0% found this document useful (0 votes)
22 views114 pages

Page

The document discusses developing natural fibre reinforced bio-epoxy composite roof sheets. It examines using flax and kenaf fibres to reinforce a bio-epoxy matrix. The effects of fibre content, water aging, and UV weathering on mechanical properties are investigated. Flax fibre composites exhibited higher tensile strength but absorbed more water. Both composites decreased in strength after weathering, but kenaf composites reduced more.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
22 views114 pages

Page

The document discusses developing natural fibre reinforced bio-epoxy composite roof sheets. It examines using flax and kenaf fibres to reinforce a bio-epoxy matrix. The effects of fibre content, water aging, and UV weathering on mechanical properties are investigated. Flax fibre composites exhibited higher tensile strength but absorbed more water. Both composites decreased in strength after weathering, but kenaf composites reduced more.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 114

THE DEVELOPMENT OF NATURAL FIBRE REINFORCED

COMPOSITES ROOF SHEET

by

MUFEBA MUSIKWA

Submitted in accordance with the requirements for the degree

MASTER OF SCIENCE (M.Sc.)

Department of Textile Science

Faculty of Science

at

NELSON MANDELA UNIVERSITY

SUPERVISOR: DR MAYA JACOB JOHN


CO-SUPERVISOR: PROFESSOR RAJESH D. ANANDJIWALA

APRIL 2018

Page |
DEDICATION

This work is dedicated to my parents, siblings and the entire Mufeba family.

Ndia livhuwa thikedzo yavho.

Page |ii
ACKNOWLEDGEMENTS

The completion of this study could not have been possible without the Almighty God who
gave me strength, courage and perseverance to finally finish it.

I am grateful to my supervisor, Dr Maya John for her patience and expertise, co-supervisor,
Professor RD Anandjiwala for challenging me and helping me develop my ideas.

I would like to extend my gratitude to the Council of Scientific and Industrial Reaserch
(CSIR) for their financial support and resources that were utilised in completing this study.

I am thankful to the friends I made at CSIR, Sandisiwe, Princeton, Mpho, Abongile, Darrel
and Mziwamadoda for their academic support and friendship. I also thank the administrative
staff for their support during this study.

Special thanks to Dr Thabang Mokhothu, Dr Teboho Mokhena, Dr Asanda Mtibe, Dr Mlando


Mvubu, Mr Lebo Maduna, Dr Cyrus Tshifularo, Mr Steve Chapple and Mr Osei Ofosu for
their support with experiments.

Finally, I would like to acknowledge with gratitude, the support and love of my family – my
parents, Joseph and Florence; my brother, Humbelani; my sister, Mulivhuweni, niece,
Mufunwa and nephew, Wandeme and my fiancé, Dr Cyrus Tshifularo and his family. They
all kept me going, and this work would not have been possible without them.

Page |iii
ABSTRACT

The study aims to develop natural fibre reinforced bio-epoxy composite for use as roof
sheets, manufacturing and characterization to evaluate its suitability for building applications.
In this study natural fibres such as flax and kenaf were selected to reinforce bio-epoxy matrix.
Different weight ratios of flax and kenaf fibres were processed by needle-punching technique
to produce nonwoven mats. The nonwoven mats and bio-epoxy matrix were prepared using
vacuum assisted resin transfer moulding (VARTM) at room temperature until dry and cured.
The effects of weathering and water aging on the static and dynamic mechanical properties of
kenaf and flax composites were investigated.

Flax fibre reinforced bio-epoxy composites were found to exhibit higher tensile strength at
25% fibre content of 41.5 MPa in comparison to the composites reinforced with kenaf fibres
(33.0 MPa). With regards to the drop weight impact results, flax fibre reinforced bio-epoxy
composites exhibited brittle failure. Water aging results showed that kenaf fibre reinforced
bio-epoxy composites absorbed less water for all fibre contents in comparison to composites
reinforced with flax fibres. The tensile strength and modulus of both the composites
reinforced with flax and kenaf fibres were reduced after water aging. However, the composite
reinforced with kenaf fibres showed the maximum reduction in tensile strength at 25% fibre
content. After UV treatment both composites reinforced with flax and kenaf fibres showed a
decrease in tensile strength of 6.25% and 30%, respectively. In comparison to kenaf, bio-
epoxy composites reinforced with flax fibres showed an increase in tensile modulus. Both
composites reinforced with flax and kenaf fibres were found to be brittle and broke easily but
no colour fading was observed after UV treatment.

The dynamic mechanical analyses results showed that the incorporation of flax and kenaf
fibres increases the storage modulus of the composites with the maximum storage modulus
value exhibited by flax fibre reinforced bio-epoxy composite at 30% fibre content. The glass
transition temperature of composites reinforced with both flax and kenaf fibres shifted to
lower temperatures of 79 °C and 69 °C respectively, in comparison to 96 °C for bio-epoxy
resin, with the incorporation of fibres.

Page |iv
TABLE OF CONTENTS

Page
Declaration i
Dedication ii
Acknowledgements iii
Abstract iv
Table of contents v
List of symbols and abbreviations viii
List of tables x
List of figures xi
CHAPTER 1: MAIN INTRODUCTION 1
1.1 Background of the study 1
1.2 Problem statement 3
1.3 Research aims and objectives 3
1.4 Outline of the thesis 4
CHAPTER 2: LITERATURE REVIEW 6
2.1 Composite materials 6
2.1.1 Polymer matrix composites (PMCs) 6
2.1.1.1 Matrices 7
2.1.1.1.1 Thermoplastic polymers 7
2.1.1.1.2 Thermoset polymers 10
2.1.1.1.3 Bio-based polymers 11
2.2 Reinforcements 13
2.2.1 Classification of natural fibres 13
2.2.2 Plant fibres 14
2.2.3 Extraction of NFs 17
2.2.4 Advantages and Disadvantages of natural fibres 19
2.2.5 Natural fibre reinforced polymer composites 20
2.2.5.1 Natural fibre reinforced thermoplastic composites 20
2.2.5.2 Natural fibre reinforced thermoset composites 21
2.3. Processing of NFCs 21
2.4 Factors affecting NFC performance 24

Page |v
2.4.1 Aspect ratio 24
2.4.2 Fibre volume fraction 25
2.4.3 Fibre Orientation 26
2.4.4 Interfacial adhesion 27
2.4.4.1 Surface modifications 28
2.4.5 Moisture absorption 34
2.4.6 Weathering in natural fibre reinforced composites 38
2.5 Applications of natural fibre reinforced composites 40
2.5.1 Building applications 43
CHAPTER 3: EXPERIMENTAL TECHNIQUES 48
3.1 Materials 48
3.1.1 Flax fibres 48
3.1.2 Kenaf fibres 49
3.1.3 Epoxy resin 49
3.2 Preparation of flax and kenaf nonwovens 51
3.2.1 Needle punching technique 51
3.3 Fabrication of composites 52
3.3.1 VARTM technique 52
3.4 Characterisation of composites 53
3.4.1 Mechanical Properties 53
3.4.1.1 Tensile testing 53
3.4.1.2 Flexural testing (three point bending) 54
3.4.1.3 Drop weight impact testing 54
3.4.2 Dynamic mechanical analysis (DMA) 55
3.4.3 Water absorption studies 56
3.4.4 Environmental ageing (UV radiation) 56
CHAPTER 4: RESULTS AND DISCUSSION 57
4.1 Mechanical properties 57
4.1.1 Effect of fibre content on flax and kenaf fibre reinforced bio-epoxy composites 57
4.1.1.1 Tensile properties of flax and kenaf fibre reinforced bio-epoxy composites 57
4.1.1.2 Flexural properties of flax and kenaf fibre reinforced composites 59
4.1.2 Drop weight impact testing for flax and kenaf fibre reinforced bio-epoxy composites 62
4.1.3 Comparison of mechanical properties of flax and kenaf fibre reinforced composites 69

Page |vi
4.2 Water absorption studies 70
4.2.1 Water absorption behaviour 70
4.2.2 Effect of water aging on the mechanical properties of flax and kenaf reinforced bio-
epoxy 75
4.2.2.1 Tensile properties 75
4.3 Accelerated weathering (UV radiation) 79
4.3.1 Tensile properties 80
4.4 Dynamic mechanical analysis (DMA) 84
CHAPTER 5: CONCLUSIONS 89
References 91

Page |vii
LIST OF SYMBOLS AND ABBREVIATIONS

⁰C Degree Celsius

% Percentage

CaCl2 Calcium chloride

df Fibre diametre

Eur/kg Euro per kilogram

FRAs Fire retardant additives

g/m2 Gram per square metre

g/m3 Gram per cubic metre

GPa Gigapascal

HDPE High density polyethylene

J/m Joule per metre

Ɩc Critical length

Ɩ/d Fibre aspect ratio

lf Fibre length

LDPE Low density polyethylene

LLDPE Linear low density polyethylene

mm Milimetre

MPa Megapascal

ηlE Stiffness

ηlS Strength

Page |viii
NF Natural fibre

NFC Natural fibre reinforced composite

OH Hydroxyl

PA Polyamide

PCL Polycaprolactone

PE Polyethylene

PEEK Polyether ether ketone

PHBV Polyhydroxybutyrate-co-hydroxyvalerate

PLA Polylactic acid

PMC Polymer matrix composite

PP Polypropylene

PS Polystyrene

PVC Polyvinyl chloride

RDP Reconstruction and Development Programme

RTM Resin transfer moulding

Tg Glass transition temperature

T max Maximum peak temperature

Tm Melting temperature

UV Ultraviolet radiation

VARTM Vacuum assisted resin transfer molding

vf Volume fraction

WPC Wood fibre reinforced polymer composite

wt % Weight percentage

Page |ix
LIST OF TABLES

Table 2.1 Common thermoplastic polymers and their properties ............................................. 9


Table 2.2 Common thermosetting polymers and their properties ........................................... 11
Table 2.3 Cost of biodegradable vs. thermoplastic polymers ................................................. 13
Table 2.4 Natural fibres in the world and their world production........................................... 15
Table 2.5 Processing techniques for thermoset and thermoplastic matrix .............................. 22

Table 3.1 Physical and mechanical properties of flax fibre .................................................... 48


Table 3.2 Physical and mechanical properties of kenaf fibre ................................................. 49
Table 3.3 Mechanical properties and processing data of Epoxy resin – from the supplier’s
technical data sheet .................................................................................................................. 50
Table 3.4 Processing parameters used for preparation of needle punched kenaf and flax
nonwovens ............................................................................................................................... 51

Table 4.1 Impact energy absorbed during impact testing ....................................................... 63


Table 4.2 Tensile and Flexural properties of flax and kenaf fibres reinforced bio-epoxy
composites at 25% fibre content. ............................................................................................. 69
Table 4.3 Water absorption, diffusion, sorption and permeation coefficients of flax and kenaf
fibre reinforced bio-epoxy composites after immersion in water. ........................................... 73

Page |x
LIST OF FIGURES

Figure 2.1 Composite material .................................................................................................. 6


Figure 2.2 Classification of biodegradable polymers ............................................................. 12
Figure 2.3 Classification of natural fibres ............................................................................... 14
Figure 2.4 (a) Composition and cross section of a fibre plant stem and (b) schematic diagram
of the bundle fibre .................................................................................................................... 16
Figure 2.5 Predictions of the fibre length efficiency factors for a) stiffness ηlE and b)
strength ηlS, based on Cox’s shear lag model (Eq. 2.1) and Kelly-Tyson’s model (Eq. 2.2),
respectively. Typical values for flax reinforced PFRPs are used in the calculations: df = 20
μm, Gm = 1 GPa, Ef = 50 GPa, vf,max,FRP = π/4, vf = 0.30, σf = 1000 MPa and τ = 30 MPa
or 15 MPa (lc = 0.333 mm or 0.667 mm, respectively). .......................................................... 25
Figure 2.6 Different fibre orientation in composites: a) unidirectional, b) random, c) biaxial
and d) multi directional ............................................................................................................ 27
Figure 2.7 Corona (a) and Plasma (b) treatment ..................................................................... 29
Figure 2.8 Schematic representation of untreated and alkali treated natural fibre ................. 30
Figure 2.9 Acetylation of natural fibre.................................................................................... 31
Figure 2.10 Hydrolysis of silanol ............................................................................................ 32
Figure 2.11 Reaction of silane with -OH groups of natural fibre ........................................... 33
Figure 2.12 Chemical reaction of cellulose fibres with maleated PP ..................................... 34
Figure 2.13 (a) Fickian diffusion at room temperature and (b) non-Fickian diffusion at
elevated temperature ................................................................................................................ 36
Figure 2.14 NFC applications in the automotive industry ...................................................... 40
Figure 2.15 Vacuum press-moulded sport seat ....................................................................... 41
Figure 2.16 Various NFC applications: (a) Biomedical applications (b) wind energy, (c)
Sports and leisure, (d) furniture ............................................................................................... 43
Figure 2.17 Schematic presentation of a monolithic house roof made of natural composites as
a hurricane-resistant roof ......................................................................................................... 46
Figure 2.18 Photographs of (a) Roof shingles, (b) Eindhoven bio-bridge, (c) the Hemp house
and (d) NFC sandwich roof construction ................................................................................. 47

Page |xi
Figure 3.1 Needle punching process to produce nonwovens ............................................................... 52
Figure 3.2 (a) Nonwoven, (b) Vacuum infusion setup and (c) Prepared composite ............................ 53
Figure 3.3 Schematic diagram of the impact test arrangement ............................................................ 55

Figure 4.1 Effect of fibre content on the tensile strength and modulus of flax fibre reinforced
bio-epoxy composites .............................................................................................................. 58
Figure 4.2 Effect of fibre content on the tensile strength and tensile modulus of kenaf fibre
reinforced bio-epoxy composites ............................................................................................. 59
Figure 4.3 Effect of fibre content on the flexural strength and flexural modulus of flax fibre
reinforced bio-epoxy composites ............................................................................................. 60
Figure 4.4 Effect of fibre content on the flexural strength and modulus of kenaf fibre
reinforced bio-epoxy composites ............................................................................................. 61
Figure 4.5 (a-h) Photographs of flax fibre reinforced composite specimens after impact
testing ....................................................................................................................................... 66
Figure 4.6 (a-h) Photographs of kenaf fibre reinforced composite specimens after impact
testing ....................................................................................................................................... 68
Figure 4.7 Water absorption behaviour of flax reinforced bio-epoxy composites with time for
different fibre content .............................................................................................................. 72
Figure 4.8 Water absorption behaviour of kenaf reinforced bio-epoxy composites with time
for different fibre content ......................................................................................................... 74
Figure 4.9 Effect of water absorption in flax/ bio-epoxy composite and its effect on fibre-
matrix interface ........................................................................................................................ 75
Figure 4.10 Effect of water aging on the tensile strength of flax fibre reinforced composites
at different fibre content........................................................................................................... 76
Figure 4.11 Effect of water aging on the tensile modulus of flax fibre reinforced composites
at different fibre content........................................................................................................... 77
Figure 4.12 Effect of water aging on the tensile strength of kenaf fibre reinforced composites
at different fibre content........................................................................................................... 78
Figure 4.13 Effect of water aging on the tensile modulus of kenaf fibre reinforced
composites at different fibre content........................................................................................ 79
Figure 4.14 Pictures of tensile specimens (a) flax composite before UV exposure, (b) flax
composite after UV radiation, (c) kenaf composite before UV exposure and (d) kenaf
composite after UV radiation ................................................................................................... 80

Page |xii
Figure 4.15 Effect of UV radiation on the tensile strength of flax fibre reinforced composites
at different fibre content........................................................................................................... 81
Figure 4.16 Effect of UV radiation on the tensile modulus of flax fibre reinforced composites
at different fibre content........................................................................................................... 82
Figure 4.17 Effect of UV radiation on the tensile strength of kenaf fibre reinforced
composites at different fibre content........................................................................................ 83
Figure 4.18 Effect of UV radiation on the tensile modulus of kenaf fibre reinforced
composites at different fibre content........................................................................................ 84
Figure 4.19 Variation of the storage modulus with temperature for flax fibre reinforced bio-
epoxy at different fibre content ................................................................................................ 85
Figure 4.20 Temperature dependence of Tan 𝛿 for flax fibre reinforced bio-epoxy at different
fibre content ............................................................................................................................. 86
Figure 4.21 Variation of the storage modulus with temperature for kenaf fibre reinforced bio-
epoxy at different fibre content ................................................................................................ 87
Figure 4.22 Temperature dependence of tan 𝛿 for kenaf fibre reinforced bio-epoxy at
different fibre content .............................................................................................................. 88

Page |xiii
CHAPTER 1: MAIN INTRODUCTION

1.1. Background of the study

The rising environmental awareness, high rate of depletion of petroleum resources and
government regulations favouring green products have led researchers to develop new
environmentally friendly materials and processes [1, 2]. Biodegradability, eco-friendliness,
abundance of raw material and light-weight have become important considerations in the
fabrication of new materials [3]. Natural fibre reinforced composites which possess the above
attributes are able to replace traditional polymer composites in non-load bearing applications
[4]. The polymer composite, produced from non-renewable oil-based resources, has
limitations such as difficulty to reuse and recycle (particularly those produced from thermoset
resins) to an extent that disposal is preferred. Furthermore, the production of traditional
composites reinforced with glass fibres are abrasive to processing machinery and hazardous
to the operator due to inhalation of glass particles [5].

Natural fibres have low density, good thermal, acoustic and insulation properties and
comparable specific strength [2] to that of glass fibres used in traditional composites. These
factors contribute to the gradually increasing applications of natural fibres in industrial
sectors such as automotive, construction and packaging. Various manufacturers have
employed wood fibre reinforced polymer composites (WPC) in the building and construction
sector for structural (load bearing walls, stairs, roof systems, and sub-flooring) and non-
structural applications (outdoor decking, window and door frames, ceiling tiles and furniture)
[6]. However, non-structural applications do not require as high mechanical properties as
structural applications [1]. Wood based composites are also used to manufacture composite
panels for partition walls in buildings. The panels made from WPC are economically
attractive because wood fibres are easily available in sawmill wastes [4] and they are
reasonably strong and lightweight [7]. Bagasse fibres obtained from sugar cane was first
utilized in America in the 1920s to manufacture composite panels [8]. Currently, building
panels and roofing sheets made from bagasse fibre reinforced phenolic composites are being
used in houses in Philippines, Jamaica and Ghana [7]. Cereal straws are also utilized in the
production of panels due to their low density, resulting in panels that are more resilient and
resistant to earthquakes [9]. Rice husks or their ash are also used in fibre cement blocks and

Page |1
other cement products. The presence of rice husks in building products helps in improving
acoustic and thermal properties [8].

In another study, sisal fibres were manually cast on corrugated sheets and the strength of the
composites was evaluated. The results showed that the strength of corrugated roof sheets was
improved upon incorporation of sisal fibres [10]. This was attributed to the fact that the fibres
act as energy absorber and also prevent crack generation and propagation[10, 11]. Recycled
paper was previously evaluated in composite sheets and structural beams produced by
VARTM and was found to exhibit the required stiffness and strength for roof construction
[12]. The accelerated and natural weathering of bagasse reinforced cement composite filled
with rice husk ash pozzolan (class of siliceous and aluminous) used as roofing sheets was
studied [13] and the authors reported that the incorporation of bagasse in cement composites
had no significant effect on the tensile and absorption properties for the bagasse treated with
2% CaCl2 and 20% replacement of cement with rice husk ash. Thus, the cement composites
reinforced with bagasse are suitable alternatives for construction of both external and internal
building panels.

Although attractive, natural fibre reinforced composites have some drawbacks such as poor
adhesion between the hydrophilic fibres and the hydrophobic polymer matrix [4].
Researchers have investigated and addressed the compatibility between fibre and the polymer
matrix [2, 4, 14, 15]. Most of the researches have been focused particularly on chemical
modification to remove surface impurities and waxes in order to improve the interfacial
adhesion between the fibre and matrix, as well as their dispersion in the matrix. Good
interfacial adhesion prevents energy absorption through interfacial cracks, thereby reducing
fracture toughness within the composite, and thus leading to improved mechanical properties
[5]. The properties of a fibre-reinforced composite depends considerably on the fibre-matrix
interface because the interface acts as a binder and transfers stress between the matrix and the
reinforcing fibres [16].

Another problem associated with the use of natural fibre composites in building materials is
their high flammability and poor fire resistance. Studies have shown that the fire retardant
behaviour of natural fibre reinforced composites (NFRC) can be improved through
incorporation of fire retardant additives (FRAs) such as soluble inorganic salts and polymers

Page |2
containing nitrogen, halogen, and phosphorus [17, 18] to expand their usage in other
applications (building, public transport, and electrical equipment). The primary function of
FRAs is to prevent, minimize, suppress or stop combustion of a material. Polymer/ resin
blends are also used to improve flame retardancy in composite materials. Polymers with high
flammability is blended with flame retardants, and the resultant blend shows improved heat
resistance and a reduction in heat release and smoke and toxic gas emission [19].
Natural fibres are a primary source of fuel during fire, but their usage in composites is also
advantageous as they can help improve char formation. Hence, the combustion of natural
fibres must be thoroughly investigated to understand their contribution to fire. Thus, efficient
flame retardant solutions for natural fibres will help develop fire retardant composites.

1.2. Problem statement

South Africa’s population is rapidly growing and the need for good quality and affordable
housing for low and middle income earners poses a huge problem for the country as a whole.
A few decades ago, government started a housing project known as the Reconstruction and
Development Programme (RDP) to address this problem. However, the houses built are of
very low standard and with significant health and safety risks to dwellers. Moreover, the
South African building and construction sector is now focusing on energy efficiency to
minimize the environmental impact of buildings by using environmentally friendly materials
and production processes, and as a result, the need to develop low cost, durable materials for
the building and construction sector is of great importance. The solution to the above
problem will be to use environmentally friendly materials to develop quality building
products for the housing sector and this will be attempted in this study.

1.3. Research aims and objectives

This study aims to develop a natural fibre based roof-sheet panel by VARTM technique. The
currently available roof panels are epoxy composites reinforced with glass fibres. In this
study, glass fibre will be replaced by flax and kenaf nonwovens. The roof panel will be
subjected to in-depth characterizations and the properties will be compared to that of the
market products such as PVC roof sheet. The objectives to achieve the set goals are:

Page |3
1. Preparation of flax and kenaf nonwovens by needle-punching technique
2. Development of natural fibre reinforced epoxy composites using VARTM technique
3. Optimise the moulding parameters to achieve the ideal epoxy resin saturation
4. Characterization of the developed natural fibre-reinforced composite roof panels as
follows:

 Mechanical properties (tensile, flexure and impact tests)


 Thermogravimetric analysis
 Dynamic mechanical analysis
 Water absorption studies
 Environmental ageing

1.4. Outline of the thesis

The study will consist of five chapters. A brief description is as follows:

CHAPTER 1: MAIN INTRODUCTION

The chapter provides a general background to the study, the objectives and outline thereof.

CHAPTER 2: LITERATURE REVIEW

This chapter will review the published literature on natural fibre composites. A broad
discussion on different types of composite materials and their applications will be reported.
The commonly used composite manufacturing processes will also be discussed.

CHAPTER 3: EXPERIMENTAL TECHNIQUES

In this chapter, the different processing techniques employed for fabricating natural fibre
composites will be highlighted. The characterization techniques for analysing flax and kenaf
reinforced bio-epoxy composites will also be reported.

CHAPTER 4: RESULTS AND DISCUSSION

Page |4
This chapter will present and discuss the mechanical, thermal and physical studies of flax and
kenaf based epoxy composites.

CHAPTER 5: CONCLUSION

The chapter will summarise the main findings of the study and will also highlight the future
studies.

Page |5
CHAPTER 2: LITERATURE REVIEW

2.1 Composite materials

When two or more components with different physical or chemical properties are combined
they produce a composite with different characteristics from the individual components, [20]
as depicted in Figure 2.1. The resulting composite is easily processed and have improved
properties (strength and stiffness) in comparison to those in individual constituents [21].

Figure 2.1 Composite material

In composites, one material is the reinforcing phase (fibre, sheet, and particle) and the other is
the matrix (polymers, ceramic, and metal). The reinforcement phase carries majority of the
load and provides strength and stiffness [7]. Mostly, the reinforcement is harder, stronger,
and stiffer than the matrix. The matrix is the continuous phase which surrounds and protects
the reinforcement from environmental factors (abrasion, wear, chemical attack) [22]. The
type of matrix such as polymer, cement and metal classifies the composite material [23].
Polymer matrix composite is mostly commercially produced, wherein the polymer is used as
a matrix with various reinforcing materials. In metal composites, metal is the matrix and the
reinforcement may be a metal, ceramic or organic constituent.

2.1.1 Polymer matrix composites (PMCs)

Polymer matrix composites consist of a polymer matrix (resin) and a dispersed phase of
reinforcing fibres. PMCs are very popular and widely used due to their low cost and simple
processing methods. Polymers or resins are classified in two types: thermoplastics

Page |6
[polypropylene (PP), polyethylene (PE), polystyrene (PS), polyvinyl chloride (PVC)] as
shown in Table 2.1, and thermosets (epoxy, phenol formaldehyde, polyester) and they can be
reinforced with different types of fibres such as natural (plant, mineral, animal) and synthetic
for various applications [23]. Reinforcing fibres may be continuous (unidirectional, cloth,
roving) and discontinuous (chopped, mat) depending on desired applications of the
composite. Some applications of PMCs include boat bodies, automotive parts, sport goods
(golf clubs, skis, tennis racquets, and fishing rods), brake and clutch linings and bullet-proof
vests [24]. Thermoplastic matrices are commonly used in automotive applications, while
thermosets are often used for high performance advanced composites as they achieve much
higher mechanical properties (strength and stiffness) in comparison to thermoplastic matrices.
As a result, thermoset polymers are more preferred in structural composites for construction
materials [15]. Another class of matrix materials is the biopolymers or bio-resins derived
from renewable resources. Currently the market for biodegradable polymers is rapidly
growing and several bio-based products are available, for example, polylactic acid (PLA),
starch, polyesters and soy-based polymers. However, the high price of these materials is one
of the restrictions for their use despite providing unique physical and chemical properties
[25].

2.1.1.1 Matrices

2.1.1.1.1 Thermoplastic polymers

Thermoplastic polymers soften when heated at high temperatures forming a viscous liquid
and retain their shape upon cooling to an amorphous solid [26]. The process is reversible by
heating to melting temperature. They are suitable to produce complex parts with flexible
designs at low processing costs compared to thermosets. Thermoplastic polymers (Table 2.1)
include high density polyethylene (HDPE), low density polyethylene (LDPE), linear low
density polyethylene (LLDE), polyamide (PA), polyether-ether ketone (PEEK) and can be
processed with techniques such as injection, extrusion and compression moulding. Due to
their low processing temperatures, thermoplastics tend to degrade at high temperatures
resulting in inferior properties and chemical instability compared to thermosets [27].
However, newly developed high performance polymers such as PEEK are more resistant to
cracking and impact damage, exhibit excellent high temperature strength and solvent

Page |7
resistance [28]. These new class of thermoplastic polymers are mainly used in automotive
applications and electrical and electronic goods, and make up a market share of 24% and
41%, respectively [29].

Page |8
Table 2.1 Common thermoplastic polymers and their properties [15, 30-32]

Property PP HDPE LDPE LLDPE PS PVC PA PEEK

(Nylon 6)

Density (g/m3) 0.90–0.92 0.94-0.96 0.91–0.93 0.94 1.04-1.06 1.41 1.12-1.14 1.3

Young’s modulus 15-700 2.0-130 90-800 837 1-2.5 60 2.9 4.14


(GPa)

Tensile Strength 26-41.4 14.5-38 40-78 24 25-69 52 43-79 103


(MPa)

Elongation at break 15-700 2.0-130 90-800 830 1-2.5 _ 20-150 20


(%)

Water absorption – 0.01-0.02 0.01-0.20 <0.02 <0.02 0.03-0.10 0.01 1.3-1.8 0.1
after 24hr

Izod impact 21.4-267 26.7-1068 >854 35 1.1 53.4 42.7-160 63


strength, notched
(J/m)

Page |9
2.1.1.1.2 Thermoset polymers

Thermoset polymers include phenolics, polyesters, epoxies, and vinyl esters as shown in
Table 2.2. Epoxies make up most of the current market for advanced composites [27]. Epoxy

resins are very quick to cure at temperatures from 5 ̊C to 150 C


̊ , depending on the hardener
used. Their advantageous properties make them out-perform other resin types in terms of
mechanical properties. One of their properties is the low shrinkage during cure which
minimises internal stress. High adhesive strength and high mechanical properties are also
enhanced by high electrical insulation and good chemical resistance. Epoxies are used as
adhesives, casting compounds, sealants, varnishes and paints, as well as composite resins for
a variety of industrial applications [28], and hence it is chosen in this study.

Polyesters and vinyl esters are used in the same temperature range as epoxies for commercial
applications but rarely for high performance composites because of their low mechanical
performance. Phenolic resins are used in high temperature applications as they have excellent
smoke and fire resistance, therefore, often used in aircraft interior components. Thermoset
polymers are synthesised from low molecular, low viscosity monomers that form a highly
crosslinked three-dimensional structure when curing. Crosslinking results from the chemical
reaction induced by heat generated in the reaction or by externally supplied heat. The reaction
accelerates as curing progresses and fills up the available space, causing a decrease in the
molecular arrangement, resulting in reduced mobility of the molecules and increase in
viscosity. The resulting structure cannot be reheated and reshaped once cured; the process of
thermoset application is irreversible unlike thermoplastics.

Page | 10
Table 2.2 Common thermosetting polymers and their properties [8, 33-35]

Property Epoxy Polyester Phenolics Vinyl ester

Density (g/m3) 1.1 - 1.4 1.2 - 1.4 1.3-1.7 1.2 - 1.5

Young’s modulus (GPa) 3-6 3.1 - 3.8 2.7-2.9 2 - 4.5

Tensile Strength (MPa) 35 - 100 69 - 83 22-24 40 - 90

Compression Strength 100 - 200 100 476 90 - 250


(MPa)

Elongation at break (%) 1-6 4-7 3 2

Water absorption – after 0.1 - 0.4 0.1 0 0.1 - 0.3


24hr

Izod impact strength (J/m) 0.3 2.5 13 0.15- 3.2

2.1.1.1.3 Bio-based polymers

Unsustainable consumption of petroleum based materials and the rising oil prices have led to
the development of materials from renewable resources with improved properties.
Researchers have explored and developed biodegradable polymers to replace their synthetic
counterparts. Biopolymers are generally classified as biodegradable polymers and can be
reinforced with either renewable (plant and animal) or synthetic fibres. Biopolymers are
classified according to the source as shown in Figure 2.2.

Page | 11
Figure 2.2 Classification of biodegradable polymers [36]

However, biopolymers show moderate strength and stiffness and for this reason their
applications are limited to non-critical and non-load bearing parts such as packaging; they
cannot be used in load bearing applications where high strength and stiffness are required [5].
They are also relatively expensive (Table 2.3) and have smaller scale production in
comparison to traditional thermoplastic polymers. Biopolymers that have been extensively
explored for various applications include PLA and its derivatives, which are commonly used
in biomedical applications such as drug delivery, vascular grafts, artificial skin, orthopaedic
implants and scaffolds for tissue engineering [37, 38]. Polyhydroxybutyrate-co-
hydroxyvalerate (PHBV) is used in motor oil containers, film formation and paper coatings
and other packaging. PLA is also used to make carpet fibres because of its unique properties
of breaking down under composting conditions. Starch polymers have the potential for use in
non-food industries where 50% of its production is currently used. Starch polymers have
replaced polystyrene in foam products such as expanded trays, shape moulded parts and
expanded layers [38].

Page | 12
Table 2.3 Cost of biodegradable vs. thermoplastic polymers [25]

Biodegradable (Eur/kg) Thermoplastic (Eur/kg)


polymers
polymers

PLA 3-4 HDPE 0.92

Starch 2-4 LLDPE 0.82

Polyesters 3.5-5 PP 0.80

BAK (polyester amide) 5.58 PS 1.08

PCL (polycaprolactone) 8.3 PVC 0.78

2.2 Reinforcements

2.2.1 Classification of natural fibres

Natural fibres are divided into three categories depending on their origin: plant based, animal
based, and mineral based. Asbestos is a mineral based naturally occurring fibre. It has been
mined and utilized all around the world specifically in building materials [39]. The main
properties of asbestos fibres include thermal, electrical and sound insulation [23]. Animal
based fibres comprise of proteins and include silk, wool, feathers, and hair. Plant based fibres
are ligno-cellulosic in nature. They comprise mainly of cellulose, hemicellulose, lignin,
pectin and waxes, and prominent examples are flax, hemp, jute, sisal, kenaf and ramie. Plant
fibres can be categorized into seed, fruit, bast/ stem, stalk, and leaf fibres as shown in Figure
2.3. The amount of cellulose in a fibre and micro-fibrillar angle determine its strength and
stiffness, which is provided by hydrogen bonds in the cellulose. Hemicellulose is responsible
for moisture absorption, biodegradation, and thermal degradation of fibre, while lignin gives
rigidity to fibres [40].

Page | 13
Figure 2.3 Classification of natural fibres [20]

2.2.2 Plant fibres

The increase in population, depletion of oil resources, fluctuating prices and shift towards
environmentally friendly products have led to natural resources being exploited to replace
synthetic materials, hence the utilization of natural fibres as reinforcement in composite
materials have gained interest. Today, different types of natural fibres have been investigated
for use in polymer composites and that includes flax, kenaf, sisal, jute, coir, agave, straw,
ramie, pineapple leaf fibre, cane (sugar and bamboo), wood fibre, grass reeds, rice husk and
kapok [41]. Sisal is widely grown in tropical African countries, the West Indies, and the Far
East with Tanzania and Brazil being the two main producing countries. Kenaf is grown
commercially in the United States and flax in the European Union as well as in many diverse
agricultural systems and environments throughout the world, including Canada, Argentina,
India, and Russia [42]. The most common and commercial natural fibres in the world are
shown in Table 2.4 as well as their worldwide production.

Page | 14
Table 2.4 Natural fibres in the world and their world production [26]

Fibre source Worldwide production


(103 tons)
Jute 2300

Kenaf 970

Flax 830

Grass 700

Sisal 375

Hemp 214

Coir 100

Ramie 100

Abaca 70

Natural fibres (NFs) are grown for textile or oil productions. NFs grown for textile purposes
are located in the bark of the plant stem and termed ‘bast fibres’. The stem is comprised of
three layers: bark, bundle and xylem as shown in Figure 2.4(a). The outer layer of the bark
act as a protective cover to prevent environmental attacks, and for the penetration of water
and other nutrients [43]. During fibre processing, the bark and xylem are eliminated to leave
the fibre bundles consisting of elementary fibres. Technical fibres are extracted by partly
separating the fibre bundles in the fibre plant and can be as long as the stem length of
approximately 1m. The schematic diagram of bundle fibres is presented in Figure 2.4(b).
Technical fibres consist of about 10-40 elementary fibres in the cross section [44, 45]. Unlike
the technical fibres, the length of elementary fibres varies between 2 and 5 cm, and the
diameter is about 19-25 µm [46]. The elementary fibres have primary and secondary cell
walls which are both cellulose material. Cellulose, hemicellulose, lignin and pectin are major
polymers involved in the composition of fibres. Cellulose is the main component wound
around the cell wall providing strength, stiffness, and structural stability. Hemicellulose is
hydrogen bonded to the cellulose micro-fibrils and act as a cementing matrix between the

Page | 15
micro-fibrils structure. It is responsible for moisture absorption, thermal degradation and
biodegradation of the fibre. Lignin encapsulates the cellulose/ hemicellulose units together
giving rigidity to the plant fibre [47, 48]. It is also reported to be very sensitive to ultraviolet
radiation and responsible for the degradation of fibres under UV radiation [20].

(a) (b)

Figure 2.4 (a) Composition and cross section of a fibre plant stem [49] and (b) schematic
diagram of the bundle fibre [50]

Separated fibres (long and short) are further processed for other textile applications. The long
fibres are used for spinning very fine yarns while short fibres are used as raw material for
coarser yarns [51]. Short fibres are further assembled in sheets resembling woven or knitted
fabric in appearance, referred to as nonwoven. The thickness of the sheets varies from 25 µm
to several centimetres, and weighs from 10 to 1000 g/m2 [52]. The fibres are passed through
rows of needles and punched to achieve mechanical adhesion between them; and they are
called needle punched fabrics.
Woven fabrics are another type of reinforcement made by weaving yarns or tows in different
patterns. Yarns are divided into two components: one is the warp which runs along the length
of the yarn and the other is the weft, which runs in cross direction. The warp and weft yarns
produce reinforcement in two directions, usually 0 and 90 degrees. A woven fabric is usually
supplied without resin or as a prepeg with resin [22].

Page | 16
2.2.3 Extraction of NFs

Plants are allowed to grow until harvesting time depending on the plant type. Harvesting is
done by pulling or mowing (manually or mechanically) of plants, thereafter, they are left in
the field to dry. Harvesting should be done such that the fibre length is preserved. As fibres
are used in composite reinforcement, the greater the aspect ratio of the fibre, the better the
mechanical properties of composites.
Natural fibres from the leaf, bast and seed of the plant are used as reinforcements in
composite materials. Fibres are extracted by a method named decortication. Leaves are
beaten and crushed by blunt knives until only the leaf fibre has remained. Thereafter, the
fibres are washed and dried in hot air. This is followed by machine combing and sorting
according to different grades. Bast fibres include flax, hemp, kenaf and jute. They act as
structural reinforcements for the plant stem and have typically the best improvements in
mechanical properties. Bast fibres are extracted from the plant stem through retting processes.
Retting process facilitates the separation of fibres from the tissues of the stem. During retting,
the removal of pectin and other nonfibre materials allows the fibres to loosen from the plant
core, as described below.

Retting

Retting is a microbial process that breaks down the chemical bonds which hold the stem
together and thereby separate the fibres from the nonfibre substances and the woody core,
which is termed shives [53]. Insufficient or under-retting results in poor separation of shives
from the fibres, leaving this material attached to the fibre bundles. Furthermore, over-retting
can weaken the cellulose fibres resulting in low fibre yield, processing efficiency, and
ultimately poor fibre quality [54]. Hence, the retting process is very vital in fibre yield and
quality. The two main methods of retting that have been historically employed to ret plant
fibres are namely, dew or field retting and water retting. Dew retting is the oldest method
used in the Western Europe to obtain high quality fibres [55]. It is inexpensive, mechanised
and does not require water. Plant stems are pulled and left in the field to rot. Moisture is
needed to encourage fungi to colonize and grow on the stems. Dew retting is the main
method for extracting high grade fibres; however the quality has declined over the years due
to noncompliant climates. Too dry weather result in poor fungal growth and lack of proper

Page | 17
retting while wet weather delays the harvest time and also interferes with fungal growth,
resulting in anaerobic degradation. These in turn produce low quality and inconsistent fibres.
Another disadvantage is that large hectares of land are occupied for weeks until retting is
complete. Dew retted fibre is dusty due to fungi and soil, posing a health risk to farmers [51].
On the other hand, water retting produces more uniform and high quality fibres but the
process is labour and capital intensive. In this process, plant stems are pulled and immersed
in water tanks or ponds for a period of five to seven days. The retted stems are then dried and
sun bleached in the field [56]. The high cost and resulting pollution, accompanied with the
stench residing in water-retted fibres led to the process being abandoned in the Western
Europe [54]. Furthermore, the process uses large volumes of clean water and produces
environmentally unacceptable fermentation waste, causing pollution. The present alternative
form of retting that has had long term consideration is the use of enzymes in controlled
reactions. The improved micro-organisms or direct use of enzymes for retting practiced in the
European countries may lead to the production of high quality bast fibres for textile
production. However, the cost of enzymes and other reasons have prevented the upscaling of
enzyme retting process and as a result, dew-retting remains the commercial process to obtain
fibres [55].

Scutching

Once the fibres are retted and thoroughly dried, they are mechanically separated in a process
known as scutching. Drying of retted fibres is vital because it allows the fibre bundles to
shrink so as to separate from the woody core. The two operations of scutching are breaking
and hackling. In the breaking process, stalks are passed between the fluted rollers to crush
and break the woody core into short shives of less than 1cm in length. Hackling is the
removal of shives and is achieved by tangential scraping of the broken stems [55]. The
process of long fibre scutching produces two types of fibres, the scutched fibre (long fibre)
which measures up to 700mm and the tow (short fibre) which are less than 200mm.

Page | 18
2.2.4 Advantages and Disadvantages of natural fibres

Advantages

The growing interest in natural fibres is mainly due to their advantages over glass fibres.
They are environmentally friendly, abundant, renewable, fully biodegradable, inexpensive
and have low density [57]. Natural fibres are lightweight (less than half the density of glass
fibres) which result in high specific strength and stiffness. They are a renewable resource; the
production of natural fibres requires little energy. Natural fibres are biodegradable; they do
not pose a hazard when disposed. Another interesting aspect is that they are natural organic
products, and are not harmful when handled and during processing in comparison to certain
synthetic fibres which can cause dermal and respiratory issues. In comparison to glass,
natural fibres are nonabrasive to processing equipment. Another advantage is that, due to
their hollow cellular structure, natural fibres exhibit good thermal and acoustic insulation
properties [25].

Disadvantages

One of the major limitations of using natural fibres as reinforcement in composites is poor
adhesion between the polar fibre and the non-polar matrix. In composite materials, the matrix
acts as a binder while the fibres transfer stress in the material. If the adhesion with the fibres
is weak, the composites will not have desired properties and may be susceptible to
environmental attacks resulting in short lifespan. Therefore, it is imperative to improve the
adhesion between the fibre and matrix by chemically or physically modifying the fibre
surface. Unfortunately, the treatments for fibre modifications do not always keep the integrity
of fibres and their natural character [58]. Another drawback of natural fibres is the variability
of their properties depending on the batch, even from the same cultivation area and the
location of fibre in the plant. For example, comparing the mechanical properties of flax fibres
located at different positions in the stem showed that the fibres located in the centre had
better mechanical strength than the ones at the higher and lower points [59].

The susceptibility of natural fibres to fungal attack poses a serious problem during storage,
transportation, composite processing, and use in humid conditions. Low degradation

Page | 19
temperature of natural fibres constitutes another drawback. The processing temperature of
natural fibres should not exceed 200 °C. Above this temperature, natural fibres undergo
degradation and therefore cannot be used with polymers that have high melting temperatures
[60]. Natural fibres such as glass tend to release volatile gases before, during and after
composite fabrication. This factor can lead to odour issues if the composite is to be used in an
enclosed environment such as in automotive applications. Lastly, the hydrophilic nature of
natural fibres is a major limitation for their use as reinforcement in composites. Natural fibres
tend to absorb moisture from the environment, leading to swelling and creation of voids at the
interface which result in poor composite performance and low dimensional stability [61].

2.2.5 Natural fibre reinforced polymer composites

Natural fibre reinforced polymer composites have attracted the attention of researchers due to
their potential to replace synthetic fibre composites such as glass and carbon fibre
composites. Natural fibre composites have outstanding advantages such as:
 Renewable characteristics
 Easy availability
 Low cost
 Light weight and
 High specific strength and stiffness.

2.2.5.1 Natural fibre reinforced thermoplastic composites

Natural fibre reinforced thermoplastic composites are a family of composite materials, where
a natural fibre (flax, wood, kenaf, hemp, sisal, etc.) is incorporated in a thermoplastic
polymer to produce a composite. Amongst various thermoplastic polymers, PP is one of the
commonly used matrix materials because of its beneficial properties such as dimensional
stability, transparency and high impact strength. Khoathane et al reinforced hemp fibre in
polypropylene resin and reported the mechanical properties. They observed that an increase
in fibre loading resulted in increased tensile strength until a decrease was noticed when fibre
content increased from 10 to 20% [62] . Sisal and kenaf reinforced polypropylene composites
were investigated and an increase in tensile strength of kenaf and sisal fibre composites was

Page | 20
observed at 15% fibre loading, followed by a decrease at 30% fibre loading due to
insufficient wetting of the fibres [63].

2.2.5.2 Natural fibre reinforced thermoset composites

Thermoset materials have the ability to cure at room-temperature by heat and chemical
reaction, coupled with good mechanical properties. Thermoset composites are produced by
impregnating fibres in resin (polyester, epoxy, phenolic etc.) and cured to a highly cross-
linked, three-dimensional network structure. These crosslinked structures are highly solvent
resistant, tough, and creep resistant. Thermoset composites are processed by techniques such
as hand layup, compression, pultrusion, resin transfer, injection, pressure bag and vacuum
forming moulding methods, with fibre loading as high as 80% [40]. A detailed overview on
thermoset composites is given in Section 2.5.

Mechanical properties of natural fibre reinforced composites

The properties of natural fibre reinforced composites generally depend on fibre properties,
such as fibre volume fraction, fibre aspect ratio, fibre-matrix adhesion, stress transfer at the
interface, and orientation of the fibres. A number of studies involve the determination of
mechanical properties as a function of fibre content, effect of fibre treatments, and the use of
coupling agents [64-66]. Both the matrix and fibre properties are important in improving the
mechanical properties of the composites. The tensile strength (the maximum stress a material
can withstand without breaking) is more sensitive to the matrix properties, whereas the
modulus (a measurement of stiffness and an indication of load bearing capacity) is mainly
dependent on the fibre properties (long or short) [67]. To improve the tensile strength, a
strong interface, low stress concentration, and fibre orientation in load direction are required
whereas fibre concentration, fibre wetting in the matrix, and high fibre aspect ratio determine
the tensile modulus. The aspect ratio, which refers to the fibre length divided by diameter
(Ɩ/d), is important for determining the fracture properties. In short-fibre-reinforced
composites, a critical fibre length is required to develop its full stressed condition in the
polymer matrix. Fibre lengths shorter than this critical length lead to failure due to de-
bonding at the interface at lower load. For fibre lengths greater than the critical length, the

Page | 21
fibre is stressed under applied load and this result in higher tensilestrength of the composite
[40].
For good impact strength, an optimum bonding level is necessary. The degree of adhesion,
fibre pull-out, and a mechanism of energy absorption are some of the parameters that can
influence the impact strength of a short-fibre-filled composite. The properties vary with
composition as per the rule of mixtures and increase linearly with composition [68]. This
linear dependence on percentage fibre content does not hold at high percentage (above 80%
fibre content), due to lack of fibre wetting in the matrix polymer.

2.3. Processing of NFCs

Natural fibre reinforced composites are processed by various methods ranging from very
simple and low cost to complex and very expensive. Moulding processes are selected based
on factors such as, the end-use properties of the product, the volume of the product to be
manufactured, and dimensional tolerance of the product. The most common manufacturing
methods for thermoset and thermoplastic matrix composites are listed in Table 2.5. Pultrusion
was popularly used for manufacturing synthetic fibres and was only recently explored for
natural fibres as it has high potential for use with natural fibres and/ or biopolymers [69].

Table 2.5 Processing techniques for thermoset and thermoplastic matrix

Thermoset matrix Thermoplastic matrix

 Resin transfer  Extrusion


 Hand lay-up and spray up  Injection moulding
 Vacuum infusion or pressure bagging  Film stacking
 Pultrusion  Blow moulding
 Filament winding  Rotational moulding
 Autoclave moulding  Thermoforming
 Hot compression moulding

The simplest and cost effective form of composite manufacturing is the hand lay-up used for
thermoset matrix. It works well with natural fibre composites as it does not involve the
Page | 22
application of heat. However, this technique allows for a low fibre volume content (less than
20% wt.%) [70].
Factors determining composite properties include temperature, pressure, and speed of
processing [71]. If the processing temperature is too high, it is possible for fibre degradation
to occur, which limits thermoplastic matrices used, to those with low melting points than the
temperature at which degradation will occur. As a result, suitable matrix systems and
manufacturing methods need to be carefully selected. Biopolymers such as PLA (Tg of ~58°C
and a melting point of, Tm of 170°C) have lower processing temperatures which make them
compatible with natural fibres [72].

Thermoplastic matrices are often used in extrusion moulding in the form of beads or pellets
which are softened and mixed with fibres and transported using a single or twin screw;
compressed and forced out of the chamber through a shaping die. High screw speed can result
in air entrapment, excessive melt temperatures and fibre breakage, whereas low speeds lead
to poor mixing and insufficient fibre wetting in the matrix. This method is used on its own or
for producing pre-cursor for injection moulding. However, twin screw extruding systems
have better fibre dispersion and mechanical properties than single screw extruders [73]. Due
to the viscosity requirements, injection moulding for such composites is generally limited to
less than 40w% fibre content.

In RTM, a liquid thermoset resin is injected into the mould containing a fibre preform. The
main variables associated with this process are temperature, injection pressure, resin
viscosity, preform architecture and mould configuration. The advantages compared with
other processes include high volume parts, lower temperature requirements and avoidance of
thermomechanical degradation. The structure and nature of fibres and due to lower degrees of
fibre alignment, the compaction during processing is generally affected. However, natural
fibre composites are less compactable than glass fibre composites [74]. Good component
strength can be achieved with this process which is suitable for low production runs [75].
According to Brouwer [76], RTM and vacuum mouldings are suitable for composites
containing higher fibre volume fractions and preformed mats. O’Donnell et al. [77] showed
that VARTM was able to cure large-volume natural fibre reinforced soybean oil composites
at room temperature. A difference in VARTM is that only one side of the component has a
smooth surface finish.

Page | 23
2.4 Factors affecting NFC performance

2.4.1 Aspect ratio

Aspect ratio is defined as the fibre length divided by the fibre diameter (Ɩ/d). It is important
for determining the fracture properties in fibre reinforced composites. Plant fibres are
categorized into two types: long and short fibres. The aspect ratio identifies the long and short
fibres, if l/d≥100, it is identified as the long fibre while l/d<100 represent short fibres [8].
Long fibres, which measures up to 700 mm can be obtained from plant origins and is used for
spinning very fine yarns for various applications. This fibre is known as continuous fibres.
Short fibre which is less than 200 mm is obtained during fibre processing as waste and can be
used as raw material for coarser yarns; these are known as discontinuous fibres.

Length efficiency factors for strength and stiffness are directly affected by the reinforcing
geometry [fibre length (lf) and aspect ratio] which can be maximized by using high aspect
ratio fibres with fibre lengths longer than the critical fibre length. Shah [48] described the
critical fibre length as a function of the fibre tensile strength, fibre diameter (df) and the
interfacial shear strength. It is important to note that the length efficiency factors for stiffness
and strength range between 0 (for lf << df) and 1 (for lf >>df). Shah [48] used a graphical
demonstration using flax fibre reinforced polymer composite (Figure 2.1) to further explain
this theory. From the graph, as lf approaches 0.5mm and aspect ratio (l/d → 25), the length
efficiency factor value is about 0.80. Thereafter, the length factor asymptotically approach
unity as the fibre length or aspect ratio increases.

The critical length of bast fibre reinforced polymer composites range between 0.2-3 mm and
the fibre length is found to be typically >30 mm and have high aspect ratios between 100-
2000. Different composite moulding process is found to influence the fibre length and aspect
ratio. Bos et al [78] studied the mechanical properties of short flax fibre reinforced
compounds, manufactured by batch kneading and extrusion moulding. They determined the
length efficiency factor after injection moulding of flax composites to be between 0.17 and
0.20.

Page | 24
Eq. 2.1

Eq. 2.2

Figure 2.5 Predictions of the fibre length efficiency factors for a) stiffness ηlE and b)
strength ηlS, based on Cox’s shear lag model (Eq. 2.1) and Kelly-Tyson’s model (Eq.
2.2), respectively. Typical values for flax reinforced PFRPs are used in the calculations:
df = 20 μm, Gm = 1 GPa, Ef = 50 GPa, vf,max,FRP = π/4, vf = 0.30, σf = 1000 MPa and
τ = 30 MPa or 15 MPa (lc = 0.333 mm or 0.667 mm, respectively) [48].

2.4.2 Fibre volume fraction

Fibre volume fraction plays an important role in the properties of natural fibre reinforced
composites. It is often observed that an increase in fibre content leads to an increase in the
strength and stiffness of the fibre-reinforced composites. The impact strength of epoxy resin
was seen to reduce with the addition of up to 25wt% fibre content, and then increases to give
an overall improvement in impact strength of 40% at fibre content of 40wt% [79]. Roe and
Ansell incorporated jute fibre in polyester resin matrix to form uniaxial reinforced composites
containing 0.6 fibre volume fraction. Increases in modulus and strength were observed,
followed by a decrease. Further increase in volume fraction resulted in a decrease in the
properties, which was attributed to insufficient wetting of the fibres [80]. In banana fibre
reinforced polyester composites, the improvement in flexural strength was only observed
when the fibre content was more than 12 wt%. At 30 wt%, the composite material shows a
Page | 25
flexural strength up to 97.3 MPa, which is 33% greater than that of the neat polyester matrix
(73.1 MPa). The flexural modulus increased from 3.8GPa to 6.5GPa, as the fibre content
increases from 6 to 30wt%. The increase in stiffness of the composite was attributed to the
high fibre modulus as compared to that of the neat polyester matrix (4.1 GPa) [81].

2.4.3 Fibre Orientation

Fibre reinforced composites are processed to have isotropic (same properties in all directions)
and anisotropic (material properties are different in all directions) properties with the control
of the fibre orientation. Orientation of fibres may be unidirectional with fibres arranged in
one direction, randomly orientated fibres with no arrangement, biaxial orientation with fibres

arranged in longitudinal and transverse directions at 90 ̊. Different types of fibre orientation

in composites are presented in Figure 2.6. Due to the natural structure (cellulose based) and
larger aspect ratio along the axis of the fibre [48], most fibres are anisotropic in nature.
Randomly orientated fibres in composites are found to produce isotropic composites;
however the overall strength is decreased.

Unidirectional composites have been proven to have the highest tensile strength in fibre
direction than that of biaxial composite, hence the latter can be used in structural applications
[48]. While short fibre composites are normally used in non-structural applications, long
fibre composites gave superior values for flexural strength, modulus and toughness compared
to short fibre composites. Wazzan [82] investigated the tensile strength and modulus of date
palm fibre (DPF) reinforced polyester composites moulded using RTM. The composites were
prepared with long unidirectional and woven DPF and subjected to mechanical testing. Long
unidirectional fibres exhibited the highest tensile strength up to 86.1 MPa and 76.9 MPa for
woven fibres. In another study, Sathish et al [83] studied the effect of fibre orientation and
stacking sequence on banana-kenaf hybrid reinforced epoxy composites. The composites
were fabricated by hand lay-up processing with different fibre orientations (vertical,

horizontal and 45 ̊ and subjected to mechanical and thermal characterisation. The authors

concluded that the hybrid composite in which the fibres were arranged at 45 ̊ had better

properties than the others.

Page | 26
Figure 2.6 Different fibre orientation in composites: a) unidirectional, b) random, c)
biaxial and d) multi directional [84]

2.4.4 Interfacial adhesion

Natural fibres consist of cellulose, hemicellulose, lignin and pectin. Cellulose is the main
component rich in hydroxyl groups. Because of this, natural fibres are strong polar materials
that exhibit hydrophilic groups [75]. Moreover, most matrix materials are non-polar and
hydrophobic. As a result this causes incompatibility between the matrix and the natural fibre
because the fibre interacts alone or with water more strongly than it interacts with the
hydrophobic matrix. Natural fibres are subjected to variable conditions during their lifetime
including unsteady hygroscopic conditions [59]. In humid environments, the hydrophilic
behaviour of natural fibres leads to high levels of moisture absorption which results in
structural modification of fibres as well as the change in mechanical properties. Such
problems must be tackled by employing appropriate methods to improve adhesion between
fibre and the matrix. Good interfacial adhesion prevents energy absorption through interfacial
cracks, thereby reducing fracture toughness within the composite.
There are essentially two strategies to achieve the goal, the first being to alter the fibre
surface and secondly to modify the matrix properties. In both strategies, the aim is to improve
wettability of the matrix on the fibre surface and promote adhesion.

Page | 27
2.4.4.1 Natural fibre surface modifications

Physical methods for surface modification

Physical treatments change the structure and surface properties of fibres and therefore
enhance mechanical bonding between fibre and the matrix. Apart from mercerisation, which
involves the use of sodium hydroxide (NaOH), physical methods are said to be ‘clean’ since
they do not involve chemical agents [75].

Corona treatment

Corona treatment (sometime called air plasma) is a surface modification technique that
changes the surface energy of plant fibres, which in turn affects the melt viscosity of
composites [Figure 2.7(a)]. The corona plasma is formed by applying high voltage to an
electrode that has a sharp tip. The plasma forms at the electrode tip. A range of linear
electrodes is often used to create a curtain of corona plasma where materials are passed
through in order to change their surface energy. Prolonged treatment times may result in
fibres with a rough surface. The instrumentation for corona treatment makes it possible for
modification so that it can be used as a continuous process in industry [75]. Corona treatment
is widely used for the surface treatment of plastic film, extrusion, and converting industries.

Plasma treatment

Plasma treatment is another physical treatment method similar to corona treatment. Plasma
treatment as shown in Figure 2.7 (b) is a surface modification technique that readily primes
the fibre surface for better compatibility with the matrix by utilizing ingredients such as
electrons, ions radicals, and excited molecules produced by electrical discharge. Low
temperature plasma can be generated under atmospheric pressure in the presence of helium.
The action involves abstraction of protons and creation of unstable radicals that convert
functional groups such as alcohols, aldehydes, ketones, and carboxylic acids. Electrical
discharge methods are used for cellulose fibre modification to decrease the melt viscosity of
the composite [42]. The difference compared to corona treatment is the need for a vacuum

Page | 28
chamber and gas feed to maintain the appropriate pressure and composition of the gaseous
mixture.

a) Corona treatment

b) Plasma treatment

Figure 2.7 Corona (a) and Plasma (b) treatment

Alkali treatment/ Mercerisation

Alkali treatment or mercerization is one of the oldest and most used methods for treating
plant fibres. The important modification done by alkali treatment is the disruption of
hydrogen bonding in the network structure, thereby increasing surface roughness. The
process involves removing some amount of lignin, hemicellulose, wax and oils covering the
fibre surface [60] by soaking in a dilute solution of NaOH, thereby creating a rough surface
topography by revealing the fibrils, as shown in Figure 2.8. Different degrees of modification
are obtained with various concentrations of alkaline solution such as the temperature and
length of treatment. Mwaikambo and Ansell [85] treated hemp, sisal, jute and kapok fibres
with various concentrations of NaOH and found that 6% was the optimum concentration in

Page | 29
terms of cleaning the fibre bundle surfaces, and hence retaining a high degree of crystallinity
index. This is because at higher alkali concentration, fibres become weaker and damaged
resulting in decreased mechanical properties.
Ouajai and Shanks [86] studied the effect of alkali treatment on hemp fibres and found that
pectin and hemicellulose were removed by the treatment. They also found that an increase in
alkaline concentration and longer enzyme scouring time disrupted the fibre crystalline
structure resulting in the presence of less ordered cellulose structure.
Studies on tensile properties of kenaf fibres treated with NaOH (concentration of 3%, 6% and
9% for 24 hours) was conducted [87] and the results showed an increase of 0.4% in tensile
properties of treated fibres compared to untreated kenaf fibres at an optimum concentration of
6% NaOH. The main finding of the study was that the treatment of fibres by alkalization
improves the mechanical interlocking and chemical bonding between the matrix and fibres
resulting in better mechanical properties.

Figure 2.8 Schematic representation of untreated and alkali treated natural fibre [88]

Chemical methods for surface modification

Esterification-based treatments

Esterification-based methods refer to the use of various chemicals that have the ability to
form ester bonds with the fibre surface [75]. Cellulose fibre surfaces are rich in OH groups,
so the chemicals used are carboxylic acids and their derivatives (anhydrides, chlorides, etc.).

Page | 30
This OH groups are responsible for the hydrophilic nature of fibres and must be coated with
molecules that have a hydrophobic nature. Acetylation, benzoylation, treatment with maleic
anhydride and treatment with stearates are some of the widely used chemical processes. Of
these four methods, acetylation is by far the most commonly used method.

Acetylation is a treatment that introduces plasticization, and has been shown to improve
dimensional stability and environmental degradation [58]. An acetyl functional group is
introduced into the reaction. The process substitutes an acetyl group for an active hydrogen
atom, and the resultant acetic acid generated from the reaction must be removed from the
fibre material before it is used.
Chemical modification with acetic anhydride substitutes the polymer hydroxyl groups of the
cell wall with acetyl groups, modifying the properties of these polymers so that they become
hydrophobic. The reaction below shows the process of acetylation.

Figure 2.9 Acetylation of natural fibre [88]

In acetylation reaction, acetic anhydride is preferred compared to acetic acid because the
latter does not sufficiently react with cellulose in fibres. Acetic anhydride is not a good
swelling agent for cellulosic fibres. Hence, the reaction must be accelerated by initially
soaking the material in acetic acid and subsequently treating with acetic anhydride to reduce
the hygroscopic nature of fibres [88].

Silane coupling agents

Coupling agents are chemical modifications that involve the introduction of a third material
that remains in the composite to act as a compatibilizer between the hydrophilic fibres and
the hydrophobic matrix. Coupling agents are molecules that possess two functions: first is to
react with OH groups of the cellulose fibre and the second is to react with functional groups
of the matrix [89]. The most common coupling agents used is silane, isocyanate and titanate

Page | 31
based compounds, of which their chemical composition allows them to react with fibre
surfaces, to form a bridge of chemical bond between the fibre and matrix.
Silanes are by far the largest group of coupling agents used in composite industries [75].
Silanes are used as coupling agents to make natural fibres adhere to a polymer matrix, and
stabilizing the composite material. Silane coupling agents may reduce the number of
cellulose hydroxyl groups in the fibre-matrix interface. In the presence of moisture,
hydrolysable alkoxy groups lead to the formation of silanols. The silanol reacts with the
hydroxyl group of the fibre, forming stable covalent bonds to the cell wall that are
chemisorbed onto the fibre surface. As a result, the hydrocarbon chains provided by the
application of silane restrain the swelling of the fibre by creating a crosslinked network due
to covalent bonding between the matrix and the fibre [90]. The coupling action of silanes is
catalysed by the presence of solvents and initiators [89] and is represented in Figures 2.9 and
2.10. Silane treatment is found to reduce moisture absorption of fibres in humid
environments, effectively modify natural fibre and polymer matrix interface, and improve
mechanical properties. España et al [91] investigated the effect of coupling agents on basalt
fibre reinforced bio-based epoxy composites made using vacuum assisted resin transfer
moulding. The authors studied the influence of different silane coupling agents on the
mechanical properties of the composites and concluded that the addition of silanes increased
the compatibility between the fibre and matrix. After the addition of these, the mechanical
properties were significantly improved compared to samples without silane treatment.

Figure 2.10 Hydrolysis of silanol [90]

Page | 32
Figure 2.11 Reaction of silane with -OH groups of natural fibre [90]

Graft copolymerisation – treatment with maleic anhydride

Graft copolymerisation is an effective method of surface chemical modification of natural


fibres [75, 88, 89]. Graft copolymerisation reaction is initiated by free radicals of the
cellulose molecule. The cellulose is treated with an aqueous solution with selected ions and
exposed to high energy radiation. Thereafter, the free radicals are formed by cracking the
cellulose molecules. Thereafter, the radical sites of the cellulose molecules are treated with a
suitable solution. The aqueous solution must be compatible with the polymer matrix, for
example, vinyl monomer (i.e. acrylonitrile, methyl methacrylate, polystyrene, etc.) [92]. The
degree of compatibility between the cellulose fibres and the polymer matrix is determined by
the grafting efficiency, grafting proportion and the grafting frequency.

One of the important factors to be remembered during grafting to the fibre is that the
monomer cannot diffuse into the ordered regions such as crystallites, but only into amorphous
or disordered regions. Since the mechanical behaviour of the fibres is largely determined by
the ordered structure, grafting does not bring about any drastic changes in the properties [89].

Treatment with maleic anhydride (MA) is the most popular grafting method. The treatment
of natural fibres with polypropylene-maleic anhydride (MAH-PP) copolymers provides
covalent bonds across the interface [93].The treatment increases the fibre surface energy to a

Page | 33
level much closer to the surface energy of the polymer matrix, thereby providing a better
wettability and a higher interfacial adhesion. The mechanism of reaction can be divided into
two steps: a) activation of the copolymer by heating and b) esterification of the cellulose

Figure 2.12 Chemical reaction of cellulose fibres with maleated PP [92]

Chemical treatments have shown to increase the interfacial adhesion between the fibre and
matrix, and hence improve the mechanical properties of the fibre composites. Treatment of
natural fibres also reduces the hydrophilicity and swelling of the fibre by creating a
crosslinked network of covalent bonds between the matrix and fibre. Alkali treatment is one
of the commonly used and efficient methods of surface modification with almost all natural
fibres with successful results.

2.4.5 Moisture absorption

Components used in building constructions are exposed to moisture and humidity during their
lifetime; hence information on moisture absorption and weathering is important in the
building sector. Natural fibres from plants are cellulose based and swell when they come into
contact with water or organic liquids, resulting in poor mechanical properties and
dimensional stability [94]. The swelling of natural fibres is dependent on parameters of the
wetting liquid (including hydrogen bonding capability, molecular size, cohesive energy,
viscosity, density, and basicity), parameters of fibres (including surface coating and
treatment, density, porosity, and crystalline structure of the fibres), the ambient temperature
and the steric effects [95]. Natural fibres can absorb and desorb moisture from the

Page | 34
atmosphere and equilibrate with the surrounding environment. Hence, natural fibres undergo
swelling in humid and dry environments affecting their dimensional stability. Water
absorption in natural fibre composites is governed by mechanisms such as [96]:
1. Hydrogen bonding of the water molecules to the free OH-groups present in the
cellulose materials
2. Diffusion of water molecules into the fibre-matrix interface caused by the capillary
action
To reduce water absorption and swelling of natural fibres, a number of studies have been
conducted including the use of coupling agents [97] to bond with OH-groups in the cellulose
material and the use of chemical treatments [98].

In polymer composites, the transport of water is facilitated by three mechanisms [93, 99]. The
first involves the diffusion of water molecules into the micro-gaps between polymer chains.
The second mechanism involves capillary transport into the gaps and flaws at the interfaces
between the fibre and polymer, resulting in poor wettability and impregnation (during
composite manufacturing). The third involves the transport of water molecules by micro-
cracks in the polymer as a result of fibre swelling (particularly in natural fibre composites).

Of these three mechanisms, diffusion mechanism majority influences the water absorption in
polymer composites. Diffusion behaviour is further classified according to the relative
mobility of the penetrant of the polymer segments which is known as Fickian diffusion
model, non-Fickian or anomalous, and an intermediate behaviour between Fickian and non-
Fickian. However, the water absorption pattern of natural fibre composites at room
temperature tend to follow the Fickian behaviour whereas at higher temperatures the
absorption of water is non-Fickian [100]. In Fickian diffusion, the rate of diffusion is much
less than that of the polymer segment mobility. Figure 2.13 present the Fickian and non-
Fickian diffusion.

Page | 35
Figure 2.13 (a) Fickian diffusion at room temperature and (b) non-Fickian diffusion at
elevated temperature [101]

Composite materials absorb water when they are exposed to humid environments. The water
absorption continues until saturation. The water absorbed in composites consists of free water
and bound water [102]. Free water are water molecules that are able to move freely through
the micro voids and holes whereas bound water are dispersed water molecules that are bound
to the hydrophilic groups of the polymer matrix. At high humidity levels, water penetrates
and attaches onto the hydrophilic groups of the fibre, establishing intermolecular hydrogen
bonding with fibres and hence, reducing the interfacial adhesion of fibre-matrix [103]. The
degradation process occurs when swelling of cellulose fibres develops stress at interface
regions causing micro-cracking of the matrix around swollen fibres and this promotes
capillary and transport of water via micro-cracks. Water soluble substances start leaching
from the fibres and eventually lead to ultimate de-bonding between the fibre and the matrix.
After long periods of water exposure, biological activities such as fungi growth degrade the
natural fibres [101].

The water absorption behaviour of a composite depends on the type of fibre and matrix, the
fibre content, fibre orientation, atmospheric conditions like temperature and humidity, water
distribution within the composite and reaction between water and matrix, permeability nature
of the fibre, and characteristics of the surfaces exposed to water [99, 104].

Page | 36
Several researchers have studied the effects of water absorption on properties of natural fibre
reinforced polymer composites. They also reported on the effects of fibre content, fibre type
and matrix type on moisture absorption properties.

Venkateshwaran and Elayaperumal [105] studied the effect of layering sequence on the
mechanical (tensile, flexural and impact) properties of woven jute/banana hybrid composites
fabricated by hand lay-up. Epoxy resin LY556 with hardener HY951 were used to produce
composites reinforced with jute, banana, jute/banana/jute (JBJ) and banana/jute/banana (BJB)
fabrics. It was observed that the composite reinforced with BJB achieved higher tensile and
flexural properties followed by composite reinforced with JBJ and the least by the composite
reinforced by only banana fibre fabric. This shows that the addition of banana fibre as skin
layer increases the strength and stiffness. Hybridisation improves the properties of the
composite; the layering sequence has much more effect on the flexural properties of the
composite. The moisture absorption were 4.6%, 6.4%, 6.8% and 7.4%, respectively; for
composites reinforced with banana fibres, jute fibres, BJB, JBJ fabrics. The results showed
that the composite reinforced with three-layers JBJ has the highest moisture absorption of
7.4% because of the moisture affinity of jute fibres. Jute fibre showed the highest sorption,
diffusion and permeability coefficients than banana fibre resulting in high moisture
absorption capability of the composite reinforced JBJ. Hybridisation of jute fibres with
banana fibres slightly decreased the moisture uptake.

In another study [104], composites produced from thermoset matrices (epoxy and polyester)
reinforced with carpet waste jute yarns were investigated. Waste jute yarns were treated with
10 and 25 wt% alkali solutions. The composites were compression moulded as single-ply (4/1
lea) and double-ply (4/2 lea) yarn composites. For flexural and impact tests, the samples were
subjected to water submersion to investigate the effects of water absorption on the
mechanical properties. The percentage weight gain of samples was measured at different
intervals and the magnitude of diffusion parameters such as n and K was evaluated to assess
the mode of water transport. It was found that the composites follow Fickian diffusion
behaviour. Both the water content (Mmax) and effective diffusion coefficient (Deff) values
increase with the increase in fibre content. The K value was also found to increase with
increasing temperatures, hence the Deff value increased remarkably. This indicated that the
interaction between water and the composites was higher at elevated temperatures, thus

Page | 37
increasing the mobility of polymer segments and creating additional voids for water to
penetrate.
The flexural strength and modulus decreased with water uptake for all composites.
Nonetheless, impact strength values increased with water absorption for all composite
samples and this was caused by the plasticizing effect of water on the matrix and hence, the
effects of water on fibre-matrix interface. The composites reinforced with double-ply yarns
showed higher water uptake than that of single-ply yarns.

2.4.6 Weathering in natural fibre reinforced composites

Natural fibre reinforced composites exhibit excellent mechanical properties especially when
their low density and price are taken into account. However, due to their unknown long term
properties when exposed to a combination of in-service loads and environments,
manufacturers are reluctant to use natural fibre composites in primary load bearing structures.
Durability of natural fibre composites upon exposure to UV-light is of great concern as UV
cause changes in the surface structure of the composites, a mechanism known as photo-
degradation. Photo-degradation ranges from mere surface discoloration that is affected by the
aesthetic appeal in indoor applications to extensive loss of mechanical properties [106]. Even
in the absence of a significant amount of UV absorption, small amounts of impurities can be
sufficient to initiate polymer degradation. Moreover, the combination of light, temperature,
and moisture in outdoor applications can completely destroy the lignocellulosic network,
limiting the performance of unprotected composites for outdoor applications [107]. The
degradation of natural fibre components such as cellulose, hemicellulose, and lignin attribute
to the photo-degradation of NFCs. Lignin and hemicellulose are more prone to degradation
than cellulose by various means. Lignin degrades upon exposure to UV-light and
hemicelluloses degrade upon moisture absorption and biological attacks. The UV degradation
process is known to be triggered by the formation of free radicals and probably starts with the
oxidation of phenolic hydroxyls [108]. Moreover, singlet oxygen that can be formed by
oxygen quenching of photo excited lignin plays a significant role in the degradation of
lignocellulosic materials like wood or lignin-rich mechanical pulps. The formed singlet
oxygen is a source of peroxides [109], which can initiate the auto-oxidation of carbohydrates
and cleavage of lignin [110, 111]. The rate of lignin degradation increases in the presence of
singlet oxygen sensitizer in pulps [112]. Degradation of polymer composites as a result of

Page | 38
photo-oxidation has undesirable effects such as loss of strength, stiffness, and surface quality.
These effects can be inhibited by adding UV stabilizers into formulations [106].

In order to study the durability of composite materials, weathering characteristics must be


thoroughly evaluated. Shokrieh et al. studied the effects of UV radiation on mechanical
properties of glass/ polyester composites. Since photo-degradation occurs on the surface and
is restricted to degradation of the mechanical properties of the resin, polyester resin properties
were evaluated. The authors prepared and treated test samples of polyester resin in an
artificial UV chamber at three different time intervals, equivalent to 3, 6 and 12 months then
subjected to tensile and compression tests. After 1000hrs of exposure the specimens showed a
decrease of up to 15% in average failure under tensile load, a decrease of up to 30% in
ultimate strength and 18% in tensile modulus. However, the specimens had no difference in
colour after exposure but behaved as brittle materials [113].

Chin et al. studied the effects of UV radiation in polymeric matrix resins including vinyl ester
and isophthalic polyester which are commonly used in construction applications. Neat
polymer films were exposed to UV radiation for 1200hrs and changes in viscoelastic and
thermal properties were evaluated through dynamic mechanical thermal analysis (DMTA)
and X-ray photoelectron spectroscopy (XPS). The authors reported that the UV exposed
samples showed no differences in glass transition temperature in comparison to the un-
exposed samples. However, the XPS showed dramatic changes in the chemical composition
of UV exposed samples [114].

Islam et al [115] studied the influence of accelerated ageing on alkali-treated hemp fibre/PLA
composites. Accelerated ageing was carried out using UV radiation and water spray at 50°C
for four different time intervals (200, 500, 750 and 1000 hrs). The tensile strength, flexural
strength, Young’s modulus, flexural modulus and fracture toughness were found to decrease
whereas the impact strength increased for both untreated and alkali treated hemp fibre/PLA
composites. The aligned untreated hemp fibre reinforced PLA composites (AUL) had the
greatest overall reduction in mechanical properties than that for the treated composites upon
exposure to accelerated ageing environment. The deterioration of the AUL composites was
found to be initiated by the fibre bridging followed by PLA film rupture through micro-
cracking and then fibre pull-out. This was attributed mainly to the stresses produced by

Page | 39
differential swelling and shrinkage of the fibre/ PLA caused by changes in moisture content.
The difference between the coefficients of thermal expansion for fibre and matrix caused
stresses to build up in the fibre-matrix interface, leading to failure of composite. Increased
exposure to weathering conditions led to the initiation of rupture in the fibres, which was due
to the degradation by UV attack. This was further supported by FTIR analysis and
crystallinity content results of the deterioration in mechanical properties [115].

When NFCs are exposed to heat, UV radiation, and moisture there is a decrease in
mechanical properties which can be attributed to a combination of several changes such as
composite surface oxidation, matrix crystallinity changes and, interfacial degradation [116].

2.5 Applications of natural fibre reinforced composites

Natural fibre reinforced composites are being targeted in different fields due to both
environmental and economic benefits.
Automotive applications: The automotive industry represents the biggest consumer [48] for
natural fibre reinforced polymers because of the low density of natural fibre that makes it
possible to produce lightweight materials that are cost and fuel effective. Compared to
synthetic fibres, natural fibres exhibit better insulation, crash and absorbent properties, which
is essential in automotive applications [117]. Door and ceiling panels, panels separating the
engine and inner passenger compartments (Figure 2.14) are some of the vehicle components
where natural fibres are used [26].

Figure 2.14 NFC applications in the automotive industry [118]

Page | 40
Recent developments include the use of a new type of resin, Acrodur, where needle-punched
kenaf and hemp fibre mats were impregnated with Acrodur, an aqueous acrylic resin
developed by BASF (Ludwigshafen, Germany), by vacuum press moulding. The resin is a
green, formaldehyde free acrylic thermoset made with a combination of polycarboxylic acid
and polyalcohol to form polyester. Acrodur is an ideal thermoset material for high
performance and emission-critical processes and applications, be it in abrasive nonwovens for
household, industrial, automotive and filter applications. At temperatures below 130 ̊C the
resin exhibit thermoplastic characteristics and gives the material good storage stability
compared to thermoset resins. Following cure (above 150 ̊C) the material exhibits thermoset
properties, becoming hydrophobic, tough and wear resistant [119]. The manufactured
material was intended to replace glass reinforced thermoplastics in automotive interior parts,
however a variety of components such as executive cases, trays, and a sport seat (Figure 2.15)
are currently manufactured out of the new fibre composite.
Acrodur is capable of producing composites with very high fibre loading of up to 70%, with
less than 1.0 g/m3 specific gravity. The produced composite is lightweight, thin and durable.

Figure 2.15 Vacuum press-moulded sport seat [119]

Medical applications: Researchers have explored various NFC for the biomedical
applications including skin tissue repair materials, drug delivery systems, blood vessel
growth, scaffolds for tissue engineering, stent covering and bone reconstruction [120]. The
use of NFC for biomedical applications should allow for long-term use in the body without
being rejected, must be biocompatible and biodegradable so as to allow the body’s own cells
to replace the implanted scaffold [37]. Polymers composites used in medical applications

Page | 41
include polyurethane and polyvinyl alcohol. Their use as bio-composites in such delicate
applications is due to their excellent stress-strain properties that were found to be similar to
the heart valve tissue [121]. Natural polymers such as bacterial cellulose have a great
potential in wound healing system such as biocompatibility, conformability, elasticity,
transparency ability to maintain moisture in the wound and absorb exudates during
inflammation. If mass produced, bacterial cellulose will be a biomaterial used to create a
wide variety of medical devices and consumer products [122].
Sports and Leisure: An acoustic guitar from flax linen fabrics reinforced with a thermoset
bio resin (Ekoa) was developed and found to exhibit excellent stiffness in two directions
better than the traditional ones. The Ekoa natural composite guitar has a glossy surface finish
that does not require finishing. The material is also durable and humidity and temperature
resistant [123]. Jute fibre composites is currently used in the longboard industry to replace
traditional glass fibre composites [124].
Other potential applications: The use of natural fibres in making furniture could reduce
waste, as the furniture industry is increasingly utilizing agricultural waste fibres [125].
BASF’s Acrodur resin is currently used in various applications as mentioned previously.
FlexForm MT (nonwoven mat) and Acrodur resin was developed to make outdoor chairs and
office furniture [126]. Natural fibre reinforced blades for rooftop wind turbines were made
from flax fibres. The Biotex flax used for turbine blades offer a unique twist less technology
that gives performance characteristics equal to or better than the blades currently in the
market [127].

Page | 42
(a) (b)

(c) (d)

Figure 2.16 Various NFC applications: (a) Biomedical applications [122] (b) wind
energy [127], (c) Sports and leisure [123, 124], (d) furniture [126]

Natural fibre applications have expanded their use in geotechnical engineering where they are
mainly used in civil engineering applications to improve soil structural performance. Recent
studies have looked at new geotextiles known as “Limited Life Geotextiles (LLGs)”, a
reinforcing fabric only required to perform its duties for a limited time. LLGs are used in
applications for temporary roads over soft land as well as basal embankment reinforcement to
help supress extreme fluctuations of soil temperature and increase soil moisture [128]. In the
marine sector, a kayak prototype made from flax fibre and PLA was manufactured by
companies together with research centres. Their primary goal was to produce high
performance green composites to be used in small pleasure boats. The durability of the
prototype has been tested in sea water and was found to be sea water resistant [117].

2.5.1 Building applications

Composite materials were first employed by people of earlier civilization for a number of
household applications. The use of natural fibres as reinforcement in composite materials
dates back to 3,000 years ago when Egyptians utilized clay reinforced with straw to build
walls [41]. Later in the years, interest in natural fibre reinforced composites diminished due

Page | 43
to introduction of more durable construction materials like metals. Then in the 1940s, glass
reinforced plastics (GRP) and unsaturated polyester resins emerged and reached commodity
status [117]. The rise of these composite materials began when glass fibre was combined with
tough rigid resins and produced in large scales.

In recent decades, wood has been extensively utilized to reinforce cement composites in
products such as non-structural building materials. Recent examples of wood cement
products include slate roofing, cladding, architectural elements, and lumber substitutes such
as fascia, trim soffit, and corner boards. Wood fibre came about as a replacement for asbestos
in cement products for the building industry due to properties such as better durability and
high fire and moisture resistance. Furthermore, this product has been treated to be resistant to
rot and insect attack [8]. Wood based composites are also used to manufacture composite
panels that are one of the important building products. Panels made from wood fibre
reinforced plastic composites (WPC) are attractive because wood fibre is easily available in
sawmill wastes [4] and extremely strong and lightweight [7].
Good properties of thin walled elements such as high strength in tension and compression,
made of sisal fibre reinforced composite, give it a wide range of application, for example,
structural building members, permanent formwork, tanks, facades, long span roofing
elements, and pipes strengthening of existing structures [129]. Bamboo fibres can also be
used in structural concrete elements as reinforcement, while sisal and coir fibre composites
have been used in roofing components in order to replace asbestos [130]. In Asia, bamboo
fibres were first utilized after the World War II for a building center in Japan. Since then,
researchers have explored a variety of agricultural wastes for building materials.

Construction material manufacturers have resorted to utilize composites in building structures


such as dams, bridge decks, roads, sidewalk, and paving, where durability is a great concern.
To improve durability, additives such as fly ash, slag and silica fume are added to decrease
the chemical interactions between the matrix and fibres. Fibres can act as a primary
reinforcement to increase the strength and toughness of composites or as secondary products
in components such as pavements and slabs to control the cracking induced by environmental
conditions [8]. In general, fibre bridges across the cracks, and make post-cracking ductility.
Fibres act as crack arrestors and absorb energy when incorporated into matrix materials,
resulting in improved strength of the overall composite.

Page | 44
Bagasse, the fibre from sugar cane processing was first utilized in America in the 1920s to
manufacture composite panels [8]. Currently, building panels and roofing sheets made from
bagasse fibre reinforced phenolic composites are being used in houses in Philippines, Jamaica
and Ghana [7]. Traditional roof tiles are predominantly made from materials such as clay,
cement and steel, which have disadvantages of weight that result in high design (dead weight)
and installation loads (live weight). Another disadvantage of such materials is the negative
environmental impact associated with high embodied energy, material waste and pollution
from the manufacturing processes (air, land and water) [57]. Roof sheet materials must be
designed to sustain dead load, live load, wind load and in other cases, snow load. The
material must be lightweight, fire resistant, water resistant and weather resistant (such as
resistance to ultraviolet light) [131]. Roof sheet materials that have substituted traditional
ones include the utilization of recycled paper reinforced cellulose fibre. Dweib et al [12] used
bio-based materials to manufacture a NFC sandwich roof structure and structural beams.
Recycled paper was incorporated into composite sheets and structural unit beams and the
resulting strength and stiffness was found to be comparable with the currently used materials
for roof construction [Figure 2.18 (d)] [12]. In another study [10], sisal fibres were manually
cast on corrugated roof sheets thereafter, strength of the sheets was evaluated. The authors
concluded that the strength of sisal fibre corrugated roof sheets (towards splitting due to
direct and impact loads) was improved compared to unreinforced corrugated sheets. John et
al [130] concluded that sisal and coir fibre composites have the potential to replace asbestos
in roofing components.

Natural fibre composites have the potential to replace petroleum based products in structural
applications by making use of green materials. A sustainable hurricane-resistant housing
structure (Figure 2.17) was developed in South Carolina of USA by the ACRES group
following Hurricane Andrew [132]. The high performance, low cost house was made of
chicken feathers, recycled paper, straws and soybean oil resin, making this a fully bio-based
composite. The roof structure composite was moulded by vacuum infusion and bonded with a
foam block to make a sandwich panel [132]. This type of roof panel provides thermal
insulation and can fulfil a primary structural function. The sustainable housing structure
proves that substitution of glass fibres by natural fibres is feasible. Compared to zinc and
iron, natural fibre reinforced composites are better insulators, renewable and are locally

Page | 45
available resources. Furthermore, traditional materials used for corrugation applications are
zinc coated. Zinc coated steel causes pollution in long term and leads to rusting. In hot areas,
steel roofs provide no insulation and heat under such roofs is unbearable.

Other uses of renewable materials in buildings and infrastructure ranges from flooring,
furniture, roofing sheets, bridges, to pavement pathways, in replacement of petroleum-based
systems. Green kitchens are developed from bio-products such as wheat/ rice straw fibres
reinforced with polyurethane resin instead of formaldehyde to manufacture cupboards, while
countertops are made from hemp fibres due to the strength required for the product. Modern
bridges called the stay-in-place bridge forms (SIP) made from bio-composites are now
installed to replace traditional steel form bridges. Bio-composite based SIP forms are porous
and breathable, leading to water evaporation through the form and preventing corrosion. In
addition, the SIP form is lighter compared to a steel form allowing for faster and cheaper
installations [125].

Figure 2.17 Schematic presentation of a monolithic house roof made of natural


composites as a hurricane-resistant roof [132]

Other examples include architectural cladding screens made from cereal straw fibre/ PLA by
cold press moulding [133] and pedestrian bridge made from hemp and flax fibres reinforced
PLA [134] [Figure 2.18 (b)]. The hemp house [135] built in South Africa is considered the
“most sustainable building in Africa” [Figure 2.18 (c)]. The house is made of layers of
hempcrete, hemp insulation and hemp plaster. Modern roofing shingles are made of post-
industrial and post-consumer recycled rubber, plastic and cellulose fibre bound together with

Page | 46
polymer additives to look like cedar shakes or slate tiles. A representation of shingles is in
Figure 2.18 (a) below. These shingles assist with lowering the building energy consumption
[125].

(a) (b)

(c) (d)

Figure 2.18 Photographs of (a) Roof shingles, [125] (b) Eindhoven bio-bridge, [134] (c)
the Hemp house [135] and (d) NFC sandwich roof construction [12]

Page | 47
CHAPTER 3: EXPERIMENTAL TECHNIQUES

3.1 Materials

3.1.1 Flax fibres

Flax fibres (Linum usitatissimum) were obtained from a farm in Graaff Reinet, Eastern Cape,
South Africa. Table 3.1 shows the mechanical and physical properties of flax fibres.

Table 3.1 Physical and mechanical properties of flax fibre [23, 136, 137]

Properties Value
Density (g/cm3) 1.4-1.5

Length (mm) 5-900

Diameter range (µm) 40-600

Tensile strength (MPa) 345-1500

Tensile modulus (GPa) 27.6-80

Elongation at break (%) 1.2-3.2

Page | 48
3.1.2 Kenaf fibres

Kenaf fibres (Hibiscus cannabinus L.) were supplied by a company in Bangladesh.

Table 3.2 Physical and mechanical properties of kenaf fibre [137, 138]

Properties Value

Density (g/cm3) 1.4

Length (mm) -

Diameter (µm) 140

Tensile strength (MPa) 223-930

Tensile modulus (GPa) 14.5-53

Elongation at break (%) 1.5-2.7

3.1.3 Epoxy resin

Epoxy resin, Super Sap (CLR 04) was supplied by Aerontec, South Africa. As mentioned in
the manufacturer’s technical data sheet, Super Sap CLR Epoxy is a modified, clear liquid
epoxy resin. As opposed to traditional epoxies that are composed primarily of petroleum-
based materials, Super Sap formulations contain bio-based renewable materials sourced as
co-products or from waste streams of other industrial processes, such as wood pulp and bio-
fuels production. The properties of the bio-based epoxy resin as obtained from the technical
data sheet are presented in Table 3.3. Super Sap INH hardener was used as the curing agent
(cycloaliphatic polyamine).

Page | 49
Table 3.3 Mechanical properties and processing data of epoxy resin – from the
supplier’s technical data sheet

Properties (Epoxy/ Hardener) Description

Mechanical*

Tensile strength (MPa)* 68

Tensile modulus (MPa) 3227 (ASTM D-638)

Elongation (%)* 5

Flexural strength (MPa) 101 (ASTM D-638)

Flexural modulus (MPa) 2965 (ASTM D-790)

Onset Tg by DSC (°C) 44 (ASTM D-3418)

Ultimate Tg by DSC (°C) 73 (ASTM D-695)

Bio-based carbon content 21 (ASTM D-6866)

Processing

Mix by ratio (by volume) 2:1

Mix by ratio (by weight) 100:48

Mixed specific density (@ 25°C) 1.09

Viscosity (A/B/mixed, cPs, @ 25°C) 1850/500/1030

Tack free time (hrs, @ 25°C) 8

Recommended full cure 7 days @ 25°C, post cure recommended

*All performance data was taken from neat resin samples that underwent an initial cure at
room temperatures for 24 hrs and a post cure at 50°C for 2hrs

Page | 50
3.2 Preparation of flax and kenaf nonwovens

3.2.1 Needle punching technique

Flax (100%) and kenaf (100%) fibres were first mechanically cottonised using a Temafa
cottonisation machine. Cottonisation is necessary to reduce the fibre roughness, although the
process reduces fibre length and diameter. The cottonised fibres were blended with
polypropylene carrier fibres in the ratio 90:10 in order to improve the cohesion between the
natural fibres. The needle punched nonwoven fabrics, shown in Figure 3.1, were processed
using a Dilo needle punching machine. (Processing parameters are given in Table 3.4).
Nonwovens produced with areal density of 200 g/m2, 250 g/m2, 300 g/m2 and 350 g/m2 were
used for producing composites.

Table 3.4 Processing parameters used for preparation of needle punched kenaf and flax
nonwovens

Areal density(g/m2)
Parameter
200 250 300 350

Feed speed 0.6 0.6 0.6 0.6


(m/min)

Output 1.20 1.15 1.10 1.0


(m/min)

Needle depth 5 5 5 5
(mm)

Stroke (min-1) 350 350 350 350

Page | 51
Figure 3.1 Needle punching process to produce nonwovens

3.3 Fabrication of composites

3.3.1 VARTM technique

Composites were fabricated from bio epoxy reinforced flax and kenaf nonwovens. The
nonwovens were cut (100 cm x 60 cm) into rectangular shapes (Figure 3.2a) and weighed
before drying in an air oven at 80°C for 24 hrs. After drying, the nonwovens were re-weighed
and carefully arranged on the clean mould to begin processing. The mixing ratios for the bio-
based epoxy system (resin-hardener) were 100:48 by weight. The resin mixture was manually
mixed using a wooden stick. This process increases the formation of air bubbles within the
mixture. As a result the resin is allowed to stand for a minute to release entrapped gas in the
bubbles.
The process begins with placing the nonwoven fabrics on the mould, followed by a release
film, and a peel ply cloth to absorb excess resin. A breather ply is placed followed by a
vacuum bag mounting the entire assembly. The edges were sealed with a tacky tape and
vacuum extracted from within the bag using an atmospheric pressure. The bio-based epoxy
was infused into the mould through an inlet port and exhausted via an outlet port located on
the parallel side, as depicted in Figure 3.2b. Following complete mould filling, the resin was
allowed to run until no bubbles were observed at the mould surface. The vacuum was then
released and the resin outlet port shut-off. The laminate (Figure 3.2c) was cured for 24hrs at
room temperature, thereafter, cured further for seven days at ambient temperatures.

Page | 52
(a) (b)

(c)

Figure 3.2 (a) Nonwoven, (b) Vacuum infusion setup and (c) Prepared composite

3.4 Characterisation of composites

3.4.1 Mechanical Properties

3.4.1.1 Tensile testing

Tensile test is one of the most important testing requirements for composites in both research
and industrial environments. The specimen is prepared and loaded under specific conditions
depending on applications of the material. The tensile strength, tensile modulus and
elongation at break are measurements of the tensile properties determined by the tensile test.
Tensile tests were conducted on rectangular shaped specimens according to EN ISO 527-
4:1997 method. An Instron universal testing machine (model: 3369) with a load cell of 50kN
was used with a crosshead speed of 2 mm/min. A total of 25 specimens cut from the
fabricated composites were tested. Five specimens (250 mm x 25mm) were tested for each

Page | 53
sample. The Instron’s Bluehill 2 Material Testing Software for data acquisition was used to
analyse the raw data. Tensile testing was carried out on the following samples:
 Water immersed samples (wet)
 Dry samples
 Aged samples (UV radiation)

3.4.1.2 Flexural testing (three point bending)

Flexural tests were conducted according to EN ISO 14125:1998 method to determine the
flexural properties of the composites. Universal Instron tester (model: 3369) with 10kN load
cell was used to carry out three-point bending tests with a span length of 64 mm between
supports and a crosshead speed of 2 mm/min. A total of 25 rectangular specimens cut from
the fabricated composites were tested. Five specimens (80 mm x 15mm) were tested for each
sample. The average values for flexural strength and flexural modulus were reported.
Flexural testing was carried out on the following samples:
 Water immersed samples (wet)
 Dry samples
 Aged samples (temperature, humidity and UV radiation)

3.4.1.3 Drop weight impact testing

Drop weight impact testing was carried out on an Instron Dynatup impact tester (model:
9250HV) depicted in Figure 3.3. The specimens were subjected to puncture testing according
to EN ISO 6603-2:2000 method. The machine was calibrated prior to specimen testing. Five
square specimens (150mm x 150mm) were tested for each sample at room temperature. The
Instron’s software calculated important parameters such as the absorbed energy, which was
used to characterise the impact damage.

Page | 54
Figure 3.3 Schematic diagram of the impact test arrangement [139]

3.4.2 Dynamic mechanical analysis (DMA)

DMA is a widely used technique to investigate the structure and dynamic properties of
polymeric materials based on viscoelasticity. It characterizes the mechanical responses of a
material by monitoring changes in property with respect to temperature or/and frequency of
oscillation. The technique separates the dynamic response of the material in two distinctive
parts: elastic (E’) and viscous (E’’). The elastic part (storage modulus) defines the energy
stored in the material during one cycle of oscillation. It also gives an idea of stiffness
behaviour and load bearing capability of a composite material, while the viscous part (loss
modulus) describes the energy dissipated as heat by the sample. The ratio of the loss modulus
to storage modulus is tan δ, and is often called damping. It is a measure of the internal
friction of the material and indicates the amount of energy lost in the material as dissipated
heat. It is related to the degree of molecular mobility in the polymer material. DMA is used to
study molecular relaxation processes (glass transition temperature and secondary transitions)
as well as to determine inherent mechanical (e.g. storage modulus, loss modulus, damping
factor) and flow properties as a function of time, temperature, or frequency [140, 141].

Page | 55
Dynamic mechanical analyses were performed on rectangular (50 mm x 10 mm) specimens
using a PerkinElmer DMA 800. The test was performed under three point flexural mode at a
frequency of 1Hz and the heating rate was 2°C /min. The temperature range was 25-140°C.
The storage modulus (E’), loss modulus (E”) and tangent delta (tan δ) of the samples were
registered.

3.4.3 Water absorption studies

Water absorption of the composites was carried out according to EN ISO 62:2008 method.
Five specimens of each sample (tensile test) were immersed in distilled water at an ambient
temperature for 28 days. The specimens were firstly dried in an air oven at 50°C for 24hrs
prior to water immersion. After 24 hrs the specimens were removed from water, dried and
immediately weighed using a digital scale. This process was repeated over 28 days. The
average weight of dry specimens and that after water immersion was obtained as follows:

𝑚𝑡−𝑚𝑖
mt (%) = × 100,
𝑚𝑖

Where mt is the weight of the sample at time t during water immersion and mi is the weight
of the dry sample at initial time.

3.4.4 Environmental ageing (UV radiation)

Natural fibre composites used for outdoor applications are usually exposed to direct sunlight
and therefore subjected to UV radiation which breaks the covalent bonds, causing yellowing,
colour fading, weight loss, surface roughening, and deterioration in mechanical properties.
The important durability properties for outdoor application of these construction products
include UV resistance [101]. To study the ageing properties of composites, accelerated
ageing chambers which simulate natural environment and the damaging effects of long term
outdoor exposures are used.

The specimens were placed in a Xenon-Arc chamber for UV radiation for 168hrs according
to EN ISO 4892-2A test standard. The weights of the specimens were recorded before and
after aging, thereafter, the tensile properties were compared to un-aged specimens.

Page | 56
CHAPTER 4: RESULTS AND DISCUSSION

4.1 Mechanical properties

The properties of natural fibre reinforced composites depend on a number of factors such as
fibre volume fraction, fibre-matrix adhesion, fibre aspect ratio, fibre orientation and stress
transfer at the interface. Theoretical models are also used to predict the composite properties
(strength and modulus) for two phase systems [142]. The properties of the fibre and matrix
are very important in improving mechanical properties of composites. The tensile strength is
more dependent on the interfacial adhesion between fibre and the matrix, while the modulus
is dependent on the fibre properties and fibre content. In order to improve the tensile strength,
a strong interface, low stress concentration and fibre orientation in load direction are required
whereas fibre wetting and high fibre aspect ratio determine the tensile modulus [5].

4.1.1 Effect of fibre content on flax and kenaf fibre reinforced bio-epoxy composites

4.1.1.1 Tensile properties of flax and kenaf fibre reinforced bio-epoxy composites

The effects of fibre content on the tensile and flexural properties (strength and modulus) of
flax and kenaf composites are evaluated. Flax and kenaf reinforced bio-epoxy composites
were produced with four different fibre content, 10%, 15%, 25% and 30%, respectively.

Flax fibre reinforced bio-epoxy composites

Figure 4.1 presents the effect of fibre content on the tensile properties of flax fibre reinforced
bio-epoxy composites. The results show that the incorporation of flax fibres in bio-epoxy
matrix increases the tensile strength followed by a decrease at 15% fibre content. With
further increase in fibre content up to 25% the tensile strength also increases to 41.5MPa. At
30% fibre content the tensile strength decreases to 34.3 MPa. The drop in tensile strength at
15% could be attributed to a processing defect in the sample and at 30% fibre content it is
attributed to agglomeration of flax fibres which causes inefficient stress transfer within the
composite. However, several factors affect the strength of fibre reinforced composites such as
the fibre strength, fibre volume fraction, fibre length and fibre orientation and in addition a

Page | 57
strong fibre-matrix adhesion is critical, which gives the maximum strength [21]. Fibres act as
load carriers in the matrix and provide strength and stiffness to the overall composite.
According to the rule of mixtures, the addition of fibres with high tensile strength to a matrix
with low tensile strength should result in an increase in tensile strength of the composite
provided there is good interfacial adhesion between fibre and matrix [48]. However, since a
decrease in tensile strength is observed it is suggested that the interfacial adhesion is poor.
Also the reduction could be due to the fact that necessary vacuum needed to produce the
composite was not sufficient which led to formation of voids within the manufactured
composite, resulting in composite failure. It can be seen that the incorporation of 25% flax
fibres in the bio-epoxy matrix increases the tensile strength by 30%.

The tensile modulus shows an increase with fibre content with maximum modulus (2541
MPa) being exhibited by composite containing 30% fibre content. A slight decrease in tensile
modulus is observed at 25%, however standard deviations are large to form any significant
conclusion. At 30% fibre content, the incorporation of flax fibres in bio-epoxy matrix
increases the tensile modulus by 43%.

2800
46 Tensile strength
Tensile modulus
44 2600
42
Tensile modulus (MPa)
Tensile strength (MPa)

40 2400
38
36 2200

34
2000
32
30 1800
28
26 1600
0 10 15 25 30
Fibre content (%)

Figure 4.1 Effect of fibre content on the tensile strength and modulus of flax fibre
reinforced bio-epoxy composites

Page | 58
Kenaf fibre reinforced bio-epoxy composites

Figure 4.2 presents the tensile strength and tensile modulus of kenaf fibre reinforced bio-
epoxy composites at different fibre content. The tensile strength shows a decrease at 10%
fibre content and as the fibre content increases, it can be seen that tensile strength increases.
Maximum tensile strength is exhibited by composite containing 30% fibre content. The
decrease in tensile strength at 10% fibre content could be attributed to the fact that at low
fibre content, the population of fibres is low to take part in stress transfer [143].

40 2800
Tensile strength
Tensile modulus
38 2600

36 2400

Tensile modulus (MPa)


Tensile strength (MPa)

34
2200

32
2000
30
1800
28
1600
26
1400
24
0 10 15 25 30
Fibre content (%)

Figure 4.2 Effect of fibre content on the tensile strength and tensile modulus of kenaf
fibre reinforced bio-epoxy composites

4.1.1.2 Flexural properties of flax and kenaf fibre reinforced composites

When load is applied on the composite the bottom edge is in tension and the top in
compression, however there is a mixed mode of tension and shear. The specimen fails when
the bending and shear stress reaches the corresponding critical value. The modes of failure of
the specimen under three point bending are obtained from the stress-strain curves which give
the overall strength of the composite. Flexural properties depend mainly on the fibre

Page | 59
orientation and the location of the resin rich areas. The flexural strength assesses the
capability of composites to withstand the bending before it ruptures.

Flax fibre reinforced bio-epoxy composites

When fibres are incorporated to the matrix they act as load carriers and stress is transferred
from the matrix to the reinforce fibres, leading to uniform distribution and improved
mechanical properties of the composite. Figure 4.3 shows the variations in flexural strength
and flexural modulus with fibre content. Both the flexural strength and modulus values are
found to increase with increasing fibre content up to a maximum of 25%. At 25% fibre
content, the incorporation of flax fibres in bio-epoxy matrix increases the flexural strength
and flexural modulus by 48% and 59%, respectively. Further increase in fibre content to 30%
decreases both the flexural strength and flexural modulus. This was attributed to the
formation of binderless agglomerates at high fibre content, which hinders stress transfer and
causes the stiffness to decrease. Similar results were observed in other studies [35, 80, 144].

75
3800
Flexural strength
Flexural modulus
70 3600

3400
Flexural modulus (MPa)
Flexural strength (MPa)

65
3200
60
3000

55 2800

2600
50
2400
45 2200

0 10 15 25 30
Fibre content (%)

Figure 4.3 Effect of fibre content on the flexural strength and flexural modulus of flax
fibre reinforced bio-epoxy composites

Page | 60
Kenaf fibre reinforced bio-epoxy composites

The effect of fibre content on kenaf fibre reinforced bio-epoxy composites is presented in
Figure 4.4. It can be seen that the flexural strength and flexural modulus increase as the fibre
content increases. The maximum flexural strength and flexural modulus at 30% fibre content
are 60.1 MPa and 3195MPa, respectively. The flexural strength and flexural modulus at 30%
fibre content are increased by 32% and 42% in comparison to bio-epoxy matrix. The
incorporation of kenaf fibres enhances the load bearing capability (tensile) and bending
strength (flexural) of the composites. The flexural strength is a combination of the tensile and
compressive strengths, which directly varies with the interlaminar shear strength [145].

64 3400
Flexural strength
Flexural modulus
62
60 3200

Flexural modulus (MPa)


Flexural strength (MPa)

58 3000
56
54 2800

52 2600
50
48 2400

46 2200
44
0 10 15 25 30
Fibre content (%)

Figure 4.4 Effect of fibre content on the flexural strength and modulus of kenaf fibre
reinforced bio-epoxy composites

Page | 61
4.1.2 Drop weight impact testing for flax and kenaf fibre reinforced bio-epoxy
composites

Impact testing

Building roof tiles and structures are exposed to most impacts from weather driven objects
such as tree debris and ice. Impact performance can be improved by increasing the thickness
of samples, however, the weight and cost need to be taken into account. Impact performance
need to be accurately quantified to optimise design of roof structures.
Impact testing are of two types: the pendulum and drop weight. In this study we used the
instrumented drop weight impact test. The advantages of drop weight impact test are:
 Testing can be done on complex moulded parts, for example corrugated roof
structures
 Direction of failure is uni-directional. Failure occurs at the weakest point in the
sample and propagates accordingly.
 Failure of samples is defined by the deformation, crack initiation or amplitude of
fracture and unlike in pendulum test, samples do not have to break into two parts
 Gives information on load on the specimen being continuously recorded as a function
of time/ specimen deflection prior to fracture.
Hence, the instrumented drop weight impact test is a better simulation of functional impact
exposures and closest to real-life conditions [146, 147].

In this study, the samples were placed in an instrumented drop weight impact tester and
impacted with a cross-head mass of 3.31kg. The response of the composite to impact is
measured and stored in the computer system. The energy absorbed in the composite is
calculated based on the conservation of energy principle and is calculated based on the
potential energy of the impactor, initial kinetic energy of the impactor at the time of impact
and instantaneous kinetic energy [139].

The energy absorbed by flax and kenaf fibre reinforced bio-epoxy composites during impact
test as well as their thickness and normalised impact energy values (energy divided by the
thickness) are presented in Table 4.1. Overall, it can be seen that composites reinforced with

Page | 62
flax fibre absorb the most energy to failure in comparison to kenaf fibre reinforced
composites.

Table 4.1 Impact energy absorbed during impact testing

Samples and their Thickness Energy absorbed Normalised impact


energy (J/mm)
fibre content (%) (mm) (J)

Flax-10 1.33 1.82 1.37

Flax-15 1.48 2.05 1.39

Flax-25 1.73 2.25 1.30

Flax-30 2.82 2.70 0.96

Kenaf-10 2.42 1.60 0.66

Kenaf-15 3.74 2.67 0.71

Kenaf-25 3.90 2.77 0.71

Kenaf-30 3.54 2.36 0.67

The total energy dissipated by the composite prior to impact damage is a measure of its
impact resistance. The impact strength of the composite is governed by two factors namely:
the capability of the fibre to absorb energy that can stop crack propagation in matrix and the
poor interfacial bonding which induces micro-spaces between fibre and matrix, resulting in
easy crack propagation [148].

It can be seen from Table 4.1 that for flax fibre reinforced composites, the impact energy
increases with increasing fibre content and the maximum is shown at 30% fibre content. On
the other hand, the impact energy for composites reinforced with kenaf fibre shows an
increase with the increase in fibre content, reaches a maximum at 25% and then registers a
decrease at 30% fibre content. Sreenivasan et al reported that the capability of polymeric
composites to absorb high amount of energy minimises the propagation of cracks formed

Page | 63
during impact [149]. Both flax and kenaf fibre composites exhibit similar values for
maximum impact energy, 2.70J and 2.77J, respectively. The fracture behaviour of the
composites are explained in the photographs showing impact damage in Figures 4.5 (a-h) and
4.6 (a-h).

Impact damage characteristics

One of the reasons for limited use of natural fibre reinforced composites is their susceptibility
to impact loading and difficulty to characterise the impact damage. Most studies on impact
damage characterisation involve the use of acoustic emission (AE) technique to monitor
mechanical behaviour of composites. The main objective for using AE is the multi-parameter
study, the use of amplitude distribution for damage characterisation and AE localisation for
crack propagation analysis [150].

Santulli et al [147] investigated the impact damage characterisation of flax fibre reinforced
epoxy composites produced by hand lay-up using different fibre content in weight (31%, 55%
and 56%). The results showed that the composite made with 55% fibre content exhibited
better properties in relation to fibre impregnation, as reflected in the presence of a reversed-
pine pattern of impact damage cracks which was induced by sufficient strong fibre-matrix
interface. In another study, dealing with low velocity impact response of hemp fibre
nonwoven reinforced composites, the authors suggest that the major damage mechanisms
during impact loading include initiation and propagation of matrix cracking, fibre-matrix
debonding, fibre pull-out and fibre breakage [151].

After impact, the damaged area was visually observed and discussed. Figures 4.5 (a-h) and
4.6 (a-h) presents the fractured patterns of the impacted specimens. It is evident that the flax
reinforced bio-epoxy composites exhibit brittle failure behaviour as compared to kenaf
reinforced bio-epoxy composites as shown in Figure 4.5 (a-h). For all the flax reinforced
specimens, crack fractures are evident on the front and back face except for 15% fibre
content. The fracture mechanism at 15% fibre content shows matrix cracking, delamination
and fibre breakage in the impacted area through the thickness of the front face to the back
face in the form of a pyramid shape. These results are in agreement with the findings by other

Page | 64
researchers, that in case of nonwoven laminates, random or discontinuous fibres cannot
efficiently transfer the impact energy in the in-plane direction [152].

Page | 65
(a) Front face - flax 10% (b) Back face - flax 10%

(c) Front face - flax 15% (d) Back face - flax 15%

(e) Front face - flax 25% (f) Back face - flax 25%

(g) Front face – flax 30% (h) Back face – flax 30%

Figure 4.5 (a-h) Photographs of flax fibre reinforced composite specimens after impact
testing

Page | 66
The images of kenaf fibre reinforced bio-epoxy composite specimens are shown in Figure 4.6
(a-h). Apart from composite containing 25% fibre content, all the other specimens exhibit
extensive matrix cracking and fibre fracture in the middle of the front face. At 10% fibre
content, an indentation cone and matrix crack damage on the front and back face of the
impacted specimen is observed. As the fibre content increases to 15% the crack moves away
from the front and back face of the impacted area, however there is no visible indentation
cone. When the fibre content increases to 25% the specimen exhibited front and back face
damage through matrix cracking and fibre fracture but the impactor did not penetrate through
the specimen, indicating that the stiffness of these composites is greater at high fibre content.
A further increase in fibre content up to 30%, result in severe damage of the specimen at the
front and back face, showing a small indentation cone, matrix cracking and fibre fracture.
Again this confirms that as the fibre content increases up to its threshold value (25%), the
energy absorbed by the composite increases and minimises the fracture mechanisms.
Furthermore, when the fibre content reaches beyond its threshold value (25%), the
insufficient fibre wetting by the matrix result in weak fibre-matrix interface. In this case the
composite cannot withstand any further impact loads [151].

Page | 67
(a) Front face – kenaf 10% (b) Back face – kenaf 10%

(c) Front face – kenaf 15% (d) Back face – kenaf 15%

(e) Front face – kenaf 25% (f) Back face – kenaf 25%

(g) Front face – kenaf 30% (h) Back face – kenaf 30%
Figure 4.6 (a-h) Photographs of kenaf fibre reinforced composite specimens after
impact testing

Page | 68
4.1.1.3 Comparison of mechanical properties of flax and kenaf fibre reinforced
composites

Table 4.2 presents the tensile properties of flax and kenaf fibre reinforced bio-epoxy
composites at 25% fibre content. It can be observed that flax fibre reinforced bio-epoxy
composites outperform kenaf fibre reinforced composites in terms of tensile strength.
However, flax fibre reinforced bio-epoxy composites exhibit lower tensile modulus in
comparison to that of kenaf fibre reinforced bio-epoxy composites.
The difference in the mechanical properties of flax and kenaf composites may be attributed to
fibre type which affects the composite properties. Plant fibres are subjected to various
environmental climates and production processes (retting process) and as such their
properties affect the overall mechanical properties of the composites. According to Habibi et
al. [153], the mechanical properties of composites are strongly affected by the cellulose
content in fibres. The cellulose content of flax fibre is around 64-71% [48] and that of kenaf
fibre ranges from 50.6-60.8% [154]. As flax fibres have higher cellulose content than kenaf
fibres, it is predicted that the tensile strength of flax fibres will be higher. This is because
cellulose is the main structural component that provides strength and stability to the plant
fibre cell walls. Bismarck et al. [155] studied the tensile properties of natural fibres and
concluded that the strength increases with the increase of cellulose content in the fibres.
The flexural strength and flexural modulus of flax reinforced bio-epoxy composites also
outperforms kenaf fibre composite as it can be seen in Table 4.2. It can be concluded that the
higher tensile properties of flax fibres compared to that of kenaf fibres resulted in improved
mechanical properties of the flax fibre composites.

Table 4.2 Tensile and Flexural properties of flax and kenaf fibres reinforced bio-epoxy
composites at 25% fibre content.

Samples @ Tensile Tensile Flexural Flexural


25% fibre strength (MPa) modulus (MPa) strength (MPa) modulus (MPa)
content
Flax 42±2.1 2373±118.7 68±3.4 3581±179.1
Kenaf 33±1.7 2504±125.2 59±3.0 2904±145.2

Page | 69
4.2 Water absorption studies

Composite materials used as roof structures are exposed to varying environmental conditions
(moisture, temperature and humidity) during use and hence it is crucial to analyse the
durability of natural fibre composites after long-term exposure to moisture. Natural fibres are
susceptible to moisture [156] and cellulose as the major fibre constituent contains numerous
hydroxyl groups that are strongly hydrophilic. Therefore natural fibre reinforced composites
can absorb water which greatly causes a reduction in their mechanical properties [157]. Also
as in the present study the matrix being bio-epoxy will also influence the water uptake in the
composites. The results obtained in the study were divided into two parts, firstly the nature of
the diffusion of water into the flax and kenaf reinforced bio-epoxy composites and secondly
evaluation of the effects of water absorption on mechanical properties of composites. This
study also enables researchers to develop strategies for controlling and minimising moisture
absorption in natural fibre composites

Generally, the water content in the natural fibre composites is about 10-20% under standard
conditions. The rate of water absorption in composites is affected by factors such as fibre-
matrix interface, temperature, reaction between water and the matrix and other variables [55].
The rate of water absorption and the total moisture absorbed at saturation depend on the
chemical structure of the resin and the crosslinking agent together with temperature and
relative humidity. The moisture absorbed can cause plasticization to the matrix and also
change the state of stress to leading to formation of cracks as a result of swelling [99].

4.2.1 Water absorption behaviour

The percentage of absorbed water for all flax and kenaf reinforced composites were
determined by calculating the difference between the weight of water immersed and dry
samples by the equation below, and results are given in Table 4.3.

𝑚𝑡−𝑚𝑖
𝑀𝑡 (%) = × 100 (1)
𝑚𝑖

Where mt is the weight of the sample at time t during water immersion and mi is the weight
of the dry sample at initial time.

Page | 70
In composite materials, water diffusion can follow Fickian or non-Fickian behaviour [158]. If
the behaviour of water absorption follow Fickian law, the water gained with respect to time
can be found as a function of diffusion coefficient (D) and is calculated in the following
equation [94].

𝑘ℎ 2
𝐷 = 𝜋 (4𝑀 ) (2)
𝑚

Where, k is the slope of the linear portion of the sorption curves, h is the sample thickness in
mm and Mm is the maximum weight gain of the composites. The variation of the diffusion
coefficients with fibre content is shown in Table 4.3. The permeability of water molecules
through the composite samples depends on the sorption of water by fibres. Hence, the
sorption coefficient which is closely related to equilibrium sorption of the penetrant is
calculated using the following equation [14].

𝑀∞
𝑆= (3)
𝑀𝑝

Where M∞ is the weight of the sample at saturation point and Mp is the weight of the sample.
Permeability coefficient (P) implies to the net effect of sorption and diffusion and it is
calculated by the given formula [14]. The values of S and P are given in Table 4.3.

P=DxS (4)

Bio-epoxy composite samples reinforced with flax and kenaf fibres were immersed in
distilled water at 23 °C for a period of 168hrs. The results obtained were compared against
the control samples. Figures 4.7 and 4.8 present the percentage of water gained as a function
of square root of time for flax and kenaf fibre reinforced bio-epoxy composites at various
fibre content. It is observed from Figure 4.7 that the water uptake process is linear in the
beginning and approaches saturation after prolonged time, following Fickian diffusion
behaviour. The maximum water uptake is shown for composite with 30% fibre content as it
contains high flax fibre content. It can be concluded that as the fibre content increases the
water uptake also increases for the composite samples.

Page | 71
20
10%
18 15%
25%
16
30%

water gained (%) 14

12

10

0
0 5 10 15 20 25 30 35 40 45 50 55
1/2
time

Figure 4.7 Water absorption behaviour of flax reinforced bio-epoxy composites with
time for different fibre content

Page | 72
Table 4.3 Water absorption, diffusion, sorption and permeation coefficients of flax and kenaf fibre reinforced bio-epoxy composites
after immersion in water.

Samples and their Water absorption at Diffusion Coefficient Sorption coefficient Permeability coefficient
-3 2
fibre content (%) equilibrium Mt (%) D x 10 (cm /min) S (g/g) P x 10-3(cm2/min)
Flax-10 8.6 1.70 1.06 1.80
Flax-15 9.0 2.76 1.07 2.95
Flax-25 11.7 3.80 1.16 4.41
Flax-30 17.2 6.04 1.11 6.70
Kenaf-10 5.6 1.0 1.16 1.16
Kenaf-15 7.4 1.94 1.11 2.15
Kenaf-25 9.6 2.09 1.08 2.26
Kenaf-30 10.3 3.44 1.11 3.82

It is observed from Table 4.3 that the diffusion and permeability coefficients increase with the increase in fibre content which characterises the
ability of water molecules to move within the polymer segments [14].

Page | 73
The effect of water absorption as a function of fibre content for kenaf fibre reinforced
composites were analysed and the results presented in Figure 4.8. The water uptake at all
fibre content increases with increasing time until saturation point. It can be observed that the
composites containing 25% and 30% kenaf fibres exhibited the same water uptake up to 4
days immersion of samples in water. Thereafter, the composite containing 25% kenaf fibres
exhibited the highest water uptake of all the fibre content. The deviation from Fickian
diffusion behaviour curve is attributed to the development of micro-cracks within the
composite.

12
10%
15%
10 25%
30%

8
Water gained (%)

0
0 5 10 15 20 25 30 35 40 45 50 55
1/2
time (hr)

Figure 4.8 Water absorption behaviour of kenaf reinforced bio-epoxy composites with
time for different fibre content

Page | 74
Figure 4.9 Effect of water absorption in flax/ bio-epoxy composite and its effect on fibre-
matrix interface [159]

4.2.2 Effect of water aging on the mechanical properties of flax and kenaf reinforced
bio-epoxy

Water aged samples were subjected to tensile testing and the results compared against the un-
aged samples.

4.2.2.1 Tensile properties

Flax fibre reinforced epoxy composites

Figure 4.10 presents the tensile strength results for flax fibre reinforced bio-epoxy composites
for control sample (before) and after water aging. There is clear indication that the tensile
strength decreases significantly for samples immersed in water for 168 hours. Of all the
samples, the composite containing 25% fibre content exhibited maximum decrease in tensile
strength of 48%. When the composite is exposed to moisture, flax fibres absorb water and
swell leading to creation of micro-cracks in the brittle bio-epoxy resin. Flax fibres have high
cellulose content, approximately 71%. The cellulose molecules are rich in polar groups which
attract water molecules through hydrogen bonding, which further lead to moisture build up in

Page | 75
the cell wall and result in fibre swelling. The swelling of flax fibres exerts internal pressure
on the bio-epoxy matrix leading to the development of micro-cracks in the matrix [160].
Water penetration is then channelled through the micro-cracks in the fibre-matrix interface,
leading to debonding and hence, reducing the mechanical properties of the composite. This
phenomenon can be seen in Figure 4.9 [159]. Haameem et al. reported similar results for
short fibre reinforced polyester composites. They reported that at dry conditions the tensile
strength of the composites was 18.2 MPa and after water immersion for a period of 23 days,
the tensile strength decreased to 8.5 MPa. The authors concluded that the immersion of
composites in water weakened the interfacial bonds leading to formation of voids between
the fibre and the matrix which created debonding and subsequently led to the deterioration of
mechanical properties [161].

50
Control
45 After 168 hrs water aging

40

35
Tensile strength (MPa)

30

25

20

15

10

0
0 10 15 25 30
Fibre content (%)

Figure 4.10 Effect of water aging on the tensile strength of flax fibre reinforced
composites at different fibre content

Figure 4.11 presents the results of the tensile modulus as a function of the fibre content for
control (before) and water aged flax fibre composites. The tensile modulus for all fibre
content decreased after immersion in water for a period of 168 hours. The water within fibre
and the matrix significantly reduces the load bearing capability of the flax/ bio-epoxy
composites [101]. These results are supported by several studies [94, 99]. Of all the samples,

Page | 76
composite containing 30% fibre content exhibited the maximum decrease in tensile modulus
of 65%.

Control
After 168 hrs water aging
3000
Tensile modulus (MPa)

2000

1000

0
0 10 15 25 30
Fibre content (%)

Figure 4.11 Effect of water aging on the tensile modulus of flax fibre reinforced
composites at different fibre content

Kenaf fibre reinforced composites

Figure 4.12 shows the tensile strength results of the control and water aged kenaf fibre
reinforced bio-epoxy composites. It is evident that the tensile strength decreases for all fibre
contents. This phenomenon can be explained by taking into account the water absorption
characteristics of natural fibres. When the composite sample is immersed in water, the
hydrophilic natural fibres swell. As a result of swelling, micro-cracking of the brittle bio-
epoxy matrix occurs and this weakens the interfacial region of the composite. As more water
penetrates into the interface through micro-cracks, it creates swelling stresses that eventually
lead to composite deterioration. The composites containing 25% and 30% fibre content
exhibited similar maximum tensile strength reduction of 50% of all the samples. In
comparison to flax fibre composites at 25% fibre content, kenaf fibre reinforced composite
shows higher tensile strength reduction.

Page | 77
40
Control
After 168 hrs water aging
35

Tensile strength (MPa) 30

25

20

15

10

0
0 10 15 25 30
Fibre content (%)

Figure 4.12 Effect of water aging on the tensile strength of kenaf fibre reinforced
composites at different fibre content

The tensile modulus results of kenaf fibre reinforced bio-epoxy composites are presented in
Figure 4.13. After water aging, the tensile modulus decreased for all composite samples. A
similar trend of decreasing tensile modulus upon water aging has been reported by other
authors [99]. The micro cracks that can appear on the bio-epoxy matrix due to fibre swelling
leads to weak fibre-matrix interfacial bonding and in turn result in composite failure. This
effect has more influence in this study because the matrix used is a bio-based epoxy (bio-
content ~ 21), bio-epoxy matrix promotes water absorption in composites due to its increased
bio-based content leading to increased deterioration in mechanical properties.

Page | 78
3000
Control
After 168 hrs water aging

Tensile modulus (MPa)

2000

1000

0
0 10 15 25 30
Fibre content (%)

Figure 4.13 Effect of water aging on the tensile modulus of kenaf fibre reinforced
composites at different fibre content

4.3 Accelerated weathering (UV radiation)

The increased use of natural fibre reinforced composites in many applications has led to
concerns about the durability, especially prolonged exposure to UV radiation since their most
promising aspect seems to be that they can replace glass fibre composites in many
applications including for outdoor use. The effects of UV radiation on tensile properties of
flax and kenaf fibre reinforced bio-epoxy composites were studied in an accelerated
weathering chamber. Such weathering experiments give an insight on the amount of damage
caused to the samples by the harsh conditions and help to set guidelines of the life span of
such materials in their applications. However, accelerated weathering tests cannot be
compared to real life weathering conditions since there is no correlation to convert the actual
time in the real world with that in the test chamber. Generally, the exposure of natural fibre
composites to sunlight results in fading of colour, weight loss and deterioration in mechanical
properties which occurs due to degradation of the fibres and the matrix. The absorption of
ultraviolet rays by the composites lead to changes in the chemical structure of the bio-epoxy
due to mechanical chain scission, surface oxidation and breakdown of molecules to form
active radicals.

Page | 79
In this study, composite samples were subjected to accelerated UV radiation for a period of
seven days according to EN ISO 4892-2A test method and the results were compared against
those for the control samples. Figure 4.14 clearly shows that there is no visible colour change
or surface roughness on both flax and kenaf reinforced composites exposed to UV radiation.
Shokrieh and Bayat found that composite materials are more brittle after UV exposure and
this is due to the surface photo-crosslinking of the matrix material [113]. Studies have also
indicated that biopolymer samples exhibit lower molecular weight after UV radiation caused
by the damaged molecular structure of the composite during radiation [162].

Before UV After UV Before UV After UV

(a) (b) (c) (d)

Figure 4.14 Pictures of tensile specimens (a) flax composite before UV exposure, (b) flax
composite after UV radiation, (c) kenaf composite before UV exposure and (d) kenaf
composite after UV radiation

4.3.1 Tensile properties

Flax fibre reinforced bio-epoxy composites

The tensile properties of un-exposed and UV radiated flax fibre reinforced bio-epoxy
composite samples are shown in Figure 4.15. As expected, the tensile strength decreased after
UV radiation except for the sample with 15% fibre content. The reduction in tensile strength

Page | 80
after UV irradiation has been reported by Matuana et al [163] on the photo aging of wood
fibre reinforced PVC composites. They suggest that natural fibres are UV activators and
accelerate photo degradation of polymeric composites. Lignin present in natural fibres is
responsible for photo degradation, it absorbs ultraviolet radiation that results in the formation
of chromophoric functional groups such as carbonyls, carboxylic acids, quinones and hydro-
peroxy radicals that cause yellowing and photo-induced breakdown of materials after UV
exposure [164]. Hence, the tensile strength decreases. The increase in tensile strength after
UV radiation at 15% fibre content may be attributed to the formation of radicals, which
subsequently increases the degree of cross-linking between neighbouring cellulose molecules,
which results in the increased strength of the composite [165].

50
Control
45 After 168hrs UV radiation

40

35
Tensile strength (MPa)

30

25

20

15

10

0
0 10 15 25 30
Fibre content (%)

Figure 4.15 Effect of UV radiation on the tensile strength of flax fibre reinforced
composites at different fibre content

The effect of UV radiation on the tensile modulus of flax/ bio-epoxy composites with
different fibre content is shown in Figure 4.16. It is observed that at all fibre content, the
tensile modulus of the composite increases after UV radiation as compared to that of the un-
exposed samples. The increase in tensile modulus after UV radiation was explained by
Zaman et al [166] who found that mechanical properties increase with UV treatment due to
two opposing factors, photo-crosslinking and photo-degradation that occur simultaneously

Page | 81
under UV radiation. At higher fibre content, free radicals are stabilized by a combination
reaction and as a result, photo-crosslinking occurs between the cellulose molecules resulting
in improved mechanical properties of flax composites. The maximum increase in tensile
modulus of 37% is observed for composite reinforced with 30% flax fibres.

4000
Control
After 168hrs UV radiation

3000
Tensile modulus (MPa)

2000

1000

0
0 10 15 25 30
Fibre content (%)

Figure 4.16 Effect of UV radiation on the tensile modulus of flax fibre reinforced
composites at different fibre content

Kenaf fibre reinforced bio-epoxy composites

The effect of UV irradiation on kenaf fibre reinforced bio-epoxy composites was evaluated
and the results presented in Figure 4.17. After UV radiation, the tensile strength for kenaf/
bio-epoxy composites decreases with increasing fibre content. However, at maximum fibre
content of 30% the tensile strength shows an increase. In a study by Chen et al [167] on
ramie fibre reinforced PLLA biocomposites subjected to UV irradiation hydrothermal aging,
the authors concluded that the reduction in mechanical properties of polymeric composites is
caused by a combination of irradiation and hydrothermal environmental factors. The cellulose
component in natural fibres absorbs light energy which induces the initial fractures of
glycosidic bonds, oxidation and decomposition between the sugar rings. The increase in
tensile strength after UV radiation was also reported by Raj et al. This could be attributed to

Page | 82
the increase in stiffness of the composite resulting from chain-scission brought by thermal
stress during UV radiation [168].

50
Control
45 After 168hrs UV radiation

40

35
Tensile strength (MPa)

30

25

20

15

10

0
0 10 15 25 30
Fibre content (%)

Figure 4.17 Effect of UV radiation on the tensile strength of kenaf fibre reinforced
composites at different fibre content

Figure 4.18 shows the effect of UV radiation on the tensile modulus of kenaf fibre reinforced
bio-epoxy composites. As it can be seen, tensile modulus decreases up to 25% fibre content.
The sample containing 30% fibre content shows an increase of 3.2% in tensile modulus after
UV radiation. The increase in tensile modulus could be attributed to inter-crosslinking
between the neighbouring cellulose molecules occurring under UV radiation, resulting in the
composite exhibiting maximum tensile modulus after UV radiation [165, 169, 170]. In
comparison to flax fibre composites at 25% fibre content, composites reinforced with kenaf
fibre shows higher reduction in tensile strength tensile modulus after 168 hours of UV
treatment.

Page | 83
3000
Control
After 168hrs UV radiation

Tensile modulus (MPa)

2000

1000

0
0 10 15 25 30
Fibre content (%)

Figure 4.18 Effect of UV radiation on the tensile modulus of kenaf fibre reinforced
composites at different fibre content

4.4 Dynamic mechanical analysis (DMA)

The dynamic mechanical properties of a composite material depend on the fibre content,
presence of the additives like filler, compatibiliser, fibre orientation and the mode of testing.
DMA is excellent in obtaining strength and stiffness values from low to high temperatures.
As NFC are used in applications where they are exposed to varying temperatures and loads,
dynamical mechanical analysis is useful for establishing a wide range of temperature
dependent material data.

Flax fibre reinforced bio-epoxy composites

Figure 4.19 compares the variation of the storage modulus (E’) for flax fibre reinforced bio-
epoxy composites, as a function of temperature. It can be seen that E’ decreases with
increasing temperature for all fibre content. At lower temperatures the molecules are in a
glassy state where they are frozen and exhibit high stiffness. As the temperature increases, the
mobility of the molecules increases leading to lower stiffness. At 25 °C, E’ for the samples
containing 10% and 30% fibre contents are 2904 MPa and 3480 MPa, respectively, indicating

Page | 84
that addition of flax fibre content increases the stiffness of the composites by 20%. The
maximum E’ is obtained for composite containing 30% fibre content. This is attributed to the
higher cellulose content that increases the stiffness of the composite.

4.00E+011 10%
15%
3.50E+011 25%
30%
3.00E+011
Storage modulus [Pa]

2.50E+011

2.00E+011

1.50E+011

1.00E+011

5.00E+010

0.00E+000

20 40 60 80 100 120 140

Temperature [°C]

Figure 4.19 Variation of the storage modulus with temperature for flax fibre reinforced
bio-epoxy at different fibre content

Figure 4.20 shows tan𝛿 versus temperature curves of flax fibre reinforced bio-epoxy
composites. tan 𝛿 relates to the impact resistance of the material. As the damping peak occurs
in the region of the glass transition region where the material changes from a rigid to a more
plastic state, it is associated with the movement of small groups and chains of molecules
within the polymer structure all of which are initially frozen in. Therefore the higher the tan
𝛿 peak value, greater is the degree of molecular mobility.
The peak of the tan 𝛿 curve indicates glass transition temperature (Tg). Tg of bio-epoxy resin
was reported to be 96 °C [171]. It can be seen that the Tg of all samples shifted to lower
temperatures, indicating reduced interfacial adhesion due to incorporation of flax fibres. The
damping factor is observed to be maximum for 30% fibre content. However, the value tan 𝛿
is observed to increase with increasing fibre content. The inverse results for epoxy matrix
reinforced ramie fibres were reported [172]. The reason for these results could be the
different interfacial strengths in flax fibres which interact with the bio-epoxy matrix. The

Page | 85
degree of fibre adhesion affects the mobility of the bio-epoxy molecular chains and hence its
temperature transition to an amorphous state.

0.40
10%
0.35 15%
25%
0.30 30%

0.25
Tan delta

0.20

0.15

0.10

0.05

0.00
20 40 60 80 100 120 140

Temperature [°C]

Figure 4.20 Temperature dependence of tan 𝜹 for flax fibre reinforced bio-epoxy at
different fibre content

Kenaf fibre reinforced bio-epoxy composites

Figure 4.21 shows the effect of fibre content on storage modulus for kenaf fibre reinforced
bio-epoxy composites. It is observed that E’ decrease with increasing temperature for all fibre
contents. The maximum E’ is shown by the sample containing 15% fibre content. E’ of bio-
epoxy as reported by Sahoo [171] is 2400 MPa. It can be observed that E’ of bio-epoxy resin
is improved at samples containing 15% and 25% fibre contents.

Page | 86
4.00E+011
10%
3.50E+011 15%
25%
3.00E+011 30%
Storage Modulus [Pa]

2.50E+011

2.00E+011

1.50E+011

1.00E+011

5.00E+010

0.00E+000

20 40 60 80 100 120 140

Temperature [°C]

Figure 4.21 Variation of the storage modulus with temperature for kenaf fibre
reinforced bio-epoxy at different fibre content

Figure 4.22 shows the tan 𝛿 curves as a function of temperature for kenaf fibre reinforced
bio-epoxy composites. It can be seen that when the fibre content increases from 15% to 30%,
the peak intensities decrease i.e. the area under the tan 𝛿 curve become smaller since the bio-
epoxy content decreases, however the peaks are shifted towards low temperatures. It can be
seen that the value of tan 𝛿 max decreases with fiber content. Incorporation of fibers acted as
barriers to the mobility of polymer chains, leading to lower flexibility, lower degrees of
molecular motion and hence lower damping characteristics. Another reason for the decrease
is that there is less matrix by volume to dissipate the vibration energy and also due to fibre
agglomeration at high levels of fibre content. Therefore one can conclude that presence of
natural fibres decreases the damping properties of the composite. It can also be observed as
fibre content increases, Tg is shifted to lower temperatures indicating weak interfacial
adhesion.

Page | 87
0.50

10%
0.45
15%
25%
0.40
30%
0.35

0.30
Tan delta

0.25

0.20

0.15

0.10

0.05

0.00
20 40 60 80 100 120 140

Temperature [°C]

Figure 4.22 Temperature dependence of tan 𝜹 for kenaf fibre reinforced bio-epoxy at
different fibre content

Page | 88
CHAPTER 5: CONCLUSIONS

The main aim of the study was to develop natural fibre (flax and kenaf) reinforced bio-epoxy
composites for replacement of the commercially available glass fibre based roof sheets.
Composites were manufactured using kenaf and flax nonwovens with bio-epoxy matrix by
the process of vacuum assisted resin transfer moulding technique. Both static and dynamic
mechanical properties of the composites were studied. The effects of weathering and water
aging on the mechanical properties of the composites were also investigated.

The tensile results showed that the incorporation of flax fibres significantly reduced the
tensile strength for 15% fibre content and this could be attributed to a processing defect in the
sample. At 30% flax fibre content the tensile strength was found to increase by 7% compared
to that of bio-epoxy. The tensile modulus of flax fibre reinforced bio-epoxy composite
increased by 43%. In the case of kenaf fibre composites, the tensile strength and tensile
modulus increased by 13% and 27%, respectively. Flax fibre reinforced bio-epoxy
composites were found to exhibit higher tensile properties compared to composites reinforced
with kenaf fibres. This was attributed to the high cellulose content in flax fibres as compared
to kenaf fibres. Flax fibre reinforced bio-epoxy composites were found to exhibit brittle
failure mechanism. This was attributed to the fact that nonwoven laminates, random or
discontinuous fibres cannot efficiently transfer the impact energy in the in-plane direction.
Based on the tensile results, it can be concluded that bio-epoxy composites reinforced with
flax and kenaf fibres can be a feasible alternative to PVC (tensile strength~30MPa) roof
sheet.

Water aging experiments revealed that bio-epoxy composites reinforced with kenaf fibres
absorbed less water for all fibre content in comparison to the composites reinforced with flax
fibres, thus indicating that flax fibres are more hydrophilic in nature. Both flax and kenaf
fibre based composites showed decreases in tensile strength and modulus after water aging.
However, kenaf fibre reinforced bio-epoxy composites showed the maximum decrease in
tensile strength at 25% fibre content. After UV radiation, both flax and kenaf fibre based
composites were found to be brittle and broke easily but no colour fading was observed after
the treatment. The tensile strength of both flax and kenaf fibre reinforced bio-epoxy
composites were found to decrease after UV treatment. In comparison to kenaf, flax fibre

Page | 89
reinforced bio-epoxy composites showed an increase in tensile modulus which was attributed
to the opposing factors of photo-crosslinking and photo-degradation that occur
simultaneously under UV radiation.

The dynamic mechanical analysis of composites reinforced with flax and kenaf fibres showed
that the incorporation of fibres resulted in an increase in storage modulus of the composites
and the maximum storage modulus was exhibited by the flax fibre reinforced bio-epoxy
composite at 30% fibre content. This was attributed to the high cellulose content in flax fibres
which increases the stiffness of the composite. The glass transition temperatures of flax and
kenaf composites shifted to lower temperatures with the incorporation of fibres. The damping
factor for flax fibre reinforced bio-epoxy composite was observed to be maximum at 30%
fibre content whereas kenaf fibre reinforced bio-epoxy composite showed a maximum
damping factor at 10% fibre content.

Recommendations for future work:

 This study shows that bio-epoxy composites reinforced with flax and kenaf fibres can
be feasible alternative especially when compared to synthetic polymer based roof tiles
(PVC). Improved properties will be obtained by developing hybrid composites using
kenaf skins and flax core.
 Flammability studies must be undertaken to understand the flammability of
composites.
 Future work should also include developing sandwich panel roof sheets containing a
foam core for thermal insulation purposes.

Page | 90
REFERENCES

1. Netravali A N, Chabba S, Composites get greener. Materials today, 2003. 6(4): p. 22-
29.
2. Al-Oqla F M, Sapuan S M, Natural fiber reinforced polymer composites in industrial
applications: feasibility of date palm fibers for sustainable automotive industry.
Journal of Cleaner Production, 2014. 66: p. 347-354.
3. Singha A S, Thakur V K, Grewia optiva fiber reinforced novel, low cost polymer
composites. Journal of Chemistry, 2009. 6(1): p. 71-76.
4. La Mantia F, Morreale M, Green composites: A brief review. Composites Part A:
Applied Science and Manufacturing, 2011. 42(6): p. 579-588.
5. Mitra B C, Environment friendly composite materials: biocomposites and green
composites. Defence Science Journal, 2014. 64(3): p. 244-261.
6. Fowler P A, Hughes J M, Elias R M, Biocomposites: technology, environmental
credentials and market forces. Journal of the Science of Food and Agriculture, 2006.
86(12): p. 1781-1789.
7. Khoathane M C, Natural fibre reinforced polyolefins composites for structural
applications. 2012, Tshwane University of Technology, PhD Thesis.
8. Yatim J M, Khalid N H B A, Mahjoub R, Biocomposites for the Construction
materials and structures. Faculty of Civil Engineering, Universiti Teknologi Malaysia
(Research University), 2010: p. 1-29.
9. Campilho R, Dipa R, State-of-the-Art Applications of Natural Fiber Composites in the
Industry, in Natural Fiber Composites. 2015, CRC Press. p. 319-340.
10. Ramakrishna G, Sundararajan T, Kothandaraman S, Strength of corrugations of a
roofing sheets reinforced with sisal fibres. Journal of Engineering and Applied
Sciences, 2011. 6(12): p. 24-32.
11. Ismail M A. Compressive and Tensile Strength of Natural Fibre-reinforced Cement
base Composites. 2006. Date accessed: 2017. 12. 12 ; Available from:
http://www.iasj.net/iasj?func=fulltext&aId=44954.
12. Dweib M A, Hu B, Shenton H W, Wool R P, Bio-based composite roof structure:
Manufacturing and processing issues. Composite Structures, 2006. 74(4): p. 379-388.
13. Omoniyi T E, Akinyemi B A, Durability based suitability of bagasse-cement
composite for roofing sheets. Journal of Civil Engineering and Construction
Technology, 2012. 3(11): p. 280-290.
14. Joseph P V, Rabello M S, Mattoso L H C, Joseph K, Thomas S, Environmental effects
on the degradation behaviour of sisal fibre reinforced polypropylene composites.
Composites Science and Technology, 2002. 62(10): p. 1357-1372.
15. Ku H, Wang H, Pattarachaiyakoop N, Trada M, A review on the tensile properties of
natural fiber reinforced polymer composites. Composites Part B: Engineering, 2011.
42(4): p. 856-873.
16. Jayaraman K, Manufacturing sisal–polypropylene composites with minimum fibre
degradation. Composites Science and technology, 2003. 63(3): p. 367-374.
17. Molaba T P, Chapple S, John M J, Aging studies on flame retardant treated
lignocellulosic fibers. Journal of Applied Polymer Science, 2016. 133(44): p. 44175 (1-
10).
18. Salmeia K A, Jovic M, Ragaisiene A, Rukuiziene Z, Milasius R, Mikucioniene D,
Gaan S, Flammability of Cellulose-Based Fibers and the Effect of Structure of
Phosphorus Compounds on Their Flame Retardancy. Polymers, 2016. 8(8): p. 293.

Page | 91
19. Chapple S, Anandjiwala R, Flammability of natural fiber-reinforced composites and
strategies for fire retardancy: a review. Journal of Thermoplastic Composite
Materials, 2010. 23(6): p. 871-893.
20. Thakur V K, Thakur M K, Gupta R K, Review: raw natural fiber–based polymer
composites. International Journal of Polymer Analysis and Characterization, 2014.
19(3): p. 256-271.
21. Kumar D, Characterization of tensile properties of treated bamboo natural fibre
polymer composite. International Journal of Mechanical Engineering and Technology,
2014. 05(04): p. 110-115.
22. Campbell F C, Structural composite materials. 2010: ASM International Technical
Book Committee.
23. Saxena M, Pappu A, Sharma A, Haque R, Wankhede S, Composite materials from
natural resources: Recent trends and future potentials, in Advances in Composite
Materials-Analysis of Natural and Man-Made Materials, D P Tesinova, Editor. 2011,
InTech.
24. Kopeliovich D. Polymer Matrix Composites Introduction. 2012 [cited 2017. 19.06];
Available from:
http://www.substech.com/dokuwiki/doku.php?id=polymer_matrix_composites_introd
uction.
25. Bogoeva‐Gaceva G, Avella M, Malinconico M, Buzarovska A, Grozdanov A, Gentile
G, Errico M, Natural fiber eco‐composites. Polymer composites, 2007. 28(1): p. 98-
107.
26. Mohammed L, Ansari M N, Pua G, Jawaid M, Islam M S, A review on natural fiber
reinforced polymer composite and its applications. International Journal of Polymer
Science, 2015. 2015: p. 1-15.
27. Congress U S, Advanced materials by design. Vol. OTA-E351. 1988, Washington DC:
U.S Government printing office: US Congress, Office of Technology assessment.
28. Midland Plastics. ; Available from:
http://www.midlandplastics.com/srtd_high_performance.htm.
29. High-performance plastics. Date accessed: 2017. 07. 12. ; Available from:
https://en.wikipedia.org/wiki/High-performance_plastics.
30. Plastics D, Product Datasheet, PEEK KT820NT, D Plastics, Editor. 2015. Date
accessed: 2017. 11. 27.
31. Polymers S, LLDPE-Product data sheet, S P-P Business, Editor. 2002. Date accessed:
2017. 11. 27.
32. Solutions I P, Polyvinyl Chloride Properties, I P Solutions, Editor. Date accessed:
2017. 11. 27.
33. Kumar R, Anandjiwala R, Process for preparing polyfurfuryl alcohol products. 2012,
Google Patents.
34. Megiatto J D, Ramires E C, Frollini E, Phenolic matrices and sisal fibers modified
with hydroxy terminated polybutadiene rubber: Impact strength, water absorption,
and morphological aspects of thermosets and composites. Industrial crops and
products, 2010. 31(1): p. 178-184.
35. Joseph S, Sreekala M, Oommen Z, Koshy P, Thomas S, A comparison of the
mechanical properties of phenol formaldehyde composites reinforced with banana
fibres and glass fibres. Composites Science and Technology, 2002. 62(14): p. 1857-
1868.
36. Anne B, Environmental-friendly biodegradable polymers and composites, in
Integrated Waste Management-Volume I. 2011, InTech.

Page | 92
37. O'Brien F J, Biomaterials & scaffolds for tissue engineering. Materialstoday, 2011.
13(3): p. 88-95.
38. Mohanty A, Misra M, Hinrichsen G, Biofibers, biodegradable polymers and
biocomposites: An overview. Macromolecular materials and Engineering, 2000.
276(1): p. 1-24.
39. Coutts R S, A review of Australian research into natural fibre cement composites.
Cement and Concrete Composites, 2005. 27(5): p. 518-526.
40. Saheb D N, Jog J P, Natural fiber polymer composites: a review. Advances in polymer
technology, 1999. 18(4): p. 351-363.
41. Taj S, Munawar M A, Khan S, Natural fiber-reinforced polymer composites.
Proceedings-Pakistan Academy of Sciences, 2007. 44(2): p. 129.
42. Mohanty A K, Misra M, Drzal L T, Natural fibers, biopolymers, and biocomposites.
2005: CRC press.
43. Charlet K, Jernot J P, Evea S, Gominaa M, Bréardb J, Multi-scale morphological
characterisation of flax: From the stem to the fibrils. Carbohydrate Polymers, 2010.
82(01): p. 54-61.
44. Snegireva A, Chernova T, Ageeva M, Lev-Yadun S, Gorshkova T, Intrusive growth of
primary and secondary phloem fibres in hemp stem determines fibre-bundle formation
and structure. Journal for Plant Sciences, 2015. 7: p. 1-14.
45. Shubhra Q T, Alam A, Quaiyyum M, Mechanical properties of polypropylene
composites: A review. Journal of thermoplastic composite materials, 2013. 26(3): p.
362-391.
46. Zhu J, Zhu H, Njuguna J, Abhyankar H, Recent development of flax fibres and their
reinforced composites based on different polymeric matrices. Materials, 2013. 6(11): p.
5171-5198.
47. Baley C, Analysis of the flax fibres tensile behaviour and analysis of the tensile
stiffness increase. Composites - Part A: Applied Science and Manufacturing, 2002.
33(7): p. 939-948.
48. Shah D U, Characterisation and optimisation of the mechanical performance of plant
fibre composites for structural applications. 2013, University of Nottingham: United
Kingdom. p. 274.
49. Zeng X, Mooney S J, Sturrock C J, Assessing the effect of fibre extraction processes
on the strength of flax fibre reinforcement. Composites Part A: Applied Science and
Manufacturing, 2015. 70: p. 1-7.
50. Rogers R N, Ultra-thin and highly superficial primary cell walls of flax and linen
fibres. 2013.
51. Hann M, Innovation in linen manufacture. Textile Progress, 2005. 37(3): p. 1-42.
52. Bhat G S, Nonwovens as three-dimensional textiles for composites. Material and
Manufacturing Processes, 1995. 10(4): p. 667-688.
53. Akin D E, Dodd R B, Foulk J A, Pilot plant for processing flax fiber. Industrial Crops
and Products, 2005. 21(3): p. 369-378.
54. Akin D E, Linen most useful: perspectives on structure, chemistry, and enzymes for
retting flax. ISRN biotechnology, 2012. 2013: p. 1-23.
55. Biagiotti J, Puglia D, Kenny J M, A review on natural fibre-based composites-part I:
structure, processing and properties of vegetable fibres. Journal of Natural Fibers,
2004. 1(2): p. 37-68.
56. Ruan P, Raghavan V, Gariepy Y, Du J, Characterization of flax water retting of
different durations in laboratory condition and evaluation of its fiber properties.
BioResources, 2015. 10(2): p. 3553-3563.

Page | 93
57. Brouwer W. Natural fibre composites in structural components: Alternative
applications for sisal. in Seminar, Commond Fund for Commodities-Alternative
Applications for Sisal and Henecuen. 2000.
58. Zhu J, Zhu H, Abhyankar H, Njuguna J, Effect of fibre treatment on water absorption
and tensile properties of flax/tannin composites. Proceedings of the ICMR, 2013.
59. Célino A, Fréour S, Jacquemin F, Casari P, The hygroscopic behavior of plant fibers:
A review. Frontiers in chemistry, 2013. 1(43): p. 1-12.
60. Sgriccia N, Hawley M, Misra M, Characterization of natural fiber surfaces and
natural fiber composites. Composites Part A: Applied Science and Manufacturing,
2008. 39(10): p. 1632-1637.
61. Mokhothu T H, John M J, Review on hygroscopic aging of cellulose fibres and their
biocomposites. Carbohydrate polymers, 2015. 131: p. 337-354.
62. Khoathane M C, Vorster O C, Sadiku E R, Hemp Fiber-Reinforced 1-
Pentene/Polypropylene Copolymer: The Effect of Fiber Loading on the Mechanical
and Thermal Characteristics of the Composites. Journal of Reinforced Plastics and
Composites, 2008. 27(14): p. 1533-1544.
63. Phiri G, Khoathane M C, Sadiku E R, Effect of fibre loading on mechanical and
thermal properties of sisal and kenaf fibre-reinforced injection moulded composites.
Journal of Reinforced Plastics and Composites, 2014. 33(3): p. 283-293.
64. Anuar H, Zuraida A, Improvement in mechanical properties of reinforced
thermoplastic elastomer composite with kenaf bast fibre. Composites Part B:
Engineering, 2011. 42(3): p. 462-465.
65. Asumani O, Reid R, Paskaramoorthy R, The effects of alkali–silane treatment on the
tensile and flexural properties of short fibre non-woven kenaf reinforced
polypropylene composites. Composites Part A: Applied Science and Manufacturing,
2012. 43(9): p. 1431-1440.
66. Moscoso F J, Martínez L, Canche G, Rodrigue D, González‐Núñez R, Morphology
and properties of polystyrene/agave fiber composites and foams. Journal of Applied
Polymer Science, 2013. 127(1): p. 599-606.
67. John M, Sikampula N, Boguslavsky L, Agave nonwovens in polypropylene composites
– Mechanical and thermal studies. Journal of Composite Materials, 2015. 49(6): p.
669-676.
68. Sanadi A, Prasad S, Rohatgi P, Sunhemp fibre-reinforced polyester. Journal of
Materials Science, 1986. 21(12): p. 4299-4304.
69. Ray D, Rout J, Thermoset biocomposites. Natural fibers, biopolymers, and
biocomposites. Ed. CRC Press, Boca Raton, 2005: p. 291-345.
70. Satyanarayana K, Pai B, Sukumaran K, Pillai S, Fabrication and properties of
lignocellulosic fiber-incorporated polyester composites. Hand Book of Ceramic and
Composite, 1990. 1: p. 339-386.
71. Pickering K L, Efendy M A, Le T M, A review of recent developments in natural fibre
composites and their mechanical performance. Composites Part A: Applied Science
and Manufacturing, 2016. 83: p. 98-112.
72. Henton D E, Gruber P, Lunt J, Randall J, Polylactic acid technology. Natural fibers,
biopolymers, and biocomposites, 2005. 16: p. 527-577.
73. Yang H-S, Wolcott M P, Kim H-S, Kim S, Kim H-J, Properties of lignocellulosic
material filled polypropylene bio-composites made with different manufacturing
processes. Polymer Testing, 2006. 25(5): p. 668-676.

Page | 94
74. Francucci G, Rodríguez E S, Vázquez A, Experimental study of the compaction
response of jute fabrics in liquid composite molding processes. Journal of Composite
Materials, 2012. 46(2): p. 155-167.
75. Pickering K, Properties and performance of natural-fibre composites. 2008: Elsevier.
76. Brouwer W, Natural fibre composites: where can flax compete with glass? Sampe
Journal, 2000. 36(6): p. 18-23.
77. O'donnell A, Dweib M, Wool R, Natural fiber composites with plant oil-based resin.
Composites science and technology, 2004. 64(9): p. 1135-1145.
78. Bos H L, Müssig J, van den Oever M J, Mechanical properties of short-flax-fibre
reinforced compounds. Composites Part A: Applied Science and Manufacturing,
2006. 37(10): p. 1591-1604.
79. Mutasher S A, Poh A, Than A M, Law J. The effect of alkali treatment mechanical
properties of kenaf fiber epoxy composite. in Key Engineering Materials. 2011. : Trans
Tech Publications.
80. Roe P, Ansell M P, Jute-reinforced polyester composites. Journal of Materials
Science, 1985. 20(11): p. 4015-4020.
81. Zhu W, Tobias B, Coutts R, Banana fibre strands reinforced polyester composites.
Journal of materials science letters, 1995. 14(7): p. 508-510.
82. Wazzan A, Effect of fiber orientation on the mechanical properties and fracture
characteristics of date palm fiber reinforced composites. International Journal of
Polymeric Materials, 2005. 54(3): p. 213-225.
83. Sathish P, Kesavan R, Ramnath B V, Vishal C, Effect of fiber orientation and stacking
sequence on mechanical and thermal characteristics of banana-kenaf hybrid epoxy
composite. Silicon, 2017. 9(4): p. 577-585.
84. Alhashmy H, Fabrication of Aluminium Matrix Composites (AMCs) by Squeeze
Casting Technique Using Carbon Fiber as Reinforcement. 2012, University of
Ottawa: Canada.
85. Mwaikambo L Y, Ansell M P, Chemical modification of hemp, sisal, jute, and kapok
fibers by alkalization. Journal of applied polymer science, 2002. 84(12): p. 2222-2234.
86. Ouajai S, Shanks R, Composition, structure and thermal degradation of hemp
cellulose after chemical treatments. Polymer degradation and stability, 2005. 89(2): p.
327-335.
87. Meon M S, Othman M F, Husain H, Remeli M F, Syawal M S M, Improving tensile
properties of kenaf fibers treated with sodium hydroxide. Procedia Engineering, 2012.
41: p. 1587-1592.
88. Mohanty A, Misra M, Drzal L T, Surface modifications of natural fibers and
performance of the resulting biocomposites: an overview. Composite Interfaces, 2001.
8(5): p. 313-343.
89. George J, Sreekala M, Thomas S, A review on interface modification and
characterization of natural fiber reinforced plastic composites. Polymer Engineering
& Science, 2001. 41(9): p. 1471-1485.
90. Kalia S, Kaith B, Kaur I, Pretreatments of natural fibers and their application as
reinforcing material in polymer composites—a review. Polymer Engineering &
Science, 2009. 49(7): p. 1253-1272.
91. Espana J, Samper M, Fages E, Sánchez‐Nácher L, Balart R, Investigation of the effect
of different silane coupling agents on mechanical performance of basalt fiber
composite laminates with biobased epoxy matrices. Polymer Composites, 2013. 34(3):
p. 376-381.

Page | 95
92. Malkapuram R, Kumar V, Negi Y S, Recent development in natural fiber reinforced
polypropylene composites. Journal of Reinforced Plastics and Composites, 2009.
28(10): p. 1169-1189.
93. Espert A, Vilaplana F, Karlsson S, Comparison of water absorption in natural
cellulosic fibres from wood and one-year crops in polypropylene composites and its
influence on their mechanical properties. Composites Part A: Applied science and
manufacturing, 2004. 35(11): p. 1267-1276.
94. Muñoz E, García-Manrique J, Water absorption behaviour and its effect on the
mechanical properties of flax fibre reinforced bioepoxy composites. International
Journal of Polymer Science, 2015. 2015: p. 1-10.
95. Masoodi R, Pillai K M, A study on moisture absorption and swelling in bio-based
jute-epoxy composites. Journal of Reinforced Plastics and Composites, 2012. 31(5): p.
285-294.
96. Ashori A, Sheshmani S, Hybrid composites made from recycled materials: moisture
absorption and thickness swelling behavior. Bioresource technology, 2010. 101(12): p.
4717-4720.
97. Ghasemi I, Kord B, Long-term water absorption behaviour of polypropylene/wood
flour/organoclay hybrid nanocomposite. Iranian Polymer Journal, 2009. 18(9): p. 683-
691.
98. Kabir M, Wang H, Lau K, Cardona F, Chemical treatments on plant-based natural
fibre reinforced polymer composites: An overview. Composites Part B: Engineering,
2012. 43(7): p. 2883-2892.
99. Dhakal H, Zhang Z, Richardson M, Effect of water absorption on the mechanical
properties of hemp fibre reinforced unsaturated polyester composites. Composites
Science and Technology, 2007. 67(7): p. 1674-1683.
100. Assarar M, Scida D, El Mahi A, Poilâne C, Ayad R, Influence of water ageing on
mechanical properties and damage events of two reinforced composite materials:
Flax–fibres and glass–fibres. Materials & Design, 2011. 32(2): p. 788-795.
101. Azwa Z, Yousif B, Manalo A, Karunasena W, A review on the degradability of
polymeric composites based on natural fibres. Materials & Design, 2013. 47: p. 424-
442.
102. Chiang T, Osman M, Hamdan S, Water absorption and thickness swelling behavior of
sago particles urea formaldehyde particle board. International Journal of Science and
Research, 2014. 3(12): p. 1375-1379.
103. Chen H, Miao M, Ding X, Influence of moisture absorption on the interfacial strength
of bamboo/vinyl ester composites. Composites Part A: Applied Science and
Manufacturing, 2009. 40(12): p. 2013-2019.
104. Karaduman Y, Onal L, Water absorption behavior of carpet waste jute-reinforced
polymer composites. Journal of Composite Materials, 2011. 45(15): p. 1559-1571.
105. Venkateshwaran N, ElayaPerumal A, Mechanical and water absorption properties of
woven jute/banana hybrid composites. Fibers and Polymers, 2012. 13(7): p. 907-914.
106. Stark N M, Matuana L M, Ultraviolet weathering of photostabilized wood‐flour‐filled
high‐density polyethylene composites. Journal of Applied Polymer Science, 2003.
90(10): p. 2609-2617.
107. Müller U, Rätzsch M, Schwanninger M, Steiner M, Zöbl H, Yellowing and IR-changes
of spruce wood as result of UV-irradiation. Journal of Photochemistry and
Photobiology B: Biology, 2003. 69(2): p. 97-105.
108. Pandey K K, Study of the effect of photo-irradiation on the surface chemistry of wood.
Polymer Degradation and Stability, 2005. 90(1): p. 9-20.

Page | 96
109. Beckert W, Lauke B, Critical discussion of the single-fibre pull-out test: does it
measure adhesion? Composites science and technology, 1998. 57(12): p. 1689-1706.
110. Argyropoulos D, Heitner C, Schmidt J, Observation of quinonoid groups during the
light-induced yellowing of softwood mechanical pulp. Research on chemical
intermediates, 1995. 21(3): p. 263-274.
111. Schmidt J, Kimura F, Gray D, IR and UV spectroscopic study of borohydride reduced
mechanical pulp during monochromatic and wide band irradiation. Research on
chemical intermediates, 1995. 21(3): p. 287-301.
112. Bonini C, D'Auria M, Mauriello G, Viggiano D, Zimbardi F, Singlet oxygen
degradation of lignin in the pulp. Journal of Photochemistry and Photobiology A:
Chemistry, 1998. 118(2): p. 107-110.
113. Shokrieh M M, Bayat A, Effects of ultraviolet radiation on mechanical properties of
glass/polyester composites. Journal of Composite materials, 2007. 41(20): p. 2443-
2455.
114. Chin J W, Nguyen T, Aouadi K, Effects of environmental exposure on fiber-reinforced
plastic (FRP) materials used in construction. Journal of Composites, Technology and
Research, 1997. 19(4): p. 205-213.
115. Islam M S, Pickering K L, Foreman N J, Influence of accelerated ageing on the
physico-mechanical properties of alkali-treated industrial hemp fibre reinforced poly
(lactic acid)(PLA) composites. Polymer Degradation and Stability, 2010. 95(1): p. 59-
65.
116. Stark N M, Matuana L M, Clemons C M, Effect of processing method on surface and
weathering characteristics of wood–flour/HDPE composites. Journal of Applied
Polymer Science, 2004. 93(3): p. 1021-1030.
117. Zini E, Scandola M, Green composites: an overview. Polymer composites, 2011.
32(12): p. 1905-1915.
118. Journal I F, Tailor-Made PP Fibers for Automotive Composites, in International Fiber
Journal. 2016.
119. Medina L, Schledjewski R, Natural-fibre thermoset composite prepregs, in JEC
composites. 2008. p. 46-47.
120. Namvar F, Jawaid M, Tanir P M, Mohamad R, Azizi S, Khodavandi A, Rahman H S,
Nayeri M D, Potential use of plant fibres and their composites for biomedical
applications. BioResources, 2014. 9(3): p. 5688-5706.
121. Rudi Dungani, Myrtha Karina, Subyakto, A. Sulaeman, Dede Hermawan, Hadiyane A,
Agricultural Waste Fibers Towards Sustainability and Advanced Utilization: A
Review. Asian Journal of Plant Sciences, 2016. 15: p. 42-55.
122. Lina F, Yue Z, Jin Z, Guang Y, Bacterial cellulose for skin repair materials, in
Biomedical Engineering-frontiers and Challenges. 2011, InTech.
123. Group J, An acoustic guitar made of flax linen fabrics, in JEC composites. Retrieved
03 November 2017. 2015.
124. JuteLab, Eco friendly long board made out of jute fiber, in Longboardism.com. 2015,
Longboardism. Retrieved 03 November 2017: Bangladesh.
125. Composite as high performance building solutions. [cited 2016. 08. 30]; Available
from: “https://greenbuildingsolutions.org/blog/composites-high-performance-building-
solutions”.
126. Technologies F, Acrodur: Innovation from BASF, in Archive for the 'Natural Fibers'
Category. 2014: Indiana, North America.
127. Technologies F, Natural fiber finds its way to renewable energy, in Archive for the
'Natural Fibers' Category. 2014: Indiana, North Amarica.

Page | 97
128. Methacanon P, Weerawatsophon U, Sumransin N, Prahsarn C, Bergado D, Properties
and potential application of the selected natural fibers as limited life geotextiles.
Carbohydrate Polymers, 2010. 82(4): p. 1090-1096.
129. Awalellu K A, A Review on Properties and Applications of Polymer Matrix
Composites. International Journal of Research and Scientific Innovation, 2016.
III(IA): p. 53-55.
130. John M J, Thomas S, Biofibres and biocomposites. Carbohydrate Polymers, 2008.
71(3): p. 343-364.
131. Ticoalu a, Aravinthan T, Cardona F, A review of current development in natural fiber
composites for structural and infrastructure applications. Southern Region
Engineering Conference, 2010(November): p. 1-5.
132. Wool R P, Hurricane-Resistant houses from Soybean oil and natural fibers, in
Biobased Polymers and Composites, R P Wool, X S Sun, Editors. 2015, Elsevier
Science and Technology Books.
133. Speckenbach M. https://issuu.com/mark.speckenbach/docs/bio1403_einzelseiten.
[cited 2016. 08. 15].
134. News B. Nominees for the Innovation Award “Biocomposite of the Year 2017” call
attention to new technologies and application fields for biocomposites. 2017 [cited
2017. 10. 02]; Available from: http://news.bio-based.eu/nominees-for-the-innovation-
award-biocomposite-of-the-year-2017-call-attention-to-new-technologies-and-
application-fields-for-
biocomposites/?lipi=urn%3Ali%3Apage%3Ad_flagship3_feed%3BmPo9Su%2FjQv
Or7ZezLEqqFg%3D%3D.
135. BTL. The Hemp House : Africa’s Most Sustainable Building. [cited 2017. 04. 03];
Available from: https://btl.co.za/the-hemp-house-africas-most-sustainable-building/.
136. Biyana N Y, Jacobs N V, John M J, Studies on flax/polypropylene-reinforced
composites for automotive applications. 2015, Nelson Mandela University, MSc
Thesis.
137. Ramesh M, Kenaf (Hibiscus cannabinus L.) fibre based bio-materials: A review on
processing and properties. Progress in Materials Science, 2016. 78: p. 1-92.
138. Shibata S, Cao Y, Fukumoto I, Press forming of short natural fiber-reinforced
biodegradable resin: Effects of fiber volume and length on flexural properties.
Polymer testing, 2005. 24(8): p. 1005-1011.
139. David-West O S, Nash D H, M B W, Impact characterisation of doubly curved
composite sturcture, in 17th International Conference on Composite Materials,
ICCM17. 2009, Strathprints: United Kingdom.
140. Chartoff R P, Menczel J D, Dillman S H, Dynamic mechanical analysis (DMA).
Thermal analysis of polymers: fundamentals and applications, 2009: p. 387-495.
141. Joseph P, Mathew G, Joseph K, Groeninckx G, Thomas S, Dynamic mechanical
properties of short sisal fibre reinforced polypropylene composites. Composites Part
A: Applied Science and Manufacturing, 2003. 34(3): p. 275-290.
142. Garcia-zetina F, Martinez E, Alvarez-castillo A, Castano V M, Numerical analysis of
the experimental mechanical properties in polyester resins reinforced with natural
fibers. Journal of Reinforced Plastics and Composites, 1995. 14(6): p. 641-649.
143. Mahjoub R, Yatim J M, Mohd Sam A R, Raftari M, Characteristics of continuous
unidirectional kenaf fiber reinforced epoxy composites. Materials and Design, 2014.
64: p. 640-649.
144. Monteiro S, Terrones L, D’almeida J, Mechanical performance of coir fiber/polyester
composites. Polymer Testing, 2008. 27(5): p. 591-595.

Page | 98
145. Mishra V, S B, Physical and mechanical properties of bi-directional jute fiber epoxy
composites. procedia Engineering, 2013. 51: p. 561-566.
146. Santulli C, Caruso A, A Comparative Study on Falling Weight Impact Properties of
Jute/Epoxy and Hemp/Epoxy Laminates. Malaysian Polymer Journal, 2009. 4(1): p. 19-
29.
147. Santulli C, Janssen M, Jeronimidis G, Partial replacement of E-glass fibers with flax
fibers in composites and effect on falling weight impact performance. Journal of
materials science, 2005. 40(13): p. 3581-3585.
148. Sumaila M, Amber I, Bawa M, Effect of fiber length on the physical and mechanical
properties of ramdom oriented, nonwoven short banana (musabalbisiana) fiber/epoxy
composite. Asian Journal of Natural & Applied Sciences, 2013. 2(1): p. 39-49.
149. Sreenivasan V, Ravindran D, Manikandan V, Narayanasamy R, Influence of fibre
treatments on mechanical properties of short Sansevieria cylindrica/polyester
composites. Materials & Design, 2012. 37: p. 111-121.
150. De Rosa I M, Marra F, Pulci G, Santulli C, Sarasini F, Tirillò J, Valente M, Post-impact
mechanical characterisation of glass and basalt woven fabric laminates. Applied
Composite Materials, 2012. 19: p. 475-490.
151. Dhakal H, Zhang Z, Richardson M, Errajhi O, The low velocity impact response of
non-woven hemp fibre reinforced unsaturated polyester composites. Composite
structures, 2007. 81(4): p. 559-567.
152. Shyr T-W, Pan Y-H, Impact resistance and damage characteristics of composite
laminates. Composite structures, 2003. 62(2): p. 193-203.
153. Habibi Y, El-Zawawy W K, Ibrahim M M, Dufresne A, Processing and
characterization of reinforced polyethylene composites made with lignocellulosic
fibers from Egyptian agro-industrial residues. Composites Science and Technology,
2008. 68(7): p. 1877-1885.
154. Ishak M R, Leman Z, Sapuan S M, Edeerozey A M M, Othman I S, Mechanical
properties of kenaf bast and core fibre reinforced unsaturated polyester composites.
IOP Conference Series: Materials Science and Engineering, 2010: p. 1-6.
155. Bismarck A, Mishra S, T. L, Plant fibres as reinforcement for green composites, in
Natural Fibres, Biopolymers, and Biocomposites, Mohanty A.K., Misra M., D L T.,
Editors. 2005, CRC Press. p. 37-108.
156. Verghese N, Hayes M, Garcia K, Carrier C, Wood J, Lesko J. Effects of temperature
sequencing during hygrothermal aging of polymers and polymer matrix composites:
the reverse thermal effect. in Second International Conference on Composites in
Infrastructure. 1998.
157. Gassan J, Bledzki A K, Effect of cyclic moisture absorption desorption on the
mechanical properties of silanized jute‐epoxy composites. Polymer Composites, 1999.
20(4): p. 604-611.
158. Loos A C, Springer G S, Sanders B, Tung R, Moisture absorption of polyester-E glass
composites. Environmental effects on composite materials, 1981. 1: p. 51.
159. Singh A A, Palsule S, Effect of water absorption on interface and tensile properties of
jute fiber reinforced modified polyethylene composites. Applied Polymer Composites,
2013. 1(2): p. 113-123.
160. Bismarck A, Aranberri‐Askargorta I, Springer J, Lampke T, Wielage B, Stamboulis A,
Shenderovich I, Limbach H H, Surface characterization of flax, hemp and cellulose
fibers; surface properties and the water uptake behavior. Polymer composites, 2002.
23(5): p. 872-894.

Page | 99
161. Haameem M J A, Majid M A, Afendi M, Marzuki H, Hilmi E A, Fahmi I, Gibson A,
Effects of water absorption on Napier grass fibre/polyester composites. Composite
Structures, 2016. 144: p. 138-146.
162. Shangguan Y-Y, Wang Y-W, Wu Q, Chen G-Q, The mechanical properties and in
vitro biodegradation and biocompatibility of UV-treated poly (3-hydroxybutyrate-co-
3-hydroxyhexanoate). Biomaterials, 2006. 27(11): p. 2349-2357.
163. Matuana L M, Kamdem D P, Zhang J, Photoaging and stabilization of rigid
PVC/wood‐fiber composites. Journal of Applied Polymer Science, 2001. 80(11): p.
1943-1950.
164. Matuana L M, Jin S, Stark N M, Ultraviolet weathering of HDPE/wood-flour
composites coextruded with a clear HDPE cap layer. Polymer Degradation and
Stability, 2011. 96(1): p. 97-106.
165. Khan M A, Haque N, Al-Kafi A, Alam M, Abedin M, Jute reinforced polymer
composite by gamma radiation: effect of surface treatment with UV radiation.
Polymer-Plastics Technology and Engineering, 2006. 45(5): p. 607-613.
166. Zaman H U, Khan M A, Khan R A, Banana fiber-reinforced polypropylene
composites: A study of the physico-mechanical properties. Fibers and Polymers, 2013.
14(1): p. 121-126.
167. Chen D, Li J, Ren J, Influence of fiber surface-treatment on interfacial property of
poly (l-lactic acid)/ramie fabric biocomposites under UV-irradiation hydrothermal
aging. Materials Chemistry and Physics, 2011. 126(3): p. 524-531.
168. Raj R, Kokta B, Daneault C, A comparative study on the effect of aging on mechanical
properties of LLDPE–glass fiber, mica, and wood fiber composites. Journal of
Applied Polymer Science, 1990. 40(5‐6): p. 645-655.
169. Benedetto R M D, Gelfuso M V, Thomazini D, Influence of UV Radiation on the
Physical-chemical and Mechanical Properties of Banana Fiber. Materials Research,
2015. 18: p. 265-272.
170. Hassan M M, Islam M R, Shehrzade S, Khan M A, Influence of Mercerization Along
with Ultraviolet (UV) and Gamma Radiation on Physical and Mechanical Properties
of Jute Yarn by Grafting with 3‐(Trimethoxysilyl) Propylmethacrylate (Silane) and
Acrylamide Under UV Radiation. Polymer-Plastics Technology and Engineering,
2003. 42(4): p. 515-531.
171. Sahoo S K, Mohanty S, Nayak S K, Synthesis and characterization of bio-based epoxy
blends from renewable resource based epoxidized soybean oil as reactive diluent.
Chinese Journal of Polymer Science, 2015. 33(1): p. 137-152.
172. Margem F M, Monteiro S N, Bravo Neto J, Rodriguez R J S, Soares B G, The
dynamic-mechanical behavior of epoxy matrix composites reinforced with ramie
fibers. Matéria (Rio de Janeiro), 2010. 15(2): p. 164-171.

Page | 100

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy