly silk, hair, and feather are picked from several sources.
Researchers have analyzed the animal fibers’ potential as reinforcement in composites.351
Generally animal fibers are costlier than plant fibers and the availability is lower than plant
fibers which make them expensive for normal applications. These composites could be used
in more sophisticated applications such as biomedical applications. Silk and wool are used
in textile industries for many purposes and it could be expensive to be used as reinforcement
in composites but chicken feathers constitute waste which is obtained from slaughter house
and could be used as reinforcement. The chemical composition and the mechanical and
thermal properties of these fibers are crucial to be used as reinforcement. Like the plant
fibers, there are variations in the composition and the properties of animal fibers which
could be an issue. There are few journal articles about each of the above-mentioned fibers
as reinforcement but there is no review paper on animal fibers as composite reinforcement
till date.

2.2.1 Silk. The potential of silk is outstanding due to its structure and properties. It consists
of highly structured proteins and a wide range of properties namely high tensile strength,
high elongation, and resistance to chemicals.351 Silk is obtained from several sources and
its properties vary. Several insects and most of the larvae of the butterfly species (approx.
140000 Lepidoptera) produce silk during metamorphosis.352 Silk is also obtained from
spiders (approx. 40000 species) and they spin seven different types of silk with unique
properties in their lifespan.351 Out the these fibers, Mulberry silk (Bombyx mori) from
silkworm and dragline silk (Nephila) from spider were widely discussed and used. This
paper focuses on these two types of silk, mulberry silk and dragline silk.
The mulberry silk (Bombyx mori) (Fig. 15) is the most commonly used silk and it is
the only silk that is popularly used for commercial applications. The two main components
of mulberry silk are fibroin (fiber) and sericin (gum). Secricin accounts for about one-
third of mulberry silk which is washed off and the fibroin is used as fiber in several
applications. The fibroin consists of two types of amino acids with 1:1 fraction in the
fiber, light chain containing 262 amino acids and a heavy chain with 5263 amino acids.351
A Review of Natural Fibers Used in Biocomposites 133

The amino acid composition is 43% glycine, 30% alanine, and 12% serine. The ratio of
crystallinity changes with change in alanine content. Mulberry silk fiber is semi-crystalline
and the degree of crystallinity is between 38 and 66% (Table 6). It shows submicron-
scale fibrillar substructure with a diameter of approximately 120 nm. The fibers could
be used below 170◦ C without degradation of fibers and thermal degradation occurs at
250◦ C (Table 6). The cocoon is intended to give shelter to the larvae during pupation.
Accordingly, the silk fibroin is very well suited to withstand the elements and this is
reflected in its chemical stability; this fiber is stable against many chemicals; insoluble
in most alcohols and acetone, resistant against mild acids, and absorbs less water. Even
hydrochloric acid needs a couple of hours of exposure for hydrolysis which takes place in
amorphous regions. The structure, crystallinity, and chemical properties of mulberry silk
were discussed in detail.351 Its production and extraction are discussed in section 4.2. The
processing and applications of mulberry silk were explained in detail.351 The mulberry
silk has tensile strength of 0.6 GPa which is higher than that of most of the plant fibers
Downloaded by [University of Boras] at 03:48 29 January 2015

(Table 7). Mulberry silk was reinforced in different blends of HDPE, LDPE, and natural
rubber.358 The tensile, tear, and morphological properties of the composites were analyzed.
It has biocompatibility, support of cell proliferation, and mechanical stability; therefore, it
is used in tissue engineering to produce scaffolds.351
The dragline silk (Nephila) is one variety of the silk that is produced by a spider in
its lifespan. This fiber consists of two highly repetitive proteins, namely, spidroin I and
spidroin II. It is the toughest and biggest fiber in the Nephila spider’s silk set.351 It is
semi-crystalline and exhibits lower degree of crystallinity than mulberry silk (Table 6), and
the fibers are stable up to 150◦ C, above which degradation starts. The tensile strength of
the dragline silk fiber is higher than that of polyamide fibers and comparable in strength
to several synthetic fibers (Table 7). It has a tensile strength of 1.1 GPa is comparable
to that of high-tensile steel (1.5 GPa) while the density of the dragline silk fiber is much
less. The extensibility of the dragline silk fiber is about 30% while steel possesses that of
only 0.8%. High tensile strength and extensibility makes this fiber unique and there are
attempts to make fibers with these qualities. Energy dissipation and mechanical hysteresis
are important properties of dragline silk fiber and the results show that 65–68% of stretching
energy is dissipated.351 A detailed study was done on the structure, properties, influence
of water and temperature, production, extraction, and processing.351 Mechanical, thermal,
and morphological properties of dragline silk reinforced composites studied by Cunniff
et al.354 High strength and elongation makes it a potential reinforcement option in composite
applications. The biggest drawback of these fibers is that they cannot be produced in large
quantities due to practical reasons.351
The silk fibers, their properties, and their applications are discussed by several au-
thors.359–362 Silk reinforced composites have not been in the interest of many due to high

Table 6
Properties of two types of silk.351,353–356

Degree of Maximum use Thermal


Fiber crystallinity (%) temperature (◦ C) degradation (◦ C)
Mulberry Silk 38–66 170 250
Dragline Silk 20–45 150 234
134 S. K. Ramamoorthy et al.

Table 7
Tensile strength and extensibility of two types of silk.357

Fiber Tensile Strength (GPa) Extensibility (%)


Mulberry Silk 0.6 18
Dragline Silk 1.1 30

cost and low availability. Low availability of the fibers makes it difficult to produce com-
posites in industrial scale.

2.2.2 Wool. Wool fiber is commonly a textile fiber which is obtained from sheep, goat,
camel, rabbit, and certain other mammals; sheep wool is the most commonly used variant in
Downloaded by [University of Boras] at 03:48 29 January 2015

commercial operations. These fibers are named based on the source of their origin such as
alpaca, angora, bison, cashmere, sheep wool, qiviut, etc., and the properties of these different
types of wool differ. Wool obtained from alpaca is lighter in weight and warmer than sheep’s
wool. The thickness of the alpaca wool fiber is between 12 and 29 μm depending upon the
quality of the fiber. The color of the alpaca fiber is mostly white, but it also comes in shades
of black and brown. Angora wool fiber is thin and soft wool fiber obtained from angora rab-
bit. The thickness of these fibers is around 12–16 μm. Bison wool fiber is separated into two
types based on their thickness and softness; coarse guard hairs (around 59 μm) and soft down
(18.5 μm). Cashmere wool is luxurious soft wool fiber from cashmere goat. The fiber must
be under 18.5 μm in diameter and be at least 3.175 cm long for it to be considered as cash-
mere. Qiviut wool is expensive smooth fiber obtained from muskox. They are 5 to 8 cm long
and 15 to 20 μm in diameter. Qiviut is approximately 8 times warmer than sheep’s wool and
does not shrink in water. However, the above-mentioned types of wool are used lesser than
sheep wool due to availability and price. Sheep wool is more versatile a choice to be used in
felting.
The annual production of sheep wool is approximately 1.2 million metric tons and
90% of the same is consumed by the textile industry.363,364 The major producers are
Australia, New Zealand, and China. The wool is cleaned to remove the “wool grease”
before being used as textile fiber. It is mainly composed of keratin (animal protein).365
Keratin is more susceptible to chemical damage and unfavorable environmental conditions
than the cellulose in the plant fibers.365 Wool’s scaling and crimp make it easier to spin the
fleece which is used in textile fabrics. It is mainly used for insulation purpose as it holds
the air and retains the heat. Wool fibers are hydrophilic like plant fibers and absorb water
to one-third of its weight.366 The nature, fineness, and color of the wool fiber change with
breed types but the most common wool is creamy white in color. The thermal properties of
the wool fibers are different from that of cotton and other synthetic fibers; the rate of flame
spread is low, the rate of heat release is low, and heat of combustion is low. A char which
is self-extinguishing is formed during combustion. Wool fiber is coarser than other textile
fibers such as cotton and silk; and the length of the fiber is dependent on coarse behavior.365
The breathability, heat generation, structure, felting, shrinking, and thermal insulation of
wool are discussed in detail by Arunkumar et al.366
It was suggested that the waste wool could be used as reinforcement instead of virgin
wool due to high cost of virgin material.366 Wool industries produce a lot of waste which
could be used as reinforcement/filler in manufacturing composite. Conzatti et al. produced
composites by reinforcing short wool fibers in polypropylene fibers.367 The morphological,
A Review of Natural Fibers Used in Biocomposites 135

thermal, and mechanical properties of wood fiber composites were analyzed. Similar re-
search was done by reinforcing raw wool fibers in polyester resin and their properties were
studied and compared to theoretical models.368,369 Homogeneous distribution of wool fibers
in composites was achieved through melt blending.369 The wool fibers could be potential
reinforcement for composites for certain applications.

2.2.3 Chicken Feather. Chicken feather (Fig. 16) is waste from poultry industry and it
constitutes 4 billion pounds annually in the Unites States.370 These feathers are dumped
in landfills at a cost of $30/ton or sold at a low price ($250/ton) as livestock feed.370 The
poultry feather cannot be used as feed in the European Union as it banned the practice
since 2001 due to health concerns. The high volume of feathers could be consumed by
composite industries which cannot be consumed by many other industries. It is reported
that chicken feathers exhibit unique properties such as low density and good thermal and
acoustic insulation.370 The chicken feathers consist of 91% keratin (protein), 1% lipids,
Downloaded by [University of Boras] at 03:48 29 January 2015

and 8% water.366 The chemical composition, elemental analysis, morphological structure,


aspect ratio, apparent specific gravity, chemical durability, and thermal insulation of these
feathers were studied by Arunkumar et al.366 Mechanical properties and the structure of
the feathers were reported previously.371 The study shows good potential of these fibers as
reinforcements.
Chicken feathers were used as whole fibers or processed to separate keratin fibers
which were used as reinforcements.372 These feathers were used as reinforcement by

Figure 16. Typical chicken feather fiber.21 © Elsevier. Reproduced by permission of Elsevier.
Permission to reuse must be obtained from the rightsholder.
136 S. K. Ramamoorthy et al.
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 17. (a) Flight chicken feather fiber; (b) down chicken feather fiber; (c) chicken feather fiber;
(d) cross section of chicken feather fiber.21 © Elsevier. Reproduced by permission of Elsevier.
Permission to reuse must be obtained from the rightsholder.

several researchers.372–375 The whole chicken feather reinforced polypropylene composites


showed higher tensile strength, E-modulus, and flexural strength than processed feather
composites.372 They also have higher sound absorption than plant fiber composites.372
Mechanical and thermal properties of PLA composites reinforced with chicken feathers
were studied by Cheng et al.373 The acoustic and electrical properties of chicken feather
reinforced composites were studied in detail.374,375 The study revealed that the composites
can be used for printed circuit boards (PCB).375 Figure 17 shows the microscopic image of
chicken feather. Detailed study of chicken fibers and their composites was done by Kock.370
The fibers that are obtained from hair, fur, and feather of various animals and birds
could be used as reinforcement in composites. The potential of human hair as reinforcement
has been explored and their composite properties were presented.376

3. Regenerated Fibers
Regenerated fibers are produced from natural source with human interference. Several re-
generated cellulose fibers were produced from wood pulp such as Lyocell, viscose, rayon,
and modal which have been used as reinforcement in composites.47–52 The inspiration for
the industrial fibers is credited to English physicist, Robert Hooke, who discussed the pos-
sibility of imitating silkworm “wires” in the 1600s.47 Cellulose was discovered in 1839 by
French professor, Anselme Payen, and he thereafter went on to discover pectin and dextrin.
A Review of Natural Fibers Used in Biocomposites 137

In the early 1850s, textile fibers from plant cellulose were made successfully using nitric
acid. In 1855, George Audemars got a patent for his work on producing cellulose fibers.
The main drawback of his cellulose fiber was that the fibers were very flammable. In 1891
British scientists Charles Cross, Edward Bevan, and Clayton Beadle discovered that cotton
or wood cellulose could be dissolved as cellulose xanthate following treatment with alkali
and carbon disulphide. Then, the solution was coagulated in ammonium sulphate and later
converted to cellulose using dilute sulphuric acid. These fibers are named “viscose.” In
the early 1900s the viscose production process was improved and further commercialized.
By the 1950s it was believed that the end was near for viscose and new direct dissolution
methods were discovered and explored. In 1969, Dee Lynn Johnson initiated the Ly-
ocell process by direct dissolution of cellulose in N-methylmorpholine N-oxide (NMMO)
solution.47
Detailed discussions on history, processes, properties, applications, and market of
regenerated cellulose fibers were done by Woodings.47 The advancement in biotechnology
Downloaded by [University of Boras] at 03:48 29 January 2015

led to the production of regenerated silk fibers.377,378 The wood based regenerated fibers
have been used to produce fiber reinforced composites while regenerated silk fibers were
used for biomedical applications such as scaffolds.379

3.1 Lyocell
Lyocell fiber is produced by direct dissolution of wood pulp in (NMMO).380 The pulp
is wetted out with amine oxide and subsequent removal of water produces homogenous
solution with minimum quantity of undissolved particles and air bubbles.47 It cannot be
absolutely called as a synthetic fiber as these fibers are produced from biomass origin.
The latest techniques allow the recycling of the solvent which could make the process
more environmentally friendly in the near future.381 The current manufacturing process
is designed to recover approximately 99% of the solvent which reduces the effluent. The
effluent produced in the Lyocell process is non-hazardous making it an environmentally-
friendly process. Initially, it was developed as Tencel by Courtaulds Fibers and these
fibers were sold to the public as a type of rayon as it is produced by a similar process.
Currently, Tencel is owned by Lenzing AG and it is the only producer of these fibers.381
The annual production of Lyocell fibers is around 50,000 tons and the total annual capacity
of the industrial units is upgraded to more than 160,000 tons. Lyocell exhibits some unique
characteristics such as softness, absorption, wrinkle resistance, strong when dry or wet, and
easy maintenence.381 It possesses good textile properties; it drapes well, can be hued with
different color dyes, texture can be altered to get suede, leather, and silk finishing.381 The
dyeing and finishing of Lyocell fibers is the key to their success.47 It is a distinctive fiber as it
exhibits beneficial characteristics of both synthetic and natural fibers; uniform mechanical,
morphological, and physical properties of synthetic fibers while biodegradability, non-
abrasiveness to equipment and low density of natural fibers.382 Lyocell fibers overcome
some of the issues of plant and animal fibers as it is highly pure, uniform, and the properties
are reproducible.383 This is one of the main reasons that the Lyocell could be used as
reinforcement in composites which demand even and reproducible properties.
The chemical composition of Lyocell change largely with the wood pulp used. Pulp
containing high α-cellulose (over 90%) or high hemicellulose (over 20%) could be used
to produce the fibers.384 The cost of pulp from high hemicellulose is cheaper when com-
pared to the high α-cellulose pulp.384 The crystallinity changes due to above reason and
several authors have reported different crystallinity.49,50 Generally, high α-cellulose pulp
has been used to produce Lyocell which in-turn gives high crystallinity (80%).48,50 The
138 S. K. Ramamoorthy et al.

Lyocell fibers exhibit high tensile properties; strength–1400 MPa, modulus–36 GPa, and
elongation–6%.385 Lyocell fibers have been used as reinforcement to produce biocom-
posites and their properties were studied.48,51,52,382,385–387 Mechanical, morphological, and
water absorption properties of the Lyocell composites were studied and compared to that
of flax composites.48 Lyocell hybrid composites were prepared to improve the properties
of the composites.51,52,382 The biodegradability of the composites reinforced with Lyocell
were studied by Shibata et al. and it was found that Lyocell/PLA composites have high
biodegradability.386 The main drawback of Lyocell fibers is that it is very expensive when
compared to natural fiber such as cotton or regenerated cellulose such as viscose. It is used
mainly in textile industries to make every day fabrics. The properties of these composites
are promising which makes Lyocell as potential reinforcement for composites.

3.2 Viscose
Downloaded by [University of Boras] at 03:48 29 January 2015

Viscose is another important regenerated cellulose fiber in terms of production volume


which is prepared by treatment of cellulose with sodium hydroxide and carbon disulfide.
There are two common types of cooking processes namely kraft and sulphite. Woodings
described each step in the viscose process in detail.47 These fibers are produced from
biomass origin similar to that of Lyocell. Viscose was produced and patented by Cross et al.
in 1891.388 It was also widely called as artificial silk. Global viscose production increased
from 14,000 in 1920 to 225,000 in 1931 ton/year mainly and continued to increase due
to the expiration of viscose patents and more than 100 companies started producing this
fiber.389 These fibers are used in industrial applications such as tyre cords and textiles.
The use of viscose increased even after the emergence of Lyocell and reached 3.5 million
ton in 2010. It was weaker than cotton and tends to shrink more easily. It lacked bulk,
resilience, and abrasion resistance when compared to wool fiber.389 Similar to Lyocell, it
exhibits beneficial characteristics of both synthetic and natural fibers. It has good properties
such as silk-like aesthetic, varied color hues, and easy drape. The viscose fiber could be
used as reinforcement due to its good properties if few production environmental issues
are addressed. Some environmental issues in the viscose process are addressed, namely the
usage and recovery of the alkali to the main circuit treatment of effluents and the recovery
of carbon disulphide.

(C6 H10 O5 )n + nNaOH → (C6 H9 O4 ONa)n + nH2 O


cellulose + alkali → sodium cellulosate + water
SNa
(C6 H9 O4 ONa)n + nCS2 →
n(SC − OC6 H9 O4
sodium cellulose + carbon disulphide → sodium cellulose xanthate

The pulp used to produce the fibers determines the chemical composition of the fibers.
Generally, the pulp used to make viscose has high α-cellulose by removal of hemicellulose
from the pulp.390,391 The crystallinity of the viscose fibers (41%) is lower than that of
Lyocell fibers (80%).50 This is one of the reasons that viscose fibers possess lower tensile
strength, lower modulus, and higher elongation than Lyocell fibers. It was also noted that the
thermal stability of viscose fiber is lower than Lyocell fiber.50 These fibers have been used
as reinforcement to produce composites.51,52,392,393 The composites produced from Lyocell
A Review of Natural Fibers Used in Biocomposites 139

and viscose were compared and it was found that superior properties of Lyocell fiber
made their composites perform better than viscose counterparts composites.51,52 But the
properties of viscose fiber reinforced composites were comparable to that of the plant and
animal fiber reinforced composites which certainly indicates the reinforcement potential of
viscose fibers.52
Other types of regenerated cellulose such as rayon and modal could be also as rein-
forcements to produce composites. The crystallinity analysis showed that modal fibers have
lower crystallinity than Lyocell fibers.49,50 The moisture absorption of modal fibers was
higher than that of Lyocell and viscose fibers, and thermal stability of modal fibers was
lower than that of the other regenerated cellulose fiber.49,50 Fiber dyeing, yarn and fabric
manufacturing of regenerated cellulose fibers were also explained.47

4. Fiber Separation and Extraction


Downloaded by [University of Boras] at 03:48 29 January 2015

4.1 Plant Fibers


The fibers should be separated and extracted from woody tissue of the fiber crops after the
plant harvesting. The plant fibers used in commercial applications, mostly bast fibers are
separated from the fiber crops by the retting process. Retting is a process which separates
the fiber bundles from central stem which loosens the fibers from woody tissue of the fiber
crops. The fiber extraction has an impact on fiber yield, fiber quality, chemical composition,
structure, and properties of the fiber.337
There are several retting processes as shown in Fig. 18a. It can be divided into four major
separation processes such as biological, mechanical, physical, and chemical. Biological
activity of microorganisms such as bacteria and fungi in the environment plays a major role
in the degradation of the pectic polysaccharides from non-fiber tissue and separated fiber
bundles. Sometimes, the retting process could be challenging with respect to the caution
involved as under-retting results in contaminated fibers.337

Dew Retting
Dew retting is the oldest and most commonly used retting process to separate fibers.
It is also called field retting. This process needs appropriate moisture and temperature
ranges, and therefore cannot be used globally. The plant remains in the field (spread in
uniform and thin non-overlapping swaths) after harvesting for the microorganisms to
separate fibers from cortex and xylem. The plants are turned over on a regular basis to
ensure homogeneous retting. The retting process should be monitored and stopped at
the right time to prevent degradation of cellulose by microorganisms (fungi); and this
is called as over-retting. The over-retting reduces the mechanical performance of the
fibers while under-retting makes further fiber processing difficult. Dew retting depends
on weather conditions and normally takes 3 to 6 weeks. There are several disadvantages
of this process such as poor quality of fibers compared to other processes, dirty dark
fibers due to contact with soil limited to geographical location, uncontrollable weather
conditions, and occupation of land until the retting is complete.337
Dew retting remains the choice of the farmers of various plant fiber crops, due to low
labor cost and being sustainable.
Stand Retting
A modified field retting was attempted to overcome the limitations of dew retting in
Northern Ireland. A pre-harvest desiccant, glyphosate (N-phosphonomethyl glycine)
was used to facilitate retting. It was found that fungal spread and retting was slower
140 S. K. Ramamoorthy et al.
Downloaded by [University of Boras] at 03:48 29 January 2015

Figure 18. (a) Classification of retting processes.337 (b) Schematic representation of different fiber
and fabric structure reinforced in matrix.

than dew retting but in comparison the fibers had better strength when the stand-retting
method was used.351
Another stand retting method is the thermally-induced stand retting. The plant growth
is terminated by open gas flames and the plant bases are heated approximately 100◦ C.
The plants dry in a couple of days which allows the replacement of weather-vulnerable
dew retting. The risk of weather and crop damage is reduced by using this method but
the cost of retting increases.337
Cold Water Retting
Traditional water retting involves steeping the fiber section of the plant in running
water such as rivers or streams. Later, this was replaced by water tanks which are
sealed or opened depending upon the weather and the water change takes place every
2 days. This process utilizes anaerobic bacteria that break down the pectin and it is
mainly followed in Eastern Europe. Cold water retting normally takes between 7 and
14 days but the duration largely depends on the temperature of the water, the water
type, and the bacterial function.
This retting process produces high quality fibers than field retting. There are some
drawbacks of this process such as high amount of water consumption, high cost of
artificial drying after retting, environmental pollution from organic wastewater, and
malodor of fermentation gases. This process has been terminated in most parts of
Europe due to environmental concerns.337
A Review of Natural Fibers Used in Biocomposites 141

Warm Water Retting


This is an accelerated water retting process which produces clean and high quality of
fibers in 3 to 4 days. The water tanks are heated between 28 and 40◦ C. The process
and limitations are similar to that of cold water retting, and has been abolished in most
parts of Europe.337 The waste water could be used as liquid fertilizer if it is treated to
remove toxic elements.
Mechanical Retting
Mechanical retting, also known as green retting, is a simple and cost-effective process
to separate the fibers. Mechanical retting is performed after field or technical drying.
Field drying normally takes about 2 to 3 days. The fibers are separated from woody
tissues mechanically. Fibers produced by mechanical retting are less fine than field or
water retting process.337
Ultrasound Retting
The crushed and washed fiber section of the plant is fed into a hot water bath to separate
the fiber. Hot water bath is kept at 70◦ C with small amounts of alkali and surfactants;
Downloaded by [University of Boras] at 03:48 29 January 2015

and exposed to high intensity ultrasound (40 kHz). The fibers are separated from the
hurds in this continuous process. The advantage of this process is that it does not need
field drying as the raw fiber section of the plant is fed directly.337
Steam Explosion
Steam explosion is a good alternative to the traditional retting process. The steam
and the additives penetrate the fiber interspaces of the fiber bundles under pressure and
high temperature; that removes the center lamella at optimum conditions. The resulting
relaxation leads to breaking up of the fiber section and results in decomposition into
fine fibers. This fibers produced by this method are fine and have good properties.337
Enzyme Retting
The pectin degrading enzymes are used to separate the fibers from the woody tissue.
This gives controlled retting of the fiber crops through selective biodegradation of the
pectinaceous substances. The enzyme activity could increase with increasing temper-
ature up to an optimum temperature above which the enzyme starts to denature. The
fibers produced by this process are of high and consistent quality.337
Chemical and Surfactant Retting
This refers to all retting processes in which the fiber section of the plant is submerged
in heated tanks with water solutions of sulphuric acid, chlorinated lime, sodium or
potassium hydroxide, and soda ash to dissolve the pectin component. The surface ac-
tive agents could be used in retting to remove the unwanted non-cellulosic components
adhering to the fibers (dispersion and emulsion forming process). Chemical and sur-
factant retting produces high quality fibers but the cost is higher than that of traditional
methods.337
There are other processes such as duralin and tuxying to separate the fibers. The fiber
extraction and cleaning takes place after the retting process where the non-fibrous
materials are removed entirely. The extraction could take place in one step or in several
steps such as breaking, milling, scotching, and decortication. The fibers are cleaned
after extraction to be used in several applications.

4.2 Animal Fibers


Silk fibers from silkworms and spiders are extracted in different ways. Cocoons from
mulberry silkworm are boiled gently in a mild soap solution which dissolves the sericin
gum binding the fibers and unravels the silk fibers. Then, the fibers can be reeled onto
142 S. K. Ramamoorthy et al.

wheels. The silk fiber is washed to remove remaining sericin and other contaminations. The
extraction of silk from spider is labor intensive, space and time consuming. The spider is
sedated at cool temperature which is then fixed upside down to expose spinnerets. A small
brush is used to provoke fiber production and a microscope is used to separate the fibers. It
has to be done for each fiber which is uneconomical for industrial scale production.351
Wool fibers are extracted by manually shearing and collecting greasy wool fibers which
have various impurities. So, skirting and classing are done manually. The greasy wool is
scoured to extract the clean wool.351
Various fibers are extracted in different ways which are then cleaned to remove the
impurities before the fibers are used.

5. Reinforcement Architecture
Fibers are reinforced in a matrix in different ways, either continuous or discontinuous
Downloaded by [University of Boras] at 03:48 29 January 2015

fibers. Fibers (short or long) could be directly reinforced in a matrix or fabrics made out
of fibers could be used as reinforcement. The fabric architecture is divided into three types
specifically nonwoven, woven, and knitted (Fig. 18b). The engineered architecture has
a major impact on the mechanical properties of fabric and the resulting composites. It
influences the method to be used for composite manufacturing.
Nonwoven fabrics are made from short or long staple fibers which are used as rein-
forcements. The nonwoven fabrics are made by entangling fibers mechanically, thermally,
or chemically. Several products such as filters, tissue paper, surgical drapes, and many other
consumer products are nonwoven fabrics. Carding and needling are well known mechani-
cal techniques to produce nonwoven fabric by entangling the fibers. Nonwoven fabric has
good tensile strength in fiber direction but poor strength in the direction perpendicular to
the fiber. Due to this reason, this structure is used in disposable tissue paper where it has to
disperse quickly on flushing. Nonwoven fabric is also used as reinforcement in composites
where the stress is applied in one direction (fiber direction).51,52
Woven fabrics are produced from yarns by weaving (yarn interlacing) on loom. Unlike
nonwovens, woven fabric has good strength in both the directions as fibers are oriented in
0 and 90◦ as warp and weft, respectively. There are several patterns in weaving such as
plain, twill, satin, basket, leno, and mock leno. Figure 19 shows some of the weave patterns.
Detailed discussions were made about the composites and their properties resulting from
different weave structures.14

Figure 19. Weave patterns: (a) Satin, (b) basket, and (c) leno.14 © Elsevier. Reproduced by permis-
sion of Elsevier. Permission to reuse must be obtained from the rightsholder.
A Review of Natural Fibers Used in Biocomposites 143

Knitted fabrics are produced from yarns by knitting (yarn interloping). It is the third
major class of fabric after woven and nonwoven fabrics. The longitudinal loops are called
wales and the horizontal loops are called courses. There are two types of knitting such
as warp and weft knitting. The knitted fabrics are mostly used in textile industry. Knitted
fabric has higher elastic property but lower dimensional stability than woven fabric. These
fabric properties have a significant effect on the resulting composites.

6. Summary
Plant fiber properties are affected by maturity, geographical cultivation, harvesting type,
weather, etc. It is not possible to generalize or reproduce the properties of natural fiber while
the same can be achieved with regenerated cellulose fibers. The fiber properties reflect in
the composites. In general, every fiber mentioned above has reinforcement potential in
composites for diverse applications. Recycled fibers could also be used as reinforcement;
Downloaded by [University of Boras] at 03:48 29 January 2015

recycled newspaper fiber was used as reinforcement to produce composites and their
properties were investigated.394,395 Waste textile fibers were reused as reinforcements in
composites and their properties were studied.240,396 Every fiber could be reinforced in
several thermoset and thermoplastic plastics, and their properties could be evaluated for
mechanical, thermal, electrical, chemical, viscoelastic, water absorption, morphological
properties, etc. The choice of matrix could be limited due to the natural fiber degradation
at elevated temperatures. A high number of publications and the broad scope of this article
were made to review few characterization techniques and go through few properties for
each fiber and their composites; however, all characterization could be done for every fiber
and their reinforced composites.
The authors have made a comprehensive review of a particular fiber and their com-
posites which gives in-depth knowledge about them.54–58,156,178,256,264,282,299 Other review
papers have considered a brief review about cellulose based composites that give a good
comparative analysis of few natural fibers and their composites.11–16 Mechanical testing
such as tensile, flexural, and impact tests are the most commonly studied aspects of biocom-
posites. Morphological analysis such as microscopy is also often used to see the fiber-matrix
interface. Mechanical and morphological analyses help to see the stress properties and adhe-
sion which is required in most of the applications. Thermal, electrical, and other properties
are vital for certain applications and these were studied on certain composites.

7. Discussion
Plant fibers are hydrophilic due to their high cellulose content that makes the fiber to
absorb high amounts of water. The poor interface between fiber and matrix is obtained
when hydrophilic fibers reinforce hydrophobic matrixes. The mechanical properties of the
composites can be improved by improving the interface which makes the composites firm
and allows good stress transfer. The chemical analysis of the fiber was done in order to study
and modify the surface of the fiber which improves the interface between fiber and matrix.
A brief review was done on surface modification of natural fiber by Mohanty et al.397 The
performance of the resulting biocomposites were also studied.
Composites from natural fibers can be fabricated by several methods such as com-
pression molding, vacuum bag molding, resin transfer molding, etc. Proper selection and
control of processing parameters such as temperature, pressure, time, etc., are a mandatory
requirement to manufacture good composites. The parameters would control fiber degra-
dation, resin flow, pore creation, cross-linking, and fiber wetting, etc., which affect the
144 S. K. Ramamoorthy et al.

composites. The composite fabrication that includes different types of molding was dis-
cussed in detail.398 3D fiber architecture of reinforced polymer composites was discussed
by Tong et al.399
Fiber and composite mechanics give good knowledge about the properties. It is possi-
ble to estimate the load bearing capacity and other properties of fiber reinforced composites
by studying micro and macro mechanics. Mechanics of composite materials were discussed
in detail.401,402 Kinetics with respect to the properties such as tensile strength and water
absorption have been done.403 Composite designing and modeling of fiber reinforced com-
posites were done to predict the properties of composites. Computer software was used to
predict the properties of the fiber-reinforced composites.45–46
This review article is believed to have covered the most important natural fibers used in
composites. In our upcoming review papers, the concentration will be on the fiber surface
modifications, various thermoset and thermoplastic matrix, composite processing, hybrid
composites, mechanics, modeling, applications, and life-cycle assessment.
Downloaded by [University of Boras] at 03:48 29 January 2015

8. Conclusions
The use of natural fibers is increasing in composites due to their environmental and economic
benefits. The market for natural fibers has seen a marked increase over the last decade and
experts foresee a continuation of this trend in the future. It also has unique properties such
as low density and good strength. Natural fiber composites have found the commercial
applications in automotive, aircraft, marine, and construction industries. Several composites
have been produced from various plant and animal fibers while only few of these fibers have
actually been used in industrial scale. Natural fiber reinforced composites have significantly
developed over the years through continuous research. High performance natural fiber
composite materials are made from decades of research. Presently, substantial research is
taking place globally on natural fibers and their composites to improve the properties. The
fibers and composites are characterized for various properties with respect to applications
with several instruments.
Further research on fibers and their reinforcement potential has to be done. The moisture
absorption of the fiber and fiber-matrix has to be addressed to improve the properties.
Currently, the natural fiber reinforced composites are mostly used in indoor applications
and its potential in outdoor applications can be enhanced if the above issues related to
natural fibers are addressed.

References
1. Chawla K. K. “Carbon fiber composites”, In: Composite Materials, 2nd Edition; Springer, New
York, 1998, pp. 252–277.
2. Strands G. F. Glass fiber reinforced composite article exhibiting longitudinal tensile and com-
pressive moduli. US Patent 3691000, 1972.
3. Brody H; Ward I. M. “Modulus of short carbon and glass fiber reinforced composites”, Polymer
Science & Engineering 1971, 11(2), 139–151.
4. Summerscales, J. “Short D. Carbon fibre and glass fibre hybrid reinforced plastics”, Composites
1978, 9(3), 157–166.
5. Harris, B.; Bunsell, A. R. “Impact properties of glass fibre/carbon fibre hybrid composites”,
Composites 1975, 6(5), 197–201.
6. Bunsell, A. R.; Harris, B. “Hybrid carbon and glass fibre composites”, Composites 1974, 5(4),
157–164.
A Review of Natural Fibers Used in Biocomposites 145

7. Manders, P. W.; Bader, M. G. “The strength of hybrid glass/carbon fibre composites”, Journal
of Material Science 1981,16(8), 2233–2245.
8. Henshaw, J. M.; Han, W.; Owens, A. D. “An overview of recycling issues for composite
materials”, Journal of Thermoplastic Composite Materials 1996, 9(1), 4–20.
9. Pickering, S. J. “Recycling technologies for thermoset composite materials–current status”,
Composites Part A: Applied Science and Manufacturing 2006, 37(8), 1206–1215.
10. Conroy, A.; Halliwell, S.; Reynolds, T. “Composite recycling in the construction industry”,
Composites Part A: Applied Science and Manufacturing 2006, 37(8), 1216–1222.
11. Bledzki, A. K.; Gassan, J. “Composites reinforced with cellulose based fibres”, Progress in
Polymer Science 1999, 24(2), 221–274.
12. Faruk, O.; Bledzki, A. K.; Fink, H. P.; Sain, M. “Biocomposites reinforced with natural fibers:
2000–2010”, Progress in Polymer Science 2012, 37(11), 1552–1596.
13. La Mantia, F. P.; Morreale, M. “Green composites: A brief review”, Composites Part A 2011,
42(6), 579–588.
14. John, M. J.; Thomas, S. “Biofibres and biocomposites”, Carbohydrate Polymers 2008, 71(3),
Downloaded by [University of Boras] at 03:48 29 January 2015

343–364.
15. Bogoeva-Gaceva, G.; Avella, M.; Malinconico, M.; Buzarovska, A.; Grozdanov, A.; Gentile,
G. E. “Natural fiber eco-composites”, Polymer Composites 2007, 28(1), 98–107.
16. Mohanty A. K.; Misra, M.; Hinrichsen, G. “Biofibres, biodegradable polymers and biocompos-
ites: An overview”, Macromolecular Materials and Engineering 2000, 276–277(1), 1–24.
17. Koronis G.; Silva, A.; Fontul, M. “Green composites: A review of adequate materials for
automotive applications”, Composites: Part B 2013, 44(1), 120–127.
18. Ku, H.; Wang, H.; Pattarachaiyakoop, N.; Trada, M. “A review on tensile properties of natural
fiber reinforced polymer composites”, Composites: Part B 2011, 42(4), 856–873.
19. George J.; Sreekala, M. S.; Thomas, S. “A review on interface modification and characterization
of natural fiber reinforced plastic composites”, Polymer Engineering & Science 2001, 41(9),
1471–1485.
20. Sgriccia, N.; Hawley, M.; Misra, C. “Characterization of natural fiber surfaces and natural fiber
composites”, Composites: Part A 2008, 39(10), 1632–1637.
21. Cheung, H. Y.; Ho, M. P. H.; Lau, K. T.; Cardona, F.; Hui, D. “Natural fibre-reinforced com-
posites for bioengineering and environmental engineering applications”, Composites Part B:
Engineering 2009, 40(7), 655–663.
22. Kabir, M. M.; Wang, H.; Lau, K.; Cardona, T. “Chemical treatments on plant-based natural fibre
reinforced polymer composites: An overview”, Composites Part B: Engineering 2012, 43(7),
2883–2892.
23. Xie, Y.; Hill, C. A. S.; Xiao, Z.; Militz, H.; Mai, C. “Silane coupling agents used for natural
fiber/polymer composites: A review”, Composites Part A: Applied Science and Manufacturing
2010, 41(7), 806–819.
24. Dittenber, D. B.; Gangarao, H. V. S. “Critical review of recent publications on use of natural
composites in infrastructure”, Composites Part A: Applied Science and Manufacturing 2012,
43(8), 1419–1429.
25. Malkapuram, R.; Kumar, V.; Negi, Y. S. “Recent developments in natural fiber reinforced
polypropylene composites”, Journal of Reinforced Plastics and Composites 2009, 28(10),
1169–1189.
26. Shubhra, Q. T.; Alam, A. K. M. M.; Quaiyyum, M. A. “Mechanical properties of polypropy-
lene composites: A review”, Journal of Thermoplastic Composite Material 2013, 26(3),
362–391.
27. Yuan, Q.; Wu, D.; Gotama, J.; Bateman, S. “Wood fiber reinforced polyethylene and polypropy-
lene composites with high modulus and impact strength”, Journal of Thermoplastic Composite
Material 2008, 21(3), 195–208.
28. Torres, F. G.; Cubillas, M. L. “Study of the interfacial properties of natural fibre reinforced
polyethylene”, Polymer Testing 2005, 24(6), 694–698.
146 S. K. Ramamoorthy et al.

29. Sobczak, L.; Brugemann, O.; Putz, R. F. “Polyolefin composites with natural fibers and wood-
modification of the fiber/filler-matrix interaction”, Journal of Applied Polymer Science 2013,
27(1), 1–17.
30. Plackett, D.; Andersen, T. L.; Pedersen, W. B.; Nielsen, L. “Biodegradable composites based
on L-polylactide and jute fibres”, Composites Science and Technology 2003, 63(9), 1287–1296.
31. Siracusa, V.; Rocculi, P.; Romani, S.; Rosa, M. D. “Biodegradable polymers for food packaging:
A review”, Trends in Food Science & Technology 2008, 19(12), 634–643.
32. Bax, B.; Mussig, J. “Impact and tensile properties of PLA/cordenka and PLA/flax composites”,
Composites Science and Technology 2008, 68(7–8), 1601–1607.
33. Meier, M. A. R.; Metzger, J. O.; Schubert, U. S. “Plant oil renewable resources as green
alternatives in polymer science”, Chemical Society Reviews 2007, 36(11), 1788–1802.
34. Galia, M.; de Espinosa, L. M.; Ronda, J. C.; Lligadas, G.; Cadiz, V. “Vegetable oil-based
thermosetting polymers”, European Journal of Lipid Science and Technology 2010, 112(1),
87–96.
35. de Espinosa, L. M.; Meier, M. A. R. “Plant oils: The perfect renewable resource for polymer
Downloaded by [University of Boras] at 03:48 29 January 2015

science”, European Polymer Journal 2011, 47(5), 837–852.


36. Sharma, V.; Kundu, P. P. “Addition polymers from natural oils: A review”, Progress in Polymer
Science 2006, 31(11), 983–1008.
37. Guner, F. S.; Yagci, Y.; Erciyes, A. T. Polymers from triglyceride oils. Progress in Polymer
Science 2006, 31(7), 633–670.
38. Bledzki, A. K.; Izbicka, J.; Gassan, J. Kunststoffe-Umwelt-Recycling; Stettin: Poland; 1995, pp
27–29.
39. Pal, P. K. “Jute reinforced plastics: A low cost composite material”, Plastics and Rubber
Processing and Applications 1984, 4(3), 215–219.
40. Winfield, A. G. “Jute reinforced polyester projects for UNIDO/government of India”, Plastics
and Rubber International 1979, 4(1), 23–28.
41. Business Communication Company. “Composites: Resins, Fillers, Reinforcements, Natural
Fibers & Nanocomposites” I. Plastics, 2007. Available at: http://www.bccresearch.com/market-
research/plastics/composites-resins-fillers-pls029b.html
42. Report NFCM. Natural Fiber Composite Market Report. Little Falls, Kline, & Company: NJ,
2004.
43. Shen, L.; Haufe, J.; Patel, M. K. Product overview and market projection of emerging bio-based
plastics, PRO-BIP; Final Report, Report No: NWS-E-2009-32. Utrecht University: Utrecht,
The Netherlands, 2009, p 243.
44. Mirous, A. “Fiber-reinforced plastics use, 2002”, In: Plastics News; Plastics News
Global Group: Akron, OH, 2002. Available at: http://www.plasticsnews.com/article/
20020826/FYI/308269999/fiberreinforced-plastics-use-2002
45. Venkateshwaran, N.; Elayaperumal, A.; Sathiya, G. K. “Prediction of tensile properties of
hybrid-natural fiber composites”, Composites: Part B 2012, 43(2), 793–796.
46. Behzad, T.; Sain, M. “Measurement and prediction of thermal conductivity for hemp fiber
reinforced composites”, Polymer Engineering and Science, 47(7), 977–983.
47. Woodings, C. Regenerated Cellulose Fibres; Woodhead Publishing: UK, 2001.
48. Carrillo, F.; Colom, X.; Canavate, X. “Properties of regenerated cellulose lyocell fiber-reinforced
composites”, Journal of Reinforced Plastics and Composites, 2010, 29(3), 359–371.
49. Colom, X.; Carrillo, F. “Crystallinity changes in lyocell and viscose-type fibres by caustic
treatment”, European Polymer Journal 2002, 38(11), 2225–2230.
50. Carrillo, F.; Colom, X.; Sunol, J. J.; Saurina, J. “Structural FTIR analysis and thermal character-
ization of lyocell and viscose-type fibres”, European Polymer Journal 2004, 40(9), 2229–2234.
51. Ramamoorthy, S. K.; Di, Q.; Adekunle, K.; Skrifvars, M. “Effect of water absorption on
mechanical properties of soybean oil thermosets reinforced with natural fibers”, Journal of
Reinforced Plastics and Composites 2012, 31(18), 1191–1200.
52. Ramamoorthy, S. K.; Kundu, C. K.; Adekunle, K.; Bashir, T.; Skrifvars, M. “Properties of
green composites with regenerated cellulose fiber and soybean-based thermoset for technical
applications”, Journal of Reinforced Plastics and Composites 2013, 33(2), 193–201.
A Review of Natural Fibers Used in Biocomposites 147

53. Staiger, M. P.; Tucker, N. “Natural-fibre composites in structural applications”, In: Pickering
K, ed.; Properties and Performance of Natural-Fibre Composites; Woodhead Publishing: UK,
2008, pp. 269–300.
54. Yan, L., Chouw, N., Jayaraman, K. “Flax fibre and its composites: A review”, Composites Part
B: Engineering 2014, 56, 296–317.
55. Shahzad A. “Hemp fiber and its composites: A review”, Journal of Composite Materials 2012,
46(8), 973–986.
56. Mohanty, A. K.; Misra, M. “Studies on jute composites: A literature review”, Polymer-Plastics
Technology and Engineering 1995, 34(5), 729–792.
57. Akil, H. M.; Omar, M. F.; Mazuki, A. A. M.; Safiee, S.; Ishak, Zam.; Abu, Bakar; Ke-
naf, A. “Fiber reinforced composites: A review”, Materials & Design 2011, 32(8–9), 4107–
4121.
58. Li, Y.; Mai, Y.W.; Ye, L. “Sisal fibre and its composites: A review of recent developments”,
Composites Science and Technology 2000, 60(11), 2037–2055.
59. Lobovikov, M.; Paudel, S.; Piazza, M.; Ren, H.; Wu, J. “World bamboo resources: A thematic
Downloaded by [University of Boras] at 03:48 29 January 2015

study prepared in the framework of the global forest resources assessment 2005”,FAO, Rome,
Italy, 2007, pp 1–37.
60. Bledzki, A. K.; Reihmane, S.; Gassan, J. “Properties and modification methods for veg-
etable fibers for natural fiber composites”, Journal of Applied Polymer Science 1996, 59(8),
1329–1336.
61. Hon, D. N. S. “Chemical modification of lignocellulosic materials: Old chemistry, new ap-
proaches”, Polymer News 1992, 17, 102–107.
62. Ugbolue, S. C. O. “Structure/property relationships in textile fibres”, Textile Progress 1990,
20(4), 1–43.
63. Kucharyov, M. S. “Properties and modification methods for vegetable fibers for natural fiber
composites”, Tekst. Prom 1993, 23, 8–9.
64. Wuppertal, E. W. Die Textile Rohstoffe (Natur und Chemiefasern); Dr. Spohr-Verlag/Deutscher
Fachverlag, Frankfurt, Germany, 1981.
65. Rowell R. M.; Schiltz, T.; Narayan, R.; eds: Emerging Technologies for Materials and Chemicals
from Biomass; Symposium Series 476; American Chemical Society: Washington D.C., USA.
1992, pp. 251–286.
66. Varma, D. S.; Varma, M.; Varma, I. K. “Coir fibers. Part I. Effect of physical and chemical
treatments on properties”, Textile Research Journal 1984, 54(12), 827–832.
67. Mukherjee, P. S.; Satyanarayana, K. G. “Structure and properties of some vegetable fibers - Part
2: Pineapple fiber”, Journal of Material Science 1986, 21(1), 51–56.
68. Bledzki, A. K.; Gassan, J. “Natural fiber reinforced plastics”, In: Handbook of engineering
polymeric materials; Marcel Dekker: New York, USA, 1997, pp. 787–810.
69. Varma, D. S.; Varma, M.; Varma, I. K. “Thermal behaviour of coir fibres”, Thermochimica Acta
1986, 108, 199–210.
70. Hattallia, S.; Benaboura, A.; Ham-Pichavant, F.; Nourmamode, A.; Castellan, A. “Adding value
to alfa grass (Stipa tenacissima L.) soda lignin as phenolic resins. 1. Lignin characterization”,
Polymer Degradation and Stability 2002, 76(2), 259–264.
71. Hoareau, W.; Trindade, W. G.; Siegmund, B.; Castellan, A.; Frollini, E. “Sugar cane bagasse and
curaua lignins oxidized by chlorine dioxide and reacted with furfuryl alcohol: Characterization
and stability”, Polymer Degradation and Stability 2004, 86(3), 567–576.
72. Marti-Ferrer, F.; Vilaplana, F.; Ribes-Greus, A.; Benedito-Borras, A.; Sanzbox, C. “Flour rice
husk as filler in block copolymer polypropylene: Effect of different coupling agents”, Journal
of Applied Polymer Science 2006, 99(4), 1823–1831.
73. Lewin, L. M. Pearce, E. M. Handbook of Fiber Science and Technology; Fiber Chemistry Vol.
4; Marcel Dekker: NY, 1985.
74. Mohanty A. K. “Graft copolymerization of vinyl monomers onto jute fibers”, Journal of Macro-
molecular Science, Part C: Polymer Reviews 1987, 27(3–4), 593–639.
75. Bisanda, E. T. N.; Ansell, M. P. “Properties of sisal-CNSL composites”, Journal of Materials
Science 1992, 27(6), 1690–1700.
148 S. K. Ramamoorthy et al.

76. Chand, N., Rohatgi, P. K. “Plant fibres”, In: Natural Fibres and Composites; Periodical Experts
Agency: Delhi, India, 1994. pp. 25–79.
77. Zeronian, S. H. The mechanical properties of cotton fibers. Journal of Applied Polymer Science:
Applied Polymer Symposium 1991, 47, 445–461.
78. Saechtling, H. International Plastics Handbook; Hanser: Munich, Germany, 1987.
79. Rowell, R. M. “Natural fibres: types and properties”, In: Properties and Performance of Natural-
Fibre Composites; Pickering, K, ed.; Woodhead Publishing: UK, 2008, pp. 3–66.
80. Gassan, J.; Chate, A.; Bledzki, A. K. “Calculation of elastic properties of natural fibers”, Journal
of Materials Science 2001, 36(15), 3715–3720.
81. Baumgartl, H.; Schlarb, A. Nachwachsende rohstoffe-perspektiven fur die chemie. Symposium,
Frankfurt, Germany, May 5–6, 1993.
82. Dewilde, B. 20 Eeuwen Vlas in Vlaanderen; Lannoo, Teilt, Bussum, 1983.
83. Kvavadze, E.; Bar-Yosef, O.; Belfer-Cohen, A.; Boaretto, E.; Jakeli, N.; Matskevich,
Meshveliani. “30,000-year-old wild flax fibers”, Science 2009, 325(5946), 1359.
84. Charlet, K.; Baley, C.; Morvan, C.; Jernot, J. P.; Gomina, M.; Breard, J. “Characteristics of
Downloaded by [University of Boras] at 03:48 29 January 2015

hermes flax fibres as a function of their location in the stem and the properties of the derived
unidirectional composites”, Composites Part: A 2007, 38(8), 1912–1921.
85. Baley, C. “Analysis of the flax fibres tensile behavior and analysis of the tensile stiffness
increase”, Composites Part: A 2002, 33(7), 939–948.
86. Lilholt, H.; Toftegaard, H.; Thomsen, A. B.; Schmidt, A. S. “Natural composites based on cellu-
losic fibres and polypropylene matrix. Their processing and characterization”, In: Proceedings
of ICCM 12, Paris, France, 1999.
87. Cristaldi, G.; Latteri, A.; Recca, G.; Cicala, G. “Composites Based on Natural Fibre Fabrics”,
Woven Fabric Engineering 2010, 317–342.
88. Khalil, A.; Rozman, H. D.; Ahamd, N. N.; Ismail, H. “Acetylated plant-fibre-reinforced polyester
composites: A study of mechanical, hydrothermal and ageing characteristics”, Polymer Plastics
Technology Engineering 2000, 39(4), 757–781.
89. Charlet, K.; Eve, S.; Jernot, J. P.; Gomina, M.; Breard, J. “Tensile deformation of a flax fiber”,
Procedia Engineering 2009, 1(1), 233–236.
90. Pillin, I.; Kervoelen, A.; Bourmaud, A.; Goimard, J.; Montrelay, N.; Baley, C. “Could oleaginous
flax fibres be used as reinforcement for polymers?” Industrial Crops and Products 2011, 34(3),
1556–1563.
91. Alix, S.; Philippe, E.; Bessadok, A.; Lebrun, L.; Morvan, C.; Marais, S. “Effect of chemical
treatment on water sorption and mechanical properties of flax fibres”, Bioresource Technology
2009, 100(20), 4742–4749.
92. Charlet, K.; Jernot, J. P.; Gomina, M.; Breard, J.; Morvan, C.; Baley, C. “Influence of an
Agatha flax fibre location in a stem on its mechanical, chemical and morphological properties”,
Composites Science and Technology 2009, 69(9), 1399–1403.
93. Charlet, K.; Jernot, J. P.; Breard, J.; Gomina, M. “Scattering of morphological and mechanical
properties of flax fibres”, Industrial Crops and Products 2010, 32(3), 220–224.
94. Le Duigou, A.; Bourmaud, A.; Balnois, E.; Davies, P.; Baley, C. “Improving the interfacial
properties between flax fibres and PLLA by a water fibre treatment and drying cycle”, Industrial
Crops and Products 2012, 39, 31–39.
95. Andersons, J.; Porike, E.; Spamins, E. “The effect of mechanical defects on the strength
distribution of flax fibres”, Composites Science and Technology 2009, 69(13), 2152–2157.
96. Baley, C. “Analysis of the flax fibres tensile behavior and analysis of the tensile stiffness
increase”, Composites Part A: Applied Science and Manufacturing 2002, 33(7), 939–948.
97. Barber, E. J. W. Prehistoric Textiles: The Development of Cloth in the Neolithic and Bronze
Ages with Special Reference to the Aegean; PUP: Princeton, NJ, 1991.
98. Catling, D.; Grayson, J. Identification of Vegetable Fibres; Chapman and Hall: London, 1982.
99. Olesen P. O.; Plackett D. V. “Perspectives on the performance of natural plant fibers”, In: Natural
Fibres Performance Forum – Plant Fibre Products – Essentials for the Future, Copenhagen,
Denmark, May 27–28, 1999.
A Review of Natural Fibers Used in Biocomposites 149

100. Kandachar, P.; Brouwer, R. “Applications of bio-composites in industrial products”, Material


Research Society Proceedings. 2001, 702, 101–112.
101. Nishino, T. “Natural fiber sources”, In: Green Composites: Polymer Composites and the Envi-
ronment; Baille, C., ed.; Woodhead Publishing: UK, 2004, pp 49–80.
102. Amaducci, S.; Errani, M.; Venturi, G. Plant population effects on fibre hemp morphology and
production. Journal of Industrial Hemp 2002, 7(2), 33–60.
103. Svennerstedt, B.; Sevenson, G. “Hemp (Cannabis sativa L.) trails in southern Sweden
1999–2001”, Journal of Industrial Hemp 2006, 11(1), 17–25.
104. Pickering, K. L.; Beckermann, G.W.; Alam, S. N.; Foreman, N. J. “Optimising industrial hemp
fiber for composites”, Composites Part A: Applied Science 2007, 38(2), 461–468.
105. Pan, N. C.; Day, A.; Mahalanabis, K. K. Properties of jute. Indian Textile Journal 2000, 110(5),
16–23.
106. Guo-Qing, L. “Archeological evidence for the use of ‘chu-nam’ on the 13th century Quanzhou
ship, Fujian Province, China”, International Journal of Nautical Archaeology 1989, 18(4),
277–283.
Downloaded by [University of Boras] at 03:48 29 January 2015

107. Tsien, T. H. “Raw materials for old papermaking in China”, Journal of the American Oriental
Society 1973, 93(4), 510–519.
108. Kundu, B. C.; Basak, K. C.; Sarcar, P. B. Jute in India; Indian Central Jute Committee: India,
1959.
109. Chakrabarty, D. Rethinking Working-Class History: Bengal, 1890–1940; Princeton University
Press: Princeton, NJ, 2000.
110. Lewington, A. Plants for People; Transworld Publisher: London, 2003.
111. Mishra, H. K.; Dash, B. N.; Tripathy, S. S.; Padhi, B. N. “A study on mechanical performance of
jute-epoxy composites”, Polymer-Plastics Technology and Engineering, 2000, 39(1), 187–198.
112. Khondker, O. A.; Ishiaku, U. S.; Nakai, A.; Hamada, H. “A novel processing technique for ther-
moplastic manufacturing of unidirectional composites reinforced with jute yarns”, Composites
Part A: Applied Science and Manufacturing 2006, 37(12), 2274–2284.
113. Mohanty, A. K.; Khan, M. A.; Hinrichsen, G. “Influence of chemical surface modification on
the properties of biodegradable jute fabrics-polyester amide composites”, Composites Part A:
Applied Science and Manufacturing 2000, 31(2), 143–150.
114. Singh, B.; Gupta, M.; Verma, A. “The durability of jute fibre-reinforced phenolic composites”,
Composites Science and Technology 2000, 60(4), 581–589.
115. Razera, I. A. T.; Frollini, E. “Composites based on jute fibers and phenolic matrices: Properties
of fibers and composites”, Journal of Applied Polymer Science 2004, 91(2), 1077–1085.
116. Aziz, S. H.; Ansell, M. P.; Clarke, S. J.; Panteny, S. R. “Modified polyester resins for natural
fiber composites”, Composites Science and Technology 2005, 65(3), 525–535.
117. Lee, S. A.; Eiteman, M. A. “Ground kenaf core as a filteration aid”, Industrial Crops and
Production 2001, 13(2), 155–161.
118. Karnani, R.; Krishnan, M.; Narayan, R. “Biofiber–reinforced polypropylene composites”, Poly-
mer Engineering & Science 1997, 37(2), 476–483.
119. Nishino, T.; Hirao, K.; Kotera, M.; Nakamae, K.; Inagaki, H. “Kenaf reinforced biodegradable
composites”, Composites Science and Technology 2003, 63(9), 1281–1286.
120. Shibata, S.; Cao, Y.; Fukumoto, I. “Lightweight laminate composites made from kenaf and
polypropylene fibres”, Polymer Testing 2006, 25(2), 142–148.
121. Liu, W.; Drzal, L. T.; Mohanty, A. K.; Misra, M. “Influence of processing methods and fiber
length on physical properties of kenaf fiber reinforced soy based biocomposites”, Composites
Part B: Engineering 2007, 38(3), 352–359.
122. Cho, D.; Lee, H. S.; Han, S. O. “Effect of fiber surface modification on the interfacial and
mechanical properties of kenaf fiber–reinforced thermoplastic and thermosetting polymer com-
posites”, Composite Interfaces 2009, 16(7–9), 711–729.
123. Kasolang, S.; Ahmad, M. A.; Ghazali, F. A.; Azmi, A. M. “Preliminary study of dry sliding
wear in kenaf epoxy and carbon epoxy composites”, Applied Mechanics and Materials 2011,
52, 464–469.
150 S. K. Ramamoorthy et al.

124. Ochi, S. “Mechanical properties of kenaf fibers and kenaf/PLA composites”, Mechanics of
Materials 2008, 40(4–5), 446–452.
125. Chin, C. W.; Yousif, B. F. “Potential of kenaf fibres as reinforcement for tribological applica-
tions”, Wear 2009, 267(9–10), 1550–1557.
126. Serizawa, S.; Inoue, K.; Iji, M. “Kenaf-fiber-reinforced poly(lactic acid) used for electronic
products”, Journal of Applied Polymer Science 2006, 100(1), 618–624.
127. Ramaswamy, G. N.; Sellers, T.; Tao, W.; Crook, L. G. “Kenaf nonwovens as substrates for
laminations”, Industrial Crops and Products 2003, 17(1), 1–8.
128. Mohamed, A.; Bhardwaj, H.; Hamama, A.; Webber, C. “Chemical composition of kenaf (Hi-
biscus cannabinus L.) seed oil”, Industrial Crops and Products 1995, 4(3), 157–165.
129. Morrison Iii, W. H.; Akin, D. E.; Archibald, D. D.; Dodd, R. B.; Raymer, P. L. “Chemical and
instrumental characterization of maturing kenaf core and bast”, Industrial Crops and Products
1999, 10(1), 21–34.
130. Ogbonnaya, C. I.; Roy-Macauley, H.; Nwalozie, M. C.; Annerose, D. J. M. “Physical and
histochemical properties of kenaf (Hibiscus cannabinus L.) grown under water deficit on a
Downloaded by [University of Boras] at 03:48 29 January 2015

sandy soil”, Industrial Crops and Products 1997, 7(1), 9–18.


131. Perry, D. R.; Appleyard, H. M.; Cartridge, G.; Cobb, P. G. W.; Coop, G. E.; Lomas, B.; Ritchie,
G. G.; Taylor, C.; Welch, M. J.; Farnfield, C. A. Identification of Textile Materials; Textile
Institute: Manchester, UK, 1985.
132. Batra, S. K. Other Long Vegetable Fibers: Abaca, Banana, Sisal, Henequen, Flax, Ramie, Hemp,
Sunn and Coir; CRC Press: Boca Raton, FL, 2007, pp. 453–520.
133. Kirby, R. H. Vegetable Fibres, Botany, Cultivation and Utilization; Hill Leonard: London, 1964.
134. Hearle, J. W. S.; Peters, R. H. Fibre Structure; Textile Institute and Butterworths: London, 1968.
135. Nam, S.; Netravali, A. N. “Green composites. I. Physical properties of ramie fibers for
environment-friendly green composites”, Fibers and Polymers 2006, 7(4), 372–379.
136. Nam, S.; Netravali, A. N. “Green composites. II. Environment-friendly, biodegradable com-
posites using ramie fibers and soy protein concentrate (SPC) resin”, Fibers and Polymers 2006,
7(4), 380–388.
137. Yu, T.; Ren, J.; Li, S.; Yuan, H.; Li, Y. “Effect of fiber surface-treatments on the properties of
poly(lactic acis)/ramie composites”, Composites Part A: Applied Science and Manufacturing
2010, 41(4), 499–505.
138. Kim, J. T.; Netravali, A. N. “Mechanical, thermal and interfacial properties of green composites
with ramie fiber and soy resins”, Journal of Agricultural and Food Chemistry 2010, 58(9),
5400–5407.
139. Summerscales, J.; Dissanayake, N. P. J.; Virk, A. S.; Hall, W. “A review of bast fibres and
their composites. Part 1: Fibres as reinforcements”, Composites Part A: Applied Science and
Manufacturing 2010, 41(10), 1329–1335.
140. Summerscales, J.; Dissanayake, N. P. J.; Virk, A. S.; Hall, W. “A review of bast fibres and
their composites”, Part 2: Composites. Composites Part A: Applied Science and Manufacturing
2010, 41(10), 1336–1344.
141. Dempsey, J. M. Fiber Crops; University Presses of Florida: Gainesville, FL, 1975.
142. Goltenboth, F.; Muhlbauer, W. Abaca–Cultivation, Extraction and Processing. Industrial Ap-
plications of Natural Fibers: Structure, Properties and Technical Applications; Wiley: UK, pp.
163–179.
143. Shibata, M.; Ozawa, K.; Teramoto, N.; Yosomiya, R.; Takeishi, H. “Biocomposites made from
short abaca fiber and biodegradable polyesters”, Macromolecular Materials and Engineering
2003, 288(1), 35–43.
144. Teramoto, N.; Urata, K.; Ozawa, K.; Shibata, M. “Biodegradation of aliphatic polyester com-
posites reinforced by abaca fiber”, Polymer Degradation and Stability 2004, 86(3), 401–409.
145. Shibata, M.; Takachiyo, K. I.; Ozawa, K.; Yosomiya, R.; Takeishi, H. “Biodegradable polyester
composites reinforced with short abaca fiber”, Journal of Applied Polymer Science 2002, 85(1),
129–138.
A Review of Natural Fibers Used in Biocomposites 151

146. Rahman, M. R.; Huque, M. M.; Islam, M. N.; Hasan, M. “Mechanical properties of polypropy-
lene composites reinforced with chemically treated abaca”, Composites Part A: Applied Science
and Manufacturing 2009, 40(4), 511–517.
147. Vilaseca, F.; Valadez-Gonzalez, A.; Herrera-Franco, P. J.; Pelach, M. A.; Lopez, J. P.; Mutje,
P. “Biocomposites from abaca strands and polypropylene. Part I: Evaluation of the tensile
properties”, Bioresource Technology 2010,101(1), 387–395.
148. Bledzki, A. K.; Mamun, A. A.; Faruk, O. “Abaca fibre reinforced PP composites and comparison
with jute and flax fibre PP composites”, Express Polymer Letters 2007, 1(11), 755–762.
149. Bledzki, A. K..; Mamun, A. A..; Jaszkiewicz, A.; Erdmann, K. “Polypropylene composites with
enzyme modified abaca fibre”, Composites Science and Technology 2010, 70(5), 854–860.
150. Takagi, H.; Itotani, K.; Fukubayashi, Y. Strength evaluation of unidirectional abaca fibre rein-
forced biocomposites. In: Proceedings of the 6th Australasian Congress on Applied Mechanics,
Engineers Australia, 2010, p 1307.
151. Niranjan, R. R.; Junaid Kokan, S.; Sathya Narayanan, R.; Rajesh, S.; Manickavasagam, V.
M.; Ramnath, B. V. “Fabrication and testing of abaca fibre reinforced epoxy composites for
Downloaded by [University of Boras] at 03:48 29 January 2015

automotive applications”, Advanced Materials Science 2013, 718, 63–68.


152. Reinking, O. A. “Review of bunchy top disease of abaca”, Plant Disease Reporter 1950 34,
63–65.
153. Wardlaw, C. W. Banana Diseases, Including Plantations and Abaca; Longmans Green, Cornell
University: New York, USA, 1961.
154. Reinking, O. A. Review of Abaca Mosaic Control Program in the Philippines, 1950–1954;
1955.
155. Magnaye, L. V.; Halos, P. M. Recorded abaca diseases in the Philippines. In: Proceedings
of the Symposium on Philippine Phytopathology 1917–1977. Department of Plant Pathology,
Auditorium, UP at Los Banos, College, Laguna, Philippine Phytopathological Society, 14–15
December 1977, pp 271–289.
156. Jarman, C. G. “Banana fiber: A review of its properties and small-scale extraction and process-
ing”, Tropical Science 1977, 19(4), 173–185.
157. Leao, A.; Sartor, S. M.; Caraschi, J. C. “Natural fibers based composites–technical and social
issues”, Molecular Crystals and Liquid Crystals 2006, 448 (1), 161–177.
158. Franck, R. R. Bast and Other Plant Fibers; 2005, p. 39, CRC Press: Boca Raton, FL.
159. Zah, R. “Curaua fibers in the automobile industry: A sustainability assessment”, Journal of
Cleaner Production, 2007, 15(11), 1032–1040.
160. Leao, A. L.; Rowell, R.; Tavares, N. “Applications of natural fibers in automotive industry in
Brazil–thermoforming process”, Science and Technology of Polymers and Advanced Materials
1998, 755–761.
161. Spinace, M. A. S.; Lambert, C. S.; Fermoselli, K. K. G.; De Paoli, M. A. “Characterization of
lignocellulosic curaua fibres”, Carbohydrate Polymers 2009, 77(1), 47–53.
162. Silva, R. V.; Aquino, E. M. F. “Curaua fiber: A new alternative to polymeric composites”,
Journal of Reinforced Plastics and Composites 2008, 27(1), 103–112.
163. Schuh, T. G.; Gayer, U. Lignocellulosic–Plastic Composites; Universdade de Sao Paulo: Sao
Paulo, Brazil, 1997.
164. Mothe, C. G.; de Araujo, C. R.; Wang, S. H. “Thermal and mechanical characteristics of
polyurethane/curaua fiber composites”, Journal of Thermal Analysis and Calorimetry 2009,
95(1), 181–185.
165. Mano, B.; Araujo, J. R.; Spinace, M. A. S.; De Paoli, M. A. “Polyolefin composites with curaua
fibres: Effect of the processing conditions on mechanical properties, morphology and fiber
dimensions”, Composite Science and Technology 2010, 70(1), 29–35.
166. Araujo, J. R.; Waldman, W. R.; De Paoli, M. A. “Thermal properties of high density polyethylene
composites with natural fibres: Coupling agent effect”, Polymer Degradation and Stability 2008,
93(10), 1770–1775.
152 S. K. Ramamoorthy et al.

167. Da Silva, L. V.; Junior, J. H. S. A.; Angrizani, C. C.; Amico, S. C. “Short beam strength of
curaua, sisal, glass and hybrid composites”, Journal of Reinforced Plastics and Composites
2013, 32(3), 197–206.
168. Junior, J. H. S. A.;Junior, H. L. O.; Amico, S C.; Amado, F. D. R. “Study of hybrid intralaminate
curaua/glass composites”, Materials & Design 2012, 42, 111–117.
169. Spinace, M. A. S.; Fermoseli, K. K.; De Paoli, M. A. “Recycled polypropylene reinforced with
curaua fibers by extrusion”, Journal of Applied Polymer Science 2009, 112(6), 3686–3694.
170. Harnnecker, F.; dos Santos Rosa, D.; Lenz, D. M. “Biodegradable polyester-based blend re-
inforced with curaua fiber: Thermal, mechanical and biodegradation behavior”, Journal of
Polymers and the Environment 2012, 20(1), 237–244.
171. Caraschi, J. C.; Leao, A. L. Mechanical properties of curaua fiber reinforced polypropylene
composites. In: ISNaPol/2000: Third International Symposium on Natural Polymers and Com-
posites and the Workshop on Progress in Production and Processing of Cellulosic Fibers and
Natural Polymers, 2000, pp. 450–453.
172. Leao, A. L.; Machado, I. S.; De Souza, S. F.; Soriano, L. Production of curaua (Ananas Erec-
Downloaded by [University of Boras] at 03:48 29 January 2015

tifolius LB SMITH) fibers for industrial applications: Characterization and micropropagation.


In: VI International Pineapple Symposium, 2007, Vol. 822, pp. 227–238.
173. Borsoi, C.; Scienza, L. C.; Zattera, A. J. “Characterization of composites based on recycled
expanded polystyrene reinforced with curaua fibers”, Journal of Applied Polymer Science 2013,
128(1), 653–659.
174. Ornaghi, H. L.; Da Silva, H. S. P.; Zattera, A. J.; Amico, S. C. “Dynamic mechanical properties
of curaua composites”, Journal of Applied Polymer Science 2012, 125(S2), E110–E116.
175. D’Almeida, A. L. F. S.; Melo Filho, J. A.; Toledo Filho, R. D. “Use of curaua fibers
as reinforcement in cement composites”, Chemical Engineering Transactions 2009, 17,
1717–1722.
176. Caraschi J. S.; Leato, A. L. “Characterization of curaua fiber”, Molecular Crystals and Liquid
Crystals 2000, 353(1), 149–152.
177. Collins, J. L. The Pineapple: Botany, Cultivation and Utilization; Interscience Publishers: New
York, USA, 1961.
178. Mishra, S.; Mohanty, A. K.; Drzal, L. T.; Misra, M.; Hinrichsen, G. “A review on pineapple
leaf fibers, sisal fibers and their biocomposites”, Macromolecular Materials and Engineering
2004, 289(11), 955–974.
179. Liu, W.; Misra, M.; Askeland, P.; Drzal, L. T.; Mohanty, A. K. “‘Green’ composites from soy
based plastic and pineapple leaf fiber: Fabrication and properties evaluation”, Polymer 2005,
46(8), 2710–2721.
180. George, J.; Bhagawan, S. S.; Prabhakaran, N.; Thomas, S. “Short pineapple-leaf-fiber-reinforced
low-density polyethylene composites”, Journal of Applied Polymer Science 1995, 57(7)
843–854.
181. Lopattananon, N.; Panawarangkul, K.; Sahakaro, K.; Ellis, B. “Performance of pineapple leaf
fiber-natural rubber composites: The effect of fiber surface treatments”, Journal of Applied
Polymer Science 2006, 102(2) 1974–1984.
182. Arib, R. M. N.; Sapuan, S. M.; Ahmad, M. M. H. M.; Paridah, M. T.; Khairul Zaman, H. M. D.
“Mechanical properties of pineapple leaf fibre reinforced polypropylene composites”, Materials
& Design 2006, 27(5), 391–396.
183. Devi, L. U.; Bhagawan, S. S.; Thomas, S. “Mechanical properties of pineapple leaf
fiber–reinforced polyester composites”, Journal of Applied Polymer Science 1997, 64(9)
1739–1748.
184. Luo, S.; Netravali, A. N. “Interfacial and mechanical properties of environment–friendly “green”
composites made from pineapple fibers and poly(hydroxybutyrate-co-valerate) resin”, Journal
of Material Science 1999, 34(15), 3709–3719.
185. Mangal, R.; Saxena, N. S.; Sreekala, M. S.; Thomas, S.; Singh, K. “Thermal properties
of pineapple leaf fiber reinforced composites”, Material Science and Engineering: A 2003,
339(1–2), 281–285.
A Review of Natural Fibers Used in Biocomposites 153

186. Vijayan, K. Technical Brief No. MT-TB-1-80, (1980), National Aeronautical Laboratory, Ban-
galore, India.
187. Datta, A. K.; Samantaray, B. K.; Bhattacherjee, S. “Mechanical and dielectric properties of
pineapple fibres”, Journal of Materials Science Letters 1984, 3(8), 667–670.
188. Dutta, A. K..; Mukherjee, P. S.; Mitra, G. B. “A dielectric study of cellulose fibres”, Journal of
Materials Science 1980, 15(7), 1856–1860.
189. “Sisal”, In: Encyclopedia Britannica, 2012. Available at: http://global.britannica.
com/EBchecked/topic/546658/sisal
190. Lock, G. W. W. Sisal; Longmans Green: London, UK, 1962.
191. Mukherjee, P. S.; Satyanarayana. “Structure and properties of some vegetable fibres”, Journal
of Material Science 1984, 19(12), 3925–3934.
192. Bisanda, E. T. N.; Ansell, M. P. “The effect of silane treatment on the mechanical and phys-
ical properties of sisal-epoxy composites”, Composites Science and Technology 1991, 41(2),
165–178.
193. Chand, N.; Tiwary, R. K.; Rohatgi, P. K. “Bibliography resource structure of natural cellulosic
Downloaded by [University of Boras] at 03:48 29 January 2015

fibres: An annotated bibliography”, Journal of Material Science 1988, 23(2), 381–387.


194. Oksman, K.; Wallström, L.; Berglund, L. “Morphology and mechanical properties of uni-
directional sisal-epoxy composites”, Journal of Applied Polymer Science 2002, 84(13),
2358–2365.
195. Bakare, I. O.; Okieimen, F. E.; Pavithran, C.; Abdul Khalil, H. P. S.; Brahmakumar, M. “Me-
chanical and thermal properties of sisal fiber-reinforced rubber seed oil-based polyurethane
composites”, Materials & Design 2010, 31(9), 4274–4280.
196. Costa, F. H. M. M.; d’Almeida, J. R. M. “Effect of water absorption on the mechanical properties
of sisal and jute fiber composites”, Polymer-Plastics Technology and Engineering 1999, 38(5),
1081–1094.
197. Joseph, P. V.; Mathew, G.; Joseph, K.; Groeninckx, G.; Thomas, S. “Dynamic mechanical
properties of short sisal fibre reinforced polypropylene composites”, Composites Part A: Applied
Science and Manufacturing 2003, 34(3), 275–290.
198. Rong, M. Z.; Zhang, M. Q.; Liu, Y.; Yan, H. M.; Yang, G. C.; Zeng, H. M. “Interfacial interaction
in sisal/epoxy composites and its influence on impact performance”, Polymer Composites 2002,
23(2), 182–192.
199. Joseph, K.; Thomas, S.; Pavithran, C.; Brahmakumar, M. “Tensile properties of short sisal
fiber-reinforced polyethylene composites”, Journal of Applied Polymer Science 1993, 47(10),
1731–1739.
200. Chand, N.; Joshi, S. K. “Temperature dependence of dielectric behavior of sisal fibre”, Journal
of Material Science Letters 1994, 13(3), 156–158.
201. Chand, N.; Hashmi, S. A. R. “Mechanical properties of sisal fibre at elevated temperatures”,
Journal of Material Science, 1993, 28(24), 6724–6728.
202. Megiatto Jr, J. D.; Silva, C. G.; Ramires, E. C.; Frollini, E. “Thermoset matrix reinforced with
sisal fibers: effect of the cure cycle on the properties of the biobased composites”, Polymer
Testing 2009, 28(8), 793–800.
203. Sydenstricker, T. H.; Mochnaz, S.; Amico, S. C. “Pull-out and other evaluations in sisal-
reinforced polyester composites”, Polymer Testing 2003, 22(4), 375–380.
204. Joseph, K.; Toledo Filho, R. D.; James, B.; Thomas, S.; Carvalho, L. H. “A review on sisal fiber
reinforced polymer composites”, Revista Brasileira de Engenharia Agricola e Ambiental 1999,
3(3), 367–379.
205. Luo, S.; Netravali, A N. “Characterization of henequen fibers and the henequen fiber/poly
(hydroxybutyrate-co-hydroxyvalerate) interface”, Journal of Adhesion Science and Technology
2001, 15(4), 423–437.
206. Cazaurang-Martinez, M. N.; Herrera-Franco, P. J.; Gonzalez-Chi, P. I.; Aguilar-Vega, M. “Phys-
ical and mechanical properties of henequen fibers”, Journal of Applied Polymer Science 1991,
43(4), 749–756.
154 S. K. Ramamoorthy et al.

207. Gonzalez-Murillo, C.; Ansell, M. P. “Mechanical properties of henequen fibre/epoxy resin


composites”, Mechanics of Composite Materials 2009, 45(4), 435–442.
208. Han, Y. H.; Han, S. O.; Cho, D.; Kim, H. I. “Henequen/unsaturated polyester biocompos-
ites: Electron beam irradiation treatment and alkali treatment effects on the henequen fiber”,
Macromolecular Symposia 2006, 245(1), 539–548.
209. Maleque, M.; Belal, F. Y.; Sapuan, S. M. “Mechanical properties study of pseudo-stem banana
fiber reinforced epoxy composite”, The Arabian Journal for Science and Engineering 2007,
32(2B), 359–364.
210. Sapuan, S. M.; Leenie, A.; Harimi, M.; Beng, Y. K. “Mechanical properties of woven banana
fibre reinforced epoxy composites”, Materials & Design 2006, 27(8), 689–693.
211. Bilba, K.; Arsene, M. A.; Ouensanga, A. “Study of banana and coconut fibers: botanical
composition, thermal degradation and textural observations”, Bioresource Technology 2007,
98(1), 58–68.
212. Harish, S.; Peter Michael, D.; Bensely, A.; Mohal Lal, D.; Rajadurai, A. “Mechanical prop-
erty evaluation of natural fiber coir composite”, Materials Characterization 2009, 60(1),
Downloaded by [University of Boras] at 03:48 29 January 2015

44–49.
213. Viswanathan, R.; Gothandapani, L.; Kailappan, R. “Water absorption and swelling characteris-
tics of coir pith particle board”, Bioresource Technology 2000, 71(1), 93–94.
214. Geethamma, V. G.; Thomas, S. “Diffusion of water and artificial seawater through coir fiber
reinforced natural rubber composites”, Polymer Composites 2005, 26(2), 136–143.
215. Rout, J.; Misra, M.; Tripathy, S. S.; Nayak, S. K.; Mohanty, A. K. “The influence of fibre treat-
ment on the performance of coir-polyester composites”, Composites Science and Technology
2001, 61(9), 1303–1310.
216. Espert, A.; Vilaplana, F.; Karlsson, S. “Comparison of water absorption in natural cellulosic
fibres from wood and one-year crops in polypropylene composites and its influence on their
mechanical properties”, Composites Part A: Applied Science and Manufacturing 2004, 35(11),
1267–1276.
217. Bismarck, A.; Mohanty, A. K.; Aranberri-Askargorta, I.; Czapla, S.; Misra, M.; Hinrichsen, G.;
Springer, J. “Surface characterization of natural fibers: Surface properties and the water up-take
behavior of modified sisal and coir fibers”, Green Chemistry 2001, 3(2), 100–107.
218. Prasad, M. “Physical, chemical and biological properties of coir dust”, International Symposium
Growing Media and Plant Nutrition in Horticulture 1996, 450, 21–30.
219. Satyanarayana, K. G.; Pillai, C. K. S.; Sukumaran, K.;Pillai, S. G. K.; Rohatgi, P. K.; Vijayan,
K. “Structure property studies of fibres from various parts of the coconut tree”, Journal of
Material Science 1982, 17(8), 2453–2462.
220. Geethamma, V. G.; Kalaprasad, G.; Groeninckx, G.; Thomas, S. “Dynamic mechanical behavior
of short coir fiber reinforced natural rubber composites”, Composites Part A: Applied Science
and Manufacturing 2005, 36(11), 1499–1506.
221. Haque, M. M.; Hasan, M.; Islam, M. S.; Ali, M. E. “Physico-mechanical properties of chemically
treated palm and coir reinforced polypropylene composites”, Bioresource Technology 2009,
100(20), 4903–4906.
222. Monteiro, S. N.; Terrones, L. A. H.; d’Almeida, J. R. M. “Mechanical performance of coir
fiber/polyester composites”, Polymer Testing 2008, 27(5), 591–595.
223. Goulart Silva, G.; De Souza, D. A.; Machado, J. C.; Hourston, D. J. “Mechanical and thermal
characterization of native Brazilian coir fiber”, Journal of Applied Polymer Science 2000, 76(7),
1197–1206.
224. Lai, C. Y.; Sapuan, S. M.; Ahmad, M.; Yahya, N.; Dahlan, K. Z. H. M. “Mechanical and elec-
trical properties of coconut coir fiber-reinforced polypropylene composites”, Polymer–Plastic
Technology and Engineering 2005, 44(4), 619–632.
225. Rout, J.; Tripathy, S. S.; Nayak, S. K.; Misra, M.; Mohanty, A. K. “Scanning electron microscopy
study of chemically modified coir fibers”, Journal of Applied Polymer Science 2001, 79(7),
1169–1177.
A Review of Natural Fibers Used in Biocomposites 155

226. Asasutjarit, C.; Charoenvai, S.; Hirunlabh, J.; Khedari, J. “Materials and mechanical properties
of pretreated coir-based green composites”, Composites Part B: Engineering 2009, 40(7),
633–637.
227. Nor, M. J. M.; Jamaludin, N.; Tamiri, F. M. “A preliminary study of sound absorption using
multi-layer coconut coir fibers”, Electronic Journal Technical Acoustics 2004, 3, 1–8.
228. Varma, D. S.; Varma, M.; Varma, I. K. “Coir fibres II: Evaluation as a reinforcement in unsat-
urated polyester resin composites”, Journal of Reinforced Plastics and Composites 1985, 4(4),
419–431.
229. MacNeish, R. S. Ancient Mesoamerica Civilization; Bobbs-Merrill: Indianapolis, Indiana, USA,
1964.
230. Huckell, L. W. “Plant remains from the Pinaleno cotton cache, Arizona”, The Kiva, 1993, 59(2),
147–203.
231. Fuller, D. Q. “The spread of textile production and textile crops in India beyond the Harappan
zone: An aspect of the emergence of craft specialization and systematic trade: Linguistics,
archaeology and the human past”, Industrial Project. Kyoto: Research Institute of Humanity
Downloaded by [University of Boras] at 03:48 29 January 2015

and Nature, 2008, pp. 1–26.


232. Basra, A. S.; Malik, C. P. “Development of cotton fiber”, International Review of Cytology
1984, 89(1), 65–113.
233. Hossain, F.; Pray, C. E.; Lu, Y.; Huang, J.; Fan, C.; Hu, R. “Genetically modified cotton and
farmers’ health in China”, International Journal of Occupational and Environmental Health
2004, 10(3), 296–303.
234. Bennett, R. M.; Ismael, Y.; Kambhampati, U.; Morse, S. “Economic impact of genetically
modified cotton in India”, AgBioForum 2004, 7(3), 96–100.
235. Su, C. I.; Fang, J. X.; Chen, X. H.; Wu, W. Y. “Moisture absorption and release of pro-
filed polyester and cotton composite knitted fabrics”, Textile Research Journal 2007, 77(10),
764–769.
236. Mwaikambo, L. Y.; Bisanda, E. T. “The performance of cotton-kapok fabric-polyester compos-
ites”, Polymer Testing 1999, 18(3), 181–198.
237. Mwaikambo, L. Y.; Martuscelli, E.; Avella, M. “Kapok/cotton fabric-polypropylene compos-
ites”, Polymer Testing 2000, 19(8), 905–918.
238. Kim, S. J.; Moon, J. B.; Kim, G. H.; Ha, C. S. “Mechanical properties of polypropylene/natural
fiber composites: A comparison of wood fiber and cotton fiber”, Polymer Testing, 2008, 27(7),
801–806.
239. Kamath, M. G.; Bhat, G. S.; Parikh, D. V.; Mueller, D. “Cotton fiber nonwovens for automotive
composites”, International Nonwovens Journal 2005, 14(1), 34–40.
240. Zou, Y.; Reddy, N.; Yang, Y. “Reusing polyester/cotton blend fabrics for composites”, Com-
posites Part B: Engineering 2011, 42(4), 763–770.
241. DeVallance, D. B.; Gray, J.; Lentz, H. “Properties of wood/recycled textile composite panels”,
Wood and Fiber Science 2012, 44(3), 310–318.
242. Mueller, D. H.; Krobjilowski, A. “New discovery in the properties of composites reinforced
with natural fibers”, Journal of Industrial Textiles 2003, 33(2), 111–130.
243. Kiple, K. F.; Ornelas, K. C. Cambridge World History of Food; Cambridge University Press:
UK, 2000.
244. Abdul Khalil, H. P. S.; Siti Alwani, M.; Ridzuan, R.; Kamarudin, H.; Khairul, A. “Chemical
composition, morphological characteristics, and cell wall structure of Malaysian oil palm fibers”,
Polymer-Plastics Technology and Engineering 2008, 47(3), 273–280.
245. Wirjosentono, B.; Guritno, P.; Ismail, H. “Oil palm empty fruit bunch filled polypropy-
lene composites”, International Journal of Polymeric Materials 2004, 53(4), 295–
306.
246. Rozman, H. D.; Lai, C. Y.; Ismail, H.; Ishak, Z. “The effect of coupling agent on the mechanical
and physical properties of oil palm empty fruit bunch-polypropylene composites”, Polymer
International 2000, 49(11), 1273–1278.
156 S. K. Ramamoorthy et al.

247. Raju, G.; Ratnam, C. T.; Ibrahim, N. A.; Rahman M. Z. A.; Yunus W. M. Z. W. “Enhancement
of PVC/ENR blend properties by poly (methyl acrylate) grafted oil palm empty fruit bunch
fiber”, Journal of Applied Polymer Science 2008, 110(1), 368–375.
248. Jayashree, E.; Mandal, P. K.; Madhava, M.; Kamraj, A.; Sireesha, K.; Sreedharan, P. K.; Ku-
mar, V.; Chulaki, B. M. “Development of decorticator for extraction of quality fibre from
oil palm empty fruit bunches”, In: Proceeding of the 15th Plantation Crops Symposium
Placrosym XV, Mysore, India, 10–13 December, Central Coffee Research Institute, 2002,
pp. 678–681.
249. Khalil, H. P. S. A.; Alwani, M. S.; Omar, A. K. M. “Chemical composition, anatomy, lignin
distribution, and cell wall structure of Malaysian plant waste fibers”, BioResources, 2007, 1(2),
220–232.
250. Sreekala, M. S.; Kumaran, M. G.; Thomas, S. “Oil palm fibers: Morphology, chemical compo-
sition, surface modification, and mechanical properties”, Journal of Applied Polymer Science
1997, 66(5), 821–835.
251. Hill, C. A. S.; Abdul, H. P. S.; Khalil, H. P. S. A. “The effect of environment exposure upon
Downloaded by [University of Boras] at 03:48 29 January 2015

the mechanical properties of coir or oil palm fiber reinforced composites”, Journal of Applied
Polymer Science 2000, 77(6), 1322–1330.
252. Sreekala, M. S.; Kumaran, M. G.; Thomas, S. “Water sorption in oil palm fiber reinforced
phenol formaldehyde composites”, Composites Part A: Applied Science and Manufacturing
2002, 33(6), 763–777.
253. Jacob, M.; Francis, B.; Thomas, S.; Varughese, K. T. “Dynamic mechanical analysis of sisal/oil
palm hybrid fiber–reinforced natural rubber composites”, Polymer Composites 2006, 27(6),
671–680.
254. Piarpuzan, D.; Quintero, J. A.; Cardona, C. A. “Empty fruit bunches from oil palm as a
potential raw material for fuel ethanol production”, Biomass and Bioenergy 2011, 35(3)
1130–1137.
255. Mohammed, M. A. A..; Salmiaton, A.; Wan Azlina, W. A. K. G.; Mohamad Amran, M. S.
“Gasification of oil palm empty fruit bunches: A characterization and kinetic study”, Bioresource
Technology, 2012, 110, 628–636.
256. Shinoj, S.; Visvanathan, R.; Panigrahi, S.; Kochubabu, M. “Oil palm fiber (OPF) and its com-
posites: A review”, Industrial Crops and Products 2011, 33(1), 7–22.
257. Reddy, G. V.; Naidu, S. V.; Rani, T. “Impact properties of kapok based unsaturated
polyester hybrid composites”, Journal of Reinforced Plastics and Composites. 2008, 27(16–17),
1789–1804.
258. Reddy, G. V.; Rani, T. S.; Rao, K. C.; Naidu, S. V. “Flexural, compressive, and interlaminar shear
strength properties of kapok/glass composites”, Journal of Reinforced Plastics and Composites
2009, 28(14), 1665–1677.
259. Boynard, C. A.; d’Almeida, J. R. M. “Water absorption by sponge gourd (luffa cylindrical)-
polyester composite materials”, Journal of Material Science 1999, 18(21), 1789–1791.
260. Boynard, C. A.; d’Almeida, J. R. M. “Morphological characterization and mechanical behavior
of sponge gourd (Luffa cylindrical)-polyester composite materials”, Polymer–Plastics Technol-
ogy and Engineering 2000, 39(3), 489–499.
261. Boynard, C. A.; Monteiro, S. N.; d’Almeida, J. R. M. “Aspects of alkali treatment of sponge
gourd (Luffa cylindrical) fibers on the flexural properties of polyester matrix composites”,
Journal of Applied Polymer Science 2003, 87(12), 1927–1932.
262. Demir, H.; Atikler, U.; Balkose, D.; Tihminlioglu, F. “The effect of fiber surface treatments on
the tensile and water sorption properties of polypropylene–luffa fiber composites”, Composites
Part A: Applied Science and Manufacturing 2006, 37(3), 447–456.
263. Kaewtatip, K.; Thongmee, J. “Studies on the structure and properties of thermoplastic
starch/luffa fiber composites”, Materials & Design 2012, 40, 314–318.
264. Verma, D.; Gope, P. C.; Maheswari, M. K.; Sharma, R. K. “Bagasse fiber composites: A review”,
Journal of Materials and Environmental Science 2012, 3(6), 1079–1092.
A Review of Natural Fibers Used in Biocomposites 157

265. Cardona, C. A.; Quintero, J. A.; Paz, I. C. “Production of bioethanol from sugarcane bagasse:
status and perspectives”, Bioresource Technology 2010, 101(13), 4754–4766.
266. Rabelo, S. C.; Carrere, H.; Maciel Filho, R.; Costa, A. C. “Production of bioethanol, methane and
heat from sugarcane bagasse in biorefinery concept”, Bioresource Technology 2011, 102(17),
7887–7895.
267. Van Zyl, C.; Prior, B. A.; Du Preez, J. C. “Production of ethanol from sugar cane bagasse
hemicellulose hydrolyzate by Pichia stipites”, Applied Biochemistry and Biotechnology 1988,
17(1–3), 357–369.
268. Luo, L.; Van Der Voet, E.; Huppes, G. “Life cycle assessment and life cycle costing of bioethanol
from sugarcane in Brazil”, Renewable and Sustainable Energy Reviews 2009, 13(6), 1613–1619.
269. Boopathy, R.; Dawson, L. “Cellulosic ethanol production from sugarcane bagasse without
enzymatic saccharification”, BioResources 2008, 3(2), 452–460.
270. Cundy, V. A.; Maples D.; Tauzin, C. “Combustion of bagasse: Use of an agricultural-derived
waste”, Fuel 1983, 62(7), 775–780.
271. Harel, P.; Baguant, J. “Bagasse combustion”, International Sugar Journal 1992, 94, 11–20.
Downloaded by [University of Boras] at 03:48 29 January 2015

272. Ramajo-Escalera, B.; Espina, A.; Garcia, J. R.; Sosa-Arnao, J. H.; Nebra, S. A. “Model-
free kinetics applied to sugarcane bagasse combustion”, Thermochimica Acta 2006, 448(2),
111–116.
273. Lamb, B. W.; Bilger, R. W. “Combustion of bagasse: Literature review”, Sugar Technology
Reviews 1977, 4(2), 89–130.
274. Pandey, A.; Soccol, C. R.; Nigam, P.; Soccol, V. T. “Biotechnological potential of argo-industrial
residues. I: sugarcane bagasse”, Bioresource Technology 2000, 74(1), 69–80.
275. Paiva, J. M. F.; Frollini, E. “Sugarcane bagasse reinforced phenolic and lignocellulosic co-
mosites”, Journal of Applied Polymer Science 2002, 83(4), 880–888.
276. Luz, S. M.; Goncalves, A. R..; Del’Arco, Jr, A. P. “Mechanical behavior and microstructural
analysis of sugarcane bagasse fibers reinforced polypropylene composites”, Composites Part
A: Appled Science and Manufacturing 2007, 38(6), 1455–1461.
277. Vazquez, A.; Dominguez, V. A.; Kenny, J. M. “Bagasse fiber-polypropylene based composites”,
Journal of Thermoplastic Composite Materials 1999, 12(6), 477–497.
278. Stael, G. C.; Tavares, M. I. B.; d’Almeida, J. R. M. “Impact behavior of sugarcane bagasse
waste-EVA composites”, Polymer Testing, 2001, 20(8), 869–872.
279. Lei, Y.; Wu, Q.; Yao, F.; Xu, Y. “Preparation and properties of recycled HDPE/natural
fiber composites”, Composites Part A: Applied Science and Manufacturing 2007, 38(7),
1664–1674.
280. Hassan, M. L.; Rowell, R. M.; Fadl, N. A.; Yacoub, S. F.; Christainsen, A. W. “Thermoplas-
ticization of bagasse. II. Dimensional stability and mechanical properties of esterified bagasse
composite”, Journal of Applied Polymer Science 2000, 76(4), 575–586.
281. Bilba, K.; Arsene, M. A.; Ouensanga, A. “Sugar cane bagasse fibre reinforced cement compos-
ites. Part I. Influence of the botanical components of bagasse on the setting of bagasse/cement
composite”, Cement and Concrete Composites 2003, 25(1), 91–96.
282. Loh, Y. R.; Sujan, D.; Rahman, M. E.; Das, C. A. “Sugarcane bagasse–The future composite
material: A literature review”, Resources, Conservation and Recycling 2013, 75, 14–22.
283. Justiz–Smith, N.; Junior Virgo, G.; Buchanan, V. E. “Potential of Jamaican banana, coconut coir
and bagasse fibres as composite materials”, Materials Characterization 2008, 59(9), 1273–1278.
284. Monteiro, S. N.; Rodriquez, R. J. S.; De Souza, M. V.; d’Almeida, J. R. M. “Sugar cane bagasse
waste as reinforcement in low cost composites”, Advanced Performance Materials 1998, 5(3),
183–191.
285. Lakkad, S. C.; Patel, J. M. “Mechanical properties of bamboo, a natural composite”, Fibre
Science and Technology 1981, 14(4), 319–322.
286. Yueping, W.; Ge, W.; Haitao, C.; Genlin, T.; Zheng, L.; Feng, X. Q.; Xiangqi, Z.; Xiaojun,
H.; Xushan, G. “Structures of bamboo fiber for textiles”, Textile Research Journal 2010, 80(4),
334–343.
158 S. K. Ramamoorthy et al.

287. Shuying, S. “A study of structure and performance of bamboo fibers”, Journal of Textile
Research 2003, 24(6), 27–28.
288. Hengshu, Z. “Study on the characteristics of bamboo fiber in spinning and weaving”, Journal
of Textile Research 2004, 25(5), 91–93.
289. Ray, A. K.; Das, S. K.; Mondal, S.; Ramachandrarao, P. “Microstructural characterization of
bamboo”, Journal of Materials Science 2004, 39(3), 1055–1060.
290. Lee, S. H.; Ohkita, T.; Kitagawa, K. “Eco–composite from poly (lactic acid) and bamboo fiber”,
Holzforschung 2004, 58(5), 529–536.
291. Takagi, H.; Kako, S.; Kusano, K.; Ousaka, A. “Thermal conductivity of PLA-bamboo fiber
composites”, Advanced Composite Materials 2007, 16(4), 377–384.
292. Xiao, Y.; Zhou, Q.; Shan, B. “Design and construction of modern bamboo bridges”, Journal of
Bridge Engineering 2009, 15(5), 533–541.
293. Xiao, Y.; Inoue, M.; Paudel, S. K. (Eds.) Modern Bamboo Structures: Proceeding of the First
International Conference; CRC Press: Boca Raton, FL, 2008.
294. Thwe, M. M.; Liao, K. “Durability of bamboo-glass fiber reinforced polymer matrix hybrid
Downloaded by [University of Boras] at 03:48 29 January 2015

composites”, Composites Science and Technology 2003, 63(3), 375–387.


295. Chen, X.; Guo, Q.; Mi, Y. “Bamboo fiber-reinforced polypropylene composites: a study of the
mechanical properties”, Journal of Applied Polymer Science 1998, 69(10), 1891–1899.
296. Mi, Y.; Chen, X.; Guo, Q. “Bamboo fiber-reinforced polypropylene composites: crystallization
and interfacial morphology”, Journal of Applied Polymer Science 1997, 64(7), 1267–1273.
297. Ghavami, K. “Bamboo as reinforcement in structural concrete elements”, Cement and Concrete
Composites 2005, 27(6), 637–649.
298. Khalil, H. P. S. A., Bhat, I. U. H., Jawaid, M, Zaidon, A, Hermawan, D, Hadi, YS. “Bamboo
fibre reinforced biocomposites: A review”, Materials & Design 2012, 42, 353–368.
299. Liu, D.; Song, J.; Andersen, D. P.; Chang, P. R.; Hua, Y. “Bamboo fiber and its reinforced
composites: structure and properties”, Cellulose 2012, 19(5), 1449–1480.
300. Sundstol, F.; Owen, E. Straw and Other Fibrous By-Products as Feed; Elsevier Science: Ams-
terdam, 1984.
301. Kaparaju, P.; Serrano, M.; Thomsen, A. B.; Kongjan, P.; Angelidaki, I. “Bioethanol, biohydrogen
and biogas production from wheat straw in a biorefinery concept”, Bioresource Technology 2009,
100(9), 2562–2568.
302. Zhang, Y.; Min, B.; Huang, L.; Angelidaki, I. “Generation of electricity and analysis of microbial
communities in wheat straw biomass-powered microbial fuel cells”, Applied and Environmental
Microbiology 2009, 75(11), 3389–3395.
303. Bainbridge, D. A. “High performance low cost buildings of straw”, Agriculture, Ecosystems &
Environment 1986, 16(3–4), 281–284.
304. Emerson, J. W. Production of high strength packaging papers from straw U.S. Patent No.
4,040,899, 1977.
305. Molina, J.; Sikora, M.; Garud, N.; Flowers, J. M.; Rubinstein, S.; Reynolds, A.; Huang, P.
“Molecular evidence for a single evolutionary origin of domesticated rice”, Proceedings of the
National Academy of Sciences 2011, 108(20), 8351–8356.
306. Datta, D. Principles and Practices of Rice Production; International Rice Research Institute:
Manila, 1981.
307. Kamel, S. “Preparation and properties of composites made from rice straw and poly (vinyl
chloride) (PVC)”, Polymers for Advance Technologies 2004, 15(10), 612–616.
308. Buzarovska, A.; Bogoeva, G. G.; Grozdanov, A.; Avella, M.; Gentile, G.; Errico, M. “Potential
use of rice straw as filler in eco-composite materials”, Australian Journal of Crop Science 2008,
1(2), 37–42.
309. Yao, F.; Wu, Q.; Lei, Y.; Xu, Y. “Rice straw fiber-reinforced high-density polyethylene com-
posite: Effect of fiber type and loading”, Industrial Crops and Products 2008, 28(1), 63–72.
310. Grozdanov, A.; Buzarovska, A.; Bogoeva, G. G.; Avella, M.; Errico, M. E.; Gentile, G. “Rice
straw as an alternative reinforcement in polypropylene composites”, Agronomy for Sustainable
Development 2006, 26(4), 251–255.
A Review of Natural Fibers Used in Biocomposites 159

311. Prasad, R.; Rao, K. M. M.; Avssks, G. “Tensile and impact behaviour of rice straw polyester
composites”, Indian Journal of Fibre & Textile Research 2007, 32, 399–403.
312. Yang, H. S.; Kim, D. J.; Kim, H. J. “Rice straw–wood particle composite for sound absorbing
wooden construction materials”, Bioresource Technology 2003, 86(2), 117–121.
313. Ahmad, I.; Yaseen, A.; Mir, M. M. “Biochemical constituents of the developing grains of wheat
cultivars”, International Journal of Biosciences 2012, 2(2), 53–55.
314. Gooding, M. J.; Davies, W. P. Wheat production and utilization: Systems, quality and the
environment; CAB International: Wallingford, Oxon, UK, 1997.
315. Jiamin, L. Z. Y. M. F.; Rui, H. “Composite of wheat straw/PP”, China Plastic Industry 2000, 4,
003.
316. Zou, Y.; Huda, S.; Yang, Y. “Lightweight composites from long wheat straw and polypropylene
web”, Bioresource Technology 2010, 101(6), 2026–2033.
317. Alemdar, A.; Sain, M. “Biocomposites from wheat straw nanofibers: Morphology, thermal and
mechanical properties”, Composites Science and Technology, 2008, 68(2), 557–565.
318. Avella, M.; Rota, G. L.; Martuscelli, E.; Raimo, M.; Sadocco, P.; Elegir, G.; Riva, R. “Poly(3-
Downloaded by [University of Boras] at 03:48 29 January 2015

hydroxybutyrate-co-3-hydroxyvalerate) and wheat straw fibre composites: Thermal, mechanical


properties and biodegradation behavior”, Journal of Material Science 2000, 35(4), 829–836.
319. Mengeloglu, F.; Karakus, K. “Thermal degradation, mechanical properties and morphology of
wheat straw flour filled recycled thermoplastic composites”, Sensors 2008, 8(1), 500–519.
320. Han, G. “Development of high–performance reed and wheat straw composite panels”, Wood
Research: Bulletin of the Wood Research Institute Kyoto University 2001, 88, 19–39.
321. Shands, H. L.; Dickson, A. D. “Barley: Botany, production, harvesting, processing, utilization
and economics”, Economic Botany 1953, 7(1), 3–26.
322. Ullrich, S. E. Barley: Production, Improvement, and Uses; Wiley: NJ, 2010.
323. Bouhicha, M.; Aouissi, F.; Kenai, S. “Performance of composite soil reinforced with barley
straw”, Cement and Concrete Composites, 2005, 27(5), 617–621.
324. White, N. M.; Ansell, M. P. “Straw-reinforced polyester composites”, Journal of Material
Science 1983, 18(5), 1549–1556.
325. Belayachi, N.; Bouasker, M.; Hoxha, D.; Al-Mukhtar, M. “Thermo-mechanical behaviour of
an innovant straw lime composite for thermal insulation applications”, Applied Mechanics and
Materials 2013, 390, 542–546.
326. Sun, R. C.; Sun, X. F. “Structural and thermal characterization of acetylated rice, wheat, rye,
and barley straws and poplar wood fibre”, Industrial Crops and Products 2002, 16(3), 225–235.
327. Della, V. P.; Kuhn, I.; Hotza, D. “Rice husk ash as an alternate source for active silica production”,
Materials Letters 2002, 57(4), 818–821.
328. Ishak, Z. A.; Yow, B. N.; Ng, B. L.; Khalil, H. A.; Rozman, H. D. “Hygrothermal aging and
tensile behavior of injection–molded rice husk-filled polypropylene composites”, Journal of
Applied Polymer Science 2001, 81(3), 742–753.
329. Premalal, H. G.; Ismail, H.; Baharin, A. “Comparison of the mechanical properties of rice husk
powder filled polypropylene composites with talc filled polypropylene composites”, Polymer
Testing 2002, 21(7), 833–839.
330. Yang, H. S.; Kim, H. J.; Son, J.; Park, H. J.; Lee, B. J.; Hwang, T. S. “Rice-husk flour filled
polypropylene composites; mechanical and morphological study”, Composite Structures 2004,
63(3), 305–312.
331. Kim, H. S.; Yang, H. S.; Kim, H. J.; Park, H. J. “Thermogravimetric analysis of rice husk flour
filled thermoplastic polymer composites”, Journal of Thermal Analysis and Calorimetry 2004,
76(2), 395–404.
332. Panthapulakkal, S.; Law, S.; Sain, M. “Enhancement of processability of rice husk filled high-
density polyethylene composite profiles”, Journal of Thermoplastic Composite Materials 2005,
18(5), 445–458.
333. Bledzki, A. K.; Mamun, A. A.; Volk, J. “Barley husk and coconut shell reinforced polypropylene
composites: the effect of fibre physical, chemical and surface properties”, Composites Science
and Technology 2010, 70(5), 840–846.
160 S. K. Ramamoorthy et al.

334. Huda, S.; Yang, Y. “Chemically extracted cornhusk fibers as reinforcement in light-weight
poly(propylene) composites”, Macromolecular Materials and Engineering 2008, 293(3),
235–243.
335. Mackin, T. A.; Sottos, N. R.; White, S. R. Corn-based structural composites. U.S. Patent No.
5,834,105. Washington, DC, 1998.
336. Bledzki, A. K.; Mamun, A. A.; Volk, J. “Physical, chemical and surface properties of wheat
husk, rye husk and soft wood and their polypropylene composites”, Composites Part A: Applied
Science and Manufacturing, 2010, 41(4), 480–488.
337. Mohanty, A. K.; Misra, M.; Drzal, L. T. (Eds.) Natural Fibers, Biopolymers, and Biocomposites;
CRC Press: Boca Raton, FL, 2005.
338. Stokke, D. D.; Wu, Q.; Han, G. Introduction to Wood and Natural Fiber Composites; Wiley:
NJ, 2013.
339. Stark, N. M.; Rowlands, R. E. “Effects of wood fiber characteristics on mechanical properties
of wood/polypropylene composites”, Wood and Fiber Science 2003, 35(2), 167–174.
340. Coutinho, F. M. B.; Costa, T. H. S.; Carvalho, D. L. “Polypropylene-wood fiber composites:
Downloaded by [University of Boras] at 03:48 29 January 2015

Effect of treatment and mixing conditions on mechanical properties”, Journal of Applied Science
1997, 65(6), 1227–1235.
341. Selke, S. E.; Wichman, I. “Wood fiber/polyolefin composites”, Composites Part A: Applied
Science and Manufacturing 2004, 35(3), 321–326.
342. Bledzki, A. K.; Faruk, O.; Huque, M. “Physico-mechanical studies of wood fiber reinforced
composites”, Polymer-Plastics Technology and Engineering 2002, 41(3), 435–451.
343. Peltola, H.; Pääkkönen, E.; Jetsu, P.; Heinemann, S. “Wood based PLA and PP composites: Ef-
fect of fiber type and matrix polymer on fibre morphology, dispersion and composite properties”,
Composites Part A: Applied Science and Manufacturing 2014, 61, 13–22.
344. Niska, K. O.; Sain, M. (Eds.) Wood–Polymer Composites; Elsevier; Amsterdam, 2008.
345. Bledzki, A. K.; Reihmane, S.; Gassan, J. “Thermoplastics reinforced with wood fillers: A
literature review”, Polymer-Plastics Technology and Engineering 1998, 37(4), 451–468.
346. Klyosov, A. A. Wood-Plastic Composites; Wiley: NJ, 2007.
347. Wallenberger, F. T.; Weston N. E. (Eds.) Natural Fibers, Plastics and Composites; Springer:
NY, 2004.
348. Blackburn, R. S. (Ed.) Biodegradable and Sustainable Fibres; Elsevier; Amsterdam, 2005.
349. Franck, R. R. (Ed.) Bast and Other Plant Fibres; CRC Press: Boca Raton, FL, 2005.
350. Thomas, S.; Pothan, L. A. (Eds.) Natural fibre reinforced polymer composites: From macro to
nanoscale. In: Archives Contemporaines. Old City Publishing: Philadelphia, PA, USA, 2009.
351. Mussig, J. (Ed.) Industrial Applications of Natural Fibres: Structure, Properties and Technical
Applications; Wiley: NJ, 2010.
352. Kaplan, D. L.; Mello, C. M.; Arcidiacono, S.; Fossey, S.; Senecal, K.; Muller, W. Silk;
Birkhäuser: Boston, 1997, pp 103–131.
353. Bhat, N. V.; Hadiger, G. S. “Crystallinity in silk fibers: Partial acid hydrolysis and related
studies”, Journal of Applied Polymer Science 1980, 25(5), 921–932.
354. Cunniff, P. M.; Fossey, S. A.; Auerbach, M. A.; Song, J. W.; Kaplan, D. L.; Adams, W. W.; Eby,
R. K.; Mahoney, D.; Vezie, D. L. “Mechanical and thermal properties of dragline silk from the
spider Nephila clavipes”, Polymers for Advanced Technologies 1994, 5(8), 401–410.
355. Magoshi, J.; Magoshi, Y.; Nakamura, S. Mechanism of fiber formation of silkworm. In: ACS
Symposium Series, 1994.
356. Glisovic, A.; Salditt, T. “Temperature dependent structure of spider silk by X-ray diffraction”,
Applied Physics A 2007, 87(1), 63–69.
357. Gosline, J. M.; Guerette, P. A.; Ortlepp, C. S.; Savage, K. N. “The mechanical design of spider
silks: from fibroin sequence to mechanical function”, Journal of Experimental Biology 1999,
202(23), 3295–3303.
358. Akhtar, S.; De, P. P.; De, S. K. “Short fiber-reinforced thermoplastic elastomers from blends
of natural rubber and polyethylene”, Journal of Applied Polymer Science, 1986, 32(5),
5123–5146.
A Review of Natural Fibers Used in Biocomposites 161

359. Kluge, J. A.; Rabotyagova, O.; Leisk, G. G.; Kaplan, D. L. “Spider silks and their applications”,
Trends in Biotechnology 2008, 26(5), 244–251.
360. Vollrath, F. “Strength and structure of spiders’ silks”, Reviews in Molecular Biotechnology
2000, 74(2), 67–83.
361. Hakimi, O.; Knight, D. P.; Vollrath, F.; Vadgama, P. “Spider and mulberry silkworm silks as
compatible biomaterials”, Composites Part B: Engineering 2007, 38(3), 324–337.
362. Verpoest, I.; Vuure, V.; El Asmar, N.; Vanderbeke, J. Silk fibre composites U.S. Patent No.
20,100,040,816. 2007.
363. Zach, J. “Possibilities of using waste sheep wool for manufacture of thermo-insulating material”,
Central for Integrated Design of Advanced Structure 2009.
364. Mokrejs, P.; Krejci, O.; Svoboda, P.; Vasek, V. “Modeling technological conditions for break-
down of waste sheep wool”, Rasayan Journal of Chemistry 4(4), 728–735.
365. Encyclopedia Britannica (Ed.), Wool; 2014.
366. Arunkumar, C.; Megwal, H. S.; Borkar, S. P.; Bhongade, A. L. “Recycling of chicken feather
and wool fibre waste into reforced multilayer composite; A review”, Technical Textile 2013.
Downloaded by [University of Boras] at 03:48 29 January 2015

367. Conzatti, L.; Giunco, F.; Stagnaro, P.; Patrucco, A.; Marano, C.; Rink, M.; Marsano, E. “Com-
posites based on polypropylene and short wool fibres”, Composites Part A: Applied Science
and Manufacturing 2013, 47, 165–171.
368. Blicblau, A. S.; Coutts, R. S. P.;Sims, A. “Novel composites utilizing raw wool and polyester
resin”, Journal of Materials Science Letters, 1997, 16(17), 1417–1419.
369. Conzatti, L.; Giunco, F.; Stagnaro, P.; Capobianco, M.; Castellano, M.; Marsano, E. “Polyester-
based biocomposites containing wool fibres”, Composites Part A: Applied Science and Manu-
facturing 2012, 43(7), 1113–1119.
370. Kock, J. W. Physical and mechanical properties of chicken feather materials. Georgia Institute
of Technology: Atlanta, GA, USA, 2006. Masters thesis.
371. Reddy, N.; Yang, Y. “Structure and properties of chicken feather barbs as natural protein fibers”,
Journal of Polymers and the Environment 2007, 15(2), 81–87.
372. Reddy, N.; Yang, Y. “Light-weight polypropylene composites reinforced with whole chicken
feathers”, Journal of Applied Polymer Science 2010, 116(6), 3668–3675.
373. Cheng, S.; Lau, K. T.; Liu, T.; Zhao, Y.; Lam, P. M.; Yin, Y. “Mechanical and thermal proper-
ties of chicken feather/PLA green composites”, Composites Part B: Engineering 2009, 40(7),
650–654.
374. Huda, S.; Yang, Y. “Feather fiber reinforced light–weight composites with good acoustic prop-
erties”, Journal of Polymers and the Environment 2009, 17(2), 131–142.
375. Zhan, M.; Wool, R. P.; Xiao, J. Q. “Electrical properties of chicken feather reinforced epoxy
composites”, Composites Part A: Applied Science and Manufacturing 2011, 42(3), 229–233.
376. Choudhry, S.; Pandey, B. “Mechanical behavior of polypropylene and human hair fibers and
polypropylene reinforced polymeric composites”, International Journal of Mechanical and
Industrial Engineering 2012, 2(1), 118–121.
377. Seidel, A.; Liivak, O.; Calve, S.; Adaska, J.; Ji, G.; Yang, Z.; Grubb, D.; Zax, D. B.; Jelinski,
L. W. “Regenerated spider silk: Processing, properties, and structure”, Macromolecules 2000,
33(3), 775–780.
378. Shao, Z.; Vollrath, F.; Yang, Y.; Thogersen, H. C. “Structure and behavior of regenerated spider
silk”, Macromolecules 2003, 36(4), 1157–1161.
379. Nazarov, R.; Jin, H. J.; Kaplan, D. L. “Porous 3–D scaffolds from regenerated silk fibroin”,
Biomacromolecules 2004, 5(3), 718–726.
380. Gannon, J. M.; Graveson, I.; Mortimer, S. A. Process for the manufacture of lyocell fibre. U.S.
Patent No. 5,725,821, 1998.
381. Borbely, E. “Lyocell: The new generation of regenerated cellulose.” Acta Polytechnica Hun-
garica, 2008, 5(3), 11–18.
382. Johnson, R. K.; Zink-Sharp, A.; Renneckar, S. H.; Glasser,W. G. “Mechanical properties of
wetlaid lyocell and hybrid fiber-reinforced composites with polypropylene”, Composites Part
A: Applied Science and Manufacturing 2008, 39(3), 470–477.
162 S. K. Ramamoorthy et al.

383. Lenz, J.; Schurtz, W. E. “Properties and structure of solvent-spun and viscose-type fibres in the
swollen state”, Colloid and Polymer Science, 1993, 271(5), 460–468.
384. Zhang, H.; Zhang, H.; Tong, M.; Shao, H.; Hu, X. “Comparison of the structures and properties
of Lyocell fibers from high hemicellulose pulp and high α-cellulose pulp”, Journal of Applied
Polymer Science 2008, 107(1), 636–641.
385. Bourban, C.; Karamuk, E.; De Fondaumiere, M. J.; Ruffieux, K.; Mayer, J.; Wintermantel, E.
“Processing and characterization of a new biodegradable composites made of a PHB/V matrix
and regenerated cellulosic fibers”, Journal of Environmental Polymer Degradation 1997, 5(3),
159–166.
386. Shibata, M.; Oyamada, S.; Kobayashi, S. I.; Yaginuma, D. “Mechanical properties and
biodegradability of green composites based on biodegradable polyesters and lyocell fabric”,
Journal of Applied Polymer Science 2004, 92(6), 3857–3863.
387. Adekunle, K.; Patzelt, C.; Kalanter, A.; Skrifvars, M. “Mechanical and viscoelastic properties of
soybean oil thermoset reinforced with jute fabrics and carded lyocell fiber”, Journal of Applied
Polymer Science 2011, 122(5), 2855–2863.
Downloaded by [University of Boras] at 03:48 29 January 2015

388. Cross, C. F.; Bevan, E. J.; Beadle, C. Improvements in dissolving cellulose and allied compounds.
British Patent: 8,700, 1892.
389. Woodings, C. (Ed.) Fibers, Regenerated Cellulose; Kirk-Othmer Encyclopedia of Chemical
Technology; 2003.
390. Krotscheck, A.; Lackner, K.; Schuster, J.; Wizani, W. Process for the production of viscose
pulp. U.S. Patent No. 5,676,795, 1997.
391. Gubitz, G. M.; Lischnig, T.; Stebbing, D.; Saddler, J. N. “Enzymatic removal of hemicellulose
from dissolving pulps”, Biotechnology Letters 1997, 19(5), 491–495.
392. Paunikallio, T.; Suvanto, M.; Pakkanen, T. T. “Composition, tensile properties, and dispersion of
polypropylene composites reinforced with viscose fibers”, Journal of Applied Polymer Science
2004, 91(4), 2676–2684.
393. Reinhardt, M.; Kaufmann, J.; Kausch, M.; Kroll, L. “PLA-viscose composites with continuous
fibre reinforcement for structural applications”, Procedia Materials Science 2013, 2, 137–143.
394. Huda, M. S.; Drzal, L. T.; Misra, M.; Mohanty, A. K..; Williams, K.; Mielewski, D. F. “A
study on biocomposites from recycled newspaper fiber and poly(lactic acid)”, Industrial &
Engineering Chemistry Research 2005, 44(15), 5593–5601.
395. Sanadi, A. R.; Young, R. A.; Clemons, C.; Rowell, R. M. “Recycled newspaper fibers as reinforc-
ing fillers in thermoplastics: Part I: Analysis of tensile and impact properties in polypropylene”,
Journal of Reinforced Plastics and Composites 1994, 13(1), 54–67.
396. Ramamoorthy, S. K.; Persson, A.; Skrifvars, M. “Reusing textile waste as reinforcements in
composites”, Journal of Applied Polymer Science 2014,
397. Mohanty, A. K.; Misra, M.; Drzal, L. T. “Surface modification of natural fibers and performance
of the resulting biocomposites: An overview”, Composites Interfaces 2001, 8(5), 313–343.
398. Akovali, G. (Ed.) Handbook of Composite Fabrication; iSmithers Rapra Publishing: Akron,
OH, USA, 2001.
399. Tong, L.; Mouritz, A. P.; Bannister, M. K. 3D Fibre Reinforced Polymer Composites; Elsevier:
Amsterdam, 2002.
400. Wambua, P.; Ivens, J.; Verpoest, I. “Natural fibres: Can they replace glass in fibre reinforced
plastics.” Composites Science and Technology 2003, 63(9), 1259–1264.
401. Jones, R. M. Mechanics of Composite Materials; CRC Press: Boca Raton, FL, 1998.
402. Gibson, R. F. Principles of Composite Material Mechanics; CRC Press: Boca Raton, FL, 2011.
403. Gouanve, F.; Marais, S.; Bessadok, A.; Langevin, D.; Metayer, M. “Kinetics of water absorption
in flax and PET fibers”, European Polymer Journal 2007, 43(2), 586–598.

You might also like