0% found this document useful (0 votes)
79 views17 pages

Review Raw Natural Fiber Based Polymer Composites

Uploaded by

waleedsikandry
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
79 views17 pages

Review Raw Natural Fiber Based Polymer Composites

Uploaded by

waleedsikandry
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 17

International Journal of Polymer Analysis and

Characterization

ISSN: 1023-666X (Print) 1563-5341 (Online) Journal homepage: https://www.tandfonline.com/loi/gpac20

Review: Raw Natural Fiber–Based Polymer


Composites

Vijay Kumar Thakur , Manju Kumari Thakur & Raju Kumar Gupta

To cite this article: Vijay Kumar Thakur , Manju Kumari Thakur & Raju Kumar Gupta (2014)
Review: Raw Natural Fiber–Based Polymer Composites, International Journal of Polymer Analysis
and Characterization, 19:3, 256-271, DOI: 10.1080/1023666X.2014.880016

To link to this article: https://doi.org/10.1080/1023666X.2014.880016

Published online: 28 Mar 2014.

Submit your article to this journal

Article views: 3843

View related articles

View Crossmark data

Citing articles: 175 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gpac20
International Journal of Polymer Anal. Charact., 19: 256–271, 2014
Copyright # Taylor & Francis Group, LLC
ISSN: 1023-666X print/1563-5341 online
DOI: 10.1080/1023666X.2014.880016

Review: Raw Natural Fiber–Based Polymer Composites

Vijay Kumar Thakur,1 Manju Kumari Thakur,2 and Raju Kumar Gupta3
1
School of Mechanical and Materials Engineering, Washington State University,
Pullman, Washington, USA
2
Division of Chemistry, Govt. Degree College Sarkaghat,
Himachal Pradesh University, Shimla, India
3
Department of Chemical Engineering, Indian Institute of Technology, Kanpur, India

The effective utilization of raw natural fibers as indispensable component in polymers for developing
novel low-cost eco-friendly composites with properties such as acceptable specific strength, low
density, high toughness, good thermal properties, and biodegradability is one of the most rapidly
emerging fields of research in polymer engineering and science. In fact, raw natural fiber–reinforced
composites are the subject of numerous scientific and research projects, as well as many commercial
programs. Keeping in mind the immense advantages of raw natural fibers, in the present article
we concisely review raw natural fiber=polymer matrix composites with particular focus on their
mechanical properties.

Keywords: Composites; Mechanical properties; Natural fibers; Polymers

INTRODUCTION

In the past few decades, a thrust towards low-cost lightweight materials has resulted in intense
interest in the field of polymers.[1–5] In that time, polymers have emerged as viable alternatives
to some traditional materials, such as metals, due to their inherent properties such as ease of
fabrication, structural control, productivity, easy availability, less physical work, and cost
reduction.[4,6–10] Indeed, due to their outstanding properties polymers have replaced most tra-
ditional materials for a number of applications and are currently playing an important role in
the economies of most nations.[11–14] Depending upon their origin, polymers are generally classi-
fied into two types: natural and synthetic.[15–18] For most applications, especially in the field of
polymer composites, they are frequently broadly classified into thermosetting and thermoplastic
polymers.[19–23]
For successful industrial applications, especially in composites, various kinds of fillers and
fibers are incorporated into the polymers to meet the high strength=high modulus requirements
of different applications ranging from electronic materials to load-bearing structural materi-
als.[24–26] Due to the intensified use of polymer composite materials in a number of industries,

Submitted 14 December 2013; accepted 24 December 2013.


Correspondence: Vijay Kumar Thakur, School of Mechanical and Materials Engineering, Washington State
University, Pullman, WA 99614, USA. E-mail: [email protected]
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/gpac.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 257

combined with rising environmental awareness throughout the globe, greater attention is being
paid to the incorporation of biopolymer-based materials as one component in polymer matrices
instead of petrochemical-based materials.[27–29] Among various biopolymer-based materials,
natural fibers are being envisioned as potential environmentally friendly materials that can
become indispensable components of polymers for various applications.[30–32] Different kinds
of natural fibers (e.g., Hibiscus sabdariffa, henequen, Grewia optiva, sisal, pine needles, flax,
jute, kenaf, coir, and many others), due to their renewable nature, offer a number of advantages
over synthetic fibers such as glass fibers, aramid fibers, and carbon fiber.[33–35] Some of the
unique properties of natural fibers include: low cost, biodegradability, recyclability, acceptable
specific strength, ease of separation, low density, high toughness, good thermal properties,
reduced tool wear, nonirritation to the skin, and enhanced energy recovery.[36,37] Compared
to traditional synthetic fibers, natural fibers are very cheap and promising. Figure 1 shows the
cost comparison between glass and natural fibers.[38]
Most natural fibers are primarily composed of cellulose, hemicellulose, and lignin.[36–39] The
effective utilization of natural fibers as reinforcement fibers in diverse polymer matrices (natur-
al=synthetic) provides a number of positive benefits, including environmental benefits.[40,41]
Natural fiber–reinforced composites are recently finding their use in a number of commercial
applications such as deck surfaces, door components, windows, sports facilities, packaging
and automotive industries, and furniture.[42,43] Along with the nature of the fibers, the properties
of the resulting composites are also influenced by the type of polymer matrix used and the
amount and dimensions of the fiber.
A number of review articles have been published on the effect of surface modification of
natural fibers on the properties of polymer composites.[44–46] However, limited information is
available in the existing literature that summarizes solely the processing and properties of raw
natural fiber–reinforced polymer composites. Thus the present article aims to concisely review
the reported works on raw natural fiber–reinforced polymer composites that include synthesis,
processing, and mechanical characterization.

FIGURE 1 Cost per weight (a) and cost per unit length (b) comparison between glass and natural fibers (from Dittenber
and GangaRao[38]). # Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.
258 V. K. THAKUR ET AL.

STRUCTURE AND PROPERTIES OF NATURAL FIBERS

Different kinds of natural fibers are in use all around the globe.[39,44–46] These natural fibers,
found directly or indirectly in everyday life, have been used by common people for a number
of applications,[47,48] One of the biggest assets of these natural fibers is that at the end of their
life, they can be degraded by composting or other methods.[47,49] Natural fibers are generally
subdivided into different types (Figure 2) depending upon their origin (e.g., whether they are
procured from minerals, animals, or plants).[50] Among these fibers, plant–based fibers are most
frequently used in a number of applications and are of very high commercial importance.[51]
Natural plant fibers include different classes of fibers such as bast, straw, seed, grass, leaf,
and wood fibers.[52]
Almost all natural fibers, except for cotton, are primarily composed of cellulose, hemicellu-
lose, lignin, waxes, and several water-soluble compounds.[53] The amount of cellulose in a given
fiber determines the strength and stiffness of the fibers, which is provided by hydrogen bonds
and other linkages in the cellulose.[54] A number of factors affect the overall properties of a parti-
cular fiber, starting from different stages of production up to final processing.[44–55] Table I sum-
marizes the factors that play a key role in affecting the overall quality during stages of natural
fiber processing.[38] The amount of hemicellulose controls moisture absorption, biodegradation,
and thermal degradation properties of all the natural cellulosic fibers,[44–55] while lignin has been
reported to be the most thermally stable among all three constituents, but is very sensitive to UV
radiation and is responsible for the degradation of fibers under ultraviolet radiation.[44–50]
Among the constituents present in any cellulosic fiber, cellulose is the most essential compo-
nent.[44–50,56] Figure 3 shows the chemical structure of cellulose that is present in all fibers. Dif-
ferent properties of natural fibers, especially mechanical properties, depend upon the type of
cellulose present in the fiber.[50] The mechanical properties depend upon the degree of polymer-
ization of the cellulose, total cellulose content in the fibers, and the microfibril angles.[44–50,57–61]

FIGURE 2 Classification of natural fibers (from Akil et al.[50]). # Elsevier. Reproduced by permission of Elsevier.
Permission to reuse must be obtained from the rightsholder.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 259

TABLE I
Factors affecting fiber quality at various stages of natural fiber production
(from Dittenber and GangaRao[38])

Stage Factors affecting fiber quality

Plant growth Local climate


Crop cultivation
Crop location
Fiber location in plant
Harvesting stage Fiber ripeness, which effects:
Cell wall thickness
Coarseness of fibers
Adherence between fibers and surrounding structure
Fiber extraction stage Decortication process
Type of retting method
Supply stage Transportation conditions
Storage conditions
Age of fiber

# Elsevier. Reproduced by permission of Elsevier. Permission to reuse must


be obtained from the rightsholder.

FIGURE 3 Chemical structure of cellulose (from Akil et al.[50]). # Elsevier. Reproduced by permission of Elsevier.
Permission to reuse must be obtained from the rightsholder.

The cell wall in cellulosic fibers is primarily composed of a hollow tube with one primary cell
wall, three secondary cell walls, and a lumen.[57–61] Each layer in a natural fiber contains cellu-
lose that is embedded in a matrix of hemicellulose and lignin.[57,58] It has been well documented
in the existing literature that natural cellulosic fibers having higher content of cellulose and
degree of polymerization and a smaller microfibrillar angle have higher tensile strength.[44–61]

NATURAL FIBER–REINFORCED POLYMER COMPOSITES

Different researchers have defined composites in different ways. In a broader manner a


composite is defined as: Two or more dissimilar materials which when combined are stronger
than the individual materials with entirely different properties from those of the individual
260 V. K. THAKUR ET AL.

components.[57–61] The matrix, which is the polymer in composite materials, plays an imperative
role in determining the overall properties of a composite.[61,62] Among various polymer compo-
sites, there has been keen interest in the synthesis, processing, and applications of natural fiber–
reinforced composites during the past few years.[63,64] Composites materials were often
employed by the people of earlier civilizations for a number of household applications.[61]
The use of natural fibers as one component in composite materials was reported as far back
as nearly 3000 years ago in ancient Egypt.[57,68] For those composite materials, Egyptians mixed
straw and clay in the fabrication of building walls. The use of natural fibers as reinforcement in
polymers has recently seen a resurgence due to rising environmental awareness.[62–65] Different
kinds of natural fibers having different origins have been reported to be potential alternatives to
traditional synthetic fibers in composite materials for a number of applications, including
structural applications.[66,67] Different types of polymer matrices, from synthetic to natural, have
been reported to be frequently used in a number of automotive applications.[57–59,68]
Polymer composites can be divided into two different types, namely totally renewable com-
posites and partly renewable composites, depending upon the nature and origin of the reinforce-
ment as well as the polymer matrix.[57–61,69] In totally renewable polymer composite materials
both the reinforcing materials and the matrix material comes from biorenewable resources.[57–61]
In partly renewable composites only one component is from biorenewable resources such as a
polymer matrix from biorenewable resources and a reinforcement from nonrenewable resources
or a polymer matrix from synthetic resources and a reinforcement from biorenewable resources.
Polymer composites using synthetic and natural polymer matrices have been covered in a
number of reviews.[44–46,50,57–61] Some of the commercially important polymer matrices include
phenolic resins, amino resins, polythene, polypropylene, polystyrene, polyester, epoxy, starch,
and polylactic acid. In the following section we discuss some of the industrially important
raw natural fiber–reinforced polymer composites.

Raw Natural Fiber=Poly (Butylene Succinate) Composites

Poly (butylene succinate) (PBS), having a low melting temperature (114 C), is an emerging
polymeric matrix generally synthesized by the polycondensation reaction of 1, 4-butanediol
and succinic acid.[70] PBS is biodegradable and deteriorates in moist soil, compost, fresh water
with activated sludge, and seawater. It is recently being explored in polymer composite applica-
tions due to its eco-friendliness and good biodegradability. Frolini et al.[70] reported their study
on the preparation of polymer composites using a traditional thermo-pressed molding technique.
They used different kinds of fibers, namely coconut, sugarcane bagasse, curaua, and sisal, as
reinforcing materials in a PBS matrix. These natural fibers were thoroughly characterized by
different characterization techniques in terms of their chemical composition, thermal stability
(thermogravimetric analysis; TGA), crystallinity (X-ray diffraction; XRD), and surface mor-
phology (scanning electron microscopy; SEM). Figure 4 shows cross section SEM micrographs
of (a) sugarcane bagasse, (b) coconut, (c) curaua, and (d) sisal. Along with these natural fibers,
the resulting composites were also characterized by mechanical tests such as Izod impact
strength, flexural resistance, and thermal stability (TGA). Thermal stability of the PBS and
fiber-reinforced composites using 20% fiber loadings was studied, and each fiber was found to
exhibit a similar influence on the thermal behavior of the polymeric matrix. The small difference
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 261

FIGURE 4 Cross sections of the fibers. SEM micrographs of (a) sugarcane bagasse, (b) coconut, (c) curaua, and (d) sisal
(from Frollini et al.[70]). # Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from
the rightsholder.

in the thermal stability above 400 C was ascribed to the different content of lignin present in the
polymer matrix. Figure 5 shows the impact and flexural strength results of different fiber-
reinforced polymer composites. It was observed that among all the fibers, sisal and curaua fibers
demonstrated a huge potential as reinforcement in PBS matrix. This behavior was attributed to

FIGURE 5 (a) Flexural modulus and (b) impact resistance results of different fiber-reinforced polymer composites (from
Frollini et al.[70]). # Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the
rightsholder.
262 V. K. THAKUR ET AL.

the superior chemical compatibility of these fibers with the aliphatic matrix along with their
surface morphology.
Short silk fiber–reinforced PBS composites were prepared by Lee et al.[71] using the
compression molding method, and different mechanical=thermal properties were studied (see
Figure 6). It was observed that even without any surface modification the chopped silk fibers
significantly improved the mechanical properties of PBS due to better interfacial adhesion
between the silk fiber and the matrix. The thermal stability of silk=PBS biocomposites was
found to be intermediate between that of the PBS matrix and that of the silk fiber.[71]

Raw Natural Fiber=Thermoplastic Composites

Natural fiber–reinforced thermoplastic composites have been studied by a number of researchers


for automotive applications and have been comprehensively reviewed, with particular emphasis
on the surface modifications of natural fibers.[44–45,57–61] Among various thermoplastic polymers,
polypropylene is perhaps one of the most widely used because of its distinct properties such as
dimensional stability, high heat distortion temperature, flame resistance, and transparency.[60]
Shubhra et al.[72] in their excellent review article on propylene composites extensively covered
the mechanical properties of different natural fiber–reinforced composites. Figure 7 shows the
tensile, bending, and impact strength results of different natural fiber–reinforced PP composites.
It was observed that flax fiber–reinforced composites demonstrated the highest mechanical
properties among all natural fibers; this behavior was attributed to the distinctive surface
morphology of the flax fibers.[72]

Raw Natural Fiber=Thermosetting Composites

Natural fiber–reinforced polymer composites with different polymer matrices such as phenol-
formaldehyde and urea formaldehyde have been prepared for structural applications.[73–78]

FIGURE 6 (a) Tensile strength and (b) flexural strength results of PBS and five silk=PBS biocomposites with various
silk fiber contents (from Lee et al.[71]). # Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be
obtained from the rightsholder.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 263

FIGURE 7 (a) Tensile strength, (b) bending strength, and (c) impact strength of different fiber-reinforced polypropylene
composites (from Shubhra et al.[72]). # Sage Publications. Reproduced by permission of Sage Publications. Permission
to reuse must be obtained from the rightsholder.
264 V. K. THAKUR ET AL.

Recently, Stoeckel et al.[79] reviewed the mechanical properties of different kinds of polymers
(e.g., phenol formaldehyde, PF; urea formaldehyde, UF; phenol resorcinol formaldehyde,
PRF; polymeric diphenylmethane diisocyanate, pMDI; polyurethane, PUR; polyvinyl acetate,
PVAc; resorcinol formaldehyde, RF; urea melamine formaldehyde, UMF) that were used as
adhesive for bonding wood. Most thermosetting polymers and their composites (micro=nano)
are processed by simple processing techniques such as hand lay-up, resin transfer, spraying,
compression, injection, and pressure bag molding operations. Figure 8 shows the elastic modu-
lus of different adhesives designed for a number of applications. Mechanical properties of dif-
ferent adhesives using nanoindentation tests were also summarized.[79]
Figure 9 shows the mechanical propertiesof different adhesives. Singha and Thakur[80–87] also
carried out an extensive study of the preparation of different kinds of natural fiber–reinforced
phenol-formaldehyde and urea formaldehyde composites. For all the natural fiber composites
(e.g., using pine needles, Hibiscus sabdariffa, Grewia optiva, and Saccaharum cilliare) both
the phenolic and amino resin polymer matrices were found to strongly bind with these fibers.
The stress was successfully transferred from one fiber to another and resulted in the formation

FIGURE 8 Elastic modulus of wood adhesives designed for panel production (right) and for solid wood bonding (left)
measured by means of nanoindentation inside bondlines (top), on adhesive films (middle), and by using macroscopic
testing methods (bottom) (from Stoeckel et al.[79]). # Elsevier. Reproduced by permission of Elsevier. Permission to
reuse must be obtained from the rightsholder.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 265

FIGURE 9 Mechanical properties of wood adhesives measured by means of nanoindentation. Elastic modulus used as a
proxy for reduced modulus Er (from Stoeckel et al.[79]). # Elsevier. Reproduced by permission of Elsevier. Permission
to reuse must be obtained from the rightsholder.

of a consolidated composite structure. They concluded that the mechanical properties of these
matrices and the quality of their adhesion to the fibers control the load transfer ability between
fibers. Overall, the matrix was found to play a significant role in the performance of these
composite materials. Effect of fiber dimensions and fiber loading was also studied in detail
for phenol-formaldehyde and urea formaldehyde composites.[80–87]

Raw Vegetable Fiber=Binder-Free Composites

Green biobased polymer composites without the help of any external binder were prepared using
fibers from the reed-like plant Typha sp.[88] Different mechanical properties such as flexural
strength, flexural modulus of elasticity, and surface textural properties of these binder-free veg-
etable fiberboards based on cattails were also studied and compared with those of other natural
fiber–reinforced composites. Table II shows the comparative flexural strength results for
binder-free fiber composites and other natural fiber composites. Along with mechanical proper-
ties, the surface properties of the binder-free composites were extraordinary good. It was con-
cluded that the binder-free cattails-based composites exhibit a high potential in the automotive
and furniture industries.[88]

Nanocellulose=Natural Rubber Composites

Along with different synthetic and natural polymers matrices, natural rubber (NR), obtained from
latex, has its own importance in a number of applications, especially in packaging. NR is a
266 V. K. THAKUR ET AL.

TABLE II
Surface roughness expressed in roughness parameters Ra and Rt, flexural modulus of elasticity (MOE), and
flexural strength of various natural fiber composites bonded with phenolic (PF) binder resin compared to
binder-free produced Typha-based panels (all values in the table were obtained from panels of the same
density) (q ¼ 1 g cm3; from Wuzella et al.[88])

Resin Flexural Flexural


Fiber mixture Binder content (%) Ra Rt MOE (MPa) strength (MPa)

Kenaf (100%) PF 15 25.27  2.18 289.04  35.73 932.5  1190 22.5  14.0
PF 30 24.00  4.88 300.23  74.51 4343  1064.5 53.8  12.1
Flax (100%) PF 15 12.89  1.18 136.45  11.07 574  230 17.4  3.2
PF 30 13.37  0.36 159.49  12.00 4839  886.5 47.5  13.5
Hemp (100%) PF 15 20.90  2.19 259.77  50.24 2276.5  230 37.9  9.0
PF 30 20.72  1.63 260.20  39.80 6186.5  500 73.3  4.5
Coco (100%) PF 15 32.57  2.87 351.82  41.85 n.d. n.d.
PF 30 34.2  1.94 444.64  11.98 2049.5  696.5 44.4  9.1
Kenaf=flax (50=50) PF 15 22.08  2.02 294.70  39.15 1488.5  792.5 29.3  5.3
PF 30 22.85  2.30 256.44  45.71 5877.5  884.5 50.3  13.1
Hemp=flax (50=50) PF 15 15.93  0.98 175.98  8.37 1420  518.5 20.4  7.4
PF 30 12.83  1.81 174.99  14.25 5524  601 67.1  7.3
Wood=flax (75=25) PF 15 17.86  2.27 237.80  54.57 n.d. n.d.
PF 30 17.14  1.61 219.25  24.95 1202  149 15.1  0.9
Typha 0 4.96  0.51 100.98  10.00 3100  92 21  2

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

polymer of cis-1, 4-polyisoprene and generally contains more than 90% cis-1, 4-polyisoprene.[89]
The rest of the content is less than 10% and includes proteins and carbohydrates. Abraham et al.[89]
studied different properties including degradation and tensile strength of nanocellulose-reinforced
natural rubber composites. Nanocellulose in this study was procured from raw jute fiber using the
steam explosion technique; different techniques were used to characterize the nanocomposite
samples. The mechanical properties of the non-cross-linked composites were found to be lower
than those of their cross-linked counterparts. Table III shows the mechanical properties of the
NR=nanocellulose composites.

TABLE III
Mechanical properties of the composites of NR=nanocellulose with and without cross-linking (from
Abraham et al.[89])

rR (MPa) E (MPa) eR (%)

Material Non-cross-linked Cross-linked Non-cross-linked Cross-linked Non-cross-linked Cross-linked

Natural rubber 0.7  0.06 1.6  0.20 0.85  0.10 1.3  0.15 816  25 912  19
2.5% composite 3.1  0.1 5.2  0.15 2.7  0.12 4.2  0.25 477  29 576  23
5% composite 4.2  0.17 6.8  0.18 3.8  0.20 6.3  0.22 325  24 413  22
7.5% composite 6.6  0.20 9.8  0.24 5.7  0.18 8.1  0.35 205  8 275  12
10% composite 9.1  0.24 12.2  0.36 7.2  0.21 9.6  0.31 101  5 144  5

# Elsevier. Reproduced by permission of Elsevier. Permission to reuse must be obtained from the rightsholder.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 267

FIGURE 10 Schematic diagram of the VARI process (from Lee et al.[90]). # American Chemical Society. Reproduced
by permission of the American Chemical Society. Permission to reuse must be obtained from the rightsholder.

Bacterial Cellulose and Nanofibrillated Cellulose=Epoxy Nanocomposites

Bacterial and nanofibrillated cellulose has been used as reinforcement in a number of polymer
matrices and was found to affect the mechanical properties of the polymer composite materi-
als.[90] The surface and bulk properties of two kinds of nanocellulose, nanofibrillated cellulose
(NFC) and bacterial cellulose (BC), were studied by Lee et al.[90] It was observed that the surface
tension as well as thermal degradation of the BC was higher than that of NFC; this behavior was
attributed to the higher crystallinity of BC determined using XRD. The effect of different NFC
and BC on the mechanical properties of epoxy polymer matrix was also studied. The nanocom-
posites were prepared using the vacuum-assisted resin infusion (VARI) technique (Figure 10).
However, in the nanocomposites no significant difference was observed in terms of the reinfor-
cing ability of NFC and BC in a polymer matrix. The BC-reinforced nanocomposites showed
slightly higher tensile strength than NFC composites by approximately 6%. The slightly higher
tensile properties (e.g., elongation at break) of BC nanocomposites than NFC composites was
attributed to the higher strain-to-failure of the BC paper. It was concluded that that both NFC
and BC exhibit strong potential to serve as novel reinforcement in number of polymer
matrices.[90]

CONCLUSIONS

Among various natural materials, natural fibers are one of the potential low-cost, environmentally
friendly materials that can be an indispensable component of polymer composite applications. Dif-
ferent kinds of natural fibers, due to their biorenewable nature and inherent eco-friendly character-
istics, offer a number of advantages over synthetic fibers such as glass fibers, aramid fibers, and
carbon fibers. The effective utilization of different kinds of such natural fibers has been discussed
in detail in the present article. Recently, in Europe some natural fibers are being used as one of the
components in industrial applications and their use as green reinforcement in polymers has proved
encouraging. The structural aspects and properties of various natural fiber=polymer matrices and
the polymer composites as discussed in this review article and appropriate knowledge might be
268 V. K. THAKUR ET AL.

used in the near future for the commercialization of polymer composite products for an assortment
of applications.

REFERENCES

1. Khalil, H. P. S. A., M. Jawaid, P. Firoozian, M. Amjad, Z. Zainudin, and M. T. Paridah. 2013. Tensile, electrical
conductivity, and morphological properties of carbon black-filled epoxy composites. Int. J. Polym. Anal. Charact.
18(5): 329–338.
2. Chandramohan, A., and M. Alagar,. 2013. Preparation and characterization of cyclohexyl moiety toughened
POSS-reinforced epoxy nanocomposites. Int. J. Polym. Anal. Charact. 18(1): 73–81.
3. Mallakpour, S., and A. Zadehnazari. 2012. New organosoluble, thermally stable, and nanostructured poly (amide-
imide)s with dopamine pendant groups: Microwave-assisted synthesis and characterization. Int. J. Polym. Anal.
Charact. 17(6): 408–416.
4. Wanasekara, N. D., V. Chalivendra, and P. Calvert. 2012. Effect of accelerated ultraviolet and thermal exposure on
nanoscale mechanical properties of nylon fibers. Polym. Eng. Sci. 52(11): 2482–2488.
5. Pinto, M. A., V. B. Chalivendra, Y. K. Kim, and A. F. Lewis. 2014. Evaluation of surface treatment and fabrication
methods for jute fiber=epoxy laminar composites. Polym. Compos. 35(2): 310–317. doi:10.1002=pc.22663.
6. Ioan, S., A. M. Necula, I. Stoica, N. Olaru, L. Olaru, and G. E. Ioanid. 2010. Surface properties of cellulose acetate.
High Perform. Polym. 22(5): 598–608.
7. Široká, B., J. Široký, and T. Bechtold. 2011. Application of ATR-FT-IR single-fiber analysis for the identification of
a foreign polymer in textile matrix. Int. J. Polym. Anal. Charact. 16(4): 259–268.
8. Jawaid, M., O. Y. Alothman, M. T. Paridah, and H. P. S. A. Khalil. 2013. Effect of fiber treatment on dimensional
stability and chemical resistance properties of hybrid composites. Int. J. Polym. Anal. Charact. 18(8): 608–616.
9. Spiridon, I., C. A. Teaca, R. Bodirlau, and M. Bercea. 2013. Behavior of cellulose reinforced cross-linked starch
composite films made with tartaric acid modified starch microparticles. J. Polym. Environ. 21(2): 431–440.
10. Vadlamani, V. K., V. B. Chalivendra, A. Shukla, and S. Yang. 2012. Sensing of damage in carbon nanotubes and
carbon black-embedded epoxy under tensile loading. Polym. Compos. 33(10): 1809–1815.
11. Thakur, V. K., M.-F. Lin, E. J. Tan, and P. S. Lee. 2012. Green aqueous modification of fluoropolymers for energy
storage applications. J. Mater. Chem. 22(13): 5951–5959.
12. Bodirlau, R., C. A. Teaca, and I. Spiridon. 2013. Influence of natural fillers on the properties of starch-based
biocomposite films. Composites Part B Eng. 44(1): 575–583.
13. Thakur, V. K., J. Yan, M.-F. Lin, C. Zhi, D. Golberg, Y. Bando, R. Sim, and P. S. Lee. 2012. Novel polymer
nanocomposites from bioinspired green aqueous functionalization of BNNTS. Polym. Chem. 3: 962–969.
14. Spiridon, I., C. A. Teaca, and R. Bodirlau. 2011. Structural changes evidenced by FTIR spectroscopy in cellulosic
materials after pre-treatment with ionic liquid and enzymatic hydrolysis. Bioresources 6(1): 400–413.
15. Nervo, R., O. Konovalov, and M. Rinaud. 2012. Chitosan-behenic acid monolayer interaction at the air-water
interface: Characterization of the adsorbed polymer layers by X-ray reflectivity. Int. J. Polym. Anal. Character.
17(1): 11–20.
16. Buruiana, L. I., E. Avram, A. Popa, I. Stoica, and S. Ioan. 2013. Influence of triphenylphosphonium pendant
groups on the rheological and morphological properties of new quaternized polysulfone. J. Appl. Polym. Sci.
129(4): 1752–1762.
17. Dobos, A. M., M. D. Onofrei, I. Stoica, N. Olaru, L. Olaru, and S. Ioan. 2012. Rheological properties and micro-
structures of cellulose acetate phthalate=hydroxypropyl cellulose blends. Polym. Compos. 33(11): 2072–2083.
18. Dobos, A. M., I. Stoica, N. Olaru, L. Olaru, E. G. Ioanid, and S. Ioan. 2012. Surface properties and biocompatibility
of cellulose acetates. J. Appl. Polym. Sci. 125(4): 2521–2528.
19. Kommula, V. P., K. O. Reddy, M. Shukla, T. Marwala, and A. V. Rajulu. 2013. Physico-chemical, tensile, and ther-
mal characterization of Napier grass (native African) fiber strands. Int. J. Polym. Anal. Charact. 18(4): 303–314.
20. Necula, A. M., I. Stoica, N. Olaru, F. Doroftei, and S. Ioan. 2011. Silver nanoparticles in cellulose acetate polymers:
Rheological and morphological properties. J. Macromol. Sci. Part B Phys. 50(4): 639–651.
21. Maheswari, C. U., K. O. Reddy, E. Muzenda, M. Shukla, and A. V. Rajulu. 2013. Mechanical properties and
chemical resistance of short tamarind fiber=unsaturated polyester composites: Influence of fiber modification and
fiber content. Int. J. Polym. Anal. Character. 18(6): 439–450.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 269

22. Darie, R. N., R. Bodirlau, C. A. Teaca, J. Macyszyn, M. Kozlowski, and I. Spiridon. 2013. Influence of accelerated
weathering on the properties of polypropylene=polylactic acid=eucalyptus wood composites. Int. J. Polym. Anal.
Charact. 18(4): 315–327.
23. Zhang, Z. Y., Z. Z. Dong, and C. F. Xiao. 2013. Estimation of ethyl cellulose layer thickness of coated polyethylene
oxide particles using differential scanning calorimetry. Int. J. Polym. Anal. Charact. 18(1): 25–29.
24. Bodirlau, R., C. A. Teaca, D. Rosu, L. Rosu, C. D. Varganici, and A. Coroaba. 2013. Physico-chemical properties
investigation of softwood surface after treatment with organic anhydride. Cent. Eur. J. Chem. 11(12): 2098–2106.
25. Spiridon, I., C. A. Teaca, and R. Bodirlau. 2011. Preparation and characterization of adipic acid-modified starch
microparticles=plasticized starch composite films reinforced by lignin. J. Mater. Sci. 46(10): 3241–3251.
26. Singha, A. S., V. K. Thakur, I. K. Mehta, A. Shama, A. J. Khanna, R. K. Rana, and A. K. Rana. 2009. Surface
modified Hibiscus sabdariffa fibers: Physico-chemical, thermal and morphological properties evaluation. Int. J.
Polym. Anal. Charact. 14(8): 695–711.
27. Maheswari, C. U., K. O. Reddy, E. Muzenda, M. Shukla, and A. V. Rajulu. 2013. Mechanical properties and
chemical resistance of short tamarind fiber=unsaturated polyester composites: Influence of fiber modification and
fiber content. Int. J. Polym. Anal. Character. 18(7): 520–533.
28. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2014. Pressure induced synthesis of EA grafted Saccaharum cilliare
fibers. Int. J. Polym. Mater. Polym. Biomater. 63(1): 17–22.
29. Thakur, V. K., M. K. Thakur, and R. K. Gupta. 2014. Rapid synthesis of graft copolymers from natural cellulose
fibers. Carbohydr. Polym. 98(1): 820–828.
30. Thakur, V. K., M. K. Thakur, and R. K. Gupta. 2013. Graft copolymers from natural polymers using free radical
polymerization. Int. J. Polym. Anal. Character. 18(7): 495–503.
31. Thakur, V. K., M. K. Thakur, and R. K. Gupta. 2013. Synthesis of lignocellulosic polymer with improved chemical
resistance through free radical polymerization. Int. J. Biol. Macromol. 61: 121–126.
32. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2013. Free radical-induced graft copolymerization onto natural
fibers. Int. J. Polym. Anal. Charact. 18(6): 430–438.
33. Thakur, V. K., A. S. Singha, I. K. Mehta, R. P. Nagarajarao, and L. P. Yang. 2010. Silane functionalization of
Saccaharum cilliare fibers: Thermal, morphological, and physicochemical study. Int. J. Polym. Anal. Charact.
15(7): 397–414.
34. Fernandesa, E. M., R. A. Piresa, J. F. Manoa, and R. L. Reisa. 2013. Bionanocomposites from lignocellulosic
resources: Properties, applications and future trends for their use in the biomedical field. Prog. Polym. Sci. 38:
1415–441.
35. Shah, D. U. 2013. Developing plant fibre composites for structural applications by optimising composite parameters:
A critical review. J. Mater. Sci. 48: 6083–6107.
36. Avella, M., A. Buzarovska, M. E. Errico, G. Gentile, and A. Grozdanov. 2009. Eco-challenges of bio-based polymer
composites. Materials 2: 911–925.
37. Koronis, G., A. Silva, and M. Fontul. 2013. Green composites: A review of adequate materials for automotive
applications. Composites: Part B 44: 120–127.
38. Dittenber, D. B., and H. V. S. GangaRao. 2012. Critical review of recent publications on use of natural composites in
infrastructure. Composites: Part A 43: 1419–1429.
39. Thakur, V. K., M. K. Thakur, and R. K. Gupta. 2013. Graft copolymers from cellulose: Synthesis, characterization
and evaluation. Carbohydr. Polym. 97(1): 18–25.
40. Atluri, R. P. V., K. M. Rao, and A. V. S. S. K. S. Gupta. 2013. Experimental investigation of mechanical properties
of golden cane fiber-reinforced polyester composites. Int. J. Polym. Anal. Charact. 18(1): 30–39.
41. Narendar, R., and K. P. Dasan. 2013. Effect of chemical treatment on the mechanical and water absorption properties
of coir pith=nylon=epoxy sandwich composites. Int. J. Polym. Anal. Charact. 18(5): 369–376.
42. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2013. Ecofriendly biocomposites from natural fibers: Mechanical
and weathering study. Int. J. Polym. Anal. Charact. 18(1): 64–72.
43. Zhao, J., C. Xiao, and N. Xu. 2012. Surface and physical mechanical properties of polypropylene=poly (butyl
methacrylate-co-hydroxyethyl methacrylate) blend fiber. Int. J. Polym. Anal. Charact. 17(8): 557–567.
44. Xie, Y., C. A. S. Hill, Z. Xiao, H. Militz, and C. Mai. Silane coupling agents used for natural fiber=polymer
composites: A review. Composites: Part A 41: 806–819.
45. Mukhopadhyay, S., and R. Fangueiro. 2009. Physical modification of natural fibers and thermoplastic films for
composites. J. Thermoplast. Compos. Mater. 22: 135–162.
270 V. K. THAKUR ET AL.

46. John, M. J., and R. D. Anandjiwala. 2008. Recent developments in chemical modification and characterization of
natural fiber-reinforced composites. Polym. Compos. 29(2): 187–207.
47. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2012. In-air graft copolymerization of ethyl acrylate onto natural
cellulosic polymers. Int. J. Polym. Anal. Charact. 17(1): 48–60.
48. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2012. Modification of natural biomass by graft copolymerization.
Int. J. Polym. Anal. Charact. 17(7): 547–555.
49. Thakur, V. K., A. S. Singha, and B. N. Misra. 2011. Graft copolymerization of methyl methacrylate onto cellulosic
biofibers. J. Appl. Polym. Sci. 122(1): 532–544.
50. Akil, H. M., M. F. Omar, A. A. M. Mazuki, S. Safiee, Z. A. M. Ishak, and A. A. Bakar. 2011. Kenaf fiber reinforced
composites: A review. Mater. Des. 32(8–9): 4107–4121.
51. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2012. Surface modification of natural polymers to impart low water
absorbency. Int. J. Polym. Anal. Character. 17(2): 133–143.
52. Thakur, V. K., and A. S. Singha. 2011. Rapid synthesis, characterization, and physicochemical analysis of
biopolymer-based graft copolymers. Int. J. Polym. Anal. Charact. 16(3): 153–164.
53. Singha, A. S., and V. K. Thakur. 2008. Saccaharum cilliare fiber reinforced polymer composites. E-J. Chem. 5(4):
782–791.
54. Thakur, V. K., and A. S. Singha. 2011. Rapid synthesis, characterization, and physicochemical analysis of
biopolymer-based graft copolymers. Int. J. Polym. Anal. Charact. 16(3): 153–164.
55. Niimura, N. 2012. Structural study of a Japanese lacquer film with thermogravimetry-linked scan mass spectrometry.
Int. J. Polym. Anal. Charact. 17(7): 540–546.
56. Goud, G., and R. N. Rao. 2011. The effect of alkali treatment on dielectric properties of Roystonea regia=epoxy
composites. Int. J. Polym. Anal. Charact. 16(4): 239–250.
57. Thakur, V. K. (ed.). 2013. Green Composites from Natural Resources. Boca Raton, Fla.: CRC=Taylor & Francis.
58. Thakur, V. K., and A. S. Singha. (ed.). 2013. Biomass Based Biocomposites. Shawbury, UK: Smithers Rapra
Technology.
59. Singha, A. S., and V. K. Thakur. (ed.). 2013. Green Polymer Materials. Houston, Tex.: Studium Press.
60. Thakur, V. K., and A. S. Singha. (ed.). 2012. Nanotechnology in Polymers. Houston, Tex.: Studium Press.
61. Thakur, V. K., A. S. Singha, and M. K. Thakur. (ed.). 2011. Green Composites from Natural Cellulosic Fibers.
Saarbrücken, Germany: Lambert Academic Publishing.
62. Ramanaiah, K., A. V. Ratna Prasad, and K. H. Chandra Reddy. 2011. Thermal and mechanical properties of
sansevieria green fiber reinforcement. Int. J. Polym. Anal. Character. 16(8): 602–608.
63. Habib, F. N., S. S. Kordestani, F. Afshar-Taromi, and Z. Shariatinia. 2011. A novel topical tissue adhesive composed
of urethane prepolymer modified with chitosan. Int. J. Polym. Anal. Charact. 16(8): 609–618.
64. Singha, A. S., and V. K. Thakur. 2008. Effect of fibre loading on urea-formaldehyde matrix based green composites.
Iran. Polym. J. 17(11): 861–873.
65. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2012. Green composites from natural fibers: Mechanical and
chemical aging properties. Int. J. Polym. Anal. Charact. 17(6): 401–407.
66. Pinto, M. A., V. B. Chalivendra, Y. K. Kim, and A. F. Lewis. 2013. Effect of surface treatment and Z-axis reinforce-
ment on the interlaminar fracture of jute=epoxy laminated composites. Eng. Fract. Mech. 114: 104–114.
doi:10.1016=j.engfracmech.2013.10.015
67. Singha, A. S., and V. K. Thakur. 2008. Synthesis and characterization of pine needles reinforced RF matrix based
biocomposites. E-J. Chem. 5(1): 1055–1062.
68. Singha, A. S., and V. K. Thakur. 2009. Grewia optiva fiber reinforced novel, low cost polymer composites. E-J.
Chem. 6(1): 71–76.
69. Ramanaiah, K., A. V. Ratna Prasad, and K. H. Chandra Reddy. 2011. Mechanical properties and thermal conductivity
of Typha angustifolia natural fiber–reinforced polyester composites. Int. J. Polym. Anal. Charact. 16(7): 496–503.
70. Frollini, E., N. Bartolucci, L. Sisti, and A. Celli. 2013. Poly (butylene succinate) reinforced with different lignocel-
lulosic fibers. Ind. Crops Prod. 45: 160–169.
71. Lee, S. M., D. Cho, W. H. Park, S. G. Lee, S. O. Han, and L. T. Drzal. 2005. Novel silk=poly (butylene succinate)
biocomposites: The effect of short fibre content on their mechanical and thermal properties. Compos. Sci. Technol.
65(3–4): 647–657.
72. Shubhra, Q. T. H., A. K. M. M. Alam, and M. A. Quaiyyum. 2011. Mechanical properties of polypropylene
composites: A review. J. Thermoplast. Compos. Mater. 26(3): 362–391.
REVIEW: NATURAL FIBER–BASED POLYMER COMPOSITES 271

73. Thakur, V. K., and A. S. Singha. 2010. Mechanical and water absorption properties of natural fibers=polymer
biocomposites. Polym.-Plast. Technol. Eng. 49(7): 694–700.
74. Singha, A. S., and V. K. Thakur. 2010. Fabrication and characterization of H. sabdariffa fiber-reinforced green
polymer composites. Polym.-Plast. Technol. Eng. 48(4): 482–487.
75. Thakur, V. K., A. S. Singha, and M. K. Thakur. 2013. Fabrication and physico-chemical properties of high-
performance pine needles=green polymer composites. Int. J. Polym. Mater. Polym. Biomater. 62(4): 226–230.
76. Thakur, V. K., and A. S. Singha. 2010. Physico-chemical and mechanical characterization of natural fibre reinforced
polymer composites. Iran. Polym. J. 19(1): 3–16.
77. Singha, A. S., and V. K. Thakur. 2010. Evaluation of Grewia optiva fibers as reinforcement in polymer composites.
Polym.–Plast. Technol. Eng. 49(11): 1101–1107.
78. Singha, A. S., and V. K. Thakur. Synthesis, characterization and study of pine needles reinforced polymer matrix
based composites. J. Reinf. Plast. Compos. 29(5): 700–709.
79. Stoeckel, F., J. Konnerth, and W. Gindl-Altmutter. 2013. Mechanical properties of adhesives for bonding wood—A
review. Int. J. Adhes. Adhes. 45: 32–41.
80. Thakur, V. K., and A. S. Singha. 2010. Natural fibres-based polymers: Part I—Mechanical analysis of Pine needles
reinforced biocomposites. Bull. Mater. Sci. 33(3): 257–264.
81. Singha, A. S., and V. K. Thakur. 2010. Renewable resource-based green polymer composites: Analysis and
characterization. Int. J. Polym. Anal. Charact. 15(3): 127–146.
82. Singha, A. S., and V. K. Thakur. 2010. Mechanical, morphological, and thermal characterization of compression-
molded polymer biocomposites. Int. J. Polym. Anal. Character. 15(2): 87–97.
83. Singha, A. S., and V. K. Thakur. 2009. Mechanical, morphological and thermal properties of pine needle-reinforced
polymer composites. Int. J. Polym. Mater. 58(1): 21–31.
84. Singha, A. S., and V. K. Thakur. 2010. Synthesis and characterization of short Grewia optiva fiber based polymer
composites. Polym. Compos. 31(3): 459–470.
85. Thakur, V. K., and A. S. Singha. 2011. Physico-chemical and mechanical behavior of cellulosic pine needle-based
biocomposites. Int. J. Polym. Anal. Charact. 16(6): 390–398.
86. Thakur, V. K., and A. S. Singha. 2010. Evaluation of Grewia optiva fibers as reinforcement in polymer biocompo-
sites. Polym.-Plast. Technol. Eng. 49(11): 1101–1107.
87. Thakur, V. K., A. S. Singha, and I. K. Mehta. 2010. Renewable resource-based green polymer composites: Analysis
and characterization. Int. J. Polym. Anal. Charact. 15(3): 137–146.
88. Wuzella, G., A. R. Mahendran, T. Bätge, S. Jury, and A. Kandelbauer. 2011. Novel, binder-free fiber reinforced
composites based on a renewable resource from the reed-like plant Typha sp. Ind. Crops Prod. 33(3): 683–689.
89. Abraham, E., P. A. Elbi, B. Deepa, P. Jyotishkumar, L. A. Pothen, S. S. Narine, and S. Thomas. 2012. X-ray
diffraction and biodegradation analysis of green composites of natural rubber= nanocellulose. Polym. Degradation
Stab. 97(11): 2378–2387.
90. Lee, K.-Y., T. Tammelin, K. Schulfter, H. Kiiskinen, J. Samela, and A. Bismarck. 2012. High performance cellulose
nanocomposites: Comparing the reinforcing ability of bacterial cellulose and nanofibrillated cellulose. ACS Appl.
Mater. Interfaces 4(8): 4078–4086.

You might also like