You are on page 1of 227

Narendra Reddy

Sustainable
Applications of
Coir and Other
Coconut
By-products
Sustainable Applications of Coir and Other
Coconut By-products
Narendra Reddy

Sustainable Applications
of Coir and Other Coconut
By-products
Narendra Reddy
Centre for Incubation, Innovation, Research
and Consultancy
Jyothy Institute of Technology
Bangalore, Karnataka, India

ISBN 978-3-030-21054-0    ISBN 978-3-030-21055-7 (eBook)


https://doi.org/10.1007/978-3-030-21055-7

© Springer Nature Switzerland AG 2019


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, express or implied, with respect to the material contained herein or for any errors
or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims
in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Coconut trees and their produce and residues are one of the most distinct crops that
have supported humans and animals for their food and nonfood needs since primi-
tive times. In addition to being one of the primary sources for food in many coun-
tries and cultures, the produce, the tree, and the by-products are utilized for a variety
of non-food applications. Unlike most other resources, almost all parts of a coconut
plant and the coproducts are considered to be of high utility and value. Coir fibers
obtained from the husks of the fruit are the primary residues and commercially
traded as a commodity product. The outer shells of the nuts, leaf, sheath, and trunk
have also been used for various applications. There is probably no other crop that
has such versatile end uses comparable to the coconut tree.
The increase in demand for biodegradable and environmentally friendly products
has put new impetus on the use of agricultural resources. Coir and other coconut
by-products are readily available and have unique properties and hence considered
as potential resources for a multitude of new applications. Cocopeat obtained during
the processing of coir is being extensively used as a soilless growth media. Similarly,
coconut shells are being converted into carbon as a high-level sorbent for the
removal of solid, liquid, and gaseous pollutants. Production of enzymes and biofu-
els and their conversion into supercapacitors and conducting fibers are some of the
newer applications of coir and their by-products. Coir fibers are also being made
into composites either alone or in combination with other fibers. Unusually high
lignin content, high durability, and good resistance to environmental degradation
make coconut by-products a preferred resource for various applications. This book
provides a comprehensive review of the structure and properties of coir fibers and
also the conventional and new uses of coir and other coconut by-products.
Narendra Reddy expresses his sincere thanks to the Director, Center for
Incubation, Innovation, Research and Consultancy and Jyothy Institute of
Technology for his support to complete this work. The assistance to Narendra Reddy

v
vi Preface

by the Department of Biotechnology, Ministry of Science and Technology,


Government of India, through the Ramalingaswami Re-entry Fellowship is also
thankfully acknowledged.

Bangalore, Karnataka, India  Narendra Reddy


Contents

1 Processing and Properties of Coconuts��������������������������������������������������    1


1.1 Introduction��������������������������������������������������������������������������������������    1
1.2 Extraction of Fibers��������������������������������������������������������������������������    1
1.2.1 Production of Coir Fibers Through Retting��������������������������    4
1.2.2 Chemical Extraction of Coir ������������������������������������������������    5
1.2.3 Microbial Extraction of Coir������������������������������������������������    9
1.2.4 Extraction of Fibers with Plasma������������������������������������������   10
1.3 Structure and Properties of Coir Fibers��������������������������������������������   11
1.3.1 Morphological Structure ������������������������������������������������������   11
1.3.2 Thermal Properties of Coir ��������������������������������������������������   12
1.3.3 Acoustic Properties ��������������������������������������������������������������   14
1.3.4 Changes in Properties Due to Geography����������������������������   16
1.3.5 Changes in Coir Properties Due
to Chemical Treatments��������������������������������������������������������   17
1.3.6 Changes Due to Loading and Unloading������������������������������   20
1.4 Chemical Modifications of Coir Fibers��������������������������������������������   20
1.5 Properties of Coir Pith����������������������������������������������������������������������   25
References��������������������������������������������������������������������������������������������������   28
2 Agricultural Applications of Coir����������������������������������������������������������   31
2.1 Introduction��������������������������������������������������������������������������������������   31
2.2 Coir Mulching����������������������������������������������������������������������������������   39
2.3 Coir Geotextiles��������������������������������������������������������������������������������   43
2.4 Miscellaneous Applications��������������������������������������������������������������   50
References��������������������������������������������������������������������������������������������������   52
3 Biotechnological Applications for Coir and Other Coconut Tree
By-products����������������������������������������������������������������������������������������������   55
3.1 Introduction��������������������������������������������������������������������������������������   55
3.2 Production of Enzymes ��������������������������������������������������������������������   55
3.3 Production of Ethanol ����������������������������������������������������������������������   59
3.4 Substrate for Preparation of Catalysts����������������������������������������������   65

vii
viii Contents

3.5 Chemicals and Pharmaceuticals��������������������������������������������������������   68


3.6 Coir Pith as Substrate for Production of Mushrooms ����������������������   69
3.7 Purification of Apple Juice����������������������������������������������������������������   70
3.8 Removal of Urea ������������������������������������������������������������������������������   71
References��������������������������������������������������������������������������������������������������   72
4 Applications of Coir Fibers in Construction ����������������������������������������   75
4.1 Introduction��������������������������������������������������������������������������������������   75
References��������������������������������������������������������������������������������������������������   91
5 Energy Applications of Coir��������������������������������������������������������������������   95
5.1 Coir as Fuel ��������������������������������������������������������������������������������������   95
5.2 Carbonization of Coconut Fiber and Shells
for Supercapacitor Applications��������������������������������������������������������   99
5.3 Gasification of Coir�������������������������������������������������������������������������� 103
5.4 Catalysts for Biodiesel���������������������������������������������������������������������� 111
References�������������������������������������������������������������������������������������������������� 112
6 Coir for Environmental Remediation���������������������������������������������������� 115
6.1 Introduction�������������������������������������������������������������������������������������� 115
6.2 Removal of Dyes������������������������������������������������������������������������������ 115
6.3 Removal of Heavy Metals���������������������������������������������������������������� 122
6.4 Chemicals and Compounds�������������������������������������������������������������� 133
6.5 Desalination�������������������������������������������������������������������������������������� 136
6.6 Sorption of Gases������������������������������������������������������������������������������ 136
6.7 Desulfurization of Diesel������������������������������������������������������������������ 137
References�������������������������������������������������������������������������������������������������� 138
7 Composites from Coir Fibers������������������������������������������������������������������ 141
7.1 Coir Fibers as Reinforcement for Synthetic Polymers �������������������� 141
7.1.1 Polyester Composites Reinforced with Coir Fibers�������������� 141
7.1.2 Polypropylene Composites Reinforced
with Coir Fibers�������������������������������������������������������������������� 144
7.1.3 Polyethylene Reinforced Coir Composites�������������������������� 147
7.1.4 Epoxy Reinforced with Coir Fibers�������������������������������������� 149
7.1.5 Phenol Formaldehyde Based Coir Fiber Composites ���������� 152
7.1.6 Rubber Composites Containing Coir Fibers ������������������������ 152
7.1.7 UV Treatment of Coir Fibers������������������������������������������������ 154
7.2 Biocomposites from Coir Fibers������������������������������������������������������ 155
7.2.1 Poly(Lactic) Acid Based Coir Biocomposites���������������������� 155
7.2.2 Poly(Caprolactone) Based Biodegradable
Coir Composites ������������������������������������������������������������������ 158
7.2.3 Coir Fiber Composites Developed using Proteins
and Starch as Matrix ������������������������������������������������������������ 159
7.3 Hybrid Composites �������������������������������������������������������������������������� 162
7.4 Coir Shell and Coir Pith Composites������������������������������������������������ 178
7.5 Coconut Sheath �������������������������������������������������������������������������������� 180
References�������������������������������������������������������������������������������������������������� 181
Contents ix

8 Miscellaneous Applications for Coir and Other Coconut


By-products���������������������������������������������������������������������������������������������� 187
8.1 Synthesis of Cellulose Nanoparticles������������������������������������������������ 187
8.2 Carbon Nanospheres and Nanotubes������������������������������������������������ 191
8.3 Synthesis of Silver Nanoparticles ���������������������������������������������������� 193
8.4 Sorption of Oil���������������������������������������������������������������������������������� 194
8.5 Cooling Pads ������������������������������������������������������������������������������������ 197
8.6 Pest Control�������������������������������������������������������������������������������������� 198
8.7 Gas Diffusion Layer�������������������������������������������������������������������������� 198
8.8 Fire Resistant Coatings and Biosorption������������������������������������������ 199
8.9 Wood Vinegar������������������������������������������������������������������������������������ 199
8.10 Paper/Boards from Coir�������������������������������������������������������������������� 199
8.11 Precursor for Ceramic Production���������������������������������������������������� 203
8.12 Brake Lining ������������������������������������������������������������������������������������ 204
8.13 Superabsorbent Hydrogels���������������������������������������������������������������� 206
8.14 Development of SiC Whiskers���������������������������������������������������������� 206
8.15 Corrosion Inhibition�������������������������������������������������������������������������� 208
8.16 Production of Bio-oil������������������������������������������������������������������������ 210
8.17 Manufacture of Conducting Fibers �������������������������������������������������� 212
References�������������������������������������������������������������������������������������������������� 214

Index������������������������������������������������������������������������������������������������������������������ 217
Chapter 1
Processing and Properties of Coconuts

1.1 Introduction

The structure and properties of coir and other coconut by-products vary consider-
ably with the variety, climatic conditions at the place of growth, and processing
conditions. In addition, the application of coconut and its by-products also varies
depending on the culture, living environments, value addition desired, and societal
needs. For example, raw coconuts are primarily consumed for the water and also
dried and processed to extract oil. A picture of the various products that can be
manufactured from coconuts is shown in Fig. 1.1. Apart from the coconuts being the
primary product, the leaves, trunk, and flowers of coconuts/trees are also used for
various applications. Typically, raw or undried coconuts contain about 50% solids
and 50% liquid. However, there are considerable variations observed between coco-
nuts, even from the same location and in fact from the same tree (Table 1.1). In
addition to the size of the coconuts, the extent of husk covering the nuts and thick-
ness of the shell vary among the nuts and cause differences in processing of the nuts.
For instance, dehusking a dry coconut took an average of 857 N compared to 745 N
for green coconuts (Varghese and Jacob 2017), whereas it takes an average breaking
force of 4427 N to break a coconut shell having a thickness of 3 mm and diameter
of 95 mm.

1.2 Extraction of Fibers

A majority of coconuts are processed commercially for obtaining oil and other
products. These coconuts are processed after the water inside gets naturally evapo-
rated and nuts are considered dry during which the size of the coconuts reduces
considerably and the coconuts turn brown. After drying, the coconuts are broken
and the inside part is collected and processed for oil. During such commodity

© Springer Nature Switzerland AG 2019 1


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_1
2 1 Processing and Properties of Coconuts

Fig. 1.1 Depiction of the many applications of coconuts

Table 1.1 Variations in the structural features of coconuts (Pandiselvam et al. 2018)
Shell Testa White
thickness, Shell thickness, Testa Coconut White kernel kernel
mm weight, g mm weight, g water, ml thickness, mm weight, g
3.7 ± 0.2 99 ± 12 1.9 ± 0.4 15.6 ± 2.6 180 ± 84 10.8 ± 0.6 275 ± 36
2.8 ± 0.8 71 ± 8 0.5 ± 0.2 9.5 ± 2.5 176 ± 43 9.9 ± 0.5 200 ± 22
3.4 ± 0.5 132 ± 13 0.8 ± 0.4 23.2 ± 2.2 236 ± 106 12.6 ± 0.4 316 ± 28
2.3 ± 0.7 115 ± 19 0.7 ± 0.1 17.4 ± 1.6 186 ± 94 12.2 ± 0.6 293 ± 47
4.0 ± 0.4 161 ± 13 1.0 ± 0.7 19.4 ± 5.3 169 ± 63 9.8 ± 1.3 251 ± 30
2.7 ± 0.3 134 ± 11 0.7 ± 0.4 16.5 ± 2.7 151 ± 44 8.0 ± 0.6 224 ± 24
3.1 ± 0.3 49 ± 4 0.8 ± 0.5 6.2 ± 3.4 118 ± 26 11.1 ± 0.3 105 ± 21
2.4 ± 0.2 36 ± 5 0.7 ± 0.2 3.3 ± 1.2 86 ± 29 10.2 ± 0.4 96 ± 19
3.8 ± 0.3 108 ± 16 2.0 ± 0.9 17.1 ± 1.4 59 ± 32 11.3 ± 0.5 228 ± 26
3.4 ± 0.4 97 ± 16 1.9 ± 0.3 15.2 ± 5.0 41 ± 23 10.7 ± 0.5 199 ± 33
Reproduced with permission from Taylor and Francis

processing, the outer layer covering the coconut shell, “husk” is removed (Fig. 1.2).
Further processing of the husk results in the brown coir fiber as one of the by-
products. Alternatively, green or immature coconuts are used for the water and the
husks of these coconuts which are green on the outside produce the white coir
fibers. The white coir fibers have a density of 1.01 g/cm3 compared to a density of
1.2 Extraction of Fibers 3

Fig. 1.2 Image of the


parts of a coconut

Table 1.2 Comparison of the structure and property differences between white and brown coconut
fibers (Valášek et al. 2018)
White coconut fibers Brown coconut fibers
Total intrusion volume, mL/g 0.6578 0.765
Total pore area, m2/g 22.62 14.37
Median pore diameter (volume), μm 555.4 751
Median pore diameter (area), nm 13.8 6.6
Average pore diameter (4 V/A), μm 0.116 0.213
Bulk density at 1.02 psi, g/mL 0.709 0.664
Apparent density, g/mL 1.32 1.35
Porosity, % 46.6 50.8
Tensile strength, MPa, raw fibers 115 ± 33 123 ± 54
Tensile strength, MPa, after alkali 151 181
treated
Tensile modulus, GPa 3.39 4.16
Reproduced with permission from Elsevier

1.29 g/cm3 for the brown fibers (Valášek et al. 2018). Morphologically, there are
differences between the brown and white coir fibers which require the fibers to be
processed and used differently. Tables 1.2 and 1.3 list some of the variations in the
structure and properties of green and brown coir fibers. Green coconut fibers from
Brazil had fibers with fineness ranging from 272 to 514 μm with 10–30% variation
along the length of the single fibers (Lomelí-Ramírez et al. 2018), and individual
fibers had a thickness of 3.5 μm and diameter of 11.7 μm.
4 1 Processing and Properties of Coconuts

Table 1.3 Properties of brown and green coir fibers reported in literature (Lomelí-Ramírez et al.
2018)
Fiber type Diameter, mm Length, mm Modulus, GPa Strength, MPa Elongation, %
Green 0.25 ± 0.02 – 1.96 ± 0.2 159 ± 26.4 41.3 ± 4.6
Brown 0.225 25-May 1.26–2.73 118.3–142.6 –
Brown 0.21 – 2.8 107 37.7
Brown 0.10–0.40 60–50 – 15–327 17.6–75
Brown 0.27 ± 0.07 50 ± 10 2.0 ± 0.3 142 ± 36 24 ± 10
Green 0.11–0.53 – 2.50–4.50 108–252 13.7–41
Brown 0.12 ± 0.05 – 3.7 ± 0.6 158 –
Green 0.1–0.4 – 16–26 174 25-Oct
Brown – – 8 95–118 –
Green – – 0.8–1.6 82–129 –
Green – – – 75–140 20–40
Brown 0.11–0.46 20–50 1.74–2.78 132–139 34.2–67.8
Brown – – – 204 3.2
Reproduced through open access publication by North Carolina State University

1.2.1 Production of Coir Fibers Through Retting

Conventional method of extracting fibers from the coconut husks is through the
process called retting. Although retting is commonly practiced and produces good
quality fibers, the process is associated with several limitations. Retting is a long
process and may take anywhere from 3 to 12 months depending on the conditions
during retting. Since coconuts are grown in coastal regions, retting using backwa-
ters is a common practice. However, natural retting in backwaters is reported to
cause health and environmental hazards (Basu et al. 2015). For example, release of
organic substances and chemicals causes increase in biological oxygen demand
(BOD) and affects organisms (Basu et al. 2015). It is estimated that 1.5 tons of BOD
is generated for every ton of coir fibers retted (Shibu et al. 2013). Due to the limita-
tions of conventional retting, several physical and chemical approaches have been
used to increase the rate of retting and attempts have also been made to even avoid
the process.
A new process of closed loop retting was developed to avoid pollution from open
water retting. In this process, an upflow anaerobic sludge blanket (UASB) was used
and coconut leachate was used for the degradation. Additional nutrients, bicarbon-
ate, and trace elements were added and appropriately aerated. Considerably lower
levels of phenolic compounds were released and the fiber obtained was whiter and
softer (Shibu et al. 2013). Although chemical retting is effective, faster and provides
fibers with better properties, disposal of the liquor (black liquor) used for retting is
a concern. A study was conducted to understand the feasibility of reusing the retting
bath. Three baths consisting of different chemical combinations were formulated.
Considerable variations were observed in the properties of the fibers after the treat-
ments. It was found that the properties of the fibers obtained using the spent baths
1.2 Extraction of Fibers 5

Table 1.4 Changes in the properties of the coir fibers after treating with the three different
chemical baths (Jose et al. 2016)
Fiber parameters Raw fiber Bath A Bath B Bath C
Diameter, μm 273 228 247 218
Length, cm 23 23.1 23.4 22.7
Length-diameter ratio 843 1011 936 1058
Linear density, tex 39.1 32.5 38.7 31.7
Breaking tenacity, cN/tex 12.6 15.3 14.4 15.7
Breaking extension, % 22.1 25.9 27 27.6
Weight loss, % – 25 18 24
Moisture regain, % 10.3 11.6 10.8 11.8
Flexural rigidity, mN/mm2 1265 686 89 643
Reproduced with permission from Taylor and Francis

were better or similar to the pristine bath (Table 1.4). Reusing the bath will therefore
be helpful to reduce costs and pollution. Darker color (lower brightness) was one of
the limitations of fibers from the reused bath (Jose et al. 2016).
• Bath A: Standard retting bath (40% sodium sulfide +10% sodium hydroxide
+25% sodium carbonate).
• Bath B: Remnant liquor of bath A +25% of original bath A chemicals.
• Bath C: Remnant liquor of bath A +50% of original bath A chemicals.

1.2.2 Chemical Extraction of Coir

One of the common methods to delignify biomass and extract fibers is through alka-
line extraction with or without the help of reducing agents, emulsifiers, steam, acids,
etc. In one such attempt, coconut husks were desalted and later treated with 10%
sodium hydroxide, 10% hydrochloric acid, or 10% acetic acid separately at 30 °C
for 4 h. Later, the treated fibers were exposed to ultraviolet radiation for 100 h
(Varma et al. 1984). Properties of the untreated and treated fibers are shown in
Table 1.5. Alkali treatment was more effective in removing non-cellulosic sub-
stances and providing fibers with higher strength. Acid conditions hydrolyzed the
fibers and led to a decrease in tensile strength, orientation of the cellulose crystals,
and percent crystallinity (Varma et al. 1984).
Green coconut fibers from Brazil were treated with three different bleaching
agents to improve wettability, tensile properties, and morphology. A reduction in
fiber diameter from 157 μm to 44 μm occurred after exposure to NaOCl/
NaOH. Corresponding increase in cellulose content was from 46 to 62%. Although
waxes, lignin, and other impurities were removed making the fibers smoother, for-
mation of pits and damage to the fibers were observed when H2O2 was used (Fig. 1.3)
(Brígida et al. 2010). X-ray photon spectroscopy (XPS) studies showed that the
surface composition of the fibers (Table 1.6) varied with the different treatments.
6 1 Processing and Properties of Coconuts

Table 1.5 Changes in the properties of coir fibers before and after the various treatments (Varma
et al. 1984)
Composition, % %
Moisture Modulus, Tenacity, Elongation, Crystallinity
Treatment Carbon Hydrogen regain, % g/den g/den % index
Untreated 49.6 6.1 8.0 38.3 2.26 28.8 25.6
Desalted 50.5 6.0 7.0 40.1 2.21 28.6 22.5
Sodium 49.2 6.3 9.0 43.6 2.08 27.4 28.9
hydroxide
Hydrochloric 50.1 6.1 16.0 33.6 1.49 21.8 23.3
acid
Acetic acid 50.5 6.1 11.0 35.3 1.9 27.3 22.8
UV – – – 32.5 1.34 19.4 –
light-
untreated
UV light- – – – 35.5 1.45 20.0 –
treated fibers

In another study, raw coconut fibers were subject to various treatments and the
changes in chemical and physical properties were studied in detail (Basu et al.
2015). Substantial differences in the physical, chemical, and mechanical properties
were observed before and after treatment (Table 1.7). In terms of the composition of
the fibers, the cellulose, lignin, or hemicellulose content did not change much due
to backwater retting but lignin content decreased by about 20% and cellulose con-
tent increased by about 42% when chemical retting was done. Crystallinity of the
fibers also increased from 37 to 54% due to the chemical treatments. It was con-
cluded that chemical retting provides fibers with better properties and decreases the
retting time from 10 months to just 2 h. However, the cost of the process and
­subsequent effluent treatment issues should also be considered. Use of steam after
alkali treatment was also found to be effective in reducing the lignin in the fibers
from 39 to 31% and enriching the cellulose content from 50 to 57% (Abraham et al.
2013). In this approach, coir fibers were treated with 2% NaOH for 6 h at 25 °C and
later subject to steam explosion at 137 Pa at temperatures between 100 and
150 °C. Fibers were also bleached and acid hydrolyzed to extract the nanofibers.
Raw coir fibers which had cellulose content of 39% and lignin content of 49%
showed remarkable changes in composition with 88% cellulose and 0.3% lignin
after bleaching. Removal of lignin and hemicellulose also improved the percentage
of crystallinity and substantially changes in fiber morphology were obtained
(Fig. 1.4). Individual fibrils with diameter between 3 and 12 μm were obtained after
bleaching and nanofibers with diameters of 100 nm were formed after the acid
hydrolysis (Abraham et al. 2013).
1.2 Extraction of Fibers 7

Fig. 1.3 Changes in the morphology of the green fibers after treating with various chemicals
(Brígida et al. 2010). NaOCl treated (a); NaOCl/NaOH treated (b); NaOCl/NaOH treated fiber
showing partial disintegration (c); NaOCl/NaOH showing wax and fatty acid residues (d); H2O2
treated fibers at 500× (e) and 1000× (f). Reproduced with permission from Elsevier

Table 1.6 Surface elemental composition of the green coconut fibers after various treatments
(Brígida et al. 2010)
Treatment C O N Si Ca Mg Cl Na O/C
Untreated 69.05 27.36 2.04 1.10 0.46 – – – 0.40
NaOCl 59.89 34.43 1.32 0.39 0.13 0.96 2.89 – 0.57
NaOCl/NaOH 63.39 30.41 0.86 0.97 0.29 0.19 0.30 3.60 0.48
H2O2 65.01 33.23 0.81 0.78 0.17 – – – 0.51
Published with permission from Elsevier
8

Table 1.7 Changes in the physical and mechanical properties of coir fibers after various treatments (Basu et al. 2015)
Physical property Mechanical property
Flexural
Weight Diameter, Fineness, L/D rigidity, cN/ Breaking Breaking Specific work of
Treatment loss, % μm tex Ratio mm2 tenacity, cN/tex elongation, % rupture, mJ/tex.m
– Control – 345 53 695 1273 11.3 21.5 12.1
– Backwater retted – 282 44 748 613 14.0 27.1 12.0
– Boil water, 2 h 2.6 381 52 689 1235 12.1 18.8 13.9
– Steam at 121 °C, 2 h 7.2 347 48 712 1177 12.2 19.2 14.5
– 20% NaOH, boil, 2 h 12.2 341 45 789 1034 13.8 24.7 19.3
– 40% Na2S, boil, 2 h 17.6 227 41 917 731 13.4 27.8 18.1
– 15% Na2CO3, boil, 2 h 11.5 359 49 742 1084 11.9 22.4 12.5
– 40% Na2S, 20% NaOH, boil 21.7 249 43 898 893 11.8 21.5 18.2
for 2 h
– 40% Na2S, 15%Na2CO3, 20.1 255 48 915 921 12.1 21.2 17.4
boil for 2 h
– 40% Na2S, 15% Na2CO3, 28.4 225 34 991 361 14.2 21.3 21.6
20% NaOH, boil for 2 h
– 3% H2O2, 2 h, 80 °C, pH 11 13.6 251 39 891 890 11.7 19.9 17.5
– 2% NaIO4, boil, 2 h 6.4 276 52 734 1062 8.4 14.3 9.5
– HCl, 30 °C, 1 h 37.4 250 35 1073 745 2.9 3.90 0.5
– H2SO4, 30 °C, 1 h – – – – – – – –
1 Processing and Properties of Coconuts
1.2 Extraction of Fibers 9

Fig. 1.4 Changes in the surface morphology of coir fibers before (a), after alkali treatment (b),
steam explosion (c), and bleaching (d) (Abraham et al. 2013). Reproduced with permission from
Elsevier

1.2.3 Microbial Extraction of Coir

Instead of using chemicals, the efficiency of microorganisms to soften and delignify


coir fibers was studied. Retted and unretted coir fibers having length of 17 cm and
diameter of 250 μm were treated with microorganisms isolated from the retting back-
waters Pseudomonas putida and Phanerochaete chrysosporium, Aspergillus flavi-
ceps and Trametes hirsuta for 20 days at 30 °C in neutral pH (Rajan et al. 2005).
Tensile strength of the fibers had increased from 164 up to 245 MPa for the retted
fibers and from 199 to 270 MPa for the unretted fibers. An increase in elongation
from 22 to 40% had also occurred. The microbial process was considered to be suit-
able for softening retted coir and resulted in softer, whiter, and stronger fibers com-
pared to chemical treatments. However, the treatment conditions such as pH, time,
temperature, and carbon source for the microorganisms should be controlled to obtain
fibers with good properties (Rajan et al. 2005). In a similar study, lignin degrading
white rot fungi Phanerochaete chrysosporium and Ceriporiopsis subvermispora,
Pleurotus eryngii and Ganoderma lucidum were studied as alternative to the conven-
tional retting process. Coconut husks were treated at 30 °C for up to 90 days, and the
changes in cellulose and lignin content were noted. In addition to the decrease in
lignin and increase in cellulose content, (Table 1.8), release of the biofouling agent
10 1 Processing and Properties of Coconuts

Table 1.8 Changes in the lignin and cellulose content of coir fibers after retting for 30, 60, and
90 days using various fungi (Suganya et al. 2007)
% Lignin % Cellulose
Treatment 30 60 90 30 60 90
Control 45.5 ± 0.16 42.4 ± 0.13 41.4 ± 0.31 43.0 ± 0.13 42.0 ± 0.08 39.2 ± 0.21
Phanerochaete 34.5 ± 1.70 30.4 ± 0.82 25.3 ± 0.50 39.8 ± 0.26 38.8 ± 0.22 37.4 ± 0.49
chrysosporium
Ceriporiopsis 36.7 ± 0.79 34.5 ± 0.59 29.8 ± 0.40 40.9 ± 0.68 39.7 ± 1.25 30.1 ± 1.18
subvermispora
Control 43.7 ± 0.19 39.5 ± 0.23 35.3 ± 0.55 41.1 ± 0.08 40.1 ± 0.07 39.9 ± 0.05
Pleurotus 38.2 ± 0.79 37.3 ± 0.18 29.2 ± 0.67 40.1 ± 0.05 39.2 ± 0.57 37.0 ± 0.82
eryngii
Control 42.6 ± 0.09 38.6 ± 0.05 35.2 ± 0.03 42.2 ± 0.05 41.3 ± 0.16 39.8 ± 0.25
Ganoderma 39.6 ± 0.1 37.4 ± 0.26 33.8 ± 0.20 42.0 ± 0.47 41.0 ± 0.08 39.2 ± 0.63
lucidum

tannin was lower by 57% and the treated fibers had considerably high dye uptake of
97% compared to 88% for the untreated fibers (Suganya et al. 2007).
A bacterial consortium isolated from coconut husk leachate was used to acceler-
ate the retting process under laboratory conditions. After 63 days of incubation, the
husk treated with inoculum peeled out easily and formed finer and less colored
fibers. It was suggested that retting in tanks using the inoculum was a better approach
than the backwater retting (Ravindranath and Bhosle 2000).

1.2.4 Extraction of Fibers with Plasma

Other than using chemicals and enzymes, delignification of coir has also been done
using plasma. Plasma treatment was done using pure oxygen and air at pressures of
10 Pa for air and 5 Pa for oxygen at power of 50 or 80 W (de Farias et al. 2017).
Considerable changes were observed both in the physical and chemical properties of
the fibers and products made from the fibers. Tensile strength of composites made
using fibers treated with 80 W and oxygen increased from 2 to 7 MPa and the elastic
modulus from 30 to 370 MPa. Extent of delignification also varied depending on the
treatment condition but led to substantial changes to the morphology and composi-
tion of the fibers (Fig. 1.5, Table 1.9). In another approach, it was shown that consid-
erable amount of hemicellulose in the coir could be removed by ultrasonic treatment
at 45 KHz and 400 W for up to 40 h (Renouard et al. 2014). After 24 h of treatment,
coir fibers had a weight loss of about 9% and had released 0.37 × 10−3 % of lignin.
Based on the FTIR absorption peaks at 1375 cm−1, 1595 cm−1, and 897 cm−1/1595 cm−1,
it was suggested that the crystalline and amorphous celluloses in the coir fibers were
not affected. However, the amount of pectin, xylan, xyloglucan, and hemicellulose
fractions clearly showed considerable decrease in hemicellulose content with overall
weight loss of 4% (Renouard et al. 2014).
1.3 Structure and Properties of Coir Fibers 11

Fig. 1.5 Morphological changes in the coir fibers after various plasma treatments (de Farias et al.
2017). Reproduced with permission from Elsevier

Table 1.9 Changes in the composition of coir fibers due to plasma treatment based on the ratio of
FTIR peaks (de Farias et al. 2017)
Lignin/cellulose ratio Lignin/cellulose ratio Hemicellulose/cellulose ratio
(1508/898 (1508/1317 (1602/898 (1602/898 (1728/898 (1728/898
Treatment cm−1) cm−1) cm−1) cm−1) cm−1) cm−1)
Untreated 3.65 ± 0.41 16.21 ± 1.02 0.28 ± 0.03 1.23 ± 0.06 6.95 ± 0.95 30.83 ± 3.08
Air 3.61 ± 0.22 1.61 ± 0.08 0.25 ± 0.03 0.11 ± 0.01 5.59 ± 0.71 2.51 ± 0.39
Oxygen 4.17 ± 0.16 0.99 ± 0.06 0.27 ± 0.02 0.06 ± 0.01 5.03 ± 0.50 1.19 ± 0.17
Reproduced with permission from Elsevier

1.3 Structure and Properties of Coir Fibers

1.3.1 Morphological Structure

Coir fibers are composed of several (200–300) elementary fibers and a large lumen
in the center (Fig. 1.6). The elementary fibers are made up to two main cell walls
and are in turn made of bundles of microfibrils which determine the microfibrillar
angle (Tran et al. 2015). Presence of the lumen also led to a porosity of 20–40% in
12 1 Processing and Properties of Coconuts

Fig. 1.6 Scanning electron microscope of the elementary fibers in the fibers show the presence of
large lumen and primary and secondary cell walls (Tran et al. 2015). Reproduced with permission
from Elsevier

the fibers. However, considerable variations have been observed in the properties of
a single fiber (Tran et al. 2015). For instance, diameter of fibers varied from 158 to
301 μm, pore area varied from 4142 to 19,000 μm2, and total fiber porosity was
between 21 and 31%. For the same fibers, measurements using SEM-CT scan pro-
vided a total porosity between 27 and 46% and elementary fiber diameter and length
between 5.6 and 15.7 and 283 and 960 μm, respectively (Tran et al. 2015). Such
variations in structure and properties are inherent in most natural materials. Further
investigations have shown that coir fibers are composed of nanofibrils which in turn
form the microfibrils (Fig. 1.7). Nanofibrils of various dimensions have been
extracted from coir fibers and generally are in the range of 6–10 nm. In addition, the
fibers in most places are curved in the form of helical ribbons with angle of 83–88°.
Visually, presence of craters and bundle of fibrils held together to form the fibers
could be observed. Typical dimensions of the coir fiber, tube fiber, and helical rib-
bon are given in Table 1.10. Based on the measurements of the different fibrous
elements, a new model for the morphological structure of coir fibers has been pro-
posed (Yu et al. 2016) (Fig. 1.7).

1.3.2 Thermal Properties of Coir

Thermal properties of coir fibers have also been extensively studied and found to
vary with the type of fibers and modifications done to the fibers. For instance, alkali-­
treated coir fibers had maximum degradation temperature of 366 °C compared to
339 °C for acetic acid treated fibers. Some of the thermal parameters that were
affected by alkali or acid treatment are given in Table 1.11 (Varma et al. 1986). In
another study, coir fibers subjected to various alkali treatments (mercerization using
1.3 Structure and Properties of Coir Fibers 13

Fig. 1.7 A schematic of the multilevel fibril packing model for coir fibers (Yu et al. 2016).
Reproduced with permission from Springer Nature

Table 1.10 Physical parameters of the coir, tube fiber, and helical ribbon (Yu et al. 2016)
Type Parameter Units, μm
Coir fiber Fiber length 140 × 103–180 × 103
Fiber diameter 130–280
Length/diameter ratio 500–1380
Thickness of cell membrane complex 3.32–4.02
Corrugated width 8.13–13.01
Arch span 15.02–18.78
Exterior margin diameter of volcano 14-Sep
Inner margin diameter of volcano 9-Jun
Tube fiber Thickness of tube fiber 0.92–1.51
Diameter of tube fiber 6.92–19.23
Inside diameter 4.59–17.87
Length of tubular fiber 250–1394
Length/diameter ratio 36–200
Helical ribbon Width of helical crystal ribbon 1.95–2.64
Inter-ribbon matrix 0.27–0.48
Pitch of the ribbon coil 2.22–3.12
Helical angle 83–88°
Thickness of wall surface layer 0.24
Thickness of crystal of ribbon coil 0.64–1.06
Thickness of inner wall layer 0.12
Reproduced with permission from Springer Nature
14 1 Processing and Properties of Coconuts

Table 1.11 Changes in the thermal behavior of coir fibers before and after treating with acids/
alkali (Varma et al. 1986)
Untreated Alkali Hydrochloric acid Acetic acid
Thermal property coir treated treated treated
Tmax-1 (°C) 292–295 – 300 295
Tmax-2 (°C) 339–367 366 368 339
Tf (°C) 310–390 385 392 –
Residual weight (%)-Tmax-1 77–85 85 85 78
Residual weight (%)-Tmax-2 44–54 50 48 50
Weight up to 150 °C (%) 3.0–5.0 4 5 5
Weight loss (%) 22–27 23 25 29
Char yield @ 450 °C, % 25–34 30 27 23
Reproduced with permission from Elsevier

5–30% NaOH at room temperature for 3 h) showed substantially different thermal


behavior. It was suggested that alkali causes decrease in crystallinity but increases
accessibility leading to changes in moisture sorption and thermal properties (Varma
et al. 1986).
Few researchers have also studied the changes in the thermal behavior after a
series of alkali treatments. As seen from Table 1.12, the peak and final degradation
temperatures increase after treating with 15–20% alkali (Mahato et al. 2013). This
indicates that the stability of the fibers increased after treating with alkali which was
also supported by the energy absorbed data. A study by Mohato et al. also suggested
that treating with alkali increased the activation energy making the fibers more sta-
ble. Thermal stability of the fibers was found to improve up to 15% alkali treatment
(Mohato et al. 1995).

1.3.3 Acoustic Properties

The presence of the hollow structure in coir fibers also provides good acoustical
properties. For example, the sound absorption coefficient of coir fibers varied from
0.1 to 0.9 depending on the frequency with maximum absorption between 1000 and
2500 Hz (Fouladi et al. 2011). However, the absorption coefficients varied consider-
ably with changes in the thickness or the processing of the coir fibers. Industrially
processed coir fibers containing a binder had maximum sorption between 3000 and
4000 Hz or between 1500 and 3500 Hz depending on the sample used (Figs. 1.8 and
1.9). It was suggested that natural coir fibers should be supplemented with binders
or other additives to enhance sound absorption, particularly in the low frequency
region (Fouladi et al. 2011).
1.3

Table 1.12 Effect of alkali treatment on the changes in initial (Ti), peak (Tp), and final (Tf) temperatures (°C) and energy absorbed after various alkali
treatments (Mahato et al. 2013)
First endotherm Second endotherm Third endotherm
Sample Ti Tp Tf Ti Tp Tf Ti Tp Tf Energy absorbed, mJ
Structure and Properties of Coir Fibers

Raw 20 105.2 185.1 185.1 257.3 297.8 297.8 336.7 381.2 1954
5% mercerized 20 114.2 300.9 – – – – 2400
10% mercerized 30 111.1 228.1 288.2 – 300.3 – 2610
15% mercerized 20 108.7 230.2 230.3 – 310 – 2845
20% mercerized 20 105.5 239.2 239.3 – 302.3 – 3467
30% mercerized 20 110 216.3 216.3 – 282 – 2222
Reproduced with permission from The National Institute of Science Communication and Information Resources
15
16 1 Processing and Properties of Coconuts

Fig. 1.8 Sound absorption profile of raw coir fibers with two different sample configurations
(Fouladi et al. 2011). Reproduced with permission from Elsevier

Fig. 1.9 Changes in the sound absorption coefficient of industrially processed coir fibers contain-
ing a binder (Fouladi et al. 2011). Reproduced with permission from Elsevier

1.3.4 Changes in Properties Due to Geography

Properties of coir fibers are also highly dependent on the geographical region. Large
variations have been observed in the composition, structure, and properties of coir
fibers obtained from different regions and also depending on the treatment of the
fibers and testing conditions used. For instance, Brazilian coconut husks were retted
for 4 months and the resulting fibers were soaked in 5% NaOH for 48 or 72 h.
Properties of the fibers were noticed to be considerably different between the two
1.3 Structure and Properties of Coir Fibers 17

Table 1.13 Changes in the tensile properties of coir fibers after treating in 5% alkali for 48 h
(Silva et al. 2000)
Tensile strength, MPa Elongation at break, % Initial modulus, GPa
Raw coir (retted) 76 ± 15 29 ± 5 2.1 ± 0.3
Alkali treated (48 h) 94 ± 12 33 ± 5 2.9 ± 0.3
Reproduced with permission from John Wiley and Sons

Fig. 1.10 The elliptical cross-section (left), large lumens (middle), and helical structure (right)
seen in coir fibers soaked in alkali (Silva et al. 2000). Reproduced with permission from John
Wiley and Sons

treatments (Silva et al. 2000) (Table 1.13). Unlike most coir fibers which have a
circular cross-section, the Brazilian variety was found to have an oval cross-section
(Fig. 1.10). White coir fibers from Vietnam had circular cross-section, whereas the
brown coir fibers were seen to have an oval cross-section (Fig. 1.9). Irrespective of
the source of the coir, the large lumen and helical configuration of the fibers could
be observed. Tensile properties of the fibers increased substantially after treating
with alkali. About 23% increase in strength and up to 33% increase in elongation
had occurred. Higher moisture content but decrease in thermal stability was also
caused by the alkali treatment (Silva et al. 2000). In another study, coir fibers found
in Caribbean had strength in the range of 118–143 MPa compared to 186–343 MPa
for coir fibers from Vietnam and 112–161 MPa for coir fibers extracted from coco-
nuts grown in India (Mathura and Cree 2016). Similarly, elongation varied from 25
to 60, 26 to 64, and 18 to 43% for Brazil, Vietnam, and Indian coir, respectively.
Chemical composition also showed a major difference between the fibers with lig-
nin content ranging from 59% for Jamaican coir compared to 35% for Indian coir.

1.3.5 Changes in Coir Properties Due to Chemical Treatments

Treatment of coir with distilled or salt water and the conditions during testing also
affected tensile properties and crystallinity (Table 1.14, Fig. 1.11). In addition to the
tensile property changes seen in table, the percentage of crystallinity of the fibers
treated with distilled water decreased to 10.8% compared to 21.6% for the unretted
fibers (Mathura et al. 2014; Mathura and Cree 2016). A considerably high alkali
18 1 Processing and Properties of Coconuts

Table 1.14 Properties of coir fibers before and after treating with distilled water (DW) or salt
water (SW) and at two different testing conditions (Mathura et al. 2014)
Retting
Retting time, weeks medium Tensile strength, MPa Modulus, GPa Strain at break, %
Gauge length = 20 mm, strain rate = 20 mm/min
0 Unretted 139 ± 31.8 1.74 ± 0.6 67.8 ± 12.9
1 DW 124 ± 36.7 1.7 ± 0.2 39.6 ± 18.6
SW 103 ± 15.2 1.3 ± 0.3 44.7 ± 2.8
4 DW 102 ± 47 1.8 ± 0.6 39.3 ± 9.2
SW 103 ± 39 1.6 ± 0.7 49.5 ± 19.5
12 DW 169 ± 17 2.5 ± 0.7 49.4 ± 10.1
SW 126 ± 14 1.7 ± 0.3 47.7 ± 10.6
Gauge length = 50 mm, strain rate = 20 mm/min
0 Unretted 133 ± 54 2.5 ± 0.6 34.2 ± 21.1
1 DW 98 ± 21 1.7 ± 0.4 36.2 ± 2.1
SW 101 ± 24 1.6 ± 0.3 34.5 ± 4.9
4 DW 85 ± 38.5 1.7 ± 0.3 31.9 ± 17.7
SW 106 ± 16 1.9 ± 0.2 36.4 ± 4.2
12 DW 109 ± 58 2.2 ± 0.7 35.6 ± 21.9
SW 132 ± 37 2.9 ± 0.8 30.0 ± 8.8
Reproduced with permission from Cambridge University Press

Fig. 1.11 Changes in the tensile properties of coir fibers subject to distilled or saline water treat-
ment and with variation in strain rates (Mathura and Cree 2016). NR denotes non-retted, R denotes
retted fibers, DW is distilled water, and SW is salt water. Reproduced with permission from John
Wiley and Sons
1.3 Structure and Properties of Coir Fibers 19

concentration (50%) and glutaraldehyde (25%) was used to treat coir fibers for 2 h
and improve the thermal stability and surface texture (Manjula et al. 2018a, b).
Strength and elongation of the fibers increased after thermal treatment for up to
5 days but decreased when the treatment was continued for 9 days. Similarly, tensile
strength increased after treating with alkali, whereas elongation decreased consider-
ably due to treating with glutaraldehyde (Manjula et al. 2018a). Based on FTIR
studies, it was suggested that alkali treatment removed the fatty acids (disappear-
ance of carboxyl group). Also, appearance of peak at 1251 cm−1 was due to the
crosslinking between cellulose, lignin, and glutaraldehyde through ether linkages.
In terms of thermal stability, the raw, NaOH, glutaraldehyde treated fibers had resid-
ual mass of 22, 4.9, and 6.1%, respectively. Chemical treatment slightly changed the
dimensions of the cellulose crystal but decreased crystallinity from 50.8% to 36%.
Treated fibers were considered to be highly suitable as filler for composites.
Coir fibers were treated with two different concentrations of alkali under tension
to understand the effect of swelling and stretching on tensile properties and physical
structure. Treating with alkali removes considerable amounts of lignin and non-­
cellulosic components, which leads to improvement in the crystal structure unlike
the cellulose in cotton where alkaline treatment disrupts the crystal structure.
Conversion of cellulose I to cellulose II also occurred when the fibers were treated
with 18% alkali. Another important parameter related to the elongation of the fibers,
the spiral angle, increased from 40 to 45° after delignification. Stretching in water
aligns the spiral and reduced the angle to 34°. However, these changes did not affect
the tensile strength or elongation of the fibers (Sreenivasan et al. 1996). In another
study, treating coir fibers with various concentrations of alkali did decrease tensile
strength from 560 to 442 cN but increased elongation substantially from 10 to 18%
(Gu 2009). Morphologically, the surface substances were removed and a porous
fiber (Fig. 1.12) was formed after treating with 4% NaOH. Such changes were sug-
gested to improve the adhesion between fibers and synthetic binders for composite
applications (Gu 2009). Coir fibers were treated with 3% NaOH solution for 1–5 h
and the changes in properties were determined (Musanif and Thomas 2015). Tensile
strength, elongation, and modulus decreased with increasing concentration of alkali.
Although the tensile properties decreased, it was reinterated that the alkali treatment
would improve interfacial interaction and provide composites with better properties
when coir is used as the reinforcement. Instead of bulk treatments, single coir fibers
were treated with 5% NaOH for 0, 1, 2, 3, and 4 h and changes in tensile properties
and moisture sorption were measured (Setyanto et al. 2013). Density of the coir
fibers decreased from 0.814 g/cm3 to 0.708 g/cm3 with increase in soaking time
from 0 to 4 h. Tensile strength and elongation increased up to 2 h of treatment,
whereas modulus did not show any appreciable change.
In addition to alkali, treating coir with sodium chlorite or acrylamide also
affected the tensile properties and thermal behavior (Khan and Alam 2012). Alkali-­
treated fiber showed decrease in strength, whereas those treated with acrylamide
increased strength to 245 MPa from 204 MPa before treatment. Degradation tem-
perature showed larger increase for alkali-treated fibers compared to those treated
with acrylamide with residual char content of 24 and 17%, respectively. In addition,
20 1 Processing and Properties of Coconuts

Fig. 1.12 Impurities on the surface of the coir fibers are removed, resulting in relatively smooth
and porous fibers after treating with 4% alkali (Gu 2009). Left image is before and right image is
after alkali treatment. Reproduced with permission from Elsevier

increasing temperature or contact time also affected the properties of the fibers.
Degree of polymerization (DP) of cellulose in the fibers was 432 for the untreated
fiber but decreased to 348 for the treated fibers. Sodium chlorite and sodium hydrox-
ide treated fibers had DP decrease from 504 to 446 and 512 to 459, respectively.
Comparatively larger changes were seen in the tensile strength of the samples with
decrease from 194 to 154 MPa (Khan and Alam 2012).

1.3.6 Changes Due to Loading and Unloading

Coir fibers having length of about 60 cm and diameter of 80–500 μm were subject
to cyclic loading and unloading to understand the effect of physical structure and
properties (Martinschitz et al. 2008). The structural changes were characterized in
situ using a wide angle X-ray scattering beamline. Radiation was done using a
wavelength of 0.082 nm and beam size of 0.5 × 1 mm2. For the tensile tests, the
fibers were strained at a rate of 1.4 × 10−4/s. Coir fibers had an average microfibrillar
angle (MFA) of about 45° and the MFA was directly related to the rate of strain
applied (Fig. 1.13). Also, the fibers could regain the strain even after repeated load-
ing and unloading cycles (Martinschitz et al. 2008).

1.4 Chemical Modifications of Coir Fibers

Several researchers have also studied the possibility of chemically modifying the
coir fibers to improve performance properties and extend applications. Coir fibers
were washed with detergent, dewaxed by treating with ethanol and benzene for 72 h
at 50 °C and later with distilled water. These fibers were again treated with 5%
1.4 Chemical Modifications of Coir Fibers 21

Fig. 1.13 Changes in the stress and MFA of coir fibers subject to various strain rates and repeated
loading and unloading cycles (Martinschitz et al. 2008). Reproduced with permission from
Springer Nature

NaOH for 30 min at 30 °C and rewashed with water. Methylmethacrylate (MMA)


was grafted onto the fibers to increase their compatibility with synthetic resins to
develop composites (Bismarck et al. 2001). Changes in the surface wettability and
moisture sorption of the fibers were determined by measuring the zeta (ζ) potential.
Dewaxed fibers had a zeta potential of −3.4 mV compared to −5.9 for the grafted
fibers. Water uptake of the fibers at 100% relative humidity varied from 0.38 to 0.80
22 1 Processing and Properties of Coconuts

and percentage of moisture content was between 6.5 to 9.2%. Maximum decompo-
sition temperature had also increased by 10 °C after treating with alkali. It was sug-
gested that zeta potential was a convenient method to determine the changes in the
surface properties and determine the suitability of the modified fibers for composite
applications. Dewaxed and alkali-treated coir fibers were also grafted with MMA
and acrylonitrile using acetone and pyridine at 60 °C for 2 h (Rout et al. 2002).
Alkali (5%) treated fibers showed considerable decrease in tensile strength to
80 MPa from 108 MPa but 10% alkali-treated fibers had strength of 260 MPa.
Grafting further improved strength, particularly for the cyanoethylated fibers that
had tensile strength of 252 MPa. Coir fibers having diameters between 200 and
240 μm, cellulose content of 33%, and klason lignin of 47% were soxhlet extracted
using cyclohexane-­ethanol for 48 h and later with water for 48 h to remove waxes
and other natural impurities. The dewaxed fibers were oxidized by treating with
chlorine dioxide (ClO2) and later combined with furfuryl alcohol and heated at
100 °C for 4 h in the presence of nitrogen for the grafting to complete (Saw et al.
2011a, b). Grafting resulted in the formation of a smooth surface on the fibers and
also considerable increase in contact angle. Untreated coir fibers had contact angle
of 71°, which increased to 87° after grafting with the alcohol. However, the contact
angle of the grafted fibers was lower for glycerol, ethylene glycol, and 1-bromo-
naphthalene (Saw et al. 2011a, b). As observed in other studies, grafted fibers had
higher strength and modulus (50% increase) with decrease in elongation (40%
decrease) (Saw et al. 2011a, b).
In another study, MMA was grafted onto chemically treated coir fibers using
copper sulfate (CuSO4) and potassium periodate (KIO4) as the initiators. Grafting
was done at different temperatures to obtain the desired grafting ratio (Rout et al.
1999). Optimum percent grafting of about 20 was obtained at 50 °C. Changes in
tensile properties of the fibers were dependent on the percentage of grafting. An
increase in FTIR peak at 1729 cm−1 with increase in percentage of grafting con-
firmed successful grafting of MMA onto coir fibers. Grafted coir fibers had strength
of 223 MPa compared to 111 and 108 for the 18% and ungrafted fibers, respectively.
Modulus of the fibers also showed a similar trend compared to strength (Rout et al.
1999). To further improve the properties of coir fibers grafted with methyl acrylate
in the presence of methanol, fibers were subject to gamma irradiation and also
exposed to UV. A gamma dose ranging from 250 to 1000 krad was used at a rate of
350 krad/h using CO60 gamma source. Alternatively, fibers were also treated with
UV irradiation (254–313 nm) for 20–25 h. After treatment, the fibers were soaked
in the monomer solution and cured for the grafting to occur (Zaman et al. 2012).
Tensile strength and elongation of the fibers were dependent on the monomer con-
centration and curing time. Strength factor showed variations from 1.3 to 1.7 and
elongation factor from 1.3 to 1.9 for grafting percentage ranging from 2.3 to 5.6%
(Zaman et al. 2012). Compared to these properties, the gamma irradiated sample
had strength and elongation factor of 2.15 and 2.4, respectively, and UV-treated
sample was 2.3 and 2.5, respectively. In addition, grafting the fibers after UV treat-
ment provided higher resistance to weathering and water (Zaman et al. 2012). In
further continuation of the study, coir fibers were also grafted with 2-hydroxyethyl
1.4 Chemical Modifications of Coir Fibers 23

Table 1.15 Effect of urea on the grafting percentage (G) and tensile properties (Strength (TS),
modulus (TM) of coir fibers (Zaman et al. 2013))
Concentration of urea, %
0.5 1 2
TM TS TM TM
Monomers % G TS (MPa) (MPa) %G (MPa) (MPa) %G TS (MPa) (MPa)
20% HEMA 24.4 173 ± 3 821 ± 14 26.5 176 ± 2 833 ± 12 25.3 174 ± 2.6 826 ± 15
30% MMA 14.3 153 ± 1.8 748 ± 15 16.8 156 ± 3 756 ± 23 15.4 154 ± 2.5 752 ± 19
25% 2-HEA 18.2 164 ± 2 786 ± 25 20.3 166 ± 2 795 ± 18 19.2 164 ± 3 792 ± 21
Reproduced with permission from Springer Nature

methacrylate (HEMA), methyl methacrylate (MMA), and 2-hydroxyethyl acrylate


(2-HEA) under UV radiation and presence of 0.5–2% of urea (Zaman et al. 2013).
Tensile strength and modulus increased for all grafted samples for a UV exposure
of up to 20 passes above which the properties decreased (Table 1.15). Amount of
urea decreased the grafting percentage and also affected tensile properties (Zaman
et al. 2013). It was suggested that grafting under UV was environmentally friendly
and could provide fibers with good properties.
Similar to MMA, acrylonitrile was also grafted onto coir fibers using CuSO4 and
NaIO4 at a temperature range of 50–70 °C (Rout et al. 2002). Grafting (6.5 or 10%)
of acrylonitrile increased tensile strength from 108 to 148 MPa at 6% grafting but
strength decreased to 97 MPa when the grafting was 10%. Elongation of the fibers
had a considerably large decrease from 15 to 9%, and moisture regain decreased
from 39 to 23%. SEM studies showed that grafting occurred not only on the surface
but also on the inside of the fibers. Presence of acrylonitrile was also responsible for
the increase in hydrophobicity of the fibers (Rout et al. 2002), which could help in
increasing the compatibility with synthetic polymers for composite and other appli-
cations. Alkali-treated coir fibers were subject to various chemical modifications
including crosslinking and cyanoethylation. Crosslinking was done using formalde-
hyde, p-phenylene diamine, and phthalic anhydride (Samal et al. 1995). There was
a decrease in moisture regain of the chemically modified coir and the extent of
decrease was dependent on the type and conditions used during modification.
Tensile strength of the fibers varied from 130 to 201 kg;/cm2 with p-phenylene
diamine (PPDA) providing the highest strength to the fibers (Samal et al. 1995).
Crosslinking also increased the resistance of the fibers to dissolution in various
chemicals. The modified fibers also had lower solubility in most solvents including
hydrogen peroxide (Table 1.16). A new process of biografting using laccase from
Trametes versicolor was used to graft natural phenol syringaldehyde onto the sur-
face of coconut fibers (Thakur et al. 2015). For grafting, the fibers were treated in
3.5% of syringaldehyde in 40 mM citrate buffer and 40 U/g of laccase and incubated
at 50 °C for 12 h. A schematic representation of the grafting process is shown in
Fig. 1.14. A maximum biografting of 15.6% was obtained when 4.5% syringalde-
hyde was used. Although good grafting was achieved, there was no major effect on
the crystallinity but the thermal stability of the fibers was increased. Syringaldehyde
24 1 Processing and Properties of Coconuts

Table 1.16 Solubility (%) of modified coir fibers after various chemical modifications (Samal
et al. 1995)
Sample H2SO4 HCl NaOH NH4OH NaCl H2O2
Coir 9.97 10.07 11.73 9.47 4.57 14.22
Coir-ONa 14.02 13.97 9.52 2.34 8.20 11.25
Coir-c-CH2O 7.92 7.68 8.12 4.21 2.78 11.21
Coir-c-PPDA 8.47 8.05 7.74 5.72 1.00 15.12
Coir-c-PhA 5.62 4.73 5.82 3.86 2.15 13.15
Reproduced with permission from John Wiley and Sons

Fig. 1.14 Depiction of the mechanism of laccase catalyzed biografting of eugenol on coconut
fibers (Thakur et al. 2015)

is known for its antioxidant, antimicrobial properties. Hence, grafted fibers did
show high inhibition of both E. coli and S. aureus. It was expected that the grafting
would also improve the adhesion between fibers and matrix and could provide com-
posites with higher mechanical properties (Thakur et al. 2015). In a similar approach,
ferulic acid was grafted onto lignin in coir fibers using laccase as the catalyst to
improve the thermal, antibacterial, and hydrophobic properties (Thakur et al. 2016).
A high grafting ratio of 54% was obtained using 4.5% ferulic acid. Surface of the
fibers had become smoother, and interaction between the fibers had also increased.
Ferulic acid also imparted antimicrobial properties to the coir fibers.
Instead of grafting, treating coir fibers with chromous sulfate (CrSO4) and/or
NaHCO3 smoothened the fiber surface and improved adhesion. Treating with both
the chemicals provided better properties than just with CrSO4. Considerable increase
in strength and modulus but decrease in elongation was observed (Mir et al. 2012).
A targeted grafting of furfuryl alcohol was done onto coir fibers after treating with
chlorine dioxide that could preferentially oxidize the guaiacyl and syringyl units of
1.5 Properties of Coir Pith 25

lignin (Saw et al. 2011a, b). Grafting was done to improve the interfacial adhesion
between the fibers and composite matrices. Oxidation increased the roughness of
the fibers by more than 100% but grafting decreases the roughness and makes the
fiber surface similar to that of the untreated fiber. Grafting also increases the hydro-
phobicity for several liquids as evident from the larger contact angles. Tensile
strength of the grafted fibers was 196 MPa compared to 145 for the untreated fibers.
Although modulus also showed similar increase (4871 MPa from 3101 MPa), the
elongation of the fibers was only 19.7% compared to 32% before treatment (Saw
et al. 2011a, b). Grafting furfuryl alcohol was suggested to make coir fibers suitable
for outdoor applications. In another study, coir fibers were mercerized and also
treated under UV radiation before grafting ethylene dimethylacrylate (Rahman and
Khan 2007). The UV-treated fibers had increased tensile strength due to intercross-
linking between the polymers. Alkali treatment increased hydrophobicity, surface
tension and also improved the tensile properties of the grafted fibers. Surface and
intercellular gaps were filled with the grafted polymer, which made the fibers
smooth and uniform.
Coir fibers were subjected to transesterification by refluxing with butylacrylate in
acetone and pyridine as catalyst at 50 °C for 6 h. After the reaction, the fibers were
washed with dilute acetic acid and acetone, methanol, and finally with distilled water
(Rout et al. 2018). Further modification of the fiber was done by refluxing the trans-
esterified coir with benzoyl peroxide (BPO) in acetone at 50 °C for 2 h. Both the
untreated (parent) and treated fibers were coated with silver to improve thermal sta-
bility. Silver coating was done by treating the fibers with silver salt of AgNO3 (0.1 M
and 0.2 M, separately) for 0.5 h under dark at 28 °C and later washing the fibers with
water and drying in a hot air oven. Silver nanoparticles of 40–210 nm were deposited
on the fibers. Fibers which were more hydrophobic had larger deposits of the parti-
cles. Also, transesterified fibers treated with BPO had more even distribution and
coating. Modified fibers had lower mass loss below 150 °C, and decomposition of
the fibers occurred between 278 and 295 °C (hemicellulose), 312 and 337 °C (cel-
lulose), and 338 and 439 °C (lignin) (Table 1.17). Among all the fiber treatments,
fibers modified with BPO and coated with the silver nanoparticles had the highest
stability. The process of modifying the fibers was suggested to be suitable for devel-
oping conducting polymers, electronic and optical devices, optical sensors, and
other applications requiring high temperature stability (Rout et al. 2018).

1.5 Properties of Coir Pith

Coir pith is the residue generated during the processing of coir fibers. Typically,
about 2 tons of pith is generated for every ton of fibers obtained (Narendar and Priya
Dasan 2014). Although large amount of pith is available, the high lignin content
makes pith difficult to degrade. Currently, pith is mainly used as fertilizer or as soil-
less potting media. The high moisture sorption (up to 200% its weight) also makes
pith to have poor compatibility with matrices and hence difficult to obtain good
properties. Various chemical modifications (Fig. 1.15) were done to improve the
26

Table 1.17 Thermal properties of coir fibers subject to various chemical modifications (Rout et al. 2018)
Char
Tmax-1 Tmax-2 Tmax-3 T final Residual weight, % Weight yield at
500 °C,
Properties 278–295 °C 312–337 °C 338–439 °C 454–694 °C Tmax-­1 Tmax-­2 Tmax-­3 at 150 °C %
Raw coir – 330 385 454 – 52 34 14.5 2.2
Parent coir 278.7 312 406 493 73 62 41 14.0 16.7
Ag NPs coated parent coir with 285 315 400 500 73.7 60.7 36.3 13.7 5.2
0.1 M AgNO3
Ag NPs coated parent coir with 283 315 404 583 71.1 63.3 40 15.2 17.0
0.2 M AgNO3
Modified coir-1 287 317 400 500 72.2 61.8 36.3 13.8 8.5
Ag NPs coated modified coir – 337 339 504 – 54.1 36.7 12.6 6.9
with 0.1 M AgNO3
Ag NPs coated modified coir 282 322 439 609 80 67.9 30 14.2 7.9
with 0.2 M AgNO3
Modified coir 2 295 316 405 520 76.7 65.7 37.7 9 14.5
Ag NPs coated modified coir 289 313 396 526 72.6 62.2 33.7 12.9 5.5
with 0.1 M AgNO3
Ag NPs coated modified coir 283 313 415 694 74.4 66.7 43.7 12.1 24
with 0.2 M AgNO3
Reproduced with permission from Springer Nature
1 Processing and Properties of Coconuts
1.5 Properties of Coir Pith 27

Fig. 1.15 Schematic depiction of the various treatments used to modify coco pith (Narendar and
Priya Dasan 2014). Reproduced with permission from Elsevier

Table 1.18 Changes in the composition of coir pith after various chemical treatments (Narendar
and Priya Dasan 2014)
Sample α-cellulose, % Hemicellulose, % Lignin, % Wax, % Moisture, %
Untreated 27.4 14.6 42 10.2 0.41
Sodium hydroxide 37.3 11.9 18.2 6.4 0.35
Dicumyl peroxide 38.6 10.9 18 4.5 0.31
Acrylic acid 39.6 10.5 17.8 4 0.29
Acetic acid 40.6 11.6 16.2 3.6 0.28
Sodium hypochlorite 41.6 9.9 15.4 3.5 0.32
Sulfuric acid 43.8 8.9 13.6 3.1 0.23
Reproduced with permission from Elsevier

properties, morphology, and compatibility of coir pith. These treatments led to con-
siderable increase in α-cellulose and corresponding decrease in hemicellulose and
lignin content (Table 1.18). In addition, the surface of the pith had become rougher
and had increased thermal stability. Treating coir pith with hydrogen peroxide (1%,
pH 11.5 for 8–24 h) and retreating with 2% peroxide also affected properties of the
coir pith (Rojith and Bright Singh 2012). Marginal decrease in crystallinity index
from 39 to 32 occurred due to the peroxide treatment. Increase in surface area and
accessibility after treatment made the fibers more suitable for processing into vari-
ous applications (Rojith and Bright Singh 2012).
28 1 Processing and Properties of Coconuts

References

Abraham E, Deepa B, Pothen LA, Cintil J, Thomas S, John MJ, Anandjiwala R, Narine SS (2013)
Environmental friendly method for the extraction of coir fibre and isolation of nanofibre.
Carbohydr Polym 92(2):1477–1483
Basu G, Mishra L, Jose S, Samanta AK (2015) Accelerated retting cum softening of coconut fibre.
Ind Crop Prod 77:66–73
Bismarck A, Mohanty AK, Aranberri-Askargorta I, Czapla S, Misra M, Hinrichsen G, Springer
J (2001) Surface characterization of natural fibers; surface properties and the water up-take
behavior of modified sisal and coir fibers. Green Chem 3(2):100–107
Brígida AIS, Calado VMA, Gonçalves LRB, Coelho MAZ (2010) Effect of chemical treatments on
properties of green coconut fiber. Carbohydr Polym 79(4):832–838
de Farias JGG, Cavalcante RC, Canabarro BR, Viana HM, Scholz S, Simão RA (2017) Surface lig-
nin removal on coir fibers by plasma treatment for improved adhesion in thermoplastic starch
composites. Carbohydr Polym 165:429–436
Fouladi MH, Ayub M, Nor MJM (2011) Analysis of coir fiber acoustical characteristics. Appl
Acoust 72(1):35–42
Gu H (2009) Tensile behavior of the coir fibre and related composites after NaOH treatment. Mater
Des 30:3931–3934
Jose S, Mishra L, Basu G, Samanta AK (2016) Study on reuse of coconut fiber chemical retting
bath. Part I: retting efficiency. J Nat Fibers 13(5):603–609
Khan GMA, Alam MS (2012) Thermal characterization of chemically treated coconut husk fibre.
Indian J Fibre Text Res 37:20–26
Lomelí-Ramírez MG, Anda RR, Satyanarayana KG, Bolzon de Muniz GI, Iwakiri S (2018)
Comparative study of the characteristics of green and brown coconut fibers for the develop-
ment of green composites. Bioresources 13(1):1637–1660
Mahato DN, Mathur BK, Bhattacherjee S (2013) DSC and IR method for determination of acces-
sibility of cellulosic coir fber and thermal degradation under mercerization. Indian J Fibre Text
Res 38:96–100
Manjula R, Raju NV, Chakradhar RPS, Johns J (2018a) Effect of thermal aging and chemical treat-
ment on tensile properties of coir fiber. J Nat Fibers 15(1):112–121
Manjula R, Raju NV, Chakradhar RPS, Kalkornsurapranee E, Johns J (2018b) Influence of chemi-
cal treatment on thermal decomposition and crystallite size of coir fiber. Int J Thermophys
39(1):3
Martinschitz KJ, Boesecke P, Garvey CJ, Gindl W, Keckes J (2008) Changes in microfibril angle
in cyclically deformed dry coir fibers studied by in situ synchrotron X-ray diffraction. J Mater
Sci 43:350–356
Mathura N, Cree D, Mulligan RP (2014) Characterization and utilization of coconut fibers of the
Caribbean. MRS Online Proc Lib Arch 1611:95–104
Mathura N, Cree D (2016) Characterization and mechanical property of trinidad coir fibers. J Appl
Polym Sci 133:43692
Mir SS, Hasan SMN, Hossain MJ, Hasan M (2012) Chemical modification effect on the mechani-
cal properties of coir fiber. Eng J 16(2):73–84
Mohato DN, Mathur BK, Bhattacherjee S (1995) Effect of alkali treatment on thermal stability and
moisture retention of coir fibre. Indian J Fibre Text Res 20:202–205
Musanif IS, Thomas A (2015) Effect of alkali treatments on physical and mechanical properties of
coir fibers. Chem Mater Eng 3(2):23–28
Narendar R, Priya Dasan K (2014) Chemical treatments of coir pith: morphology, chemical com-
position, thermal and water retention behavior. Compos Part B 56:770–779
Pandiselvam R, Manikantan MR, Kothakota A, Rajesh GK, Beegum S, Ramesh SV, Niral V,
Hebbar KB (2018) Engineering properties of five varieties of coconuts (Cocos nucifera L.) for
efficient husk separation. J Nat Fibers:1–9
References 29

Rahman MM, Khan MA (2007) Surface treatment of coir (Cocos nucifera) fibers and its influence
on the fibers’ physico-mechanical properties. Compos Sci Technol 67(11–12):2369–2376
Rajan A, Senan RC, Pavithran C, Abraham TE (2005) Biosoftening of coir fiber using selected
microorganisms. Bioprocess Biosyst Eng 28(3):165–173
Ravindranath AD, Bhosle S (2000) Development of a bacterial consortium for coir retting. JSIR
59:140–143
Renouard S, Hano C, Doussot J, Blondeau J-P, Lainé E (2014) Characterization of ultrasonic
impact on coir, flax and hemp fibers. Mater Lett 129:137–141
Rojith G, Bright Singh IS (2012) Delignification, cellulose crystallinity change and surface
modification of coir pith induced by oxidative delignification treatment. Int J Environ Bioene
3(1):46–55
Rout J, Misra M, Tripathy SS, Nayak SK, Mohanty AK (2002) Surface modification of coir fibers.
II. Cu (II)-IO initiated graft copolymerization of acrylonitrile onto chemically modified coir
fibers. J Appl Polym Sci 84(1):75–82
Rout J, Misra M, Mohanty AK (1999) Surface modification of coir fibers. I. Studies on graft
polymerization of methyl methacrylate on to chemically modified coir fibers. Polym Adv
Technol 10:336–344
Rout SK, Tripathy BC, Ray PK (2018) Significance of nano-silver coating on the thermal behavior
of parent and modified agro-waste coir fibers. J Therm Anal Calorim 131(2):1423–1436
Samal RK, Panda BB, Rout SK, Mohanty M (1995) Effect of chemical modification on FTIR
Spectr. I. Physical and chemical behavior of coir. J Appl Polym Sci 58:745–752
Saw SK, Sarkhel G, Choudhury A (2011a) Surface modification of coir fibre involving oxidation
of lignins followed by reaction with furfuryl alcohol: characterization and stability. Appl Surf
Sci 257:3763–3769
Saw SK, Sarkhel G, Choudhury A (2011b) Surface modification of coir fibre involving oxidation
of lignins followed by reaction with furfuryl alcohol: characterization and stability. Appl Surf
Sci 257(8):3763–3769
Setyanto RH, Diharjo K, Setyono P, Made Miasa I (2013) A preliminary study: the influence of
alkali treatment on physical and mechanical properties of coir fiber. J Mater Sci Res 2(4):80
Shibu VY, Ajit H, Manilal VB (2013) Closed retting; a green technology for controlling coir retting
pollution of backwaters. J Environ Res Dev 7(44):1523–1530
Silva GG, Souza DAD, Machado JC, Hourston DJ (2000) Mechanical and thermal characteriza-
tion of native Brazilian coir fiber. J Appl Polym Sci 76:1197–1206
Sreenivasan S, Iyer PB, Iyer KRK (1996) Influence of delignification and alkali treatment on the
fine structure of coir fibers (Cocos Nucifera). J Mater Sci 31:721–726
Suganya DS, Pradeep S, Jayapriya J, Subramanian S (2007) Bio-softening of mature coconut husk
for facile coir recovery. Indian J Microbiol 47(2):164–166
Thakur K, Kalia S, Kaith BS, Pathania D, Kumar A (2015) Surface functionalization of coconut
fibers by enzymatic biografting of syringaldehyde for the development of biocomposites. RSC
Adv 5(94):76844–76851
Thakur K, Kalia S, Pathania D, Kumar A, Sharma N, Schauer CL (2016) Surface functionalization
of lignin constituent of coconut fibers via laccase-catalyzed biografting for development of
antibacterial and hydrophobic properties. J Clean Prod 113:176–182
Tran LQN, Minh TN, Fuentes CA, Chi TT, Van Vuure AW, Verpoest I (2015) Investigation of
microstructure and tensile properties of porous natural coir fibre for use in composite materials.
Ind Crop Prod 65:437–445
Valášek P, D’Amato R, Müller M, Ruggiero A (2018) Mechanical properties and abrasive wear of
white/brown coir epoxy composites. Compos Part B 146:88–97
Varghese A, Jacob J (2017) A study of physical and mechanical properties of the Indian coconut
for efficient dehusking. J Nat Fibers 14(3):390–399
Varma DS, Varma M, Varma IK (1984) Coir fibers: Part I: effect of physical and chemical treat-
ments on properties. Text Res J 54:827–832
30 1 Processing and Properties of Coconuts

Varma DS, Varma M, Varma IK (1986) Thermal behavior of coir fibers. Thermochim Acta
108:199–210
Yu W, Xie W, Du Z (2016) Structure of the right-handed helical crystal ribbon and multilevel fibrils
in a tube fiber from a coir fiber. Cellulose 23(5):2841–2852
Zaman HU, Khan MA, Khan RA, Ghoshal S (2012) Effect of ionizing and non-ionizing preirradia-
tions on physico-mechanical properties of coir fiber grafting with methacrylate. Fibers Polym
13(5):593–599
Zaman HU, Khan RA, Khan MA, Beg MDH (2013) Physico-mechanical and degradation properties
of biodegradable photografted coir fiber with acrylic monomers. Polym Bull 70(8):2277–2290
Chapter 2
Agricultural Applications of Coir

2.1 Introduction

One of the largest consumption of coir and its by-products is for agricultural appli-
cations. Examples of such applications include mulching to prevent erosion, as fer-
tilizer for plants, soilless growing media, etc. It is surprising to know the extent to
which coir is applied for agricultural applications. Extensive studies are still being
conducted to study and develop coir for newer agricultural applications. From bed-
ding for poultry farms, as substrate for mushroom cultivation, soil amendment,
mulch, and compost are the most common agricultural applications of coir (Prabhu
and Thomas 2002). Coir is preferred as compost due to its neutral pH and low elec-
trical conductivity. Coir dust was considered to have good moisture absorption and
degradability and hence considered as replacement for Canadian sphagnum peat for
growing Dieffenbachia maculate in a greenhouse. The coir dust was combined with
three different compost mixes, and the electrical properties and plant growth were
monitored. Adding coir substantially increased the water holding capacity of the
peat. Air- and water-filled pore space varied with changing coir content (Stamps and
Evans 1997) (Table 2.1). Cornell media supplemented with coir provided the big-
gest, heaviest, and highest grade plants although higher level of potassium in coir
could have affected plant growth. It was concluded that coir dust was an adequate
substitute for sphagnum peat and could be used for growing D. maculate. Similar
results were also obtained when coir dust was considered as replacement for sphag-
num peat for the production of Spathiphyllum. In addition, coir dust was reported to
reduce spathiphyllum root damage due to Cylindrocladium spathiphylli (Stamps
and Evans 1999). Physical, chemical, and morphological features of coir dust also
affected properties as soilless media. Coir dust obtained from different sources also
had significant variations compared to sphagnum peat (Fornes et al. 2003) as seen
from Table 2.2. However, the surface porosity (about 40%), internal porosity (about
40%), and pore diameter (44 μm) were similar for the two coco peats but different
than that of the sphagnum peat. It was suggested that the pore size was the major

© Springer Nature Switzerland AG 2019 31


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_2
32

Table 2.1 Physical characteristics of growing media containing coir dust and other materials
Bulk density (g/ Air-filled pore Water-filled pore Water holding
Components (% by vol) cm−2) space (% by vol) space (% by vol) capacity (% by mass)
Mix Peat Coir Vermiculite Pine bark Perlite A B A B A B A B
Cornell 50 0 25 0 25 0.11 0.14 23.6 13.1 62.1 75.4 575 554
0 50 25 0 25 0.10 0.13 11.6 3.6 64.6 86.2 650 692
Hybrid 40 0 30 30 0 0.16 0.19 27.7 8.5 54.4 66.2 344 342
0 40 30 30 0 0.12 0.15 21.3 5.5 62.6 83.4 538 536
UF-2 50 0 0 50 0 0.13 0.15 21.9 8.7 57.7 70.9 430 449
0 50 0 50 0 0.14 0.14 13.0 7.6 65.1 77.0 480 565
Significance
GM ∗∗∗ ∗∗∗ ∗∗∗ ∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗
Peat vs. ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗
2

coir
GM × P/C ∗∗∗ ∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ ∗∗∗ NS
Changes in the composition during plant implantation and after 5 months are listed below (Stamps and Evans 1997). Reproduced from American Society for
Horticultural Science
A before treatment, B after treatment, NS non-significant
∗Significant at P ≤ 0.05; ∗∗P ≤ 0.01; ∗∗∗P ≤ 0.001
Agricultural Applications of Coir
2.1 Introduction 33

Table 2.2 Comparison of the properties of coir dust from Mexico (M) and Sri Lanka (S) with
sphagnum peat obtained from Finland (Fornes et al. 2003)
Water Total water
Coarseness Total pore Air Easily buffering holding
index, % space, % content, available capacity, % capacity, %
Material w/w v/v % v/v water, % v/v v/v w/w
Coir 62.4 97.4 83.3 2.6 0.3 520
dust-M
Coir 31.5 94.1 31.7 22.5 5.3 668
dust-S
Peat, 62.9 94.2 41.2 22.5 4.4 739
Finland
Reproduced with permission from CSIRO Publishing

contributing factor for the differences in the structure of the peat and coir dust. In a
more exhaustive study, coir dusts from six different countries were compared for
their structure and properties against the conventional peat (Abad et al. 2002).
Although the physico-chemical properties of the coir dust varied significantly
between the sources, these variations were not considered to be of practical concern.
However, the differences in the concentration of sodium and chloride ions could
affect some of the salt sensitive plants. Some varieties of dust from the same country
showed large differences in electrical conductivity (Table 2.3).
The effect of salt stress in pepper plants grown using coconut coir dust in a con-
trolled polyhouse environment was studied by Rubio et al. (2010). The extent of
salinity, season of producing the pepper plants, and irrigation frequency were varied
to understand the changes in production. It was observed that the yield and fruit
quality were not affected due to the presence of coconut dust. Electrical conductiv-
ity is considered as one of the important parameters to evaluate the performance of
coir-based products for agricultural applications since the capacitance-based soil
sensors are able to determine the volumetric water content and predict quality. The
changes in the physical properties of coconut coir dust combined with various per-
lite mix was studied (Rhie and Kim 2017). As can be observed from Table 2.4,
substrates containing higher amounts of perlite had larger particles, water holding
capacity, and electrical conductivity. However, the optimum conductivity of water
and air should be decided based on the type of crop and climatic condition in the
particular region. In addition, it has also been reported that particle size influences
the physical and chemical properties of coir pith as a soilless medium (Jeyaseeli and
Raj 2010). For instance, electrical conductivity decreased with increasing particle
size and higher cation exchange (168 meq/100 g) occurred in pith containing
925 μm particles. Ash and phenol contents also changed considerably with chang-
ing particle size from 925 to 1100 μm (Jeyaseeli and Raj 2010). Similarly, pH lower
than 4 makes the micronutrients to be mobile and absorbed in excess compared to
pH greater than 9 leading to deficiency. Coir pith having particles with size ranging
from 800 to 1112 μm and cation exchange capacity above 140 meq/100 g was found
34 2 Agricultural Applications of Coir

Table 2.3 Variation in the properties of coir dust obtained from different countries and in different
varieties compared to peat (Abad et al. 2002)
Electrical Cation exchange Organic matter C/N
Source pH conductivity (mS/m) capacity (cmol/kg) content, % dry weight ratio
Costa Rica
– CR1 5.95 293 69.8 94.2 149
– CR2 5.89 39 60.4 95.4 117
Ivory Coast
– IC1 5.83 230 38.8 96.9 113
India
– IN1 6.14 166 89.2 90.7 110
Mexico
– ME1 5.74 597 52.5 91.5 110
– ME2 5.73 567 37.8 93.6 113
– ME3 5.98 387 35.8 94.2 137
– ME4 5.68 449 31.7 93.7 134
Sri
Lanka
– SL1 4.90 247 95.4 89.3 97
– SL2 5.06 70 92.9 94.4 75
Thailand
– TH1 5.46 451 70.4 94.4 160
– TH2 5.26 296 50.1 93.4 177
– TH3 5.16 482 70.6 93.5 186
Peat 3.17 21 99.6 97.9 48
Reproduced with permission from Elsevier

Table 2.4 Properties of coconut coir dust mixed with perlite in various ratios (Rhie and Kim
2017)
Coir Air Bulk Electrical
dust:perlite Container space, Total density, conductivity
(V/V) capacity, % % porosity, % g/cm3 pH (dS/min)
1:0 64.3 26.4 90.7 0.06 4.99 0.84
8:2 60.6 27.0 87.5 0.08 5.12 0.76
6:4 58.0 26.7 847 0.10 5.24 0.52
4:6 53.5 26.3 79.9 0.12 5.46 0.41
2:8 38.4 39.9 78.3 0.13 6.62 0.18
0:1 24.5 45.0 69.5 0.14 8.74 0.06
Reproduced with permission from American Society for Horticultural Science

to be ideal for horticultural applications (Jeyaseeli and Raj 2010). In a study on the
effect of coco peat and perlite on the quality and antioxidant properties of bell pep-
per, a substantial increase in amounts of vitamin C concentrations was observed.
The coir peat mixture used had pH of 6.4, electrical conductivity of 11 mS/cm,
2.1 Introduction 35

particle density of 1.4 g/cm3, and total pores of 91%. Also, different varieties of bell
pepper were found to have different levels of increase in the phenolic contents,
DPPH activity, etc. (Aslani et al. 2016). Vitamin concentrations ranged from 100 to
211 mg/g, DPPH radical scavenging ability increased from 0.3 to 1.1 μmol BHT
equivalents, and the best ratio of coir pith to perlite was found to be 80:20. Various
organic matters were added as a supplement to coconut coir dust used as soilless
growth medium. Empty fruit bunches, peat moss, and burnt rice husk were com-
bined in various ratios with coir, and the development of rock melon was monitored.
The combination of 70% coir dust and 30% empty fruit bunches provided the best
plant growth, leading to a rock melon having diameter of 14.2 cm, fruit weight of
1483 g, and total soluble solids of 15.3% (Wira et al. 2011).
Although coir is extensively used for agricultural applications, the high lignin
content and carbon/nitrogen (C/N) ratio reduced biodegradability. In order to
increase the biodegradation of coir, cow manure was combined and the changes in
C/N ratio and other chemical parameters were monitored (Tripetchkul et al. 2012).
The C/N ratio largely affected the organic matter and total nitrogen losses, which
did not show much effect on temperature or pH of the compost pile. Highest rate of
biodegradation (k = 0.309/day) and total nitrogen loss was highest for the 25 C/N
ratio, whereas highest nitrogen loss of 7.3% was observed in the samples containing
C/N ratio of 30 and was considered ideal for coir pith composting (Tripetchkul et al.
2012) (Table 2.5).
Although coir pith is extensively used as compost, the rate of degradation of coir
is considered to be slow. Attempts have been made to increase the rate of composit-
ing of coir (Ghosh et al. 2007). Coir pith was treated with a bioformulation contain-
ing mushrooms and placed in layers along with urea. Appropriate mechanism was
made to provide necessary oxygen for the composting. Moisture in the mixture was
maintained at 200% and the material was exposed to the environment for 30–45 days
until the pith had become black. The compost or manure obtained was enriched in
nitrogen, phosphorus, and potassium, and the entire process of compositing was
completed in 21 days (Ghosh et al. 2007). Similar studies were also conducted using
a blend of coconut coir with coco peat. It was found that coir was less acidic and
electrical conductivity increased with increasing coir content. It was suggested that

Table 2.5 Evaluation of coir pith quality based on the C/N ratio after 35 days of degradation
(Tripetchkul et al. 2012)
Standard organic
Parameter C/N ratio 1:30 C/N ratio 1:25 C/N ratio 1:20 composite
Nitrogen (%) 2.17 ± 0.10 2.15 ± 0.30 2.13 ± 0.30 >1.0
Phosphorous (%) 0.16 ± 0.002 0.240 ± 0.002 0.18 ± 0.002 >0.5
Potassium (%) 2.49 ± 0.30 3.03 ± 0.30 2.32 ± 0.10 >0.5
pH 7.10 ± 0.00 7.30 ± 0.00 6.90 ± 0.00 7.0–8.0
Organic matter (%) 66.20 ± 3.40 67.90 ± 1.40 63.70 ± 0.10 25–50
C/N ratio 17.00 ± 0.90 17.60 ± 0.40 16.40 ± 0.20 0–20:1
Reproduced using open access from Springer Nature
36 2 Agricultural Applications of Coir

adding coir to peat up to ratio of 1:1 will significantly increase crop yield, particu-
larly shoot and root growth (Vavrina et al. 1996). However, addition of coir did not
promote the growth of tomato plants. In another study, it was shown that a combina-
tion of coir, peat, and vermiculite promoted better growth of tomato transplants.
This study also suggested that a 50:50 ratio of vermiculite and coir was more appro-
priate for growing the transplants and higher coir content led to decrease in nitrogen
content and hence inhibited plant growth (Arenas et al. 2012). In a unique study,
ability of various mulch materials to amend bacterial and fungal growth in forest fire
burnt soil was studied (Barreiro et al. 2016). Coconut fiber had relatively high car-
bon/nitrogen ratio (271) than the other mulches and also provided less respiration to
the soil. However, there was considerably low bacterial growth on coconut contain-
ing soil during the initial 2 weeks of incubation but stimulated growth was observed
after prolonged periods. Similarly, fungal growth was also lower on coconut fiber
supplemented soil due to the lower availability of carbon (Barreiro et al. 2016).
These results could be useful to design soil erosion prevention systems.
In addition to coir, chars obtained from forest waste and olive mill waste were
mixed together with coir and used to cultivate tomatoes. A combination of coir and
oil mill waste showed increased electrical conductivity from 1 to 3.5 ds/m and mois-
ture content increased from 35 to 70%. Yield of tomato was dependent on the dose
percentage and varied from 2.9 to 5.8 kg/plant. Similarly, fruit weight was also var-
ied between 117 to 131 g and 177 to 216 g depending on the dosage and variety of
the tomato plant (Fornes et al. 2017). It was suggested that particle size of the
growth substrate was also important in addition to the percentage and ratio of the
two components.
In another study, coconut coir dust was combined with peat, oil palm frond as
soilless growth media for cultivation of cauliflower. Combination of coir dust and
oil palm frond was observed to provide the best output with a cauliflower curd
weight of 302 g/plant. Also, plants receiving the combination showed earlier matu-
rity and considerable increase in chlorophyll content from week 2 to week 8.
Electrical conductance values were within acceptable limits throughout the growth
period and hence the dust and frond were considered suitable for plant growth
(Erwan et al. 2013). Supplementing growing media with other constituents has been
found to benefit plant growth. For instance, coconut dust used as growing media
was supplemented with biochar prepared by pyrolyzing rice hull at 500 °C. The
biochar (pH 10.2, electrical conductivity of 0.82 dS/m) was mixed with growing
media in 1, 2, and 5% and used to grow kale (Brassica oleracea L) for 25 days (Kim
et al. 2017). Both the physical and chemical properties of the growing media were
improved due to the addition of the biochar (Table 2.6). Consequently, considerable
increase in nutrient retention and dry weight of shoot and root was observed
(Fig. 2.1). It was suggested that adding biochar would be useful as a supplement to
improve the productivity and properties of crops (Kim et al. 2017). In another study,
sewage sludge that is rich in nitrogen, phosphorous, and trace elements such as
cadmium, lead, copper, nickel, or zinc was combined with coir pith and studied for
growing forage maize (Somasundaram et al. 2016). Since application of only sludge
increases metal contaminants in the plants, a 50/50 ratio of coir and sludge was
2.1 Introduction 37

Table 2.6 Physical properties of the coir dust supplemented with various ratios of biochar (Kim
et al. 2017)
Biochar, Bulk density, Particle Total Water Air
% kg/m3 density, kg/m3 pore space, % volume, % capacity, %
0% 158 ± 0.9 2072 ± 3.7 92.4 ± 0.04 63.5 ± 0.94 28.9 ± 0.95
1 162 ± 0.2 2053 ± 11.5 92.1 ± 0.04 67.1 ± 1.20 25.0 ± 1.21
2 164 ± 0.9 2040 ± 8.7 92.0 ± 0.04 71.4 ± 0.39 20.6 ± 0.38
5 174 ± 2.9 2031 ± 7.1 91.4 ± 0.15 73.2 ± 1.15 18.2 ± 1.29
Reproduced with permission from Springer Nature

Fig. 2.1 Increase in the dry weight of the shoot and root of kale due to the addition of 0, 1, 2, or
5% of biochar (Kim et al. 2017). Reproduced with permission from Springer Nature

combined and added in different quantities to pots. Considerable decrease in the


amount of trace elements and increase in the dry matter yield up to 55% was
observed when 4.8 g of the mixed material was added per pot. Electrical conductiv-
ity was 0.23 dS/m for the control and increased to 0.37 dS/m for the mixture con-
taining 4.8 g of media (Somasundaram et al. 2016). Combining coir pith and sludge
would also decrease the risk of heavy metal contamination in plants.
Other than the normal plants, coir pith was also used to improve the yield and
properties of the medicinal plant Coleus forskohlii. Three different levels (5, 10, and
15 tons per ha) of coir pith were added as the fertilizer. As seen from Table 2.7, the
plant height, stem girth, and even the weight of the tuber were considerably higher
38

Table 2.7 Increased plant parameters after adding different levels of coir pith (Reghuvaran and Ravindranath 2010)
Carbohydrate,
Plant height, Stem girth, Lamina length, Lamina breadth, Tuber length, Fresh weight, mg/g Protein, mg/g
Treatment cm cm cm cm cm g Leaves Tubers Leaves Tubers
Untreated 24.8 5.1 4.04 3.3 7.6 92.3 2.03 6.12 1.80 5.15
Treated, 5t/h 30.6 5.8 5.20 4.22 13.2 145.1 2.34 6.97 2.24 6.30
Treated, 34.4 7.4 6.40 5.18 17.4 246.5 2.65 7.63 2.45 6.88
10t/h
Treated, 37.5 8.3 6.88 5.8 19.1 309.9 3.21 8.10 2.69 7.03
2

15t/h
Reproduced with permission from National Institute of Science Communication and Information Resources
Agricultural Applications of Coir
2.2 Coir Mulching 39

Table 2.8 Comparison of the physical and chemical parameters of coir with sedge and sphagnum
based peat (Meerow 1994)
Electrical
Water holding conductivity
Air space, % capacity, % pH (ds.m−1)
Medium Initial Final Initial Final Initial Final Initial Final
Coir 13.7 11.0 35.7 39.2 5.6 6.3 3.1 1.6
Sedge 23.1 9.7 29.8 45.4 5.6 6.6 2.4 1.7
Sphagnum 14.5 8.5 36.9 43.4 4.9 6.1 2.6 1.4
Reproduced with permission from American Society for Horticultural Science

on the plants that had the coir pith as fertilizer (Padamadevi et al. 2016). Similarly,
the content of the medicinal component forskohlii had marginally increased from 3
to 3.5 wt%. In another study, coir pith was supplemented to garden soil to increase
the carbohydrate and chlorophyll contents of A. paniculata. It was suggested that
addition to coir pith along with bacteria could support the development of medicinal
plants (Reghuvaran and Ravindranath 2010).
A mixture of coir peat, sand, and soil in various ratios was considered for the
cultivation of hyacinth. Considerably higher plant growth was observed when soil
was substituted by coco peat. Better rate of photosynthesis and increased water use
efficiency were provided by coco peat (Nazari et al. 2011). A comparison was done
to understand the potential of using coconut coir as a substitute for peat to grow
various native Australian plants. Different ratios of coir, peat, perlite, and sand were
prepared, and the growth of plants such as Pultenaea parviflora, Eucalyptus mel-
liodora, and Callicoma serratifolia was studied. After 14 months of study, it was
concluded that there was no major difference between the different potting mixtures
studied and coir was suitable to replace peat as a soilless growth medium (Offord
et al. 1998). Root and shoot regrowth was similar with root length ranging from 139
to 155 mm and total dry root weight between 6.9 and 13.4 mg. Contrary results have
also been published by Meerow on the ability of coir pith as a substitute to peat for
the growth of Pentas lanceolata and Ixora coccinea L. (Meerow 1994). Growth
index and top and root dry weight for both crops were found to be significantly
higher in the coir-based medium compared to the sedge peat based medium for
lanceolata, whereas they were considerably lower for Ixora. Coir-based medium
maintained comparatively less variations in physical and chemical properties
(Table 2.8) throughout the study and was hence considered suitable to replace
sphagnum or sedge peat based soilless growth media (Meerow 1994).

2.2 Coir Mulching

Mulching is a common approach to avoid weeds, retain moisture, and increase crop
yield. Coir in various forms and blends has been considered as an ideal material for
replacing plastic-based mulching. Rubberized coir, black and needled felted coir
40 2 Agricultural Applications of Coir

were compared with transparent polyethylene as mulching for two different crops.
Results showed that the plant growth and yield on rubberized coir mulching was
similar to that of transparent polyethylene for both pineapple and bhindi crops.
However, coir-based mulching will be biodegradable and will not pose environmen-
tal issues in disposal and hence preferable over plastic mulching (Beena and Anil
2011). Considerable increase in the production of cotton was noticed when coir pith
was used as the mulch under drip irrigation (Ramesh et al. 2006). Coir mulched
plots were able to retain higher soil moisture and helped enhance plant height and
consequently yields by about 15%. It was suggested that coir mulching with drip
irrigation could lead to substantially higher income to cotton farmers (Ramesh et al.
2006). Both coir dust and saw dust prevented water evaporation and reduced water
stress in tomato plants cultivated at different temperatures (Dishani and Silva 2016).
Considerable increase in yield was observed in the mulched plots even at 34 °C,
which was also reflected in the size of the tomatoes obtained (Fig. 2.2). It was sug-
gested that coir-based mulching could be a viable approach to obtain high yields
and decrease water consumption, particularly as the global temperatures keep
increasing in future years.
Similar results were also reported when coir dust was used as mulching for the
growth of groundnuts exposed to temperature stress (Aiome and Silva 2014).
Although pod yield decreased with increasing temperature, the plants protected
with coir mulch had higher yield than the unprotected and straw and saw dust
mulched plants (Fig. 2.3). Higher cation exchange ability of coir dust (50 meq/100 g
dry matter) and better nutrient binding capacity due to considerably lower nitrogen
leaching were suggested to be the factors for the higher output from plants mulched
with coir dust (Aiome and Silva 2014). Considerable decrease in evapotranspiration
and about 10.4% increase in maize productivity were achieved when coir dust was
used as mulch (Muti et al. 2017). A 10 cm thick layer of coir dust was sufficient to

Fig. 2.2 Digital image showing the difference in size of tomatoes with the addition of the coir
mulching and with and without stress (Dishani and Silva 2016). Ambient conditions with mulching
(a), 32 °C without mulch and no water stress (b) and 34 °C with mulch and no water stress.
Reproduced under open access license from Open University of Sri Lanka
2.2 Coir Mulching 41

90
80
70
60
pod Yield (g)

50 No mulch
40 Coir dust
30 Straw
20 Saw dust
10
0
Ambiant Tem 32°C Max tem 34°C Max tem

Fig. 2.3 Influence of temperature and type of mulch on yield of groundnuts (Aiome and Silva
2014). Reproduced with open access license from Open University of Sri Lanka

control the soil conditions and achieve significant increase in yield of PH7 maize.
Mulching was considered to be necessary to minimize the effect of temperature
stresses and future changes due to global warming. Various types of biomass includ-
ing coir pith and dried coconut leaves were considered as mulches for cultivation of
ginger (Zingiber officinale Roscoe). Although not as efficient as paddy straw in
terms of reducing weeds and increasing yield, dried coconut leaves had ­considerably
higher benefit cost ratio of 2.0 and yield increase was about 11% (Thankamani et al.
2016). Coir pith used as mulch at a rate of 20 tons per ha provided increased straw-
berry weight of 9.6 g compared to 8.3 g without mulch. However, the yield and
weight of strawberries obtained using organic mulches including coir was substan-
tially lower than that of polyethylene mulches (Mathad and Jholgiker 2005).
Pineapple was grown using organic and polyethylene based mulches. Although
coir dust was able to retain higher levels of moisture, the yield per hectare was lower
compared to polyethylene mulch (Alwis and Herath 2012). It was suggested that the
low cost of coir could be an advantage to use as mulch despite the lower fruit pro-
ductivity. However, contradictory results were reported by Martinez et al. who
showed that use of coir fiber as growth medium increased the yield of strawberry
per plant and the quality of the fruit was enhanced (Martinez et al. 2017). The prop-
erties of the coir fiber used as growth medium compared to the soil are given in
Table 2.9. Fruit weight was 20 g with coir treatment compared to 18 g without. Yield
per plant was also substantially higher at 318 g compared to 210 g (Martinez et al.
2017) (Table 2.9). Combining coco peat with burnt rice husk ash and composted
sewage sludge was found to be highly effective to increase the productivity and
quality of Chrysanthemum plants (Singh et al. 2016) (Table 2.10). Physical and
chemical properties of the coco peat showed considerable variation depending on
the type of mix. A mixture of 50% coco peat, 25% soil, and 25% sewage sludge
provided the most optimum growth conditions, resulting in good root zone and ideal
appearance of chrysanthemum pots (Singh et al. 2016).
42 2 Agricultural Applications of Coir

Table 2.9 Properties of coco Coir


peat used as growth medium Parameter Soil fiber
compared to soil (Martinez
pH 7.21 6.27
et al. 2017)
Electrical conductivity, 0.46 1.84
mS/cm
Total organic matter, % 0.52 0.96
Total organic carbon, % 0.30 0.48
Carbon/nitrogen ratio 8.68 8.0
(C/N)
Sodium, % 9.47 108.0
Nitrogen, % 46.78 1.66
Phosphorus, ppm 44.5 29.6
Potassium, % 12.78 372.7
Calcium, % 65.05 <5
Magnesium, ppm 12.7 <5
Iron, ppm 17.09 0.27
Manganese, ppm 3.48 0.05
Zinc, ppm 1.89 <0.04
Copper, ppm 1.62 <0.04
Boron, ppm 0.15 0.13
Reproduced with permission from Elsevier

Table 2.10 Properties of coco peat supplemented with various ratios of rice husk ash and sewage
sludge (Singh et al. 2016)
Ratio of Bulk
soil:coco Conductivity density, Total
Supplement peat pH (dS/m) %N %P %K mg/m3 porosity, %
Rice husk ash 75:0 7.92 0.182 0.17 0.84 2.20 1.21 52.9
50:25 7.53 0.293 0.14 1.54 3.38 0.94 60.8
25:50 7.33 0.555 0.12 2.37 4.31 0.87 62.1
0:75 6.72 0.895 0.09 2.71 6.28 0.32 80.9
Sewage sludge 75:0 7.31 0.795 0.92 2.68 1.46 1.18 48.7
50:25 7.10 1.241 0.76 2.85 2.59 1.13 51.9
25:50 6.98 1.748 0.62 3.28 3.34 0.87 59.5
0:75 6.48 4.28 0.51 3.26 4.44 0.40 77.4
Reproduced with permission from Taylor and Francis

However, coir-based mulching was found to provide relatively poor plant growth
and yield for Okra compared to polyethylene or rice straw based mulching (Shamla
et al. 2017). Larger pore size in coir mat led to weed growth, which led to reduction
in yield. In a unique study, the ability of coir, pea straw, and tree bark mulches to
prevent contamination of domestic greywater by E. coli was studied (Boyte et al.
2017). The mulches were treated with solution containing E. coli and cultivated for
up to 50 days. As seen from Fig. 2.4, there was considerable decrease in the
2.3 Coir Geotextiles 43

Fig. 2.4 Concentration of E. coli on the three types of mulches after 0–50 days of inoculation
(Boyte et al. 2017). Reproduced with permission from John Wiley and Sons

­concentration of E. coli on the coir mulch. High C:N ratio of 159 was suggested to
be one of the reasons for the inhibition of the microorganism on the coir mulch. No
major increase in root dry weight was observed when composted coir was used as
mulch for cotton grown with different irrigation system. About 6.9 g of dry root
weight obtained for coir mulched cotton similar to that of unmulched plants (6.6 g),
whereas polyethylene or sugarcane mulch produced about 6.9–8.0 g of dry root
(Thukkaiyannan et al. 2005).

2.3 Coir Geotextiles

Coir mats have been extensively studied for geotextile applications, particularly to
stabilize soil. Woven and non-woven coir in different sizes, shapes, and physical
parameters have been found to be useful for reinforcing soil, sand, and even roads.
A coir mat having a mesh opening of 6 mm × 10.5 mm and a density of 0.7 kg/m2
was used to stabilize soil, and the degradation behavior of the coir geotextile was
studied. The geotextile was able to retain 43% of its strength in pH 11 media and
60% in pH 3, and corresponding strength retention was 34 and 26%, respectively,
after being in the soil for 7 months. Although the strength of the geotextile decreased,
the ability to sorb moisture had increased 2.5 times, suggesting that degradation of
coir would help to increase water retention. Coir geotextile was able to prevent any
erosion of soil during the study period (Vishnudas et al. 2012). Two types of woven
and non-woven geotextiles were subject to various treatments and studied for
44 2 Agricultural Applications of Coir

Table 2.11 Properties of the woven and non-woven geotextiles (Vivek and Raman Part 2018)
Property Woven Non-woven
Gram per square meter 769 461 1255 1000
Aperture size 12 mm × 10 mm 20 mm × 19 mm 796 μm 974 μm
Warp × weft yarns 83 × 60 46 × 40 – –
Reproduced with permission from Taylor and Francis

Table 2.12 Changes in the tensile properties of the woven and non-woven textiles before and after
the chemical treatments (Vivek and Raman Part 2018)
Woven Woven Non-woven Non-woven
Property Before After Before After Before After Before After
Tensile strength, kN/m-warp 10.45 1.5 6.0 0.4 3.35 4.35 5.9 4.8
Tensile elongation, %-warp 43 21.6 45.4 10.3 32.5 59.2 41.7 61.0
Tensile strength, kN/m-weft 10.1 0.5 4.5 0.21 1.8 2.1 2.4 2.3
Tensile elongation, %-weft 40.6 10.2 36.0 9.9 23.6 32.0 35.7 33.0
Reproduced with permission from Taylor and Francis

changes in structure and properties (Vivek and Raman Part 2018). Some of the
properties of the geotextiles are listed in Table 2.11. Both the geotextiles were
treated with sodium periodate, p-aminophenol, and sodium hydroxide. Initially, the
geotextiles were treated with 40% aqueous solution of sodium periodate for 6 h at
50 °C and later with 7% p-aminophenol for 4 h at 70 °C and finally washed with 1%
solution of sodium hydroxide. Considerable changes were observed in the proper-
ties of the geotextiles after treatment (Table 2.12).
Considerable amounts of industrial wastes are generated and dumped, leading to
ecological and health implications. Solid wastes generated from a sponge iron plant
were disposed in a dump, which was effectively restored using coir mats (Maiti and
Maiti 2015). The waste consisted of fly ash, dolochar, and slag in the form of a
mound having a surface of 5 ha, 40–50 m height, and an inclining slope of about
>70° (Fig. 2.5). To effectively handle the waste material, a new geotextile approach
was attempted. In this study, the waste mound was blanketed with top soil for about
75–100 cm at the top and the slope was covered with a coir mat. Various legumes
and grasses were planted along the dump. After 18 months, the entire slope had
stabilized and very healthy growth of the vegetation was observed (Fig. 2.5). The
average cost of the restoration was about US$ 4 m−2 considered to be inexpensive
compared to other methods. Similarly, the dry mulch accumulation on the upper and
lower slope was 14.8 and 12.6 tons per ha, considerably higher than the litter fall on
forest floor (Maiti and Maiti 2015). Coir fabrics having a mass of 720 g/m2, ultimate
load of 12.6 kN/m, 22% elongation, and thickness of 6.4 mm were also used to
reinforce sand (Lal et al. 2017a, b). The soil had a specific gravity of 2.65, dry den-
sity of 16.4 KN/m3, and angle of friction of 38.5°. Number of layers of the coir
geotextile was increased, which led to an increase in the stiffness of the soil. A
heave reduction of up to 98.7% was possible with three layers of the geotextiles.
2.3 Coir Geotextiles 45

Fig. 2.5 Digital image of the waste dump restored using coir fibers as geo textiles (left) and C.
citratus, C. juncea and A. indica grown to stablize the soil system (Maiti and Maiti 2015)

Fig. 2.6 Three types of coir fabric rolls used to reduce sediment transport from hill slopes
(Sutherland and Zeigler 2007). Reproduced with permission from Elsevier

Three different types of coir systems were used for erosion control. Left image
(Fig. 2.6) is a random fiber coil system, middle image shows an open weave coir
with low mass per unit area and low surface cover, and right image is an open weave
coir mat with high mass per unit area and high surface cover (Sutherland and Zeigler
2007). Some of the properties of the coir used are given in Table 2.13. The sediment
yield (g/m2/h) was considerably low (5.8, 5.3) for the rolled coir fiber system for
rainfall and overland flow phase, respectively, compared to 434 and 4080 when no
erosion control system was used. Hence, it was suggested that coir in different
forms can be effective in controlling the erosion of soil. In another study, a coir
blanket was used to reinforce and stablize soil in a real-world example. Unlike most
other studies, coir, jute, and a synthetic geogrid were used for erosion control on
steep slopes (45° and 60 °C). The coir blanket used had a roll size of 2 × 33 (m × m),
46 2 Agricultural Applications of Coir

Table 2.13 Properties of three types of coir system used for erosion control (Sutherland and
Zeigler 2007)
Open weave, low Open weave, high Random fiber coir
Property mass mass system
Thickness, mm 7.4 7.6 5.0
Mass/area, g/m2 470 710 270
Light transmission, % 45 23 7
Functional longevity, year 4–6 4–6 3
Tensile strength—machine 11.4 25.4 3.1
direction, kN/m
Tensile strength—traverse 10.9 17.2 3.1
direction, kN/m
Permissible shear stress 145 215 108
Coir mat opening dimensions, 1.9 × 1.9 1.3 × 1.3 –
cm
Reproduced with permission from Elsevier

Fig. 2.7 Growth of vegetation on slopes reinforced with coir, jute, and geogrids at 45 and 60°
1 month after hydroseeding (Alvarez-Mozos et al. 2014a, b). Reproduced with permission from
Elsevier

thickness of 8 mm, mass per unit area of 271 (g/m2), open area of 7%, water holding
capacity of 0 0.84 mm, and tensile strength of 3 kN/m (Alvarez-Mozos et al. 2014a,
b). Coir and jute geotextiles produced two to three times larger runoff volumes on
both the slopes but lower soil loss and erosion rates. It was suggested that the coir
geotextile performed better when buried as compared to laying on the surface
(Alvarez-Mozos et al. 2014a, b). When the geotextiles were assessed for vegetation
growth over a period of 18 months, it was observed that coir geotextile had consid-
erably lower vegetation growth rate (Fig. 2.7) with only 5 and 56% green cover
index for the 60° and 45° slopes, respectively. It was found that at steep slopes, coir
and other geotextiles had poor runoff enhancements and hence not suitable for
hydroseeded steep roadside slopes (Alvarez-Mozos et al. 2014b).
A coir mat having mesh opening of 6 × 10.5 mm2, density of 0.7 kg/m2, and
tensile strength of 18 kN/m was used to prevent soil erosion in a test plot having a
slope of 26°. Specially fabricated erosion control drums were also placed at the bot-
2.3 Coir Geotextiles 47

tom portion of the test plot. Lemon grass was seeded and observed for growth on the
test plot and control plot which did not have any geotextile. After 1 year, the soil loss
on the coir-protected plot was observed to be 94.9% less compared to the unpro-
tected plot. The coir netting had lost about 78% of its strength after 7 months in the
soil and considered to be sufficient to promote the growth of the lemon grass root
system and hence stabilize the soil (Lekha 2004). Ability of coir geotextiles to pro-
tect the embankment with 70° slope of a small water body (pond) by covering with
grass was investigated. Coir mat used in the study had a mesh opening of 6 × 6 mm2
and a density of 0.74 kg/m2 (Vishnudas et al. 2006). Considerably thi.ck vegetation
was observed on the coir reinforced section of the embankment compared to the
control plot. In addition to being highly effective, it was also reported that soil ero-
sion control using coir geotextile could be inexpensive at 1 euro/m2 compared to 2.5
euros for syntetic polymer based geotextiles (Vishnudas et al. 2006). The coir
­geotexile showed gradual decrease in strength and retained about 19% of the
strength after being in soil for 9 months (Fig. 2.8). However, some studies have sug-
gested that coir is more prone to degradation and is inferior as a geotextile compared
to jute fiber in terms of runoff reduction and peak discharge on a 29° slope under
natural conditions (Kalibova et al. 2016).
In addition to soil, the potential of using coir getotextiles to prevent sand erosion
has also been studied (Lal et al. 2017a, b). The coir mat used for preventing sand ero-
sion had mass per unit area of 768 g/m2, ultimate load of 16 kN/m, failure strain of
17.5%, modulus of 94 kN/m, and thickness of 6.47 mm. Laboratory-based erosion
studies were done using two configurations (planar and geocell) of the mat. Geocell
form provided better reinforcement and overall considered suitable for controlling

13.75 14
100 strength retained (%)
Tensile strenght (kN/m)
12
Tensile strength (kN/m)

80
strength retained (%)

10
100

60 7.65 8

6
55.6

40
4.52
4
32.87

2.60
20
2
18.9

0 initial 2 4 6 8 10
Month (1-June)

Fig. 2.8 Decrease in the tensile strength of the coir mat after being in soil for 9 months (Vishnudas
et al. 2006). Reproduced with permission through Creative Commons License
48 2 Agricultural Applications of Coir

Fig. 2.9 Images showing the materials used for installing the geotextiles on clay dikes and a site
image of the actual clay dyke reinforced with coir geotexile (Lekha and Kavitha 2006). Reproduced
with permission from Elsevier

sand erosion in beaches and seashores (Lal et al. 2017a, b). Coir geotexile sand
interface was reported to have better bond resistance than coir–soil interface and
differences were observed even between coarse and fine sand. However, it was con-
cluded that mesh size, weave type, and soil grain size were major factors to be
considered before deciding on the appropriate geotextile (Subaida et al. 2008). Coir
geotextile was also found to be highly effective to reinforce clay dykes by prevent-
ing leakage of soil particles but allowed drainage of water. As seen in Fig. 2.9, rolls
(logs) of coir were stretched along the dyke and fixed to the soil using iron hooks
and bamboo. Coir reinforced clay dykes were able to provide good penetration
resistance even after 1 year. The coir reinforced paddy fields were able to avoid
ingress of salt water but at the same time drained out excess water (Lekha and
Kavitha 2006). In addition to preventing erosion of soil, geotextiles are helpful to
control and prevent growth of weeds along the banks of water bodies. In a field
study, prevegetated coir was used to reinforce a streambank of Teton river, Idaho
and prevent the growth of reed canary grass. Rapid establishment of native vegeta-
tion and prevention of canary grass were possible even after 3–6 years of initial
installation (Hook et al. 2009).
A particular type of ecosystem called Cerrado in Brazil is prone to severe soil
erosion and instability. Coir geotextile was studied as a substrate to promote the
growth of grasses and legumes and prevent soil erosion and improve soil organic
carbon and total nitrogen, phosporous, and postassium in the Cerrado region
(Marques et al. 2016). For this study, coir geotextiles having width of 1.5–3 m, den-
sity of 300 g/m2, and tensile strength of 60 kgf/m was used to cover the soil on to
which one type of grass and two types of legumes were seeded (Fig. 2.10).
Considerable increase in the total soil carbon and labile nutrients was observed with
the use of the geotextiles (Table 2.14). Further, it was found that the coir geotextile
retained 23% of its original strength after 12 months of exposure to natural environ-
mental conditions. Treating the geotextile with lime was also found to decrease the
degradation during the first 3 months but the cellulose content and modulus of the
fibers decreased rapidly after 3 months leading to a low strength retention of 19%
(Marques et al. 2014).
Natural coir is susceptible to microorganisms and also has high moisture regain,
which limits its geotextile applications. To improve the durability, coir yarns were
first treated with either copper sulfate or boric acid and later immersed in cashew
2.3 Coir Geotextiles 49

Fig. 2.10 Picture of the soil that was reinforced with the geotextile (b, d) and anchoring the geo-
textile with J-clips (c) (Marques et al. 2016). Reproduced with permission from Elsevier

Table 2.14 Changes in the composition of the soil with and without the use of the geotextile
(Marques et al. 2016)
Total nitrogen (N) Phosphorous (P) Potassium (K)
Sample Soil organic carbon (SOC) mg/kg mg/kg mg/kg
Control 3.7 ± 0.11 1670 ± 120 1206 ± 51 560 ± 52
Coir grass 4.9 ± 0.58 2130 ± 120 1136 ± 30 643 ± 14
Coir–legume 6.0 ± 0.12 2500 ± 690 1294 ± 20 694 ± 3.9
Control 3.1 ± 0.23 1600 ± 100 1230 ± 73 512 ± 32
Coir grass 3.6 ± 0.01 1330 ± 250 1434 ± 79 546 ± 17
Coir–legume 3.7 ± 0.12 1630 ± 60 1304 ± 61 544 ± 70
Reproduced with permission from Elsevier

nut shell liquid dissolved in kerosene and nitric acid (HNO3) as the polymerizing
agent. Various concentrations (10–50%) of CNSL was used for coating to achieve
the desired hydrophobicity (Sumi et al. 2017). Changes in the properties of the
fibers due to the treatment are shown in Table 2.15. Substantial increase in antimi-
50 2 Agricultural Applications of Coir

Table 2.15 Changes in the antimicrobial resistance and properties of coir yarns treated with
various concentrations of CNSL (Sumi et al. 2017)
Fiber CFU/mL % Water absorption Breaking load, N Breaking strain, %
Untreated 220 80 158 ± 12 17.0 ± 1.5
10% CNSL 30 70 178 ± 11 11.0 ± 1.8
20% CNSL 10 55 185 ± 14 9.8 ± 1.1
30% CNSL 10 53 182 ± 12 9.3 ± 0.8
40% CNSL 10 52 183 ± 18 8.4 ± 0.8
50% CNSL 10 52 180 ± 8 6.7 ± 0.8
Reproduced with permission from Taylor and Francis

crobial resistance, lower water absorption, and marginally higher tensile strength
were suggested to improve the performance of the coir for geotextile applications
(Sumi et al. 2017).

2.4 Miscellaneous Applications

Retention of moisture and providing adequate nutrition are necessary for obtaining
high yields from crops and maximizing the use of natural resources. Most fertilizer
and nutrients are directly added into the soil, which leads to burst releases and
improper application. Studies were done to make coco peat superabsorbent, have
high rate of loading and provide sustained release of nutrients. To achieve this, coir
pith was first pretreated with adipic acid dihydrazide and later modified with maleic
anyhydride and grafted with poly(acrylic acid) (PPA) superabsorbent as per the
scheme shown (Fig. 2.11). Considerably large number of amine groups were added
on to the surface of coco peat, which assisted in acheving high grafting efficiency of
75%. The superabsorbent developed was able to retain up to 470 g of water per g of
material, which is substantially higher than common absorbents. Grafting of PAA to
the modified coco peat was done by heating at 55 °C using APS as the initiator
under nitrogen atmosphere (Xu et al. 2016). Ability of the modified and grafted
coco peat to absorb and release nutrients (NPK) was determined by adding the
superabsorbent into 500 mL of water containing 0.03 mol/L of KH2PO4 and NH4NO3
for 6 h, so that equilibrium sorption could be obtained. Morphologically, the surface
of the modified sample was changed into a coarse and wrinkled surface containing
small pores and fine particles. Due to these modifications and presence of the syn-
thetic polymers, the contact angle had reduced to 37° from 68° and further to 0°
after grafting which also led to increased microporous holes. A grafting efficiency
of up to 75% could be achieved. Extent of water absorbency was directly related to
the grafting efficiency (Table 2.16). In addition to the very high sorption, there was
a gradual desorption indicating good water holding capacity for up to 50 h. Nutrients
(N, P, and K) loaded onto the modified peat also had a desired slow and constant
2.4 Miscellaneous Applications 51

Fig. 2.11 Schematic representation of the mechanism of developing the poly(acrylic acid) based
superabsorbent (Xu et al. 2016). Reproduced with permission from RSC publishing

Table 2.16 Water absorbency of coco peat grafted with PAA and at different ratios of acrylic acid
(Xu et al. 2016)
Water absorbency, g/g
Ratio of grafted peat to Distilled 0.9% NaCl
acrylic acid Grafting efficiency, % water Tap water solution
1:3 70.9 ± 0.9 445 311 55
1:5 72.1 ± 1.1 461 330 58
1:7 75.2 ± 1.3 470 350 61
1:9 69.5 ± 0.8 458 334 57
1:11 68.3 ± 1.2 416 308 50
Reproduced with permission from RSC publishing

release profile. However, N was released much more faster than P and K but at a rate
preferable for plant growth. High water and nutrient sorption and controlled release
were suggested to be ideal for adopting simultanoues irrigation and fertilization for
horticultural applications (Xu et al. 2016).
52 2 Agricultural Applications of Coir

References

Abad M, Noguera P, Puchades R, Maquieira A, Noguera V (2002) Physico-chemical and chemical


properties of some coconut coir dust for use as a peat substitute for containerized ornamental
plants. Bioresour Technol 82:241–245
Aiome GVN, Silva CS (2014) Effect of mulch on soil properties and yield of groundnut plants
exposed to temperature stress. http://www.ou.ac.lk/ours/wp-content/uploads/2017/05/
Agriculture-1.pdf
Alvarez-Mozos J, Abad E, Gimenez R, Campo MA, Goni M, Arive M, Casali J, Diez J, Diego
I (2014a) Evaluation of erosion control geotextiles on steep slopes. Part 2: influence on the
establishment and growth of vegetation. Catena 121:195–203
Alvarez-Mozos J, Abad E, Gimenez R, Campo MA, Goni M, Arive M, Casali J, Diez J, Diego I
(2014b) Evaluation of erosion control geotextiles on steep slopes. Part 1: effects on runoff and
soil loss. Catena 118:168–178
Alwis A, Herath HKMSK (2012) Impact of mulching on soil moisture, plant growth and yield of
Mauritius pineapple (Ananas comosus L. Merr). J Food Agric 2(1)
Arenas M, Vavrina CS, Cornell JA, Hanlon EA, Hochmuth GJ (2012) Coir as an alternative to peat
in media for tomato transplant production. Hortic Sci 37(2):309–312
Aslani L, Mobli M, Keramat J (2016) Comparison of antioxidant activities and fruit quality attri-
butes in four sweet bell pepper (Capsicum annum L.) cultivars grown in two soilless media.
J Hortic Sci Biotechnol 91(5):497–505
Barreiro A, Bååth E, Díaz-Ravina M (2016) Bacterial and fungal growth in burnt acid soils
amended with different high C/N mulch materials. Soil Biol Biochem 97:102–111
Beena KS, Anil KR (2011) Effectiveness of coir geotextiles in soil moisture conservation. Int
J Eng Tech 3(3):200–207
Boyte S, Quaife S, Horswell J, Siggins A (2017) Survival of Escherichia coli in common garden
mulches spiked with synthetic grey water. Lett Appl Microbiol 64:386–391
Dishani PTN, Silva CS (2016) Effect of simulated temperature and water stress on growth, physi-
ological and yield parameters of tomato grown on mulch. OUSL J 11:37–51
Erwan, Ismail MR, Sariah M, Saud HM, Habib SH, Kausar H, Naber L (2013) Effect of oil palm
frond compost amended coconut coir dust soilless growing media on growth and yield of cau-
liflower. Int J Agric Biol 15:731–736
Fornes F, Belda RM, Abad M, Noguera P, Puchades R, Maquieira A, Noguera V (2003) The micro-
structure of coconut coir dusts for use as alternatives to peat in soilless growing media. Aust
J Exp Agric 43:1171–1179
Fornes F, Belda RM, Fernández de Córdova P, Cebolla-Cornejo J (2017) Assessment of biochar
and hydrochar as minor to major constituents of growing media for containerized tomato pro-
duction. J Sci Food Agric 97(11):3675–3684
Ghosh PK, Sarma US, Ravindranath AD, Radhakrishnan S, Ghosh P (2007) A novel method for
accelerated composting of coir pith. Energy Fuels 21:822–827
Hook PB, Salsbury KM, Klausmann JM (2009) Revegetation of reed canarygrass infested ripar-
ian areas: performance of pre-vegetated coir after 3 to 6 years. Proc Am Soc Min Reclam
2009:597–629
Jeyaseeli DM, Raj SP (2010) Chemical characteristics of coir pith as a function of its particle size
to be used as soilless medium. Ecoscan 4(2&3):163–169
Kalibova J, Jacka L, Petro J (2016) The effectiveness of jute and coir blankets for erosion control
in different field and laboratory conditions. Solid Earth 7:469–479
Kim HS, Kim KR, Yang J-E, Ok YS, Kim WI, Kunhikrishnan A, Kim K-H (2017) Amelioration of
horticultural growing media properties through rice hull biochar incorporation. Waste Biomass
Valori 8(2):483–492
Lal D, Sankar N, Chandrakaran S (2017a) Effect of reinforcement form on the behavior of coir
geotextile reinforced sand beds. Soils Found 57:227–236
References 53

Lal D, Sankar N, Chandrakaran S (2017b) Surface heave behaviour of coir geotextile reinforced
sand beds. J Instit Eng India A 98(1–2):121–125
Lekha KR (2004) Field instrumentation and monitoring of soil erosion in coir geotextile stabilized
slopes—a case study. Geotext Geomembr 22:399–413
Lekha KR, Kavitha V (2006) Coir geotextile reinforced clay dykes for drainage of low lying areas.
Geotext Geomembr 24:38–51
Maiti SK, Maiti D (2015) Ecological restoration of waste dump by topsoil blanketing, coir matting
and seeding with grass-legume mixture. Ecol Eng 77:74–84
Marques AR, Patrico PSO, Santos FS, Monteiro ML, Urashima DC, Rodrigues CS (2014) Effects
of the climatic conditions of the southern Brazil on degradation of fibers of coir-geotextile:
evaluation of mechanical and structural properties. Geotext Geomembr 42(1):76–82
Marques AR, Vianna CR, Monteiro ML, Pires BOS, Urashima DC, Pontes PP (2016) Utilizing coir
geotextile with grass and legume on soil of Cerrado, Brazil: an alternative strategy in improv-
ing the input of nutrients in degraded pasture soil. Appl Soil Ecol 107:290–297
Martinez F, Oliveira JA, Calvete EO, Palencia P (2017) Influence of growth medium on yield,
quality indexes and SPAD values in strawberry plants. Sci Hortic 217:17–27
Mathad JC, Jholgiker P (2005) Effect of synthetic and organic mulches in improving growth, yield
and quality of strawberry under subtropical ecosystem. Acta Hortic 696:56–61
Meerow AW (1994) Growth of two subtropical ornamentals using coir (coconut mesocarp pith) as
a peat substitute. Hortic Sci 29(12):1484–1486
Muti SM, Kibe AM, Ngetich W, Muindi E (2017) Effects of coir dust mulch on evapotranspiration
of PH4 maize in coastal region of Kenya. J Agric Ecol Res 10(3):1–16
Nazari F, Farahmand H, Khosh-Khui M, Salehi H (2011) Effects of coir as a component of potting
media on growth, flowering and physiological characteristics of hyacinth. Int J Agri Food Sci
1(2):34–38
Offord CA, Muir S, Tyler JL (1998) Growth of selected Australian plants in soilless media using
coir as a substitute for peat. Aust J Exp Agric 38:879–887
Padamadevi SN, Bai RSM, William SPM, Sunithakumari K (2016) Organic cultivation of medic-
inal plants: influence of composted coir pith on the growth and yield of Coleus forskohlii
(willd.) Briq. Compost Sci Util 24(4):266–272
Prabhu SR, Thomas GV (2002) Biological conversion of coir poth into a high value added resource
and its application in agri-hortuculture: current status, prospects and perspective. J Plant Crops
30(1):1–47
Ramesh K, Gurumurthy S, Veerabaran V, Senthilvel S, Shanmugasundaram K (2006) Impact of
irrigation regimes, irrigation frequencies and coirpith mulching on the economic productivity
of drip irrigated summer cotton SVPR-2. Res J Agric Biol Sci 2(6):447–451
Reghuvaran A, Ravindranath AD (2010) Efficacy of biodegraded coir pith for cultivation of medic-
inal plants. J Sci Indust Res 69:554–559
Rhie YH, Kim J (2017) Changes in physical properties of various coir dust and perlite mixes and
their capacitance sensor volumetric water content calibrations. Hortic Sci 52(1):162–166
Rubio JS, Rubio F, Martinez V, Garcia-Sanchez F (2010) Amelioration of salt stress by irrigation
management in pepper plants grown in coconut coir dust. Agric Water Manag 97:1695–1702
Shamla K, Sindhu PV, Menon MV (2017) Effect of weed management practices on growth and
yield of okra (Abelmoschus esculentus (L.) Moench.). J Trop Agric 55(1):57–62
Singh S, Dubey RK, Kukal SS (2016) Nitrogen supplemented cocopeat-based organic wastes as
potting media mixtures for the growth and flowering of chrysanthemum. Commun Soil Sci
Plant Anal 47(16):1856–1865
Somasundaram J, Krishnasamy R, Savithri P, Vassanda Coumar M, Arun Kumar V, Satish Kumar
B, Indirani R et al (2016) Effect of sewage sludge–coir pith pellets on dry matter yield and
trace metal concentration in various plant parts of forage maize. J Plant Nutr 39(11):1556–1569
Stamps RH, Evans MR (1997) Growth of Dieffenbachia maculate “Camille” in growing media
containing Sphagnum peat or coconut coir dust. Hortic Sci 32(5):844–847
Stamps RH, Evans MR (1999) Growth of Dracaena marginata and Spathiphyllum ‘petite’ in
Sphagnum peat and coconut coir dust based growing media. J Environ Hortic 17(1):49–52
54 2 Agricultural Applications of Coir

Subaida EA, Chandrakaran S, Sankar N (2008) Experimental investigations on tensile and pullout
behavior of woven coir geotextiles. Geotext Geomembr 26:384–392
Sumi S, Unnikrishnan N, Mathew L (2017) Surface modification of coir fibers for extended
hydrophobicity and antimicrobial property for possible geotextile application. J Nat Fibers
14(3):335–345
Sutherland RA, Zeigler AD (2007) Effectiveness of coir based rolled erosion control systems in
reducing sediment transport from hillslopes. Appl Geogr 27:150–164
Thankamani CK, Kandiannan K, Hamza S, Saji KV (2016) Effect of mulches on weed suppression
and yield of ginger (Zingiber officinale roscoe). Sci Hortic 207:125–130
Thukkaiyannan P, Singh SDS, Prabhakaran NK (2005) Studies on the irrigation methods and
mulching on root characters, water saving and yield of summer irrigated cotton. Madras Agric
J 92(7–9):398–403
Tripetchkul S, Pundee K, Koonsrisuk S, Aekprathumchai S (2012) Cocomposting of coir pith
and cow manure: initial C/N ratio vs physio-chemical changes. Int J Recycl Org Waste Agric
1:15–23
Vavrina CS, Armbrester K, Arenas M, Pena M (1996) Coconut coir as an alternative to peat media
for vegetable transplant production. SWFREC Station Report, Immokalee, Florida 96, no. 7
Vishnudas S, Savenije HHG, Zaag PV, Anil KR, Balan K (2006) The protective and attractive
covering of a vegetated embankment using coir geotextiles. Hydrol Earth Syst Sci 10:565–574
Vishnudas S, Savenije HHG, Zaag PV, Anil KR (2012) Coir geotextile for slope stabilization
and cultivation—a case study in a highland region of Kerala, South India. Phys Chem Earth
47-48:135–138
Vivek RKD, Raman Part I (2018) Effect of chemical treatment on the tensile strength behavior of
coir geotextiles. J Nat Fibers:1–15. https://doi.org/10.1080/15440478.2018.1503132
Wira AB, Razi MR, Jamil ZA (2011) Composts as additives in coconut coir dust culture for grow-
ing rockmelon (Cucumis melo I.). J Trop Agric Food Sci 39(2):229–237
Xu X, Su X, Bai B, Wang H, Suo Y (2016) Synthesis of adipic acid dihydrazide-decorated
coco peat powder-based superabsorbent for controlled release of soil nutrients. RSC Adv
6(105):103199–103209
Chapter 3
Biotechnological Applications for Coir
and Other Coconut Tree By-products

3.1 Introduction

Coir and other coconut by-products have been extensively studied for their potential
use in biofuel production, as substrates for enzyme and chemical production and cul-
tivation of mushrooms. Low cost, relatively longer degradation times, and easy avail-
ability are the primary reasons for using coconut by-products for biotechnological
applications. Among the various studies, production of enzymes and using the by-
products as a source for cultivation of mushrooms seems more feasible. Although
several studies have been done to use coir for biofuel production, coir has 30–40%
lignin, which is highly recalcitrant and hence conversion of the lignocellulose into
glucose and subsequently into ethanol is not economical. However, researchers have
attempted to utilize coir and other coconut by-products for various biotechnological
applications other than biofuels. This chapter provides an overview of the applications
that coir and coir by-products have been utilized in biotechnology and related areas.

3.2 Production of Enzymes

Due to the abundant availability and low cost of coir, studies have been done to
understand the potential of using coir as a substrate for enzyme production. Three
species of bacteria, namely Cellulomonas, Bacillus, and Micrococcus spp., were iso-
lated from coir retting effluents. Using coir powder as a substrate, various concentra-
tions of bacteria, pH, and temperature during treatment were studied. Enzyme
activity ranged from 0.0087 to 0.0071 U/mL at 20 °C depending on the type of
bacteria used. Similar variations were also observed with change in pH, time, or
temperature. Under the conditions studied, it was found that Cellulomonas sp. pro-
duced the highest enzymes followed by Bacillus and Micrococcus spp. (Immanuel
et al. 2006). Two enzymes endoglucanase and β-glucosidase were produced using

© Springer Nature Switzerland AG 2019 55


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_3
56 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

Aspergillus ochraceus MTCC 1810 extracted from coir pith (Asha et al. 2016).
While endoglucanase had a molecular mass of about 78 kDa, it was substantially
lower at 43 kDa for β-glucosidase. Corresponding activity levels were 35.6 and
15.2 U/mL at pH 6 and 40 °C, respectively. It was suggested that the two enzymes
could act synergistically and achieve complete cellulose saccharification leading to
higher conversion efficiencies (71%) (Asha et al. 2016). It has also been shown that
the type and yield of bacterial and fungal strains from coir depends on the location
and type of coir. Compared to wood and peat, coir has considerably higher number
of microbes with up to 124 operational taxonomic units (Montagne et al. 2017). The
type and range of microorganisms that could be extracted from coir fibers (CoF) and
other biomasses are shown in Fig. 3.1.

Fig. 3.1 Depiction of the type (a) and amount (b) of microorganisms that could be extracted from
coir fibers (Montagne et al. 2017). Reproduced with permission from Springer Nature
3.2 Production of Enzymes 57

To obtain low-cost cellulose degrading enzymes, two common types of fungi


Aspergillus niger and A. fumigatus were isolated from soil in a coir retting ground
and studied for their potential to produce enzymes using coir fibers as a base. The
fungal cultures were maintained in a Czapek–Dox agar broth at 30 °C for 5 days.
Cellulose production was maximum at pH 5 (0.092 IU/mL) when A. niger was used
but the level of production decreased when the pH was increased. A. fumigatus pro-
vided an enzyme production of 0.198 IU/mL also at pH 5. Similar results were also
obtained when enzymes were produced using the FPA and DNS methods. However,
the different methods affected the molecular weight of the enzymes obtained. A.
fumigatus produced enzymes with molecular weights of about 32 and 21 kDa com-
pared to 36 and 23 kDa bands produced by A. niger. It was suggested that the
enzymes produced could be used for production of low-cost ethanol from various
sources of biomass (Immanuel et al. 2007).
Ability of Aspergillus niger to produce cellulase in both submerged and solid
state fermentation was studied by Mrudula et al. using coir fibers and other biomass
as substrates (Mrudula and Murugammal 2011). After the desired culture time, the
enzymes were purified and the cellulase activity was determined based on standard
assays. Production of cellulase was higher (14.6 times) in SSF and also required
shorter time (72 h) compared to 96 h when submerged fermentation was used.
Enzyme activities were 8.89 and 3.56 U/g for CMCase and FPase, respectively. In a
similar study, Trichoderma harzianum was isolated from soil and used to evaluate
the potential to produce cellulase and xylanase using the solid state fermentation
approach (Namasivayam et al. 2015). Effect of moisture content ranging from 0 to
50% on the production of the enzyme was also studied (Namasivayam et al. 2015).
Similar studies have also shown that cellulase can be obtained from coir pith using
Aspergillus nidulans (AJSU04). Some of the variables used and resulting properties
of the enzymes produced are given in Table 3.1 (Anuradha Jabasingh et al. 2014).
Solid state fermentation was also used to produce cellulase from coir pith. The
pith was crushed to about 250 μm and later delignified and treated with 1% hydro-
gen peroxide. Various culture conditions were studied to understand their effect on
cellulase production using T. viride as the microorganism. Enzyme activity changed

Table 3.1 Production and activity of cellulase produced using coir pith as a substrate (Anuradha
Jabasingh et al. 2014)
Specific Total
Purification activity Volume protein, Total Recovery, Purification
step (U/mg) fraction, mL mg activity, U % fold
Crude 31.17 15 29.0 908 100 1
enzyme
(NH4)2SO4 50.92 10 11.5 586 65 1.6
Anion 53.6 01 1.03 55 6 1.7
exchange
Sephadex 60.58 0.5 0.36 22 2.4 1.9
G-200
Open access publication
58 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

15

Enzyme activity IU/ gm

10

x x x
x x
0
0 5 10 15 20

Fig. 3.2 Influence of the amount of nutrients on enzyme activity when the coir pith was treated
with three different culture media (Muniswaran et al. 1994)

substantially depending on the conditions during culture (Fig. 3.2). A highest


enzyme production of 4.75 IU/g was obtained and the enzyme was found to be
capable of hydrolyzing various cellulosic biomass (Muniswaran and Charyulu
1994). It was also demonstrated that cellulase could be obtained from A. niger using
coir pith as the substrate. A highest filter paper activity (FPA) of 4.11 IU/g, carboxyl
methyl cellulose (CMCase) activity of 15.6 IU/g, and cellobiase activity of 9.3 IU/g
was possible after 7–8 days of fermentation (Muniswaran et al. 1994). Inoculum
size of 20–50% and coconut pith particle size of 375 μm were found to be most
optimum for achieving higher yields. Coir pith was used as a basal medium to
­produce laccases using Pleurotus florida (Packiyam and Ragunathan 2017) and in
the presence of copper sulfate. Optimum pH for production of laccase was 6 and
temperature was 50 °C at an enzyme activity of 10.56 IU/mg. Activity and protein
content of laccase increased with increasing concentration of copper sulfate with
highest protein content obtained being 0.51 mg at 200 UM copper sulfate. Coir pith
was considered to be an ideal substrate for laccase production.
Coconut oil cake is another coproduct obtained during the conversion of the fruit
into oil. This cake has high availability at low cost. Solid state fermentation was
used to produce α-amylase using coconut oil cake (Ramachandran et al. 2004). The
substrate was treated with salt solution and autoclaved at 121 °C for 20 min. Later,
A. oryzae was inoculated and incubated for up to 120 h on the coconut cake. Up to
3.3 Production of Ethanol 59

1372 U/gds was obtained in 24 h. An increase of up to 1911% and as high as 3388 U/


gds was obtained when the fermentation was supplemented with 0.5% starch and
1% peptone, respectively. Low cost of coconut cake and high yield of α-amylase
were considered to make the system suitable for commercial scale production
(Ramachandran et al. 2004).

3.3 Production of Ethanol

Many studies have been conducted to explore the possibility of converting cellulose
in coir into glucose and subsequently fermented into ethanol. As seen from Fig. 3.3,
coir fibers were processed through the simultaneous saccharification and fermenta-
tion route to obtain ethanol using S. cerevisiae under anaerobic conditions (Ebrahimi
et al. 2017). Various pretreatments were also done to improve the digestion of coir
and obtain high ethanol yields. Maximum glucan digestibility obtained was around
80% and the highest ethanol yield was 9%. However, pretreatment conditions
played a significant role in determining the extent of digestibility and hence ethanol
yield (Table 3.2). Although the yield of ethanol was relatively low, it was suggested
that other microorganisms should be tested to obtain higher yields. Instead of using
different microorganisms, green coconut mesocarp (husk including fibers) was sub-
ject to various pretreatments and ability to generate ethanol was studied using S.
cerevisiae. A simple pretreatment with NaOH solutions ranging from 1 to 4% and
treatment time of up to 24 h was done to improve digestibility. Similarly, the enzyme
loading during saccharification varied from 3.8 to 15%. A maximum ethanol yield
of 3.7% was obtained when the enzyme loading was 15 FPU/g (Soares et al. 2016).
Further increase in the yield of ethanol up to 4.3% was achieved when fed batch
process and ethanolic yeast was used for fermentation (Soares et al. 2017).

Fig. 3.3 Schematic representation of the production of ethanol from coir fibers using the simulta-
neous saccharification and fermentation approach using S. cerevisiae (Ebrahimi et al. 2017).
Reproduced with permission from Elsevier
60 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

Table 3.2 Various pretreatment conditions and their effect on the removal of lignin, glucan
digestibility, and ethanol yield (Ebrahimi et al. 2017)
T, S/L Time, Solid Lignin Glucan Ethanol
Pretreatment °C ratio min recovery, % removal, % digestibility concentration, %
AAG 130 1:20 15 78.3 24.3 68.9 6.86
– – – 30 69.5 37.7 79.7 8.81
– – – 60 68.7 38.8 76.2 8.41
AAG 130 1:30 15 78.5 24.1 69.1 7.03
– – – 30 68.1 37.6 81.8 8.97
– – – 60 67.2 40.9 75.3 8.29
AG 130 1:30 30 94.3 7.3 12.6 2.38
– – – 60 93.8 10.1 13.5 2.66
Untreated – – – 100 – 11.8 1.67
coir
Reproduced with permission from Elsevier

Pretreatment conditions were observed to have a large influence on the extent to


which the coir fibers could be converted into sugars. The best conversion ratio of
7.6 g/L of reducing sugar was achieved by treating with 11% NaOH solution at
100 °C, which resulted in a lignin removal of 14% and an increase in cellulose con-
tent of 50% (Fatmawati et al. 2013). However, the sugar yield obtained here was
lower than that observed in other studies. To further improve the yield of sugars
from hydrolysis of coir, the fibers were subject to UV irradiation and also exposed
to ultrasound. The treatment was considered to increase the delignification, create
more accessible regions and hence increase the yield (Subhedar et al. 2018). For the
ultrasound-assisted delignification, about 0.5% of powder coir fibers were immersed
in 1 N alkali solution and subject to a microwave power of 100 W for up to 70 min.
Similarly, ultrasonication at 20 kHz, power dissipation of 80 W, substrate loading of
3% (w/v), and enzyme loading of 0.08% were used for enzymatic hydrolysis. There
was a continuous increase in % delignification of up to 45% with increase in treat-
ment time, and 80% delignification could be achieved within 70 min. Sugar yields
(25 g/L) were higher for coir than groundnut or pistachio shells, and the conversion
time was also substantially shorter (Fig. 3.4) (Subhedar et al. 2018).
A comparison of the ability of green coconut shell (GCS), mature coconut fiber
(MCF), and mature coconut shell (MCS) to be converted into ethanol was studied
using simultaneous and semi-simultaneous fermentation processes. To achieve high
yield, pretreatment was done using alkaline hydrogen peroxide and later with
sodium hydroxide (Fig. 3.5). For the peroxide treatment, 0.4 g of the materials were
mixed with 31.75 mL of hydrogen peroxide at 7.35% (v/v), 25 °C for 1 h at pH 11.5.
Later, the treated materials were further delignified by mixing with 4% (w/v) solu-
tion of NaOH and heating at 100 °C for 1 h. Fermentation was done using P. stipitis,
S. cerevisiae, and Z. mobilis. SSF was done by treating 4% of the material in 48 mL
of sodium citrate buffer (pH 5.0) with cellic Ctec2 and HTec2 at enzyme loading of
30 FPU, 75 CBU, and 130 IU per gram (Table 3.3). The broth was incubated for
48 h at 30 °C to allow the fermentation into ethanol to complete. Similarly, SSSF
3.3 Production of Ethanol 61

Fig. 3.4 Yield of sugars from ultrasound-assisted treatment of coir fibers (Subhedar et al. 2018)

Cactus (CAC) Green coconut Mature coconut Mature coconut


shell (GCS) fibre (MCF) shell (GCS)

Sequential (Alk-H2O2/NaOH)
pretreatment

Pre-hydrolytic
SSF strategy
period

SSSF strategy

Fuel-ethanol

Fig. 3.5 Process of producing ethanol from coconut and its by-products (Gonçalves et al. 2014).
Reproduced with permission from Elsevier
62 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

Table 3.3 Comparison of the ability of two different processes and three different microorganisms
to produce ethanol from mature coconut fiber (Gonçalves et al. 2014)
Ethanol Ethanol productivity,
Process Microorganism Ethanol yield, % concentration, % (g/Lh)
SSF S. cerevisiae 84 ± 0.6 8.4 ± 0.06 0.18 ± 0.01
P. stipitis 79 ± 1.6 9.1 ± 0.18 0.19 ± 0.00
Z. mobilis 82 ± 0.6 8.3 ± 0.06 0.17 ± 0.00
SSSF S. cerevisiae 89 ± 0.7 9.3 ± 0.08 0.19 ± 0.01
P. stipitis 85 ± 0.5 10.2 ± 0.06 0.21 ± 0.00
Z. mobilis 86 ± 1.0 8.9 ± 0.1 0.19 ± 0.00
Reproduced with permission from Elsevier

was done by using 8 h pre-hydrolysis and 40 h SSF process. Considerable variations


in the composition of the shell and fiber lead to various levels of changes (Table 3.4).
Consequently, the yield of ethanol was also found to vary considerably depending
on the source and extent of pretreatment. The presaccharification step used in this
study was found to be necessary and increased ethanol yields in mature coconut
fibers from about 79–84% to 85–89% depending on the microbial strain (Gonçalves
et al. 2014). To further increase the conversion efficiency, the mature coconut fibers
were subject to a hydrothermal pretreatment catalyzed by NaOH. Considerable
increase in ethanol yield to 90–91% was obtained depending on the microbial
strains (Fig. 3.6 and Table 3.5). In addition, the process was able to separate lignin
and other phenolics in the coir for various other uses (Gonçalves et al. 2016).
A process called autohydrolysis which involves treating the coconut by-products
at temperatures ranging from 160 to 200 °C from 10 to 50 min was used to increase
the sugar and ethanol yield from coconut by-products. It was expected that the sur-
face area, porosity, and amount of lignin would change due to autohydrolysis which
would increase the susceptibility of the biomass leading to higher conversion ratios
(Gonçalves et al. 2015). An ethanol yield of 90% was obtained with a productivity
of 0.21 g/Lh when green coconut shells were treated with S. cerevisiae using the
semi-simultaneous saccharification and fermentation process (Gonçalves et al.
2015). A chemical route has also been used to improve the yield of ethanol from
coir. Fibers were chopped to obtain particles with size of 0.84 mm which were later
treated with two different (20 and 25%) concentrations of ammonium carbonate for
4, 8, or 12 h at 80 °C. Treated samples were hydrolyzed using a cellulase enzyme
and incubated at 50 °C at 150 rpm for 72 h. Conversion into ethanol was done using
the simultaneous saccharification and fermentation process by adding 5% glucan,
1% yeast, 2% peptone, 0.05 M citrate buffer, and 10FPU/g of cellulase enzyme. The
amount of glucan recovery and digestibility varied with the pretreatment conditions
and reached as high as 85 and 77%, respectively. Similarly, the lignin recovery and
removal percentage was highest at 68 and 64%, respectively (Ebrahimi et al. 2018)
(Table 3.6). Corresponding to the changes in the composition and digestibility, the
ethanol concentration after the SSF varied between 1.7 g/L for untreated to 8.7 g/L
for the treated samples. The ethanol yield obtained from coir fibers in this study
3.3 Production of Ethanol 63

Table 3.4 Composition of the various coconut by-products before and after the two treatments
used to improve conversion (Gonçalves et al. 2014)
Components Untreated Alkali-H2O2 Alkali-H2O2/NaOH
Mature coconut fiber
Cellulose 31.6 ± 0.5 41.5 ± 0.9 51.8 ± 0.8
Hemicellulose 26.3 ± 0.9 28.4 ± 0.7 25.8 ± 0.5
Insoluble lignin 25.0 ± 0.8 16.5 ± 0.3 8.8 ± 0.2
Soluble lignin 1.7 ± 0.2 0.8 ± 0.2 0.09 ± 0.01
Extractable 5.4 ± 0.2 0.4 ± 0.1 0.05 ± 0.04
Ash 3.3 ± 0.3 2.9 ± 0.2 2.98 ± 0.14
Green coconut shell
Cellulose 32.9 ± 0.9 51.6 ± 0.9 54.1 ± 0.14
Hemicellulose 26.5 ± 0.5 27.9 ± 0.9 28.4 ± 0.3
Insoluble lignin 25.4 ± 0.8 9.1 ± 0.04 7.6 ± 0.4
Soluble lignin 1.4 ± 0.3 0.6 ± 0.01 0.25 ± 0.01
Extractable 3.3 ± 0.2 0.8 ± 0.28 0.26 ± 0.45
Ash 4.3 ± 0.2 2.0 ± 0.06 1.1 ± 0.07
Mature coconut shell
Cellulose 30.5 ± 0.9 37.2 ± 0.7 53.9 ± 0.4
Hemicellulose 25.4 ± 0.3 29.3 ± 0.7 23.0 ± 0.6
Insoluble lignin 31.0 ± 0.2 18.1 ± 0.3 9.3 ± 0.2
Soluble lignin 2.1 ± 0.1 1.4 ± 0.14 0.9 ± 0.03
Extractable 2.7 ± 0.3 0.84 ± 0.04 0.53 ± 0.03
Ash 4.8 ± 0.1 4.3 ± 0.3 3.80 ± 0.07
Reproduced with permission from Elsevier

using the carbonate pretreatment was higher compared to previous attempts on con-
verting coir into ethanol (Table 3.7).
Coconut husks having a cellulose content of 39% cellulose, 16% hemicellulose,
and 30% lignin were pretreated with 20–30% NaOH at 100 °C for 2–3 h under pres-
sure. Conversion of cellulose to glucose using Celluclast and Novozymes 188 pro-
vided a yield of about 21.2 and 20.7% glucose, respectively. The yeast S. cerevisiae
was used for fermentation for both the SSF and SHF processes. About 85% of the
theoretical conversion from glucose to ethanol was possible in this process
(Vaithanomsat et al. 2011). The possibility of increasing glucose yield from coir by
acid hydrolysis was investigated. Coir fibers were treated with nitric acid and/or
acetic acid from 0.2 to 0.5% (w/v) at temperatures from 140 to 160 °C for 10 to
30 min (Amenaghawon et al. 2015). Total sugar yield was dependent on the treat-
ment conditions (Fig. 3.7) and varied from as low as 4.5 g/L up to 21.7 g/L which
was close to the yield predicted using Box–Behnken design.
Since coir has considerable amounts of lignin and is difficult to be digested, a
treatment with ionic liquid was done to improve digestion and subsequent yield of
ethanol (Ribeiro et al. 2018). In this approach, coir fibers were dried at 80 °C for
72 h and later milled to form a powder. The treated fibers were pulped using 9%
64 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

Fig. 3.6 Variations in the


yields of sugar and ethanol
from mature coconut fibers
using three different
microorganisms. (a) is S.
cerevisiae, (b) is P. stipitis,
and (c) is Z. mobilis
(Gonçalves et al. 2016).
Reproduced with
permission from Elsevier
3.4 Substrate for Preparation of Catalysts 65

Table 3.5 Yield of ethanol from hydrothermally pretreated mature coir fibers with different
processes and microorganisms (Gonçalves et al. 2016)
Ethanol Ethanol Ethanol productivity,
Process Microorganism yield, % concentration, % (g/Lh)
SSF S. cerevisiae 85 ± 1.2 10.9 ± 0.15 0.30 ± 0.15
P. stipitis 87 ± 1.8 10.9 ± 0.22 0.23 ± 0.22
Z. mobilis 85 ± 1.7 10.8 ± 0.2 0.30 ± 0.21
SSSF S. cerevisiae 91 ± 1.5 11.7 ± 0.19 0.32 ± 0.19
P. stipitis 90 ± 1.9 11.3 ± 0.24 0.24 ± 0.24
Z. mobilis 91 ± 0.7 11.6 ± 0.09 0.32 ± 0.09
Reproduced with permission from Elsevier

(w/w) sodium hydroxide solution at temperature of 137 °C and pressure of 2.5


atmospheres. After pulping, the material was treated with the ionic liquid
n-­butylammonium acetate at various temperature and reaction times. Conversion of
the pulp into glucose showed significant difference depending on the reaction time
and temperature (Table 3.8). Similar changes in % glucose yield from 0.5 to 27.0%
were observed when the pH and temperature were varied. A highest yield of 32%
was obtained after the coir was treated with ionic liquids and subsequently used for
ethanol production. In another study, coconut coir dust (Sangian et al. 2015) was
treated with ionic liquids to improve hydrolysis and ethanol yield. The ability of
methymethylimidazolium dimethyl phosphate [(Immim) (dmp)] to be recycled and
reused for the hydrolysis at 60 °C at pH 3 for 48 h was investigated. Yield of sugar
was between 0.15 and 0.19 g per gram of cellulose + hemicellulose when the fibers
were not pretreated with alkali but increased to 0.28 g/g when pretreated with alkali.
It was suggested that the ability to increase the yields and use recycled ionic liquids
will help to decrease the cost of conversion and make biofuel production from coir
more profitable (Sangian et al. 2015). Ionic liquids were also used to pretreat coir
fibers before acid hydrolysis (de Andrade Neto et al. 2016). Coir fibers chopped into
3 mm particles were pulped by treating in 8% sodium hydroxide for 6 h at
137 °C. The pulped coir fiber was treated with the ionic liquid n-butylammonium
acetate for 2 h at 90 °C and later hydrolyzed using acidic conditions. A maximum
glucose yield of 22% was obtained when 37.5% acetic acid was used and the yield
obtained was about 800% higher than that obtained from the untreated sample (de
Andrade Neto et al. 2016).

3.4 Substrate for Preparation of Catalysts

In a unique study, coir was used as a substrate to produce solid acid catalysts which
were used to improve the conversion of waste palm oil into biodiesel (Thushari and
Babel 2018). The catalysts were prepared by burning the coir husks in a muffle
66
3

Table 3.6 Changes in the composition and digestibility of coir fibers after different pretreatments (Ebrahimi et al. 2018)
Ammonium carbonate, Treatment Solid recovery, Coir composition, % Glucan Lignin Lignin Glucan
w/w, % time % Glucan Lignin Xylan recovery, % recovery, % removal, % digestibility, %
20 4 76.4 33.3 31.9 16.5 73.2 68.1 31.9 61.2
8 74.6 39.9 27.7 13.5 85.5 57.8 42.2 77.0
12 69.8 36.8 21.4 11.2 73.4 41.7 58.4 70.4
25 4 75.7 34.6 32.2 15.6 75.2 68.0 31.9 63.4
8 72.2 38.6 24.1 11.9 80.0 48.6 51.4 71.4
12 65.9 34.2 19.5 9.9 64.7 35.9 64.2 68.9
Untreated 100 34.8 35.8 18.3 100 – – 11.8
Reproduced with permission from Taylor and Francis
Biotechnological Applications for Coir and Other Coconut Tree By-products
3.4 Substrate for Preparation of Catalysts 67

Table 3.7 Comparison of the production of bioethanol from coir fibers using different pretreatment
conditions using Saccharomyces cerevisiae (Ebrahimi et al. 2018)
Composition,
% Process/ Bioethanol
Raw material Pretreatment Glucan Lignin condition yield
Mature coir H2O2-NaOH, 25 °C 51.8 8.9 SSF, 30 °C 8.4 g/L
fiber
Green coir shell Autohydrolysis, 200 °C 41.9 41.3 SSF, 30 °C 7.4 g/L
Mature coir Hydrothermal-NaOH, 56.4 11.7 SSF, 30 °C 10.9 g/L
fiber 180 °C
Coconut husk NaOH, 121 °C 55.2 29.9 SSF, 30 °C 7.9 g/L
Coconut husk NaOH, 100 °C – – SSF, 37 °C 1 g/100 mL
Coconut coir Acidified glycerol, 130 °C 48.6 32.8 SSF, 37 °C 8.97 g/L
fiber
Coconut coir Ammonium carbonate, 39.9 27.7 SF, 37 °C 8.7 g/L
fiber 80 °C
Reproduced with permission from Taylor and Francis

Fig. 3.7 Effect of acid concentration and hydrolysis time on the yield of sugars from the biomass
(Amenaghawon et al. 2015). Open Access Publication

Table 3.8 Ability to generate Temperature,


glucose from coir fiber pulp Time, h °C Glucose, %
after subjecting to various
1.37 70.00 8.68 ± 0.04
treatment times and
temperature (Ribeiro et al. 8.00 50.00 9.67 ± 0.59
2018) 8.00 90.00 13.68 ± 0.23
24.00 98.30 13.13 ± 0.16
24.00 70.00 11.73 ± 0.05
24.00 70.00 11.00 ± 0.01
24.00 42.00 9.11 ± 0.05
40.00 50.00 9.39 ± 0.60
40.00 90.00 17.79 ± 0.17
46.63 70.00 11.90 ± 0.15
Reproduced with permission from Springer
Nature
68 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

furnace up to 400 °C for 5 h. Later, the carbonized powder was mixed with 25%
sulfuric acid and shaken at 200 °C for 24 h for sulfonation to occur. The catalyst
obtained was neutralized and used for the biodiesel production. A surface area of
61 m2/g, mean pore volume of 14 cm3/g, and mean pore diameter of 2.6 nm were
obtained for the catalyst. A fatty acid methyl ester (FAME) yield of up to 85% was
obtained depending on the catalyst concentration and reaction conditions (Thushari
and Babel 2018), suggesting that coir husk can be used as a potential carbon source
for preparing catalysts for biodiesel production.

3.5 Chemicals and Pharmaceuticals

Efforts have also been made to understand the capability of various coconut by-
products to produce chemicals and products. Potential of producing vanillin, a valu-
able aromatic compound used in food, beverages, and pharmaceutical industries,
was studied (Barbosa et al. 2008). Fermentation was done on powdered coconut
husks using Phanerochaete chrysosporium and culturing at 28 °C for 5 days.
Amount of vanillin produced was determined using HPLC and LCMS. Concentration
of vanillin obtained varied depending on the form of coconut husk (sun-dried or
mechanically pressed) and the incubation time. Up to 45 μg of vanillin per gram of
coconut fiber used was obtained after 24 h of incubation (Fig. 3.8) (Barbosa et al.
2008).
A medicinally important drug kojic acid that is used as pain killer, anti-­
inflammatory and antimicrobial drug was synthesized using coir and other bio-
masses as substrates and Aspergillus oryzae RMS2 as the microorganism (Kumar
and Jayalakshmi 2016). Up to 106 g of kojic acid was produced per kilogram of the
biomass source used after incubation for 14 days, pH 5.0 at 30 °C. It was suggested
that coir was an efficient and inexpensive source to produce kojic acid.

Fig. 3.8 Vanillin 50


concentration (µg/g of the support)

production from coconut sun-dried


shells was influenced by mechanical-pressed
40
the pretreatment of the
husk and the incubation
time (Barbosa et al. 2008). 30
Reproduced with open
access license from North 20
Carolina State University

10

0
24 48 72 96
Time (hours)
3.6 Coir Pith as Substrate for Production of Mushrooms 69

Table 3.9 Parameters for production of lactic acid from mutant strain M6 (Almaliki et al. 2016)
Final cell Consumed xylose Lactic acid yield Acetic acid yield
Parameter Value growth (g/L) (g/L) (g/L)
pH 5 1.7 21.1 4.9 1.2
6 3.1 67 21.7 12.1
7 3.4 94 40.7 31.1
8 2.7 48.4 23.1 14.5
Temperature 30 3.5 94.9 38.2 28.5
(°C) 35 3.5 96.8 42.4 30.0
40 2.8 43.1 18.4 11.0
45 1.1 8.1 4.1 2.2
Agitation (rpm) 50 3.5 93.4 38.2 30.0
75 3.4 94.8 40.2 33.1
100 3.5 96.5 43.0 31.7
150 3.4 94.5 39.2 32.0
Reproduced with permission through Creative Commons Attribution 3.0 License

Effluent from the coir processing industry was used to isolate various strains of
bacteria that could convert xylose into lactic acid. One particular strain (L-3) pro-
duced considerably higher lactic acid (15 mM) and acetic acid (11 mM). This gram
positive strain identified as Lactobacillus, particularly Lactobacillus sp. 16 s rRNa
had cell size of 1.3–0.9 and could grow under both micro-aerobic and anaerobic
conditions. Effect of the production conditions on yield of lactic and acetic acid is
shown in Table 3.9 (Almalki 2016).

3.6 Coir Pith as Substrate for Production of Mushrooms

Coir pith was studied as a substrate for its potential to grow mushroom using mono-
cultures of Pleurotus florida (Shashirekha and Rajarathnam 2007). The coir pith
was chopped and treated with hot water at 70 °C for 20 min. Beds of the pith were
prepared and the culture was maintained until the mushrooms spawned and grew to
the desired extent. Coir pith was also treated by soaking in 40 L of water containing
0.1–2% of phosphoric acid or NaOH for 3 h at room temperature. Further, the pith
was mixed with rice straw and/or horse gram to promote higher growth. Yield of
mushrooms (g/kg of dry substrate) varied from 200 to 350 but was much lower than
that obtained using rice straw (1000 g/kg). However, supplementing mushroom
with rice straw increased the yields substantially even up to 1250 g/kg (Shashirekha
and Rajarathnam 2007).
Coir fibers were studied for their potential to be used as substrates for production
of poly-β-hydroxybutyrate (PHB) (Sathesh Prabu and Murugesan 2010). The fibers
were first delignified by autoclaving at 120 °C for 20 min and later hydrolyzed using
cellulase at 37 °C for 24 h. A culture of A. beijerinckii (5% V/V) was grown for the
70 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

production of PHB in the culture medium at 37 °C for 48 h. Obtained coir was


treated with sodium hypochlorite and later dried and stored. It was found that A.
beiherinckii was capable to produce 2.4 g/L of PHB with a yield of 48%. In another
study, the strain KSN5 Bacillus sp. was used to produce polyhydroxyalkanoate
(PHA) using coir pith as a source for carbon (Kalaivani 2016). Depending on the
amount of carbon source (coir pith), 19–38% of PHA was generated similar to that
seen when coir fibers were used.
In addition to fruits and vegetables, coco peat has also been considered for sup-
plementing the growth of rose plants (Roosta and Rezaei 2014). Rose plants were
cultivated in pots containing 75% peat and 25% perlite and were nourished with
nutrients with pH variations from 4.5 to 8. Many parameters including number of
buds, chlorophyll content, leaf phosphorous, iron, manganese, and copper were
highest at pH 6.5 and lowest at pH 8.0. Coco peat supplemented growth media was
suggested to be ideal for growing rose plants (Roosta and Rezaei 2014).

3.7 Purification of Apple Juice

In a unique application, coir fibers were used as substrate to immobilize laccase,


which was then used for purification of apple juice. Fibers were first dried and later
treated with 2% sulfuric acid or 2% NaOH and heated to 121 °C for 15 min. Residue
obtained was further treated with ethanol to remove impurities. Activation of the
fibers was done by treating the hydrolyzed and milled residue with 1.7 M NaOH
containing 0.792 g of sodium borohydride (de Souza Bezerra et al. 2015). The coir
fibers were further activated with glyceryl or glyoxyl groups by addition of 1.43 mL
of NaIO4 solution. Similarly, the amino groups were activated by treating with gly-
oxal and ethylenediamine (1 M at pH 10 for 2 h). Glutaraldehyde was also immobi-
lized on the coir fibers by treating with 0.05 M glutaraldehyde for 2 h and then
washed with 20 volumes of phosphate buffer and water. Some of the properties of
the laccase from T. versicolor are given in Table 3.10. It was concluded that the low

Table 3.10 Immobilization yield (IY), expressed activity (EA), and stabilization factor (SF)
obtained from the coir fibers modified with various approaches (de Souza Bezerra et al. 2015)
Half-time,
Derivatives IY EA OT Op hours SF
Glycerol 60 ± 0.1 59 ± 0.9 40 ± 0.2 3.0 ± 0.02 53.5 ± 0.03 16.5 ± 0.00
Glyoxyl 90 ± 0.9 40 ± 0.2 60 ± 0.08 3.0 ± 0.04 18.6 ± 0.01 5.8 ± 0.05
Glutaraldehyde— 97 ± 0.6 54 ± 0.1 50 ± 0.05 2.2 ± 0.01 22.2 ± 0.05 6.9 ± 0.1
alkali treated
Glutaraldehyde— 98 ± 1.0 44 ± 0.0 40 ± 0.15 3.0 ± 0.01 5.10 ± 0.08 1.6 ± 0.08
acid treated
Reproduced through open license access
3.8 Removal of Urea 71

cost of coir could be used to remove phenolic compounds from apple juice and later
the enzyme could also be easily removed since it was attached to the substrate.

3.8 Removal of Urea

Activated carbon obtained from coconut shells was used to immobilize urease
which was then used to remove urea from a packed bioreactor (Wang et al. 2013).
Urease was extracted from jack bean with a molecular weight of about 480 kDa and
50,000–100,000 units/g. Activation of the coconut shells was done by oxidation,
aminosilylation, and aldehyde linking as seen from Fig. 3.9. For the immobilization
of the urea, the activated fibers were placed in 0.5 mg/mL of urease solution and
shaken in an incubator for 48 h at 4 °C. Concentration of the enzymes was deter-
mined at an absorbance of 275 nm. Amount of bound urease per gram of substrate
was used to define the loading capacity (Wang et al. 2013). Loading capacity of the
carbon was found to be 79 mg/g and the enzymes attached to the carbon showed a
maximum activity at 70 °C at pH 7.2. It was suggested that catalytic degradation of
urea could be effectively achieved in the packed bed reactor. More importantly, the
urease on the activated carbon retained more than 80% of the activity and was hence
reusable.

Fig. 3.9 Process of preparing the coconut shell activated carbon into laccase-modified sorbent for
removal of urea (Wang et al. 2013). Reproduced with permission from American Chemical Society
72 3 Biotechnological Applications for Coir and Other Coconut Tree By-products

References

Almalki MA (2016) Production of medically important lactic acid by lactobacillus pentosus: a


biological conversion method. Indian J Sci Technol 9(4):1–8
Amenaghawon NA, Osayuki-Aguebor O, Egharevba LP (2015) Effect of acid mixtures on the
hydrolysis of coconut coir for recovery of fermentable sugars. J Appl Sci Environ Manag
19(3):345–351
Anuradha Jabasingh S, Varma S, Garre P (2014) Production and purification of cellulase from
Aspergillus nidulans AJSU04 under solid-state fermentation using coir pith. Chem Biochem
Eng Q 28(1):143–151
Asha P, Divya J, Bright Singh IS (2016) Purification and characterisation of processive-type endo-
glucanase and β-glucosidase from Aspergillus ochraceus MTCC 1810 through saccharification
of delignified coir pith to glucose. Bioresour Technol 213:245–248
Barbosa d S, Elisabete DP, Vendramini AL d A, Leite SGF (2008) Vanillin production by
Phanerochaete chrysosporium grown on green coconut agro-industrial husk in solid state fer-
mentation. Bioresources 3(4):1042–1050
Bezerra d S, Milena T, Bassan JC, Santos VT d O, Ferraz A, Monti R (2015) Covalent immobiliza-
tion of laccase in green coconut fiber and use in clarification of apple juice. Process Biochem
50(3):417–423
de Andrade Neto JC, de Souza Cabral A, de Oliveira LRD, Torres RB, Morandim-Giannetti A d
A (2016) Synthesis and characterization of new low-cost ILs based on butylammonium cation
and application to lignocellulose hydrolysis. Carbohydr Polym 143:279–287
Ebrahimi M, Caparanga AR, Ordono EE, Villaflores OB (2017) Evaluation of organosolv pre-
treatment on the enzymatic digestibility of coconut coir fibers and bioethanol production via
simultaneous saccharification and fermentation. Renew Energy 109:41–48
Ebrahimi M, Caparanga AR, Villaflores OB (2018) Weak base pretreatment on coconut coir
fibers for ethanol production using a simultaneous saccharification and fermentation process.
Biofuels:1–7. https://doi.org/10.1080/17597269.2018.1468979
Fatmawati A, Agustriyanto R, Liasari Y (2013) Enzymatic hydrolysis of alkaline pretreated coco-
nut coir. Bull Chem React Eng Catal 8(1):34–39
Gonçalves FA, Ruiz HA, Nogueira C d C, dos Santos ES, Teixeira JA, de Macedo GR (2014)
Comparison of delignified coconuts waste and cactus for fuel-ethanol production by the simul-
taneous and semi-simultaneous saccharification and fermentation strategies. Fuel 131:66–76
Gonçalves FA, Ruiz HA, dos Santos ES, Teixeira JA, de Macedo GR (2015) Bioethanol production
from coconuts and cactus pretreated by autohydrolysis. Ind Crop Prod 77:1–12
Gonçalves FA, Ruiz HA, Silvino dos Santos E, Teixeira JA, de Macedo GR (2016) Bioethanol
production by Saccharomyces cerevisiae, Pichia stipitis and Zymomonas mobilis from deligni-
fied coconut fibre mature and lignin extraction according to biorefinery concept. Renew Energy
94:353–365
Immanuel G, Dhanusha R, Prema P, Palavesam A (2006) Effect of different growth parameters
on endoglucanase enzyme activity by bacteria isolated from coir retting effluents of estuarine
environment. Int J Environ Sci Technol 3(1):25–34
Immanuel G, Bhagavath CMA, Iyappa Raj P, Esakkiraj P, Palavesam A (2007) Production and
partial purification of cellulase by Aspergillus Niger and A. fumigatus fermented in coir waste
and sawdust. Internet J Microbiol 3(1):1–20
Kalaivani R (2016) A study on impact of agronomical carbon sources in bacterially produced
PHA. Int J Curr Micobiol Appl Sci 5(12):910–918
Kumar CR, Jayalakshmi S (2016) Use of coconut coir fibers as an inert solid support for kojic acid
production under solid state fermnetion. Int J Curr Microbial Appl Sci 6(12):256–264
Montagne V, Capiaux H, Barret M, Cannavo P, Charpentier S, Grosbellet C, Lebeau T (2017)
Bacterial and fungal communities vary with the type of organic substrate: implications for
biocontrol of soilless crops. Environ Chem Lett 15(3):537–545
References 73

Mrudula S, Murugammal R (2011) Production of cellulase by Aspergillus Niger under submerged


and solid state fermentation using coir waste as a substrate. Braz J Microbiol 42(3):1119–1127
Muniswaran PKA, Charyulu NCLN (1994) Solid substrate fermentation of coconut coir pith for
cellulase production. Enzym Microb Technol 16(5):436–440
Muniswaran PK, Selvakumar P, Charyulu NCL (1994) Production of cellulases from coconut coir
pith in solid state fermentation. J Chem Technol Biotechnol 60(2):147–151
Namasivayam SKR, Babu M, Arvind Bharani RS (2015) Evaluation of lignocellulosic agro wastes
for the enhanced production of extracellular cellulase and Xylanase by Trichoderma harzia-
num. Nat Environ Pollut Technol 14(1):47–56
Packiyam JE, Ragunathan R (2017) Production, purification, characterization, and copper induc-
tion of laccase isoenzyme in the lignolytic fungus pleurotus Florida. J Microbiol Biotechnol
Res 4(2):10–16
Ramachandran S, Patel AK, Nampoothiri KM, Francis F, Nagy V, Szakacs G, Pandey A (2004)
Coconut oil cake––a potential raw material for the production of α-amylase. Bioresour Technol
93(2):169–174
Ribeiro d O, Cesar W, Lima AC d S, Morandim-Giannetti A d A (2018) Optimizing treatment con-
dition of coir fiber with ionic liquid and subsequent enzymatic hydrolysis for future bioethanol
production. Cellulose 25(1):527–536
Roosta HR, Rezaei I (2014) Effect of nutrient solution pH on the vegetative and reproductive
growth and physiological characteristics of rose cv. ‘Grand Gala’ in hydroponic system. J Plant
Nutr 37(1):2179–2194
Sangian HF, Kristian J, Rahma S, Dewi HK, Puspasari DA, Agnesty SY, Gunawan S, Widjaja A
(2015) Preparation of reducing sugar hydrolyzed from high-lignin coconut coir dust pretreated
by the recycled ionic liquid [mmim][dmp] and combination with alkaline. Bull Chem React
Eng Catal 10(1):8–22
Sathesh Prabu C, Murugesan AG (2010) Effective utilization and Management of Coir Industrial
waste for the production of poly-Î2-hydroxybutyrate (PHB) using the bacterium Azotobacter
Beijerinickii. Int J Environ Res 4(3):519–524
Shashirekha MN, Rajarathnam S (2007) Bioconversion and biotransformation of coir pith for eco-
nomic production of Pleurotus Florida: chemical and biochemical changes in coir pith during
the mushroom growth and fructification. World J Microbiol Biotechnol 23(8):1107–1114
Soares J, Demeke MM, Foulquié-Moreno MR, Van de Velde M, Verplaetse A, Fernandes AAR,
Thevelein JM, Fernandes PMB (2016) Green coconut mesocarp pretreated by an alkaline pro-
cess as raw material for bioethanol production. Bioresour Technol 216:744–753
Soares J, Demeke MM, Van de Velde M, Foulquie-Moreno MR, Kerstens D, Sels BF, Verplaetse A,
Fernandes AAR, Thevelein JM, Fernandes PMB (2017) Fed-batch production of green coconut
hydrolysates for high-gravity second-generation bioethanol fermentation with cellulosic yeast.
Bioresour Technol 244:234–242
Subhedar PB, Ray P, Gogate PR (2018) Intensification of delignification and subsequent hydroly-
sis for the fermentable sugar production from lignocellulosic biomass using ultrasonic irradia-
tion. Ultrason Sonochem 40:140–150
Thushari I, Babel S (2018) Preparation of solid acid catalysts from waste biomass and their appli-
cation for microwave-assisted biodiesel production from waste palm oil. Waste Manag Res
36(8):719–728
Vaithanomsat P, Apiwatanapiwat W, Chumchuent N, Kongtud W, Sundhrarajun S (2011) The
potential of coconut husk utilization for bioethanol production. Kasetsart J 45:159–164
Wang L, Wang S, Deng X, Zhang Y, Xiong C (2013) Development of coconut shell activated
carbon-tethered urease for degradation of urea in a packed bed. ACS Sustain Chem Eng
2(3):433–439
Chapter 4
Applications of Coir Fibers
in Construction

4.1 Introduction

Coir fibers have been extensively studied for application in civil and construction
materials. Primarily, the fibers are used as reinforcement in cement or as replace-
ment for aggregates to improve tensile properties, reduce moisture absorption, and
improve other performance properties. In one such study, coir fibers were cut to
about 2 cm in length and treated with 5% alkali solution for 2 h at room tempera-
ture. Mortar mixtures were prepared according to TS EN 196-1 standard in which
1350 g of reference sand, 450 g of cement, 225 g of water, and different levels of
fibers (0.4, 0.6, and 0.75%) were used. The mixed mortars were cast into
40 × 40 × 160 mm3 molds and allowed to stay for 24 h. Cast molds were later cured
in water for 28 days before testing for their properties (Andiç-Çakir et al. 2014).
Alkali-treated fibers increased amount of O containing groups in the surface of coir
fibers. Similarly, C-O groups increase but the C-C groups decrease, making the fiber
more hydrophilic. Hence, the water absorption of mortar containing untreated fibers
decreased from 6.3 to 3.9%, whereas that of treated fibers increased from 6.5 to
8.9%. Compressive strength of the mortar showed marginal increase between 45
and 52 MPa, but there was a large increase in toughness from about 28–52 MPa mm1/2
(Fig. 4.1). Thermal conductivity of the mortar showed a marginal change from 1.8
to 1.767 W/mK when the amount of fibers was 0.75% (Andiç-Çakir et al. 2014). In
another study, coir fibers having an average length of 50 mm was treated using 5%
NaOH solution at 20 °C for 30 min and later thoroughly washed to remove any
alkali present. Treated fibers were combined with plain concrete using a
cement:water:gravel:sand ratio of 1:0.6:3.7:2.46. Amount of fibers used in the
cement was controlled at 1%. Cylinders having diameters of 100 mm and height of
200 mm and beams having length of 500 mm, width of 100 mm, and depth of
50 mm were fabricated. Changes in the properties of the cementitious composites
containing treated and untreated fibers are given in Table 4.1. As seen from the
table, fracture energy and flexural toughness increased to 550 and 424%,

© Springer Nature Switzerland AG 2019 75


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_4
76 4 Applications of Coir Fibers in Construction

Fig. 4.1 Changes in the toughness of mortar reinforced with various amounts of coir fibers
(Andiç-Çakir et al. 2014). Reproduced with permission from Elsevier

respectively, due to the addition of coir fibers. Coir fibers were able to bridge adja-
cent fibers and micro-cracks and served as secondary reinforcement and prevented
crack propagation and failure of the beam (Fig. 4.2). Alkali treatment further
increased the properties of the concrete but did not change the failure pattern.
However, fiber breakage, fiber pullout, and debonding were noticed, indicating that
the fibers were not very compatible with the cement matrix (Yan et al. 2016).
The long-term performance of coir fiber reinforced concrete was studied by Li
et al. (2012). Coir fibers were treated with 1% NaOH for 48 h at 20 °C and later
trimmed into 40 mm and 20 mm length and mixed with mortar in a ratio of 1:3:0.43
(cement:sand:water). Samples of 50 × 50 × 175 mm3 were developed and condi-
tioned for 24 h. Inclusion of the fibers increased flexural strength and flexural tough-
ness with initial curing but decreased as the ageing time increased (Fig. 4.3). Coir
fiber reinforced cementitious composites had up to 734% increase in toughness
index but the long-term performance (400 days) was also dependent on the type of
ageing. Freeze-thaw ageing had lower effect on the composites compared to water-­
ageing or natural weather fluctuations (Li et al. 2012).
Coir fibers of 20 mm length were also used (0.5 or 1%) as reinforcement in com-
parison with steel (0.5 and 1%) to prepare normal strength and high strength con-
crete (Islam et al. 2012). Compressive strength of the normal strength concrete
decreased with increasing ratio of fibers. At 0.5% reinforcement, steel and coir pro-
vided similar compressive strength improvement to the normal strength concrete
(9.8 and 8.2 MPa, respectively). However, the extent of increase was dependent on
4.1 Introduction

Table 4.1 Properties of concrete reinforced with untreated and treated coir fibers (Yan et al. 2016)
Compressive strength, Modulus of Compressive Energy absorption, Flexural stress, Fracture energy, Flexural
Specimen MPa elasticity, GPa strain, % N.m MPa N.m toughness,
Unreinforced 22.4 ± 0.8 28.6 ± 1.0 0.2 ± 0.2 32.7 ± 5.2 9.8 ± 0.8 1.6 ± 0.4 1.21
Raw fibers 23.8 ± 1.3 27.4 ± 1.8 0.29 ± 0.6 54.5 ± 8.9 11.2 ± 1.0 10.4 ± 2.6 6.34
Treated 24.0 ± 1.4 27.9 ± 1.7 0.34 ± 0.7 58.4 ± 9.7 11.9 ± 1.1 11.3 ± 2.0 6.64
fibers
Reproduced with permission from Elsevier
77
78 4 Applications of Coir Fibers in Construction

Fig. 4.2 Coir fibers are able to bridge adjacent cracks and prevent complete failure of the beams.
Left is CFRC and right is unreinforced concrete (Yan et al. 2016). Reproduced with permission
from Elsevier

Fig. 4.3 Changes in the strength and toughness of the composites reinforced with untreated and
alkali-treated 20 and 40 mm length coir fibers (Li et al. 2012). Reproduced with permission from
Taylor and Francis
4.1 Introduction 79

the number of curing days. After 30 days of curing, the strength of the steel and coir
reinforcements were similar but after 90 days, the reinforcements did not provide
adequate strength compared to the plain concrete. Although compressive strength of
high strength concrete was lower when reinforced with coir fibers, the flexural
strength was higher even compared to steel reinforcement. However, reinforcing
with coir fibers decreased workability of concrete but could be useful to improve
ductility and toughness (Islam et al. 2012).
Most studies on using coir fibers as reinforcement for concrete have been done
with natural sand as one of the major components. Due to the decreasing availability
and restrictions on mining of natural sand, manufactured sand (M-sand) is gaining
importance. Ability of coir fibers to reinforce concrete made from M-sand was
investigated by varying its ratio with river sand (Neeraja et al. 2017). Compressive
strength (Fig. 4.4) and split tensile strength of the concrete increased, whereas com-
pressive strength decreased with increase in the amount of fibers (Neeraja et al.
2017). Split strength of the concrete having 1% coir fibers was highest at 3.28 N/
mm2 after 28 days of curing.
Coir fibers were studied for their potential to replace coarse aggregates in cement
panels. Untreated and fibers washed in water were used along with cement and sand
to prepare composite panels. Weight of the fibers was changed from 0 to 9% with
corresponding decrease in the quantity of sand and maintaining the same level of
cement (Abdullah et al. 2011). Addition of untreated coir increased compressive
strength and water absorption, whereas the treated coir fibers considerably decreased
the strength of the concrete panels. A maximum compressive strength of 43.8 MPa
was achieved. Moisture sorption was 4.1% for untreated and 12.3% for treated
fibers (Abdullah et al. 2011). Treatment of coir fibers was also considered necessary
to improve setting of cement. The fibers were either treated in room temperature
water for 72 h, immersed in 80 °C water for 90 min or exposed to 5% NaOH solu-
tion for 72 h to reduce the lignin, holocellulose, and other extractives (Ferraz et al.
2012). CaCl2 was added to the concrete containing untreated fibers to simulate prac-

60

50
Strength in N/mm2

40
7 days curing
30
28 days curing
20

10

0
1 2 3 4 5
Sample No.

Fig. 4.4 Compressive strength of concrete reinforced with 1% coir fibers after 7 and 28 days of
curing (Neeraja et al. 2017). Reproduced under open access publication
80 4 Applications of Coir Fibers in Construction

Table 4.2 Inhibitory index between untreated and treated coir fibers and cement (Ferraz et al.
2012)
Time for max Maximum Max increment Inhibitory
Treatment temp, °C temperature, °C temperature, °C index, %
Cement 11.00 55.06 – –
Untreated 1.08 30.47 28.21 186.28
fibers
Cold water 20.02 32.92 1.64 20.85
treated
80 °C water 17.56 34.15 1.94 14.07
CaCl2 14.86 42.88 2.89 2.97
NaOH 12.87 46.96 3.65 2.24
Reproduced under Creative Commons Attribution License

tical use conditions. Treating with NaOH was more effective in reducing the lignin
and increasing holocellulose content. Inhibitory index (Table 4.2) which suggests
the binding affinity between cement and the fibers was lowest for the NaOH-treated
fibers due to the surface becoming rougher and fibers becoming finer (Ferraz et al.
2012). Various chemical treatments were performed on coir fibers to develop
­tamarind starch based low-cost housing composites (Kiruthika and Veluraja 2017).
Tensile strength and elongation of the coir fibers were considerably affected by the
treatments. Hence, the composites developed had tensile strength ranging from 1.7
to 3.0 MPa and elongation from 22 to 40.5% depending on the type of coating used.
The composites developed had poor durability but were found to resist water and
considered suitable for use in outdoor environments (Kiruthika and Veluraja 2017).
Similar to reinforcing concrete, coir fibers were found to reduce brittleness and
improve peformance properties of mortar used as plaster. Coir fibers having length
of 20 mm and diameter of 20 μm were used for the study (Sathiparan et al. 2017).
To prepare the mortar, cement, lime, and sand along with the fibers (0.125–0.75%)
were mixed together. Higher fiber contents decreased the workability of the mortar
and were hence avoided. Cubes having dimensions of 150 × 150 × 150 mm3 and
beam prisms of 100 × 100 × 500 mm3 were developed using the mixture. Both the
bulk and dry density of the mortar decreased with increase in fiber content, whereas
water sorption and porosity steadily increased. However, compressive strength and
flexural strength increased initially but decreased at higher fiber content (Sathiparan
et al. 2017). Flexural ductility and mortar toughness which are important for plaster
applications increased. As seen from Fig. 4.5, mortar containing coir fibers provided
considerable resistance to compression. Marginal decrease in the alkaline resistance
of the mortar was also observed. Coir and polypropylene fibers were combined
together to develop fiber reinforced concrete. Coir fibers used had length of 100 mm,
diameter of 200 μm, and tensile strength of 140–150 MPa (Gnanapragasam and
Jesuraj 2016). Volume fraction of the fibers in the concrete was varied between 0.15
and 0.45%. Compressive and split tensile strength of the concrete was dependent on
the amount of fibers and curing time (Table 4.3). A combination of coir and poly-
propylene was found to provide better workability for the concrete. In another study,
4.1 Introduction 81

Fig. 4.5 Image depicting the better resistance of coir fiber reinforced (0.5%) mortar (b) compared
to unreinforced mortar (a) (Sathiparan et al. 2017). Reproduced with permission from Elsevier

Table 4.3 Increase in the compressive and tensile strength of concrete reinforced with various
ratios of coir fibers (Gnanapragasam and Jesuraj 2016)
Compressive strength, Split tensile strength,
MPa MPa
Curing time, days Fiber volume fraction 1 2 3 1 2 3
7 0.15 20.89 21.78 21.33 2.55 2.69 2.69
0.25 22.67 22.67 23.56 2.83 2.83 2.97
0.35 24 23.56 24.44 3.11 3.26 3.26
0.45 25.3 26.77 27.11 3.96 3.68 3.68
14 0.15 24.56 24.89 25.33 2.97 3.26 3.11
0.25 26.22 26.77 27.11 3.4 3.54 3.54
0.35 28.44 29.78 28.89 3.68 3.96 3.96
0.45 29.78 30 31.56 4.25 4.10 3.96
28 0.15 27.56 27.11 28.44 3.68 3.54 3.68
0.25 30.22 29.78 30.22 3.82 3.96 3.96
0.35 32.44 32.0 31.56 4.1 4.10 4.25
0.45 34.22 33.78 34.67 4.39 4.39 4.53

a combination of coir and polypropylene fibers provided a compressive strength


between 32.9 and 40.9 MPa and flexural strength between 7.7 and 13.1 MPa
(Sarangi and Sinha 2016).
In addition to reinforcing concrete with coir by direct addition, external strength-
ening of the beams was done using multiple layers of flax fiber and coir reinforced
epoxy composites. Coir fibers were washed with water and found to have tensile
modulus of 2.74 GPa, tensile stress of 286 MPa, and strain at break of 20.8%. These
fibers were added into cement, sand, and aggregates to prepare concrete beams (Yan
82 4 Applications of Coir Fibers in Construction

Table 4.4 Comparison of the performance of unreinforced, coir fiber reinforced and flax fabric-­
epoxy strengthened beams (Yan et al. 2015)
Type of Peak load, Flexural stress, Fracture Deflection at peak
Specimen bending kN MPa energy, J load, mm
Plain concrete 3-point 3.7 ± 0.3 10.0 ± 0.8 0.5 ± 0.3 0.7
Plain concrete 4-point 5.4 ± 0.2 9.8 ± 0.4 1.5 ± 0.8 0.8
Coir 3-point 4.0 ± 0.5 10.8 ± 1.4 7.4 ± 2.4 0.8
reinforced
Coir 4-point 6.2 ± 0.8 11.2 ± 1.4 10.4 ± 1.9 0.9
reinforced
2L plain 3-point 7.4 ± 0.7 19.9 ± 1.9 18.0 ± 2.8 2.8
concrete
2L plain 4-point 12.9 ± 1.0 23.2 ± 1.8 28.1 ± 3.8 3.2
concrete
2L coir 3-point 10.7 ± 2.5 28.9 ± 5.4 59.4 ± 9.2 3.9
reinforced
2L coir 4-point 19.8 ± 2.0 35.6 ± 3.6 69.3 ± 7.8 4.2
reinforced
4L plain 3-point 12.2 ± 1.3 32.9 ± 3.5 21.5 ± 3.4 2.7
concrete
4L plain 4-point 19.2 ± 1.7 34.6 ± 3.1 38.9 ± 4.4 3.2
concrete
4L coir 3-point 18.0 ± 2.8 48.6 ± 7.5 72.1 ± 10.6 4.1
reinforced
4L coir 4-point 26.4 ± 3.3 47.5 ± 5.9 85.5 ± 13.5 4.6
reinforced
6L plain 3-point 15.8 ± 2.1 42.6 ± 5.6 41.3 ± 6.7 3.4
concrete
6L plain 4-point 25.4 ± 2.4 45.7 ± 4.3 147.6 ± 19.6 4.1
concrete
6L coir 3-point 23.6 ± 2.0 63.7 ± 5.4 86.1 ± 13.2 4.5
reinforced
6L coir 4-point 31.0 ± 3.9 55.8 ± 7.0 168.4 ± 21.5 5.2
reinforced
Reproduced with permission from Elsevier

et al. 2015). Similarly, flax fiber reinforced epoxy composites were prepared using
a bidirectional flax fabric having weight/unit area of 550 g/m2, tensile strength of
154 MPa compared to tensile strength of epoxy which was 87.8 MPa. The epoxy
was reinforced with 2, 4, or 6 layers of the fabric. Concrete beams cured for 28 days
were polished and later externally glued with the epoxy-flax composites. Substantial
improvement (Table 4.4) in the properties of the concrete beams was observed due
to the inclusion of coir fibers and particularly with the external composite
­reinforcement. Extent of improvement was dependent on the number of layers of
flax fabric in the epoxy composite (Fig. 4.6). Concrete beams reinforced with coir
fibers showed better performance than those without the fibers. The rough surface
of coir fibers was found to provide good interfacial adhesion for the cement matrix
4.1 Introduction 83

Fig. 4.6 Reinforcing concrete with coir provided better flexural strength to concrete beams exter-
nally reinforced with various layers of flax fabrics in an epoxy composite (Yan et al. 2015).
Reproduced with permission from Elsevier

and hence improved properties. Supporting the beams with external composite lay-
ers led to improved resistance to fire and external environmental factors (Yan et al.
2015). An increase of 12.2% in ultimate laod, 77.9% in deflection, and 146% in
ductility was observed when coir fibers were used in steel fiber reinforced concrete
(Yan and Chouw 2013). Coir fibers were also able to provide bridging and reduce
the brittleness of flax fiber reinforced concrete beams (Yan and Chouw 2013).
Performance of coir and other types of natural fibers in concrete was studied
under different static and dynamic loads (Hamood Al-Masoodi et al. 2016). Coir
fibers (1, 3, and 5%) were combined with steel wires, fine and coarse aggregates,
and sand and cement to form 100 mm3 cubes. Extent of improvement in properties
was dependent on the type and amount of fibers (Table 4.5). Coir fiber reinforced
concrete was found to be suitable for roofing materials, wall panels/boards, etc.
Coir fibers have been found to improve mechanical properties of concrete, mor-
tar, etc., but most of these studies have been done under standard testing or atmo-
spheric conditions. However, concrete structures are extensively used in harsh
environments where they are exposed to chemicals, high temperatures, water, etc.
Few researchers have studied the ability of coir fibers to sustain or support concrete
structures under harsh conditions. Ramli et.al studied the performance of coir fiber
reinforced concrete under three types of aggressive environments which were a
tropical climate (A-series), alternative air and seawater environments (14 day cycle
with 4 days wet and 10 days drying) called the N-series, and continuous immersion
in seawater (W-series). Coir fibers used in this study had diameter of 0.32 mm,
length of 20–30 mm, tensile strength of 176 MPa, and elastic modulus of 22.4 GPa.
Compressive strength of the specimens differed based on the type of environment
exposed to, duration of exposure, and the amount of fibers in the concrete (Fig. 4.7).
84 4 Applications of Coir Fibers in Construction

Table 4.5 Changes in the properties of concrete with different levels of coir fibers (Hamood
Al-Masoodi et al. 2016)
Increase in Increase Increase Increase in Increase in Increase in
compressive in flexural in split dynamic dynamic dynamic
strength, strength, strength, stress, % strain, % toughness, %
Specimen % % % 2 MPa 3 MPa 2 MPa 3 MPa 2 MPa 3 MPa
1% coir −0.34 4.22 19.5 20.26 14.8 −4.26 19.07 72.6 13.4
3% coir 6.9 5.63 1.5 31.9 38.64 13.07 69.3 57.5 59.8
5% coir 2.23 10.73 34.7 72.19 76.6 1.74 126.3 94.5 124.4

90.00 90.00

85.00 85.00

80.00 80.00
90 A 180 A
75.00 75.00
90 N 180 N

70.00 90 W 70.00 180 W

65.00 65.00

60.00 60.00
CTRL 0.6 CF 1.2 CF 1.8 CF 2.4 CF CTRL 0.6 CF 1.2 CF 1.8 CF 2.4 CF

90.00 90.00

85.00 85.00

80.00 80.00
365 A 546 A
75.00 75.00
365 N 546 N

70.00 365 W 70.00 546 W

65.00 65.00

60.00 60.00
CTRL 0.6 CF 1.2 CF 1.8 CF 2.4 CF CTRL 0.6 CF 1.2 CF 1.8 CF 2.4 CF

Fig. 4.7 Changes in the compressive strength of concrete reinforced with coir fibers and exposed
to different environments and at various fiber contents (Ramli et al. 2013). Reproduced with per-
mission from Elsevier

In all conditions studied, the properties of the concrete increased up to fiber content
of 1.2–1.8% but decreased when 2.4% fibers were used. Addition of coir fibers also
increased resistance to penetration of chloride ions at low fiber concentrations.
Morphological studies (Fig. 4.8) suggested that there was poor interaction between
the fibers and concrete (existence of gaps) and considerable salt deposits on the
surface of cement and fibers were observed after the concrete containing 2.4% was
exposed to sea water for 546 days.
Coir fibers were used to reinforce blast-furnace slag cement and made into panels
and adopted for indoor and outdoor walls (John et al. 2005). Samples were exposed
to indoor and outdoor conditions for 12 years. Samples collected after 12 years
showed that no significant deterioration had occurred to the coir fibers. Although the
4.1 Introduction 85

Fig. 4.8 SEM image depicts poor interaction between the fibers and concrete (left). Substantial
salt deposits can be observed on the surface of the concrete and fibers after immersion in sea water
environment (Ramli et al. 2013). Reproduced with permission from Elsevier

cement was fully carbonated, no changes were observed in the morphology or the
lignin content (John et al. 2005). A relatively high coir fiber content of up to 15 wt%
and 20% fly ash were used to develop cementitious composites to avoid/decrease the
use of sand (Nadzri et al. 2012). Changes in the properties of the cement composite
with addition of fibers and different curing times are shown in Fig. 4.9. Cementitious
composites were developed using short coir fibers along with fly ash, sand, and
ground blast-furnace slag (GBFS) instead of cement. Raw coir fibers having density
of 1.28 g/cm3 were washed, boiled for 2 h and dried at 100 °C for 24 h. After treat-
ment, the fibers were cut into lengths of 17 mm and used as reinforcement with ratios
ranging from 1 to 4% (Hwang et al. 2016). Compressive strength decreased progres-
sively with increase in fiber content and also water to binder (W/B) ratio (Fig. 4.10).
However, flexural strength showed an opposite trend. Also, water absorption of the
composites containing 4% fibers was about 30% higher than that compared to com-
posites without any fibers. Toughness index and crack resistance also increased sub-
stantially with the addition of the fibers into the composites (Hwang et al. 2016).
Differences in the ability of coir fibers and coir fabric to change the depth of
reinforcement in a sandbed were determined (Sridhar and Prathap Kumar 2018).
Peak stress of the sand was similar for the fibers and mat (Fig. 4.11) at the different
depth of reinforcement to width of footing (u/B ratio) studied. However, peak strain
was considerably higher for the fibers compared to the sand. Reinforcing fibers
increased ductility and also settlement of sand which was suggested to be useful for
developing concrete structures (Sridhar and Prathap Kumar 2018).
Similar to coir fibers, coir pith has also been studied as potential reinforcement
in concrete (Brasileiro et al. 2013). Coir pith consisting of 30.7% lignin, 35.6% cel-
lulose, and 33.7% hemicellulose was made into particles of 2.4 mm in size using a
processing called quartering. Considerable changes were observed in the properties
of the composites with the addition of the coir pith particles. Compressive strength,
modulus of elasticity, and toughness were lower for the coir reinforced concrete
(Fig. 4.12).
86 4 Applications of Coir Fibers in Construction

Fig. 4.9 Changes in properties of cement reinforced with fibers after curing for three different
time periods (Nadzri et al. 2012). Reproduced with permission from Taylor and Francis

Addition of coir pith reduced the bulk density of the composites and increased
the porosity which provides better thermal and acoustic insulation. The composites
also have higher ductility and hence support loads for a longer time before rupture
or disintegration. Tensile and compressive properties of the coir reinforced concrete
composites were found to be suitable for lightweight construction materials
(Brasileiro et al. 2013). In another study, it was suggested that chemical pretreat-
ments were necessary for coir pith to improve adhesion and provide concrete with
better properties (Koňáková et al. 2015). Coir pith containing 2–3% short fibers,
4.1 Introduction 87

Fig. 4.10 Decrease in the compressive strength of the composites with increasing fiber content
and water/binder (W/B) ratio (Hwang et al. 2016). Reproduced with permission from Elsevier

Fig. 4.11 Changes in the peak stress for the coir fiber or mat containing sand at different u/B ratios
(Sridhar and Prathap Kumar 2018). Reproduced with permission from Taylor and Francis

30% lignin, and 27% cellulose was subject to NaOH or acetylation treatments.
Cementitious composites were prepared using 5 or 10% of the untreated (A), NaOH
modified (B), or acetylated (C) fibers. Properties of the composites varied depend-
ing on the amount and type of modification done to the fibers. Porosity of the fibers
increased by about 56% for the alkali modified and acetylated fibers. However,
bending strength decreased drastically due to the chemical modifications (Table 4.6)
(Koňáková et al. 2015). In addition to the mechanical properties, the thermal behav-
ior of the samples also showed considerable improvement. However, high coir con-
tent adversely affected properties and hence it was suggested that the amount of coir
fibers should be limited to 5% to obtain composites with properties suitable for
construction applications (Koňáková et al. 2015).
88 4 Applications of Coir Fibers in Construction

Fig. 4.12 Changes in the tensile properties of concrete with and without coir pith particles as the
fine aggregates (Brasileiro et al. 2013).CP is cement paste, SCM is sand cement mortar, CCC is
coir pith cement composite, and CSCC is coir pith sand cement composite. Reproduced with per-
mission from Elsevier

Table 4.6 Effect of treatments on the bulk and mechanical properties of cementitious composites
reinforced with various coir fibers (Koňáková et al. 2015)
Bulk Matrix Open Bending Water absorption
density, density, porosity, Compressive strength, coefficient
Specimen kg/m3 kg/m3 % strength, MPa MPa (kg m−2 s-1/2)
Control 2072 2526 18.0 63.1 9.8 0.018
A-5% 1857 2307 19.5 40.1 9.3 0.019
A-10% 1602 2214 27.6 15.2 4.3 0.193
B-5% 1846 2287 19.3 47.6 8.5 0.019
B-10% 1564 2178 29.0 12.8 4.1 0.199
C-5% 1874 2303 18.6 48.3 9.5 0.020
C-10% 1588 2208 28.1 14.5 4.5 0.241
Reproduced with permission through the Creative Commons Attribution License

Coir fibers were used as insulation on concrete roof tops to decrease outdoor
concrete and indoor room temperatures. At 12 noon, the temperature of the concrete
surface was 41.8 °C, whereas that covered with coir fibers was only 28.7 °C. However,
the extent of difference between covered and uncovered roofs decreased and was
only about 3 °C at 6 pm. Interestingly, the coir fibers covered roof released heat into
the cool night sky at a slower pace, suggesting the potential of the fibers to hold
heat. Using coir fibers as insulation could reduce energy consumption from 3 to 9%
and reduce environmental stress (Mintorogo et al. 2015).
In another application, coir fibers having length of 1, 2.5, and 4 cm were com-
bined in 2, 4, or 6% with mud and made into unfired soil lime bricks. Bricks devel-
oped were conditioned under ambient conditions for 14 days before testing
4.1 Introduction 89

(Purnomo et al. 2014). Addition of fibers increased bending strength and modulus
but decreased compressive strength. However, moisture absorption also increased
and it was necessary to obtain uniform distribution of the fibers in the soil (Purnomo
et al. 2014). Bricks were also made using ordinary Portland cement and reinforced
with untreated or treated coir fibers. Fibers were ground into particles and treated
with chemicals including gelatin-hexamine solution, linseed oil or sodium metasili-
cate and aluminum sulfate. Chemical treatments reduced the water retention values
of the fibers from 80 to 42% (Amaylia and Lee 2015). Coir reinforced bricks had
better compressive strength, whereas linseed oil provided better adhesion and
mechanical properties. Although the coir bricks developed had properties similar to
that of commercially available bricks, it was suggested that further studies would be
required to determine the durability and performance under specific environmental
conditions (Abdullah and Lee 2017). Coir fibers made into particles of different size
were combined with rubber and used as reinforcement for red sand bricks. Coir
fibers were also treated with sodium hydroxide solution for 4 h at 50 °C. The coir
fibers or particles were thoroughly mixed with the red sand in predetermined propo-
sitions and compacted with a laboratory compacting machine at a pressure of 20 kN
for 5 min (Oluwole et al. 2015). A picture (Fig. 4.13) of the red sand bricks rein-
forced with coir is shown in Fig. 4.13. Use of natural rubber as binder provided
better flexural properties compared to using water as binder. Similarly, addition of

Fig. 4.13 Digital picture of the red sand brick reinforced with coir fibers (Oluwole et al. 2015).
Reproduced with permission from University POLITEHNICA Timisoara
90 4 Applications of Coir Fibers in Construction

Fig. 4.14 Picture of the


coir and polyester resin
reinforced red mud
composites subject to
buckling test (Rachchh
et al. 2015). Reproduced
with permission from Iran
Polymer and Petrochemical
Institute

coir fibers/particles improved the flexural properties and water repellency of the
bricks. Coir fibers along with polyester resin were used to reinforce red mud which
is obtained as a byproduct during the processing of alumina. Coir fibers were cut
into lengths of 0.1 to 0.3 mm and had density between 1.12 and 1.15 g/cm3. Red
mud was of 70–90 μm diameter and density of 1.36–1.6 g/cm3 compared to 1.35 ­g/
cm3 and elastic modulus of 3.23 GPa (Rachchh et al. 2015). Methyl ethyl ketone
peroxide was the hardener and cobalt naphthalene was used as the accelerator. The
amount of red mud was controlled at 10% and fiber content was varied between 10
and 25%. Materials were combined together in the desired ratio and placed in a
mold and cured for 24 h. Samples obtained were subject to buckling (Fig. 4.14),
flexural, and other tests. Compressive strength decreased with increasing fiber con-
tent and varied between 64 and 31 MPa and flexural modulus between 1315 and
1850 MPa. Overall, 20% fiber content provided the most optimum properties
(Rachchh et al. 2015).
In a unique application, three types of coconut husks (Fig. 4.15) were studied for
their potential to remove pollutants in the air by adopting the fibers in walls. To
increase absorption, activated carbon was also added into the coir-based green wall
biofilters with a specific surface area of 0.75 m2/g and water holding capacity of
5.5 g/g of dry material (Pettit et al. 2018). Common pollutants in air and volatile
organic components were passed through the coconut husk based wall, and single pass
References 91

Fig. 4.15 Three types of coconut husks used as biofilters. Left is husks of about 0.5 mm diameter,
middle is 5–15 mm, and right is 8–35 mm (Pettit et al. 2018)

Fig. 4.16 Absorption of different particulate matter by the biofilters made using coconut husks
and coconut husk-carbon substrate (Pettit et al. 2018)

removal efficiency (SPRE) was determined. Considerably high removal of particulate


matter was observed in the biofilters. The filters containing only the fibers and without
the carbon showed higher absorption of all particle sizes except 0.3–0.5 (Fig. 4.16).

References

Abdullah AC, Lee CC (2017) Effect of treatments on properties of cement-fiber bricks utilizing
rice husk, corncob and coconut coir. Procedia Eng 180:1266–1273
Abdullah A, Jamaludin SB, Anwar MI, Noor MM, Hussin K (2011) Assessment of physical and
mechanical properties of cement panel influenced by treated and untreated coconut fiber addi-
tion. Phys Procedia 22:263–269
92 4 Applications of Coir Fibers in Construction

Al-Masoodi AHH, Kawan A, Kasmuri M, Hamid R, Khan MNN (2016) Static and dynamic prop-
erties of concrete with different types and shapes of fibrous reinforcement. Constr Build Mater
104:247–262
Amaylia CA, Lee CC (2015) Effect of treatment on surface modifier and water retention value
(WRV) of natural fiber. Aust J Basic Appl Sci 9(25):101–104
Andiç-Çakir Ö, Sarikanat M, Tüfekçi HB, Demirci C, Erdoğan ÜH (2014) Physical and mechanical
properties of randomly oriented coir fiber–cementitious composites. Compos Part B 61:49–54
Brasileiro GAM, Vieira JAR, Barreto LS (2013) Use of coir pith particles in composites with
Portland cement. J Environ Manag 131:228–238
Ferraz JM, Del Menezzi CHS, Souza MR, Okino EYA, Martins SA (2012) Compatibility of pre-
treated coir fibres (Cocos nucifera L.) with Portland cement to produce mineral composites. Int
J Poly Sci 2012:1–7. https://doi.org/10.1155/2012/290571
Gnanapragasam A, Jesuraj P (2016) Mechanical properties of high strength concrete with artificial
and natural fibers. Adv Nat Appl Sci 10(1):48–56
Hwang C-L, Tran V-A, Hong J-W, Hsieh Y-C (2016) Effects of short coconut fiber on the mechani-
cal properties, plastic cracking behavior, and impact resistance of cementitious composites.
Constr Build Mater 127:984–992
Islam SM, Hussain RR, Morshed MAZ (2012) Fiber-reinforced concrete incorporating locally
available natural fibers in normal-and high-strength concrete and a performance analysis with
steel fiber-reinforced composite concrete. J Compos Mater 46(1):111–122
John VM, Cincotto MA, Sjöström C, Agopyan V, Oliveira CTA (2005) Durability of slag mortar
reinforced with coconut fibre. Cem Concr Compos 27(5):565–574
Kiruthika AV, Veluraja K (2017) Physical properties of plant fibers (sisal, coir) and its composite
material with tamarind seed gum as low-cost housing material. J Nat Fibers 14(6):801–813
Koňáková D, Vejmelková E, Čáchová M, Siddique JA, Polozhiy K, Reiterman P, Keppert M, Černý
R (2015) Treated coconut coir pith as component of cementitious materials. Adv Mater Sci Eng
2015:1. https://doi.org/10.1155/2015/264746
Li Z, Wang L, Wang X (2012) The long-term performance of cementitious composites reinforced
with coir fibre. J Text Inst 103(8):912–920
Mintorogo DS, Canadarma WW, Juniwati A (2015) Application of coconut fibres as outer eco-­
insulation to control solar heat radiation on horizontal concrete slab rooftop. PhD dissertation,
Petra Christian University
Nadzri NIM, Jamaludin SB, Noor MM (2012) Development and properties of coconut fiber rein-
forced composite cement with the addition of fly ash. J Sustain Cem Based Mater 1(4):186–191
Neeraja D, Wani AI, Kamili Z, Agarwal K (2017) Study on strength characteristics of concrete
using M-Sand and coconut fibers. IOP Conf Ser Mater Sci Eng 263(3):11–17
Oluwole OI, Avwerosuoghene OM, Oluwatobi AJ (2015) The effect of natural rubber on the flex-
ural properties of coconut coir (Cocos Nucifera) reinforced red sand composites. Acta Tech
Corviniensis Bull Eng 8(2):87–94
Pettit T, Irga PJ, Torpy FR (2018) Functional green wall development for increasing air pollutant
phytoremediation: substrate development with coconut coir and activated carbon. J Hazard
Mater 360:594–603
Purnomo H, Priadi D, Ausias G, Lecompte T, Lumingkewas HR, Perrot A (2014) Effect of coconut
fibers addition to early age unfired soil lime bricks strength. Key Eng Mater 594:471–476
Rachchh NV, Misra RK, Roychowdhary DG (2015) Effect of red mud filler on mechanical and buck-
ling characteristics of coir fibre-reinforced polymer composite. Iran Polym J 24(3):253–265
Ramli M, Kwan WH, Abas NF (2013) Strength and durability of coconut-fiber-reinforced concrete
in aggressive environments. Constr Build Mater 38:554–566
Sarangi S, Sinha A (2016) Mechanical properties of hybrid fiber reinforced concrete. Indian J Sci
Technol 9(30):1–4
Sathiparan N, Rupasinghe MN, HM Pavithra B (2017) Performance of coconut coir reinforced
hydraulic cement mortar for surface plastering application. Constr Build Mater 142:23–30
Sridhar R, Prathap Kumar MT (2018) Experimental investigation of load settlement behavior of
coir mat and coir fiber reinforced sand. J Nat Fibers 15(3):452–463
References 93

Yan L, Chouw N (2013) A comparative study of steel reinforced concrete and flax fibre rein-
forced polymer tube confined coconut fibre reinforced concrete beams. J Reinf Plast Compos
32(16):1155–1164
Yan L, Su S, Chouw N (2015) Microstructure, flexural properties and durability of coir fibre rein-
forced concrete beams strengthened with flax FRP composites. Compos Part B 80:343–354
Yan L, Chouw N, Huang L, Kasal B (2016) Effect of alkali treatment on microstructure and
mechanical properties of coir fibres, coir fibre reinforced-polymer composites and reinforced-­
cementitious composites. Constr Build Mater 112:168–182
Chapter 5
Energy Applications of Coir

5.1 Coir as Fuel

Abundant availability and low cost makes coir a highly valuable source for fuel.
However, the bulkiness of coir makes it inconvenient to use. Both ripe (brown) and
unripe (white) coir fibers were studied for pelletizing properties and potential use as
fuel briquettes. A high heating value of 18–19 MJ/kg was obtained without much
difference between the two types of fibers. However, brown coir pellets had higher
compressive strength (1200–1400 N) compared to white coir fibers (900–1500 N). It
was suggested that brown coir could be made into pellets for fuel applications (Stelte
et al. 2018). Although, one of the major by-products, coconut leaves are rarely stud-
ied for any large-scale applications. In a unique study, conversion of coconut leaves
into fuel by a process called torrefaction (Fig. 5.1) to improve the calorific value was
done (Pestaño and Jose 2016). Calorific value of the coconut leaves increased from
21 to 27 MJ/g as the torrefaction temperature and holding time were varied.
Correspondingly, weight of the samples also decreased from 39 to 24%.
A comparative study was done to understand the benefits of using coir to produce
solid biochars in comparison to conventional wood (Liu and Han 2015). The bio-
chars were prepared by heating the materials in a furnace at 200–330 °C at a heating
rate of 15 °C per minute. Figure 5.2 shows the digital images of the biochars
obtained at various temperatures. Some of the properties of the biochars are given
in Table 5.1 and the combustion behavior of coir biochar is compared with pine-
wood biochar in Table 5.2. It was shown that low temperature pyrolysis, particularly
at 300 °C provided coir biochar with the most optimum combustion values (Liu and
Han 2015).
In addition to biochars, coir fibers were made into hydrochars using hydrother-
mal conditions and studied for combustion range and efficiency with and without
lignite. For preparing the hydrochar, the coir fibers were treated in an autoclave at
250 °C for 20 min in the presence of water (Liu et al. 2012). Hydrochar formed was
dried at 105 °C and later milled to about 125 μm for combustion studies with and

© Springer Nature Switzerland AG 2019 95


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_5
96 5 Energy Applications of Coir

Fig. 5.1 Process of increasing the calorific value of coconut leaves using a process called torrefac-
tion (Pestaño and Jose 2016). Reproduced with permission through Creative Commons Attribution
4.0 International License

Fig. 5.2 Images of the pine wood and coir biochars prepared at different temperatures (Liu and
Han 2015). Reproduced with permission from Elsevier
5.1 Coir as Fuel

Table 5.1 Characteristics of the coir biochar compared to pine wood biochar (Liu and Han 2015)
Proximate analysis, % Ultimate analysis, % Energy analysis
Sample Volatile matter Fixed carbon Ash N C H S O Yield HHV (MJ/kg) Energy yield, %
CF-200 78.26 15.04 6.70 1.35 50.42 5.20 0.45 42.6 83.1 19.93 87.5
CF-225 76.98 15.90 7.12 1.02 51.83 4.99 0.30 41.9 80.1 20.36 81.9
CF-250 75.58 16.37 8.05 1.13 53.77 4.76 0.28 40.1 69.2 21.03 71.5
CF-275 70.20 21.04 8.76 1.12 58.08 4.55 0.28 35.9 63.9 22.63 68.7
CF-300 64.21 25.71 10.08 1.21 61.38 4.08 0.32 33.0 55.7 23.80 58.5
CF-330 60.4 28.4 10.98 1.18 65.36 4.12 0.29 29.05 50.2 25.36 53.5
PW-200 80.43 18.62 0.95 1.33 51.16 6.54 0.18 40.79 91.0 20.74 99.8
PW-225 78.25 20.77 0.98 1.34 51.74 6.50 0.17 40.25 89.2 20.00 90.3
PW-250 75.84 23.09 1.07 1.34 52.00 6.35 0.17 40.14 79.2 21.06 79.4
PW-275 70.33 28.52 1.16 1.65 53.37 6.20 0.18 38.60 72.0 21.66 74.1
PW-300 58.34 40.37 1.29 1.68 55.52 6.07 0.22 36.51 65.2 22.60 68.0
PW-330 50.71 47.90 1.39 1.60 59.86 5.81 0.20 32.53 61.4 24.49 63.3
Lignite 48.76 40.98 10.26 1.74 61.64 5.72 0.77 30.13 – 25.31 –
Reproduced with permission from Elsevier
97
98 5 Energy Applications of Coir

Table 5.2 Comparison of the combustion parameters of coir biochar and pinewood biochar (Liu
and Han 2015)
Temp, °C Mass loss, % Peak temp, °C
DTGmax (%C)
Stage
Sample Stage I Stage II Stage I Stage II Stage I Stage II Stage I II Residue, %
CF-200 264–412 412–500 53.31 25.28 309 454 0.54 0.42 8.64
CF-250 270–415 415–505 52.51 34.41 312 455 0.50 0.44 10.26
CF-300 282–416 416–500 44.44 39.81 370 459 0.37 0.56 12.29
CF-330 281–405 405–490 42.86 39.47 367 453 0.40 0.60 13.99
PW-200 301–385 385–480 51.47 22.00 331 464 1.71 0.32 2.56
PW-250 314–391 391–484 58.76 26.34 336 473 1.70 0.44 3.12
PW-300 328–402 402–486 47.82 35.96 346 480 1.58 1.38 3.80
PW-330 326–381 381–472 29.11 56.66 337 458 0.69 2.66 3.95
Reproduced with permission from Elsevier

Table 5.3 Composition and calorific value of the coir fibers compared to eucalyptus leaves and
lignite (Liu et al. 2012)
Coir Coir biochar Eucalyptus Eucalyptus biochar Lignite
Proximate analysis
Volatile matter, % 8.85 67.92 79.21 70.97 48.76
Fixed carbon, % 11.10 27.11 10.31 22.17 40.98
Ash, % 8.05 4.97 10.48 6.86 10.26
Moisture content, % 12.94 2.51 6.59 3.20 13.46
Heating value, MJ/kg 18.41 26.73 18.93 24.93 24.98
Ultimate analysis, %
 C 47.75 67.10 46.96 62.30 61.64
 H 5.61 5.20 6.22 5.47 5.72
 N 0.90 0.98 1.23 1.44 1.74
 S 0.23 0.29 0.77 0.44 0.77
 O 45.51 26.43 44.82 30.35 30.13
Reproduced with permission from Elsevier

without lignite. Composition and properties of the biomass and hydrochar obtained
from coir and eucalyptus leaves are shown in Table 5.3. Combustion of the hydro-
char was influenced by the amount of lignite. Raw coir fibers had the lowest thermal
stability, which increased substantially after formation of the hydrochar (from 247
to 358 °C). Addition of lignite did not increase the combustion parameters of the
coir hydrochar but increased for the eucalyptus biochar (Liu et al. 2012).
Interestingly, the maximum weight loss rate was 8.4 for the two hydrochars, much
higher than that for the unmodified biomass and lignite. Further, it has also been
shown that the biochars produced from coir can be pyrolyzed to generate hydrogen
cyanide (HCN) or ammonia (NH3) gas. Pyrolysis was done in a horizontal fixed bed
tubular quartz reactor at temperatures between 600 and 900 °C for 20 min. The
5.2 Carbonization of Coconut Fiber and Shells for Supercapacitor Applications 99

Fig. 5.3 Comparison of the release of hydrogen cyanide and ammonia from the raw biomass and
biochar (Liu and Balasubramanian 2014). Reproduced with permission from Elsevier

amount of gases generated during the pyrolysis was collected by absorption in


20 mmol/L NaOH and 2 mmol/L nitric acid solutions (Liu and Balasubramanian
2014). The amount of nitrogen distribution for the unmodified and modified b­ iochars
is shown in Fig. 5.3. Lower temperatures provide higher gas release for all the sam-
ples. Hydrogen cyanide was more prominently released from the raw biomass,
whereas higher amount of ammonia was released from the pyrolyzed biochars. It
was suggested that combining biochar and lignite reduces environmental pollution
compared to using the raw biomass for combustion.

5.2  arbonization of Coconut Fiber and Shells


C
for Supercapacitor Applications

Coir and other coconut byproudcts have been extensively studied for energy
applications. One of the most common approaches in using coir is to convert it
into biocarbon and study its potential for supercapacitors and other areas. In one
such attempt, coir pith was precarbonized at 350 °C and later activated with KOH
at three different temperatures (800, 850, and 900 °C). Activated coir pith carbon
(CPC) was made into electrodes and the electrochemical characteristics were
measured using the CPC as anode against sodium metal as the cathode
(Mullaivananathan et al. 2017a, b). Some of the performance parameters of the
carbon developed are in Table 5.4. CPC-850 was able to withstand current of 1,
100 5 Energy Applications of Coir

Table 5.4 Performance of the CPC anode when subject to extended cycles at 50 mA/g current
density in SIBs (Mullaivananathan et al. 2017a, b)
CPC-800 CPC-850 CPC-900
Charge Discharge Charge Discharge Charge Discharge
capacity capacity capacity capacity capacity capacity
Cycles mA h g−1 mA h g−1 mA h g−1 mA h g−1 mA h g−1 mA h g−1
1 243 797 291 826 228 769
2 247 279 286 288 220 251
10 229 236 280 280 244 231
20 214 221 282 283 203 209
30 209 211 267 263 222 211
40 207 215 266 272 201 201
50 200 207 260 261 212 204
Reproduced through Creative Commons Attribution-Noncommercial 3.0 Unported License

2, 5, and 10 A/g and delivered 168, 152, 150, and 128 F/g up to 10,000 cycles.
Similarly, specific capacitance values of 168, 152, and 150 F/g at 1, 2, and 5 A/g
current density were observed. It was suggested that CPC-850 was a suitable
electrode for use in energy storage devices including LIBs and Li-S systems
(Mullaivananathan et al. 2017a, b).
Instead of converting coir using chemical or physical approaches, a method was
developed to use raw coir fibers as electrode materials. Coir fibers from a shell were
collected and coated with a layer of gold to form the electrode (Mondal et al. 2013).
A bundle of fibers having diameter in the range of 130–200 μm were sputter coated
with gold for about an hour until the fibers turned brown. Morphological features and
chemical composition of the fibers are shown in Fig. 5.4. Ends of the fibers were
coated with conducting silver paste and connected to a copper wire to form the elec-
trodes. Cyclic voltammetry studies were done in a ferrocene carboxylic acid aqueous
medium and in acetonitrile medium (Mondal et al. 2013) at scan rates ranging from
10 to 150 mV/s. Gold coating increased the strength of the fibers from 250 to 350 MPa
and the elongation from 10 to 15%. An electrical resistivity of 4.4 × 10−4 Ω cm was
observed for the fibers. Coir fibers had similar performance as conventional elec-
trodes in both aqueous and non-aqueous media (Mondal et al. 2013).
Three parts of the coconut tree, namely the coconut fiber (CF), coconut leaves
(CL), and coconut stick (CS), have been studied for their potential to be used as
electrodes for potential application in supercapacitors (Divyashree et al. 2016). The
three parts were ground to a particle size of 65 μm and then pyrolyzed at 650 °C for
120 min at a heating rate of 10 °C under nitrogen atmosphere. These nanoparticles
were made into a slurry by combining with polyvinylidene difluoride (PVDF) in
1-methy 2-pyrrolidione in 80/15 ratio and 5 wt% solution. The slurry was then
applied onto a carbon paper of area 1 × 1 cm2 and used as the working electrode in
a three electrode system (Divyashree et al. 2016). Particles obtained from the fiber
had average size of 20 nm, coir leaves was 60 nm, and those from coir shells was
about 40 nm. Surface area of the samples was 7, 4, and 8 m2/g for the fiber, leaf, and
shell carbon, respectively. Similarly, the pore diameters were 5.59 nm, 13.2 nm, and
5.7 nm, respectively.
5.2 Carbonization of Coconut Fiber and Shells for Supercapacitor Applications 101

Fig. 5.4 Optical image of the fibers (a), SEM of uncoated fiber (b), image of the gold-coated coir
fiber electrode (c), polyaniline deposited on the coir fibers to form electrode (d), and EDS spectra
of the coir electrode (e) (Mondal et al. 2013). Reproduced with permission from American
Chemical Society

In another study, coir pith derived carbon was studied for potential use as anode
material (Mullaivananathan et al. 2017a, b). The carbon contained randomly ori-
ented layers of graphene sheet like structures. Specific area of the carbon after acti-
vation with KOH reached 2500 m2/g and had a steady state capacity of 837 mAh/g
after the 50th cycle at 100 mA/g. It was suggested that the CPC could be used as an
anode up to 5.4 °C rate with good capacitance and retention capabilities
(Mullaivananathan et al. 2017a, b).
Carbonization and activation with KOH is commonly used to improve the per-
formance of materials intended for supercapacitor applications (Yin et al. 2016).
Ground coir fibers were carbonized by heating to 700 °C and maintaining this tem-
perature for 1 h. Later, the carbonized material was combined with KOH in different
102 5 Energy Applications of Coir

Fig. 5.5 Schematic representation of the process used to prepare the activated coir fiber carbon
(Yin et al. 2016). Reproduced with permission from Elsevier

Table 5.5 Changes in the porosity and pore volume of coir fiber carbon before and after various
levels of activation with KOH (Yin et al. 2016)
Sample SBET (m2/g) Smic (m2/g) Smeso (m2/g) Vt (cm3/g) Vmic (cm3/g) Vmeso (cm3/g)
Untreated 33 7 26 0.08 0.02 0.06
KOH-1 1734 1608 126 0.81 0.68 0.13
KOH-2 1966 1825 141 0.89 0.76 0.13
KOH-3 2495 2038 457 1.32 1.05 0.27
KOH-4 2898 1937 961 1.59 1.11 0.48
KOH-5 2589 1590 999 1.49 1.08 0.41
Reproduced with permission from Elsevier

ratios and heated again at a rate of 3 °C/min up to 850 °C for 1 h. A schematic


illustration of the process is shown in Fig. 5.5. A symmetrical two-electrode cell
was developed to study the potential of the activated coir fiber carbon as superca-
pacitor (Yin et al. 2016). Untreated coir fiber has considerably low surface area and
pore volumes. Activating with KOH increases the properties of the carbon up to
several levels of magnitude as seen from Table 5.5. Highest specific surface area of
2898 m2/g with a pore volume of 1.59 cm3/g was obtained. More importantly, high
specific capacitance of 266 F/g at 0.1 A/g with KOH as electrode and of 155 F/g
with EMIMBE4 as electrolyte were obtained. A combined power of 53 Wh/kg and
high energy density of 8224 W/kg could be drawn using the modified coir fiber
carbon (Yin et al. 2016). Some of the images of the precursor and activated carbon
are shown in Fig. 5.6.
5.3 Gasification of Coir 103

Fig. 5.6 Scanning electron images of the coir fibers (a); carbonized coir fibers (b) and activated
porous carbon (c) and its corresponding TEM image (d) (Yin et al. 2016). Reproduced with per-
mission from Elsevier

5.3 Gasification of Coir

Generation of producer gas using various biomasses has been considered as a means
to produce clean energy (Ramadhas et al. 2006, 2008). Coir pith is a high calorific
biomass and hence used to generate producer gas. Some of the properties of coir
pith including calorific value are compared to other common sources of biomass
(Table 5.6). Coir pith was used as a source to generate producer gas and run a dual
fuel engine. It was suggested that coir pith can be used directly in the gasifier as a
fuel source for engines either with wood chips or rubber seed oil (Ramadhas et al.
2006, 2008). In another attempt, coir pith was used as a source for producer gas
(Sheeba et al. 2013). Various ratios of oxygen/carbon and heating values were stud-
ied for their effect on overall thermal efficiency. A hydrogen yield of 43% was
obtained at a steam/biomass ratio of 4.5 and a temperature of 1008 °C. A gas yield
of 1.55 Nm3/kg with a steam to biomass ratio of 4.5 and a temperature of 805 °C
was obtained. A tar yield of 17 g/Nm3 was achieved with an S/B ratio of 4.5 at a
temperature of 641–647 °C.
Coir dust was subject to carbonization under hydrothermal conditions, and the
influence of the treatment conditions on formation of curved graphene structures
was investigated (Barin et al. 2014). High resolution TEM images (Fig. 5.7) showed
104 5 Energy Applications of Coir

Table 5.6 Composition and calorific values of coir pith and other biomasses used for production
of biochar (Ramadhas et al. 2006, 2008)
Bulk density, Ash content, C, H, N, O, Calorific value,
Biomass kg/m3 % % % % % kJ/kg
Bagasse 74 4.0 47 6.5 0.0 42.5 17.6
Coir pith 47 13.6 41 4.0 1.5 39.6 16.8
Groundnut 165 3.1 34 2.0 1.1 59.9 18.9
shell
Saw dust 177 1.2 52 5.2 0.5 40.9 18.5
Straw 80 15.5 36 5.3 0.2 43.1 15.5
Wood 330 1.5 52 5.2 0.5 42.0 18.5
Reproduced with permission from Elsevier

Fig. 5.7 (a) and (c) HRTEM images of graphite-like structures prepared from coconut coir dust
through hydrothermal carbonization followed by pyrolysis. (b) and (d) FFT image of selected area
from HRTEM image showed in (a) and (c), respectively. Transmission electron microscopy (TEM)
images of the graphite like structure formed from the coir dust (Barin et al. 2014). Reproduced
with permission from Elsevier
5.3 Gasification of Coir 105

that direct pyrolysis of the coconut results in the formation of overlapped sheet
structure, whereas an additional hydrothermal treatment resulted in well-defined
lattice fringes, suggesting a high degree of crystallinity. In addition, the lattices
demonstrated distinct curves, indicating that the carbon can be curved or rolled up.
About 70–80% of the coir dust could be obtained as the graphite structure. However,
it was found that hydrothermal treatment affected formation of the graphite struc-
ture and should be avoided (Barin et al. 2014).
Ability of coconut shells to be used as carbon for electrodes has also been stud-
ied. Coconut shells were heated up to 1000 °C at a heating rate of 10 °C per min and
converted into carbon. This carbon was later activated using a microwave tubular
furnace at a frequency of 2.4 GHz. Activation was done at a temperature of 900 °C
using three different means (steam, CO2, and a mixture of the two (Yang et al. 2010).
Considerable variations (Table 5.7) were observed in the properties of the carbon
obtained depending on the treatment conditions used. Surface area of the carbon
varied anywhere from 1011 to 2288 m2/g. Steam activation provided higher surface
area than the other two treatments. The carbon mostly contained mesopores and less
than 20% micropores. Microwave heating substantially decreased the carbonization
and activation time without affecting the yield or the chemical composition and
hence suggested to be useful to develop electrodes from coconut shell and other
biomass (Yang et al. 2010).
In a similar study, effects of different heating methods on the properties of coco-
nut shell carbon were studied (Duan et al. 2012). To carbonize the shells, they were
heated in a horizontal tube furnace up to 600 °C at a heating rate of 10°/min and

Table 5.7 Some of the parameters obtained during the different processing conditions used to
obtain carbon from the coconut shells (Yang et al. 2010)
Sample SBET (m2/g) Vtot (cm3/g) Vmicro (cm3/g) V meso (cm3/g)
MS 900-15 1011 0.585 0.5179 0.0671
MS 900-30 1363 0.8454 0.6727 0.1727
MS 900-45 1677 1.0250 0.8339 0.1911
MS 900-60 1888 1.1570 0.9483 0.2087
MS 900-75 2079 1.2120 0.9735 0.2385
MC 900-60 1162 0.7159 0.5703 0.1496
MC 900-90 1425 0.8820 0.7022 0.1798
MC-900-120 1703 1.0320 0.8153 0.2167
MC-900-150 1905 1.204 0.9365 0.2675
MC-900-180 2080 1.2700 0.9974 0.2726
MC-900-210 2288 1.2990 1.0120 0.2870
MCS 900-15 1139 0.6911 0.5869 0.1042
MCS 900-30 1424 0.8276 0.7229 0.1047
MCS 900-45 1761 1.1020 0.8773 0.2247
MCS 900-60 2020 1.2480 1.0080 0.2400
MCS 900-75 2194 1.2930 1.0100 0.2830
Reproduced with permission from Elsevier
106 5 Energy Applications of Coir

Fig. 5.8 Comparison of the BET surface area and pore volumes of carbons obtained using differ-
ent methods of heating and activation (Duan et al. 2012). Reproduced with permission from
Springer Nature

held at that temperature for 2 h under nitrogen atmosphere. Further, steam and CO2
activation was done both in the conventional and microwave oven at a temperature
of 900 °C. For the steam heating, a flow rate of 1.35 g/min was used and for the CO2
activation a flow rate of 600 cm3/min was used. Yield of carbon varied from 20 to
70% depending on the type and duration of heating. Similarly, BET surface area and
pore volume also showed significant changes (Fig. 5.8). Overall, microwave heating
decreased the activation temperature and time substantially and provided higher
yield. Similarly, CO2 activation provided higher pore volume and surface area than
steam activation. Activated carbon obtained using the microwave heating and steam
activation had a surface area of 1684 m2/g compared to 1716 m2/g for the CO2 acti-
vated carbon (Duan et al. 2012).
Porous graphene like nanosheets were developed from coconut shells using a
simple and unique approach (Sun et al. 2013). In this approach, a one-step activa-
tion and graphitization by adding the graphitic catalyst precursor (FeCl3) and acti-
vating agent (ZnCl2) along with the coconut shell for carbonization was done. The
components were graphitized and activated by heating in a tubular furnace under N2
atmosphere at a heating rate of 5 °C/min for up to 900 °C for 1 h. Prepared graphitic
carbon had an exceptionally high surface area of 1874 m2/g and large pore volume
of 1.21 cm3/g (Table 5.8). In terms of electrical parameters, the carbon had high
specific capacitance of 168 F/g and columbic efficiency of 99.5% even after
5000 cycles in KOH. Even higher performance of 196 F/g at 1 A/g was obtained in
organic electrolytes. An energy density of 54.7 W h/kg was obtained at a high power
density of 10 kW/kg (Sun et al. 2013).
Several modifications and additions have been done to improve the specific
capacitance of materials developed for supercapacitor applications. Metal oxides
have been found to improve the intra- and inter-particle electrical conduction
(Dandekar et al. 2005). Among the various metal oxides, ruthenium oxide has been
found to be very effective to improve conduction since it can accept and donate
protons from electrolytes. It was also found that the oxide should be in an amor-
phous state to increase the specific capacitance. Cyclic voltammetric studies showed
5.3 Gasification of Coir 107

Table 5.8 Properties of the porous graphitic carbon prepared from coconut shells (Sun et al. 2013)
Samples SBET (m2/g) Pore volume (cm3/g) Pore size, nm
PGNS-3-700 1281 0.35 2.3
PGNS-3-800 1419 0.89 2.6
PGNS-3-900 1874 1.21 3.0
PGNS-3-1000 1568 0.93 3.2
PGNS-1-900 896 0.26 2.2
PGNS-5-900 1668 0.97 3.6
GC-900 138 0.12 6.6
AC-900 2007 0.94 1.7
Reproduced with permission from Royal Society of Chemistry

a b
0.006 0.014 st
1 Cycle
10 mV/s
0.012 th
10 Cycle
5 mV/s
0.010 th
20 Cycle
0.004 1 mV/s
0.008
th
30 Cycle
th
40 Cycle
0.006 th
50 Cycle
0.002 0.004
0.002
-2

I/A cm-2

0.000
I/A cm

0.000 -0.002
-0.004
-0.002 -0.006
-0.008
-0.010
-0.004 -0.012
-0.014
-0.016
-0.006
-0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0

E/V E/V

Fig. 5.9 Cyclic voltammograms of composite sample at various scan rates (a) and capacitance
cycles (b) (Dandekar et al. 2005). Reproduced with permission from Elsevier

that the specific capacitance increased from 100 to 250 F/g when 9% ruthenium
(Fig. 5.9), (Table 5.9) was used to treat the material compared to without using any.
However, BET surface area, meso- and micropore area and size decreased with
increase in percentage of ruthenium (Dandekar et al. 2005).
Similar to the coir fibers, coconut shells were also carbonized and activated using
KOH in a microwave-assisted process. In a typical process, coconut shell powder
was impregnated with 50% KOH and heated in quartz reactor placed inside a micro-
wave at frequency of 2.5 GHz using power from 100 to 1000 W for 20 min under
nitrogen atmosphere. After treatment, the powder obtained was washed with 50%
hydrochloric acid and water to remove KOH and stored for further analysis (Nayai
et al. 2016). Considerable variations were observed in the surface area, pore size,
and specific capacitance with increase in microwave power and temperature of car-
bonization. Highest capacitance obtained was 126 F/g when 600 W and 635 °C
were used. Corresponding surface area and pore size were 1769 m2/g and 2.7 nm,
respectively. Also, the yield of the carbon was also relatively high at 57% when
108 5 Energy Applications of Coir

Table 5.9 Changes in the properties of the activated carbon when different percentage of
ruthenium was used (Dandekar et al. 2005)
Average
Surface Micropore Total pore pore
Ruthenium, area, Micropore Mesopore volume, volume, diameter, Capacitance,
% m2/g area, m2/g area, m2/g cm3/g cm3/g A° F/g
0 1340 1266 77.54 0.52 0.52 18 100
1.6 878 786.1 91.8 0.32 0.42 19 125
5 498 455.9 42.3 0.18 0.27 21 176
7 446 428.3 18.03 0.16 0.19 17 205
9 460 435.9 28.8 0.17 0.20 17 250
Reproduced with permission from Elsevier

compared to that obtained using other approaches. A relatively high supercapaci-


tance of 186 F/g at an energy density of 11 Wh/kg was obtained by preparing meso-
porous carbon from coconut shells using a two-step process (Barzegar et al. 2016).
The carbon obtained was considerably porous with a high surface area of 1416 m2/g.
After activation with KOH, the supercapacitor was found to have a specific capaci-
tance of 186 F/g with energy and power densities of 11 Wh/kg and 325 W/kg and
current density of 0.5 A/g. The device did not show any drop in performance even
after floating for 100 h or cycling for 10,000 cycles (Barzegar et al. 2016).
In another study, possibility of developing activated carbon from coir pith was
investigated using phosphoric acid as the activating agent (Thitame and Shukla
2016). For the activation, the coir pith was treated with boiling phosphoric acid for
1 h and carbonized using temperatures between 400 and 800 °C. Activated carbon
obtained had particle size of about 50 μ and average pore size was between 3 and
5 nm and micropore volume was 4.6%. The activated carbon had high sorption for
dyes and was also suggested to be suitable for supercapacitor applications (Thitame
and Shukla 2016).
An unprecedentedly high surface area of 3000 m2/g was obtained for carbon
made from coconut shells using 25% alcohol solution of melamine as precursor
(Jurewicz and Babeł 2010). Pretreated shells were pyrolyzed at 500 °C for 1 h and
later activated using KOH at 1:4 ratio and temperature of 725–900 °C. A second
activation was performed using KOH and heating up to 350 °C. A total pore volume
of 2 cm3/g and highest specific capacitance of 368 F/g for the negative electrode in
alkaline solution and 293 and 307 F/g for the positive and negative electrode in the
acidic solution were obtained. In addition to the surface area and suitable pore struc-
tures, it was suggested that favorable chemical composition in terms of nitrogen and
oxygen functionality was responsible for the exceptional performance of the carbon
activated using KOH at 900 °C (Jurewicz and Babeł 2010). Activated carbon derived
from coconut shells having a mesoporosity of 60% was used to achieve an extraor-
dinary energy density of 69 Wh/kg, maximum specific capacitance of 159 F/g, and
very high cycleability of 2000 cycles in a Li-ion hybrid electrochemical superca-
pacitor. Both physical and chemical hydrothermal carbonization were used to acti-
5.3 Gasification of Coir 109

vate the carbon. Relatively high surface area (1652 m2/g) and total pore volume of
up to 0.98 ml/g were obtained. The approach used for carbonization and activation
in this study and electrical performance obtained were considered to be unique and
suitable for fabrication of high performance electro-active materials for energy stor-
age applications (Jain et al. 2013).
A one-step thermal treatment consisting of combined pyrolysis and steam activa-
tion was done on coconut shells. The shells were heated at a rate of 3 °/min up to
400 °C and maintained at that temperature for 1 h and further heated at a rate of
5 °C/min up to three different temperatures (750, 800, and 850 °C). Once the desired
temperature was obtained, the samples were treated with either steam or nitrogen
for 30, 60, or 90 min in a tubular furnace. Samples obtained were analyzed to deter-
mine the thermal and morphological properties. Similarly, the porous structure and
size and distribution of pores were studied. Cyclic voltammetry studies were also
done to determine the current voltage characteristics of the samples developed (Mi
et al. 2012). The yield of biocarbon decreased (from 20 to 7%) considerably when
the temperature of activation was increased from 750 to 850 °C, whereas the poros-
ity of the samples increased. Morphologically, pyrolysis followed by activation pro-
vided carbon a rougher and irregular surface. Activation was observed to damage
the inner and outer surfaces of the carbon leading to an increase in total pore volume
and surface area (Mi et al. 2012) (Fig. 5.10). Surface area of the samples ranged
from 826 to 1559 m2/g and pore sizes ranged from 2 to 8 nm. In terms of the electri-
cal performance, high capacitance retention of 93% could be achieved at a current
density of 5 A/g. Correspondingly, the highest capacitance achieved was 228 F/g
when activated with 6.0 mol/L KOH. Average energy density was 38.5 Wh/kg and
ESR of 1.9 Ω at 0.5 A/g. With these properties, the coconut shell carbon was con-
sidered to be suitable for supercapacitor applications.

Fig. 5.10 Comparison of the performance of coconut shell based carbons prepared using different
activation temperatures. Nitrogen adsorption isotherms are shown in (a) and BJH pore size distri-
bution is shown in (b) (Mi et al. 2012). Reproduced with permission from American Chemical
Society
110 5 Energy Applications of Coir

Table 5.10 Parameters show that the nitrogen- or urea-treated activated carbon has better
performance than normal activated coconut shell carbon (Hulicova-­Jurcakova et al. 2009)
SBET V<5Ȧ V<10Ȧ Vmic Vmeso Vt DDA VCO2 SCO2
Sample (m2/g) (cm3/g) (cm3/g) (cm3/g) (cm3/g) (cm3/g) [Ȧ] (cm3/g) (m2/g) L, [Ȧ]
S 898 0.0180 0.277 0.454 0.029 0.483 15.7 0.250 635 5.8
S-O 804 0.0024 0.251 0.410 0.013 0.423 15.9 0.294 770 6.3
S-U 808 0.0046 0.248 0.406 0.026 0.432 15.8 0.241 631 5.4
S-UO 844 0.0262 0.258 0.423 0.027 0.450 15.6 0.276 722 6.2
S-M 732 0.0073 0.239 0.372 0.014 0.386 15.6 0.252 660 5.2
S-MO 829 0.0304 0.246 0.413 0.024 0.437 15.6 0.290 760 6.2
Reproduced with permission from John Wiley and Sons
S is the coconut shell carbon without any treatment, S-O is oxidized carbon, S-U is urea-treated
carbon, S-M is melamine-treated carbon. UO and MO are pre-­oxidized samples

Fig. 5.11 Performance of the electrode in terms of specific gravimetric capacitance against cur-
rent load and capacitance retention (Hulicova-Jurcakova et al. 2009). Reproduced with permission
from John Wiley and Sons

Chemical modifications have been done to improve the properties and perfor-
mance of coconut shell based carbon (Hulicova-Jurcakova et al. 2009). Carbon from
the shells was initially oxidized using 50% HNO3 for 4 h. Later, nitrogen groups
were introduced onto the carbon by treating with urea or melamine suspension and
stirred at room temperature for 5 h. Further, the samples were heated in nitrogen at
10 °C/min up to 950 °C for 0.5 h. Some of the performance properties of the acti-
vated carbon are given in Table 5.10. Surface area of the samples decreased after
modification. The pores having size less than 5 Å showed considerable variation but
those with 10 Å had relatively less variation. The S-O treated sample had the highest
capacitance but the difference between the other samples was marginal (Fig. 5.11).
It was found that the surfaces of the samples were changed considerably due to
5.4 Catalysts for Biodiesel 111

nitrogen treatment which also showed considerable pseudocapacitance (Hulicova-­


Jurcakova et al. 2009).
Another major source/component from coconut trees, the leaves of the tree have
also been studied for potential use as electrochemical double layer capacitors.
Coconut leaves carbon was prepared by carbonizing the leaves at 400 °C under
nitrogen flow for 3 h (Sulaiman et al. 2016). Activation of the carbon was done by
passing carbon dioxide from temperatures between 700 and 1000 °C. It was
observed that the surface area increased with increasing activation temperature.
Most optimum performance for the electrodes was obtained when activated at
900 °C offering a capacitance of 133 F/g at a current density of 200 mA/g.
Ability of activated carbon obtained from coir pith to be used as an electrode was
studied by Balakumar et al. (2016). Coir pith was precarbonized by treating at
350 °C for 2 h, combined with KOH and activated by heating at 800 °C for 2 h.
About 28% carbon was obtained after the treatment and some of the carbon was also
treated with sulfur to prepare a composite for electrode application. Up to 70%
­sulfur could be added onto the carbon due to its mesoporous structure. Compared to
the raw carbon, the sulfur containing carbon composite has progressive capacity of
695 mAh/g for up to 50 cycles. In addition, the capacity value was 450 mAh/g and
coulombic efficiency of greater than 95% could be achieved (Balakumar et al.
2016).

5.4 Catalysts for Biodiesel

Inexpensive and sulfonated solid acid catalysts were prepared from coir husk by
partially carbonizing at 400 °C for 5 h. The carbon formed was mixed with 25%
sulfuric acid (H2SO4) for the sulfonation to occur. Catalyst obtained from coir husks
had surface area of 61 m2/g, mean pore volume of 14 cm3/g, and mean pore diameter
of 2.6 nm (Thushari and Babel 2018). When used as a catalyst for biodiesel produc-
tion, an increase in yield was observed. About 88.5% yield was obtained using coir
husk as catalyst and microwave-assisted biodiesel production.
In another application related to biodiesel, coir fibers were used as a substitute to
water to purify biodiesel from excess methanol, unreacted raw materials, or residual
catalyst. It was found that the coir fibers could remove soap in concentrations above
3000 ppm, could be used effectively for up to 5 years, did not increase fatty acid
levels or add Ca2+ or Mg2+, removed free glycerol and reduced methanol even at
high concentrations. Coir fibers also had economical advantage due to their low cost
at about $1 per kilogram compared to $6 to $27 for common dry sorbents used for
biodiesel purification (Ott et al. 2018).
112 5 Energy Applications of Coir

References

Balakumar K, Sathish R, Kalaiselvi N (2016) Exploration of microporous bio-carbon scaffold for


efficient utilization of sulfur in lithium-sulfur system. Electrochim Acta 209:171–182
Barin GB, de Fátima Gimenez I, da Costa LP, Filho AGS, Barreto LS (2014) Influence of hydro-
thermal carbonization on formation of curved graphite structures obtained from a lignocellu-
losic precursor. Carbon 78:609–612
Barzegar F, Khaleed AA, Ugbo FU, Oyeniran KO, Momodu DY, Bello A, Dangbegnon JK,
Manyala N (2016) Cycling and floating performance of symmetric supercapacitor derived from
coconut shell biomass. AIP Adv 6(11):115306
Dandekar MS, Arabale G, Vijayamohanan K (2005) Preparation and characterization of composite
electrodes of coconut-shell-based activated carbon and hydrous ruthenium oxide for superca-
pacitors. J Power Sources 141(1):198–203
Divyashree A, Manaf SABA, Yallappa S, Chaitra K, Kathyayini N, Hegde G (2016) Low cost, high
performance supercapacitor electrode using coconut wastes: eco-friendly approach. J Energy
Chem 25(5):880–887
Duan X-h, Srinivasakannan C, Yang K-b, Peng J-h, Zhang L-b (2012) Effects of heating method
and activating agent on the porous structure of activated carbons from coconut shells. Waste
Biomass Valoriz 3(2):131–139
Hulicova-Jurcakova D, Seredych M, Lu GQ, Bandosz TJ (2009) Combined effect of nitrogen-and
oxygen-containing functional groups of microporous activated carbon on its electrochemical
performance in supercapacitors. Adv Funct Mater 19(3):438–447
Jain A, Aravindan V, Jayaraman S, Suresh Kumar P, Balasubramanian R, Ramakrishna S, Madhavi
S, Srinivasan MP (2013) Activated carbons derived from coconut shells as high energy density
cathode material for Li-ion capacitors. Sci Rep 3:3002
Jurewicz K, Babeł K (2010) Efficient capacitor materials from active carbons based on coconut
shell/melamine precursors. Energy Fuel 24(6):3429–3435
Liu Z, Balasubramanian R (2014) A comparative study of nitrogen conversion during pyrolysis
of coconut fiber, its corresponding biochar and their blends with lignite. Bioresour Technol
151:85–90
Liu Z, Han G (2015) Production of solid fuel biochar from waste biomass by low temperature
pyrolysis. Fuel 158:159–165
Liu Z, Quek A, Kent Hoekman S, Srinivasan MP, Balasubramanian R (2012) Thermogravimetric
investigation of hydrochar-lignite co-combustion. Bioresour Technol 123:646–652
Mi J, Wang X-R, Fan R-J, Wen-Hui Q, Li W-C (2012) Coconut-shell-based porous carbons
with a tunable micro/mesopore ratio for high-performance supercapacitors. Energy Fuel
26(8):5321–5329
Mondal D, Perween M, Srivastava DN, Ghosh PK (2013) Unconventional electrode material pre-
pared from coir Fiber through sputter coating of gold: a study toward value addition of natural
biopolymer. ACS Sustain Chem Eng 2(3):348–352
Mullaivananathan V, Packiyalakshmi P, Kalaiselvi N (2017a) Multifunctional bio carbon: a
coir pith waste derived electrode for extensive energy storage device applications. RSC Adv
7(38):23663–23670
Mullaivananathan V, Sathish R, Kalaiselvi N (2017b) Coir pith derived bio-carbon: demonstration
of potential anode behavior in Lithium-ion batteries. Electrochim Acta 225:143–150
Nayai MIM, Ismail K, Ishak MAM, Zaharudin N, Nawawi WI (2016) Fabrication and charac-
terization of porous activated carbon from coconut shell by using microwave-induced KOH
activation technique. Appl Mech Mater 835:289–298
Ott LS, Riddell MM, O’Neill EL, Carini GS (2018) From orchids to biodiesel: coco coir as an
effective dry wash material for biodiesel fuel. Fuel Process Technol 176:1–6
Pestaño LDB, Jose WI (2016) Production of solid fuel by torrefaction using coconut leaves as
renewable biomass. Int J Renew Energy Dev 5(3):187–197
References 113

Ramadhas AS, Jayaraj S, Muraleedharan C (2006) Power generation using coir-pith and wood
derived producer gas in diesel engines. Fuel Process Technol 87(10):849–853
Ramadhas AS, Jayaraj S, Muraleedharan C (2008) Dual fuel mode operation in diesel engines using
renewable fuels: rubber seed oil and coir-pith producer gas. Renew Energy 33(9):2077–2083
Sheeba KN, Babu JSC, Jaisankar S (2013) Steam gasification characteristics of coir pith in a
circulating fluidized bed gasifier. Energy Sources A Recov Utiliz Environ Eff 35(2):110–121
Stelte W, Barsberg ST, Clemons C, Morais JPS, de Freitas Rosa M, Sanadi AR (2018) Coir fibers
as valuable raw material for biofuel pellet production. Waste Biomass Valoriz:1–9. https://doi.
org/10.1007/s12649-018-0362-2
Sulaiman KS, Mat A, Arof AK (2016) Activated carbon from coconut leaves for electrical double-­
layer capacitor. Ionics 22(6):911–918
Sun L, Tian C, Li M, Meng X, Wang L, Wang R, Yin J, Fu H (2013) From coconut shell to porous
graphene-like nanosheets for high-power supercapacitors. J Mater Chem A 1(21):6462–6470
Thitame PV, Shukla SR (2016) Porosity development of activated carbons prepared from wild
almond shells and coir pith using phosphoric acid. Chem Eng Commun 203(6):791–800
Thushari I, Babel S (2018) Preparation of solid acid catalysts from waste biomass and their appli-
cation for microwave-assisted biodiesel production from waste palm oil. Waste Manag Res
36(8):719–728
Yang K, Peng J, Srinivasakannan C, Zhang L, Xia H, Duan X (2010) Preparation of high sur-
face area activated carbon from coconut shells using microwave heating. Bioresour Technol
101(15):6163–6169
Yin L, Chen Y, Zhao X, Hou B, Cao B (2016) 3-dimensional hierarchical porous activated carbon
derived from coconut fibers with high-rate performance for symmetric supercapacitors. Mater
Des 111:44–50
Chapter 6
Coir for Environmental Remediation

6.1 Introduction

Coir has been one of the predominant agricultural residues studied for environmen-
tal remediation, particularly removal of heavy metals and dyes from polluted water.
Easy availability, low cost, and high moisture sorption capabilities make coir an
ideal choice for removing various contaminants from polluted water. In addition to
contaminants in water, studies have also been conducted to understand the feasibil-
ity of using coir as a sorbent for gases. Coir has also been chemically and physically
modified to improve its efficiency as a sorbate. Carbonization of coir and coir pith
has been done to improve its sorption ability.

6.2 Removal of Dyes

Dyes and waste water from textile dyeing plants are a major source of water pollu-
tion. Coir pith powdered into 75–150 μm sieve sizes were used to remove four dif-
ferent dyes (Namasivayam et al. 2001). Sorption studies were conducted in batch
mode by varying the dye concentration, pH, time, and temperature. Considerable
variations were observed in the removal efficiency for the 4 dyes. Dye uptake was
7.3, 5.6, 94.7, and 120 mg/g for acid violet, brilliant blue, rhodamine B, and methy-
lene blue, respectively. Adsorption rate constants for the same dyes using different
sorbates indicated that the coir pith had considerably high sorption rate of 12.7 × 10−2
(Table 6.1). About 200 mg rhodamine B was absorbed per g, suggesting that the
potential of coir to remove dyes was not very high (Namasivayam et al. 2001).
Mesoporous carbon prepared from activated coir dust was also found to have high
sorption for methylene blue and remazol yellow representing the cationic and
anionic dyes (Macedo et al. 2006). BET analysis showed that the carbon had a high
surface area of 1884 m2/g and pore diameter was in the range 2–4 nm. Quick

© Springer Nature Switzerland AG 2019 115


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_6
116 6 Coir for Environmental Remediation

Table 6.1 Comparison of the adsorption rate constants for four dyes using different raw materials
(Namasivayam et al. 2001)
Adsorbent Dye Adsorption rate constant, kad/min
Biogas residual slurry Congo red 2.6 × 10−2
Rhodamine B 2.4 × 10−2
Acid violet 1.4 × 10−2
Methylene blue 6.1 × 10−2
Banana pith Congo red 1.1 × 10−1
Rhodamine B 4.8 × 10−2
Acid violet 3.7 × 10−2
Acid brilliant blue 3.5 × 10−2
Orange peel Congo red 4.9 × 10−2
Rhodamine B 3.0 × 10−2
Procion orange 8.3 × 10−2
Coir pith Acid violet 2.7 × 10−2
Acid brilliant blue 1.5 × 10−2
Rhodamine B 5.5 × 10−1
Methylene blue 12.7 × 10−2
Carbonized coir pith Acid violet 1.2 × 10−1
Rhodamine B 2.9 × 10−1
Reproduced with permission from Elsevier

sorption, low cost, and a multimode sorption mechanism were suggested to be ideal
properties of the carbon prepared from coir dust. Highest adsorption of 14.4 mg/g
was obtained for methylene blue (Macedo et al. 2006). In yet another report on
removing methylene blue from waste water, coir was converted into activated car-
bon by treating with 20% ZnCl2 and later carbonizing at 700 °C for 2 h. The carbon
obtained had a surface area of 205 m2/g, total pore volume of 0.0246 cm3/g, and
mean pore diameter of 41 Ȧ. Sorption conducted at different conditions showed that
the dye removal was through first-order kinetics having a rate constant of adsorption
of 1.15 × 10−2/min. Efficiency of dye removal increased from 74 to 94% with
increase in temperature. Highest removal of the dye was about 15.5 mg of dye/g of
carbon (Sharma et al. 2009). Substantial variations in the amount of adsorption and
kinetic parameters were observed when coir pith carbon was used to remove methy-
lene blue (Table 6.2) (Kavitha and Namasivayam 2007a, b). However, the sorption
capacity reported was relatively low at 5.9 mg/g and the change in entropy and heat
of adsorption was reported to be 117 kJ/mol/K and 31 kJ/mol, respectively.
Ability of carbon prepared from coir pith to sorb congo red dye was studied by
Namasivayam et.al (Namasivayam and Kavitha 2002). About 200 mg of adsorbent
was added into a solution containing 50 mL of dye solution. The dye solutions at
different concentrations and pHs were treated at 35 °C and 200 rpm for the sorption
to take place. After the desired sorption time, the coir carbon was collected and
measured for changes in the dye concentration, which was used to calculate the
6.2

Table 6.2 Variations in the kinetic parameters during the sorption of methylene blue onto activated carbon prepared from coir fibers (Kavitha and Namasivayam
2007a, b)
Concentration K1 (min−1) qc (mg/g) R2 Temp °C K1 (min−1) qc (mg/g) R2
Removal of Dyes

Pseudo-first-order constants
10 0.1089 0.3663 0.9929 35 0.117 2.314 0.967
20 0.0525 0.6108 0.9864 40 0.124 1.992 0.989
30 0.0173 0.7803 0.9655 50 0.103 1.314 0.971
40 0.0233 1.3259 0.9780 60 0.081 0.586 0.927
Pseudo-second-order constants
h(mg/g min) K2(g/mg min) R2 Temp °C h(mg/g min) K2(g/mg min) R2
10 1.6913 0.6049 0.9998 35 1.9331 0.09 0.999
20 2.3924 0.2545 0.9998 40 2.4825 0.11 0.999
30 1.9534 0.1112 0.9999 50 3.6178 0.16 0.999
40 1.845 0.0646 0.9992 60 7.813 0.34 1.000
Intra-particle diffusion constants
Concentration kid(mg/g min) R2 Temp °C kid(mg/g min) R2
10 0.060 0.9464 35 0.0936 0.9414
20 0.1058 0.9379 40 0.616 0.9583
30 0.074 0.9859 50 0.0466 0.9566
40 0.1447 0.9558 60 0.0383 0.9922
Reproduced with permission from Elsevier
117
118 6 Coir for Environmental Remediation

efficiency of sorption and dye removal (Namasivayam and Kavitha 2002). Both
first- and second-order kinetic models showed that the adsorption rate constants
increased with increasing initial dye concentration and temperature. Removal effi-
ciency of 100% was achieved when the absorbent dosage was above
400 mg/50 mL. However, the removal was drastically reduced when the pH of the
solution was changed. Adsorption of congo red onto coir pith carbon was found to
be about 6.7 mg/g and followed both Langmuir and Freundlich isotherms
(Namasivayam and Kavitha 2002). Another common dye acid yellow 99 also
showed high levels of removal by coir pith. Up to 442 mg/g of dye could be adsorbed
by the coir through the dual mode of Langmuir–Freundlich models. To further
improve the sorption of dyes, two types of chemical modifications were done.
Acetylation of the fibers was done by treating coir with 0.1% acetic anhydride and
another esterification was done by treating with iodomethane for 4 h at 30 °C (Khan
et al. 2011). It was observed that the dye was adsorbed on the surface of the fibers
and was also able to penetrate into the micropores and macropores. However, chem-
ically modified fibers had lower percentage removal than the unmodified fibers. It
was suggested that the sorption of acid dye onto coir was both through electrostatic
and complexation reactions (Khan et al. 2011). Coconut pith carbon was also used
to remove procion orange dye from aqueous solutions. Some of the properties of the
coconut carbon obtained are given in Table 6.3. Adsorption capacity was found to
be 2.6 mg/g of carbon and increasing temperature increased sorption. However, the
sorption levels achieved in this research were considerably low compared to sorp-
tion of other dyes by coir pith or carbonized coir pith (Kavitha and Namasivayam
2007a, b).
In another study, coconut coir dust was demonstrated to have an adsorption
capacity of up to 29 mg/g for methylene blue depending on the sorption conditions
(Etim et al. 2016). It was found that percentage of sorption was directly dependent
on amount of adsorbent and dye concentration but indirectly related to pH. Some of
the kinetic parameters for the adsorption of methylene blue are given in Table 6.4.
Unlike the results shown in Table 6.4, in a study by Tan et al., a considerably high
sorption (435 mg/g at 30 °C) of methylene blue was achieved using KOH-activated

Table 6.3 Characteristics of coir pith carbon (Kavitha and Namasivayam 2007a, b)
Parameter Value Parameter Value
Specific surface area, m2/g 167 Iron, % 0.18
Bulk density, g/mL 0.12 pH 10.1
Conductivity, mS/cm 2.3 Ash content 80
Moisture content, % 5.88 Porosity, % 93
Specific gravity 1.74 Volatile matter, % 58
Decolorising power, mg/g 21.0 Fixed carbon, % 42
Iodine number, mg/g 101.5 Ion exchange capacity –
Sodium, % 0.14 Potassium, % 0.18
Calcium, % 0.22 Phosphorous, % 0.01
Reproduced with permission from Elsevier
6.2 Removal of Dyes 119

Table 6.4 Kinetic parameters of methylene blue dye adsorption by coconut coir dust (Etim et al.
2016)
Pseudo-first order Second order order
Conc. of dye (mg/g) Qc exp (mg/g) K1/min Qc cal (mg/g) R2 K2/min Qc cal (mg/g) R2
10 2.48 0.29 0.24 0.74 3.41 2.49 1.00
20 4.94 0.20 0.63 0.99 1.04 4.98 1.00
30 7.31 0.25 1.08 0.98 0.66 7.36 1.00
40 7.63 0.31 1.71 0.92 0.69 7.70 1.00
50 11.84 0.33 1.54 0.95 0.76 11.90 1.00
Reproduced with permission from Elsevier

Fig. 6.1 Changes in the percentage removal of methylene blue and orange 7 by acid-treated coir
at initial dye concentrations between 50–120 mg/l (a) and 5–20 mg/l (b) (Ong et al. 2013).
Reproduced with permission from Taylor and Francis

carbon prepared from coconut husks (Tan et al. 2008). The BET surface area of the
activated carbon was 1940 m2/g and total pore volume was 1.143 cm3/g. Considerable
number of pores was also observed, suggesting that the activation increased the
accessibility of the fibers. Ramesh et al. suggested that raw coir pith powdered into
600–300 μm could sorb considerable amounts of methylene blue. Sorption followed
second-order kinetics and equilibrium sorption could be achieved in 60 min. Amount
of dye sorbed was 143 mg/g, which was considered to be good compared to other
untreated biomasses (Ramesh et al. 2014). In another study, Ong et al. studied the
potential of removing one cationic (methylene blue) and one anionic dye (acid
orange 7) using acid-treated coconut coir (Ong et al. 2013). Acid treatment was
performed by treating shredded coir in sulfuric acid at 60 °C for 3 h. Removal of
both dyes increased with time up to about 45 min for blue and after 120 min for
orange dyes (Fig. 6.1). Initial concentration of the absorbent influenced the sorption
of orange than the blue dye. Based on Langmuir and Freundlich isotherms, maxi-
mum sorption was noted at 121 mg/g for the methylene blue and 9.7 mg/g for
120 6 Coir for Environmental Remediation

orange 7. Higher sorption of the cationic dye was suggested to be due to the nega-
tive charge on the bioabsorbent (Ong et al. 2013).
To increase sorption efficiencies and capacity, coconut coir was carbonized and
activated using H2SO4 (Aljeboree et al. 2017). The two dyes studied had higher
sorption under acidic conditions and the sorption levels increase with increase in
initial dye concentration and contact time but decreased with increase in tempera-
ture or concentration of adsorbent. A multilayer absorption was observed with the
process being spontaneous and endothermic. In a similar study, activated coir pith
carbon was used to remove three reactive dyes from waste water. Absorption of dyes
by the carbon was found to follow Freundlich model and had Lagergren rate con-
stants of 1.77 × 10−2 to 2.7 × 10−2 per min. It was also reported that the sorbed car-
bon could be regenerated several times by treating with 1 M NaOH. In addition,
treating with activated carbon also reduced the COD levels in textile waste water
(Santhy and Selvapathy 2006). Coir pith was activated using phosphoric acid and
later turned into carbon by heating in a muffle furnace at 600 °C for 2 h (Thitame
and Shukla 2016) and used for sorption of reactive dyes (C.I. Reactive Red 2 and
C.I. Reactive Yellow 145A). Morphologically (Fig. 6.2), a large number of pores
with pore size ranging from 17 to 28 μm were formed after activation and carbon-
ization, leading to a surface area of 1210 m2/g. Maximum sorption by the modified
coir pith was 2023 mg/g for reactive red and 1694 mg/g for reactive yellow. However,
the amount of sorption was highly dependent on the initial concentration, tempera-
ture, and other conditions. Some of the thermodynamic parameters for the sorption

Fig. 6.2 SEM images reveal the morphology and diameter of the pores in the carbon prepared
from coir pith (Thitame and Shukla 2016). Reproduced with permission from Springer Nature
6.2 Removal of Dyes 121

Table 6.5 Thermodynamic parameters of absorption of two dyes onto coir pith (Thitame and
Shukla 2016)
Dye Adsorbent Temp, °C ∆G° (kJ/mol) ∆H° (kJ/mol) ∆S° (kJ/mol)
RR2 AC-AS 30 −7.9 18.48 0.087
45 −9.2
60 −10.5
AC-CP 30 −9.5 20.28 0.098
45 −11
60 −12.5
RY 145A AC-AS 30 −6.6 19.52 0.086
45 −7.9
60 −9.2
AC-CP 30 −7.3 26.02 0.110
45 −8.9
60 −10
Reproduced with permission from Springer Nature

of the two dyes are given in Table 6.5 (Thitame and Shukla 2016). In addition to
carbonization, coir pith was also modified on the surface by treating with a cationic
surfactant hexadecyltrimethylammonium bromide to increase the sorption of
anionic pollutants. The modified pith was used as an absorbent for direct red 12B
(anionic) and rhodamine B (cationic) (Sureshkumar and Namasivayam 2008). The
raw coir pith used in this study was reported to have 35% cellulose, 25% lignin,
7.5% pentosans, 1.85% fats and resins, and about 8.7% ash. Sorption of the dyes
from various types of polluted water showed that the highest removal efficiency
(90%) was for direct 12B when the pH was between 5 and 8. Compared to 12B,
maximum sorption for rhodamine B was observed between pHs 7 and 10. In terms
of sorption capacity, the modified pith had 76 mg/g sorption for 12B and 15 mg/g
for rhodamine B. The sorption rates achieved in this study was considerably low
compared to other reports using modified coir pith. In addition, desorption of dyes
from the coir pith was also considerably low at around 10% in pH between 2 and 12.
It was suggested that such low desorption was due to the chemical interactions
between the specific dyes and the modified fibers (Sureshkumar and Namasivayam
2008).
In a similar approach, holocellulose prepared from coir fibers was used as a sor-
bent for reactive orange 16 (Ray et al. 2014). The holocellulose extracted from coir
was also grafted with poly(acrylamide) to obtain a grafting efficiency of about 58%.
Extent of dye removal was dependent on the level of grafting and other reaction
conditions. Although about 72% dye could be removed by the grafted holocellulose
compared to only 42% for the ungrafted cellulose, the amount of dye sorbed was
lower than that achieved using coir in other studies (Ray et al. 2014).
122 6 Coir for Environmental Remediation

6.3 Removal of Heavy Metals

In addition to dyes, coir and modified carbon prepared from coir pith has also been
used to remove organic chemicals and heavy metals. It has been shown that ZnCl2
activated carbon could efficiently remove various heavy metals with removal effi-
ciencies ranging from 41 to 100%. However, unactivated carbon had negligible
removal efficiency for the inorganic anions and heavy metals but showed up to 76%
removal for organics and 92% for dyes (Tables 6.6 and 6.7) (Namasivayam and
Kavitha 2006; Namasivayam and Sangeetha 2006). One of the rare metals galla-
nium was also successfully extracted from aqueous solution of gallium nitrate using
unmodified and oxidized coir (Suryavanshi and Shukla 2009). Oxidizing the coir
led to about 1.5× increase in surface area and consequently the sorption was also
high (70%) and the sorption capacity was about 19 mg/g. Adsorption of the Ga (III)
onto coir was found to follow Langmuir isotherm and had pseudo-second-order
kinetics. Also, near complete recovery of the coir was possible just by treating with
hydrochloric acid (Suryavanshi and Shukla 2009). However, the sorption of Ga(III)
on coir was lower compared to other ions such as Fe(II) and Ni(II). To further
improve removal of Co(II), coir pith was chemically modified by esterification
using succinic anhydride. A relatively high removal of 33 mg/g was obtained since
the modification provided additional sites on the surface for adsorption. Complete
removal of the sorbed metal ions was possible by treating with 1 N HCl (Parab et al.
2008). A study on the ability of coir pith to remove Co(II), Cr(III), and Ni(II) from
single ion solutions as well as from a mixture showed a maximum sorption of 12.8,
11.6, and 16.0 mg/g for the three metal ions, respectively. Since these metal ions are
present in a mixture, sorption from nuclear power plant coolant water was also stud-
ied to determine any selective sorption (Parab et al. 2006). Under competitive sorp-
tion, it was found that Cr was taken up more aggressively than the other two ions

Table 6.6 Removal efficiencies of various entities for untreated and ZnCl2 activated coir pith
carbon (Namasivayam and Kavitha 2006; Namasivayam and Sangeetha 2006)
% Removal
Adsorbate Initial pH Adsorbent dose (mg/50 mL) ZnCPC CPC
Nitrate 6.2 400 95 –
Phosphate 4.0 450 91 –
Vanadium (V) 4.0 250 99 –
Thiocyanate 6.2 250 100 –
Molybdate 4.0 200 99 –
Sulfate 4.0 300 70 –
Selenite 6.0 600 74 –
Chromium (VI) 2.0 50 99 18
Nickel (II) 5.0 600 41 58
Mercury (II) 5.0 200 100 80
Reproduced with permission from Elsevier
6.3 Removal of Heavy Metals 123

Table 6.7 Removal efficiencies of different chemicals by the activated coir pith carbon
(Namasivayam and Kavitha 2006; Namasivayam and Sangeetha 2006)
% Removal
Adsorbate Initial pH Adsorbent dose (mg/50 mL) ZnCPC CPC
Phenol 2.0 100 98 65
Resorcinol 4.0 100 99 61
4-Nitrophenol 4.0 100 100 50
2-Chlorophenol 4.0 50 99 67
Bisphenol A 2.0 50 97 57
Catechol 2.0 100 98 54
O-Cresol 2.0 50 98 52
Quinol 2.0 50 100 76
2-Aminophenol 2.0 50 100 50
Reproduced with permission from Elsevier

12

10

8
q(mg/g)

0
0 20 40 60 80 100 120
time (min)

Co Ni Cr

Fig. 6.3 Level of sorption of metal ions under competitive conditions shows preference for Cr
(III). Uptake studies were conducted using a metal ion concentration of 50 mg/l and adsorbent dose
of 0.1 g/50 mL and temperature of 300 k (Parab et al. 2006). Reproduced with permission from
Elsevier

(Fig. 6.3). However, the sorption levels achieved in this study were considerably
low compared to using other biomasses (Parab et al. 2006).
Similar studies have also been done to adsorb Ni(II), Zn(II), and Fe(II) from
waste water using unmodified and hydrogen peroxide treated coir fibers. For the
oxidation using hydrogen peroxide, fibers were immersed in 50% solution at
pH 10.5 and heated up to 85 °C for 2 h. Both percentage of adsorption (efficiency)
124 6 Coir for Environmental Remediation

and amount of adsorption showed considerable difference depending on the type


of metal ions present. All the three metals had considerably high adsorption after
oxidation. Among the three metals, Zn(II) had higher removal (7.9 mg/g) followed
by Fe(II) (7.5 mg/g) and Ni (II) (4.3 mg/g). Regeneration studies showed that same
level of sorption was possible even after three cycles when the desorption was done
using NaOH-treated fibers (Shukla et al. 2006). It was suggested that the ability to
absorb and desorb effectively by the modified coir fibers could be used as an ion
exchanger for purification of different types of waste water (Shukla et al. 2006).
Parab et al. studied the effectiveness of coir to remove cobalt from waste water
(Parab et al. 2010). Coir fibers powdered to about 300–600 μm were added into
water containing various concentrations of Co(II). Batch studies were conducted to
determine the sorption kinetics, efficiency, and metal uptake under different condi-
tions. Increasing the concentration of cobalt decreased, whereas increasing coir
concentration increased sorption efficiency at equilibrium (Parab et al. 2010). Based
on Langmuir adsorption models, the maximum sorption capacity was found to be
14 mg/g of cobalt. A selective and complete removal of cobalt from a mixture of
metals present in nuclear power plant coolant water was achieved. Also, the sorbed
material could be regenerated by simple acid treatment. Hence, coir was considered
to be an ideal material to remove cobalt from waste water (Parab et al. 2010). A
biofiltration system was developed to remove heavy metal ions from urban storm
water runoff systems (Lim et al. 2015). Coir fibers having a pH of 4.8 and limiting
oxygen index (%) of 96 were placed into columns, and water containing various
dosages of different metals was passed through the columns. Coir fibers were able
to sorb more than 90% of Zn and Cd but comparatively lower Cu. It was suggested
that coir may have to be mixed with other substances to efficiently remove metals
from storm water runoff (Lim et al. 2015). Both coir fibers and coconut shell were
used to prepare activated carbon for sorption of Fe3+ ions in aqueous solutions
(Chukwuma et al. 2013). Removal efficiencies (%) ranged from 86 to 97.5 and 20
to 98% depending on the concentration of the absorbent. Generally, the shell had
higher sorption than the fibers because of the higher porosity.
In addition to having high sorption, ability to desorb and resorb for multiple
cycles is necessary for economical recovery of contaminants. Coir fibers were
washed with water, powdered and treated with 0.1 M NaCl for 12 h before using for
sorption of chromium (III) and chromium (VI) in contaminated waste water. It was
suggested that chemical interactions between the hydroxyl groups in the carbohy-
drates and central chromium ions led to good sorption. However, reports have sug-
gested that a pH of 1.5 is more suitable for removal of chromium ion but the pH of
the waste water was 7 and hence needed reduction in pH to at least 4.5. There was
considerable reduction of the Cr(VI) sorbed through a two-step process, and the
material could be recovered and reused for multiple sorptions making the coir fibers
suitable for developing biofilters for waste water treatment (Henryk et al. 2016).
The potential of using coconut coir for removal of copper and subsequent desorp-
tion in different mediums were investigated. Hydrochloric acid medium provided
good desorption, up to three cycles (Chang et al. 2016). Most of the copper was in
the form of CUO and was sorbed onto the surface of the coir fibers due to higher
6.3 Removal of Heavy Metals 125

Fig. 6.4 Ability of


coconut coir to adsorb and
desorb copper (Chang et al.
2016). Reproduced with
permission from Elsevier

porosity and void content. In addition, the CUO could be desorbed easily even by
washing with water (Fig. 6.4). In addition to the fibers, it was also shown that yarns
made from coir fibers and grafted with acrylic acid and phosphoric acid could effi-
ciently remove copper from CuSO4 solutions (Zaman et al. 2013). Grafting increased
the strength of the fibers and the amount of Cu uptake. Unmodified fibers had
0.58 mg/g of Cu uptake compared to 0.96 after treating with acrylic acid. A
­combination of acrylic acid and phosphoric acid provided an increased uptake of
3.85 mg/g. In addition, high regeneration was possible when the threads were
treated with HCl solution. About 0.12 mg and 0.98 mg of Cu could be recovered per
gram of coir after treating in water and acid for 2 h, respectively (Zaman et al. 2013).
Activated carbon was also prepared from coconut buttons (unmatured coconuts)
treated with acid. For carbonization, acid-treated coconut buttons were placed in a
muffle furnace and heated up to 400 °C for 1 h to obtain a yield of about 40% carbon
(Anirudhan and Sreekumari 2011). Detailed characterization of the activated carbon
showed that the material has a specific surface area of 479 m2/g, micropore volume
of 0.14 mL/g, and average particles with diameter of 0.096 mm. Removal efficiency
of all three metals (Cu(II), Hg (II), and Pb(II)) increased with increasing adsorbent
dosage and reached about 90–100% (Fig. 6.5) (Anirudhan and Sreekumari 2011).
Highest sorption capacity was found to be 93 mg/g for Pb(II), 79 mg/g for Hg(II),
and 74 mg/g for Cu(II). Unlike most absorbents, considerable high regeneration was
obtained even after three adsorption and desorption cycles. A simple treatment of
coir fibers with 18% sodium hydroxide solution at 60 °C for 4 h was highly effective
to increase sorption of Cu(II) and Pb(II) in a fixed bed column (Singh et al. 2017).
Removal efficiency and sorption capacity varied with changing conditions and
ranged from 70 to 81% and 10 to 24 mg/g depending on the metal sorbed (Table 6.8).
A high regeneration efficiency of about 80–82% was obtained even after three
cycles of treatment (Singh et al. 2017). Similar results were also obtained when coir
fibers were oxidized and used to remove Cu(II) in waste water using a fixed bed
column. Oxidation using hydrogen peroxide increased the accessibility of the coir
126 6 Coir for Environmental Remediation

Fig. 6.5 Efficiency of Pb(II), Hg (II), and Cu(II) ion removal from activated carbon prepared from
coconut buttons (Anirudhan and Sreekumari 2011). Reproduced with permission from Elsevier

Table 6.8 Sorption conditions and parameters for the removal of two metals in waste water by
coir fibers (Singh et al. 2017)
Cycle Breakthrough Exhaustion Breakthrough Regeneration
Metal no. time, min time, min uptake, mg/g efficiency, %
Cu(II) 1 300 950 9.06 –
2 388 794 8.15 90
3 346 696 7.19 79
Pb(II) 1 1271 2082 23.33 –
2 1168 1910 21.32 92
3 964 1785 19.04 82

from 19 to 44% as evident from the SEM image (Fig. 6.6) (Shukla et al. 2009). It
was suggested that the oxidizing treatment results in the formation of carboxylic
acid groups which will be in their sodium salt in the alkaline medium used during
sorption. Maximum sorption of 471 mg/g was possible and more than 90% of the
metal could be recovered even after eight sorption-desorption cycles (Shukla et al.
2009).
Modified and unmodified coir fibers were studied for potential removal of Zn
and Pb ions. For the treatment, fibers were immersed in 0.01 M NaNO3, whereas for
sorption of Zn, the fibers were swollen in 100 mL of polyethylene with electrolyte
(Conrad and Hansen 2007). Sorption studies were conducted by varying the con-
centration of the metal ions, pH, time, temperature, etc. Highest sorption obtained
6.3 Removal of Heavy Metals 127

Fig. 6.6 Changes in the morphology of the coir fibers before (top) and after oxidizing with hydro-
gen peroxide (Bottom) (Shukla et al. 2009). Reproduced with permission from Taylor and Francis

for Pb was 18.9 mg/g and 8.6 mg/g for Zn which is considerably lower than the
sorption by other biomasses (Table 6.9). Desorption of the metals from the fibers
was also low at <13% for Zn and <1% for Pb (Conrad and Hansen 2007).
Batch sorption studies on coir fiber for the removal of four different metal ions
Cu(II), Pb(II), Ni(II), and Fe(II) were done by Shukla and Shukla (2013). Coir fibers
were first washed with detergent and later treated with 18% NaOH solution at 60 °C
128 6 Coir for Environmental Remediation

Table 6.9 Comparison of the maximum sorption reported for Pb and Zn using various sorbents
(Conrad and Hansen 2007)
Sorbent Pb (mg/g) Zn (mg/g) pH Electrolyte
Coir 18.9 8.6 5.6 NaNO3
Coir 48.8 – 4.5 Water
Barley straw 15.2 – 5.5 Water
Rice husk 4.0 _ 6.0 Water
Hazel nut shell 1.8 _ 4.0 Water
Oak stem 0.8 _ 5.2 Water
Barley straw – 5.7 6.3 Water
Coniferous bark – 7.4 4.0 Water
Peat – 11.7 4.5 Unknown
Reproduced with permission from Elsevier

Table 6.10 Maximum sorption capacity of the metal ions by untreated and alkali-treated coir
fibers (Shukla and Shukla 2013)
Metal ion Ionic radius (Ȧ) Untreated coir Alkali-treated coir
Cu(II) 0.73 2.77 9.43
Pb(II) 1.19 10.41 29.41
Ni(II) 0.69 2.65 8.84
Fe(II) 0.78 2.94 11.11
Reproduced with permission from Taylor and Francis

for 12 h to remove some of the surface substances and increase sorption capacity.
Maximum sorption capacity obtained was considerably higher after treating with
alkali. Table 6.10 shows the amount of metal ions sorbed by the untreated and
treated coir fibers. Similarly, the desorption capacity and the ability to recycle the
coir were also high (Table 6.11), suggesting that the material was ideal for removal
of metals (Shukla and Shukla 2013). A considerably high removal of 22.7 mg/L of
Cr (VI) was possible when electroplating industry waste water was used for sorp-
tion. A thorough investigation was done to understand the mechanism by which Cr
(VI) is absorbed by coir fibers (Shen et al. 2010). Complete removal of 50 mg/L of
Cr(VI) was achieved in 12 h, and the Cr(VI) was converted into Cr(III) in solution
(21%) or to the Cr that was bound to the coir fibers (79%). XPS studies showed that
Cr bound to the fibers was predominantly in the trivalent form. A reductive transfor-
mation from Cr(VI) to Cr (III) made the metal to be bound onto the coir fibers.
Although cellulose and hemicellulose appeared not to be affected by the absorption,
oxidation of lignin was apparent based on the decrease in the intensity of the FTIR
peaks at 1443 and 1281 cm−1. The carboxyl and carbonyl groups formed were also
supposed to act as the binding sites and increase the absorption of Cr(III). It was
suggested that a biomass with high lignin content would be beneficial to achieve
higher Cr removal from waste water (Shen et al. 2010).
6.3 Removal of Heavy Metals 129

Table 6.11 Extent of desorption (%) of the various metal ions by untreated and alkali-treated coir
fibers when treated with different concentrations of acid (Shukla and Shukla 2013)
Conc. Cu(II) Pb(II) Ni(II) Fe(II)
of acid Untreated Treated Untreated Treated Untreated Treated Untreated Treated
0.05 7.3 82.7 62.6 73.3 71.9 80.4 70.6 79.3
0.10 84.6 93.0 75.5 88.1 83.6 92.5 82.5 91.2
0.15 92.1 96.8 86.1 94.2 91.3 95.5 90.1 94.9
0.25 92.3 96.7 90.5 96.2 92.1 95.7 91.5 95.1
0.50 94.1 96.7 91.1 97.2 93.2 96.1 92.1 95.9
Reproduced with permission from Taylor and Francis

Table 6.12 Physiochemical Coir Granulated


characteristics of two forms Characteristic carbon coir carbon
of coir carbon (Santhy and
Moisture content, % 8.9 8.4
Selvapathy 2004)
Ash content, % 2.04 2.17
Soluble matter, % 1.26 0.98
Acid soluble matter, % 3.04 2.18
pH 8.2 8.5
Decolorizing power (mg/g) 30 26
Ion exchange capacity (meq/g) 2.34 2.27
Iron content, % 0.048 0.026
Phenol number 134 149
Apparent density (g/cm3) 0.19 0.52
Surface area (m2/g) 877 826
Reproduced with permission from Taylor and Francis

Coir activated using potassium hydroxide and carbonized at 700–750 °C was


studied for potential removal of Cu(II), Cd(II), and Zn(II) using batch and column
studies. Some of the properties of the coir carbon are listed in Table 6.12 (Santhy
and Selvapathy 2004). Coir dust composed of 36% cellulose, 54% lignin, and 25%
moisture was studied for potential removal of zinc from aqueous solution (Israel
and Eduok 2012). Extent of sorption was dependent on time, pH, and particle size
with increase in size decreasing the sorption. A pseudo-second-order sorption that
fit Freundlich isotherm was observed along with intra-particle diffusion. Based on
the Gibb’s free energy change, it was suggested that spontaneous sorption of Zn(II)
ions was possible onto coir dust (Israel and Eduok 2012).
Coir pith made into particle of about 75 μm was used as a sorbent for hexavalent
chromium (Cr (VI)) found in electroplating waste water (Suksabye et al. 2007). To
study the sorption capabilities, powdered coir was placed in electroplating waste
water (pH 2) for 5 min to 24 h at 30 °C and the concentration changes in the solution
were measured using a spectrophotometer at a wavelength of 540 nm. The coir pith
130 6 Coir for Environmental Remediation

had average pore size of 77 Ȧ and total pore volume of 0.0144 cm3/g. Surface area
of the coir pith powder was considerably low at 7.4 m2/g compared to activated
carbon obtained from coconut shells (1113 m2/g). pH of the solution and contact
time affected the removal efficiency that was as high as 99.99% at pH 2. Most of the
chromium was bound onto the coir in the form of Cr (III) and it was suggested that
lignin was involved in the sorption and the entire process was mainly through che-
misorption (Suksabye et al. 2007).
Efficient removal of lead (PbII) was also possible using activated coir pith
(Kumar et al. 2013). For the study, coir pith was washed with sulfuric acid, neutral-
ized and carbonized in a muffle furnace by heating at 600 °C for 1 h. Optimum
conditions for lead removal were found to be pH 5, adsorbent dose of 1 g/L and
solution concentration of 20 mg/L. Similarly ∆H° and ∆S° were found to be 2.54 kJ/
mol and 17.6 J/mol/K, respectively, indicating the sorption was physical in nature
and there was good affinity of Pb(II) to the activated coir pith. A highest sorption of
51 mg/g was reported (Kumar et al. 2013). Chemical modification of coir has also
been considered for increasing the efficiency of metal removal. For instance, coco-
nut coir pith was modified using amines and grafted with polyacrylamide to add the
functional groups (NH3+Cl−). Grafting of acrylamide increased thermal stability
and also activation energy for thermal degradation. However, decrease in crystallin-
ity and consequently strength of the fibers was observed. High removal efficiency of
99.4% and maximum adsorption of 12.4 mg/g were achieved using an initial con-
centration of 25 mg/L Cr (VI) solution at 30 °C having 2 g/L of the adsorbent
(Unnithan et al. 2004).
Coir pith was grafted with polyacrylamide using potassium peroxydisulfate as
initiator and N,N′-methylenebisacrylamide as crosslinking agent to remove Pb(II),
Hg(II), and Cd(II) ions from aqueous solutions (Anirudhan et al. 2007). Such graft-
ing provided a carboxylate functional group at the end of the coir chain and increased
sorption. The modified coir pith had a porous structure with surface area of 42 m2/g.
Considerably high sorption of metals was possible in the pH range of 6.0–8.0 with
maximum adsorption being 254, 189, and 64 mg/g for Pb(II), Hg(II), and Cd(II)
ions, respectively, which was two to five times higher compared to the sorption by
the ungrafted coir pith. Such high absorption was also seen for waste water gener-
ated by battery manufacturing industry, fertilizer industry, and chlor-alkali industry.
Also, desorption and regeneration studies showed that 89–94% adsorption and
88–91% desorption were possible even after the third cycle and hence coir pith was
considered to be an ideal sorbent for heavy metal removal (Anirudhan et al. 2007).
Arsenic (As(V)) is one of the dangerous but notoriously difficult contaminant to
remove from waste water. A new anion exchanger was prepared from coir pith by
adding dimethylaminohydroxylpropyl as a weak functional group on the fibers. The
modification was achieved by the reaction between coir, epichlorohydrin, and
dimethylamine, and later treating with hydrochloric acid. FTIR studies confirmed
the presence of −NH+(CH3)2Cl− which favors absorption of As (V). Maximum
removal efficiency of 99.2% was achieved in batch studies. In addition, high absorp-
tion was also seen from ground water and it was possible to obtain recovery rates
higher than 93% even after 4 cycles suggesting that coir was highly effective for
6.3 Removal of Heavy Metals 131

Fig. 6.7 Level of mercury sorption by the untreated and untreated coir pith and coir fibers (Johari
et al. 2016). Reproduced with permission from Taylor and Francis

removal of Ar(VI) from water (Anirudhan and Unnithan 2007). In a similar


approach, a cationic exchanger was created when coir pith was grafted with
poly(hydroxyethylmethacrylate) using potassium peroxydisulfate as initiator and
N,N′-methylenebisacrylamide as the crosslinking agent for potential use as an sor-
bent for mercury (Hg(II)). More than 99% of mercury was removed using a pH
range of 5.5–8.0 both for laboratory water and chlor-alkali industry waste water.
Regeneration of coir was also considerably high with more than 90% sorption even
after the fourth cycle (Anirudhan et al. 2008). Coir pith was also subject to alkali
treatment and bleaching and studied for its potential to remove mercury (Johari
et al. 2016) in comparison to coir fibers. Mercury sorption level ranged from 431 to
956 ng/g depending on the type of modification used and also varied between coir
pith and fibers (Johari et al. 2016). Pith showed higher sorption before and after the
modifications (Fig. 6.7) (Johari et al. 2016) with highest sorption of 956 ng/g.
Studies were also done to determine if coconut pith and coconut fiber with and
without surface treatments would be able to remove elemental mercury (Johari et al.
2014, 2016). Mercerization was done by treating the coir with 5% NaOH and later
bleached using hydrogen peroxide or treated with NaOCl (soaking for 2 h at 30 °C)
to improve the morphology and sorption ability. Amount of mercury sorbed was
dependent on the pH and type of fibers. An increase in absorption was observed for
both treated and untreated fibers up to pH 5 above which the sorption decreased.
Similar phenomenon was also observed when temperature or initial concentration
was varied. Some of the thermodynamic parameters for the mercury sorption onto
the coir fibers are presented in Table 6.13 (Johari et al. 2014).
Coir pith has been studied for potential removal of radioactive isotopes such as
cesium and strontium. Potential of using coir pith with and without modification to
remove cesium was explored (Parab and Sudersanan 2010). Modification of coir
132 6 Coir for Environmental Remediation

Table 6.13 Thermodynamic parameters of Hg(II) sorption onto coir fibers (Johari et al. 2014)
Fiber Temperature, °C ∆G° (kJ/mol) ∆H° (kJ/mol) ∆S° (kJ/mol-K)
Raw fibers 30 −7.2272 −68.0251 8.1168
45 −5.0568
60 −1.1788
Modified fibers 30 −5.5725 −54.8415 8.1577
45 −1.3910
60 −0.8528
Reproduced with permission from Taylor and Francis

was done by the incorporation of nickel hexacyanoferrate. Nearly 100% increase in


uptake from 33 mg/g to 66 mg/g was possible due to the modification. Low cost and
rapid sorption (less than 30 min) were suggested to be good properties for using
modified coir pith as a sorbent for cesium (Parab and Sudersanan 2010). In another
study, coir pith was subject to gamma irradiation using a dosage range of 1.2–
3.6 MGy (60Co source) at a dosage rate of 5 kGy/h for potential removal of stron-
tium (Parab et al. 2016). After modification, the coir pith was immersed in Sr (II)
solution for 5 h and strontium uptake and removal was calculated using atomic
absorption spectroscopy at a wavelength of 461 nm. There was no apparent change
in the morphological features of coir fibers due to irradiation but decrease in thermal
stability of hemicellulose and cellulose was observed. Coir pith before irradiation
sorbed about 26 mg of Sr per gram of coir (53%) but sorption decreases to 20 mg/g
(40%) after irradiation in water due to degradation by radiolysis (Parab et al. 2016).
Similar to arsenic, phosphates released from industrial process and agricultural
practices are responsible for toxic algal bloom, frothing, and many other issues.
Coir pith impregnated with iron was shown to be highly effective for removal of
phosphinates from both aqueous solutions and sewage water (Krishnan and Haridas
2008). For the treatment, coir was modified using Fe(NO3)3.9H2O solution and
batch sorption experiments were done. FTIR and X-ray diffraction studies showed
that Fe was sorbed on the surface of coir pith. Exhaustion percentage varied from 0
up to 90% and amount of sorption was as high as 68 mg/g for phosphates in aqueous
solutions. In the case of sewage sludge, an adsorbent dosage of 2 g/L was sufficient
to remove 82% of the phosphate and 100% removal was possible when 4 g/L of
modified coir was used for sorption (Krishnan and Haridas 2008). In another study,
coir pith was activated using zinc chloride and made into carbon for removal of
phosphate under various sorption conditions (Namasivayam and Sangeetha 2004).
Up to 90% removal of phosphate was possible and sorption capacity was 5.1 mg/g
at optimal pH between 3 and 10. The cocopith carbon was also able to desorb up to
50% of the phosphate but at pH 2 and 11. Good sorption could be achieved even
when various other ions were present in the solution, suggesting that coir could
remove phosphates from industrial effluents (Namasivayam and Sangeetha 2004).
Some of the parameters for the diffusion process are given in Table 6.14.
Presence of high levels of lignin in coir pith was considered an advantage to
remove nickel from waste water. To improve removal efficiencies, the coir pith was
6.4 Chemicals and Compounds 133

Table 6.14 Diffusion parameters for phosphate removal from aqueous solutions using activated
coir pith carbon (Namasivayam and Sangeetha 2004)
Eq. Eq.
Phosphate time, t1/2 D × 1013, Phosphate Temp, time, t1/2 D × 1013,
conc. Mg/L min (s) cm2/s conc. Mg/L °C min (s) cm2/s
10 80 2400 4.418 10 35 80 2400 4.418
20 140 4200 2.525 40 80 2400 4.418
30 160 4800 2.209 50 60 1800 5.891
40 160 4800 2.209 60 40 1200 8.836
Reproduced with permission from Elsevier

also modified by treating with 0.1 N NaOH. Extent of nickel removal was 9.5 and
39 mg/g (91% removal) for the modified and unmodified coir pith, respectively
(Ewecharoen et al. 2008). Presence of lignin and holocellulose contributed to the
sorption of the metal and a process of chemisorption was identified.

6.4 Chemicals and Compounds

Phenol is one of the common and toxic pollutants found in water, mainly released
from petrochemical, plastics, paint, pharmaceutical, and steel industries (Din et al.
2009). Coconut shells were carbonized and later activated with KOH for potential
removal of phenol in batch experiments. Carbonization of the shells was done at
700 °C and ratio of KOH to carbon was 1:1 to achieve a total surface area of
1026 m2/g, micropore area of 943 m2/g, total pore volume of 0.57 cm3/g, and aver-
age pore diameter of 2.2 nm. Adsorption of phenol onto activated carbon followed
both Langmuir and Freundlich isotherms. A highest sorption of 205 mg of phenol
per gram of carbon was achieved (Din et al. 2009). A considerably high sorption of
716 mg/g of 2,4,6-trichlorophenol (TCP) was obtained for coconut husk carbonized
and activated with KOH (Hameed et al. 2008). Sorption capacity increased with
increase in initial concentration of the absorbate or agitation time but decreased
when the pH was increased from acidic to alkaline. More than 7–100 times higher
sorption of TCP was achieved in this study compared to previous studies that had
coir pith in various forms (Hameed et al. 2008). A pseudo-second-order kinetic
model absorption was detected and particle diffusion was also observed. Langmuir
and Freundlich isothermal adsorption was observed in pseudo-second-order form
for 2,4-dichlorophenol in waste water when coir pith was used as the absorbent
(Namasivayam and Sangeetha 2004). Sorption was directly dependent on parame-
ters such as concentration of drug, absorption time, pH, and temperature. Nearly
100% removal was achieved and the coir pith were suggested to be as good as com-
mercially available carbon sorbents but at a considerably low price (Namasivayam
and Sangeetha 2004). In a similar study, carbon made from coir pith was considered
for removal of p-chlorophenol and p-nitrophenol from water and effluents released
134 6 Coir for Environmental Remediation

from the petroleum industry (Anirudhan et al. 2009a, b). Coir pith containing 48%
cellulose, 24% lignin, and 17% hemicellulose was treated with phosphoric acid to
increase surface area and hence sorption efficiencies. Activated carbon with specific
surface area ranging from 406 to 474 m2/g and pore volume from 0.28 to 0.41 mL/g
was obtained by varying the carbonizing conditions. Due to these variations in the
properties of the carbon and also sorption conditions, the amount of phenol sorbed
varied from 25 to 100% and was mainly dependent on the initial concentration and
pH (Fig. 6.8). Absorption-desorption studies showed that the carbon can be regener-
ated up to four cycles by treating with alkali or heat with about 10–15% decrease in
sorption after each cycle (Anirudhan et al. 2009a, b).
In addition to removal of solid contaminants such as metals and minerals, coco-
nut shell has been activated using zinc chloride, carbonized and used as an sorbent
for p-Cresol (4-methylphenol) which is responsible for the malodor at swine farms
(Zhu and Kolar 2014). Psuedo-second-order adsorption was observed and the maxi-
mum monolayer adsorption was 30.2 mg/g, which increased up to 33 mg/g when
the temperature was 313 K. The level of adsorption obtained in this research was
considered to be high when compared to previous reports (Zhu and Kolar 2014)
(Table 6.15). Although effective removal of p-cresol was reported, it was not clear

Fig. 6.8 Effect of increasing pH and concentration of effluent on the adsorption of phenols by
activated carbon prepared from coconut coir pith (Anirudhan et al. 2009a, b). Reproduced with
permission from Elsevier
6.4 Chemicals and Compounds 135

Table 6.15 Comparison of the absorption capacity of activated coconut shell powder and other
common absorbents (Zhu and Kolar 2014)
Maximum adsorption capacity,
Adsorbent Qd (mg/g) Adsorbate Temperature (K)
Commercial activated carbon 4.74 Cresol 303
Fly ash 6.7 Cresol 301
Pine wood activated carbon 6.97 p-cresol 298
Clay and membrane 8.8 o-cresol 298
Swine manure car 14.99 p-cresol 308
Tobacco residue 17.83 phenol Unknown
Amberline XAD-4 resin 27.59 p-cresol 308
Coconut shell activated char 32.77 p-cresol 313
Commercial activated 56.61 p-cresol 298
alumina
Parthenium-based activated 62.91 p-cresol 301
carbon
Diatomite/carbon composites 86 p-cresol Room
temperature
Reproduced with permission from Elsevier

if the removal was sufficient to mitigate the bad odor. An antibacterial agent levo-
floxacin was removed using alumina-doped coconut coir fibers (Limbikai et al.
2016). Coir fibers were cut into pieces and impregnated with Al2O3 by in situ com-
bustion of coir, aluminum nitrate, and urea at 400 °C for 15 min. No significant
changes were observed in the morphology of the fibers. Sorption increased with pH
up to 7 and intial concentration. Highest sorption obtained was about 70% and fol-
lowed Langmuir isotherm model (Limbikai et al. 2016).
A novel approach of preparing a metal complex and grafting polyacrylamide
onto coir pith was done and used to remove phosphate from water and waste water
(Anirudhan et al. 2009a, b). Sorption of phosphate was highly dependent on pH
with greater than 99% removal in the pH range between 4.0 and 7.0. Highest amount
of sorption was 14.4 mg/L from the fertilizer industry waste water. Treating the
sorbent with 0.1 M HCl reintroduced the Fe3+ ions which help in sorption of the
phosphates. The sorbent was reusable for several cycles without any apparent loss
in sorption capacity (Anirudhan et al. 2009a, b). Much higher levels of phosphate
removal were obtained when sulfuric acid activated carbon was prepared from coir
pith. Obtained carbon had an average pore area of 727 m2/g and pore size of 19Ȧ.
However, highest removal was obtained between pH 6 and 10 with adsorption fol-
lowing Temkin isotherm and was spontaneous and exothermic (Kumar et al. 2010).
Some of the thermodynamic parameters of the adsorption of phosphates by the
activated coir pith carbon are shown in Table 6.16 below.
136 6 Coir for Environmental Remediation

Table 6.16 Thermodynamic parameters of activated coir pith carbon for sorption of phosphates
(Kumar et al. 2010)
Temperature (K) KD -∆G0 (kJ/mol K) ∆H0 (kJ/mol) ∆S0 (J/kmol)
308 3.0502 2.856 3.8843 21.885
323 3.2766 3.187 – –
333 3.4175 3.402 – –
Reproduced with permission from Taylor and Francis

6.5 Desalination

Coconut coir dust was activated by treating with 50% phosphoric acid by boiling for
1 h and drying in an oven at 100 °C for 48 h. Treated coir dust was later pyrolyzed
by heating at 450 °C for 1 h (Hettiarachchi et al. 2016) to obtain activated coir car-
bon for potential desalination applications. Both iodine value (601 mg/g) and meth-
ylene blue value (331 mg/g) were considerably higher for the activated coir carbon
compared to commercially available carbon. Morphologically, the carbon had nano-
porous structure with a BET surface area of 1235 m2/g. Adsorption capacity of the
carbon for Na+ and Mg2+ was dependent on the amount of carbon and varied from
5.3 to 44 mg/g and 0.4 to 5.5 mg/g for Na+ and Mg2+, respectively. Under the opti-
mized conditions, up to 80% of Na+ and 72% of Mg2+ was achieved after repeated
filtration of the seawater. It was also possible to regenerate the carbon after washing
with distilled water (Hettiarachchi et al. 2016).

6.6 Sorption of Gases

In addition to heavy metals, dyes, and other pollutants found in waste water, pres-
ence of CO2 and other gases in the atmosphere are highly harmful to plant and ani-
mal lives. Several attempts have been made to remove CO2 from the environment by
using sorbents at the site of release or off-site locations. Coir fibers were treated to
obtain activated carbon with surface area of 214 m2/g, pore diameter of 1.5 Ȧ, bulk
density of 350 kg/m3, and particle diameter of 0.92 mm. The activated carbon was
used to sorb CO2 released from flue gas in a laboratory setting (Hauchhum et al.
2015). Adsorption isotherms showed that up to 3 mol/kg of CO2 could be removed
by the coconut carbon depending on the pressure and temperature (Fig. 6.9). In
addition, the sorption was reversible and complete regeneration was observed for
five cycles. Low cost, high sorption, and recyclability were considered to be very
favorable for removal of CO2 from flue gas by the coconut activated carbon
(Hauchhum et al. 2015).
6.7 Desulfurization of Diesel 137

Fig. 6.9 Adsorption isotherm and recyclability of activated carbon prepared from coir fibers for
removal of CO2 from flue gas (Hauchhum et al. 2015). Reproduced with permission from Springer
Nature

6.7 Desulfurization of Diesel

Removing sulfur from diesel will reduce emissions and decrease environmental
burden. Several sorption technologies, primarily, adsorptive desulfurization has
been commonly adopted. In order to develop low cost and effective sorbent, coir
was modified using H3PO4 at 373 K for 1 h and later carbonized at 873 K for 1 h.
About 54% of coir could be converted into the carbon with a particle size of 33.3 μm.
The absorbent had surface area of 1254 m2/g and average pore diameter of 3.2 nm.
Amount of sulfur removed was dependent on the absorbent dose and contact time.
Enthalpy and entropy values suggested spontaneous absorption. An average sulfur
removal efficiency of 65% was obtained and it was also possible to regenerate the
138 6 Coir for Environmental Remediation

absorbent. Compared to activated carbon cost of USD 173 per kg, it was estimated
that the activated coir based sorbent can be prepared for about $10.7. Such low cost,
high removal efficiency, and ability to regenerate were considered to be ideal for
removal of oil, diesel, and other contaminants (Ahmed and Ahmaruzzaman 2015).

References

Ahmed MJK, Ahmaruzzaman M (2015) Adsorptive desulfurization of feed diesel using chemi-
cally impregnated coconut coir waste. Int J Environ Sci Technol 12(9):2847–2856
Aljeboree AM, Alshirifi AN, Alkaim AF (2017) Kinetics and equilibrium study for the adsorption
of textile dyes on coconut shell activated carbon. Arab J Chem 10 (S3381–S3393
Anirudhan TS, Sreekumari SS (2011) Adsorptive removal of heavy metal ions from indus-
trial effluents using activated carbon derived from waste coconut buttons. J Environ Sci
23(12):1989–1998
Anirudhan TS, Unnithan MR (2007) Aresenic (V) removal from aqueous solutions using an anion
exchanger derived from coconut coir pith and its recovery. Chemosphere 66:60–66
Anirudhan TS, Unnithan MR, Divya L, Senan P (2007) Synthesis and characterization of poly-
acrylamide grafted coconut coir pith having carboxylate functional group and adsorption abil-
ity for heavy metal ions. J Appl Polym Sci 104:3670–3681
Anirudhan TS, Divya L, Ramachandran M (2008) Mercury (II) removal from aqueous solutions
and wastewaters using a novel cation exchanger derived from coconut coir pith and its recov-
ery. J Hazard Mater 57:620–627
Anirudhan TS, Rijith S, Divya L (2009a) Preparation and application of a novel functional-
ized coconut coir pith as a recyclable adsorbent for phosphate removal. Sep Sci Technol
44:2774–2796
Anirudhan TS, Sreekumari SS, Bringle CD (2009b) Removal of phenols from water and petroleum
industry refinery effluents by activated carbon obtained from coconut coir pith. Adsorption
15:439–451
Chang N-B, Houmann C, Lin K-S, Wanielista M (2016) Fate and transport with material
response characterization of green sorption media for copper removal via desorption process.
Chemosphere 154:444–453
Chukwuma FO, Evbuomwan BO, Egwu CN (2013) Adsorption equilibrium for the removal of Fe3+
from aqueous solution using activated coconut waste. Int J Res Chem Environ 3:334
Conrad K, Hansen HCB (2007) Sorption of zinc and lead on coir. Bioresour Technol 98:88–97
Din AT, Hameed BH, Ahmad AL (2009) Batch adsorption of phenol onto physiochemical activated
coconut shell. J Hazard Mater 161:1522–1529
Etim UJ, Umoren SA, Eduok UM (2016) Coconut coir dust as a low cost adsorbent for the removal
of cationic dye from aqueous solution. J Saudi Chem Soc 20:S67–S76
Ewecharoen A, Thiravetvan P, Nakbanpote W (2008) Comparison of nickel adsorption from elec-
troplating rinse water by coir pith and modified coir pith. Chem Eng J 137:181–188
Hameed BH, Tan IAW, Ahmad AL (2008) Adsorption isotherm, kinetic modeling and mechanism
of 2, 4, 6-trichlorophenol on coconut husk-based activated carbon. Chem Eng J 144(2):235–244
Hauchhum L, Mahanta P, De Wilde J (2015) Capture of CO2 from flue gas onto coconut fibre-­
based activated carbon and zeolites in a fixed bed. Transp Porous Media 110(3):503–519
Henryk K, Jarosław C, Witold Ż (2016) Peat and coconut fiber as biofilters for chromium adsorp-
tion from contaminated wastewaters. Environ Sci Pollut Res 23(1):527–534
Hettiarachchi E, Perera R, Chandani Perera ADL, Kottegoda N (2016) Activated coconut
coir for removal of sodium and magnesium ions from saline water. Desalin Water Treat
57(47):22341–22352
References 139

Israel U, Eduok UM (2012) Biosorption of zinc from aqueous solution using coconut (Cocos
nucifera L) coir dust. Arch Appl Sci Res 4(2):809–819
Johari K, Saman N, Song ST, Heng JYY, Mat H (2014) Study of Hg (II) removal from aqueous
solution using lignocellulosic coconut fiber biosorbents: equilibrium and kinetic evaluation.
Chem Eng Commun 201(9):1198–1220
Johari K, Saman N, Song ST, Chin CS, Kong H, Mat H (2016) Adsorption enhancement of ele-
mental mercury by various surface modified coconut husk as eco-friendly low-cost adsorbents.
Int Biodeterior Biodegrad 109:45–52
Kavitha D, Namasivayam C (2007a) Experimental and kinetic studies on methylene blue adsorp-
tion by coir pith carbon. Bioresour Technol 98:14–21
Kavitha D, Namasivayam C (2007b) Recycling coir pith, an agricultural solid waste, for the
removal of procion orange from waste water. Dyes Pigments 74:237–248
Khan MR, Ray M, Guha AK (2011) Mechanistic studies on the binding of acid yellow 99 on coir
pith. Bioresour Technol 102:2394–2399
Krishnan KA, Haridas A (2008) Removal of phosphate from aqueous solutions and sewerage using
natural and surface modified coir pith. J Hazard Mater 152:527–535
Kumar P, Sudha S, Chand S, Srivastava VC (2010) Phosphate removal from aqueous solution
using coir-pith activated carbon. Sep Sci Technol 45:1463–1470
Kumar P, Rao R, Chand S, Kumar S, Wasewar KL, Yoo CK (2013) Adsorption of lead from aque-
ous solution onto coir-pith activated carbon. Desalin Water Treat 51(13–15):2529–2535
Lim HS, Lim W, Hu JY, Ziegler A, Ong SL (2015) Comparison of filter media materials for heavy
metal removal from urban stormwater runoff using biofiltration systems. J Environ Manag
147:24–33
Limbikai SS, Deshpande NA, Kulkarni RM, Khan AAP, Khan A (2016) Kinetics and adsorption
studies on the removal of levofloxacin using coconut coir charcoal impregnated with Al2O3
nanoparticles. Desalin Water Treat 57(50):23918–23926
Macedo JS, Junior NBZ, Almeida E, Vieira EFS, Cestari AR, Gimenez IF, Carreno NLCC, Barreto
LS (2006) Kinetic and calorimetric study of the adsorption of dyes on mesoporous activated
carbon prepared from coconut coir dust. J Colloid Interface Sci 298:515–522
Namasivayam C, Kavitha D (2002) Removal of Congo Red from water by adsorption onto acti-
vated carbon prepared from coir pith, an agricultural solid waste. Dyes Pigments 54(1):47–58
Namasivayam C, Kavitha D (2006) IR, XRD and SEM studies on the mechanism of adsorption of
dyes and phenols by coir pith carbon from aqueous phase. Macrochem J 83:43–48
Namasivayam C, Sangeetha D (2004) Equilibrium and kinetic studies of adsorption of phosphate
onto ZnCl2 activated coir pith carbon. J Colloid Interface Sci 280(2):359–365
Namasivayam C, Sangeetha D (2006) Recycling of agricultural solid waste, coir pith: removal of
anions, heavy metals, organics and dyes from water by adsorption onto ZnCl2 activated coir
pith carbon. J Hazard Mater 135(1–3):449–452
Namasivayam C, Dinesh Kumar M, Selvi K, Ashruffunissa Begum R, Vanathi T, Yamuna RT
(2001) Wastecoir pith—a potential biomass for the treatment of dyeing wastewaters. Biomass
Bioenergy 21(6):477–483
Ong S-A, Ho L-N, Wong Y-S, Zainuddin A (2013) Adsorption behavior of cationic and anionic
dyes onto acid treated coconut coir. Sep Sci Technol 48(14):2125–2131
Parab H, Sudersanan M (2010) Engineering a lignocellulosic biosorbent—coir pith for removal
of cesium from aqueous solutions: equilibrium and kinetic studies. Water Res 44(3):854–860
Parab H, Joshi S, Shenoy N, Lali A, Sarma US, Sudersanan M (2006) Determination of kinetic
and equilibrium parameters of the batch adsorption of Co(II), Cr (III) and Ni(II) onto coir pith.
Process Biochem 41:609–611
Parab H, Joshi S, Shenoy N, Lali A, Sarma US, Sundersanan M (2008) Esterified coir pith as an
adsorbent for the removal of Co(II) from aqueous solution. Bioresour Technol 99:2083–2086
Parab H, Joshi S, Sudersanan M, Shenoy N, Lali A, Sarma U (2010) Removal and recovery of
cobalt from aqueous solutions by adsorption using low cost lignocellulosic biomass-coir pith.
J Environ Sci Health A 45:603–611
140 6 Coir for Environmental Remediation

Parab H, Devi PSR, Shenoy N, Kumar SD, Bhardwaj YK, Reddy AVR (2016) Gamma irradiation
stability studies of coir pith: a lignocellulosic biosorbent for strontium. J Radioanal Nucl Chem
308(1):323–328
Ramesh ST, Gandhimathi R, Elavarasi TE, Thamizh IR, Sowmya K, Nidheesh P (2014)
Comparison of methylene blue adsorption from aqueous solution using spent tea dust and raw
coir pith. Global NEST J 16(1):146–159
Ray SK, Saha P, Nur HP, Saha D, Hoque AI, Saha S (2014) Study on the preparation of polymer-­
grafted coir fibre based adsorbent and its application to remove a reactive dye from aqueous
solution. Bangladesh J Sci Indust Res 48(4):271–280
Santhy K, Selvapathy P (2004) Removal of heavy metals from waste water by adsorption on coir
pith activated carbon. Sep Sci Technol 39(14):3331–3351
Santhy K, Selvapathy P (2006) Removal of reactive dyes from wastewater by adsorption on coir
pith activated carbon. Bioresour Technol 97(11):1329–1336
Sharma YC, Uma, Upadhyay SN (2009) Removal of cationic dye from wastewaters by adsorption
on activated carbon developed from coconut coir. Energy Fuels 23:2983–2988
Shen Y, Wang S, Huang S, Tzou Y, Huang J (2010) Biosorption of Cr(VI) by coconut coir: spec-
troscopic investigation on the reaction mechanism of Cr(VI) with lignocellulosic material.
J Hazard Mater 179:160–165
Shukla PM, Shukla SR (2013) Biosorption of Cu(II), Pb(II), Ni(II) and Fe(II) on alkali treated coir
fibers. Sep Sci Technol 48:421–428
Shukla SR, Pai RS, Shendarkar AD (2006) Adsorption of Ni(II), Zn(II) and Fe (II) on modified coir
fibres. Sep Pur Technol 47(3):141–147
Shukla SR, Gaikar G, Pai RS, Suryavanshi US (2009) Batch and column adsorption of Cu(II) on
unmodified and oxidized coir. Sep Sci Technol 44:40–62
Singh S, Gondhalekar S, Shukla SR (2017) Continuous column study of Cu(II) and Pb(II) ions on
alkali treated coir. Desalin Water Treat 57(37):17440–17453
Suksabye P, Thiravetyan P, Nakbanpote W, Chayabutra S (2007) Chromium removal from electro-
plating wastewater by coir pith. J Hazard Mater 141(3):637–644
Sureshkumar MV, Namasivayam C (2008) Adsorption behavior of direct red 12B and rhodamine
B from water onto surfactant-modified coconut coir pith. Colloids Surf A Physicochem Eng
Asp 317(1–3):277–283
Suryavanshi US, Shukla SR (2009) Adsorption of Ga(III) on oxidized coir. Ind Eng Chem Res
48:870–876
Tan IAW, Ahmad AL, Hameed BH (2008) Adsorption of basic dye on high-surface-area activated
carbon prepared from coconut husk: equilibrium, kinetic and thermodynamic studies. J Hazard
Mater 154(1–3):337–346
Thitame PV, Shukla SR (2016) Adsorptive removal of reactive dyes from aqueous solution
using activated carbon synthesized from waste biomass materials. Int J Environ Sci Technol
13(2):561–570
Unnithan MR, Vinod VP, Anirudhan TS (2004) Synthesis, characterization as a chromium (VI)
adsorbent of amine modified polyacrylamide grafted coconut coir pith. Ind Eng Chem Res
43:2247–2255
Zaman HU, Al-Mamun M, Khan MA, Khan RA (2013) Preparation of adsorbent for extracting
heavy metal ions by coir yarn. J Thermoplast Compos Mater 26(6):820–830
Zhu Y, Kolar P (2014) Adsorptive removal of p-cresol using coconut shell activated char. J Environ
Chem Eng 2:2050–2058
Chapter 7
Composites from Coir Fibers

7.1 Coir Fibers as Reinforcement for Synthetic Polymers

7.1.1 Polyester Composites Reinforced with Coir Fibers

Sound absorbing acoustic panels with different perforations were developed using
coir as reinforcement, kenaf as filler, and polyester resin as matrix (Thye et al. 2012).
It was found that addition of coir improved sound absorption but the level of absorp-
tion varied depending on the frequency (Table 7.1). Sound absorption coefficient
tests done according to ISO 10534-2 suggested that the most optimum absorption
was obtained when 10% fibers and 10% perforations were available. Instead of
using virgin polyester, discarded plastic bottles were reinforced with untreated and
modified coir fibers. The PET from the plastic bottles was converted into unsatu-
rated polyester resin through glycolysis. In this process, the PET was treated with
ethylene glycol along with 0.5% zinc acetate as catalyst. The mixture was refluxed
in nitrogen atmosphere at 190 °C for 8 h (Abdullah and Ahmad 2012). The unsatu-
rated polyester resin was obtained by subjecting the glycolyzed product to polyes-
terification reaction. Later, the resin was combined with styrene monomer to obtain
the unsaturated polyester resin (UPR) containing about 40% styrene monomer.
Similarly, the coir fibers were also subject to chemical treatments using alkali or
silane solution. Composites were formed using the hand lay-up technique with an
initiator and accelerator. Curing of the composites was done at room temperature for
24 h and later post curing was done by treating in an oven at 60 °C for 6 h. Tensile
and impact resistance were both higher for the composites containing coir fibers that
were silanated after alkali treatment (Fig. 7.1) (Abdullah and Ahmad 2012).
Various pretreatments have been done to improve the properties of coir and coir-­
based composites. Coir fibers were treated with 1, 5, or 10% NaOH solution for
17 h at room temperature. Treated fibers were immersed in 5% PVA solution for 2 h
and later dried at 40 °C for 2 h for the PVA to be infused and form the composite
fiber (Guo et al. 2016). Considerable increase in mechanical properties of the coir

© Springer Nature Switzerland AG 2019 141


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_7
142 7 Composites from Coir Fibers

Table 7.1 Acoustic properties of composite panels containing various levels of coir fibers and
polyester resin with kenaf fibers as fillers (Thye et al. 2012)
Panel composition Frequency, kHz
(coir:polyester) Perforation area, % 1 2 3 4 5 6
10:90 0 0.001 0.001 0.310 0.002 0.190 0.660
20:80 0 0.001 0.015 0.390 0.060 0.040 0.160
30:70 0 0.020 0.001 0.050 0.260 0.210 0.010
10:90 10 0.001 0.001 0.450 0.900 0.810 0.690
20:80 10 0.000 0.001 0.300 0.650 0.260 0.640
30:70 10 0.000 0.001 0.400 0.020 0.300 0.100
10:90 20 0.000 0.001 0.300 0.650 0.260 0.640
20:80 20 0.000 0.001 0.140 0.050 0.160 0.010
30:70 20 0.000 0.001 0.290 0.010 0.190 0.700
10:90 30 0.000 0.001 0.020 0.690 0.680 0.800
20:80 30 0.000 0.001 0.310 0.690 0.380 0.490
30:70 30 0.000 0.001 0.320 0.620 0.670 0.240

Fig. 7.1 Effect of various chemical treatments on the tensile and impact properties of coir fiber
reinforced UPR composites (Abdullah and Ahmad 2012). Reproduced with permission through
Creative Commons Attribution License

fibers after treatment and combining with PVA was observed (Fig. 7.2). As seen
from the figure, the tensile strength and toughness increase between 100 and 200%
when combined with PVA. Similarly, elongation of the fibers also shows consider-
able increase. Compared to most other natural cellulose fibers, the modified coir
fibers showed substantially higher specific strength and elongation. Since coir fibers
have low elongation, modifying it with PVA could be a viable approach to develop
composite fibers for various applications (Guo et al. 2016). In a similar approach,
coir fibers (10 mm length) were treated with 5% NaOH and combined with polyes-
ter fibers from 10 to 30%. The fibers and resin were compression molded at 85 °C
to form composites of 2–6 mm in thickness (Prasad et al. 2018). Increase in impact
strength from 0.4 to 1.6 N m was obtained depending on the fiber and thickness of
composites prepared (Prasad et al. 2018).
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 143

Fig. 7.2 Changes in the properties of PVA reinforced with treated and untreated coir fibers (Guo
et al. 2016). Stress-strain curve (a); effect of various treatments on strength (b), modulus (c) and
toughness (d). Reproduced with permission from American Chemical Society

Table 7.2 Properties of the neat epoxy and UPE matrix and composites obtained after reinforcing
with 30% alkali-treated coir fibers (Santos et al. 2018)
Impact
Tensile Tensile Flexural Flexural strength,
Sample modulus, GPa strength, MPa modulus, GPa strength, MPa kJ/m2
Neat epoxy 2.32 47.2 2.26 69.26 6.33
Treated—24 h 2.56 17.90 2.59 40.09 6.07
Treated—96 h 2.59 20.91 2.72 43.05 5.97
Treated—168 h 2.60 18.68 2.67 40.07 6.82
Neat UPE 2.41 12.83 2.40 55.9 5.29
Treated—24 h 2.96 11.68 2.68 29.25 17.64
Treated—96 h 3.12 14.40 2.82 30.42 17.99
Treated—168 h 3.12 13.32 2.81 24.51 16.93

Coir fibers treated with sodium carbonate (10%) at room temperature for
12–168 h were used as reinforcement for polyester and epoxy resins. Composites
having dimensions of 310 × 310 × 3 mm were prepared using the hand lay-up
approach. Alkali treatment increased the density but decreased thermal stability of
the fibers. Tensile, flexural and impact strength showed different variations depend-
ing on the extent of fiber treatment and type of matrix used (Table 7.2). Although the
144 7 Composites from Coir Fibers

properties of the polyester-coir composites were lower, (except for the impact
strength) than that of the neat matrix, it was suggested that treating with sodium
carbonate would avoid the corrosion, environmental and other hazards associated
with other common chemicals used for modifying the surface of the fibers (Santos
et al. 2018). Lower composite properties were also found when polypropylene was
reinforced with coir fibers (Staffa et al. 2017). However, addition of a compatibil-
izer (maleic anhydride) was found to provide improved adhesion between the fibers
and matrix leading to higher mechanical properties by about 20–30%. Also, addi-
tion of photostabilizing agents such as HALS also decreased the mechanical proper-
ties whereas presence of hydroxybenzophenone UV absorbers did not have any
adverse effect (Staffa et al. 2017). Hence, appropriate compatibilizers and additives
have to be chosen to obtain desired composite properties. Coir fibers also provided
higher thermal stability to the composites due to their higher thermal degradation
temperature (Abitha and Rane 2015).
In a different approach, coir ropes or coir fabric were used as reinforcement for
unsaturated polyester resin (Yao et al. 2014). The coir fabric had a weight per unit
area of 1220 g/m2. To improve adhesion with the polyester resin, the fabrics were
treated with NaOH (2–8%) for 24–120 h. The polyester resin was used as matrix
with methyl ethyl ketone peroxide as catalyst and cobalt octoate as accelerator
(Jayabal et al. 2013). Composites were formed by compression molding at 500 N
for 1 h. Treating with NaOH and immersing in water changed the properties of the
composites (Table 7.3). Major changes were observed in the flexural strength, par-
ticularly after treating in water for 96 h. When coir ropes were used, the average
coarseness of each coir rope was 0.3 cm and average mass was 7 g/m2. The coir
ropes were immersed in the matrix for 5 min and made into composites using 2%
hardener, 1% accelerator, and 1% catalyst and maintaining the coir fiber ratio
between 10 and 30%. Composites were formed in a semi-open mold and cured at
room temperature. Compared to neat UPE, tensile strength and modulus of the com-
posites decreased due to the addition of the coir ropes. Based on the experimental
results obtained, a mixed law model was developed to accurately predict the proper-
ties of the composites using the properties of the raw materials (Yao et al. 2014).

7.1.2 Polypropylene Composites Reinforced with Coir Fibers

To improve the compatibility between fibers and polypropylene used as matrix, coir
fibers were first treated with sodium periodate and later with urea. Cut and washed
coir fibers were immersed in sodium periodate to form cellulose dialdehyde. Later,
the treated fibers were immersed in 20% urea solution (Haque and Islam 2013).
Modified fibers were combined with polypropylene granules in 10–35 wt% and
passed through an extruder at 165 °C. Composites formed were shaped and used for
determining the mechanical properties and moisture sorption (sample size of
39 mm × 10 mm × 4.1 mm). Except for the tensile strength, the flexural strength and
modulus, impact strength and hardness values increased with increasing fiber
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 145

Table 7.3 Comparison of the properties of the polyester-coir fabric reinforced composites after
treating with alkali and immersion in water for various periods of time (Jayabal et al. 2013)
Flexural strength,
Sample Tensile strength, MPa MPa Impact strength, kJ/m2
Untreated fibers 19.9 ± 0.5 31.3 ± 5.6 49.9 ± 4.2
2% NaOH treated
–24 18.56 24.19 48.02
–48 19.58 32.22 53.78
–72 21.94 21.86 56.86
–96 22.83 48.08 54.25
−120 17.98 29.24 49.23
5% NaOH treated
–24 19.34 26.77 55.24
–48 21.55 24.42 58.45
–72 23.56 39.85 54.97
–96 18.34 24.86 52.41
−120 17.19 24.31 50.05
8% NaOH treated
–24 20.92 18.72 58.75
–48 23.17 44.45 54.23
–72 21.42 21.91 52.78
–96 19.76 21.19 49.43
−120 18.59 20.42 48.26
Reproduced with permission from Springer Nature

content. Treated fibers showed better fiber–matrix interfacial interaction and bond-
ing, leading to improved properties including lower water uptake percentage (Haque
and Islam 2013). In another approach, coir fibers were oxidized and also imidiated
(Fig. 7.3) to improve adhesion with polypropylene and obtain composites with
enhanced properties (Haque et al. 2012a). Modified fibers were combined with
polypropylene in various ratios and extruded at 165 °C to form granules. These
granules were later compression molded into composites at a temperature of
165 °C. Tensile strength decreased but modulus, impact strength, and hardness
increased continually with increase in percentage of coir. Considerable decrease in
water sorption was seen after the chemical treatment. Similarly, the fiber modifica-
tions done decreased microvoids and increased interfacial adhesion between the
fibers and the matrix (Haque et al. 2012a).
Coir fibers treated with 2, 4, 8, or 12% NaOH solutions for 2 h at 100 °C were
compression molded into composites using polypropylene as the resin and maleic
anhydride as the compatibilizer. Alkali treatment at low concentrations increased
surface roughness and high concentrations resulted in extensive delignification.
Presence of compatibilizer and high alkali treatment were considered to be neces-
sary to produce composites with good properties (Bettini et al. 2015). In a similar
study, coir fibers were treated with alkali and also functionalized using a silane
146 7 Composites from Coir Fibers

CH2CH
CH2CH H O O
H O O NaIO4 H n
H n
O
OH H H
O H O O
H OH

NH2

CH2O-Na NaOH
H O O
H n
OH
O H

N N

ONa ONa

Fig. 7.3 Schematic representation of the chemical modifications between coir and NaIO4 and
p-aminophenol (Haque et al. 2012a). Reproduced with permission from Springer Nature

(tetramethoxyorthosilicate) to increase compatibility between coir fibers and matrix.


Compared to the neat PP strength of 26 MPa, the coir fiber reinforced composites
had tensile strength of 34 MPa which further increased to 42–48 MPa when silane-­
treated coir fibers were used (Zaman and Beg 2014). Similar phenomenon was also
observed for tensile modulus and impact strength, where values increased from
478 MPa upto 2100 MPa and 7 kJ/m2 to 22 kJ/m2, respectively. It was suggested that
silane treatment would be preferable when using coir as reinforcement for polypro-
pylene. Similar silane treatment was also used to improve the adhesion between coir
and polypropylene and curaua fibers and polypropylene (de Oliveira and Marques
2014). Unlike coir that contains high levels of lignin, curaua fibers are composed of
71–74% cellulose, 9.9% hemicellulose, and 11% lignin. Both the fibers were either
treated with 5% NaOH for 1 h or 63.5% sulfuric acid for 2 h at 45 °C. After the
initial treatment of the fibers, they were immersed in vinyltrimethoxysilane solution
for 72 h for silanation to occur. Maleic acid grafted polypropylene was combined
with 20% grafted fibers and compounded in a twin screw extruder to form the com-
posites. Increase in the intensity and broadening of the peak between 3200 and
3500 cm−1 and 1600 and 1410 cm−1 confirmed changes in the chemical structure of
the fibers due to the alkali/acid treatment. The peak at 1740 cm−1 corresponding to
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 147

hemicellulose also reduced due to the removal of hemicellulose. New peaks at 1030
and 756 cm−1 (siloxane group) and increase in the shoulders at 1200 and 965 cm−1
due to the Si-O-C linkages confirmed silane modifications. Morphological studies
showed that coir fiber treated with acid had better binding with the matrix.
Comparatively, acid-treated curaua fibers had poor interactions with the matrix even
after the addition of a compatibilizing agent (de Oliveira and Marques 2014).
Polypropylene with a melt flow index of 10 g/10 min, specific gravity of 0.90–
0.91, and melting temperature between 165 and 171 °C was combined with coir
fibers having length of 1–2 cm and diameter of about 250–300 μm (Haque et al.
2012b). The coir fibers were immersed with alkali and diazonium salt at different
pHs. The modified fibers were combined with PP granules and passed through an
extruder at 135 ± 5 °C. The composite granules were compression molded into ten-
sile and flexural test specimens at temperature of 165 °C. Tensile strength of the
composites decreased continually, whereas flexural modulus continually increased
with increasing fiber content. Fibers treated at higher pH (10.5) provided better
tensile strength than those treated at lower pHs. However, flexural strength of the
composites did not show any appreciable change above a fiber content of 10%. It
was supposed that alkali treatment of fibers increased properties of the composites
and 30% fiber reinforced composites had the optimum composite properties (Haque
et al. 2012b).

7.1.3 Polyethylene Reinforced Coir Composites

Coir fibers were subject to series of chemical treatments and combined with high-­
density polyethylene having a density of 0.959 g/cm3 and melting temperature
between 180 and 220 °C (Arrakhiz et al. 2012). The coir fibers were cut and ground
into 2–3 mm length and treated with 1.6 mol/L for 48 h and later neutralized using
acetic acid. In the second modification, coir fibers were treated with dodecane bro-
mide and NaOH for 12 h. Another fiber modification was to treat the fibers with
0.5 wt% of silane coupling agent 3-(trimethoxysilyl)propylamine for 45 min at
65 °C. Modified coir fibers (20%) were added into HDPE and compounded at
200 °C at 60 rpm for 3 min and the pellets obtained were compression molded at
190 °C for 30 s according to ISO 527-1 standards (Arrakhiz et al. 2012). Addition
of fibers into the neat HDPE matrix increased both tensile strength and modulus
(Fig. 7.4). Similarly, the fibers treated with bromide had 120% increase in modulus.
However, alkali treatment of the fibers before grafting the ether chains was found to
provide better properties. Both polypropylene and polyethylene were reinforced
with chemically modified coir fibers and made into composites (Mir 2013). Coir
fibers were modified by treating with chromium sulfate (Cr2(SO4)3.12(H2O)),
sodium carbonate (NaHCO3), and HCl at different pHs (8–10). Treated fibers were
combined with poly(ethylene) or polypropylene in 0–20% weight ratios and com-
pression molded at 170 °C for 30 min using 30 kN pressure. Increasing ratio of
fibers above 10% decreased the tensile strength but the tensile modulus increased
148 7 Composites from Coir Fibers

1600 30
Young’s modulus
Tensile strength
Young’s modulus (MPa) 1400 28

Tensile strength (MPa)


1200 26

1000 24

800 22

600 20
PE

PE

PE

PE

PE
D

D
tH

r/H

r/H

r/H

r/H
ea

oi

oi

oi

oi
C

C
N

aw

LK

ne

12
C
la
R

Si

Fig. 7.4 The tensile strength and modulus of HDPE composites reinforced with coir fibers modi-
fied using various chemical treatments (Arrakhiz et al. 2012). Reproduced with permission from
Elsevier

from 1.5 to 2.7 GPa with increasing fiber content. Flexural and impact strength also
increased substantially. Moisture sorption of the composites was lower for the mod-
ified fibers. Between the two polymers used as matrix, polypropylene composites
provided better properties than polyethylene composites (Mir 2013). In another
study, in addition to treating fibers with alkali (2% NaOH at 25 °C for 15 min),
fibers were immersed in 1% maleated polyethylene dissolved in xylene at 10% fiber
ratio. After treatment, the fibers were combined with a linear medium density poly-
ethylene (LMDPE) resin having melt flow index of 5 g/10 min and density of 0.93 g/
cm3 and made into composites using the rotomolding approach (Cisneros-López
et al. 2017). Hardness and impact strength of the composites decreased with increas-
ing fiber content. Also, there was no major difference in properties of the compos-
ites containing the treated and untreated fibers.
Instead of the usual approach of modifying the coir fibers, polyethylene was
treated with plasma and made into composites. Before using as the reinforcement,
coir fibers were treated with 5% NaOH for 2 h and also later subject to oxidation
using hydrogen peroxide at 85 °C for 2 h. Two methods of preparation (melt mixing,
mechanical mixing, and hot pressing) were used to develop the composites. Plasma
treatment decreased the contact angle and increased surface energy from 21 to 32 J,
and interfacial energy decreased from 23 to 16 mJ/m2 for polyethylene. Due to these
changes, considerable effects on the properties of the composites were observed.
The method of preparation of the composites also affected the composite properties
(Table 7.4). Tensile strength of the composites decreased as the proportion of the
fibers increased irrespective of the treatments (Fig. 7.5) (Sari et al. 2015).
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 149

Table 7.4 Properties of untreated (PE) and plasma treated polyethylene (PPE) composites
containing 10% coir fibers prepared using two methods (Sari et al. 2015)
Melt mixed Hand mixed
PE PPE PE PPE
Tensile strength, MPa 12.8 ± 1.0 11.9 ± 0.9 10.5 ± 1.3 13.7 ± 1.8
Modulus, MPa 765 ± 56 689 ± 47 585 ± 85 1045 ± 59
Elongation, % 13.4 ± 0.3 9.3 ± 0.4 13.2 ± 0.3 14.3 ± 0.3

Fig. 7.5 Tensile strength of the composites decreased with increasing ratio of the fibers (Sari et al.
2015)

7.1.4 Epoxy Reinforced with Coir Fibers

Coir fibers (30%) were combined with the epoxy resin (diglycidyl ether bisphenol-
­A) and hardener (triethylenetetramine) as matrix. Length of the coir fibers in the
matrix was varied from 2 to 43 mm (Luz et al. 2018). Tensile strength of the com-
posites was 28.7 MPa, elastic modulus was 3.2 GPa, and impact strength was
111 J/m. However, there was poor interfacial resistance due to which debonding
between fiber and matrix was also noticed (Luz et al. 2018). Using a bicomponent
epoxy resin (RL 3135LV) as the matrix, coir fibers treated with 1% sodium
150 7 Composites from Coir Fibers

Fig. 7.6 Effect of fiber length and content on the tensile strength and modulus of coir fiber rein-
forced epoxy composites (Das and Biswas 2016a, b). Reproduced with permission from Creative
Commons Attribution 3.0 license

hydroxide for 1 h were made into mats using 30% reinforcement. To form the com-
posites, the resin and fibers were placed in a resin transfer mold and cured at 60 °C
for 7 h (de Carvalho Benini et al. 2017). Treating with alkali not only removed non-
cellulosic substances but also increased surface roughness which improved the
adhesion between matrix and fibers. Considerably good binding between the epoxy
resin and coir fibers can be observed from Fig. 7.6. Such good adhesion also resulted
in higher thermal behavior of the composites. Untreated fiber showed degradation
starting at 228 °C compared to 335 °C for the alkali-treated fibers. Overall weight
loss of the composites containing treated fibers was also lower at 58% compared to
77% for the untreated fibers. However, there was no difference in the flexural modu-
lus but flexural strength had increased to 51 MPa compared to 48 MPa before treat-
ing the fibers. It has also been shown that the length and content of fibers in the
epoxy matrix affect the mechanical properties and water sorption of the composites
(Das and Biswas 2016a, b). Although the void content in the composites increased
with increasing fiber length and content, there was gradual increase in both the ten-
sile strength and modulus (Fig. 7.7). Similar increase in impact strength to 13 kJ/m2
from 10 kJ/m2 was observed. Since coir fibers are hydrophilic, increase in moisture
sorption from 2 to 13% had occurred with the addition of 20% fibers (Das and
Biswas 2016a, b).
White and brown coir fibers were separately treated with 6% NaOH for 12 h and
combined with an epoxy matrix that had epoxy index of 0.51–0.56 mol/1000 g,
viscosity of 500–900 mPa s at 25 °C, and density of 1.12–1.16 g/cm3 (Valášek et al.
2018). The matrix was also hardened using a cycloaliphatic polyamine. Although
the composition and surface morphology of the white and brown fibers was similar,
there were considerable difference in the porosity and other properties (Table 7.5).
In terms of the mechanical properties of the composites, the brown coir fibers pro-
vided higher tensile strength than white fibers. In addition, alkali treatment increased
the strength and modulus of the brown fibers by 47 and 74%, respectively, which
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 151

Fig. 7.7 Scanning electron image shows good interfacial adhesion between the matrix and the
fiber (de Carvalho Benini et al. 2017). Reproduced with permission from John Wiley and Sons

Table 7.5 Properties of White Brown


white and brown coir fibers Property fibers fibers
used to develop the
Total intrusion volume, mL/g 0.6578 0.7645
composites (Valášek et al.
2018) Total pore area, m2/g 22.62 14.37
Median pore diameter, volume, μm 555.4 751
Median pore diameter, area, nm 13.8 6.6
Average pore diameter (4 V/A), μm 0.1163 0.2128
Bulk density, g/mL 0.709 0.664
Apparent density, g/mL 1.33 1.35
Porosity, % 46.64 50.8
Stem volume used, % 36 38
Reproduced with permission from Elsevier

also lead to an increase in about 3.7% in the composite tensile strength. Difference
in the properties of the composites was suggested to be due to the porosity and pore
size variations between the two fibers (Valášek et al. 2018).
Various lengths and proportions of short coir fibers were combined with epoxy
to form composites. A filler (Al2O3) having particle size of 80–100 μm and density
152 7 Composites from Coir Fibers

Fig. 7.8 Changes in the tensile strength and flexural strength of coir fiber reinforced epoxy com-
posites at different lengths of the fibers and proportion of fibers in the composites (Das and Biswas
2016a, b). Reproduced with permission from Creative Commons Attribution 3.0 license

of 3.69 g/cm3 was also used at 10 wt%. Composites were fabricated by hand casting
and later cut into desired shapes for mechanical testing (Das and Biswas 2016a, b).
Increasing fiber content up to 15% improved strength but decreased at 20% fiber
content. However, modulus showed a different trend and increased continually with
the proportion of the fibers (Das and Biswas 2016a, b). Optimum composite condi-
tions were obtained using 15% fibers and 12 mm length of the fibers (Fig. 7.8).

7.1.5 Phenol Formaldehyde Based Coir Fiber Composites

Fiber boards made from coir and phenol formaldehyde resin were coated with flame
retardant agents to make the composites suitable for interior and other building
applications (Rejeesh and Saju 2018). To impart flame resistancy, fibers were treated
in an aqueous solution containing a mixture of 40% boric acid and 60% borax at
100 °C for 30 min. Treated boards showed increased mechanical properties and
thermal resistance but lower thermal conductivity (Table 7.6). It was suggested that
the compressive strength and fire resistant values of the treated coir fiber boards
were suitable for interior applications (Rejeesh and Saju 2018). Up to 70% coir
fibers were used as reinforcement for phenol formaldehyde resin (Adnan et al.
2014). The coir fibers used had tensile strength of 32 MPa, modulus of 1.5 GPa, and
porosity of 96%. Modulus of rupture, impact strength, and hardness increased with
increasing fiber content, and optimum fiber content was found to be 70%.

7.1.6 Rubber Composites Containing Coir Fibers

Coir fibers were carbonized and combined with natural rubber and made into com-
posites and properties compared with commercially available carbon black and with
uncarbonized coir fibers (Aguele and Madufor 2012; Aguele et al. 2014). To
7.1 Coir Fibers as Reinforcement for Synthetic Polymers 153

Table 7.6 Changes in the Treated Untreated


properties of phenol Properties sample sample
formaldehyde composites
Density, kg/m3 846 650
made using untreated and
flame resistant chemical Moisture content, % 7 5
treated coir fibers (Rejeesh Tensile strength, MPa 1.26 0.84
and Saju 2018) Compressive strength, MPa 11.6 15.7
Flexural strength, MPa 33.9 17
Flexural modulus, MPa 3.03 2.16
Tensile modulus, MPa 4846 2469
Thermal conductivity, W/m-K 0.0938 0.183
Thermal resistance, m2 K/W 0.117 0.060
Open access

Table 7.7 Composition of Ingredient Content, %


the coir carbon/carbon black
Natural rubber 100
reinforced composites
(Aguele and Madufor 2012) Zinc oxide 5.0
Stearic acid 2.5
TMQ 1.0
Processing oil 2.0
Sulfur 2.5
TMTD 1.0
MBT 1.0
Filler (coir carbon/carbon black) 0–50
Open access publication

carbonize the coir fibers, the fibers were treated in an oven at 125 °C and later
heated to 800 °C. Natural rubber was masticated and mixed with additives such as
coir carbon fillers or with commercially available carbon black. The materials used
to develop the coir or carbon filled rubber composites are given in Table 7.7.
Properties (Table 7.8) of the composites were directly dependent on the amount of
fibers used as reinforcement.
Hardness, tensile strength, and modulus increased with increasing ratio of the
reinforcement. Carbon black reinforced composites had better mechanical proper-
ties than coir carbon containing composites. Similarly, composites reinforced with
uncarbonized coir had the highest mechanical properties compared to carbonized
coir. Although good improvement in properties was achieved, it was suggested that
lower carbonization temperature should be used and that the carbon has to be modi-
fied or combined with a compatibilizer or coupling agent to further improve the
properties (Aguele et al. 2014). In another study, coir fibers were cut into 10 mm
length and treated with 5% NaOH for 48 h. These fibers were combined with vulca-
nized rubber and milled into composite sheets using 10 g zinc, 10 g mercaptoben-
154 7 Composites from Coir Fibers

Table 7.8 Properties of composites reinforced with coir fibers (CC) and carbonized coir (CB)
(Aguele and Madufor 2012)
Tensile strength,
Flex fatigue Modulus MPa Elongation
Filler loading CC CB CC CB CC CB CC CB
0 21,000 21,000 1.00 1.00 3 3 880 880
10 15,268 10,878 1.00 1.80 4 17 690 600
20 11,750 10,214 1.05 2.40 4.3 21 630 520
30 10,600 6765 0.90 2.50 4.7 26 580 520
40 6540 3821 1.20 3.80 5.0 30 565 290
50 4713 2326 1.25 5.70 5.2 38 500 250

zodithiazole, 5 g tetraethylthiuram, and 5 g sulfur. Ability of the natural rubber to


withstand hot and cold treatments for up to 8 h before and after adding the coir
fibers was studied. An increase in load by 121% and tensile strength by 90% was
observed due to the addition of the reinforcement (Patil and Sharma 2018). Instead
of using natural rubber, maleated rubber was produced by grafting maleic anhydride
at three different levels. Similar to the rubber, coir fibers were also treated with 5%
NaOH (5%) for 1 h. Later, the rubber and coir fibers were combined in a twin screw
extruder and made into composites (Ujianto et al. 2017). Better binding and strong
matrix fiber interaction were achieved due to the chemical modifications of coir
fibers.

7.1.7 UV Treatment of Coir Fibers

Coir fibers were subject to a series of treatments such as UV irradiation and mercer-
ization to improve the properties of ethylene glycol dimethacrylate (EGDMA)
based composites. Initially, coir fibers were washed with acetone and then irradiated
under ultraviolet light for various lengths of time. Similarly, mercerization was done
by treating the fibers in 5–50% alkali solution for 3–20 min at 0–100 °C (Roy et al.
2014). Composites were formed by immersing the treated or untreated coir fibers in
the monomer and reirradiating at 254–313 nm with a power of 2 kW at 50 Ȧ. A
general trend of increasing tensile properties with increase in UV radiation and
alkali concentration was observed. However, a sharp decrease in strength and elon-
gation was observed above a UV irradiation of 125 passes at all mercerization con-
ditions, particularly above 10% alkali concentration. Both mercerization and UV
treatment were suggested to produce composites with good performance properties
(Roy et al. 2014).
7.2 Biocomposites from Coir Fibers 155

7.2 Biocomposites from Coir Fibers

7.2.1 Poly(Lactic) Acid Based Coir Biocomposites

PLA is one of the most commonly used biodegradable synthetic polymer for com-
posite applications. Biocomposites were developed by combining coir fibers
(1.15 g/cm3) with polylactic acid (PLA) having a density of 1.25 g/cm3. The coir
fibers had a diameter of 150–250 μm. The fibers were cut into length between 6 and
8 mm and treated with 2% NaOH solution for 1 h at 70 °C. The fibers (5 to 30%)
and PLA were heated in an oven at 180 °C for 15 min and later compressed using a
pressure of 1.5 MPa for 15 min (Dong et al. 2014). The tensile properties, particu-
larly, the elongation of the composites was dependent on the fiber content and
NaOH treatment (Fig. 7.9). However, thermal data (Table 7.9) of the composites
indicated that the amount of fibers did not have a major impact on the melting tem-
perature but the melting enthalpy changed from 6.4 to 16.7 J/g. Based on the ther-
mal degradation curves, the percent crystallinity of the composites varied between
6.9 and 20.6%. Composites containing 10% alkali-treated coir fibers had higher
biodegradation level of 35% (Dong et al. 2014). In another study, coir/sisal fibers
were combined with poly(lactic) acid as the resin and made into composites (Duan
et al. 2018). PLA used in the study had a density of 1.24 g/cm3, crystallization peak
temperature of 145–160 °C, tensile strength of 62 MPa, and tensile modulus of
1.27 MPa. Sisal leaves were immersed in distilled water for 6 days and later in
NaOH for 2 days and dried. Further, the sisal and coir fibers were cut into 10 mm
length and treated with 10% NaOH for 4 h and washed and dried. Composites were
manufactured by compounding and milling different ratios of the two fibers. The
composite pellets were molded into sheets in a compression mold at 185 °C, 10 MPa
pressure for 5–6 min. Tensile and flexural properties of the composites reveal that
increasing the concentration of the fibers in the composites decreased both flexural
and tensile strength but increased tensile and flexural modulus (Fig. 7.10). Alkali-­
treated fibers showed improved performance in all the tests conducted.
Coir fibers were cut into lengths of 5–15 mm and boiled in water for 1 h. Later,
the fibers were treated with 6% NaOH solution at 95 °C for 3 h. After the alkali
treatment, the fibers were immersed in 5% acrylic acid for 0.5–3 h to reduce the
moisture absorption. Modified fibers were placed between sheets of PLA (40/60
ratio) and compression molded at 175 °C and pressure of 1.3 MPa for 15 min
(Suardana et al. 2011). The acrylic acid treated fibers had higher tensile strength
than the alkali treated and untreated fibers. However, these composites had lower
level of biodegradation (2.5 wt%) compared to the untreated fibers. Amount of
water sorbed by the composites increased with immersion time and was lower for
the acrylic acid-treated fibers compared to the alkali only treated fiber (Suardana
et al. 2011).
To improve compatibility between PLA and reinforcement, the coir fibers were
coated with epoxidized polybutadiene (Nuthong et al. 2013). Fibers and matrix
were compounded in a twin screw extruder and made into pellets which were later
156 7 Composites from Coir Fibers

Fig. 7.9 Tensile and flexural properties of coir fibers reinforced PLA composites (Dong et al.
2014). Reproduced with permission from Elsevier

molded into standard samples suitable for testing. Impact strength of the composites
increased after addition of treated coir fibers. Using a similar approach, PLA (length
100–450 μm and density of 1.24 g/cm3) was reinforced with coir fibers that were cut
into 4.6 mm in length and later treated with hydrogen peroxide and sodium hydrox-
ide. Modified and unmodified coir fibers were combined with the PLA resin and
injection molded into pellets and later made into composites (Sun et al. 2017).
7.2 Biocomposites from Coir Fibers 157

Table 7.9 Thermal behavior of the PLA coir composites with untreated (UFC) and treated (TFC)
fibers (Dong et al. 2014)
Material Tg (°C) Tcc (°C) Tm (°C) ΔHm (J/g) Xc (%)
Neat PLA 54.78 119.0 152.9 6.44 6.93
5% UFC 53.76 112.6 151.4 8.37 9.48
10% UFC 50.83 110.0 150.5 8.65 10.33
20% UFC 54.44 112.4 147.4 6.55 8.81
30% UFC 51.49 109.5 147.8 7.86 12.08
5% TFC 53.64 114.8 150.9 6.52 7.38
10% TFC 49.60 110.0 151.3 16.66 19.91
20% TFC 51.43 114.3 148.9 15.36 20.64
30% TFC 52.77 113.3 150.6 9.42 14.47
Reproduced with permission from Elsevier

Fig. 7.10 Tensile and flexural properties of PLA reinforced with various ratios of untreated (UCF)
and alkali-treated (ACF) coir fibers. Tensile and flexural strength (a) and tensile and flexural modu-
lus (b), (Duan et al. 2018). Reproduced with permission from John Wiley and Sons
158 7 Composites from Coir Fibers

Amount of fibers in the PLA matrix was controlled between 1 and 7%. Tensile
strength of the composites decreased, whereas tensile modulus increased with
increase in fiber content. Considerable increase in adhesion between fiber and
matrix was observed for the treated fibers mostly due to improved storage modulus.
However, the composites were rigid due to the addition of the fibers. Thermal
­analysis and crystallinity data did not show any major difference between the neat
PLA and the composites (Sun et al. 2017). Biocomposites were developed using
poly(lactic acid) as the matrix and coir fibers as reinforcement (Pérez-Fonseca et al.
2016b). The matrix used had a density of 1.24 g/cm3, melt flow index of 80 g per
10 min, and melting temperature of 188–210 °C. Fibers were combined with the
matrix at three different ratios and pelletized in a twin screw extruder. The pellets
were later injection molded into dog bone specimens for tensile testing and also into
rectangular bars. No major difference was observed in the thermal behavior or crys-
tallinity of the samples with increase in fiber content from 10 to 30%. However,
dynamic modulus increased from 1400 to 1800 MPa and from 16 to 42 MPa when
measurements were done at 40 and 70 °C, respectively. Comparatively lower
improvement was observed in the tensile strength and modulus due to the fiber addi-
tion but thermal annealing assisted in improving the properties of the composites
(Pérez-Fonseca et al. 2016b).

7.2.2  oly(Caprolactone) Based Biodegradable Coir


P
Composites

Poly(caprolactone) (PCL) based composites have been mainly used for biomedical
applications but do not have desired mechanical properties (Sarasini et al. 2015). To
improve the properties of PCL-based scaffolds, three natural fibers including coir
were used as reinforcement although coir is not considered to be biocompatible.
PCL having molecular weight of 50,000 g/mol and melting temperature of 58–60 °C
was combined with the fibers in a twin screw microextruder with fiber to PCL ratio
ranging from 10 to 30%. Tensile strength of the composites containing coir fibers
was between 18.7 and 19.9 MPa, lower than that of neat PCL (21.1 MPa). Similarly,
the melting enthalpy and melting temperature had decreased. However, modulus
and shore hardness (55–57) of the composites were higher after reinforcing with
coir fibers. Crystallinity of the composite was between 30 and 41% compared to
62% for neat PCL due to the inherent low crystallinity of coir fibers (Sarasini et al.
2015). Properties of the coir reinforced PCL composites were similar to that of
commercial materials (Orfit, Rolyan) used as orthotics. PCL-based composites can
also be prepared using solvent casting technique rather than compression molding
(Obasi et al. 2018). Coir fibers were milled to obtain particles of 224 μm size. These
particles (2–5%) were combined with PCL dissolved in dichloromethane and cast
onto glass plates to form the composites. Tensile strength and elongation of the
composites was lower than that of neat PCL, whereas about 13% increase in modu-
lus was obtained when 2.5% fibers were used as reinforcement. It was suggested
that higher levels of fibers resulted in agglomeration and poor properties and
7.2 Biocomposites from Coir Fibers 159

compatibilizers were necessary to achieve better results. However, compression


molding provides better properties and also is relatively easier compared to solvent
casting approach used in this report (Obasi et al. 2018).
Coir fibers obtained from Brazilian dwarf coconut trees were pretreated and later
combined with cashew shell liquid to form biocomposites (Barreto et al. 2013). Coir
fibers cut to about 40 cm in length were washed with detergent and water and later
treated with 5 and 10% NaOH solutions at 60 to 70 °C for 6 h for delignification.
The delignified fibers were bleached using sodium hypochlorite and used as rein-
forcement. To prepare the resin, cashew nut shell liquid (cardanol) was combined
with formaldehyde and cured with epoxy resin using 2-(dimethylamino) ethyl ben-
zoate as the catalyst. Composites prepared using the resin and hand lay-up tech-
niques were found to have better thermal and X-ray diffraction properties when
modified fibers were used. Similarly, the yield strength of composites containing
10% NaOH-treated fibers was 18 MPa compared to 6.4 MPa for the unmodified
fibers. Tensile strength and breaking load were also about twice higher for the com-
posites containing the modified fibers. Interestingly, biodegradation (weight loss)
under laboratory compositing conditions of the composites containing 5% NaOH-­
treated fibers was 33%, much higher than the untreated fibers (Barreto et al. 2013).

7.2.3  oir Fiber Composites Developed using


C
Proteins and Starch as Matrix

Instead of using a biofiber and a synthetic biopolymer, biodegradable composites


can also be developed using a natural binder such as wheat gluten. Coir fibers hav-
ing a density of 1.15 g/cm3 and tensile strength between 131 and 175 MPa were
used to develop wheat gluten based composites. Before using for composites, the
fibers were cut into 40 mm length and treated with 2.5–10% alkali solution for 4 h
at room temperature. Additionally, the fibers were also modified with silane
(3-triethoxysilylpropyl)-t-butylcarbamate by dipping the fibers in the solution for
2 h (Hemsri et al. 2012). After modification, the fibers were combined with wheat
gluten (85 wt%) and compression molded at 150 °C for 10 min. Alkali treatment
removed surface impurities and also helped to have higher amount of silane bind to
the fibers (Table 7.10). The modifications resulted in a 25% increase in the stress at
first failure and 80% increase in the maximum stress. An improved adhesion and

Table 7.10 Comparison of the elemental composition and binding energies of coir fibers before
and after the alkali and silane treatment (Hemsri et al. 2012)
Elemental composition, %
Fiber C(1 s) (284 eV) O(1 s) (531 eV) Si(2p) (102 eV) N(1 s) (399 eV)
Coir fibers 72.6 24.3 1.94 1.14
Alkali treated 73.7 23.3 1.74 1.29
Silane treated 69.8 23.1 4.53 2.55
Alkali + silane 69.2 21.8 5.95 3.10
Reproduced with permission from Elsevier
160 7 Composites from Coir Fibers

reduced fiber pullouts were also observed. It was reported that wheat gluten changed
from brittle to ductile due to the addition of the fibers (Hemsri et al. 2012).
Coir fibers were used as reinforcement for toughened wheat gluten and made
into composites with thiolated poly(vinyl alcohol) as a crosslinking agent (Diao
et al. 2014). For toughening the protein, wheat gluten was dispersed in acetic acid
and combined with thiolated PVA and stirred overnight and later freeze dried and
ground into powder. Similarly, coir fibers were also modified by treating with 5%
sodium hydroxide solution for 4 h, washed and dried. Further, the alkali-treated
fibers were silanated using (3-triethoxysilylpropyl)-t-butylcarbamate to improve
adhesion and consequently mechanical properties. Composites were prepared by
compression molding the mixture at 150 °C for 10 min at a pressure of 1.1 × 108 N/
m2. The work per unit volume and modulus of the composites were highest for the
mixture made of wheat gluten, thiolated PVA, and silanated coir fibers. These val-
ues for the composites were much higher compared to neat wheat gluten or epoxy
due to the reinforcing effects. Similarly, the first failure strength also improved for
the blends but was lower than required for many applications. It was suggested that
stiffer and stronger fibers were necessary to develop practically useful products
from wheat gluten (Diao et al. 2014).
Green coconut fibers available in Brazil were used as reinforcement for starch-­
based biocomposites (Ramírez et al. 2011). The fibers were milled to about 10 mm
and combined with cassava starch and 30% glycerol. Amount of fibers in starch was
controlled between 5 and 30%. Compression molding into composites was done at
160 °C for 50 min at a pressure of 410 kgf/cm2. Annealing of the composites was
also done by treating in an oven at 60 °C for 12 h. The composites formed were
transparent and revealed the presence of the coir fibers (Fig. 7.11). Tensile strength
of the composites increased as the fiber content increased or due to the thermal
treatment but elongation showed an opposite trend. Starch which has a water sorp-
tion of about 250% decreased to about 70% when 30% fibers were included. Similar
results were also seen for the swelling behavior. Composites containing 30% fibers
also provided highest tensile strength of 3.3 MPa and modulus of 60 MPa, which are
considerably lower compared to other coir fiber containing composites. Also, the
starch-based matrices would be susceptible to moisture, resulting in decreased shelf
life and poor performance (Ramírez et al. 2011). No major changes were observed
in the chemical properties of starch but addition of fibers increased crystallinity
from 39 to 62% and also enhanced thermal stability and storage modulus from 2.2
to 3.2 GPa (Lomelí-Ramírez et al. 2014). Considerable changes were also observed
in the thermal behavior and dynamic mechanical properties of the composites as the
fiber content increased (Table 7.11).
To further improve properties, coir fibers were treated with plasma (pure oxygen
and air) to surface etch the fibers, remove lignin, and improve adhesion with bio-
polymer starch as the matrix (de Farias et al. 2017). Different surface treatment
conditions were used to obtain various levels of delignification and surface changes
(Fig. 7.12). After treatment, the fibers were compounded with starch and glycerol to
obtain thermoplastic starch which was later used to prepare the composites using
compression molding. Plasma treatment increased surface roughness and hence
7.2 Biocomposites from Coir Fibers 161

Fig. 7.11 Digital pictures of the coir fiber reinforced starch composites containing 5% (a) and
30% (b) coir fibers and image of the composites at higher magnification (c) (Ramírez et al. 2011).
Reproduced from Elsevier open access

Table 7.11 Thermal behavior and DMTA analysis results of thermoplastic starch composites
containing various amounts of coir fibers (Lomelí-Ramírez et al. 2014)
Starch, Onset Degradation Weight
% Td (°C) temperature (°C) Weight loss, % temperature (°C) loss, %
0 319.4 295.9 89.0 568.1 10.3
5 314.9 293.7 88.9 562.9 11.3
10 312.2 293.1 86.7 565.3 12.8
15 311.9 292.4 85.6 571.0 13.2
20 310.6 291.8 83.8 575.5 14.8
25 309.1 290.2 83.3 576.8 15.2
30 305.9 288.8 82.2 564.3 16.2
Material β-Relaxation, Intensity α-Relaxation, Intensity Modulus,
°C °C MPa
0 −27 −0.15 73 −0.29 2027
5 −27 −0.10 75 −0.20 2066
10 −25 −0.09 79 −0.20 2587
15 −24 −0.08 82 −0.20 2908
20 −24 −0.08 86 −0.20 3198
25 −25 −0.06 89 −0.20 3534
30 −25 −0.05 94 −0.20 3215
Reproduced from Elsevier open access. Reproduced with permission from Elsevier
162 7 Composites from Coir Fibers

Fig. 7.12 Changes in the surface morphology of the fibers with different levels of plasma treat-
ments with air (a/b) and oxygen (c/d), at 50 W power (a/c) and 80 W power (b/d) (De Farias et al.
2017). Reproduced with permission from Elsevier

increased surface properties. A drastic decrease in lignin/cellulose ratio from 16.2 to


0.99 was observed after treating with oxygen plasma. Consequently, a 300%
increase in strength and 2000% increase in modulus were observed due to the
­inclusion of the modified fibers. An increase in compatibility factor from 0.0087 to
0.2847 was observed after plasma treatment (De Farias et al. 2017).

7.3 Hybrid Composites

Several agricultural residues (coir pith, rice husk, and groundnut shell) were made
into particles and combined with epoxy resin to form hybrid composites. Particles
having diameters between 10 and 60 μm and epoxy resin having a melt flow index
of 1.8 g/min and density of 1.36 g/cm3 were used (Prithivirajan 2015). Rice husk
provided better adhesion to the matrix and groundnut shell reinforcement increased
strength directly. However, addition of coir particles increased water sorption but
the mechanical properties were inferior compared to the other two agricultural resi-
dues mainly due to the poor crystallinity and higher moisture sorption (Prithivirajan
2015). Coir fibers (10%) were used as reinforcement for epoxy resin with 20% of
7.3 Hybrid Composites 163

walnut particles (1.6–2.7 μm) as fillers. Incorporating the biomass into neat epoxy
decreased the tensile stress from 5.6 to 4.9 MPa. Similarly, modulus of elasticity
decreased from 1633 MPa to 1411 MPa (Kumar Rao and Chandra Gope 2015).
Extensive fiber pullout and fiber cracking were observed, which led to lower
mechanical properties (Kumar Rao and Chandra Gope 2015). In another study,
coconut fibers having length of 3, 5, and 9 mm were combined with Caesar weed
fibers in a polymer matrix and made into composites using the hand lay-up approach.
Fiber length and volume increased the mechanical strength when the ratio of coir
and Caesar fibers was 1:1 and ratio of total fiber to matrix was 30–70%. Tensile
strength and flexural strength of the composites were 55 and 91 MPa, respectively,
and impact energy was 6.8 J (Nelson et al. 2015), considerably higher than that of
the neat polymer.
A combination of oil palm and coir fibers was used to reinforce high-density
polyethylene. Coir fibers were ground to obtain an average length of 730 μm and an
average width of 140 μm. In addition, a coupling agent (maleic anhydride grafted
polyethylene) was used to improve fiber matrix adhesion (Kakou et al. 2015). Fibers
and the matrix were compounded in a twin screw extruder to obtain pellets. The
pellets were later compression molded in a carver press at 175 °C. Morphological
images clearly showed better adhesion due to the presence of the compatibilizer
(Kakou et al. 2015). Compared to the modulus of the neat polyethylene at 350 MPa,
using 40% coir fibers with compatibilizer increased the modulus to 700 MPa.
Overall, coir fibers provided better properties than oil palm fibers when used indi-
vidually, and hybrid composites had properties between that of the individual oil
palm or coir fiber reinforced composites (Kakou et al. 2015). In a similar study, coir
and palm fibers were boiled in 25% NaOH at 100 °C for 3 h. Treated coir fibers were
later subject to bleaching in 50% hydrogen peroxide and the palm fibers with 30%
hydrogen peroxide in the presence of 1% sodium silicate, 0.5% magnesium sulfate,
and 1.5% NaOH. Morphologically, the coir fibers were different than that of the
palm fibers (Fig. 7.13). Treated fibers were placed between the LDPE sheets and hot
pressed at a temperature of 180 °C. Single fiber pullout tests showed that the
bleached palm fibers had higher stress than coir fibers. Young’s modulus of the
composites was highest when both fibers were used as the reinforcement and
increased with increasing fiber content and also after bleaching.
A hybrid composite was manufactured using coir, E-glass fiber, and epoxy resin.
The components were mixed up manually using 5–10% coir fibers between 5 and
20 mm in length and 20% glass fibers of 10 mm in length (Bhagat et al. 2014). A
maximum flexural strength of 63 MPa was observed for composites having 10%
fiber loading at 15 mm fiber length. A maximum hardness value of 21.3 Hv was also
obtained for the 10 wt% fibers.
Instead of using fibers, coir mats were used alone and also in combination with
glass fabrics to develop hybrid composites (Hamouda et al. 2017). The mats had
weight per unit area of 763 g/m2 and glass fabrics were of 800 g/m2. Coir and glass
fabrics were placed in various configurations and placed in a vacuum assisted resin
transfer mold. Properties of the composites were dependent on the configuration of
the coir and glass fabrics as seen from table. Generally, composite properties
164 7 Composites from Coir Fibers

Fig. 7.13 Morphology of the coir (top four) and palm fibers (bottom four) before and after the
alkali treatment (Chollakup et al. 2013). Reproduced with permission from Taylor and Francis
7.3 Hybrid Composites 165

Table 7.12 Influence of the composite structure on the thickness and tensile properties (Hamouda
et al. 2017)
Sample Thickness, mm Break load, kN Tensile strength, MPa
C 3.7 0.239 4.2
CC 7.8 0.862 8.2
GCG 5.2 4.279 69.2
GCCG 8.4 4.973 43.3
GCGCG 8.9 10.701 70.0
Reproduced with permission from John Wiley and Sons

increased due to the inclusion of glass fabrics (Table 7.12). For instance, flexural
strength increased from 31.8 to 131.5 MPa. Instead of using glass in the form of
fabrics, glass spheres, functionalized clay, and alkali-treated coir fibers were used to
reinforce epoxy (Muthu et al. 2018). The glass spheres had a particle size between
9 and 13 μm and density of 1.05 g/cm3 compared to 60–90 μm and 2.02 g/cm3 for
the clay particles, respectively. Clay used in the study was a natural zeolite
­(clinoptilolite), which was considerably hydrophilic and difficult to be combined
with the hydrophobic matrix. To decrease hydrophilicity, the clay particles were
functionalized using 3-amino-propyltriethoxysilane for 30 min. Composites were
fabricated using the coir fibers (7, 10, 15 and 20%) and epoxy and also with the
glass and clay fillers (15% each by weight) by the vacuum assisted resin transfer
molding (VARTM) process (Muthu 2017). Unlike in many other studies, addition of
coir fibers increased both the tensile strength and elastic modulus (Fig. 7.14).
Similarly, addition to clay particles also increased the strength and modulus up to a
clay content of 4% but 8% glass spheres were found to provide the optimum
properties.
In order to obtain composites with good dielectric properties, coir dust was com-
bined with fly ash and used as reinforcement for epoxy (Epoxy LY 556) along with
HY-951 as the hardener. Hybrid composites had a hardness value of 20. Pure fly ash
had the highest dielectric loss in the entire frequency range studied. Similarly, a
hybrid of coir dust and fly ash had better dielectric loss than the other composites
(Dhal and Mishra 2013) with values ranging from 2.25 to 0.25. In another study,
coir as fibrous reinforcement and rice husk as particulate reinforcement were used
to develop vinyl ester composites (Ramprasath et al. 2016). Inclusion of rice husk
particles improved adhesion and hence provided better properties. A high impact
strength of 43 kJ/m2 was obtained when 30 mm fibers and 15% particulates were
used.
Two natural fibers, coir and banana were pretreated and combined with epoxy
resins to manufacture plates (Hariprasad et al. 2013). Coir fibers were immersed in
10% NaOH solution for 3 h at 70 °C and banana fibers were treated in 6% NaOH
solution for 2 h at room temperature. The treated fibers were combined with a resin
diglycidyl ether (EP 306) and a hardener diethyltetraamine (EH 758) having vis-
cosities of 9500–12,500 cps and 20–50 cps at 25 °C, respectively. Total amount of
fibers in the epoxy resin was considerably low at 23% and the ratio of the coir to
166 7 Composites from Coir Fibers

Fig. 7.14 Changes in the strength and modulus of the coir-epoxy reinforced composites rein-
forced with various levels of clay particles (Muthu et al. 2018)

banana fibers was 45:55. Marginal increase in tensile strength of about 10% and no
significant increase in modulus were observed. Flexural strength had decreased by
about 45%, whereas impact strength had increased by about 200% after the treat-
ment (Hariprasad et al. 2013). In another study, coir and banana fibers were com-
bined and used as reinforcement for phenol formaldehyde resin (Wang and Hu
2016). Coir and banana fibers were cut into 1–4 cm length before using as
­reinforcement. The fibers were mixed with 14% resin and liquid paraffin and later
compression molded into desired shape at 160 °C at a pressure of 3–10 MPa.
Composite boards obtained had a density of 0.88 g/cm3 and dimensions of
320 mm × 220 mm and 5 mm thickness. Mechanical properties of the composites
(Table 7.13) were directly dependent on the ratio of coir and banana fibers. Coir rich
7.3 Hybrid Composites 167

Table 7.13 Properties of coir:banana fiber reinforced composite boards at various ratios of the
two fibers (Wang and Hu 2016)
Coir:banana fiber Modulus of Modulus of Internal bond Water
ratio rupture, MPa elasticity, GPa strength, MPa sorption, %
10:0 22.7 1.53 1.27 10.7
9:1 25.1 1.89 1.12 11.8
8:2 30.0 2.23 0.92 12.8
7:3 33.3 2.34 0.89 14.3
6:4 36.7 2.49 0.82 15.5
5:5 39.5 2.68 0.63 18.4
0:10 50.6 3.71 0.31 21.9
Reproduced with permission from Springer Nature

composites had comparatively low properties except for internal bond strength. The
coir fibers composites also sorbed about 50% less water than the banana fiber rich
composites. The optimum ratio of coir to banana fibers was found to be 8:2 and the
boards developed satisfied the performance requirements as per national standards
GB/T4897.4 (2003) and GB/T17657 (2013) (Wang and Hu 2016). In another study,
hybrid composites were developed by combining coir and banana fibers with low
density polyethylene having melt flow index of 34 g/10 min and density of 0.930–
0.945 g/cm3. Coir and banana fibers in the desired ratio were combined with poly-
ethylene along with maleic anhydride modified low density polyethylene as the
compatibilizer (Prasad et al. 2017). Composites were formed by compression mold-
ing at 180 °C for 10 min at a pressure of 20 MPa. Both tensile and flexural proper-
ties of the hybrid composites were higher compared to those manufactured using the
individual fibers (Prasad et al. 2017). Morphological studies indicated that the addi-
tion of coir fibers into banana fibers composites decreased fiber pullout and improved
fiber/matrix interfacial bonding. Hybrid composites developed using coir and sisal
fibers with poly(lactic acid) as the matrix showed improved composite properties
due to hybridization (Fig. 7.15). The most optimum mechanical properties were
obtained with a combination of 70% sisal and 30% coir fibers treated with alkali. It
was suggested that inclusion of coir improved compressive strength, whereas sisal
fibers provided better tensile and flexural properties (Duan et al. 2018).
Perez-Fonseca et al. studied the effect of coupling agents on the mechanical
properties and water sorption of coir and agave fiber reinforced high-density poly-
ethylene composites (Pérez-Fonseca et al. 2016a). Reinforcing fibers were washed
thoroughly in water and milled into 50–70 mm for agave fibers and 100 and 140 mm
for coir. Later, the fibers were combined with polyethylene in different weight ratios
with or without maleated polyethylene (Fusabond M603) (3% based on weight of
fiber) as the coupling agent. The materials were first compounded in a twin screw
extruder to form pellets which were later injection molded between 130 and 195 °C
(Pérez-Fonseca et al. 2016a). Increasing proportion of the fibers in the composites
continually decreased the tensile strength (Fig. 7.16). Maximum tensile strength
(22 MPa) was obtained when higher agave fiber content was used for the 20% fiber
containing composites, whereas coir fibers provided higher strength (23 MPa) for
168 7 Composites from Coir Fibers

Fig. 7.15 Effect of fiber composition on the mechanical properties of the PLA composites rein-
forced with various levels of alkali-treated sisal (ASF) and alkali-treated coir fibers (ACF) (Duan
et al. 2018). Reproduced with permission from John Wiley and Sons

the 30% fiber composites. However, the tensile properties of the composites were
lower than that of neat HDPE (28 MPa), suggesting poor interaction between the
fibers and matrix (Pérez-Fonseca et al. 2016a). Flexural properties of the compos-
ites also followed a similar trend to tensile properties. Hybridization provided better
impact strength but inclusion of the compatibilizer only had a marginal improve-
ment in flexural properties at 30% fiber content and no change was observed at 20%
fiber content (Pérez-Fonseca et al. 2016a). Similar to using agave fibers, coir fibers
were also blended with sisal fibers using a process called commingled yarn prepara-
tion. The blend fibers contained 80% coir and 20% sisal and subject to various treat-
ments to improve compatibility and composite properties. Fibers were treated with
KMnO4 in acetone, vinyltrimethoxy silane (VTMO) in water-ethanol, MAPP in
acetone, and toluene diisocyanate (TDI) in chloroform. Coir and sisal fiber blended
yarns along with PP were wound on a metal plate and compressed into composites
at a temperature of 210 °C and 0.5 MPa pressure for 9 min, maintaining the ratio of
the fibers at about 13, 21, or 31%. Tensile strength and modulus of the composites
increased with increasing fiber content. Fibers modified with MAPP provided both
increased strength and modulus, higher than that of the other modifications (Arya
et al. 2015).
Higher lignin content in coir fibers was considered to be suitable for modification
and to increase the interaction between the reinforcement and matrix. Hence, coir
7.3 Hybrid Composites 169

Fig. 7.16 Changes in the properties of coir-agave fiber hybrid composites before and after treating
in water at 20% (a) and 30% (b) fiber content (Pérez-Fonseca et al. 2016a). Reproduced with per-
mission from John Wiley and Sons

and jute fibers were used in a sandwich as reinforcement with epoxy novalac as the
matrix. Properties of the jute and coir fibers used in developing the composites are
given in Table 7.14. Jute fabrics were washed in 5% NaOH solution for 2 h at 50 °C
and the coir fibers were soxhlet extracted with hexane-ethanol and later treated with
aqueous chlorine dioxide. Further, the oxidized fibers were modified by grafting
with furfuryl alcohol at 100 °C for 4 h under nitrogen atmosphere. The best
170 7 Composites from Coir Fibers

Table 7.14 Properties of the Physical properties Jute fiber Coir fibers
jute and coir fibers used to
Density, g/cm3 1.45 ± 0.1 1.3 ± 0.1
prepare the hybrid
composites (Saw et al. 2012) Cellulose content, % 65 ± 2 33.3 ± 1.2
Hemicellulose content, % 20 ± 2 2.67 ± 1.44
Lignin content, % 12 ± 2 46.8 ± 0.8
Microfibrillar angle, ° 8.1 39 ± 5
Diameter, μm 54 ± 6 200 ± 10
Tensile strength, MPa 473 144.6
Young’s modulus, MPa 19,500 3101
Elongation at break, % 1.17 ± 0.2 32.3 ± 0.2
Aspect ratio (L/D) 365 ± 10 100 ± 5
Reproduced with permission from John Wiley and
Sons

Fig. 7.17 Changes in the


impact strength with
increase in fiber content for
the kenaf–coir novolac
resin hybrid composites
(Saw et al. 2012).
Reproduced with
permission from John
Wiley and Sons

p­ roperties were found for the jute–coir ratio of 50/50 with tensile and flexural
strength of 25 and 63 MPa for the modified fibers. The impact strength of the fibers
increased up to 35% fiber content (Fig. 7.17) and decreased at higher fiber ratios. It
was suggested that fiber modifications were necessary to obtain hybrid composites
with good properties (Saw et al. 2012).
In another study, the feasibility and potential of combining coir and kenaf fibers
to reinforce polypropylene were investigated (Islam et al. 2015a). The amount of
polypropylene in the composites was fixed at 70% and the proportion of the two
natural fibers was varied from 0 to 30%. Also, montromollite having diameters
7.3 Hybrid Composites 171

between 16 and 22 μm was used as a compatibilizer. Before using as reinforcement,


the fibers were treated with 2% NaOH at room temperature for 5 h. Composites
were developed by combining the fibers with PP granules in a Brabender extruder
and later compression molding on a hot press. Samples were prepared according to
ASTM 638 and ASTM 790 standards for testing. Morphological analysis showed
that there was poor dispersion of the clay and interfacial bonding, leading to fiber
pullouts from the composites. Although hybridization with two natural fibers did
not show any significant change in morphology, addition of the nanoclay increased
the interfacial interaction, resulting in considerably low fiber pullouts and micro-
fractures. In terms of tensile properties, hybridization increased tensile strength but
did not affect the modulus (Fig. 7.18). Coir reinforced PP had lowest tensile strength
and modulus compared to kenaf mostly likely due to the inherently poor coir fiber
properties compared to kenaf. Inclusion of coir fibers decreased the biodegradabil-
ity (Fig. 7.19) of the composites due to the high lignin content in the fibers but addi-
tion of the nanoclay increased the degradability of coir containing composites while
that of the kenaf composites decreased (Islam et al. 2015b).
Another approach to improve the properties of coir-based hybrid composites was
to add nanoparticles synthesized from coconut shells as fillers (Abdul Khalil et al.
2017). Nanofillers were prepared by successively grinding coconut shells and later
extracting in n-hexane. Three different ratios (1, 3, and 5%) of nanofillers (15–
140 nm) were added into vinyl ester resin and methyl ethyl ketone peroxide as the
catalyst and cobalt naphthenate (0.2% wt) resin as the accelerator. Coir containing
composite also had lower biodegradability compared to kenaf reinforced compos-
ites because of the high lignin content in coir. Hybrid nanocomposites were pre-
pared by placing layers of kenaf–coir–kenaf in the resin solution and cold pressed
at 200 psi at room temperature for 24 h. Tensile, flexural strength and modulus of
the composites increased with increase in the nanofillers content from up to 3% and
later decreased at 5% filler concentration. Morphological images showed improved
better adhesion between the matrix and fibers due to the addition of the nanofillers
based on the fiber pullout behavior. A filler concentration of 3% was found to be
optimum for the coir/kenaf hybrid composites.
Another unique fiber (luffa) was also combined with coir and used as reinforce-
ment for polypropylene. Blending coir and luffa resulted in lower tensile strength
but higher hardness number (Sakthivel et al. 2014).
Considerably different results were obtained when coir fibers and sugarcane
bagasse were combined and made into hybrid composites using vinyl ester as the
resin (Stalin and Athijayamani 2016). Amount of fibers in the resin was varied from
10 to 50%, maintaining the coir to bagasse ratio at 1:1. Length of the fibers in the
resin was fixed at 15 mm. Tensile stress and modulus increased continually when
the ratio of fibers was increased from 10 to 40%. Similar results were also observed
for the flexural properties. Higher amount of fibers also increased the wear resis-
tance with 40% fiber containing composites providing the most optimum resistance
(Stalin and Athijayamani 2016). In another study, coir fibers were combined with
kenaf fibers after treating with 5% NaOH for 4 h. An epoxy (LY556) was used as the
resin for the composites formed using the hand lay-up approach. Fibers used had
172 7 Composites from Coir Fibers

16

12
Tensile Strength (MPa)

P T P T P T
f/P M r/P M r/P M
na P/
M oi P/
M oi P/
M
C /C
Ke f/P r/P af r/P
a oi n oi
K en C Ke f/C
na
Ke

500

400
Young’s Modulus(MPa)

300

200

100

0
P T P T P T
f/P M r/P M r/P M
na P/
M oi P/
M oi P/
M
C /C
Ke f/P r/P af r/P
a oi n oi
K en C Ke f/C
na
Ke

Fig. 7.18 Changes in the tensile strength and modulus of the polypropylene reinforced with indi-
vidual, hybrid fibers with and without nano montmorillonite (Islam et al. 2015a). Reproduced from
North Carolina State University
7.3 Hybrid Composites 173

25 30 days

60 days

19.37 90 days
20
17.28 17.5
16.97
15.71
Weight Loss (%)

14.68 15 14.61
15 14.03
13.76
12.67 12.9 12.79
11.76
10.53 10.96

10 9.09
8.61

0
Kenaf/PP Kenaf/PP/MMT Coir/PP Coir/PP/MMT Kenaf/Coir/PP Kenaf/Coir/
PP/MMT

Fig. 7.19 Extent of biodegradation of the various biocomposites (Islam et al. 2015b). Reproduced
with permission from Elsevier

length of 6 mm and 10% fibers were used as reinforcement in the composites. A


50:50 blend of coir and kenaf fibers was found to provide the most optimum proper-
ties (Vijayakumar 2014) but the composites had low charpy impact energy.
Unsaturated polyester resin was combined with coir fibers and glass fibers to study
the reinforcing effect of glass fibers on the properties of the composites. The glass
fibers were made into a fabric with density of 930 g/m2, and the coir mat had thick-
ness of 10 mm. Fibers were immersed in the unsaturated polyester resin and later
extruded to control the resin content to be between 80 and 85% in the composites.
Various configurations of the two fabrics were used to prepare the composite which
determined the properties. Incorporating the glass fibers into the composites
increased all the properties by several magnitudes (Table 7.15). However, poor
interfacial adhesion due to incompatibility between coir and matrix caused the com-
posites to have decreased properties when coir fibers were used as reinforcement.
Inclusion of the glass fibers resulted in 419% increase in flexural strength, 708%
increase in modulus of elasticity, and 562% increase in impact strength compared to
the composites without the glass fibers. Although glass fibers increased the
­mechanical properties, it also results in composites with considerably high density
(Cheng et al. 2014).
A combination of coir and oil palm empty fruit bunch fibers was used to develop
polypropylene hybrid composites (Zainudin et al. 2014). The two fibers had distinct
properties (Table 7.16) and were mixed manually in different ratios. Later, the fibers
were compounded with polypropylene in an extruder and the pellets obtained were
injection molded into the desired shape. Tensile strength and modulus of the hybrid
174 7 Composites from Coir Fibers

Table 7.15 Conditions during preparation and properties of the composites with coir and glass
fibers at various compression pressures (Cheng et al. 2014)
Resin, Pressure, Flexural Modulus of Impact
% MPa strength, MPa elasticity, MOE strength, kJ/m2
Coir fibers 83.4 0.4 24.0 0.6 4.4
79.4 0.8 26.4 1.9 8.0
76.5 1.2 35.6 2.3 10.2
66.6 1.6 28.8 1.9 4.4
Coir fibers + glass 66.4 0.4 176.6 16.2 48.2
fibers 62.9 0.8 185.0 18.3 67.2
61.4 1.2 132.8 16.0 61.4
60.3 1.6 153.9 15.6 60.3
Reproduced with permission from Springer Nature

Table 7.16 Properties of the Properties Coir Oil palm


coir and oil palm empty fruit Cellulose, % 32–43 49.6–65
bunch fibers (Zainudin et al.
2014) Hemicellulose, % 0.15–0.25 19
Lignin, % 40–45 19–21.2
Tensile strength, MPa 144 130
Young’s modulus, GPa 4–6 1–9
Elongation at break, % 15–40 14
Microfibrillar angle, ° 41–45 46
Reproduced with permission from John Wiley
and Sons

composites were lower than that of neat PP. The highest strength and modulus were
obtained for the composites containing 50/50 ratio of the two fibers. Although a
weak interaction between the fibers and matrix was noticed, addition of coir reduced
fiber pullout indicating increased interaction. Overall properties of the composites
are given in Table 7.17.
Coir fibers were combined with snake grass and made into hybrid composites
using isophthalic polyester resin as the matrix. Density of the resin was 1.12 g/cm3,
tensile strength was 18 MPa, and modulus was 0.8 GPa. A trilayer composite was
formed by placing snake grass in the middle and coir fiber layers above and below
(Sathishkumar et al. 2013). Ratio of fibers in the composites was varied from 0 to
30% and ratio of grass to coir was 1:1. Tensile strength and modulus both increased
with increasing fiber content. Coir reinforced composites had higher tensile proper-
ties compared to snakegrass. Flexural strength also had a similar trend to tensile
strength, whereas flexural modulus was lower than that of snakegrass fibers or even
bagasse. Coir reinforcement (20%) was found to provide more suitable flexural
strength than the other reinforcing materials (Sathishkumar et al. 2013).
7.3 Hybrid Composites 175

Table 7.17 Properties of the PP composites reinforced with various ratios of coir and palm fruit
fibers (PF) (Zainudin et al. 2014)
Composition,
wt% Tensile Modulus, Flexural Flexural Impact
Coir PF PP strength, MPa GPa strength, MPa modulus, GPa resistance, J/m
– – 100 41.5 2.9 54.7 2.0 28.1
0 30 70 24.9 1.5 42.3 1.9 27.9
7.5 22.5 70 24.5 1.4 40.8 1.7 26.3
15 15 70 30.8 2.1 53.0 2.9 29.2
22.5 7.5 70 29.9 1.5 52.6 2.4 22.6
30 0 70 26.6 1.9 50.5 2.9 24.7
Reproduced with permission from John Wiley and Sons

Two distinctly different natural materials (wood fiber and coir fibers) were com-
bined with MMT nanoclay having dry particle size of 16–22 mm and made into
composites using polypropylene as the matrix and compression molding at 170 °C
for 15 min. The natural fibers were also treated by immersing in 2% NaOH to
improve adhesion and mechanical properties (Islam et al. 2015a). The fibers and
wood had relatively rough surface and poor interfacial interaction. Addition of the
nanoclay increased the smoothness and also the interfacial adhesion. The coir/PP
composites had lower tensile strength, whereas the hybrid composites consisting of
the wood/coir/PP and MMT had highest tensile strength of about 12 MPa.
Interestingly, the coir/pp. composites had lower weight loss after biodegradation for
90 days, whereas the hybrid composites had weight loss of about 18% after 90 days
(Islam et al. 2017).
In a unique approach, coir fibers were grafted with a natural phenol (syringalde-
hyde) using a laccase from Trametes versicolor and used as reinforcement for
poly(butylene succinate). For the modification, coir fibers were treated with laccase,
syringaldehyde, and citrate buffer at pH 3.4 at 50 °C for 12 h (Thakur et al. 2016).
To prepare the composites, PBS and the coconut fibers were combined in various
ratios and heated in a compression mold at 125 °C for 20 min at 100 MPa. The
extent of biografting was dependent on the grafting conditions and varied from 6 to
16%. Composites reinforced with the modified fibers had considerably higher
mechanical properties than the composites reinforced with ungrafted fibers or com-
pared to neat PBS (Thakur et al. 2016). In a similar approach, coir fibers were modi-
fied by laccase mediated biografting of eugenol (Fig. 7.20), and the modified fibers
were used as reinforcement for PBS to improve hydrophobicity and antimicrobial
properties. Although only 0.5 and 1% of the modified fibers were used, substantial
activity against both E. coli and S. aureus was observed (Fig. 7.21) (Thakur et al.
2016). Considerable decrease (12 to 5% at 55% RH and 18 to 7% at 75% RH) in
hydrophilicity was observed after grafting the fibers with eugenol. Similarly, tensile
strength was found to increase from 5 to 35 MPa (Fig. 7.22). However, coir was
found to provide considerably lower increase in mechanical properties and also
176 7 Composites from Coir Fibers

Fig. 7.20 Schematic representation of the grafting of eugenol onto the coir fibers and preparation
of the PBS composites (Thakur et al. 2016). Reproduced with permission from Elsevier

Fig. 7.21 Antibacterial activity of the PBS composites reinforced with eugenol grafted coir fibers
(Thakur et al. 2016). PBS composites without any fibers (a and d), b and e (ungrafted coconut
fibers), and c and f (eugenol biografted fibers). Reproduced with permission from Elsevier
7.3 Hybrid Composites 177

Fig. 7.22 Changes in the mechanical properties of the PBS containing grafted and ungrafted coir
fibers (a) 0.5% fiber content, (b) 1% fiber content and (c) effect of fiber content on tensile strength.
(Thakur et al. 2016). Reproduced with permission from Elsevier

thermal resistance for PBS-based composites compared to other lignocellulosic


sources such as bagasse, curaua, and sisal (Frollini et al. 2013).
Hybrid composites in the form of beads (Fig. 7.23) were developed using sodium
alginate, natural rubber, and coir. The polymers in the form of beads were cross-
linked using calcium chloride (Riyajan and Tangboriboonrat 2014). Coir fibers were
milled to 80–250 μm and later treated with 10% NaOH for 12 h and bleached with
5% hydrogen peroxide for 5 h. Beads were prepared by adding rubber latex into
sodium alginate solution containing the treated coir fibers and extruding (using a
needle) the mixture into the crosslinking (CaCl2) solution. NMR studies showed
chemical bonding between the sodium alginate and calcium chloride solution.
Beads formed had diameter between 1.1 and 1.2 mm. Inclusion of the fibers resulted
in higher surface roughness, which was supposed to improve adsorption of Pb2+
ions. Up to 99% of the ions could be absorbed by the beads (Riyajan and
Tangboriboonrat 2014).
178 7 Composites from Coir Fibers

Fig. 7.23 Images of the beads made using 7.5% coir fibers when observed using a digital and
scanning electron microscope (Riyajan and Tangboriboonrat 2014). Reproduced with permission
from Society of Plastics Engineers

7.4 Coir Shell and Coir Pith Composites

In addition to using coir fibers, coconut shell powder was added as filler for polyes-
ter resin based composites. Coir fibers having length of 1–70 cm and coconut shell
powder of 300 μm size were mixed with methyl ethyl ketone peroxide as hardener
and poured into a mold to form the composites. Extent of fibers and shell powder in
the composite was varied between 13 and 27%. Tensile strength of the composites
increased, whereas the compressive strength decreased with increase in fiber con-
tent. Flexural strength also decreased but hardness numbers increased from 54 to 88
(Davis 2016). Ability of coconut shell powder and coir fibers to improve the wear
resistance of polyester composites was investigated (Ibrahem 2016). Increasing pro-
portion of coir powder or fiber decreased the friction coefficient. Similarly, wear
rate decreased from 0.0035 g/m to 0.00018 g/m with increasing reinforcement
(Fig. 7.24).
Similar to coir fibers, coir pith has also been used to prepare composites. Coir
pith having moisture content of 10–15% was made into particles of different sizes
and mixed with phenol formaldehyde (16.7%) or urea formaldehyde (20.4%) resins
and cured for 26 and 17 min at 138 °C for the two resins, respectively (Viswanathan
and Gothandapani 1999). Tensile strength of the composites was dependent on the
particle size. Increasing particle size increased the strength of composites for phenol
formaldehyde but decreased for the urea formaldehyde resin. Composites devel-
oped using the phenol resin and particle size of 2.1µm had good nail holding and
7.4 Coir Shell and Coir Pith Composites 179

Fig. 7.24 Effect of coconut reinforcement on the wear rate of polyester composites (Ibrahem
2016)

Table 7.18 Ability of the coir-based composites to withstand screw and nail holding capacity
(Viswanathan and Gothandapani 1999)
Along face, N Along edge, N
Particle size, mm Phenol Urea Phenol Urea
Screw holding capacity
0.45 400 347 163 113
0.80 590 487 267 147
1.20 747 637 387 197
2.10 957 713 577 413
Only resin 693 690 313 267
Mixed particle sizes 637 577 253 173
Nail holding capacity
0.45 396 277 83 47
0.80 490 347 127 53
1.20 520 410 157 87
2.10 560 460 220 150
Only resin 507 337 183 130
Mixed particle sizes 437 327 153 113
Reproduced with permission from Elsevier

screw holding capacity both along the face and edge of the composites (Table 7.18).
In another study, coir pith was combined with ethylene propylene-diene monomer
and polypropylene with fiber ratios from 0 to 20% (Abitha and Rane 2015). Tensile
strength and elongation decreased as the fiber content was increased, whereas
impact strength increased from 5.5 to 10.5 kJ/mm2 and tear strength also increased
from 10.5 to 14 kg/cm when the fiber proportion was varied from 0 to 20%.
180 7 Composites from Coir Fibers

Coir pith has also been used to develop hybrid biocomposites. Alkali-treated (5%
NaOH for 1 h at room temperature) coir pith was combined with nylon fabric hav-
ing an epoxy coating and placed between polyethylene sheets. The pre-preg was
compressed to form the composites. Inclusion of treated coir pith increased impact
strength and water resistance of the composites. It was found that alkaline treatment
was necessary to improve the properties of coir pith reinforced composites (Narendar
and Dasan 2013). Hydroxyapatite extracted from bones of sea bass was combined
with chitosan and coir pith and made into composites. To prepare the composites,
coir pith was powdered and added into chitosan solution dissolved in 2% acetic acid
and added into hydroxyapatite solution prior to adding the fibers. Tensile strength of
the composites was 1.58 MPa compared to 0.06–0.16 for urea or phenol formalde-
hyde and coir pith based composites. FTIR spectrum shifts indicate some interac-
tion between the two components. Composites obtained had an average compressive
strength of 14 MPa and tensile strength of 1.5 MPa and a relatively high water sorp-
tion of 32.7%. Soil burial tests indicated that a weight loss of up to 6.4% was pos-
sible after 120 days (Rajesh et al. 2014). A combination of coir pith, nylon, and
epoxy was used to develop hybrid composites with good thermal insulation
­properties. Coir pith fibers were initially treated with NaOH and later subject to a
series of chemical modifications using various chemicals (Narendar et al. 2018).
Modified coir pith was mixed with resin in a blender and placed in the mold. To
form hybrid composites, different layers of nylon fabrics treated with resin and
hardeners were placed over the coir pith and compression molded to form the com-
posites. Thermal stability of the hybrid composites containing treated coir pith was
substantially higher than those of untreated fibers. Also, with a Limiting Oxygen
Index (LOI) between 0.6 and 1, the coir pith/epoxy composites were considered to
have good flame resistance. Both impact strength and heat deflection temperature
were higher for the hybrid composites and considered suitable for commercial
applications (Narendar et al. 2018). Other studies have also shown a considerable
decrease in the impact strength of coir pith based epoxy/nylon composites (Narendar
and Dasan 2015) particularly when exposed to water. However, hybridization with
nylon provides improved water resistance and makes the composites suitable for
various applications. In another study, coir pith was combined with groundnut shell
and rice husks and made into composites with epoxy as the matrix (Prithivirajan
et al. 2016a, b). Tensile strength of the composites varied from 9 to 26 MPa and
flexural strength from 21 to 32 MPa depending on the proportion of the three rein-
forcements. The composites were susceptible to water and the sorption increased
with immersion time. Lowest swelling was obtained for composites containing
22.5% rice husk and 7.5% groundnut shells.

7.5 Coconut Sheath

In addition to coir fibers and coir pith, the sheaths of coconut tree leaves have also
been used as reinforcement for composites (Kumar et al. 2014). Fibers were
extracted from coconut sheaths by treating them in 5% sodium hydroxide solution
References 181

Fig. 7.25 SEM images of the untreated (left) and alkali-treated coconut sheath fibers (right)
(Kumar et al. 2014). Reproduced with permission from Elsevier

for 1 h and later neutralizing using acetic acid. Layers of the untreated and treated
fibers were coated with epoxy resin, stacked together and compression molded at
3.9 MPa for 24 h at room temperature. Alkali treatment of the fibers removed sub-
stantial impurities and resulted in a clean fiber surface (Fig. 7.25). An average ten-
sile strength of 48.4 MPa, flexural strength of 65 MPa, and impact strength of 4.5 J/
mm were obtained for the untreated sheath fibers compared to 58.6, 76.8, and 5.6,
respectively, for the treated fibers (Kumar et al. 2014). Composites made with the
treated fibers also had higher thermal stability and lower char yield due to the better
fiber–matrix binding.

References

Abdul Khalil HPS, Masri M, Saurabh CK, Fazita MRN, Azniwati AA, Sri Aprilia NA, Rosamah E,
Dungani R (2017) Incorporation of coconut shell based nanoparticles in kenaf/coconut fibres
reinforced vinyl ester composites. Mater Res Exp 4(3):035020
Abdullah NM, Ahmad I (2012) Effect of chemical treatment on mechanical and water-sorption
properties coconut fiber-unsaturated polyester from recycled PET. ISRN Mater Sci 2012:134683
Abitha VK, Rane AV (2015) Studies in mechanical, thermal and morphological analusis of EPDM/
PolyPropylene coconut pith composites. Moroccan J Chem 3(2):3–2
Adnan NAM, Saidin WANW, Musa MA, Zaidi AMA, Mohideen SR (2014) Mechanical perfor-
mance of coconut coir fiber reinforced urea formaldehyde composites. Aust J Basic Appl Sci
8(15):205–210
Aguele FO, Madufor CI (2012) Effects of carbonised coir on physical properties of natural rubber
composites. Am J Polym Sci 2(3):28–34
Aguele FO, Madufor CI, Adekunle KF (2014) Comparative study of physical properties of polymer
composites reinforced with uncarbonised and carbonised coir. Open J Polym Chem 4(3):73–77
Arrakhiz FZ, El Achaby M, Kakou AC, Vaudreuil S, Benmoussa K, Bouhfid R, Fassi-Fehri O,
Qaiss A (2012) Mechanical properties of high density polyethylene reinforced with chemically
modified coir fibers: impact of chemical treatments. Mater Des 37:379–383
182 7 Composites from Coir Fibers

Arya A, Tomlal JE, Gejo G, Kuruvilla J (2015) Commingled composites of polypropylene/


coir-sisal yarn: effect of chemical treatments on thermal and tensile properties. E-Polymers
15(3):169–177
Barreto ACH, Junior AEC, Freitas JEB, Rosa DS, Barcellos WM, Freire FNA, Fechine PBA,
Mazzetto SE (2013) Biocomposites from dwarf-green Brazilian coconut impregnated with
cashew nut shell liquid resin. J Compos Mater 47(4):459–466
Bettini SHP, Biteli AC, Bonse BC, Morandim-Giannetti A d A (2015) Polypropylene composites
reinforced with untreated and chemically treated coir: effect of the presence of compatibilizer.
Polym Eng Sci 55(9):2050–2057
Bhagat VK, Biswas S, Dehury J (2014) Physical, mechanical, and water absorption behavior of
coir/glass fiber reinforced epoxy based hybrid composites. Polym Compos 35(5):925–930
Cheng F, Hu Y, Yuan J (2014) Preparation and characterization of glass fiber-coir hybrid compos-
ites by a novel and facile Prepreg/Press process. Fibers Polym 15(8):1715–1721
Chollakup R, Smitthipong W, Kongtud W, Tantatherdtam R (2013) Polyethylene green composites
reinforced with cellulose fibers (coir and palm fibers): effect of fiber surface treatment and fiber
content. J Adhes Sci Technol 27(12):1290–1300
Cisneros-López EO, González-López ME, Pérez-Fonseca AA, González-Núñez R, Rodrigue D,
Robledo-Ortíz JR (2017) Effect of fiber content and surface treatment on the mechanical prop-
erties of natural fiber composites produced by rotomolding. Compos Interf 24(1):35–53
Das G, Biswas S (2016a) Effect of fiber parameters on physical, mechanical and water absorption
behaviour of coir fiber–epoxy composites. J Reinf Plast Compos 35(8):644–653
Das G, Biswas S (2016b) Physical, mechanical and water absorption behaviour of coir fiber
reinforced epoxy composites filled with Al2O3 particulates. IOP Conf Ser Mater Sci Eng
115(1):012012
Davis D (2016) Fabrication and mechanical behaviour study of coconut coir based polymer com-
posite. Int J Adv Res 4(4):1267–1272
de Carvalho Benini KCC, Brocks T, Montoro SR, Odila Hilário Cioffi M, Jacobus Cornelis
Voorwald H (2017) Effect of fiber chemical treatment of nonwoven coconut fiber/epoxy com-
posites adhesion obtained by RTM process. Polym Compos 38(11):2518–2527
de Farias JGG, Cavalcante RC, Canabarro BR, Viana HM, Scholz S, Simão RA (2017) Surface lig-
nin removal on coir fibers by plasma treatment for improved adhesion in thermoplastic starch
composites. Carbohydr Polym 165:429–436
de Oliveira PF, Marques M d FV (2014) Comparison between coconut and curaua fibers chemi-
cally treated for compatibility with PP matrixes. J Reinf Plast Compos 33(5):430–439
Dhal JP, Mishra SC (2013) Investigation of dielectric properties of a novel hybrid polymer com-
posite using industrial and biowaste. J Polym Compos 1(1):22–27
Diao C, Dowding T, Hemsri S, Parnas RS (2014) Toughened wheat gluten and treated coconut
fiber composite. Compos A: Appl Sci Manuf 58:90–97
Dong Y, Ghataura A, Takagi H, Haroosh HJ, Nakagaito AN, Lau K-T (2014) Polylactic acid (PLA)
biocomposites reinforced with coir fibres: evaluation of mechanical performance and multi-
functional properties. Compos A: Appl Sci Manuf 63:76–84
Duan J, Wu H, Fu W, Hao M (2018) Mechanical properties of hybrid sisal/coir fibers reinforced
polylactide biocomposites. Polym Compos 39:E188–E199
Frollini E, Bartolucci N, Sisti L, Celli A (2013) Poly (butylene succinate) reinforced with different
lignocellulosic fibers. Ind Crop Prod 45:160–169
Guo F, Wang N, Cheng Q, Hou L, Liu J, Yu Y, Zhao Y (2016) Low-cost coir fiber composite with
integrated strength and toughness. ACS Sustain Chem Eng 4(10):5450–5455
Hamouda T, Hassanin AH, Kilic A, Candan Z, Safa Bodur M (2017) Hybrid composites from coir
fibers reinforced with woven glass fabrics: physical and mechanical evaluation. Polym Compos
38(10):2212–2220
Haque MM, Islam MN (2013) A study on the mechanical properties of urea-treated coir reinforced
polypropylene composites. J Thermoplast Compos Mater 26(2):139–155
Haque MM, Ali ME, Hasan M, Islam MN, Kim H (2012a) Chemical treatment of coir fiber rein-
forced polypropylene composites. Ind Eng Chem Res 51(10):3958–3965
References 183

Haque MM, Islam MS, Islam MN (2012b) Preparation and characterization of polypropylene
composites reinforced with chemically treated coir. J Polym Res 19(5):9847–9855
Hariprasad T, Dharmalingam G, Praveen Raj P (2013) A study of mechanical properties of banana–
coir hybrid composite using experimental and fem techniques. J Mech Eng Sci 4:518–531
Hemsri S, Grieco K, Asandei AD, Parnas RS (2012) Wheat gluten composites reinforced with
coconut fiber. Compos A: Appl Sci Manuf 43(7):1160–1168
Ibrahem RA (2016) Friction and wear behaviour of fibre/particles reinforced polyester composites.
Int J Adv Mater Res 2:22–26
Islam MS, Ahmad MB, Hasan M, Aziz SA, Jawaid M, Haafiz MKM, Zakaria SAH (2015a) Natural
fiber-reinforced hybrid polymer nanocomposites: effect of fiber mixing and nanoclay on physi-
cal, mechanical, and biodegradable properties. Bioresources 10(1):1394–1407
Islam MS, Hasbullah NAB, Hasan M, Talib ZA, Jawaid M, Haafiz MKM (2015b) Physical,
mechanical and biodegradable properties of kenaf/coir hybrid fiber reinforced polymer nano-
composites. Mater Today Commun 4:69–76
Islam MS, Talib ZA, Hasan M, Ramli I, Haafiz MKM, Jawaid M, Islam A, Inuwa IM (2017)
Evaluation of mechanical, morphological, and biodegradable properties of hybrid natural fiber
polymer nanocomposites. Polym Compos 38(3):583–587
Jayabal S, Velumani S, Navaneethakrishnan P, Palanikumar K (2013) Mechanical and machinability
behaviors of woven coir fiber-reinforced polyester composite. Fibers Polym 14(9):1505–1514
Kakou CA, Essabir H, Bensalah M-O, Bouhfid R, Rodrigue D, Qaiss A (2015) Hybrid composites
based on polyethylene and coir/oil palm fibers. J Reinf Plast Compos 34(20):1684–1697
Kumar Rao D, Chandra Gope P (2015) Fracture toughness of walnut particles (Juglans regia L.)
and coconut fiber-reinforced hybrid biocomposite. Polym Compos 36(1):167–173
Kumar SMS, Duraibabu D, Subramanian K (2014) Studies on mechanical, thermal and dynamic
mechanical properties of untreated (raw) and treated coconut sheath fiber reinforced epoxy
composites. Mater Des 59:63–69
Lomelí-Ramírez MG, Kestur SG, Manríquez-González R, Iwakiri S, de Muniz GB, Flores-­
Sahagun TS (2014) Bio-composites of cassava starch-green coconut fiber: part II—structure
and properties. Carbohydr Polym 102:576–583
Luz d, Santos F, Ramos FJHTV, Nascimento LFC, Ben-Hur da Silva Figueiredo A, Monteiro SN
(2018) Critical length and interfacial strength of PALF and coir fiber incorporated in epoxy
resin matrix. J Mater Res Technol 7(4):528–534
Mir SS, Nafsin N, Hasan M, Hasan N, Hassan A (2013) Improvement of physico-mechanical
properties of coir-polypropylene biocomposites by fiber chemical treatment. Mater Des
(1980–2015) 52:251–257
Muthu J, Priscilla J, Odeshi A, Kuppen N (2018) Characterisation of coir fibre hybrid composites
reinforced with clay particles and glass spheres. J Compos Mater 52(5):593–607
Narendar R, Dasan KP (2013) Effect of chemical treatment on the mechanical and water absorp-
tion properties of coir pith/nylon/epoxy sandwich composites. Int J Polym Anal Charact
18(5):369–376
Narendar R, Dasan KP (2015) Development of coir pith based hybrid composite panels with
enhanced water resistant behavior. Environ Prog Sustain Energy 34(5):1481–1487
Narendar R, Priya Dasan K, Rajendran K (2018) Coir pith/nylon/epoxy hybrid composites and
their thermal properties: Thermogravimetric analysis, thermal ageing, and heat deflection tem-
perature. J Vinyl Addit Technol 24(4):297–303
Nelson E, Dagwa IM, Mudiare E (2015) Effect of Fiber length on the mechanical properties
of hybrid Ceasar weed/coir reinforced polyester composite. Usak University J Mater Sci
4(2):29–36
Nuthong W, Uawongsuwan P, Pivsa-Art W, Hamada H (2013) Impact property of flexible epoxy
treated natural fiber reinforced PLA composites. Energy Procedia 34:839–847
Obasi HC, Chaudhry AA, Ijaz K, Akhtar H, Malik MH (2018) Development of biocomposites from
coir fibre and poly (caprolactone) by solvent casting technique. Polym Bull 75(5):1775–1787
Patil Y, Sharma S (2018) Effect of coir fiber reinforcement on mechanical properties of vulcanized
natural rubber composites. Sci Eng Compos Mater 25(3):517–528
184 7 Composites from Coir Fibers

Pérez-Fonseca AA, Arellano M, Rodrigue D, González-Núñez R, Robledo-Ortíz JR (2016a) Effect


of coupling agent content and water absorption on the mechanical properties of coir-agave
fibers reinforced polyethylene hybrid composites. Polym Compos 37(10):3015–3024
Pérez-Fonseca AA, Robledo-Ortíz JR, González-Núñez R, Rodrigue D (2016b) Effect of thermal
annealing on the mechanical and thermal properties of polylactic acid-cellulosic fiber biocom-
posites. J Appl Polym Sci 133(31):43750–43759
Prasad GLE, Gowda BSK, Velmurugan R (2017) A study on impact strength characteristics of coir
polyester composites. Procedia Eng 173:771–777
Prasad N, Agarwal VK, Sinha S (2018) Hybridization effect of coir fiber on physico-­mechanical
properties of polyethylene-banana/coir fiber hybrid composites. Sci Eng Compos Mater
25(1):133–141
Prithivirajan R, Jayabal S, Bharathiraja G (2015) Bio-based composites from waste agricultural
residues: mechanical and morphological properties. Cellul Chem Technol 49(1):65–68
Prithivirajan R, Jayabal S, Kalayana Sundaram S, Pravin Kumar A (2016a) Hybrid bio com-
posites from agricultural residues: mechanical and thickness swelling behavior. Cellulose
35(31):27–31
Prithivirajan R, Jayabal S, Sundaram SK, Sangeetha V (2016b) Hybrid biocomposites from
agricultural residues: mechanical, water absorption and tribological behaviors. J Polym Eng
36(7):663–671
Rajesh R, Ravichandran YD, Nambi Raj NA, Senthilkumar N (2014) Development of a biodegrad-
able composite (hydroxyapatite-chitosan-coir pith) as a packing material. Polym-Plast Technol
Eng 53(11):1105–1110
Ramírez MGL, Satyanarayana KG, Iwakiri S, de Muniz GB, Tanobe V, Flores-Sahagun TS (2011)
Study of the properties of biocomposites. Part I. Cassava starch-green coir fibers from Brazil.
Carbohydr Polym 86(4):1712–1722
Ramprasath R, Jayabal S, Kalyana Sundaram S, Bharathiraja G, Munde YS (2016) Investigation
on impact behavior of rice husk impregnated Coir-Vinyl Ester composites. Macromol Symp
361(1):123–128
Rejeesh CR, Saju KK (2018) Effect of fire-retardant treatment on mechanical properties of
medium-density coir composite boards. Wood Fiber Sci 50(1):113–118
Riyajan S-A, Tangboriboonrat P (2014) Novel composite biopolymers of sodium alginate/natural
rubber/coconut waste for adsorption of Pb (II) ions. Polym Compos 35(5):1013–1021
Roy JK, Nousin Akter HU, Zaman KMA, Sultana S, Shahruzzaman NK et al (2014) Preparation
and properties of coir fiber-reinforced ethylene glycol dimethacrylate-based composite.
J Thermoplast Compos Mater 27(1):35–51
Sakthivel M, Vijayakumar S, Ramesh S (2014) Production and characterization of luffa/coir rein-
forced polypropylene composite. Procedia Mater Sci 5:739–745
Santos d, Cesar J, Siqueira RL, Vieira LMG, Freire RTS, Mano V, Panzera TH (2018) Effects
of sodium carbonate on the performance of epoxy and polyester coir-reinforced composites.
Polym Test 67:533–544
Sarasini F, Tirillo J, Puglia D, Kenny JM, Dominici F, Santulli C, Tofani M, De Santis R (2015)
Effect of different lignocellulosic fibres on poly (ε-caprolactone)-based composites for poten-
tial applications in orthotics. RSC Adv 5(30):23798–23809
Sari PS, Spatenka P, Jenikova Z, Grohens Y, Thomas S (2015) New type of thermoplastic bio
composite: nature of the interface on the ultimate properties and water absorption. RSC Adv
5(118):97536–97546
Sathishkumar TP, Navaneethakrishnan P, Shankar S, Kumar J (2013) Mechanical properties of
randomly oriented snake grass fiber with banana and coir fiber-reinforced hybrid composites.
J Compos Mater 47(18):2181–2191
Saw SK, Sarkhel G, Choudhury A (2012) Preparation and characterization of chemically
modified jute–coir hybrid fiber reinforced epoxy novolac composites. J Appl Polym Sci
125(4):3038–3049
References 185

Staffa LH, Agnelli JAM, de Souza ML, Bettini SHP (2017) Evaluation of interactions between
compatibilizers and photostabilizers in coir fiber reinforced polypropylene composites. Polym
Eng Sci 57(11):1179–1185
Stalin B, Athijayamani A (2016) The performance of bio waste fibres reinforced polymer hybrid
composite. Int J Mater Eng Innov 7(1):15–25
Suardana NPG, Lokantara IP, Lim JK (2011) Influence of water absorption on mechanical prop-
erties of coconut coir fiber/poly-lactic acid biocomposites. Mater Phys Mech 12(2):113–125
Sun Z, Zhang L, Liang D, Xiao W, Lin J (2017) Mechanical and thermal properties of PLA
biocomposites reinforced by coir fibers. Int J Polym Sci 2017:2178329. https://doi.
org/10.1155/2017/2178329
Thakur K, Kalia S, Kaith BS, Pathania D, Kumar A, Thakur P, Knittel CE, Schauer CL, Totaro G
(2016) The development of antibacterial and hydrophobic functionalities in natural fibers for
fiber-reinforced composite materials. J Environ Chem Eng 4(2):1743–1752
Thye TK, Tahir MFBM, Ismail AR, Nor MJM (2012) Automotive noise insulation composite
panel using natural fibres with different perforation areas. Appl Mech Mater 165:63–67
Ujianto O, Noviyanti R, Wijaya R, Ramadhoni B (2017) Effect of maleated natural rubber on ten-
sile strength and compatibility of natural rubber/coconut coir composite. IOP Conf Ser Mater
Sci Eng 223(1):012014
Valášek P, D’Amato R, Müller M, Ruggiero A (2018) Mechanical properties and abrasive wear of
white/brown coir epoxy composites. Compos Part B 146:88–97
Vijayakumar S, Nilavarasan T, Usharani R, Karunamoorthy L (2014) Mechanical and microstruc-
ture characterization of Coconut spathe fibers and Kenaf bast fibers reinforced epoxy polymer
matrix composites. Procedia Mater Sci 5:2330–2337
Viswanathan R, Gothandapani L (1999) Mechanical properties of coir pith particle board.
Bioresour Technol 67(1):93–95
Wang J, Hu Y (2016) Novel particleboard composites made from coir fiber and waste Banana stem
fiber. Waste Biomass Valoriz 7(6):1447–1458
Yao J, Ma L, Lu W, Tan H (2014) Tensile property analysis and prediction model building for coir
rope reinforced unsaturated polyester composite. BioResources 10(1):697–708
Zainudin ES, Yan LH, Haniffah WH, Jawaid M, Alothman OY (2014) Effect of coir fiber loading
on mechanical and morphological properties of oil palm fibers reinforced polypropylene com-
posites. Polym Compos 35(7):1418–1425
Zaman HU, Beg MDH (2014) Preparation, structure, and properties of the coir fiber/polypropylene
composites. J Compos Mater 48(26):3293–3301
Chapter 8
Miscellaneous Applications for Coir
and Other Coconut By-products

8.1 Synthesis of Cellulose Nanoparticles

Coir fibers have been studied for their potential to generate cellulose nanofibers,
nanoparticles, and nanocrystals. Unripe coconut fibers were ground into powder
and treated with acetosolv solution to remove lignin and obtain pulp. The deligni-
fied pulp was bleached using hydrogen peroxide and sodium hydroxide. Later,
fibers were treated with KOH for 120 min at 90 °C to obtain bleached cellulose pulp
(do Nascimento et al. 2016). Four different methods (Fig. 8.1) were used to extract
nanocellulose from the bleached pulps.
In one of the extraction procedures, acidic hydrolysis was done to the cellulose
pulp by treating with high and low concentrations of sulfuric acid and varying the
treatment time from 45 to 360 min. In another approach, pulp was placed in a steel
jacketed reactor and treated with high power ultrasound at 20 kHz and 1200 W for
20 min. Yield of the cellulose nanocrystals obtained varied from 33 to 60% depend-
ing on the procedure used with acid treatment providing lower yields. Similarly, the
size of the nanocrystals also varied from 116 to 307 nm (length) and 4.9 to 116 nm
(width) (do Nascimento et al. 2016), and crystallinity indexes varied from 68 to 82.
It was concluded that the ultrasound high power sonication provided cellulose nano-
crystals with better properties than the acid hydrolysis approach. Life cycle analysis
also confirmed that ultrasonic approach was environmentally friendlier than using
chemicals. Although the cellulose content in coir fibers is relatively lower compared
to other lignocellulosic fibers, it was estimated that cellulose nanocrystals could be
obtained from the fibers at about $11 per gram which would add substantial value
to the coir fibers (do Nascimento et al. 2016). In a similar approach, pretreated coir
fibers were subject to alkali and acid digestion and later ball milled or ground using
liquid nitrogen. Further treatment with NaOCl and HCl was also done to obtain cel-
lulose nanofibers with diameters between 30 and 90 nm (Krishnan and Ramesh
2013). Increase in crystallinity and cellulose content was observed and the cellulose
nanofibers obtained were considered to be suitable for developing films for water

© Springer Nature Switzerland AG 2019 187


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7_8
188 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.1 Different procedures and steps used to extract nanocellulose from coir fibers (do
Nascimento et al. 2016). Reproduced with permission from Elsevier

purification and other filtration applications. It was also shown that nanocellulose
could be obtained from coir fibers using a combination of chemical treatments and
steam explosion. Coir fibers were treated with 2% NaOH in an autoclave at 20lbs
pressure and temperature of 110–120 °C for 1 h. Residue obtained was delignified
by bleaching with sodium chlorite and again acid hydrolyzed using 5% oxalic acid
for 3 h in an autoclave at a pressure of 20 lbs. Fibers formed were further treated by
mechanical stirring for 6 h to obtain a nanocellulose yield of 23.5% with fiber length
between 40 and 90 nm and average width of 14 ± 5 nm. Cellulose nanofibers
obtained had a crystallinity index of 84.5% and crystal size of 3.4 nm.
In another study, using a similar pulping and ultrasonication approach, cellulose
nanofibers with diameters in the range of 2–4 nm and having high percentage of
crystallinity of 54% were obtained from coir fibers (Yue and Qian 2018). Rheological
studies showed that the nanocellulose had predominantly elastic behavior which
was relatively stable between 20 and 80 °C. Coir nanocellulose was considered to
have better properties than those obtained from cotton. Coir fibers bleached with
chlorite and treated with 50% sulfuric acid were able to form cellulose nanowhis-
kers with length of about 200 nm and width of 10–30 nm (Ikhuoria et al. 2015).
Whiskers obtained had % crystallinity of 79% and a crystallinity index of 64 and
were considered to be suitable for fabrication of materials.
Instead of directly using the fibers, ethanolic extracts from coir fibers were used
to mediate the formation of Ni-Pd nanocrystals (Fig. 8.2) (Elango and Roopan
2015; Elango et al. 2016). To prepare the nanoparticles, coir fibers were powdered
and treated with ether and later extracted using 500 mL of ethanol (Elango and
Roopan 2015) in several cycles. Each cycle of methanol treatment provided about
30 mg of residue. The methanolic residue obtained was mixed with lead acetate
solution with continuous stirring at room temperature to form the nanoparticles.
Average particle size obtained was 47 nm. Although the nanoparticles synthesized
8.1 Synthesis of Cellulose Nanoparticles 189

Fig. 8.2 Process of synthesizing the Ni-Pd nanoparticles using coir fiber extracts (Elango et al.
2016). Reproduced with permission from Elsevier

had lower antimicrobial effect, they had good biodegradation and were also able to
degrade carcinogenic dyes through photocatalytic degradation (Elango and Roopan
2015). In further studies, these metallic nanoparticles were studied for their pesti-
cidal and larvicidal activities. Coir fiber extracts could be effectively used to synthe-
size metallic nanoparticles for biological applications. In another study, Ni-Pd
nanoparticles of 62 nm size were synthesized using coir fiber methanolic extract.
The NPs showed excellent anti-feedent, ovicidal activity and also had oviposition
deterrent properties (Elango et al. 2016).
An environmentally friendly approach was used to extract nanofibers from coir
(Abraham et al. 2013). Coir fibers were treated with 2% alkali for 6 h at a tempera-
ture of 25 °C. Later the fibers were exposed to steam in an autoclave at a pressure of
137 Pa and temperatures between 100 and 150 °C. The alkali treated and steam
exploded fibers were bleached by treating with sodium chlorite solution (NaClO2)
at pH 2.3 for 1 h at 50 °C for complete delignification (Abraham et al. 2013).
Continuing the treatment further, the fibers were soaked with mild acid (5% oxalic
acid) and re-exposed to the steam explosion for 1 h at 137 Pa. Nanofibers obtained
were washed with water and later sonicated (Abraham et al. 2013). Cellulose con-
tent in the fibers progressively increased from 39% in the raw fibers to 93.7% after
final treatment with corresponding decrease in lignin and hemicellulose. A nanofi-
ber yield of 23% was obtained. The dimensions of the nanofibers (Fig. 8.3) formed
190 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.3 A picture of the


cellulose nanofibers
obtained from the coir
fibers (Abraham et al.
2013). Reproduced with
permission from Elsevier

were in the range of 4–100 nm with a zeta potential of −18.3 mV. These nanofibers
were suggested to be suitable for various industrial, biomedical, and nanotechnol-
ogy applications (Abraham et al. 2013).
Similar to preparing nanocrystals, cellulose nanowhiskers were prepared from
white coir fibers using the acetosolv process (Nascimento et al. 2014; do Nascimento
et al. 2018). In this process, the coir fibers were treated with 93% AcOH and 0.3%
HCl in a hydrothermal reactor at 100 °C at 1:10 fiber:solution ratio for 180 min.
After the treatment, the pulp obtained was retreated with 30% sulfuric acid at 60 °C
for 360 min. The mixture was also ultrasonicated using a 500 W disruptor at 90%
power for 5 min to allow the cellulose to form nanowhiskers. About 12% of cellu-
lose nanowhiskers having length of 172 nm and width of 8 nm were obtained
(Fig. 8.4). The cellulose nanowhiskers had crystallinity index of 82% and crystallite
size of 5 nm. Micro- and nanofibrillar cellulose was extracted from coir fibers
through autohydrolysis and microemulsion treatments (Tripathi et al. 2017). For the
fibrillation to occur, fibers were treated with distilled water and digested at 180 °C
for 60 min. A microemulsion treatment was done on the treated fibers using an
anionic surfactant and R-limonene. Fibrillation was done by subjecting the treated
fibers to microfluidization 20 times in an intensifier pump at a pressure of 2000 bar.
Figure 8.5 shows the morphological changes to the coir fibers due to the defibrilla-
tion process. Fibers obtained had diameter in the range of 20–70 nm and 1–3 μm.
Crystallinity of the fibers was not affected but substantial differences were observed
in the thermal behavior of the fibers due to the defibrillation (Tripathi et al. 2017).
Coir fibers are one of the highest lignin (up to 35%) containing biomasses (Juikar
and Vigneshwaran 2017). A study was conducted to know the possibility of extract-
8.2 Carbon Nanospheres and Nanotubes 191

Fig. 8.4 Transmission electron microscopy images show the morphology of the cellulose nanofi-
bers (Nascimento et al. 2014). Reproduced with permission from Elsevier

ing nanolignins from coir fibers by treating them initially with NaOH solution at
170 °C followed by acid precipitation (pH 2.0). The lignin obtained was subject to
high shear homogenization at 10,000 rpm for 60 min and ultrasonication at 30 W
and 37 kHz for 60 min, resulting in the formation of nanolignins. In addition to the
chemical approach, a microbial technique was also used for lignin extraction. Bulk
lignin obtained after alkali treatment of fibers was used as a carbon source to grow
Aspergillus sp. at 31 °C for 15 days. Nanolignin formed in the solution was filtered
and collected for further analysis (Juikar and Vigneshwaran 2017). Amount of nano-
lignin obtained varied from 1 to 50% depending on the activity of microbial degra-
dation and biomass yield (Fig. 8.6). Size of the nanoparticles obtained using
microbial hydrolysis was lower (22 nm) compared to homogenization or ultrasoni-
cation (26–35 nm). However, ultrasonication provided nanolignins with the small-
est crystallite size (1.5 nm) and microbial hydrolysis the highest (3.9 nm).
Microbially extracted lignins also had good thermal stability and were suggested to
be suitable for use in textile, biomedical, and environmental applications (Juikar
and Vigneshwaran 2017).

8.2 Carbon Nanospheres and Nanotubes

Carbon nanotubes and nanospheres have also been synthesized using activated car-
bon from coir fiber. Coir was converted into carbon by treating at 700–1100 °C and
later activated by treating with ethanol vapor. Nanospheres were obtained when
higher temperatures were used but nanotubes were obtained between 800 and
900 °C. Nanosphere diameters were dependent on the temperature of carbonization
and varied from 40 to 250 nm. The nanotubes obtained from coir carbon were found
192 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.5 Dimensions of the cellulose nanofibers and their structural arrangement obtained using
autohydrolysis of coir fibers (Tripathi et al. 2017). Reproduced with permission from American
Chemical Society
8.3 Synthesis of Silver Nanoparticles 193

Fig. 8.6 Dependence on the nanolignin yield and LiP activity on the duration of hydrolysis (Juikar
and Vigneshwaran 2017). Reproduced with permission from Elsevier

to have crystallinity and thermal behavior similar to that of commercially available


CNTs (Adewumi et al. 2018). Coir dust was able to form unique hollow carbon
nanostructures through a hydrothermal process (Barin et al. 2014). The dust was
mixed with clay and heated in an autoclave at 250 °C for 4 h at a heating rate of
10 °C/min. Later, a hydrothermal treatment was conducted using concentrated
hydrofluoric acid at 150 °C for 2 h to assist in demineralization. It was observed that
nanocages were formed with organic materials on the walls. Presence of clay par-
ticles and chemical etching significantly affected the morphology. Each carbon cage
formed had diameter between 10 and 20 nm (Fig. 8.7) (Barin et al. 2014).

8.3 Synthesis of Silver Nanoparticles

In a novel study, extracts obtained after washing coir with distilled water was used
to extract silver nanoparticles from silver nitrate. In this approach, aqueous silver
nitrate was added into coir fiber extract at room temperature with constant stirring
at 200 rpm (Roopan et al. 2013) at different pHs and incubation at 60 °C. The mor-
phology of the silver nanoparticles and their size distribution are shown in Fig. 8.8.
Nanoparticles had an average diameter of 23 nm and were mostly spherical. It was
necessary to maintain a pH of 11 to obtain uniform and homogenously distributed
particles compared to pH 2 where no particle formation was observed. The synthe-
sized nanoparticles showed excellent larvicidal activity with LC50 values of 87 mg/L
and 49.8 mg/L for A. stephensi and C. quinquefasciatus, respectively. The
194 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.7 Morphology of the carbon nanocages obtained from coir dust after hydrothermal treat-
ment (a, b) and after acid etching (c–f) (Barin et al. 2014). Reproduced with permission from
Springer Nature

corresponding LC90 values were 231 and 84 mg/L. The extent of larvicidal activity
was considered to be suitable for using the silver nanoparticles for treating malaria
(Roopan et al. 2013). In another study, coir fibers were used to synthesize silver
nanoparticles using transesterification. In this process, coir fibers were refluxed with
butylacrylate in acetone using pyridine as catalyst at 50 °C for 6 h (Rout et al. 2017).
Further, the transesterified fibers were cured with benzoyl peroxide or transesteri-
fied coir. After modifications, the fibers were reacted with silver nitrate (AgNO3) to
coat the fibers with the silver nanoparticles by stirring in the dark for 0.5 h. The size
of silver nanoparticles absorbed on to the fibers was in the range of 40–210 nm.
FTIR studies showed that there was a weak chemical bonding between oxygen
atoms of coir and silver nanoparticles. Higher level of coating was achieved on the
transesterified fibers compared to the untreated coir. The coated fibers were sug-
gested to be suitable for fabrication of novel conducting polymers, electronic and
optical devices, and also many other electrical applications (Rout et al. 2017).

8.4 Sorption of Oil

Oil spills cause considerable environmental damage and many materials have
been tried as sorbents for oil removal. Unmodified coir fibers and those modified
with butyl acrylate to different grafting levels were studied for their potential for
8.4 Sorption of Oil 195

Fig. 8.8 Morphological images of silver nanoparticles and their chemical composition. Scanning
electron image (a), elemental composition (b), TEM image (c) and distribution of nanoparticle size
(d) (Roopan et al. 2013). Reproduced with permission from Elsevier

oil removal. Sorption of up to 8–13 g of oil per gram of coir fiber was obtained
depending on the percentage of grafting (Teli et al. 2017) (Table 8.1). The level of
oil sorption possible using coir fibers is comparatively lower than other biomass.
Also, recyclability of the coir fibers was also poor with about 30% of the oil per-
sistent in the fibers (Teli et al. 2017). In another study, powdered coir fiber was
esterified by treating with N-bromosuccinimide prepared using DMaC/LiCl and
heated up to 100 °C (Yusof et al. 2015). Later, oleoyl chloride was added and reac-
tion continued for 4 h for the esterification to complete. Coir fibers with 30%
grafting provided the most optimum sorption with maximum level being 12 g/g.
A unique approach of using a bacterial consortium immobilized onto coco peat
was studied for the potential to sorb oil from seawater. Morphological images
(Fig. 8.9) showed considerable deposits of oil on the surface of the coir peat
treated with bacteria (Nunal et al. 2014). After 60 days in seawater, there was a
significant increase in the degradation of the sorbed oil for the coir peat contain-
ing bacterial consortium compared to the untreated peat. In a similar approach,
196 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.1 Ability of coir fibers to sorb oil depending on the percentage of grafting (Teli et al.
2017)
Time, Temp, Monomer to fiber Initiator, Crosslinker, Graft add-on, Oil sorption,
h °C ratio % % % g/g
2 70 1.5:1 0.1 0.05 13.86 10.68
3 70 1.5:1 0.1 0.05 15.53 13.45
4 70 1.5:1 0.1 0.05 15.15 13.31
3 60 1.5:1 0.1 0.05 12.75 9.39
3 70 1.5:1 0.1 0.05 15.50 13.45
3 80 1.5:1 0.1 0.05 15.31 13.12
3 70 1:1 0.1 0.05 10.28 8.22
3 70 1.5:1 0.1 0.05 15.54 13.45
3 70 2:1 0.1 0.05 15.58 13.31
Reproduced with permission from Taylor and Francis

Fig. 8.9 SEM images of the rice hull (a, c) and coco peat (b, d) used to sorb oil from water before
and after modifications (Nunal et al. 2014). Reproduced with permission from The Society for
Antibacterial and Antifungal Agents through open access license

ability of coco peat to degrade oil in seawater in the presence of six different
microorganisms was studied (Nunal et al. 2014). A relatively low oil biodegrada-
tion of 36% was obtained from the coir pith. Although the degradation was caused
8.5 Cooling Pads 197

by the microorganisms, it was suggested that a carrier such as peat was necessary
to obtain efficient degradation of the oil (Nunal et al. 2014).
Laboratory scale adsorption experiments were conducted to determine the ability
of coir fibers to sorb vegetable oil and biodiesel (Ifelebuegu and Momoh 2015).
Sorption levels for vegetable oil and diesel were similar and ranged from 4.5 to
7.5 g per gram of coir depending on the temperature during sorption.

8.5 Cooling Pads

An evaporative cooling pad that is capable of absorbing moisture in the air was
developed using coir fibers as an alternative to paper-based cooling pads (Rawangkul
et al. 2008). The pads were made by combining 2 × 2 × 2 cm pieces coir into boxes
of two different configurations (Fig. 8.10). Cooling and saturation efficiencies of the
pads were determined at different air velocities. A decrease in temperature by about
2–3 °C and an increase in relative humidity between 10 and 15% were observed
(Table 8.2), which was suggested to be suitable for commercial applications. In a
related application, ability of treated and untreated coir to sorb moisture in a desic-
cator was studied. Coir fibers were boiled for 0.5 or 1 h to reduce the moisture
content in the coir to 2%. The pore size, surface area, and specific heat capacity
varied due to the boiling treatment. Water adsorption affinity of coir fibers was
found to be inversely proportional to the water vapor concentration. Also, untreated
coir had the lowest adsorption due to the presence of hydrophobic and non-­cellulosic
components on the surface. Maximum sorption of 39%, 32%, and 23% were
obtained at 25, 35, and 50 °C, respectively (Rawangkula et al. 2010).

Fig. 8.10 Digital images of the coconut coir blocks used for air cooling at commercial establish-
ments (Rawangkul et al. 2008). Reproduced with permission from Taylor and Francis
198 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.2 Changes in the temperature and humidity of air before (in) and after (out) passing
through a coir pad (Rawangkul et al. 2008)
Dry bulb-in, Relative humidity-in, Dry bulb-out, Relative humidity-out,
Month °C % °C %
January 26.3 65.3 23.8 80.9
February 27.9 66.1 25.4 81.4
March 29.4 66.9 26.9 81.8
April 30.4 68.4 28.0 82.7
May 29.6 72.7 27.6 85.2
June 29.2 72.9 27.3 85.4
July 28.8 73.7 26.9 85.9
August 28.6 74.1 26.7 86.1
September 28.2 76.7 26.6 87.6
October 28.0 75.2 26.3 86.7
November 27.1 69.2 24.9 83.2
December 25.8 64.7 23.4 80.5
Reproduced with permission from Taylor and Francis

8.6 Pest Control

Coir fibers used as mulch was found to effectively control B. odoriphaga in several
crops and at the same time also promote the growth of C. attenuata which is a bio-
logical pest control agent. The number of developmental days, body weight, and
length were affected due to the use of coir (Zou et al. 2017).

8.7 Gas Diffusion Layer

Proton exchange membrane fuel cells (PEMFCs) are considered as alternative


power source because they have high power density under rapid load changes. A gas
diffusion layer which primarily facilitates the exchange of hydrogen and oxygen gas
is an integral part of a PEMFC. Woven and non-woven carbon paper or fabrics are
generally used as gas diffusion layers. Carbon obtained from coir fibers was studied
for their ability to act as the gas diffusion layer (Indayaningsih et al. 2016). To
obtain the carbon, coconut husks were pyrolyzed at 900 °C or 1300 °C for 1 h. The
carbon obtained was combined at various ratios with ethyl vinyl acetate or polyeth-
ylene glycol to form the membrane. Additionally, the membranes were coated with
polytetrafluoro ethylene (PTFE) to improve hydrophobicity and mechanical proper-
ties. Membranes obtained had thickness between 200 and 400 μm and pore diame-
ters between 1 and 30 μm. Electrical conductivity of the membranes increased from
about 8–55 S/m depending on the ratio of carbon and temperature of carbonization.
Properties of the membranes were considered to be ideal for replacing the gas dif-
fusion layer in current use (Indayaningsih et al. 2016).
8.10 Paper/Boards from Coir 199

8.8 Fire Resistant Coatings and Biosorption

A new method of preparing fire resistant coatings called “intumescent” was devel-
oped using coir fibers to prevent steel structures from fire damage (Souza et al.
2016). Performance of the coir fiber coating was compared with peach stone and
wood waste using epoxy, melamine, triphenyl phosphate, and boric acid. Addition
of the coir fiber coating increased the thermal insulation and was considered eco-
nomical and suitable to improve the fire safety of materials (Souza et al. 2016).
A biofilter was developed using coir as the substrate in a batch and continuous
reactor for degradation of toluene with T. asperellum as the microorganism. A maxi-
mum elimination of 257 gm−3 h−1 could be achieved with a column packing height
of 100 cm and empty bed resistance time of 5 min (Gopinath et al. 2016). It was
suggested that coir could be an inexpensive substrate for degradation of toluene and
development of bioreactors.
Raw and activated coir fibers were used as absorbents to remove uranium, ameri-
cium, and cesium from radioactive liquid organic waste (Ferreira et al. 2018). Coir
fibers were powdered into particles of 0.297–0.500 mm and also activated by treat-
ing with HNO3 and with NaOH (0.75 mol/L). Activated coir had a density of 1.6 g/
cm3 compared to 1.7 g/cm3 before treatment but a much higher surface area (28%
higher) leading to considerably higher sorption levels for the fibers after treatment
(Table 8.3), which were considered to be suitable for practical applications (Ferreira
et al. 2018).

8.9 Wood Vinegar

In a unique study, possibility of producing wood vinegar from coconut shells and
coir was explored (Wititsiri 2011). Different combinations (Table 8.4) of coconut
shells and coir were used to prepare the vinegar by carbonization in a 200 L fuel
tank by treating at temperatures between 300 and 400 °C. Termiticidal and pesti-
cidal activity of the wood vinegar were determined using termites and Ferrisia vir-
gata (mealy bugs), respectively. Some of the properties of the vinegar are listed in
Table 8.4. It was suggested that wood vinegar prepared from coconut shells and
fibers could be suitable for treating termites and bugs in wood and housing con-
struction (Wititsiri 2011).

8.10 Paper/Boards from Coir

Most lignocellulosic sources have been studied for production of pulp for paper.
However, there have been relatively less attempts on converting coir into pulp due
to the high lignin content which causes difficulties in delignification. The suitability
200

Table 8.3 Comparison of the biosorption capacity of the raw and activated coir fibers for three different elements found in radioactive waste (Ferreira et al.
2018)
Uranium Am-241 Cs-137
Coir fibers mg/g mmol/g mg/g mmol/g mg/g mmol/g
Unmodified 0.66 ± 0.10 (27.7 ± 4.2) × 10−7 (46.3 ± 7.8) × 10−6 (19.2 ± 3.2) × 10−11 (44.7 ± 8.9) × 10−9 (32.6 ± 6.5) × 10−14
Activated 1.82 ± 0.04 (76.5 ± 1.7) × 10−7 (73.4 ± 5.3) × 10−6 (30.5 ± 2.2) × 10−11 (37.7 ± 4.6) × 10−9 (27.7 ± 3.4) × 10−14
Reproduced with permission from Springer Nature
8 Miscellaneous Applications for Coir and Other Coconut By-products
8.10 Paper/Boards from Coir 201

Table 8.4 Properties of the wood vinegar prepared using various combinations of coir shell and
fibers (Wititsiri 2011)
Specific Quantity, Termite Mealy bugs
Source pH gravity ml mortality, % mortality, %
Coconut shell, 30 kg 2.9 1.02 850 81.7 95.12
Coconut shell + coir fibers 2.5 1.03 696 52.4 95.12
(15:15 kg)
Coconut shell + coir + basil 3.4 1.01 898 51.2 91.9
(10:10:10)
Open access publication

Fig. 8.11 Influence of pulping conditions on the burst index (a) and tensile index (b) of paper
obtained from coir fiber pulp (Main et al. 2015). Reproduced with permission from North Carolina
State University through open access publication

of coir fibers for pulp and paper production was studied by Main et al. Using a soda
anthraquinone pulping approach, coir fibers were treated with alkali concentrations
between 18 and 22% at temperature of 170 °C for 90 °C to 150 min (Main et al.
2015). The pulp obtained was made into handmade paper (60 g/m2). Properties of
the fiber obtained were dependent on the conditions during pulping. A burst index
ranging from 2.6 to 3.2 kPa m2/g and tensile index ranging from 26 to 30 N.m/g
were obtained (Fig. 8.11). Brightness of the paper ranged from 23 to 32% and scat-
tering coefficient was between 28 and 41.7 m2/kg. Using another approach, paper
was produced from coir fibers through chemical and mechanical pulping and later
bleaching using elementary chlorine process involving chlorine dioxide and alkali
extraction (Jani and Rushdan 2014). The process is generally called the DEDED
(chlorine dioxide-extraction-chlorine dioxide-extraction-chlorine dioxide) sequence
to produce pulp with 80% brightness. Some of the properties of the bleached and
unbleached paper obtained from the coir fibers are given (Jani and Rushdan 2014).
The properties of the paper obtained were considered similar to those produced
from other lignocellulosic sources. However, considerably harsh approach has to be
202 8 Miscellaneous Applications for Coir and Other Coconut By-products

followed to get good quality pulp from coir fibers (Jani and Rushdan 2014). Using
the soda-AQ approach, pulp obtained from coir fibers was made into liner boards. A
yield of 49%, viscosity of 11.7 cP, and kappa number of 41 were achieved for the
pulp. Beating the pulp between 1000 and 8000 revolutions increased fiber conform-
ability and external and internal fibrillation. A burst strength of 4.6 kPa m2/g and a
ring crush test of 1.8 Nm2/g were obtained which was above the minimum require-
ment for industrial applications (Main et al. 2015).
Coir-based cardboards were developed using coir and recycled paper box (Bahari
et al. 2013). Coir fibers were treated with 1 M NaOH for 3–5 days and then com-
bined with recycled paper in three different ratios. Paper was formed using a mold
and later drying in natural sunlight for 5 h. Cardboard obtained using materials
soaked for 5 days had the optimum Young’s modulus of 2.25 GPa and was consid-
ered suitable for commercial applications (Bahari et al. 2013). In a unique approach,
binderless fiber boards were developed from unripe coconut husks (Junior et al.
2017). The unripe husks were dried for 5 days and milled to about 4 mm in size.
Coir fiber particles were combined with white coir pith in 30/70 (%w/w) ratio hav-
ing 8% moisture and compressed into boards on an hydraulic press at temperatures
between 210 and 240 °C (Junior et al. 2017) to achieve a density of 1.2 g/cm3
(Fig. 8.12). Temperature of the compression also influenced the moisture uptake
and mechanical properties (Table 8.5) with increasing temperature decreasing the
MOE or MOR. Although the possibility of preparing the boards without any binders
was demonstrated, the mechanical properties particularly at higher moisture may
not be sufficient for practical applications (Junior et al. 2017).

Fig. 8.12 Digital images of the coir fiber boards manufactured at four different temperatures
(Junior et al. 2017). Reproduced with permission from Springer Nature

Table 8.5 Effect of temperature on the sorption ability and mechanical properties (Junior et al.
2017)
Density, Water Modulus Modulus
T °C kg m−3 uptake, % Swelling, % of elasticity, MPa of rupture, MPa
210 1372 ± 31 48 ± 13 49 ± 9 3410 ± 142 18.0 ± 1.4
220 1297 ± 52 27 ± 5 27 ± 5 2323 ± 440 16.6 ± 1.0
230 1294 ± 26 24 ± 7 24 ± 7 1934 ± 294 9.6 ± 2.0
240 1245 ± 7 19 ± 2 19 ± 2 2092 ± 158 12.0 ± 1.5
Reproduced with permission from Springer Nature
8.11 Precursor for Ceramic Production 203

8.11 Precursor for Ceramic Production

Biomorphic SiC ceramics have been developed using coir fibers as precursors. Coir
fibers having cellulose content of 44% and lignin content of 45% were cut into length
of 1–1.5 cm and made into fiber boards using cellulose acetate as binder. Later, the
boards were heated at 135 °C and pressure of 130 kg/cm2. These precursors were
later pyrolyzed at 800 °C under nitrogen atmosphere and further heated up to 1550 °C
along with molten silica to form the SiC ceramics. Images of the ceramics formed
are shown in Fig. 8.13. Ceramics that are about 98% dense and having SiC and Si
crystalline phases were formed with a flexural strength of 166 MPa and Young’s
modulus of 167 GPa (Maity et al. 2012). Another study reported that ceramics
obtained from coir fiber boards had a density of 2.7 ± 0.06 g/cm3 and a porosity of
8.98% when composed of 67% of SiC and 24% of Si. These ceramics had a flexural
strength of 138 MPa, elastic modulus of 151 GPa, and fracture toughness of 2.6 MPa.
Diffraction studies of the coir-based ceramics showed presence of cubic SiC and Si
in crystalline phases (Fig. 8.14). In terms of lattice parameters, the ceramics had cell
volume of 80.9 Å3 and 160 Å3 for the two phases, respectively. However, a non-
uniform distribution of the two phases was observed with large masses of SiC along
with free areas of Si (Maity et al. 2012). The entire siliconization reaction occurred
exothermically at activation energy of 2793 kJ/mol predominantly through diffusion.
Density and other parameters of the ceramics obtained were directly correlated to the
density of the coir board and the carbon template (Maity et al. 2014).
In a process called reaction engineering, coir fibers were mixed with cellulose
acetate and hot pressed at 135 °C and 130 kg/cm2 pressure to produce precursors.
These precursors were later pyrolyzed at 800 °C under nitrogen atmosphere and the
carbon obtained was reacted with molten silica at 1550 °C to form SiC ceramics
(Maity et al. 2016). Coir-based ceramics had fracture strength of 126 MPa, elastic
modulus of 161 GPa, fracture toughness of 2.5 MPa, and Vickers hardness of
19 GPa. Fracture strength increased at higher temperatures and it was suggested that

Fig. 8.13 Images of the coir fiber board (a), carbon template (b) and SiC ceramic obtained (Maity
et al. 2012). Reproduced with permission from Elsevier
204 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.14 X-ray diffraction pattern of the coir fiber board based SiC ceramics (Maity et al. 2012).
Reproduced with permission from Elsevier

the coir-based SiC ceramics had properties similar to that of SiSiC ceramics (Maity
et al. 2016).
Using a similar method, coir-based ceramics were obtained with fracture strength
of 126 MPa, considerably lower than that of a bamboo fiber board (226 MPa).
Similarly, elastic modulus (161 GPa) was also lower but fracture toughness was
similar (2.6 MPa) and Vickers hardness was comparatively much higher (19.4 GPa)
(Maity et al. 2016). It was suggested that non-uniform microstructure and presence
of pores were responsible for the inferior properties of the coir fiber based ceramics
(Maity et al. 2016).

8.12 Brake Lining

Coir fibers (0–20%) have been used to develop (Fig. 8.15) composite materials for
brake lining/pad applications as alternative to the conventionally used asbestos
which has limited life and is also reported to be a cancer causing agent (Maleque
and Atiqah 2013). Porosity of the composite developed varied from 14 to 22% and
hardness values were between 64 and 26, depending on the extent of coir fibers in
the samples. Study of the wear properties (Fig. 8.16) showed that the 5% coir con-
taining composites had the highest wear resistance, similar to that of the asbestos
containing brake pad (Maleque and Atiqah 2013).
8.12 Brake Lining 205

Fig. 8.15 Schematic of the process used to develop the brake lining/pad from coir fibers (Maleque
and Atiqah 2013). Copyright © 2012, King Fahd University of Petroleum and Minerals. Springer
Nature

Fig. 8.16 Changes in the wear resistance and weight loss of samples containing different levels of
coir fibers (Maleque and Atiqah 2013). Copyright © 2012, King Fahd University of Petroleum and
Minerals. Springer Nature
206 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.6 Properties of brake lining made using coir fiber whiskers as reinforcement (Pai et al.
2017)
Properties Lining with 0% Cold set lining Hot set lining
Density, kg/m3 1235 1228 1241
Tensile strength, MPa 12.3 13.3 21.7
Tensile elongation, % 5.4 6.1 23.4
Vickers hardness 32 37 35
Compressive strength, MPa 34 36.8 47.6
Wear rate (mm3/Nm) 3.03 × 104 2.67 × 104 2.24 × 104
Friction coefficient 0.31 0.27 0.48
Reproduced with permission from Taylor and Francis

In another study, coir fiber whiskers having thickness of 100 μm and length of
16–34 mm were treated with 4% NaOH for 6 h at room temperature and used as
reinforcement for friction lining materials. The liner was made using 10% coir
whiskers using hot and cold settings and properties of the material developed are
shown in Table 8.6 (Pai et al. 2017).

8.13 Superabsorbent Hydrogels

Unique superabsorbent composites were prepared by grafting poly(acrylic acid)


onto coco peat (Su et al. 2016). To prepare the superabsorbent, maleic anhydride
and trimethylolpropane were added into beaker containing coco peat and acrylic
acid and the contents were exposed to UV irradiation at 365 nm for several minutes.
Hydrogels formed (Fig. 8.17) had a maximum swelling of 523 g/g in distilled water
and 40.5 g/g in 0.9 wt% NaCl solution. In addition, the hydrogels were also able to
sorb 325 g/g of urea solution which was released (85%) in about 360 min. It was
suggested that the ability to sorb high amounts of sorbents and slow release made
the hydrogels suitable for many applications (Su et al. 2016).

8.14 Development of SiC Whiskers

Carbon rich tadpole like 3C-SiC whiskers were prepared from coir fibers using Fe
catalyst and spark plasma assisted thermal treatment using silicon slurry waste as the
whisker growth substrate (Cao et al. 2017). Coir fibers were placed over the sub-
strates in a spark plasma system and heated to 1100–1600 °C under argon atmo-
sphere at a heating rate for 100 °C/min for 10 min. Morphology (Fig. 8.18) of the
nanowhiskers obtained was dependent on the temperature at which the nanowhiskers
8.14 Development of SiC Whiskers 207

Fig. 8.17 Images of the dry (a) and superabsorbed (b) acrylic acid-coco peat hydrogels (Su et al.
2016). Reproduced with permission from John Wiley and Sons

were formed. The whiskers were composed of C, Si, and O with atomic percentage
of 63, 37, and 0.09%. The whiskers mainly consisted of single crystal cubic 3C-SiC
structures which particularly grew along the 111 plane with a spacing of 0.25 nm
(Cao et al. 2017). SiC fibers could also be made using coir fibers as template and sili-
con slurry waste using the spark plasma assisted chemical vapor siliconizing
approach (Cao et al. 2018). The silicon slurry waste was subject to various treat-
ments and the coir fibers were cleaned using an ultrasonic bath. To prepare the SiC
fibers, the slurry and fibers were placed into graphite crucible at temperatures
between 1000 and 1650 °C with a heating rate of 100 °C/min and a holding time of
10 min under argon atmosphere (Cao et al. 2018). Coir fibers were found to be com-
pletely carbonized below 1000 °C and composed of amorphous carbon. Temperature
above 1100 °C resulted in the appearance of β-SiC but temperatures above 1100 °C
resulted in the disappearance of the Si phase. Considerable changes were observed
in the morphological features of the carbon and whiskers as the temperature was
increased. At 1100 °C, protuberances with diameter of approximately 5 μm were
found uniformly across the fibers. Heating at temperatures above 1500 °C caused
small holes of 1–2 μm on the fiber surface, and particle size of about 1 μm was
formed above 1600 °C. SiC whiskers formed also had honeycomb like holes of
5–10 μm in size at temperatures between 1100 and 1650 °C but the structures col-
lapsed upon prolonged heating at 1600 °C. A temperature of 1500 °C was found to
be most optimum, and the SiC obtained at this temperature had electromagnetic
wave absorption with minimum reflection loss of −21.9 dB (Cao et al. 2018) and
was considered to be suitable for use in absorbing devices.
208 8 Miscellaneous Applications for Coir and Other Coconut By-products

Fig. 8.18 Morphology of the SiC whiskers obtained at different temperatures: (a) 1100 °C, (b)
1200 °C, (c) 1300 °C, (d) 1400 °C, (e) 1500 °C, and (f) 1600 °C (Cao et al. 2017). Reproduced
with permission from American Chemical Society

8.15 Corrosion Inhibition

Corrosion of iron and steel is a major concern for industries and infrastructure.
Extracts obtained from coir dust using acetone, methanol, or water were studied for
their potential to inhibit corrosion (Umoren et al. 2012, 2014). Acetone extracts
8.15 Corrosion Inhibition 209

from the coir had good corrosion inhibition for acid-induced corrosion of aluminum
which followed the Langmuir adsorption isotherm. A corrosion rate of (9.9–
4.05) × 10−3 mm/year and inhibition efficiency of 31–59% were observed. Similarly,
the extracts obtained from coir fibers using methanol or water contained O-H, N-H,
and C-O functional groups and also aromatic rings considered to be necessary for
corrosion inhibition. Rate of corrosion and corresponding inhibition efficiency were
dependent on the extract concentration and temperature at which the corrosion was
studied (Fig. 8.19). In another study, it was also found that ethanol and acetone
extracts from coir dust showed corrosion inhibition for mild steel in 0.5 H2SO4 solu-
tion at 30–60 °C (Umoren et al. 2014). The extent of corrosion resistance was
dependent on the type of solvent used and extract concentration and/or presence of
additional ions (Tables 8.7 and 8.8).

Fig. 8.19 Effect of extract concentration and temperature on the corrosion rate (left) and inhibi-
tion efficiency for the ethanol and water extracts from coir (Umoren et al. 2014). Reproduced with
permission from Elsevier
210 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.7 Electrochemical impedance parameters for the coir dust extracts in 0.5 M sulfuric acid
with and without iodide ions (Umoren et al. 2014)

Rs (Ω/ Y0 Rct (Ω/ Cdl ɳ,


System cm2) (ΩSn cm−2) × 10−6 n cm2) (μF cm2) %
Blank 5.16 90.19 0.92 39.09 237.9 –
Ethanol extract (0.5 g/L) 4.76 89.43 0.88 158.6 15.3 75.4
Ethanol extract +1 mM 4.88 23.44 0.88 1248.0 0.26 96.9
KI
Ethanol extract +5 mM 4.59 13.75 0.89 1603.0 0.16 97.6
KI
Acetone extract 5.65 83.58 0.90 103.9 34.4 62.4
(0.5 g/L)
Acetone extract +1 mM 5.19 25.29 0.89 770.0 0.68 94.9
KI
Acetone extract +5 mM 4.61 11.68 0.86 1763.0 0.14 97.8
KI
Reproduced with permission from Elsevier

Table 8.8 Extracts obtained from coir using various solvents and their mild steel inhibition
efficiency (Umoren et al. 2014)
Inhibition efficiencies, %
Extract concentration, Weight loss Impedance Polarization
Solvent g/L method method method
Methanol 0.5 87.4 94.3 66.5
Ethanol 0.5 79.9 75.4 58.9
Acetone 0.5 69.8 62.4 36.8
Water 0.5 45.1 41.7 43.4
Reproduced with permission from Elsevier

8.16 Production of Bio-oil

Coir fibers were subject to fast pyrolysis and the ability to generate bio-oil was studied
(Almeida et al. 2013). During the process of pyrolysis, the biomass decomposes into
organic material releasing vapors, chars, ashes, and gases. This mixture after cooling
turns into a dark brown and dense liquid called bio-oil. With this approach, biomasses
can be converted into maximum valorizable energy provided appropriate conditions
are used. Pyrolysis of coir fibers was done by heating at 700 °C at a heating rate of
100 °C/min at a N2 flow rate of 1 mL/min and maintaining at that high temperature for
5 min. The bio-oil formed was condensed and used for further analysis. A bio-oil yield
of 10.15% was obtained along with 33.3% residual solids. Further separation using
aqueous phase produced organic matter called AQ-LLE (aqueous extracts obtained
using liquid–liquid extraction). The compositional analysis of the bio-oil is shown in
Fig. 8.20. Composition of the oil and AQ-LLE is given in Table 8.9. Since the major
8.16 Production of Bio-oil 211

Fig. 8.20 Composition of the dried bio-oil obtained after pyrolyzing the coir fibers (a); 3D image
of the enlarged region is shown in figure (Almeida et al. 2013). Reproduced with permission from
American Chemical Society

Table 8.9 Comparison of the major compounds obtained after the pyrolysis of coir fibers (Almeida
et al. 2013)
Chemical class Compound Isomers DBO AP-LLE
Alcohols Furanmethanol 1 3.8 4.1 1.0 0.9
Aldehydes Furfural 1 8.3 10.2 – 0.1
Furfural C1 1 2.5 3.1 – –
Furfural hydroxyl C1 1 2.1 – – –
Vanillin 1 0.4 – 1.1 0.2
Anhydrides Maleic anhydride 1 – 2.0 – –
Lauric anhydride 1 – – 1.1 –
Ketones Cyclopentanone 2 0.4 2.2 – –
Cyclopentanone, C1 2 1.6 2.1 – –
Cyclopentanone, C2 4 0.9 1.5 – –
Cyclopentanedione, C1 2 1.7 – 2.2 1.2
Hexenone 1 – – – –
Ethers Benzofuran, C2 3 1.5 0.7 – –
Benzene, trimethoxy 1 1.0 0.1 1.8 0.4
Benzene, trimethoxy, C1 1 22.6 0.2 0.6 0.2
(continued)
212 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.9 (continued)


Chemical class Compound Isomers DBO AP-LLE
Phenols Phenol 1 7.9 37.6 34.9 37.9
Phenol, C1 3 8.2 6.7 15.6 23.4
Phenol, C2 8 2.7 3.9 9.5 11.0
Phenol, dimethoxy 1 5.2 0.4 5.3 –
Phenol, methoxy 3 2.1 3.6 – 13.9
Phenol, methoxy C1 5 3.0 1.3 3.2 3.6
Phenol, methoxy C2 2 1.2 1.0 2.4 1.3
Phenol, methoxy C4 1 6.8 – – –
Phenol, methoxypropenyl 4 2.0 0.4 3.9 0.4
Phenol, dimethoxypropenyl 3 3.0 – 0.4 –
Benzenediol, C1 4 – 3.9 0.4
Ethanone, furanyl 1 1.6 1.2 – –
Furanone, dihydro 1 1.3 0.2 – –
Hydrocarbons Toluene 1 – 4.6 – 1.6
Heptane 1 – 5.2 – –
Reproduced with permission from American Chemical Society

component of the bio-oil and AQ-LLE was phenols, it was suggested that pyrolysis of
coir fibers could be an promising alternative to obtain new products and chemicals
with minimum harm to the environment (Almeida et al. 2013).

8.17 Manufacture of Conducting Fibers

Coir fibers were coated with poly(aniline) under various conditions and used to
manufacture pressure sensitive devices (de Souza Junior et al. 2014). To chemically
modify the fibers, coir fibers were treated with 5 mL of aqueous H2O2 solution for
6 h at room temperature. Later, the fibers were immersed in sulfuric acid solution
(1 M H2SO4), and 5.37 × 10−3 mol of alanine was added slowly. After cooling the
fibers using an ice bath, 10 mL of H2SO4 containing desired amount of ammonium
persulfate (APS) or Ce(IV) was added and the polymerization was carried out for
24 h (de Souza Junior et al. 2014). Some of the electromechanical properties of the
modified coir fibers are given in Table 8.10. The new method of preparation affected
the electromechanical properties and sonication/Ce(IV) but was able to provide bet-
ter properties than stirring/APS system. It was suggested that the new method devel-
oped could be used for rapid development of fibers for use in pressure sensitive
devices (de Souza Junior et al. 2014). In another approach, coir fibers were made
conductive by in situ oxidative polymerization of aniline in the presence of iron (III)
chloride hexahydrate (FeCl3. 6H2O) or (APS) as an oxidant (Merlini et al. 2014).
8.17 Manufacture of Conducting Fibers 213

Table 8.10 Electromechanical properties of conducting fibers prepared from coir fibers using
various treatments (Merlini et al. 2014)
Ce(IV)/stirring Ce(IV)/sonication
ScompmaxC ScompmaxE ScompmaxC ScompmaxE
Cycle (% MPa) (% MPa) h (%) (% MPa) (% MPa) h (%)
1 90 101 36 421 441 108
2 86 90 21 312 326 60
3 79 83 18 248 260 43
4 72 75 14 212 221 29
5 67 68 13 187 193 25
6 62 64 13 171 175 20
7 58 61 10 157 160 20
8 55 57 10 147 149 16
9 53 54 10 138 142 13
10 51 52 10 130 134 15
Avg 67 ± 14 71 ± 16 15 ± 8 212 ± 92 220 ± 98 35 ± 30
APS/stirring APS/sonication
ScompmaxC ScompmaxE ScompmaxC ScompmaxE
Cycle (% MPa) (% MPa) h (%) (% MPa) (% MPa) h (%)
1 507 565 226 161 182 77
2 466 516 151 142 158 48
3 440 475 132 126 139 41
4 422 449 123 117 130 37
5 406 431 104 107 120 34
6 394 415 103 101 107 32
7 383 401 91 93 102 27
8 376 398 83 87 99 27
9 365 385 84 83 90 27
10 357 372 75 79 85 23
Avg 412 ± 48 441 ± 62 117 ± 45 110 ± 27 121 ± 31 37 ± 16
Reproduced with permission from Elsevier

The conductive fibers prepared were used as conductive fillers in polyurethane com-
posites with 5–25 wt% of fibers. Some of the performance parameters including
electrical conductivity are given in Table 8.11. Substantial increase in conductivity
was observed in the composites. The conductivity was observed to be dependent on
the extent of PANI. An electrical conductivity of 1.5 × 10−1 and 1.9 × 10−2 S/cm was
obtained. Interestingly, the electrical resistivity of the materials varied with com-
pressive stress (Fig. 8.21), indicating that they can be used for pressure sensitive
applications (Merlini et al. 2014).
214 8 Miscellaneous Applications for Coir and Other Coconut By-products

Table 8.11 Thermal behavior and electrical conductivity of the unmodified and PANI-modified
coir fibers (Merlini et al. 2014)
Sample Residue from TGA, % PANI content, % Electrical conductivity (S cm−1)
Coir fibers 29.95 – (3.13 ± 0.27) × 10−10
PANI.FeCl3 49.86 – 4.47 ± 0.39
PANI.APS 57.67 – 1.13 ± 0.23
CF-PANI.FeCl3 36.10 30.9 (1.5 ± 0.39) × 10−1
CF-PANI.APS 56.01 94.2 (1.9 ± 0.35) × 10−2
Reproduced with permission from Elsevier

Fig. 8.21 Changes in the electrical resistivity of the PANI-treated coir fibers (Merlini et al. 2014).
Reproduced with permission from Elsevier

References

Abraham E, Deepa B, Pothen LA, Cintil J, Thomas S, John MJ, Anandjiwala R, Narine SS (2013)
Environmental friendly method for the extraction of coir fibre and isolation of nanofibre.
Carbohydr Polym 92(2):1477–1483
Adewumi GA, Inambao F, Eloka-Eboka A, Revaprasadu N (2018) Synthesis of carbon nanotubes
and Nanospheres from coconut fibre and the role of synthesis temperature on their growth.
J Electron Mater 47(7):3788–3794
Almeida TM, Bispo MD, Cardoso ART, Migliorini MV, Schena T, de Campos MCV, Machado
ME, López JA, Krause LC, Caramão EB (2013) Preliminary studies of bio-oil from fast pyroly-
sis of coconut fibers. J Agric Food Chem 61(28):6812–6821
References 215

Bahari AR, Fudzi MSM, Ahmad NU, Rahman HA (2013) Material properties characterization of
coir cardboard. ARPN J Eng Appl Sci 10(20):9521–9526
Barin GB, de Fátima Gimenez I, da Costa LP, Filho AGS, Barreto LS (2014) Hollow carbon nano-
structures obtained from hydrothermal carbonization of lignocellulosic biomass. J Mater Sci
49(2):665–672
Cao Y, Xiang D, Li H, Ren R, Xing Z (2017) Growth of SiC whiskers onto carbonizing coir fibers
by using silicon slurry waste. ACS Sustain Chem Eng 5(11):10563–10569
Cao Y, Xiang D, Li H, Xing Z, Zhao S, Fan Z, Ren R (2018) Rapid preparation of SiC fibers by
spark plasma assisted vapor silicon infiltration. J Alloys Compd 740:836–843
de Souza Junior FG, Carlos Pinto J, Alves Garcia F, de Oliveira GE, Bruno Tavares MI, da Silva
AM, Daher Pereira E (2014) Modification of coconut fibers with polyaniline for manufacture
of pressure-sensitive devices. Polym Eng Sci 54(12):2887–2895
do Nascimento DM, Dias AF, Junior CP d A, Rosa M d F, Morais JPS, de Figueirêdo MCB (2016)
A comprehensive approach for obtaining cellulose nanocrystal from coconut fiber. Part II: envi-
ronmental assessment of technological pathways. Ind Crop Prod 93:58–65
do Nascimento DM, Almeida JS, Vale M d S, Leitão RC, Muniz CR, de Figueirêdo MCB et al
(2018) A comprehensive approach for obtaining cellulose nanocrystal from coconut fiber. Part
I: proposition of technological pathways. Indust Crops Prod 93:66–75
Elango G, Roopan SM (2015) Green synthesis, spectroscopic investigation and photocatalytic
activity of lead nanoparticles. Spectrochim Acta A Mol Biomol Spectrosc 139:367–337
Elango G, Roopan SM, Al-Dhabi NA, Arasu MV, Dhamodaran KI, Elumalai K (2016) Coir medi-
ated instant synthesis of Ni-Pd nanoparticles and its significance over larvicidal, pesticidal and
ovicidal activities. J Mol Liq 223:1249–1255
Ferreira RVP, Silva EA, Canevesi RLS, Ferreira EGA, Taddei MHT, Palmieri MC, Silva FRO,
Marumo JT (2018) Application of the coconut fiber in radioactive liquid waste treatment. Int
J Environ Sci Technol 15(8):1629–1640
Gopinath M, Mohanapriya C, Sivakumar K, Baskar G, Muthukumaran C, Dhanasekar R (2016)
Biodegradation of toluene vapor in coir based upflow packed bed reactor by Trichoderma
asperellum isolate. Environ Sci Pollut Res 23(5):4129–4137
Ifelebuegu AO, Momoh Z (2015) An evaluation of the adsorptive properties of coconut husk for oil
spill cleanup. Adsorption 6(7):8–15
Ikhuoria EU, Omorogbe SO, Agbonlahor OG, Etiuma RA (2015) Nanocellulose crystals from coir
fibre for template application. Am Chem Sci J 9(1):1–11
Indayaningsih N, Zulfia A, Priadi D, Hendrana S (2016) Preparation of carbon composite from
coconut fiber for gas diffusion layer. Ionics 22(8):1445–1449
Jani SM, Rushdan I (2014) Effect of bleaching on coir fibre pulp and paper properties. J Trop Agric
Food Sci 42(1):51–61
Juikar SJ, Vigneshwaran N (2017) Extraction of nanolignin from coconut fibers by controlled
microbial hydrolysis. Ind Crop Prod 109:420–425
Junior CPA, Coaquira CAC, Mattos ALA, de Souza M d SM, Feitosa JP d A, de Morais JPS, Rosa
M d F (2017) Binderless fiberboards made from unripe coconut husks. Waste Biomass Valoriz
9(11):2245–2254
Krishnan VN, Ramesh A (2013) Synthesis and characterization of cellulose nanofibers from coco-
nut coir fibers. J Appl Chem 6(3):18–23
Main NM, Talib RA, Ibrahim R, Abdul Rahman R, Mohamed AZ (2015) Linerboard made
from soda-anthraquinone (soda-AQ) treated coconut coir fiber and effect of pulp beating.
Bioresources 10(4):6975–6992
Maity A, Kalita D, Kayal N, Goswami T, Chakrabarti O, Gangadhar Rao P (2012) Synthesis of
biomorphic SiC ceramics from coir fibreboard preform. Ceram Int 38(8):6873–6881
Maity A, Das H, Kalita D, Kayal N, Goswami T, Chakrabarti O (2014) Studies on formation and
siliconization of carbon template of coir fibreboard precursor to SiC ceramics. J Eur Ceram
Soc 34(15):3499–3511
Maity A, Kayal N, Chakrabarti O (2016) Mechanical behaviour of reaction processed SiC ceram-
ics from artificial precursor from plant. Ceram Int 42(8):10058–10065
216 8 Miscellaneous Applications for Coir and Other Coconut By-products

Maleque MA, Atiqah A (2013) Development and characterization of coir fibre reinforced compos-
ite brake friction materials. Arab J Sci Eng 38(11):3191–3199
Merlini C, Barra GMO, Schmitz DP, Ramôa SDAS, Silveira A, Araujo TM, Pegoretti A (2014)
Polyaniline-coated coconut fibers: structure, properties and their use as conductive additives in
matrix of polyurethane derived from castor oil. Polym Test 38:18–25
Nascimento DM, Almeida JS, Dias AF, Figueirêdo MCB, Morais JPS, Feitosa JPA, Rosa M d F
(2014) A novel green approach for the preparation of cellulose nanowhiskers from white coir.
Carbohydr Polym 110:456–463
Nunal SN, Leon SMSS-D, Bacolod E, Koyama J, Uno S, Hidaka M, Yoshikawa T, Maeda H
(2014) Bioremediation of heavily oil-polluted seawater by a bacterial consortium immobilized
in cocopeat and rice hull powder. Biocontrol Sci 19(1):11–22
Pai A, Kini MV, Pokharel V (2017) Influence of a novel hardener p-toluene sulfonic acid
on mechanical and wear response of phenolic-based friction materials. Tribol Trans
60(5):770–780
Rawangkul R, Khedari J, Hirunlabh J, Zeghmati B (2008) Performance analysis of a new sustain-
able evaporative cooling pad made from coconut coir. Int J Sustain Eng 1(2):117–131
Rawangkula R, Khedaria J, Hirunlabhb J, Zeghmatic B (2010) Characteristics and performance
analysis of a natural desiccant prepared from coconut coir. Chem Anal 1269(1):11
Roopan SM, Madhumitha G, Abdul Rahuman A, Kamaraj C, Bharathi A, Surendra TV (2013)
Low-cost and eco-friendly phyto-synthesis of silver nanoparticles using Cocos nucifera coir
extract and its larvicidal activity. Ind Crop Prod 43:631–635
Rout SK, Tripathy BC, Padhi P (2017) Bikash Ranjan Kar, and Krushna Gopal Mishra. A green
approach to produce silver nano particles coated agro waste fibers for special applications. Surf
Interf 7:87–98
Simons KL, Ansar A, Kadali K, Bueti A, Adetutu EM, Ball AS (2012) Investigating the effec-
tiveness of economically sustainable carrier material complexes for marine oil remediation.
Bioresour Technol 126:202–207
Souza d, Milena M, de Sá SC, Zmozinski AV, Peres RS, Ferreira CA (2016) Biomass as the
carbon source in intumescent coatings for steel protection against fire. Ind Eng Chem Res
55(46):11961–11969
Su X, Bai B, Xu X, Ding C, Wang H, Suo Y (2016) Fabrication and properties of a novel superab-
sorbent composite based on coco peat and poly (acrylic acid) cross-linked trimethylolpropane
trimaleate under ultraviolet irradiation. Polym Adv Technol 27(9):1179–1190
Teli MD, Valia SP, Mifta J (2017) Application of functionalized coir fibre as eco-friendly oil sor-
bent. J Text Inst 108(7):1106–1111
Tripathi A, Ferrer A, Khan SA, Rojas OJ (2017) Morphological and thermochemical changes
upon autohydrolysis and microemulsion treatments of coir and empty fruit bunch residual
biomass to isolate lignin-rich micro-and nanofibrillar cellulose. ACS Sustain Chem Eng
5(3):2483–2492
Umoren SA, Eduok UM, Israel AU, Obot IB, Solomon MM (2012) Coconut coir dust extract:
a novel eco-friendly corrosion inhibitor for Al in HCl solutions. Green Chem Lett Rev
5(3):303–313
Umoren SA, Obot IB, Israel AU, Asuquo PO, Solomon MM, Eduok UM, Udoh AP (2014)
Inhibition of mild steel corrosion in acidic medium using coconut coir dust extracted from
water and methanol as solvents. J Ind Eng Chem 20(5):3612–3622
Wititsiri S (2011) Production of wood vinegars from coconut shells and additional materials
for control of termite workers, Odontotermes sp. and striped mealy bugs, Ferrisia virgata.
Songklanakarin J Sci Technol 33(3):349–354
Yue D, Qian X (2018) Isolation and rheological characterization of cellulose nanofibrils (CNFs)
from coir fibers in comparison to wood and cotton. Polymers 10(3):320
Yusof NA, Mukhair H, Malek EA, Mohammad F (2015) Esterified coconut coir by fatty acid chlo-
ride as biosorbent in oil spill removal. Bioresources 10(4):8025–8038
Zou D, Coudron TA, Xu W, Gu X, Wu H (2017) Development of immature tiger-fly Coenosia
attenuata (Stein) reared on larvae of the fungus gnat Bradysia impatiens (Johannsen) in coir
substrate. Phytoparasitica 45(1):75–84
Index

A Atomic absorption spectroscopy, 132


Absorbents, 50 Autohydrolysis, 62
Activated carbon
ability, 111
CO2, 106 B
coconut shells, 108 β-glucosidase, 55
coir pith, 108 Biochars, 36, 37, 95, 97–99, 104
nitrogen/urea-treated, 110 Biocomposites
precursor, 102 PCL, 158–160, 162
ruthenium, 108 PLA, 155–158
Agricultural applications thermal behavior and DMTA analysis, 161
acidic and electrical conductivity, 35 Biodegradable mulch, 40
biochar, 36, 37 Biodegradation, 173
C/N ratio, 35 Biodiesel, 111
coir and by-products, 31 Biofiltration system, 124
coir dust, 31–34 Biofuels, 103
coir pith and sludge, 33, 37 Biological oxygen demand (BOD), 4
cultivation of cauliflower, 36 Bio-oil, 210–212
cultivation of hyacinth, 39 Brake linings, 204–206
degradation of coir, 35 By-products
effect of salt stress, 33 biotechnological applications, 55
electrical conductivity, 36 producing ethanol, 61
geotextile, 43–50 various coconut, 63, 68
grafting efficiency, 50
medicinal plant, 37
mulching, 39–43 C
organic matters, 35 Calorific value, 95, 96, 98, 103, 104
PAA, 50, 51 Carboxyl methyl cellulose (CMCase), 57, 58
perlite, 33, 34 Cellulose nanofibers, 187, 188, 191, 192
physico-chemical properties, 33 Cellulose nanoparticles
plant parameters after adding coir pith, 38 synthesis, 187–190
poly(acrylic acid) based superabsorbent, 51 Ceramics, 203, 204
sedge and sphagnum based peat, 39 Coco peat, 31, 34, 35, 39, 41, 42, 50, 51
3-Amino-propyltriethoxysilane, 165 Coconut coir, 197
Ammonium persulfate (APS), 212 Coconut husk, 10
Apple juice, 70 Coconut sheath, 180, 181

© Springer Nature Switzerland AG 2019 217


N. Reddy, Sustainable Applications of Coir and Other Coconut By-products,
https://doi.org/10.1007/978-3-030-21055-7
218 Index

Coconut shells chemical pretreatments and treatments,


ability, 105 86, 89
carbon, 105, 108–110 and concrete, 85
electrodes, 105 compressive strength, 79, 81, 84, 85
KOH, 107 covered and uncovered roofs, 88
nanosheets, 106 curing times, 85
porous graphitic carbon, 107 electromechanical properties, 212, 213
pyrolysis and steam activation, 109 ethanolic extracts, 188
Coconuts fiber content and water/binder ratio, 87
applications, 2 flax fabric-epoxy strengthened beams, 82
brown and green coir fibers, 4 flexural ductility and mortar toughness, 80
by-products, 1 flexural strength, 83
coir indoor and outdoor walls, 84
acoustic properties, 14, 16 intumescent, 199
chemical extraction, 5–9 levels, 84
chemical modifications, 20, 22–26 long-term performance, 76
chemical treatments, 17–20 mechanical properties of concrete, 83
geographical region, 16, 17 metallic nanoparticles, 189
loading and unloading, 20, 21 methyl ethyl ketone peroxide, 90
microbial extraction, 9, 10 micro- and nanofibrillar cellulose, 190
morphological structure, 11–13 mixed mortars, 75
plasma, 10, 11 morphological studies, 84
production, 4, 5 nanocellulose, 188
properties, coir pith, 25, 27 NaOH solution, 89
thermal properties, 12, 14, 15 NaOH-treated fibers, 80
extraction of fibers, 1, 3 Ni-Pd nanoparticles, 189
parts, 3 PANI-modified, 214
raw/undried, 1 PANI-treated, 214
structural features, 2 performance of unreinforced, 82
structure and property, 3 poly(aniline), 212
Coir polyester resin, 90
fuel, 95–99 and polypropylene, 80
gasification, 103–111 pyrolysis, 210–212
Coir dust, 31, 33–37, 40, 41 reaction engineering, 203
Coir fibers, 41, 45 red sand brick reinforced, 89
ability, 79, 85, 196, 197 resistance, 80, 81, 83
alkali treatment, 76 sand at different u/B ratios, 87
A-, N- and W-series, 83 split tensile strength, 79
application, 75, 88 SPRE, 91
aqueous silver nitrate, 193 static and dynamic loads, 83
autohydrolysis, 192 steel and coir reinforcements, 79
beams, 76, 78, 81 substantial improvement, 82
biofilters, 91 suitability, 199
biomorphic SiC ceramics, 203 tensile properties of concrete, 88
biosorption capacity, 200 3C-SiC whiskers, 206
bleached and unbleached paper, 201 toughness of mortar reinforced, 76
bleached and unbleached pulp, 201 treatment, 188
brake lining/pad, 205, 206 types of coconut husks, 90, 91
butylacrylate, 194 unmodified, 194
carbon, 191 untreated and treated, 77, 80
cellulose content, 187 water adsorption affinity, 197
cellulose nanofibers, 190 wear resistance and weight loss, 205
cement panels, 79 X-ray diffraction pattern, 204
cementitious composites, 75, 76, 78, 88 Coir geotextiles, 43–50
Index 219

Coir mulching, 39–43 Conducting fibers, 212–214


Coir pith carbon (CPC), 99–101, 178, 180 Cooling pad, 197, 198
Coir shell, 178, 180 Corrosion inhibition, 208, 209
Compatibilizers, 144, 145, 153, 159, 163, 167, Czapek–Dox agar broth, 57
168, 171
Composites reinforcement
epoxy resin D
bicomponent, 149 Degree of polymerization (DP), 20
cycloaliphatic polyamine, 150
hydrophilic, 150
mechanical testing, 152 E
non-cellulosic substances, 150 Electrode
properties, white and brown coir alkaline solution, 108
fibers, 151 aqueous and non-aqueous media, 100
scanning electron image, 151 carbon, 105
tensile and flexural strength, 152 coconut shell and biomass, 105
phenol formaldehyde, 152, 153 coir pith, 111
polyester CPC-850, 100
acoustic properties, 142 EDS spectra, 101
alkali/silane solution, 141 gold-coated coir fiber, 101
alkali treatment, 143 gravimetric capacitance, 110
chemical treatments, 142 KOH, 102
coir ropes, 144 supercapacitors, 100
compatibilizer (maleic anhydride), 144 Elemental composition, 159
epoxy and UPE matrix, 143 Environmental remediation
glycolysis, 141 contaminants, 115
natural cellulose fibers, 142 desalination, 136
photostabilizing agents, 144 diesel desulfurization, 137, 138
pretreatments, 141 dyes removal
PVA, 142, 143 adsorption rate, 115, 116
sound absorption coefficient tests, 141 BET analysis, 115
polyethylene cationic and anionic dye, 119
acetic acid, 147 characteristics, coir pith carbon, 118
HDPE, 148 coconut coir dust, 118, 119
LMDPE, 148 electrostatic and complexation
plasma treatment, 148 reactions, 118
polymers, 148 first-order kinetic models, 118
rotomolding approach, 148 holocellulose, 121
polypropylene kinetic parameters, 117
alkali treatment, 145 mesoporous carbon, 115
cellulose dialdehyde, 144 methylene blue, 118, 119
fiber–matrix interfacial interaction, 145 micropores and macropores, 118
granules, 145 multilayer absorption, 120
hemicellulose, 147 second-order kinetic models, 118
maleic acid, 146 SEM images, 120
p-aminophenol, 146 thermodynamic parameters, 120, 121
silane (tetramethoxyorthosilicate), water pollution, 115
145–146 gases sorption, 136
tensile and flexural test, 147 heavy metals removal
rubber, 152–154 acrylic and phosphoric acid, 125
UV treatment, 154 activated carbon, 125
Compressive strength, 75, 76, 79–81, 83–85, adsorption and desorption cycles, 125
87, 89, 90 anion exchanger, 130
Concrete reinforcement, 76, 79, 85 aqueous solution, 129
220 Index

Environmental remediation (cont.) Fire resistant coatings, 199


biofilters, 124 Flexural strength, 76, 79–81, 84, 85
biomasses, 123
cationic exchanger, 131
cellulose and hemicellulose, 128 G
chemical modification, 130 Gas diffusion layer, 198
chemisorption, 130 Geotextile, 43–50
coir fibers, 124, 127 Gibb’s free energy, 129
diffusion process, 132, 133 Glucose yield, 63, 65
efficiencies, 122, 123 Grafting
electroplating waste water, 129 after UV treatment, 22
FTIR and X-ray diffraction, 132 cyanoethylated fibers, 22
gallanium, 122 different temperatures, 22
hydrochloric acid, 124 effect of urea, 23
hydrogen peroxide, 123 furfuryl alcohol, 24
inorganic anions, 122 SEM studies, 23
Langmuir isotherm, 122 Green coconut shell (GCS), 60
lignin, 132 Ground blast-furnace slag (GBFS), 85
mercerization, 131
metal ions, 123, 128, 129
NaOH-treated fibers, 124 H
phosphinates, 132 Hand lay-up techniques, 141, 143, 159, 163, 171
physiochemical characteristics, 129 Hemicellulose, 6
polyacrylamide, 130 Hybrid composites
pseudo-second-order kinetics, 122 agricultural residues, 162
radioactive isotopes, 131 antibacterial activity, PBS, 176
SEM image, 126 ASTM 638 and ASTM 790 standards, 171
sorbents, 128 beads, 177, 178
spectrophotometer, 129 biodegradability, 171
succinic anhydride, 122 biografting, 175
thermal degradation, 130 clay particles, 165
thermodynamic parameters, 131, 132 coir and jute fibers, 168–170
waste water, 126 coir and oil palm, 174, 175
Langmuir and Freundlich isotherms, 133 commingled yarn preparation, 168
p-chlorophenol and p-nitrophenol, 133 components, 163
p-cresol, 134 compression pressures, 174
phenols, 133, 134 coupling agents, 167
polluted water, 115 epoxy resins, 165
Temkin isotherm, 135 eugenol, 176
thermodynamic parameters, 135, 136 fiber/matrix interfacial bonding, 167
Enzymes flexural strength and modulus, 174
cellic Ctec2 and HTec2, 60 glass fabrics, 165
cellulase, 62 hydrophobicity and antimicrobial
production, 55, 57, 58 properties, 175
saccharification, 59 isophthalic polyester resin, 174
Erosion control, 45–47 kenaf–coir novolac resin, 170
Ethanol lignocellulosic sources, 177
production, 57, 59–63, 65–67 low density polyethylene, 167
Ethylene glycol dimethacrylate (EGDMA), 154 luffa fiber, 171
magnitudes, 173
maximum tensile strength, 167
F mechanical properties, 166
Fatty acid methyl ester (FAME), 68 MMT nanoclay, 175
Fermentation, 57–60, 62, 63, 68 n-hexane, 171
Filter paper activity (FPA), 57, 58 oil palm and coir fibers, 163
Index 221

optimum resistance, 171 P


polypropylene, 170, 172, 173 Phenol, 133
snakegrass and coir fibers, 174 Plasma
structure, 165 fibers, 10, 11
TDI, 168 Poly-β-hydroxybutyrate (PHB), 69
trilayer composite, 174 Poly(caprolactone) (PCL), 158–160, 162
unsaturated polyester resin, 173 Polylactic acid (PLA), 155–158
VARTM, 165 Polytetrafluoro ethylene (PTFE), 198
vinyl ester, 165 Polyvinylidene difluoride
Young’s modulus, 163 (PVDF), 100
Hydrochar, 95, 98 p-phenylene diamine (PPDA), 23
Hydrolysis, 60, 62, 63, 65, 67 Properties, coir fibers
Hydrothermal treatments, 105 acoustic, 14, 16
Hydroxybenzophenone, 144 geography, 16, 17
2-Hydroxyethyl acrylate (2-HEA), 23 thermal, 12, 14, 15
2-Hydroxyethyl methacrylate (HEMA), 23 Proton exchange membrane fuel cells
(PEMFCs), 198
Pyrolysis, 95, 98, 105, 109
K
Kojic acid, 68
R
Retting
L backwaters, 9
Langmuir adsorption models, 124 coir fibers, 10
Langmuir–Freundlich models, 118 production of coir fibers, 4, 5
Lignocellulose, 55 Rice husk ash, 41, 42
Limiting Oxygen Index (LOI), 180
Linear medium density polyethylene
(LMDPE), 148 S
Saccharification, 56, 59, 62
Saccharomyces cerevisiae, 59, 60,
M 62–64, 67
Mature coconut fiber (MCF), 60 Sewage sludge, 36, 41, 42
Mature coconut shell (MCS), 60 Single pass removal efficiency
Methylmethacrylate (MMA), 21–23 (SPRE), 91
Microfibrillar angle (MFA), 20, 21 Soilless growth media, 31, 36, 39
Mulching, 39–43 Solid state fermentation, 57, 58
Mushrooms, 69, 70 Solvent casting technique, 158
Split strength, 79
Structure, coir fibers
N morphological, 11–13
Nanocellulose, 187, 188 Substrates
Nanocrystals catalysts, 65, 68
cellulose, 187 coir, 55
Ni-Pd, 188 coir pith, 58
Nanolignin, 191, 193 enzyme and chemical production, 55
Nanotubes, 191 immobilize laccase, 70
Nanowhiskers, 188, 190, 206 loading capacity, 71
Natural reinforcing materials, 83, 90 mushroom, 69, 70
Ni-Pd nanoparticles, 189 Superabsorbent hydrogels, 206, 207
Non-woven geotextiles, 43, 44 Supercapacitors
carbonization of coconut fiber and shells,
99, 100, 102, 103
O Li-ion hybrid electrochemical, 108
Oil sorption, 195 specific capacitance, 106, 108
222 Index

T V
3C-SiC whiskers, 206, 207 Vacuum assisted resin transfer molding
Toluene diisocyanate (TDI), 168 (VARTM), 165
Torrefaction, 95, 96

W
U Wood vinegar, 199, 201
Unsaturated polyester resin (UPR), 141 Woven geotextiles, 44
Upflow anaerobic sludge blanket (UASB), 4
Urea, 71

You might also like