Anda di halaman 1dari 23

Flammability of Natural Fiber-reinforced

Composites and Strategies for Fire


Retardancy: A Review

S. CHAPPLE1,* AND R. ANANDJIWALA1,2


1
CSIR, Materials Science and Manufacturing, Polymers and Composites,
Nonwovens and Composites Research Group
PO Box 1124, Port Elizabeth, 6000, South Africa
2
Department of Textile Science, Nelson Mandela Metropolitan University
Port Elizabeth, South Africa

ABSTRACT: Natural fiber-reinforced composites are finding new applications


in many sectors. In certain industries, such as building and transport, reduced
material flammability is a key requirement. Knowledge of the flammability of
natural fiber-reinforced composites and the methods used to improve their fire
resistance is necessary to ensure their use in these industries. The purpose of this
review is to examine important aspects of the flammability of natural fiber-
reinforced composites and to outline some of the more recent strategies used to
improve their fire performance.

KEY WORDS: natural fiber, composite, polymer, flammability, fire retardant.

INTRODUCTION

NE OF THE first natural fiber-reinforced composites (NFRC), Gordon


O Aerolite, was developed by Aero Research Limited in the mid 1930s [1].
This flax fiber/phenol formaldehyde matrix composite was used to fabricate
the spar for the Bristol Blenheim bomber and was the first major structural
composite material to be considered seriously for aircraft construction.
Development of NFRC was set to increase; however, in the 1940s, Owens-
Corning produced glass–fiber reinforcement for plastic laminates [2], and
with their increasing use in aeroplanes and boats, research into the use of
natural fibers in composite materials rapidly declined [3].

*Author to whom correspondence should be addressed. E-mail: schapple@csir.co.za

Journal of THERMOPLASTIC COMPOSITE MATERIALS, Vol. 23—November 2010 871


0892-7057/10/06 0871–23 $10.00/0 DOI: 10.1177/0892705709356338
ß The Author(s), 2010. Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


872 S. CHAPPLE AND R. ANANDJIWALA

The last two decades have seen a resurgence of interest in NFRC,


which has resulted in extensive research and newer industrial applications.
Initially, technical requirements were low and applications were found
mainly in nonstructural parts [4]. More recently, however, research efforts,
in areas such as fiber–matrix compatibility and improving impact strength,
has led to the wider use of NFRC, in areas such as construction materials,
boat hulls, and automotive parts [5,6], and again research is being conducted
to extend the use of NFRC in aerospace applications [7–9] where
flammability is a key issue.
The use of natural fibers for reinforcing composite materials is attractive
because the fibers come from a renewable resource and are biodegradable
[4,5,10]. These are both necessary qualities in a society looking for green
solutions in manufacturing, as witnessed by initiatives such as the European
Union Ecoefficient Technologies and Products Based on Natural Fiber
Composites project (ECOFINA) and Clean-Sky [11]. In addition, unlike
glass fibers, the thermal recycling of the composite may be possible. Natural
fibers can be produced at a relatively low cost, offer friendly processing (low
tool wear and little skin irritation), and have good thermal and acoustic
properties. They also have a low specific weight, which results in a higher
specific strength and stiffness compared to glass fibers. There are, however,
some disadvantages to using natural fibers including variable quality and
price fluctuations, moisture absorption, lower durability, lower impact
strength, and restricted processing temperatures, which often limit the type
of polymer available for the matrix. They also have poor fire resistance,
which is a major drawback in certain applications, for example, in the
transportation sector in general but for aerospace in particular.
The physical and mechanical disadvantages of the natural fibers have
received a great deal of attention in recent research [10,12]. Fire resistance,
however, has received less attention. In 2004, Matkó et al. [13] reported that
the thermal stability and fire retardancy had hardly been studied and, in
2007 and 2008, Hapuarachchi et al. [14] and Kozlowski et al. [15],
respectively, reported that there had been only a small number of studies
looking into the fire performance of NFRC. Many of the fire retardancy
strategies used by researchers are adapted from strategies used to improve
the fire retardancy of thermoplastic and thermoset polymers/composites,
and natural fiber-based textiles. For thermoplastic composites, this includes
the use of a variety of fire-retardant additives and intumescent systems, for
thermosets, the use of coatings and intumescent systems, and for reinforcing
natural fibers, a flame-retardant treatment of the fibers prior to incorpora-
tion in the matrix. The purpose of this review is to examine the flammability
of NFRC and consider some of the strategies used or available for
improving their fire retardancy.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 873

FLAMMABILITY OF NFRC

As with composites reinforced with glass or other synthetic fibers, NFRC


comprise a polymer matrix that is reinforced with natural fibers. A wide range
of fibers, mainly plant, and polymer matrices utilized have been reported
in Table 1 [6,16]. The resultant mechanical properties and functionalities of
the composite are different and enhanced compared to the individual
component materials. Likewise, the flammability properties of the composite
are different from the component materials and factors such as the structure
of the composite, adhesion between the fibers and the matrix, the type of
matrix polymer, and the type of fiber all play roles in determining these
properties [17,18].

Polymer Flammability

There are a variety of polymers available for use in NFRC including


thermoset, thermoplastic, and biopolymers. The flammability properties
of these polymers are different but all are flammable to a varying
degree [17].
When exposed to heat, polymers undergo thermal and thermal–oxidative
decomposition [19,20]. Initially, heat, smoke, and volatiles are produced.
The volatiles consist of a mixture of combustible volatiles, such as
monomers, hydrocarbons and carbon monoxide, and noncombustible

Table 1. Examples of thermoplastic and thermoset matrix–natural fiber


composites (adapted from Puglia et al. [6] and Blicblau et al. [16]).

Polymer matrix
Thermoplastic Thermoset
Fiber PP PE PA66 PS PVC Epoxy PET Vinylester Phenolic
Cellulose* X X X X X X
Flax X X X
Jute X X X X X X
Sisal X X X X X
Kenaf X X X
Ramie X
Hemp X X X X
Bagasse X X
Bamboo X X X
Pineapple X X X
Wood flour/fiber X X X X X
Wool X

*Includes cotton.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


874 S. CHAPPLE AND R. ANANDJIWALA

gases, including carbon dioxide, hydrogen halides, and others. The actual
volatiles produced depend on the chemical nature of the polymer.
Flammable volatiles formed come into contact with oxygen to produce
highly reactive H and OH radicals, which are involved in the
further decomposition and the sustained burning of the polymer. Some
polymers may breakdown completely during thermal decomposition,
whereas others may form carbonaceous or inorganic residues. The
decomposition products, rate, and mechanisms of thermal decomposition
depend not only on the chemical composition of the polymer but also on its
physical properties [21].
Physical properties, such as glass transition, melting, and decomposition
temperatures, influence the thermal decomposition processes, as it is at these
temperatures that the polymer undergoes phase transitions, which result in
changes in physical properties, such as thermal conductivity, viscosity,
density, and modulus [19]. In the case of many aromatic thermosetting
polymers, the decomposition temperature is lower than the melting
temperature and therefore resulting physical transformations before
decomposition are minimal [21].
The limiting oxygen index (LOI), measured according to ASTM D-2863
[22], is often used to illustrate the relative flammability of different polymers
and, within the parameters of the LOI test, a polymer with a higher LOI
would have lower flammability than one with a lower LOI [19,23]. LOI values
of some commonly used polymers are given in Table 2 [19,23–25].
From the LOI values presented in Table 2, polyvinyl chloride, with a LOI
value of 23–45, would appear to be suitable for use as a matrix polymer from
a flammability point of view. When it burns, however, polyvinyl chloride
produces a heavy smoke and corrosive hydrogen chloride gas, and this
precludes its use as a matrix material for composites for end-use in ships and
aircraft [26,27]. Generally, the LOI as a measure of polymer flammability
should be used with caution because in real fire situations factors, such as
lower oxygen accessibility and higher air velocity and temperatures,
compared to those used in the LOI test according to ASTM D-2863
method, can influence the LOI of the polymer [19]. In addition, there is little
correlation between LOI and other fire properties, such as the heat release
rate (HRR) [28], smoke density, and others.
The formation of char during burning is an important criterion in gauging
the flammability property of a matrix polymer. Polymers, such as phenolics,
which are highly cross-linked, polyvinyl chloride, in which cross-linking may
occur during the decomposition process, and casein, a protein-based
biopolymer, form char during combustion [18,21,29]. The char is formed
at the expense of possible flammable volatiles, and it also acts as a
heat barrier between the heat source and lower layers of the polymer.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 875

Table 2. LOI of different polymers [19,23–25].

Polymer LOI (vol%)


Polyethylene 17
Polypropylene 17
Polystyrene 18
Polybutylene terephthalate 18
Vinyl ester 20–23
Polyethylene terephthalate 21–33
Polyamide (Nylon 6, 6) 21–30
Polyamide (Nylon 6) 21–34
Polyvinyl chloride 23–45
Epoxy 23–27
Phenolic 25–57
Polyamide (Nylon 11) 25–32
Polycarbonate 26
Acrylonitrile butadiene styrene 29–35
Polysulfone 30
Polyimide 32
Polyphenylene sulfone 34
Polyether sulfone 34–38
Polyether ether ketone 35
Polyamide-imide 42–50
Polyether-imide 44–47
Polyphenylene sulfide 44–53
Polyvinylidene chloride 60
Polytetrafluoroethylene 90

Note: Where a range of values is given, the higher value generally refers
to grades with mineral or glass–fiber filler and/or fire retardant.

A char-forming polymer, therefore, has a lower flammability and from this


aspect is more desirable for use as a matrix polymer.
Manfredi et al. [30] illustrated the effect of different polymer matrices on
the flammability properties of jute fiber composites of similar construction.
In a composite utilizing a modified acrylic, time to ignition (TTI), peak heat
release (PHRR), and total heat release (THR) were 72 s, about 920 kW/m2
and 74.2 MJ/m2, respectively, as measured by cone calorimeter. The
corresponding figures for a composite using a polyester matrix were 51 s,
about 525 kW/m2 and 77.6 MJ/m2. The modified acrylic–jute composite had
delayed ignition and lower THR than the polyester–jute composite but had
a higher PHRR, which indicated a more rapid fire growth and hence a
greater fire risk for this composite. The average specific area extinction for
the two composites was 225 and 736 m2/kg, respectively. The lower smoke
production from the modified acrylic composite was due to the greater char
forming capability of this resin.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


876 S. CHAPPLE AND R. ANANDJIWALA

Table 3. Examples of some plant-based fibers


(adapted from Kozhowski et al. [31]).

Type Examples
Bast Flax, hemp, kenaf, jute, ramie
Leaf Pineapple, banana, sisal, agave, palm
Grasses Bamboo, bagasse, reed, wheat, rice
Fruit Luffa, coir
Wood Hardwoods, softwoods
Seed Cotton, kapok

Flammability of Natural Fibers

The type of natural fibers available for use in NFRC is vast, examples of
plant-based fibers are shown in Table 3 [31]. Natural fibers are nonthermo-
plastic and as such their decomposition (pyrolysis) temperature is less than
their glass transition and/or melting temperatures [32]. Unlike protein fibers,
plant fibers has poor fire resistance; cotton, for example, has a LOI value of
18–20, whereas wool has a LOI value of 25 [19]. Consequently, there has been
more research into the flammability of plant fibers, notably cotton.
The thermal degradation of plant fibers involves a number of processes
including the desorption of adsorbed water, cross-linking of cellulose chains
with the evolution of water to form dehydrocellulose, decomposition of the
dehydrocellulose to yield char and volatiles, formation of levoglucosan, and
decomposition of the levoglucosan to yield flammable and nonflammable
volatiles and gases, tar, and char [10,32,33].
The thermal degradation of flax, sisal, and jute fibers, using thermogravi-
metric analysis (TGA), was reported by Manfredi et al. [30]. They found that
the thermal degradation of sisal and jute was similar, with the main peak at
3408C, whereas the flax started to degrade at higher temperatures, with
the main peak at 3458C. Slight variations in the flammability of different
plant fibers can be attributed in part to the chemical composition of the
fibers. Natural fibers are composed primarily of cellulose, hemicellulose, and
lignin, with the balance being made up of pectins, water soluble compounds,
waxes, and inorganic, nonflammable substances, which are generally referred
to as ash. Examples of the chemical composition of some plant fibers are
shown in Table 4 [31].
The decomposition of cellulose, between 2608C and 3508C, results in the
formation of flammable volatiles and gases, noncombustible gases, tars, and
some char [32,34–36]. A high content of cellulose tends to increase the
flammability of the fiber. Hemicellulose decomposes between 2008C and
2608C but forms more noncombustible gases and less tar than cellulose.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 877

Table 4. Chemical composition of some plant fibers


(adapted from Kozhowski et al. [31]).

Cellulose Lignin Hemicellulose Ash


Fiber (%) (%) (%) (%)
Flax 64–71 2–5 18.6–20.6 5
Kenaf 44–57 15–19 22–23 2–5
Sisal 67.5–78 8–11 10–24 0.6–1
Bamboo 26–43 21–31 15–26 5–17
Rice 28–48 12–16 23–28 15–20
Coir 36–43 41–45 0.15–0.25 2 [88]
Deciduous 38–49 23–30 19–26 51
Coniferous 40–45 26–34 7–14 51
Cotton 85–90 7–16 1–3 0.8–2

Lignin starts decomposing from about 1608C and continues to decompose


until about 4008C. At the lower temperatures, relatively weak bonds break,
whereas at higher temperatures, cleavage of bonds in the aromatic rings of
the lignin takes place [37]. Lignin contributes more to char formation than
either cellulose or hemicellulose [34,38]. Therefore, based only on chemical
composition, coir, with a low cellulose (36–43%) and high lignin (41–45%)
content, and deciduous wood fiber, with a low cellulose (38–49%), high
lignin (23–30%), and hemicellulose (19–26%) content, should have lower
flammability than, for example, cotton, which has a very high cellulose
(85–90%) content. Manfredi et al. [30] showed the importance of lignin in
the thermal decomposition of sisal, flax, and jute fibers, having lignin
contents of 9.9%, 2.0%, and 11.8%, respectively. They concluded that the
lower lignin content in flax contributed to a higher decomposition
temperature but resulted in a lower oxidation resistance, which would be
provided by the aromatic structure of lignin [38]. High silica content,
evidenced by a high ash content, tends to increase fire resistance [39] and rice
straw, with a relatively high ash content of between 15% and 20% and also
having a relatively low cellulose (28–48%) and high hemicellulose (23–28%)
content, should also have better fire resistance than, for example, cotton.
Besides the chemical composition, the fine structure of the fiber also plays a
role in the flammability of the fiber [33,40]. For cellulosic fibers in particular,
higher levels of crystallinity results in higher levels of levoglucosan during
pyrolysis and consequently increased flammability. A higher ignition
temperature, however, also results from increased crystallinity as more
energy (energy of activation) is required to decompose the crystalline
structure. For example, the activation energy of amorphous celluloses is
about 120 kJ/mol, whereas it is about 200 kJ/mol for crystalline cotton and
ramie [40]. The degree of polymerization and fibrillar orientation also

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


878 S. CHAPPLE AND R. ANANDJIWALA

influence fiber flammability [33,40]. Increased orientation or degree of


polymerization results in decreased pyrolysis. Basch et al. [33,41] attributed
the high thermal stability of ramie, when compared to cotton, to its very
high orientation since the degree of polymerization and crystallinity of
the two fibers are similar. Orientation controls the amount of oxygen
penetrating into the fiber – the higher the orientation, the lower the
permeability of the fiber to oxygen [33]. A fiber with low crystallinity and a
high degree of polymerization and orientation would, therefore, from
a flammability aspect, be the best choice to use as reinforcement in a
composite material.
Blends of fibers, for example, plant and protein based, would also be
another possible way to reduce the flammability of the fiber reinforcement.
There are, however, other factors such as mechanical properties to consider
when selecting a natural fiber as reinforcement in a composite material.

Composite Flammability

The flammability of a composite not only depends on the matrix polymer


and the type of fiber but also on interactions between the two [18].
A number of studies [42,43] have been conducted on the burning behavior of
composites reinforced with glass, aramid, or graphite fibers. The studies
have shown that certain combinations can reduce flammability, whereas
others may increase flammability of the composite compared to the
components. An increase can, in some cases, be a result of a ‘scaffolding
effect’ where the molten decomposition products from the polymer are held
in contact with the heat source by the fiber reinforcement [18,44]. In the case
of NFRC, Helwig et al. [45] conducted a study of the flammability of
polypropylene composites containing varying amounts of flax fiber. In cone
calorimeter testing, they found that at a flax fiber content of 12.5%, the
peak heat release rate (PHRR) was about 35% lower than that of neat
polypropylene. Compared to neat polypropylene, the composites had a
shorter TTI and increased smoke production. Mass loss rates (MLRs) were
lower but the samples burnt for a longer time. According to the study, the
characteristics of the composite became more like that of lignocellulosic
materials when fiber content was increased above 20%. In a later study of
polypropylene composites having 50% pine wood particles, Borysiak et al.
[46] found that the composites had a shorter TTI, lower PHR and MLR
compared to unfilled polypropylene. Smoke generation was slightly reduced.
They also found that the melt flow index of the polypropylene influenced the
flammability properties of the composites.
Although the burning behavior of the composite is largely determined by
the properties of the matrix and the reinforcing fibers and any synergistic or

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 879

antagonistic effects between them, the flammability of the composite is also


influenced by the construction of the composite [18,47]. The thickness of the
composite, for example, can affect the surface flammability up to a certain
limiting value. Thermally thick samples, in which heat wave penetration is
less than the thickness of the sample, show an increased TTI and duration of
burning. The MLRs are slower and PHRRs are lower for thermally thick
samples compared to thermally thin samples.
Core materials, when present, can melt, burn, or char depending on the
type of material and hence can also influence the burning properties of the
composite [18]. TTI, for example, can be influenced by the thermal
conductivity of the core and the contact or lack thereof between the core and
the skin [28].
The bonding between the fiber reinforcement and the matrix polymer is
critical, not only for mechanical properties but also for the stability of the
composite when exposed to heat and fire. Improved compatibility between
the fibers and the matrix increases the thermal stability of the composite
[17]. Albano et al. [48], in a study on the thermal stability of polyolefin/sisal
composites, found that acetylated sisal fiber had a higher activation
energy compared to the untreated fiber and hence greater thermal stability.
They also found that the activation energy and maximum decomposition
temperature of polypropylene with acetylated fiber was higher than that for
polypropylene with untreated fiber. They concluded that the inclusion of
acetylated fiber results in greater stability of the polyolefin composites as a
result of greater polymer–fiber interaction.
All fires are different and the intensity of the fire, or incident heat flux,
also influences the flammability characteristics of a composite [18,49]. Fires
progress through a number of stages [49–51], as illustrated in Figure 1.
The initial stage of a fire is ignition. Ignition can occur over a long period
of time, with a high production of gases and low heat generation
(smoldering ignition), or over a short period, characterized by flaming
combustion and high heat release (flaming ignition). Flaming ignition may
be spontaneous (autoignition), when the flammable mixture of fuel volatile
gases and air reaches a sufficiently high enough temperature, or piloted,
when the flammable mixture is ignited by a localized heat source, such as a
spark or open flame. The heat flux and ambient temperatures are
low, ventilation is high, and flames are relatively small (510 cm). A period
of alternating flaming and smoldering may follow, influenced by the type
of fuel, interaction with surroundings, type of combustion, and access
to oxygen. The fire may then grow to a point of established burning, usually
at an external heat flux of about 20 kW/m2. The fire will now be in
the growth stage that is characterized by an external heat flux of around
20–60 kW/m2, higher ambient temperatures (700–900 K), continuing high

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


880 S. CHAPPLE AND R. ANANDJIWALA

Fully developed fire

Flashover
Temperature

Ignition Developing fire

Smoldering Decay
/ flaming

Time
Heat flux: Low 20–60 kW/m2 > 50 kW/m2
Temperature: 600– 700 K 700– 900 K > 900 K
Ventilation: High High Low

Figure 1. The stages of a typical fire and some fire properties at various stages (redrawn
from Schartel and Hull [49], Karlsson and Quintiere [50], and Fitzgerald [51]).

ventilation – the fire is still controlled by the fuel, and large flames (10–100 cm).
The transition to a fully developed fire is known as the flashover point, and
this may occur at temperatures in the region 773–873 K. A fully developed
fire is characterized by high ambient temperatures (4900 K), high external
heat flux (450 kW/m2), low ventilation – oxygen availability is now the
limiting factor, and large flames (4100 cm). The fully developed fire
continues until the fuel source is depleted. The fire will then enter a
period of decay, when temperatures drop, and eventually it will extinguish.
The properties of a fire, such as heat flux, therefore, vary considerably
during the course of a fire and consequently the flammability properties of
the composite will also vary. Scudamore [52] has shown, in cone calorimeter
testing, that the effect of thickness on TTI is largely negated at higher heat
fluxes (75 and 100 kW/m2). The availability of oxygen can also influence
TTI – increasing or decreasing the TTI depending on the oxygen
concentration and makeup of the composite [28].

STRATEGIES FOR FIRE RETARDANCY OF NFRC

The fire resistance of composites can be improved by strategies that


include reducing the flammability of the matrix and/or the fiber reinforce-
ment and/or the composite as a whole. There are a number of recent articles
and reviews that provide in-depth information on developments around
flame retardants and mineral fillers for polymers, for example, phosphorus-
based [53] and halogen-free [54] flame retardants, nanocomposites, and

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 881

mineral fillers [55,56] and flame retardants for specific polymers such as
polyolefins [57], polystyrenes [58], and poly vinyl chlorides [59]. The flame
retardancy of textiles [60–62], including natural fibers, and composites
[15,17,18,20] have also been reviewed. In this review, we will specifically
examine some recent methods from the literature that have been or could be
employed for improving the fire resistance of NFRC.

Polymer Matrix

AMMONIUM POLYPHOSPHATE
Ammonium polyphosphate (APP), [NH4PO3]n, has been in use as a flame
retardant for many years [26,63]. It has found use in intumescent systems
and as an additive in plastics, paper, textiles, and so forth. When heated,
APP is reported to decompose in three stages, starting at around 1658C [64].
During decomposition, ammonia and water are evolved, and polymeric
phosphoric acid and phosphorous oxides are formed [63,64]. In most
polymers, APP acts mainly in the condensed phase where acid catalyzed
dehydration reactions and cross-linking promote char formation [20,63]. In
some polymers, APP may also act in the gas phase [20]. Phosphorus radicals
are released from the polymer at temperatures below that required for
decomposition of the polymer. The radicals terminate the combustion
process by reacting with H and OH radicals in the flame. In addition,
heavy phosphorus-containing volatiles may form a vapor-rich phase at the
polymer surface that restricts oxygen access. Matkó et al. [13] investigated
the flammability of different NFRC using APP as a polymer additive. In a
polypropylene/wood flake composite containing either 10% or 20% APP,
they found that with the addition of 10% wood flake, the LOI increased
slightly compared to that of pure polypropylene. There were hardly any
changes in the LOI when the level of wood flake was increased between 10%
and 40%, but at 50%, the LOI values were 30% and 32% for the composites
with 10% and 20% APP, respectively. The 20% APP samples performed
slightly better compared to the 10% APP samples. They also compared
polyurethane composites having either corn shell or wood flake as a filler.
Incorporation of 20% APP in the composites increased the LOI value from
20% to 27% for the corn shell composite and from 23% to 31% for the
wood flake composite. They concluded that the polyurethane acted as a
char-forming agent that contributed to improved flame retardancy. They
also showed that the LOI value for a plasticized starch-based biopolymer
can be increased from 23% to 33% by the addition of 10% APP. The LOI
value can be further increased to 60% by the addition of 30% APP. The
polyol nature of the matrix and plasticizer contributed to the improved
flammability through char formation.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


882 S. CHAPPLE AND R. ANANDJIWALA

EXPANDABLE GRAPHITE
Expandable graphite (EG) expands up to 300 times its initial volume at
around 9008C [65,66]. The flameless oxidation of the expanded graphite
consumes oxygen that results in the smothering of flames. The expanded
graphite also acts as a heat insulating layer and reduces dripping. Schartel
et al. [66] compared the thermal behavior and fire response of polypropy-
lene/flax composites using either EG or APP as the flame-retardant additive.
In TGA, all samples showed two-step decomposition: the first being
decomposition of the flax and the second for decomposition of the
polypropylene. The addition of 25% APP to the composite in place of
polypropylene resulted in the first-step decomposition occurring at a lower
temperature (a decrease of around 558C) and the second-step decomposition
occurring at a higher temperature (an increase of around 98C). The higher
temperature of the second-step decomposition was attributed to the
formation of a heat insulating char. The addition of 15% or 25% EG to
the composite instead of APP, however, resulted in enhanced char formation
in the first-step decomposition but little change in the decomposition
temperature. The temperature of the second-step decomposition was
increased (5–98C) as a result of the thermal insulation properties of the
expanded graphite. In cone calorimeter testing, they showed that both APP
and EG as additives reduced the PHRR and THR; however, APP also
resulted in an increase in smoke and carbon monoxide production per unit
mass loss. They concluded that EG reduced both fire risk and fire hazard,
whereas APP reduced fire risk but increased fire hazard.

ALUMINIUM TRIHYDRATE
Aluminium trihydrate (ATH), 2Al(OH)3, is an active filler for reducing
the flammability of composites [18,20]. It is an effective, low-cost flame
retardant and is widely used even though a high loading (450% for
some polymers) and is often required to achieve adequate flame retardancy.
ATH starts decomposing at around 2208C, so composite processing
temperatures should be below 2008C. ATH decomposes by the endothermic
reaction as follows:

2AlðOHÞ3 ! Al2 O3 þ3H2 O

The main endothermic peak occurs at about 3008C, which means that it is
absorbing heat from the polymer at a temperature below that at which most
composite polymers degrade. The water released during decomposition
restricts oxygen access to the composite surface and dilutes the concentration
of flammable gases evolved. It is thought that it may promote char formation
in some polymers by a condensation phase mechanism. In their study on an

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 883

unsaturated polyester/hemp sheet moulding compound, Hapuarachchi et al.


[14] used ATH (40 wt%) as a fire retardant. They showed that ATH delayed
TTI and reduced PHRR (Table 5) and concluded that, with an addition of an
adequate level of flame retardant to the sheet mould compound paste, a
natural fiber-reinforced sheet moulding compound can compete, in terms of
fire performance, with some alternative building materials.

MAGNESIUM HYDROXIDE
Another endothermic flame retardant is magnesium hydroxide [20],
Mg(OH)2, which decomposes by the endothermic reaction as follows:

MgðOHÞ2 ! MgO þ H2 O

It is stable up to 330–3408C and so can be processed at higher temperatures


compared to ATH, but its decomposition temperature is similar to that of
some matrix polymers, such as polyesters, vinyl esters, and epoxies, so it is less
effective than ATH. Like ATH, magnesium hydroxide also absorbs heat from
the polymer, slowing down burning, and the water released during
decomposition reduces the concentration of flammable gases and H and
OH radicals in the flame. The magnesium oxide (MgO) produced also has an
insulating effect. Sain et al. [67] investigated the effect of magnesium
hydroxide on sawdust and rice husk-filled polypropylene composites.
Replacement of 25% of the natural filler with magnesium hydroxide reduced
the horizontal burning rate (ASTM D 635) by around 50% for both fillers.
The LOI was higher for the composites with flame retardant in comparison to
that without flame retardant and polypropylene on its own.

OTHER ADDITIVES
Other strategies regarding the addition of flame retardants could include
the addition of catalysts or nanocatalysts [40]. According to Lewin [40], it is
thought that no chemical interaction takes place between endothermic fire
retardants, such as magnesium hydroxide and ATH, and the polymer during
pyrolysis and combustion. However, in the presence of a catalyst such as

Table 5. TTI and peak heat release values of hemp fiber/


unsaturated polyester sheet molding compounds [14].

Irradiance 25 kW/m2 Irradiance 50 kW/m2


TTI (s) PHRR (kW/m2) TTI (s) PHRR (kW/m2)
No ATH 185 290 54 362
40 wt% ATH 290 159 78 180

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


884 S. CHAPPLE AND R. ANANDJIWALA

manganese borate, ferrocene, and nickel II oxide, it was found that chemical
interactions occurred between the fire retardant and the polymer, which
resulted in increased char and oxygen index.

INTUMESCENT SYSTEMS
Intumescents can be used as additives for thermoplastic and thermoset
resins, acting as char promoters [20,47]. An intumescent system is illustrated
in Figure 2.
The system consists of a dehydrating agent that releases acid during
thermal decomposition, a carbonizing substance, containing a large number
of carbon and hydroxyl groups, which undergoes esterification with the acid
and later decomposes to form carbon, acid, water, and CO2, and a foam
producing substance, which evolves large quantities of nonflammable gases.
An intumescent char is then formed from the decomposition products of the
carbonizing agent and the nonflammable gases [20,47,68]. The char forms an
insulating layer that retards heat transfer and retards oxygen access. Le Bras
et al. [69] studied the effect of an intumescent system, comprising APP,
pentaerythritol, and melamine, in a polypropylene/flax composite. In UL-94
testing [70], they found that virgin polypropylene and a polypropylene
composite burnt completely and had a no classification (NC) rating. A
polypropylene/flax/APP composite formed a carbonaceous char but still
had a NC rating, whereas the polypropylene/flax composite with the
intumescent system formed a stable intumescent char and achieved a V-0
rating. They showed that an optimized flame-retardant formulation could
be obtained by the association of a char-forming cellulose material with an
intumescent char structure.

Decomposition of inorganic salt


(100 – 250°C)

Decomposition of carbonizing substance Decomposition of foaming agent

Hot, viscous char Nonflammable gases

Intumescent char
(50 – 200 times original size)
Figure 2. Intumescent system (redrawn from Mouritz and Gibson [20] and Kandola and
Nazare [47]).

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 885

RESIN BLENDS
Kandola et al. [47] reported on the use of resin blends to improve flame
retardancy. This involves blending of polymers with poor flame retardancy
with polymers having good flame retardancy. For example, Chiu et al. [71]
studied the thermal degradation and combustion properties of an
unsaturated polyester resin blended with a phenolic resin. The blended
resin showed improved heat resistance and a reduction in heat release and
smoke and toxic gas emissions.

NANOCOMPOSITES
Another option is the incorporation of nanoparticles such as clays,
organically modified clays, or low melting glasses [15,18,40,47,72,73].
The clays are not flame retardant in themselves; however, their introduction
into a polymer can reduce PHRR, although THR remains the same. There are
several mechanisms in play, and they impact on different fire properties. It is
proposed that as the polymer is exposed to heat, platelets migrate to the
surface providing a protective layer that delays gasification of the surface
below the layer. The clays also assist in formation of a carbonaceous-silicate
char, which acts as a heat barrier, increasing thermal stability, and a physical
barrier to mass transport. The nanoparticles can also reduce the melt flow of
the polymer, and the change in melt viscosity and barrier formation results in
reduced flame spread. Nanoparticles can be used in combination with other
flame retardants. There is a synergistic effect, which gives improved flame
retardancy and would allow for a reduction in the amount of flame retardant
used. Kozlowski et al. [74] reported on the flammability of poly(lactic acid)/
organically modified montmorillonite nanocomposites. In UL-94 horizontal
burn testing, they found that the clays decreased the rate of burning. Marosfoi
et al. [75] examined the combustion properties of sepiolite, an organically
treated sepiolite, both needle-like clays, and montmorillonite, a layered clay,
in combination with magnesium hydroxide in a polypropylene matrix. They
found that the PHRR and TTI of the organically modified clay/magnesium
hydroxide combination were lower than when either of the additives was
used alone. For example, the PHRR of polypropylene with 10% magnesium
hydroxide was 471 kW/m2 and that of polypropylene containing 10%
magnesium, hydroxide, and 5% organically modified sepiolite was 328 kW/
m2. They proposed that the clay had a catalytic effect in promoting charring
and a micromechanical effect, acting as a fibrous reinforcement, which
improved the mechanical properties of the char. Combination of sepiolite and
montmorillonite clays with the flame retardant resulted in increased TTI and
decreased HRR when compared to the flame retardant on its own. They
concluded that the two types of clay worked synergistically with the flame
retardant and a further advantage could be gained by combining the two.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


886 S. CHAPPLE AND R. ANANDJIWALA

Reinforcing Fibers

DIAMMONIUM PHOSPHATE
Diammonium phosphate (DAP), (NH4)2HPO4, is a condensed phase
flame retardant that has been used for a long time as a nondurable flame
retardant for cellulosic textiles [32,44,76,77]. DAP decomposes at around
1558C, well below the decomposition temperature of cellulose, forming
phosphoric acid and ammonia. The phosphoric acid formed can phosphor-
ylate the primary hydroxyl group of cellulose to form a phosphorous
ester. These esters catalyze the dehydration of cellulose, promoting the
formation of char and water at the expense of levoglucosan. The acid
may also cross-link with the cellulose changing the normal pathway
of pyrolysis to yield less flammable products. Matkó et al. [13] investigated
the flame retardancy of polyurethane/cellulosic composites using DAP
as a flame retardant for the cellulosic component. A polyurethane/wood
flake composite had a LOI value of 23% and a polyurethane/corn
shell composite had a LOI value of 20%. When DAP was used as the
flame-retarding agent, the composites had LOI values of 30% and 29%,
respectively. UL 94 ratings improved from H-B (no flame retardant) to V-0
(DAP). They found little difference in the level of flame retardancy
when using as an additive in the resin instead of treatment of the cellulosic
with DAP.

TIN FLAME RETARDANTS


The first patent for the use of tin-based flame retardants for textiles was
filed by Perkin in 1902 [78]. Inorganic tin chemicals have also been used for
treatments on wool [79] and polymers [80,81]. Their mode of action is
complex but may involve radical scavenging, char promotion, and soot
oxidation on the flame. In Perkin’s method, the flame-retardant treatment is
effected by acidification of an alkali stannate, which results in deposition of
insoluble metastannic acid in the fiber [62]. Metastannic acid can also be
produced by hydrolysis of stannous chloride. The use of stannous chloride
as part of a silicon-inorganic flame retardant was reported by Lomakin et al.
[82]. They proposed a reaction of stannous chloride with silicon to produce
silicon chloride and hydrochloric acid. Misra et al. [83] investigated the use
of a stannous chloride treatment of brominated coir fiber in an epoxy-based
composite. They found that the tin-based halogen compound acted as a fire
retardant at very low levels; approximately 5% of the fire retardant reduced
smoke density by 25% and increased the LOI value of the untreated coir/
epoxy microcomposite from 35.64 to 38.90. The LOI value for the unfilled
epoxy microcomposite was 31.50. No mechanism for the action of the
stannous chloride was put forward.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 887

OTHER FLAME-RETARDANT OPTIONS


The possibility exists to examine other flame-retardant treatments for
natural fibers, as a means of improving the flame retardancy of NFRC,
including those that may have lost favor in textile applications because of
their limited durability or those that have so far had limited use. Sulfamation
with sulfamic acid salts, such as ammonium sulfamate, offers possibilities
[40]. The crystalline regions of sulfamated cellulosics are more easily
hydrolyzed during pyrolysis compared to those in phosphorylated cellulosics.
The reduction in crystallinity decreases the amount of levoglucosan and hence
the flammability of the fiber.

FIBER MODIFICATION
The usefulness of acetylation of sisal in improving thermal stability has
already been mentioned [48]. Thermal stability can also be improved by
grafting. Sabaa [84] reported that the grafting of acrylonitrile on sisal fibers
improved thermal stability, as indicated by reduced weight loss and rate of
degradation and an increased initial degradation temperature. An increased
degradation temperature from 1708C to 2808C for jute grafted with
acrylonitrile was reported by Mohanty et al. [85].
Novel fiber modifications to improve fire resistance include a keratin-
cellulose filament, produced from wool fiber wastes and cellulose acetate.
The filament, with improved fire resistance, was reported by Aluigi et al. [86].
White et al. [87] reported on a spun fiber that contained 93% cotton and 7%
clay. The composite fiber was found to have a heat tolerance of 30–358C
greater than that of either cotton or regenerated cellulose fibers. The fibers
formed a char outer layer when heated.

COMPOSITES
The fire resistance of the composite can be improved by applying a fire
retardant or thermal barrier coating [18,20,47]. Commonly, these are only
applied to thermoset composites such as those based on epoxy, polyester,
and vinyl ester resins and challenges exist for their use on thermoplastic-
based composites. A coating has the advantage that it does not modify the
intrinsic properties of the composite. The coating can be in the form of an
organic polymer such as a phenolic or alkyd resin, with or without a flame-
retardant filler. These coatings delay ignition and prevent flame spread and
heat penetration. They are generally low cost and effective, but there is a
limit to the thickness that can be applied and hence the level of protection
obtained. Inorganic polymers such as geopolymers and polyhedral
oligomeric silsesquioxanes can also be used. They provide a better level of
protection compared to organic polymers because of their greater resistance
to pyrolysis and heat conduction.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


888 S. CHAPPLE AND R. ANANDJIWALA

Thermal barrier coatings, such as ceramic or mineral fibers, which are


bonded to the composite, are another option. They reflect heat away from
the composite and increase ignition resistance. They are generally easy to
apply, economical, and efficient [20,47].
Intumescent coatings, which form a high-volume char on exposure to
heat, may also be used. This char provides an insulating layer that retards
heat transfer and restricts oxygen access to the underlying composite [20,47].
Other flame-retardant coatings could also possibly be considered such as
polyvinyl chloride, polyurethane, silicone, and fluoropolymers; however, the
use of coatings on NFRC would be dependant on a number of factors
including fire performance requirements, type of matrix polymer, the level of
bonding of the coating to the matrix polymer, thickness required to be
effective, processing conditions, cost, and aesthetic and environmental
concerns.

SUMMARY

The use of natural fibers in composite materials is gaining momentum


because of their renewability, biodegradability, cost, and mechanical
properties. From their first application in Gordon Aerolite, it is only recently
that there has been renewed interest in their application in the aerospace
industry. Application in this and other industries where material flammability
and safety are important poses new challenges. Regulatory requirements
around flammability, smoke, and toxic gases must be met. Recent research
has shown that the fire resistance of NFRC can be improved. These advances
have shown that THR, PHRR, CO yields, and smoke emissions can be
reduced and char production can be increased. This has been achieved by
application of knowledge gained in the fire retardancy of thermoset and
thermoplastic polymers, textiles and composites, and by some new and
innovative developments. However, there is still much work to be done on the
flammability of NFRC to ensure their successful application.

REFERENCES

1. Bishopp, J.A. (1997). The History of ReduxÕ and the Redux Bonding Process, International
Journal of Adhesion and Adhesives, 17: 287–301.
2. Invent Now. Hall of Fame. Russell Games Slayter. Available at: http://www.invent.org
(accessed January, 2009).
3. Hill, C. (2007). Natural Fibre Reinforced Composites Opportunities and Challenges, Keynote
Address, In: Proceedings of Fibre Reinforced Composites Conference 2007, Port Elizabeth.
4. van Rijswijk, K., Brouwer, W.D. and Beukers, A. (2003). Application of Natural Fibre
Composites in the Development of Rural Societies, Food and Agriculture Organisation of the
United Nations, Rome.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 889

5. Wang, B., Panigrahi, S., Tabil, L., Crerar, W., Powell, T., Kolybaba, M. and Sokhansanj,
S. (2003). Flax Fibre-reinforced Thermoplastic Composites, Paper number: RRV03-0003,
In: The 2003 CSAE/ASAE Annual Intersectional Meeting, Fargo, North Dakota.
6. Puglia, D., Biagiotti, J. and Kenny, J.M. (2004). A Review on Natural Fibre-based
Composites- Part II: Application of Natural Reinforcements in Composite Materials for
Automotive Industry, Journal of Natural Fibers, 1(3): 23–65.
7. Anandjiwala, R.D., John, M.J., Wambua, P., Chapple, S., Klems, T., Doecker, M.,
Goulain, M. and Erasmus, L. (2007). Bio-based Structural Composite Materials for
Aerospace Applications, In: 2nd SAIAS Symposium, Stellenbosch.
8. Giancaspro, J., Papakonstantinou, C. and Balaguru, P. (2009). Mechanical Behavior of
Fire-Resistant Biocomposite, Composites: Part B, 40(3): 206–211.
9. Giancaspro, J., Papakonstantinou, C. and Balaguru, P. (2007). Fire Resistance of
Inorganic Sawdust Biocomposite, Composites Science and Technology, 68: 1895–1902.
10. Biagiotti, J., Puglia, D. and Kenny, J.M. (2004). A Review on Natural Fibre-based
Composites- Part I: Structure, Processing and Properties of Vegetable Fibres, Journal of
Natural Fibers, 1(2): 37–68.
11. The Clean Sky: Joint Technology Initiative. Available at: http://www/cleansky.eu (accessed
March 2009).
12. Nabi Saheb, D. and Jog, J.P. (1999). Natural Fibre Polymer Composites: A Review,
Advances in Polymer Technology, 18(4): 351–363.
13. Matkó, Sz., Toldy, A., Keszei, S., Anna, P., Bertalan, Gy. and Marosi, Gy. (2005). Flame
Retardancy of Biodegradable Polymers and Biocomposites, Polymer Degradation and
Stability, 88: 138–145.
14. Hapuarachchi, T.D., Ren, G., Fan, M., Hogg, P.J. and Peijs, T. (2007). Fire Retardancy of
Natural Fibre Reinforced Sheet Moulding Compound, Applied Composite Materials, 14:
251–264.
15. Kozlowski, R. and Wladyka-Przybylak, M.W. (2008). Flammability and Fire Resistance of
Composites Reinforced by Natural Fibres, Polymers for Advanced Technologies, 19:
446–453.
16. Blicblau, A.S., Coutts, R.S.P. and Sims, A. (1997). Novel Composites Utilizing Raw Wool
and Polyester Resin, Journal of Materials Science Letters, 16: 1417–1419.
17. Horrocks, A.R. and Kandola, B.K. (2005). Flammability and Fire Resistance of
Composites, In: Long, A.C. (ed.), Design and Manufacture of Textile Composites,
pp. 330–363, Woodhead, Cambridge.
18. Kandola, B.K. and Kandare, E. (2008). Composites Having Improved Fire Resistance,
In: Horrocks, A.R. and Price, D. (eds), Advances in Fire Retardant Materials, pp. 398–442,
Woodhead, Cambridge.
19. Mark, H.F., Atlas, S.M., Shalaby, S.W. and Pearce, E.M. (1975). Combustion of Polymers
and its Retardation, In: Lewin, M., Atlas, S.M. and Pearce, E.M. (eds), Flame-retardant
Polymeric Materials, pp. 1–17, Plenum Press, New York.
20. Mouritz, A.P. and Gibson, A.G. (2006). Flame Retardant Composites, In: Gladwell,
G.M.L. (series ed.), Fire Properties of Polymer Composite Materials, pp. 237–286, Springer,
London.
21. Hirschler, M.M. (2000). Chemical Aspects of Thermal Decomposition of Polymeric
Materials, In: Grand, A.F. and Wilkie, C.A. (eds), Fire Retardancy of Polymeric Materials,
pp. 28–79, CRC Press, Boca Raton.
22. ASTM D 2863 (2008). Standard Test Method for Measuring the Minimum Oxygen
Concentration to Support Candle-like Combustion of Plastics (Oxygen Index), ASTM
Publication, Philadelphia.
23. Kourtides, D.A., Gilwee Jr, W.J. and Parker, J.A. (1979). Thermochemical
Characterisation of Some Thermally Stable Thermoplastic and Thermoset Polymers,
Polymer Engineering and Science, 19: 24–29.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


890 S. CHAPPLE AND R. ANANDJIWALA

24. Brydson, J.A. (ed.) (1999). Relation of Structure to Chemical Properties, In: Plastics
Materials, pp. 76–109, Butterworth-Heinemann, Oxford.
25. Flame Retardants for Electrical Equipment (2008). Premier Farnell. Available at: http://
de.farnell.com (accessed March 2009).
26. Lyons, J.W. (ed.) (1970). Synthetic Polymers with All-carbon Backbones, In: The
Chemistry & Uses of Fire Retardants, pp. 281–344, Wiley-Interscience, New York.
27. Vinyl Council of Australia (2003). PVC and Building Fires, Vinyl Council of Australia.
Available at: http://www.vinyl.org.au (accessed January, 2009)
28. Mouritz, A.P. and Gibson, A.G. (eds.) (2006). Fire Reaction Properties of Composites, In:
Gladwell, G.M.L. (series ed.), Fire Properties of Polymer Composite Materials, pp. 59–101,
Springer, London.
29. Roff, W.J. and Scott, J.R. (eds.) (1971). Regenerated Proteins, In: Fibres, Films, Plastics
and Rubbers: Handbook of Common Polymers, pp. 197–204, Butterworths, London.
30. Manfredi, L.B., Rodrı́guez, E.S., Wladyka-Przybylak, M. and Vázquez, A. (2006). Thermal
Degradation and Fire Resistance of Unsaturated Polyester, Modified Acrylic Resins and
their Composites with Natural Fibres, Polymer Degradation and Stability, 91: 255–261.
31. Kozlowski, R. and Wladyka-Przybylak, M. (2004). Uses of Natural Fibre Reinforced
Plastics, In: Wallenberger, F.T. and Weston, N.E. (eds), Natural Fibres, Plastics and
Composites, pp. 249–274, Kluwer Academic, Dordrecht.
32. Horrocks, A.R. (1983). An Introduction to the Burning Behaviour of Cellulosic Fibres,
Journal of the Society of Dyers and Colourists, 99(7/8): 191–197.
33. Lewin, M. and Basch, A. (1978). Structure, Pyrolysis and Flammability of Cellulose,
In: Lewin, M., Atlas, S.M. and Pearce, E.M. (eds), Flame-retardant Polymeric Materials,
Vol. 2, pp. 1–41, Plenum Press, New York.
34. Russell, L.J., Marney, D.C.O., Humphrey, D.G., Hunt, A.C., Dowling, V.P. and Cookson, L.J.
(2004). Combining Fire Retardant and Preservative Systems for Timber Products in Exposed
Applications – State of the Art Review, Forest and Wood Products R&D Corporation, Victoria.
35. Shafizadeh, F. (1984). The Chemistry of Pyrolysis and Combustion, In: Rowell, R. (ed.),
The Chemistry of Solid Wood, Advances in Chemistry Series, Vol. 207, pp. 489–529,
American Chemical Society, Washington.
36. LeVan, S.L. and Winandy, J.E. (1990). Effects of Fire Retardant Treatments on Wood
Strength: A Review, Wood and Fiber Science, 22(1): 113–131.
37. Ferdous, D., Dalai, A.K., Bej, S.K. and Thring, R.W. (2002). Pyrolysis of Lignins:
Experimental and Kinetics Studies, Energy & Fuels, 16: 1405–1412.
38. Nikitin, N.I. (1966). Thermal Decomposition of Wood, In: Schmorak, J. (translator),
The Chemistry of Cellulose and Wood, pp. 570–596, Israel Program for Scientific
Translations, Jerusalem.
39. Mo, X., Wang, D. and Sun, X.S. (2005). Straw-based Biomass and Biocomposites,
In: Mohanty, A.K., Misra, M. and Drzal, L.T. (eds), Natural Fibres, Biopolymers and
Biocomposites, pp. 476–478, CRC Press, Boca Raton.
40. Lewin, M. (2005). Unsolved Problems and Unanswered Questions in Flame Retardance of
Polymers, Polymer Degradation and Stability, 88: 13–19.
41. Basch, A. and Lewin, M. (1973). The Influence of Fine Structure on the Pyrolysis of
Cellulose. I. Vacuum Pyrolysis, Journal of Polymer Science – Polymer Chemistry Edition,
11(12): 3071–3093.
42. Hshieh, F.Y. and Beeson, H.D. (1996). Flammability Testing of Flame-Retarded Epoxy
Composites and Phenolic Composites, In: Proceedings of the International Conference on
Fire Safety, San Francisco, Vol. 21, pp. 189–205.
43. Brown, J.R., Fawell, P.D. and Mathys, Z. (1994). Fire-hazard Assessment of Extended-
chain Polyethylene and Aramid Composites by Cone Calorimetry, Fire Materials, 18:
167–172.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 891

44. Schindler, W.D. and Hauser, P.J. (eds.) (2004). Flame-retardant Finishes, In: Chemical
Finishing of Textiles, pp. 98–116, Woodhead, Cambridge.
45. Helwig, M. and Paukszta, D. (2000). Flammability of Composites Based on Polypropylene
and Flax Fibres, Molecular Crystals & Liquid Crystals, 354: 373–380.
46. Borysiak, S., Paukszta, D. and Helwig, M. (2006). Flammability of Wood-Polypropylene
Composites, Polymer Degradation and Stability, 91: 3339–3343.
47. Kandola, B.K. and Nazare, S. (2007). An Overview of Flammability and Fire Retardancy
of Fibre Reinforced Unsaturated Polyester Composites, In: Proceedings of Fibre Reinforced
Composites Conference, Port Elizabeth.
48. Albano, C., González, J., Ichazo, M. and Kaiser, D. (1999). Thermal Stability of Blends of
Polyolefins and Sisal Fibre, Polymer Degradation and Stability, 66: 179–190.
49. Schartel, B. and Hull, T.R. (2007). Development of Fire-retardant Materials –
Interpretation of Cone Calorimeter Data, Fire and Materials, 31: 327–354.
50. Karlsson, B. and Quintiere, J.G. (2000). A Qualitative Description of Enclosed Fires,
In: Lilley, D.G and Gupta, A.K (series eds), Enclosed Fire Dynamics, pp. 17–18, CRC
Press, Boca Raton.
51. Fitzgerald, R.W. (ed.) (2004). Design Fires, In: Building Fire Performance Analysis, pp. 95–
130, John Wiley & Sons, New York.
52. Scudamore, M.J. (1994). Fire Performance Studies on Glass-reinforced Plastic Laminates,
Fire and Materials, 18: 313–325.
53. Levchik, S.V. and Weil, E.D. (2006). A Review of Recent Progress in Phosphorus-based
Flame Retardants, Journal of Fire Sciences, 24: 345–364.
54. Lu, S.-Y. and Hamerton, I. (2002). Recent Developments in the Chemistry of Halogen-
Free Flame Retardant Polymers, Progress in Polymer Science, 27: 1661–1712.
55. Le Bras, M., Wilkie, C.A. and Bourbigot, S. (2005). Fire Retardancy of Polymers. New
Applications for Mineral Fillers, Royal Society of Chemistry, Cambridge.
56. Bourbigot, S. and Duquesne, S. (2007). Fire Retardant Polymers: Recent Developments
and Opportunities, Journal of Materials Chemistry, 17: 2283–2300.
57. Weil, E.D. and Levchik, S.V. (2008). Flame Retardants in Commercial Use or
Development for Polyolefins, Journal of Fire Sciences, 26: 5–43.
58. Weil, E.D. and Levchik, S.V. (2007). Flame Retardants for Polystyrene in Commercial Use
or Development, Journal of Fire Sciences, 25: 241–265.
59. Weil, E.D., Levchik, S. and Moy, P. (2006). Flame and Smoke Retardants in Vinyl
Chloride Polymers – Commercial Usage and Current Developments, Journal of Fire
Sciences, 24: 211–236.
60. Horrocks, A.R., Kandola, B.K., Davies, P.J., Zhang, S. and Padbury, S.A. (2005).
Developments in Flame Retardant Textiles – A Review, Polymer Degradation and Stability,
88: 3–12.
61. Weil, E.D. and Levchik, S.V. (2008). Flame Retardants in Commercial Use or
Development for Textiles, Journal of Fire Sciences, 26: 243–281.
62. Horrocks, A.R. (1986). Flame-retardant Finishing of Textiles, Review of Progress in
Coloration, 16: 62–101.
63. EFRA (2007). Flame Retardants Fact Sheet: Ammonium Polyphosphate, European Flame
Retardants Association, Brussels. Available at: www.cefic-efra.eu (accessed May, 2008)
64. Camino, G., Costa, L. and Trossarelli, L. (1985). Study of the Mechanism of Intumescence
in Fire Retardant Polymers: Part V – Mechanism of Formation of Gaseous Products in the
Thermal Degradation of Ammonium Polyphosphate, Polymer Degradation and Stability,
12(3): 203–211.
65. Schilling, B. (1997). Expandable Graphite, Kunstoffe Plast Europe, 87(8): 16–17.
66. Schartel, B., Braun, U., Schwarz, U. and Reinemann, S. (2003). Fire Retardancy of
Polypropylene/Flax Blends, Polymer, 44: 6241–6250.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


892 S. CHAPPLE AND R. ANANDJIWALA

67. Sain, M., Park, S.H., Suhara, F. and Law, S. (2004). Flame Retardant and Mechanical
Properties of Natural Fibre-PP Composites Containing Magnesium Hydroxide, Polymer
Degradation and Stability, 83: 363–367.
68. Wladyka-Przybylak, M. and Kozlowski, R. (1999). The Thermal Characteristics of
Different Intumescent Coatings, Fire and Materials, 23: 33–43.
69. Le Bras, M., Duquesne, S., Fois, M., Grisel, M. and Poutch, F. (2005). Intumescent
Polypropylene/Flax Blends: A Preliminary Study, Polymer Degradation and Stability, 88:
80–84.
70. UL-94, Standard for Flammability of Plastic Materials for Parts in Devices and
Appliances, Underwriters Laboratories. Available at: http://www.ul.com (accessed
March 2009).
71. Chiu, H.T., Chiu, S.H., Jeng, R.E. and Chung, J.S. (2000). A Study of the Combustion and
Fire-Retardance Behaviour of Unsaturated Polyester/Phenolic Resin Blends, Polymer
Degradation and Stability, 70: 505–514.
72. Yang, F., Yngard, R. and Nelson, G.L. (2005). Flammability of Polymer-Clay and
Polymer-Silica Nanocomposites, Journal of Fire Sciences, 23: 209–226.
73. Marosfo00 i, B.B., Marosi, Gy.J., Szép, A., Anna, P., Keszei, S., Nagy, B.J., Martvonova, H.
and Sajó, I.E. (2006). Complex Activity of Clay and CNT Particles in Flame Retarded
EVA Copolymer, Polymers for Advanced Technologies, 17: 255–262.
74. Kozlowski, M., Iwanczuk, A., Frackowiak, S. and Szczurek, T. (2007). Biocomposites of
Functional Properties, In: International Sustainable Materials, Polymers & Composites
Conference, Warwick.
75. Marosfoi, B.B., Garas, S., Bodzay, B., Zubonyai, F. and Marosi, G. (2008). Flame
Retardancy Study on Magnesium Hydroxide Associated with Clays of Different
Morphology in Polypropylene Matrix, Polymers for Advanced Technologies, 19: 693–700.
76. Lewin, M. and Sello, S.B. (1975). Technology and Test Methods of Flame Proofing of
Cellulosics, In: Lewin, M., Atlas, S.M. and Pearce, E.M. (eds), Flame-retardant Polymeric
Materials, pp. 19–136, Plenum Press, New York.
77. Lyons, J.W. (ed.) (1970). Cellulose: Textiles, In: The Chemistry & Uses of Fire Retardants,
pp. 165–240, Wiley-Interscience, New York.
78. Perkin, W.H. (1907). Fireproofing of Textile Material, US Patent No. 844,042, February
12.
79. Cusack, P.A., Smith, P.J., Brooks, J.S. and Smith, R. (1979). A Study of Flame-
resist Treatments of Wool by Inorganic Tin Chemicals, Journal of the Textile Institute, 7:
308–315.
80. Atkinson, P.A., Haines, P.J. and Skinner, G.A. (2000). Inorganic Tin Compounds as
Flame Retardants and Smoke Suppressants for Polyester Thermosets, Thermochimica
Acta, 360: 29–40.
81. Xu, J., Jiao, Y., Zhang, B., Qu, H. and Yang, G. (2006). Tin Dioxide Coated Calcium
Carbonate as Flame Retardant for Semirigid Poly(vinyl chloride), Journal of Applied
Polymer Science, 101(1): 731–738.
82. Lomakin, S.M., Zaikov, G.E. and Artsis, M.I. (1996). Polypropylene Flame Retardant
System Based on Si-SnCl2, International Journal of Polymeric Materials, 32: 203–211.
83. Misra, R.K., Kumar, S., Sandeep, K. and Misra, A. (2008). Some Experimental and
Theoretical Investigations on Fire Retardant Coir/Epoxy Micro-composites, Journal of
Thermoplastic Composite Materials, 21: 71–101.
84. Sabaa, M.W. (1991). Acrylonitrile as Thermal Stabilizer for Sisal Fibres, Polymer
Degradation and Stability, 32: 209–274.
85. Mohanty, A.K., Pattanaik, S., Singh, B.C. and Misra, M. (1989). Graft Copolymerisation
of Acrylonitrile Onto Acetylated Jute Fibres, Journal of Applied Polymer Science, 37:
1171–1181.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015


Flammability of Natural Fiber-reinforced Composites 893

86. Aluigi, A., Vineis, C., Ceria, A. and Tonin, C. (2008). Composite Biomaterials from Fibre
Wastes: Characterization of Wool-Cellulose Acetate Blends, Composites. Part A, Applied
Science and Manufacturing, 39(1): 126–132.
87. White, L.A. and Delhom, C.D. (2004). Cellulose-based Nanocomposites: Fibre Production
and Characterization, In: Polymeric Materials: Science and Engineering Preprints, 227th
American Chemical Society Meeting, Anaheim, California, Vol. 90, pp. 45–50.
88. Sarma, U.S. and Kumararaja M. (1997). Manifold Potentials of Coir Geotextiles,
Coir News, 26(4): 27–30.

Downloaded from jtc.sagepub.com at DALHOUSIE UNIV on January 5, 2015

Anda mungkin juga menyukai