Materials Research Innovations
ISSN: 1432-8917 (Print) 1433-075X (Online) Journal homepage: https://www.tandfonline.com/loi/ymri20
Flame and fire retardancy of polymer-based
composites
Radhashyam Giri, Lalatendu Nayak & Mostafizur Rahaman
To cite this article: Radhashyam Giri, Lalatendu Nayak & Mostafizur Rahaman (2020): Flame
and fire retardancy of polymer-based composites, Materials Research Innovations, DOI:
10.1080/14328917.2020.1728073
To link to this article: https://doi.org/10.1080/14328917.2020.1728073
Published online: 15 Feb 2020.
Submit your article to this journal
Article views: 13
View related articles
View Crossmark data
Full Terms & Conditions of access and use can be found at
https://www.tandfonline.com/action/journalInformation?journalCode=ymri20
MATERIALS RESEARCH INNOVATIONS
https://doi.org/10.1080/14328917.2020.1728073
Flame and fire retardancy of polymer-based composites
Radhashyam Giria, Lalatendu Nayakb and Mostafizur Rahamanc
a
c
Department of Plastics Technology, Central Institute of Plastic Engineering and Technology, India; bPhillips Carbon Black Limited, Kolkata, India;
Department of Chemistry, College of Science, King Saud University, Riyadh, Saudi Arabia
ABSTRACT
ARTICLE HISTORY
In this article, the recent development concerning the flame and fire-retardant characteristics of
polymer-based composites has been discussed. The process of polymer combustion and its inhibition, and the chemistry behind the fire retardancy are mentioned in the background. Focussed are
given on mainly inorganic hydroxide, halogen, phosphorus, nitrogen, silicon, boron, clay-layered
silicates, metal hydroxides, and carbons based flame retardants as the use of these different
nanofiller additives designed the polymer for enhancing flame retardancy. Emphases are given
on how these nanofillers are beneficial to retard the flame from spreading during the development
of fire. The usefulness to incorporate the conventional flame retardants in nanocomposites is also
discussed. Finally, the mechanism of flame retardancy for clay-and carbon-based nanocomposites is
reported.
Received 29 September 2019
Accepted 3 February 2020
1. Introduction
1.1. Polymers and fire
Polymers have been extensively used in many applications
because of their light-weight. Many commodity and engineering polymers are replacing conventional metal and ceramic materials, as they comparatively possess many
advantages. But the usage of these materials has greatly
increased the risk of fire because of their ability to catch
fire and possible release of toxic byproducts. However, the
majority of organic compounds will burn readily in air or
oxygen. The polymers, having high or even lower flammability have a difficult issue and extremely confine their
applications [1,2]. Lately, the fire security department has
expressed their concern about the necessity of suitable materials in wide area of applications, which result in increasing
demand for improved flame or fire-retardant materials at
stringent condition [3–5]. The polymer/plastic materials are
used in houses, business and transportation sectors.
Consequently, the fire-retardant matter can be directly associated with the nature of polymers [6]. The regular fire
retardants are halogen-based exacerbates those are prudent
and can upgrade the fire retardancy of polymers without
debasing their physical property, for example, quality. In
any case, lethal species, for example, dioxins and furans,
which are evolved during the ignition of halogencontaining composites, may cause serious environmental
contamination [7–9]. Therefore, developing halogen-free,
low-smoke, and environmental-friendly fire-retardant carbon composites has become increasingly important in recent
years. Inorganic hydroxides, such as aluminium hydroxide
or magnesium hydroxide are examples of the most extensively used inorganic fire retardants at the present time due
to their non-toxic and environmental-friendly characteristics
[10–14]. In any case, to meet the purposes, high amount of
loading (30 wt%-60 wt %) is expected that prompts extra
cost, processing difficulty and a decrease in physical
CONTACT Mostafizur Rahaman
Riyadh 11451, Saudi Arabia
Nanocomposites; flame and
fire retardancy; nanofillers;
layered silicates
properties of polymers. Consequently, the improvement of
new exceptionally successful green fire retardants has incited
much consideration amid the last ten years [5,6,13,15].
Nano-scale layered metal materials, such as clay and layered
double hydroxides (LDHs) have been tested as potential fire
retardants that could improve the flame retardancy, while
improving also mechanical properties (like tensile strength,
elongation at break, etc.) [16–18]. In layered material/polymer frameworks, simultaneous upgrades of numerous properties are ordinarily accomplished, for example,
combustibility and biodegradability [19,20]. These enhancements of polymer-nanocomposite properties strongly
depend on the dispersion of added substances within the
polymer matrix.
1.2. Polymer combustion process
Numerous polymers are subjected to some appropriate ignition sources, will experience self-managed ignition in air or
oxygen [21]. All in all, non-polymeric materials (e.g.,
matches, cigarettes, lights, or electric circular segments) are
all the basic source of ignition, but polymers themselves are
most often in charge of the propagation of fire. A consuming
polymer constitutes an exceedingly complex ignition system.
Compound responses may occur in three associated regions:
inside the condense stage, at an interface between the consolidated stage and gas stage, and in the gas stage.
The chemistry of fire retardancy of polymers depends on
its chemical structural repeat units and their interaction
when a polymer is burnt. It has been mentioned that the
highly aliphatic polymers that is having sp3 carbon bonds in
their backbone structure are more susceptible to high heat
release [22]. On the contrary, the polymer with more aromatic in nature that is having sp2 carbon bonding exhibits
low level of heat release. The bonding interaction is high in
case of sp2 carbon bonding, and hence in required more heat
energy to break it. As a result, polymers with sp2 carbon
[email protected];
[email protected]
© 2020 Informa UK Limited, trading as Taylor & Francis Group
KEYWORDS
Department of Chemistry, College of Science, King Saud University,
R. GIRI ET AL.
2
Figure 1. Schematic diagram of polymer burning (reproduced from Ref [24].).
bonding or aromatic ring will be more fire retardant compare to polymers is having sp3 carbon bonding in their backbone structure.
The polymer burning happens like a series of interrelated
occasions [23]:
(1)
(2)
(3)
(4)
Heating of the polymer
Decomposition
Ignition
Combustion
Initially, the polymer is warmed to a temperature at where it
starts to disintegrate and produce vaporous items that are
typically combustible. These items at that point diffuse into
flame zone over the consuming polymer. At the time when
there is an ignition source, they will experience burning in
the gas stage and release heat. Under the intense and increasing state of consuming conditions, a part of the heat is
exchanged with the polymer surface, creating more unstable
polymer sections to maintain the heating cycle. This process
is described in Figure 1 [24].
There are two types of ignition included when polymers
are burned: flaming combustion and non-flaming combustion [25]. Flames are self-propagating responses which
involve both the fuel and oxidants in the gas stage. Most
polymers are based on hydrocarbons; hence, the flame above
the surface of consuming polymers is generally a flame containing hydrocarbon. The vital responses in flames are freeradical responses. The most essential radicals present in
hydrocarbon flames are basic species, for example H•, O•,
and OH•, and a little measure of HO2•, HCO•, and CH3•.
Chain-stretching responses in the ignition procedure, for
instance H• + O2 → HO• + O• can quicken the consumption
of polymers by producing more radicals. The smoke which is
formed in flames depends largely on the gaseous fuel’s structure and on the fuel to oxidant ratio. Regularly, polymers
containing simply aliphatic auxiliary units deliver moderately little smoke, while polymers containing aliphatic structural units in the fundamental chain deliver intermediate
amount of smoke. Non-flaming combustion, includes
smouldering and glowing ignition, propagates through the
polymer with help of warm front or wave, including the
surface oxidation of the pyrolysis items [23]. Shiny burning
contrasts form seething ignition; it is joined by pale frames of
carbon consuming to form carbon monoxide. Seething ignition, as a rule, happens along with polymeric materials of
high surface zone that can shape a carbonaceous scorch. It is
joined by the era of smoke because of pyrolysis at the surface
or close to it. Glowing burning happens after the underlying
roasting of the polymeric material. From a handy perspective, it is additionally vital to take into consideration the
related fire risks. The impacts coming about because of
polymer burning, which can debilitate human life, incorporate oxygen exhaustion, flame, heat, smoke, hot and lethal
burning gasses. The two major causes of fire-related deaths
are inhalation of toxic gases and burns [26,27].
1.3. Fire-retardant chemistry
When a polymer degrades and a thermal decomposition
occurs, the following types of products may be formed [23]:
(1) Combustible gases, e.g. carbon monoxide, methane
and ethane;
(2) Liquids which are usually the polymers are partially
degraded;
(3) Non-Combustible gases, e.g., carbon dioxide, hydrogen bromide and hydrogen chloride;
(4) Finely divided solid particles which consist of decomposing polymer fragments or soot in the combustion
gases;
(5) Discrete solids in the form of char or carbonaceous
residue.
The advancement of extensive volumes of exceedingly
combustible gasses will no doubt tend to build the combustibility of polymers. Fluid items are not as precisely flammable
as vaporous items; however, fluids can spread heat to adjoining parts of the polymeric structure. Along these lines, diminishing the sum and rate of production of burnable gasses and
avoiding the spreading of flame are the most essential procedures to accomplish flame retardancy. For the most part,
there are two instruments by which the polymer burning
can be repressed [26]. One is strong phase inhibition, including changes in polymer structure. Low thickness, highporosity chars have a tendency to be the most attractive
deterioration items from the non-combustibility. The chars
can have the valuable capacities including saving basic structural integrity, protecting the basic polymer from the heat of
the flame and preventing creation of new fuel and further
consuming. However, a char once in a while experiences
seething ignition, prompting a large amount of smoke
MATERIALS RESEARCH INNOVATIONS
formation. This impact typically happens just at elevated
temperatures. Another imperative factor in the polymer ignition cycle is the net heat of combustion of the gaseous items,
which is the heat discharged by the consolidated burning
responses reduce the heat expected to bring the polymer
from its underlying state to the ignition. In this way, if
a framework can evolve water and some other noncombustible gasses or have vast endothermic requests for
thermal disintegration, the flammability can be decreased.
Another flame-retardant instrument is vapour phase inhibition, including changes in chemistry of flame. In these frameworks, reactive species, for example, bromine and chlorine are
incorporated with the polymer and are changed into freeradical inhibitors that are volatile while burning, for example,
hydrogen chloride (HCl). These materials diffuse into the
flame and inhibit the radical responses by changing highly
reactive species H• and HO• into less reactive Br• and Cl•.
Two cases of radical scavenging responses are H• + HCl →
2H + Cl• and HO• + HCl → 2HO + Cl•. Thus, the combustion cycle is interfered. Some materials include both solid and
vapour phase inhibition. It is practically difficult to recognise
among the distinctive components by which the flaming
combustion of a natural polymer is being inhibited.
1.4. Inhibition of polymer combustion
The best technique by which we can prevent polymer combustion is to make inherently fire-safe polymers. Such polymers should have high thermal stability, resistance to flame
spreading and low consuming rate event if the heat flux is
high [28]. Nonetheless, these materials are difficult to process
and are exceptionally costly. Another methodology is to
utilise flame-retardant additive to hinder the ignition of
polymers, particularly for the commodity polymers. Below
are the details about these two choices.
1.4.1. Intrinsically fire-resistant polymers
Maximum polymers which have high thermal resistance are
naturally fire safe. They have elevated decomposition temperature and hence, their initial breakdown will be viably
anticipated and process of combustion will not be started.
This property of polymers with high-temperature stability
can be improved by expanding the associations between
chains of polymer or by stiffening of chains [29].
Interactions between the chains can be enhanced by a few
means, for example, increasing the crystallinity, the presence
of polar group and hydrogen bonding. Stiffening of chains
can be achieved modifying the polymer backbone by addition of aromatic or hetrocyclic groups, for example, in poly
(p-phenylene), aromatic polyamides, and polyesters.
Moreover, polymers containing extensive quantities of
Figure 2. Intrinsically fire-resistant polymers.
3
aromatic groups in the basic units have incredible tendency
of producing char when heated. They, consequently, deliver
less combustible gaseous products into flame. Taking all
things together, polymers having high thermal stability and
producing less combustible volatiles are the most wanted
fire-safe polymers.
For intrinsically fire resistant polymers there are three
types of structures: linear single-strand polymers comprising
of a succession of heterocyclic or cyclic aromatic structures,
inorganic polymers, semi-organic polymers and ladder polymers [29]. Up until this point, most fire-resistant polymers
based on carbon are prepared by fusing very rigid, stable,
aromatic or heterocyclic ring structure specifically in the
polymer chain [30–32], for example, polyimide (PI), polybenzoxazole (PBO), polybenzimidazole (PBI), and polybenzthiazoles (PBZT) (Figure 2). The synthetic ways make
such polyaromatic heterocyclic polymers include a two-stage
process in which a high molecular weight, soluble prepolymers are synthesised and after that stable, rigid rings are
formed by chemically or thermally induced build-up of reactive groups on polymer chain.
Ladder polymers are an exceptionally extraordinary kind of
rigid chain polymers [33,34]. Such polymers are doublestranded structures comprising of the two polymer chains
which are intermittently bonded together by chemical bonds,
for example, cyclised polyacrylonitrile. On a fundamental
level, these materials should show higher thermal stability as
of the fact that the polymer chains cannot be disjoined by
breaking a single bond. The synthesis of semi-organic and
inorganic polymers has aimed for the generation of polymers
which have linear chains and stable structures and comprising
of repeat units like boron-nitrogen, phosphorus-nitrogen and
silicon-nitrogen [35,36]. The non-burning characteristics of
inorganic components and the preparation of noncombustible protective coatings are the two principle explanations behind the fire resistance of polymers.
1.4.2. Flame-retardant additive
From the manufacturing perspective, the introduction of flameretardant additives without a doubt constitutes the simplest
method for making polymer less flammable. There are two
sorts of additives: the reactive flame retardants and additive
flame retardants [37]. The reactive flame retardants are the
compounds which contain heteroatoms giving some level of
flame retardance; further, they are incorporated chemically in
the polymer chains. On the other hand, the additive flame
retardants can be physically blended with given polymer. For
this situation, there is no chemical reaction between compound
and polymer. Most abundantly utilised flame retardants at the
present time are based on six components: bromine, chlorine,
aluminium, phosphorous, boron and antimony. Also, silicon
4
R. GIRI ET AL.
and nitrogen can likewise give some degree of flame retardance.
Other components and their compounds have turned out to be
comparatively less effective. Compounds of flame retardants
regularly have synergistic or hostile impacts. Once in a while
a heteroatom already present in polymer backbone may interface with flame retardant, along these lines, display synergism or
hostility. Although additive flame retardants are broadly utilised
as a part of polymers, there are a few restrictions, for example,
poor compatibility, high volatility, and harmful consequences
for the properties of polymers, an increase in generation of
smoke and carbon monoxide (CO) [38]. Flame-retardant additives can work by a different mechanism in either gas phase or
condensed phase [37]. They can end the free-radical reactions in
the condensed phase and act as heat sinks because of their heat
capacity, form a non-combustible protective layer or char to
protect the combustible polymer from the source of the oxidant
and heat and intrude on the combustion of flame in gas phase. It
is troublesome, to unequivocally attribute a single mode of
action to a particular additive or class of additives. Number of
flame retardants gives off an impression of being equipped for
working simultaneously by a few distinct components, frequently relying upon the idea of the natural polymers.
1.4.2.1. Flame retardants based on inorganic hydroxide
(hydrated additives). Inorganic hydroxides are considered
to be an important class of flame retardants as they have low
cost, low harmfulness and ease of handling [39,40].
Aluminium oxide trihydrate is utilised in the biggest
amounts in polymers as an organic flame retardant [41]. In
order to acquire huge flame-retardant impact, these are
added into the polymers in large amount (50% by weight).
The addition of flame retardants reduces the measure of
flammable materials accessible for disintegration. During
decay, this filler compound goes about as a heat sink and
along these lines, delays the polymer from achieving its
degradation temperature [42]. Whenever heat is supplied to
it, it decomposes to form anhydrous alumina and discharges
water, which is an endothermic response. The heat from the
substrate is removed by utilisation of this energy and thus
keeps the substrate below its ignition temperature.
Additionally, the concentration of combustible gas is diluted
by the water discharged into the vapour phase. The oxide
deposit produced during decomposition has heat capacity,
which can lessen the heat transfer to the substrates. Another
advantage of utilising inorganic hydroxides is that they can
decrease the measure of smoke created on combustion [43].
Because of its low thermal stability, aluminium oxide trihydrate ought to be utilised beneath 200°C. Other inorganic
hydroxides and hydroxycarbonates [44] likewise have some
flame-retardant activity. For instance, magnesium hydroxide
is more thermally steady and can be utilised over 300°C.
1.4.2.2. Flame retardants based on halogens. Halogencontaining flame retardants contributes to additives in the
plastic industry at large scale. Just like reactive flame retardants, other groups containing halogens like alkenes,
cycloalkenes and styrene can also be copolymerised with
non-halogenated monomers [45]. The organic halogenated
compounds are additive flame retardants and are usually
used with metal oxides like Antimony Oxide or with phosphorous compounds. The stability of halogen compound
decreases from fluorine to iodine as F > Cl > Br > I. Iodine
compounds are not steady enough and thus we cannot use
them at industrial level, but fluorine compounds are highly
stable and hence can be useful. Bromine and chlorine are the
most widely used halogen-containing flame retardants.
Bromine compounds are more effective than chlorine compounds on a weight basis; however, they are impressively
more costly. Halogen-containing flame retardants may work
either in the vapour phase or in the condensed phase [45].
The structure of the additive as well as of the polymers
influences the mechanism of flame retardant. For the most
part, the radicals delivered by thermal degradation of
a halogenated flame retardant can interact with the polymer
to form hydrogen halide. Radical propagation reaction is
inhibited by hydrogen halides. This reaction occurs in
flame in response to the dynamic radicals like H• and OH•.
It also should be noticed that aromatic brominated compounds can create vast amount of char. In spite of the fact
that halogen compounds are generally utilised alone in flame
retardants, free-radical initiators and antimony trioxide
increase its effectiveness. Halogen systems containing antimony can influence the combustion of polymers as they have
the capacity to act in both phases i.e. gas phase and condensed phase. Despite increasing number of laws, brominated and chlorinated flame retardants have occupied the
largest share of flame-retardant market.
1.4.2.3. Flame retardants based on phosphorous. Both
inorganic and natural phosphorous compounds are valuable
for granting flame retardant to numerous polymers.
Phosphorous flame retardants incorporate essential red
phosphorus, water solvent, insoluble ammonium phosphate,
insoluble organophosphates and phosphonates, inorganic
phosphates, phosphine oxides and chloroaliphatic and bromoaromatic phosphates [46]. Both additive and reactive
flame retardants are commercially available. Additive compounds, particularly phosphates, are generally utilised for
much hydroxylated polymers, for example, cellulose. The
most widely utilised reactive flame retardants are phosphorus containing polyols which are used in polyurethane
foam [47]. Other examples of reactive flame retardants are
the one incorporated with vinyl and allyl phosphonates [48].
The flame retardance systems of these phosphorus compounds include free radical restrain for most part proposed
for halogen, formation of surface glass which protects the
substrate from flame and oxygen and advancing of char [45].
The flame-retardant system for phosphorus relies upon the
structure of polymer compound and the type of phosphorous
compound. As there is development of phosphorus halides
or oxyhalides or P-N bonds the phosphorus flame retardants
containing nitrogen or halogen are regularly expressed to
display synergistic conduct on decay [49,50].
Recently, the interest for intumescent systems has
increased, which can build up a char on the surface of
polymer while burning. The compound of melamine, dipentaerythritol and ammonium phosphate is the most ordinarily
utilised intumescent flame-retardant framework [51]. For
the most part, intumescences require a char former for
example, an acid source, a polyol or a catalyst, a phosphate
and a spumific (gas generator) like nitrogen source. The
mechanism includes decomposition of phosphate to phosphoric acid, esterification of polyol followed by decomposition and recovery of phosphoric acid. The forming char is
blown to porous item with help of ammonium polyphosphate with high alkali content. This surface char protects
MATERIALS RESEARCH INNOVATIONS
the substrate from heat, oxygen and flame. There are also
some self-intumescing compounds that contain every one of
the three required functions in single molecule (Figure 3).
Such intumescent coatings can satisfy ecological and toxicity
quality issues as the coatings are halogen free and the decomposition gasses are ammonia and water.
1.4.2.4. Flame retardants based on nitrogen. The presence
of nitrogen in regular polymers seems to impart some
degree of flame retardance, as appeared by the generally
low combustibility of fleece, leather and silk [52]. Synthetic
polymers containing nitrogen are not all that impervious to
combustion. Various nitrogen-containing natural compounds are utilised in certain polymers as receptive flame
retardants, for example, urea, guanidine, cyanuric acid derivatives, etc., [53,54]. Some of these compounds are also
utilised as additive flame retardants, frequently with phosphorus compounds, to decrease the combustibility of cellulosic materials. In the later cases, the nitrogen seems to act
to a considerable extent by strengthening the attachment of
phosphorus to the polymer, yet nothing is certain about the
mechanisms of activity. One conceivable clarification is that
the release of nitrogen or alkali dilutes the unstable polymer
decomposition products making them less combustible.
Metal amine complexes and ammonium salts are also
broadly utilised as flame retardants for many applications,
for example, ammonium phosphates for wood [55,56].
1.4.2.5. Flame retardants based on silicon. Use of siliconbased flame retardants instead of halogens or phosphorus in
Figure 3. Intumescent compounds (reproduced from Ref [38].).
5
gaining interest [57]. Silicones like silicas, organosilanes,
silsequioxanes and silicates are all been studied.
The flame retardant which is recognised most widely in
view of silicon is as polyorganosiloxane, in specific, polydimethylsiloxane (PDMS). A significant decrease has been
observed in flammability of block copolymer of different
types of polycarbonate (PC) and polyetherimide with PDMS
[58]. Silicon can likewise be fused into the polymer chain’s
branches [59]. Under specific cases, the expansion of silica can
likewise influence the combustibility properties of materials
[60]. The arrangement of a protective surface layer which is
based on silicon seems, by all accounts to be the flameretardant mechanism for silica and silicone systems.
Polycarbosilane, polysilastyrene and polysilsesquioxane preceramic polymers, appeared in (Figure 4), are likewise used to
mix with different thermoplastics [61,62]. These studies
demonstrate that they are all flame retardants with high
efficiencies. They can reduce the peak of heat release rate
(HRR) and normal HRR; however, the total amount of discharged stays unaltered. The essential explanation behind the
low HRR for the blend is the decrease is mass loss rate. Mass
loss rate is the rate at which fuel discharged into the gas phase
is reduced because of the presence of ceramic charge. Recently
there has been great increase in the combustibility properties
of polymer-earth (layered-silicate) nanocomposites [63].
Cone-calorimetry information demonstrate that both the
peak and the normal HRR are reduced fundamentally for
delaminated and intercalated nanocomposites having low
silicate mass (3%-5% by weight), yet there is small change in
the yield of char. Polymer-clay nanocomposites are the materials that may satisfy the necessities for a superior, additive-
6
R. GIRI ET AL.
Figure 4. Some preceramic polymers.
type flame-retardant system. In general, a condense phase
mechanism, including a protective surface layer is proposed
for silicon-based flame retardants.
1.4.2.6. Flame retardants based on boron. Borate treatments were the only one to be broadly connected to cotton
and afterwards to wool initially. Boric acid and borax are as
often as possible utilised together [64]. On evaporation of the
water of hydration, the polymers swell and an intumescent
covering is formed, which protects the main part of the
polymer from the heat source. The increase in the development of char, the endothermic dehydration process and the
dilution of the gaseous breakdown products by the water
discharged could be the explanations behind the flame retardancy of additives containing boron. For cellulose cyclic
borate is used as durable additives whereas boric acid and
polyols have been fused into rigid polyurethane foams.
1.4.2.7. Flame retardants based on polymers. Polymeric
flame retardants have many advantages but are comparatively less studied than their small molecular counterparts.
The physical and mechanical properties of polymers are less
influenced by addition of polymeric flame retardants. It can
also avoid a strategic distance from the outward dissemination in the system and consequent risk of environmental
contamination. Examples of such flame retardants are polydibromostyrene and polyphosphazenes [3,65]. In a broad
sense, all the fire-resistant polymers can be utilised as polymeric flame retardants to be mixed with some different
polymers to upgrade fire retardancy. As a matter of certainty,
this is a highly advantageous approach to modify polymer
combustibility by organisation.
2. Flame retardants based on polymer/layered
silicate nanocomposites
Ordinary flame retardants, as per their nature interfere with
the combustion cycle through chemical or physical methods
of activity. The principal mechanisms of a flame retardant to
reduce combustion in composites are:
(i) The quantity of heat transferred to polymer is
reduced by arrangement of a protective impervious
coating, which limits the diffusion of oxygen to the
area of decomposition and impedes,
(ii) (ii) The escape of volatiles and inert gasses, at the
surface of the consuming polymer which dilute the
oxygen supply, and
(iii) Advancement of endothermic reactions which cool
the substrate to a temperature beneath that required
to propagate consuming.
Also, the chemical methods of action by which flame
retardants work are:
(i) carbonaceous layer on the surface of polymer by
advancing low energy solid state reactions which
leads to the carbonisation of the polymer to the
expense of volatiles generation,
(ii) Speeding up of degradation of polymer, causing pronounced dripping and hence, the withdrawal of the
fuel from the flame source,
(iii) Prevention of the oxidation reactions occurring in
the gas phase by trapping of free-radical species
(particularly, H• and OH•) created during polymer’s
decomposition.
MATERIALS RESEARCH INNOVATIONS
The chain-branching reaction is caused by the hydrogen radical which propagates the combustion of fuel (H•
+O2→OH• +O•) while the hydroxyl radical is included
with the most exothermic reaction (OH• +CO→H•
+CO2) which gives maximum energy preserving the
flame. Now, it should be noticed that more often than
not these mechanisms do not happen independently however in combination and along these lines, the path
according to which a flame retardant works is normally
a complicated procedure comprising of single stages with
one of them dominating. The determination of an appropriate additive for every step by expecting its mechanism,
relies upon factors like its tendency to migrate, stability,
toxicity, impact on electrical properties, ability to be
coloured, cost-effectiveness and capacity to cause corrosion [66–69].
Progressively restrictive legislation regarding the toxic
release of halogen-containing materials has made two
categories of flame retardants: halogenated and nonhalogenated. But this characterisation is excessively unsophisticated for an advanced analysis over the addition of
conventional flame retardants into polymer nanocomposite. Consequently, inside the system of this study, the
category of halogen-free compounds was broken down
into littler classes, mostly used in literatures: nitrogenbased compounds, metal hydroxides, intumescent systems
and phosphorus flame retardants. In spite of the fact that
intumescent systems may contain phosphorous or nitrogen substances, they work uniquely in contrast to their
individual segments, hence they are studied independently.
All together to justify the more practical approaches of
this survey, in every given classifications, the flameretardant nanocomposites inspected are arranged as per
the type of matrix of polymer. Components that do not
go under the previously mentioned gatherings of flame
retardants were incorporated into a specific segment. The
study is finished displaying detailed cases of the nanocomposites which contain layered silicates alongside new type
of nanoparticles.
2.1. Nanocomposites with halogenated compounds
The combustion cycle is interfered by halogen-based flame
retardants principally through the mechanism of free-radical
scavenging which restrains the propagation of flame in the
vapour phase. More in particular, they respond by the degradation radicals of polymers (HX + H• →X• + H2 and HX +
OH• →X• + H2O, where X is a halogen) resulting in substantially less active species. Common examples include bromine, chlorine, iodine and fluorine containing compounds,
7
with the initial two containing in the terms of tonnage, the
most generally utilised individuals from this group [69,70].
2.1.1. Polyolefin/layered silicate-based nanocomposites
The primary-halogenated flame retardant to be checked,
going for the reduced nanocomposite ignitability was
decabromodiphenyl oxide (DB) added with antimony trioxide (AO). Metal oxides like AO are incorporated in the
compounds that are halogenated as the interaction
between them enhances the flame retardancy by generating halides (e.g. antimony trihalide); halides involve
strong Lewis acid catalyst capable for promotion of dehydrogenation charring reactions, work like radical scavengers and form a blanket which acts like a barrier between
the flue gasses and the condensed phase [71,72]. Zanetti
et al. [73] consolidated the DB-AO system in poly (propylene – maleic anhydride) (PP-g-MA) nanocomposites
and calculated its productivity by testing with cone
calorimetry experiments. The TEM micrographs, XRD
patters and tensile properties (100% increase of modulus)
of the prepared samples advocate the nanodispersion of
clay in the matrix of polymer. The flame retarded polymer (PP-g-MA + 22% by wt DB + 6% by wt AO)
displayed HRR peaks having bimodal shape (major peak
at 170 s with a weak shoulder at 85 s); while the significant peak vanished when 5 wt % organically treated
clay was added, proposing an improved flammability
behaviour which is more uniform. Also, when the base
polymer was compared with the non-FR nanocomposite,
a huge reduction in peak heat release rate (PHRR) (by
60% and 33% separately) and in the normal HRR (by
62% and 66%, separately) was recorded, while the ignition time was expanded. Given the labyrinth impact of
layered silicates, the collaboration with the flameretardant system was because of the postponed release
of the gas-phase species that were actively developed
during combustion (Figure 5), for example, HBr created
by DB, SbBr3 created by DB-AO or by the reaction
amongst NaCl and AO (show in clay as pollution) and
the SbBr3 – RNH3Br created by decomposition of organic
modifier [74]. Without the clay, these species were
quickly liberated and consumed towards the start of fire,
as confirmed by the two-peak curve of the flame retarded
polymer (Figure 6). Flame retarded PP/layered silicate
nanocomposites, consisting of AO, have been synthesised
by Lee et al. [75] in a twin-screw extruder. Because of the
poor compatibility between the organoclay and the
matrix, PP-g-MA was utilised as a compatibilizer to aid
the formation of intercalated/exfoliated morphologies. It
was illustrated; indeed, that a more successful reduction
Figure 5. Evolution of active gas-phase species during the combustion of polymer nanocomposite containing DB and AO (reproduced from Ref [73].).
8
R. GIRI ET AL.
Figure 6. HRR plots of PP-g-MA and PP-g-MA nanocomposite containing DB and AO (heat flux = 35 kW/m2) (reproduced from Ref [73].).
in PHRR can be given by the co-expansion of the ingredients given.
2.1.2. Polyamide/layered silicate-based nanocomposites
Hu et al have studied the polyamide 6 based selfextinguishing fire-retardant nanocomposites [76] using
5 wt% of organically modified MMT (OMMT), which was
treated with a hexadecyl-trimethyl-ammonium salt. They
found partially exfoliated/intercalated nanocomposites. The
amount of flame-retardant system (15 wt% DB + 5 wt% AO)
was somewhat lesser than that utilisation in the investigation
of Zanetti et al. [73], yet, sufficient for accomplishing a V-0
order. The PHRR of the PA6 matrix reduced from 1220 to
673kW/m2 when clay filler was present and even lower to
390kW/m2 when the flame retardant was added confirming
the synergistic action of the additives.
2.1.3. Acrylonitrile-butadiene-styrene/layered
silicate-based nanocomposites
The DB-AO system was utilised well by Hu and his associates
[77], with a defined aim to prepare flame-retardant system
consisting of ABS/OMMT nanocomposites, which could
pass the thorough UL94 test. Accordingly, a similar sort of
clay (5 wt%) was dispersed in ABS along with 15 wt% DB and
3wt% AO utilising a two roll mill process and the fire properties of the subsequent blend were completely contemplated.
In opposition to the neat ABS and the pertinent nanocomposite, the sample containing both the clay filler and the
flame-retardant system managed to get a V-0 rating, displaying additionally a LOI value (27.5 vol. %) rose by around 50%
and 28%, respectively. Moreover, the performance in cone
calorimeter was incredibly enhanced as probed by the 78%
lower PHRR of the flame-retardant PLSN (polymer/layered
silicate nanocomposite) in respect to that of pure ABS. The
challenge to upgrade the thermal stability of ABS nanocomposites along with their resistance to ignite was confronted
by Ma et al. [78], replacing DP with a brominated epoxy tar
(BER). BER [72] (Figure 7) is considered to be high molecular weight gas-phase flame retardant with 53 wt% bromine
content, composed by ICL Industrial Products (beer-Sheva,
Israel) and utilised usually in PC/ABS blends with synergist
like AO. In this study BER particles had more affinity due to
their high polarity than ABS towards clay particles,
Figure 7. Chemical structure of BER (reproduced from Ref [78].).
encouraging the formation exfoliated and in this way more
thermally stable structure. However, the immense preferred
standpoint of the connected way was that, with little measure
of the halogenated compound (12 wt% BER+4wt%AO), the
LOI of ABS containing 2wt% clay raised from 20.5 to 31.4
vol.%, which is a long way past 24 vol.%, the LOI value
normally required for a material to acquire a V-0 rating
[79]. The combustion procedure and the synergistic impact
of clay and BER–AO are introduced schematically in (Figure
8). In addition to the OMMT–DB–AO system, the cooperative energy between OMMT–BER–AO got from silicates
forming barriers that stopped BER pyrolysis and the reactions between BER–AO occurring at reduced temperatures.
Subsequently, in vapour phase, constant flame retardancy
could be achieved all through combustion. In addition, it
was recommended that the alkylammonium cations dwelling
in the interlayer degraded near 200°C, to parts which could
expand clay layers and volatilise promoting silicates dispersion. Exfoliated structures, showing enhanced barrier properties over their intercalated analogues, could postpone all
the more effectively heat and mass transfer. Inversely, the
reaction between the surfactant’s decayed items and DB–AO
taking place at elevated temperatures that brought about the
development of radical scavengers.
2.1.4. Polystyrene/layered silicate-based nanocomposites
Different cases of organically modified clay nanocomposites
mixed with the halogenated compounds were studied in the
article of Wang et al. [80]. The creators melt blended PP, PS
and PE with 3 wt% of clay (Cloisite 30B, Southern Clay
Products, and USA), while acrylic acid pentabromobenzyl
ester (ACPB), methacrylate acid pentabromobenzyl ester
(MEPB), butyric acid pentabromobenzyl ester (BUPB) or
MATERIALS RESEARCH INNOVATIONS
9
Figure 8. Schematic demonstration of the synergistic effect between clay and BER-AO incorporated in ABS (reproduced from Ref [78].).
pentabromobenzyl ester polyacrylate (PBPA) and were
incorporated as the flame retardant. Also, mass polymerisation was connected to plan flame retarded styrene nanocomposites and nanocomposites of styrene and dibromostyrene
copolymer. For this situation, two optional clays were additionally utilised, precisely a fluorine containing ammonium
and a dimethyl-n-hexadecyl-4- vinyl benzyl ammonium
altered MMT. The two ways used for the sample preparation
were ended up being effective for the exfoliation/intercalation of the clay as it was checked through XRD characterisations. Using cone calorimetry, it was brought up that the
flame retardancy of PP and PS nanocomposites, with limited
concentration of bromine (6 wt %), is increased discernibly.
The best execution was displayed by styrene nanocomposites mass polymerised when dibromostyrene was present;
the PHRR reduced essentially and a rating of V-0 was
achieved with majority of the blends prepared. Whereas,
the fire properties of PE-based specimens were not
enhanced, inferring that interaction between the clay, polymer and bromine compound, experienced in PS and PP
formulations, is absent for this situation. An alternate route
increasing the THR and TTI of Polystyrene nanocomposite,
rather than the regular polymerclay-flame retardant blending, was studied by Chigwada et al. [81]. The creators at first
delivered novel organically/modified clays utilising ammonium salt which containing an oligomeric unit consisting of
vinyl benzyl chloride, dibromostyrene and styrene.
Therefore, nanocomposites were prepared both by means
of melt intercalation and in situ polymerisation at different
clay ratios and consequently, bromine. XRD estimations
joined by intensive examination of TEM photomicrographs
at various magnifications gave a decent idea for the development of intercalated structures. The materials which were
polymerised in situ were comparatively more thermally
stable, presented inferior THR, lower PHRR and sometimes
preserved TTI, compared to the virgin polymer. The drop in
PHRR, usually found in nanocomposites, was because of the
presence of silicates and not bromine, which, on the other
hand, was in charge of the THR drop. It is worth noticing
that these improvements were recorded with less than 4 wt%
10
R. GIRI ET AL.
Table 1. Fire properties of PMMA nanocomposite containing DB and AO
(reproduced from Ref [82]).
OMMT
Sample (wt%)
1
2
3
5
DB
(wt
%)
20
20
AO
LOI pHRR
(wt UL94 (vol. (kW/
%) rating %)
m2)
- Failed 16.9 1456.8
5 Failed 23.3 490.4
5
V-0 25.6 359.4
Average
HRR(kW/
m2)
598.5
306.9
235.9
Average
MLR(g/s)
0.138
0.177
0.125
bromine, an amount smaller than that typically utilised for
fire-resistant compositions. Then again, the outcomes got
with dissolve intercalated nanocomposites showed up not
to be so idealistic.
2.1.5. Polyester/layered silicate-based nanocomposites
As the combined impact of the DB–AO system and clay was
uncovered not to happen only in particular polymer matrix,
Si et al. [82] recently investigated the likelihood of planning
to make a PMMA nanocomposites which is less flammable,
using the same idea; the authors melt blended a commercial
organoclay (Cloisite 20A, Southern Clay Products Inc., USA)
with DB – AO and the given polymer. By applying secondary
ion mass spectrometry (SIMS) and TEM, the addition of clay
preset in the matrix in exfoliated/intercalated form was
demonstrated valuable for increasing the dispersion of the
flame retardants. Also, through Dynamic Mechanical
Analysis (DMA), it was interpreted that the viscosity of
polymer is increased by silicates, which is important to
keep up its shape while burning. The fire tests showed that
the PMMA samples of all the three components (5 wt% clay,
20 wt% DB, and 5 wt% AO) showed enhanced properties
(Table 1), related to those accomplished with clay or with the
flame retardant as it were. A comprehensive examination of
specimen that was serially burned, confirmed that in solid
phase there were no interactions between the clay filler and
flame-retardant system. Also that the silicates play a triplicate
role in quenching fire; they promote development of char,
encourage the flame retardants to disperse and catalyse.
2.2. Nanocomposites based on non-halogenated
compounds
2.2.1. Nanocomposites with phosphorous-based
compounds
The phosphorus-containing substances involve, value wise
the second biggest class and presumably the fastest developing section of flame retardants because of the expanding environmental awareness of the issues emerging from
the utilisation of additives based on halogen [83].
Phosphorous compounds increase the flame retardancy
essentially by advancing the development of carbon
layer (char) on polymer’s surface, giving rise to phosphoric anhydrides and anhydrides of related acids acting as
dehydrating agents; dehydrating reactions causes the double bond formation, which results in carbonised or
Figure 9. Chemical structure of PEBI (reproduced from Ref [85].).
crosslinked structures at elevated temperatures. Volatile
phosphorous compounds (PO•, PO2•, and HPO•) can
perform well in the gas phase, displaying the properties
of free-radical trapping. The generated acids can also
form a superficial liquid or thin glassy coating on the
condensed phase, bringing down the oxygen diffusion
and also mass and heat transfer [67,72,84].
2.2.1.1. Polyolefin/layered silicate-based nanocomposites.
The flame retardancy of PP nanocomposites was studied
by Song et al with an aim to supports the concept that
the silicates and phosphorus combinedly add some flame
retardancy in polymers [85]. Because of the non-polarity
of PP lattice, the creation of nanocomposites of PP/clay
consistently needs to use a compatibilizer, for example,
PP-g-MA which offers polar groups can be intercalated
into silicates producing affinity between the filler and the
matrix. In polymer matrix, the dispersion of clay can be
additionally enhanced by high amount of shearing, for
example by using a co-intercalating monomer having
large steric volume. Considering this, the fundamental
thought of the authors was to utilise a phosphorouscontaining monomer, with the goal to achieve high
flame retardancy accomplished by the improvement in
clay dispersion. To confirm the validity of this theory,
N-imidazol-O-(bicyclo
pentaerythritol
phosphate)O-(ethyl methacrylate) phosphate (PEBI) (Figure 9) was
utilised and its impact on combustion performance and
thermal stability of melt compounded PP nanocomposites
was assessed. As expected, the interlayer spacing of silicates was increased due to addition of bulky PEBI and
therefore encouraged the intercalation of polymer chains
into galleries, upgrading the thermal properties of PP
matrix. Enhanced dispersion alongside PEBI using as an
internal plasticisation agent, the glass transition temperature (Tg) was shifted to lower values because of DMA.
The cone calorimetry (Table 2) displays that the flame
retardancy of prepared nanocomposites was increased by
phosphorous component of PEBI and it further enhances
all the fire properties of PP; even the average SEA (specific extinction area) reduced, reducing a significant
smoke suppression.
Table 2. Fire properties of PP nanocomposite containing PEBI (reproduced from Ref [85].).
Sample
1
2
3
4
PP (wt
%)
100
97
97
82
OMMT (wt
%)
3
-
PEBI-OMMT (wt
%)
3
3
PP-g-MA (wt
%)
15
TTI
(s)
30
30
32
32
pHRR (kW/
m2)
390
185
170
161
Average pHRR (kW/
m2)
205
104
92
85
Time to
pHRR(s)
85
90
94
96
MLR (g/
s)
0.035
0.029
0.028
0.025
Average SEA (m2/
kg)
555
440
400
380
MATERIALS RESEARCH INNOVATIONS
2.2.1.2. Polyamide/layered silicate-based nanocomposites.
Red phosphorus (known part RP) is a well-of this group of
flame retardants. Since the revelation in year 1965 about its
to capacity protect polymers against fire, RP has achieved lot
of popularity and its interest increased because of its advantages like thermal stability, low concentration efficiency and
low impact on polymer’s mechanical performance [86]. In
the work by Hao et al. [87] RP and an organically modified
MMT were extruded with polyamide 66 (PA 66) to make
flame-retardant nanocomposites. Two formulations were
created containing 5 wt% clay and 10 or 15 wt% RP. When
compared with virgin PA66 it was found that the LOI of first
formulation increased from 25.5 to 30.4 vol. %, while no
impact was observed in mechanical properties. The increase
in the LOI of 5 wt% clay – 15 wt% RP – PA66 was less intense
(8%); however, UL94/V-0 was accomplished, recommending
the likelihood of synergism amongst RP and clay particles at
certain loading levels. Other than the known activity of
MMT (promotes the formation of superficial coating), RP
served for flame retardancy playing its settled double role:
(1) RP decomposes to form volatile white phosphorus
during combustion, migrates to the surface, and
then it is oxidised to derivatives of phosphoric acid
which act as char-forming agents, forming a coat that
protect the substrate along with silicates.
(2) RP occurring on polymer’s surface ingests ambient
oxygen generating PO• species that trap-free radicals
like H• and HO• responsible for sustaining combustion. Here, silicates were additionally recognised to
synergise well with phosphate flame retardant (PFR)
for diminishing the ignitability of Polyamide 6.
Characteristically, the LOI of Polyamide 6 with 3 wt
% clay was 20.2 vol. %, that of the PA6 with 5 wt%
PFR was 22.5 vol. % while the LOI of the correlated
flame-retardant nanocomposite was 24.3 vol. %.
11
2.2.1.3. Polystyrene/layered silicate-based nanocomposites. The types of phosphorus compounds utilised as
flame retardants are extremely wide since this component
exists in many oxidation states. Especially, the phosphorus
compounds based on organic substances have for quite some
time been utilised to supply styrenic polymers along with fire
resistance [86]. The case of polystyrene (PS) nanocomposites
was handled by Chigwada and Wilkie [88] who created
various phosphorous-containing PS/clay nanocomposites
by bulk polymerisation process. More than 30 diverse phosphorous-based substances were utilised and assessed as
potential effective flame retardants by means of a high
throughput procedure created by the authors. In short, this
strategy included screening for flammability resistance small
amount (1.5 g) of the samples, after exposure (1 min) to
a methane flame. Samples that did not burn by any means
were viewed as promising and in this manner, its flame
retardancy was explored more altogether. Every formulation
included 15% wt of flame retardant and 3% wt of commercial
clay filler (Cloisite 10A, Southern Clay Products, USA). Out
of all specimens, lack of ignition was observed just for those
containing tricresylphosphate (TCP), trixylylphosphate
(TXP) or resorcinoldiphosphate (RDP); the effectiveness of
these compounds was then inspected through vast scale tests,
including UL94 and cone calorimetry. In cone calorimeter,
the flame-retarded nanocomposites indicated critical
decrease in the PHRR also, THR, relative to the concentration of the added phosphate (Table 3). Regardless of the
flame-retardant type, the amount of reduction in PHRR
and THR was beyond that achieved by incorporating each
component separately. Then again, the TTI corresponding
well with the TGA information varied and as reflected by
SEA, the smoke creation in all cases increased even by 100%.
To accomplish a V-0 classification in the UL 94 test, 10 wt%
of clay alongside 30 wt% of phosphate were required, negatively yet influenced the elongation of the polymer; there was
Table 3. Fire properties of PP nanocomposite containing PEBI (reproduced from Ref [85].).
Sample
PS
PS+3%Clay
15%TCP+PS
15%TCP+3%Clay
30%TCP+3%Clay
30%TCP+5%Clay
30%TCP+10%Clay
5%TCP+3%Clay
10%TCP+3%Clay
10%TCP+5%Clay
RDP5%+3%Clay
RDP5%+5%Clay
RDP15%+3%Clay
RDP30%+3%Clay
RDP15%+PS
RDP15%+3%Clay
RDP15%+5%Clay
RDP15%+10%Clay
RDP30%+PS
RDP30%+3%Clay
RDP30%+5%Clay
RDP30%+10%Clay
15%TXP+PS
15%TXP+3%Clay
15%TXP+5%Clay
15%TXP+10%Clay
30%TXP+PS
30%TXP+5%Clay
30%TXP+10%Clay
TTI (s)
62
57
59
59
43
53
55
60
49
48
67
59
68
75
63
68
74
73
77
75
55
63
64
69
58
61
57
38
59
pHRR (kW/m2) (% reduction)
1419
610 (56)
1122 (20)
495 (65)
378 (74)
342 (76)
324 (79)
704 (50)
485 (65)
508 (64)
502 (64)
458 (67)
474 (66)
358 (74)
710 (49)
474 (66)
433 (69)
424 (70)
499 (64)
358 (74)
110 (92)
307 (78)
890 (36)
390 (72)
449 (68)
475 (66)
864 (38)
313 (78)
372 (73)
THR (MJ/m2)
109.7
85.5
63.4
59.1
49.5
45.8
47.3
75.3
62.4
70.7
69.8
79.1
58.3
42.3
56.8
58.3
57.5
60.1
41.0
42.3
43.1
44.7
58.5
62.4
59.4
63.2
53.9
45.5
49.4
MLR (g/sm2)
17
14
14
14
14
14
14
15
15
14
14
14
14
14
15
14
14
14
14
14
14
14
14
12
13
13
15
13
13
Average SEA (m2/kg)
1097
1695
1560
1803
2401
2310
2285
1560
2159
1660
2057
2641
1995
2157
1551
1995
2391
1905
1852
2157
2322
1892
1443
1763
1882
1700
2122
2287
2028
12
R. GIRI ET AL.
Figure 10. Chemical structure of polystyrene-co-vinyl phenyl phosphatecovinyl phenyl ammonium salt (R: phosphate moiety) (reproduced from Ref [89].).
no change in elongation just when the substance of phosphate was 15 wt% at maximum. Proceeding with the study
on PS nanocomposites, Zheng and Wilkie [89] endeavoured
to check the interaction between phosphates and clay fillers
by adding the phosphate in the organic treatment of the clay
filler instead of incorporating it into the polymer matrix
directly. In that regard, pristine sodium MMT was ion
exchanged with an ammonium salt which contained an
oligomeric material constituting of styrene, vinyl benzyl
chloride and vinyl phosphate responding with dimethy
l hexadecylamine (Figure 10). A while later, the polymer
was blended with the modified clay by either melt mixing
or solution blending to plan PS nanocomposites, which were
found to introduce intercalated morphology and higher heat
stability. Before continuing to cone-calorimetry experiments,
thermogravimetric analysis combined with Fourier transform infrared spectroscopy (TGA/FTIR) was applied on the
materials, as a device to pick up bits of knowledge in the
flame-retardant activity. It was demonstrated that when the
thermal degradation occurred, the phosphate species were
freed from the clay surfactant and responded with the
degrading polymer to scavenge radical species in the gas
phase. Besides, the cone calorimetry information clearly
showed that the amount and impact of the imparted fire
resistance depended exceptionally on the constituents of
the phosphate, and on preparation technique, announcing
best approach to be the melt-blending technique. The
synergy between the layered silicates and the phosphate
was again significantly apparent, looking at that as
a reduction of 70% or even 80% in PHRR was accomplished
differentiated to 60% which is generally obtained when
a typical clay (without adding phosphate) is embedded in
a PS matrix. Nevertheless, the increased levels of smoke
evolved and the loss of mechanical integrity caused by the
huge quantity of phosphate (acts as a plasticiser decreasing
melting/softening point) were the other symptoms to deal
with.
2.2.1.4. Acrylonitrile–butadiene–styrene/layered silicatebased nanocomposites. The process which was applied for
imparting flame retardancy in PS/clay nanocomposite, Kim
et al. [90] intercalated triphenylphosphate (TPP) in the displays of a business naturally adjusted MMT (Cloisite 30B,
Southern Clay Products, USA) and melt mixed the clay and
ABS to subsequent plan nanocomposites. This process has an
advantage that by protecting the typically volatile TPP present in the silicates, can suppress its evaporation during melt
compounding, taking into consideration more proficient
flame retardancy and a larger scope of processing conditions.
However, if this approach enabled nanocomposite preparation was not certified in the article as no characterisation of
the materials morphology was reported. The incorporation
of clay increased the thermal stability of ABS because of the
late release of TPP but the LOI remained practically
unaltered.
A huge enhancement in LOI was accomplished when an
epoxy-novolac system was considered. Also, an improvement in thermal properties was considered in samples consisting of silane agents, which also favoured coupling
between epoxy and silicates. It is worth seeing that the LOI
of 85/9/6 wt% ABS/epoxy/(clay – TPP + silane) formulation
raised to 41.2 vol.% while the corresponding value for the
sample of 15 wt% TPP was 20.2 vol.%, only 9% over that of
the virgin polymer (18.2 vol.%). Optical micrographs of the
combustion residue uncovered that the change in flammability was related to the development of a more coherent
char, lacking of gaps and crevices (Figure 11). The synergistic
impact of this system was likewise affirmed when rather than
TPP, a tetra-2, 6-dimethylphenyl resorcinol diphosphate
(DMP – RDP) was utilised.
2.2.2. Nanocomposites with nitrogen-based compounds
Because of the increasing requirement for flame retardants
that are halogen free, the contribution of the overall industry
of compounds based on nitrogen, chiefly melamine and its
subsidiaries, has experienced a critical development over the
earlier time. This is mainly due to its ability because of which
it can operate in all stages of burning process via different
methods like endothermic reactions, free-radical scavenging,
dilution of inert gas and the fact that materials containing
flame retardants based on nitrogen can be recycled easily [67
69,]. Specifically, melamine cyanurate (MC), a salt of cyanuric acid and melamine, includes the most vital agent, broadly
utilised as a part of the nylon industry since the mid-1980s.
Over 320°C, MC separates to its parts by interfering with the
combustion cycle as a heat sink and an inert gas source. The
produced melamine sublimes by absorbing the heat generated to inert gasses which dilute the oxygen and combustible
volatiles. Then, again evolve melamine and cyanuric acid or
the depolymerisation catalysed by ammonia, upgrading melt
flow, in this manner making the polymer drip away from the
flame source [90–92].
2.2.2.1. Polyamide/layered silicate-based nanocomposites.
Hu et al. made the first attempt to investigate the influence of
MC on nanocomposites [76,93]. He prepared flame retarded
PA6 nanocomposites (of intercalated/exfoliated structure)
and characterised the IR combustion properties by cone
calorimetry experiments and UL94 tests. At first, with
a specific end goal to reveal the dominating mechanism of
MC for flame hindering polyamides without silicates, 15 wt%
MC were added to the polymer. The PHRR of PA6/MC
reduced just by 95kW/m2 in its MLR curve and was practically similar to that of virgin polymer, showing the condense
phase fire retarded activity of MC. Then again, the incorporation of wt% MMT in a 15 wt% MC-containing sample
prompted a reduction in PHRR from 925 to 515kW/m2,
because of the carbonaceous char formed on the nanocomposite’s surface, which limits the degradation of polymer
(Figure 12). Concerning UL94 test, a clear antagonism was
detected between the flame retardant and the clay filler,
MATERIALS RESEARCH INNOVATIONS
13
Figure 11. Optical micrographs of the chars formed after LOI test: (a) ABS/Clay 99/1; (b) ABS/NanoTPP 85/15; (c) ABS/(NanoTPP/Silane) 85/15; (d) ABS/Epoxy/(Nano
TPP/Silane) 85/9/6 (reproduced from Ref [90].).
Figure 12. HRR vs. time of the PA6 and PA6 nanocomposite containing MC (heat flux = 50 kW/m2) (reproduced from Ref [76].).
bringing about materials that could not qualify the V-0 order
proved unable to be accomplished when nanocomposite was
incorporated with 25 wt% MC. This behaviour originated
from the melt viscosity of nanocomposites which was higher
than that compared with unfilled polymers. The dripping
effect of MC was also nullified because of it. The barrier
properties play a small role in UL94 tests, while alterations
in the dripping characteristic becoming fundamental factor,
this property was endeavoured by Kiliaris et al. [93] and he
tried to improve the dripping capacity of PA6
nanocomposites, studying the efficiency of MC in increasing
the melt flow of polyamides due to polymerisation. Blends of
varying composition 1–5 wt% (Nanofil 9, Süd-Chemie,
Germany) and 2.5–5 wt% MC were extruded and studied
like flammability and other vital engineering properties, for
example, thermal properties, mechanical properties and
molecular weight. Nanocomposites of blended intercalated/
exfoliated morphology were synthesised, with reduced fraction of exfoliation/intercalation in the presence of MC. It was
observed that while extruding the polymer, MC promoted
14
R. GIRI ET AL.
the development of PA6 macromolecules of increased size
and complexities, excessively bulky groups, making it impossible to diffuse into the galleries of clay. Besides, it was
observed that though clay particles made a confined environment for the crystallisation of PA6, MC expanded crystalline assuming a heterogeneous nucleation part.
Mechanical tests show that both the additives emphatically affected tensile modulus but the ductility reduced.
Regarding flammability, as found out by UL94 test, it was
proved that MC was not capable at the concentrations
examined of increasing the tendency of nanocomposites
to flow; the specimens were completely burned. Another
study on the flame retardancy of PA6 nanocomposites
containing MC was done by Zhang et al. [94].
Formulations with OMMT (Cloisite 25A, Southern Clay
Products, USA) or pristine MMT were inspected, while
he also considered the effect of incorporation of Polyvinyl
Pyrrolidone (PVP) (Table 4). The highest UL94 (V-0)
rating and LOI were gotten with 13 wt% MC and kept
up to 0.2 wt% clay content, over which the viscosity
increased bringing about bigger and heavier drops, contrarily influenced flammability. Because of the incorporation of PVP (5 wt %), a partial restoration of UL94 rating
held, as there was decrease in viscosity of PVP nanocomposite melt by limiting the miscibility of polymer–clay
composite. Essentially, the not well-dispersed pristine
clay, connecting with polymer macromolecules to lower
degree did not bring fundamental changes in the dripping
characteristics of the flame retarded polymer, safeguarding the UL94 classification.
Regarding cone calorimetry, a gradual reduction in PHRR
was observed when OMMT content was increased and the
similar method was used for TTI also. The MLR curves
which were practically indistinguishable to HRR blends
showed that the activity of MC does not take place in the
gas phase, but rather is expected to the weakening of the
network and of the flame by ammonia and sublimed melamine. The advancement of melamine was ceased on the
account of nanocomposite’s viscous melts, prompting
a progressive increment of the effective heat o combustion
(EHC), when the clay was fused with MC formulations. The
previously mentioned properties appeared not to be fundamentally influenced when PVP was present, while on
account of pristine clay, the qualities acquired were like
those of PA 6–13 wt% MC. Tensile tests performed on the
flame retarded nanocomposites showed decrease in elongation and tensile strength along with a slight increase in
modulus when compared with the neat polymer.
NR, not rated; BC, burns to clamp aPristine clay
The above findings tell that the ignorable conflict between
the effect of MC (decrease in melt viscosity) and that of
silicates (increase of melt viscosity) is witnessed only when
the organically modified clay is added at small concentrations. Additional proof of this statement was brought by Hao
et al. [87], who concentrated on PA66. They blended
Polyamide 66 with 10 wt% MC and 7 wt% clay in a twinscrew extruder and compared the resultant material and the
virgin polymer on the premise of ignitability. The LOI of
flame-retardant nanocomposite as indicated by LOI test was
22.8 vol. %, recognisably lower than the virgin PA66 (25.5
vol. %).
2.2.2.2. Polyester/layered silicate-based nanocomposites.
The blend of an organophillic clay and MC was likewise
researched by Gianelli et al. [95] on account of poly (butylene
terephthalate) (PBT) and a commercial copolyester elastomer with the trade name PIBIFLEX E5601 (P-Group,
Ferrara, Italy). The clay was first ion exchanged with
dimethyl-hydrogenated tallow-benzyl-ammonium chloride
followed by melt compounding with the matrix of polymer
using a twin-screw extruder under the conditions which were
demonstrated very efficient for creating basically intercalated
nanocomposites, as dictated by TEM and XRD characterisations. The fire performance was analysed by method of cone
calorimetry for samples containing 5 wt% clay and 10 wt% of
the flame retardant, showing a critical decrease in PHRR of
both PBT (69%) and co-polyester elastomer (71%).
Reduction was additionally studied for the THR, around
15% and 11%, individually. Moving on to the ignition resistance, the TTI of PBT increased from 38 to 64 s, when the
additives were present, though it was unaltered on account of
the co-polyester elastomer. A rise in TTI was followed (from
44 s to 55 s) when 7.5 wt% clay and 20 wt% melamine
cyanurate were included, while the various properties of the
co-polyester elastomer were not really influenced (Table 4).
When compared with the samples including just one of the
two additives, the properties of quaternary formulations
were enhanced to the best, thus indicating a synergistic
action of silicates and MC in the course of the cone calorimeter. Nevertheless, flammability tests were not performed
in this examination to evaluate whether the hostility between
these two parts confirms on a case of polyamide matrices (as
depicted in Section 2.2.2.1), and is likewise met in polyesters.
2.2.3. Nanocomposites with metal hydroxides
Metal hydroxides contain the most important segment in the
market of flame retardants. They are inorganic compounds
which serve to release heat by discharging extensive amounts
Table 4. Fire properties of PA6 nanocomposite containing MC (reproduced from Ref [94].).
Sample
1
2
3
4
5
6
7
8
9
10
11
12
13
MC (wt%)
13
13
13
13
13
13
13
13
13
13
13
-
OMMT (wt%)
0.2
0.5
1
3
5
3
5
0.2a
0.5a
3a
5
PVP (wt%)
5
5
-
UL94 rating
Failed
V-0
V-0
V-2
V-2
Failed
Failed
V-2
V-2
V-0
V-2
V-2
Failed
LOI (vol.%)
25.3
32.8
29.2
21.0
21.2
21.9
22.6
21.7
21.5
32.0
31.6
30.2
22.3
TTI (s)
55
57
45
42
39
27
55
36
pHRR (kW/m2)
1629
1543
1171
905
706
676
1450
834
Average MLR (g/s)
0.11
0.094
0.095
0.089
0.074
0.083
0.097
0.077
EHC (MJ/kg)
30.2
25.3
27.2
28.0
28.9
26.7
26.2
31.2
MATERIALS RESEARCH INNOVATIONS
of water at a similar temperature go or beneath polymer’s
disintegration temperature. The produced water vapours can
also dilute combustible gasses and may incite the build-up of
an oxygen displacing protective layer. Another contribution
of metal hydroxides to flame retardancy is smoke suppression. The correct mechanism has not been yet illustrated,
however, the most possible clarification is by all accounts
that carbon, resulting from polymers’ degradation, is deposited on the oxide (produced by the decomposition of metal
hydroxide) and after that, this is volatilised as carbon dioxide, discharging no smoke [68,69,96].
2.2.4. Nanocomposites with miscellaneous compounds
Other than the previously mentioned cases, distinctive sorts
of flame retardants have been moreover joined in PLSN.
Among them, silicon-containing flame retardants have
been accounted to synergise well with clay fillers. For quite
a while, silicon compounds were known to be proficient coadditives in flame-retardant systems. Recent advances, however, have shifted a great deal of attention to their use
(especially polysiloxanes) for preparation of highperformance protective coatings. Coatings which are based
on inorganic siloxane or organic–inorganic siloxane hybrids
are inherently resistant to temperature and oxidation. They
also provide protection to polymers exposed to fire [97]. Few
years ago Quede et al. [98] demonstrated that the utilisation
of organosilicon coating can also offer a gainful path for
additionally enhancing the flame retardancy of PA6 nanocomposites. The dual advantage of this approach is that the
mechanical and physical properties of materials remain
unchanged without employing an additive-type flame retardant and secondly that the protection is focused on the part
of the material most affected by fire, i.e. the surface. The
coating of organosilicon was a homogeneous film created by
the polymerisation of 1, 1, 3, 3-tetramethyldisiloxane monomer premixed with oxygen, applying a cool remote nitrogen
plasma process. Expanding the film thickness (from 0.6 to
18.1 m) reduced nanocomposites flammability, as reflected
by the LOI values; the most maximum value (48 vol. %,
increase of 130%) was gotten for the coating of
1.5 m thickness. Despite what might be expected, the coating,
regardless of the thickness, the LOI of the neat polymer was
barely influenced, which remained practically unaltered at 22
vol. %.
The fire properties of coated nanocomposites, measured
in a cone calorimeter, did not show spectacular enhancement
but the synergic activity between the film and silicates was
still clear. The deposit of the coating on the surface of
nanocomposite brought an extra reduction of 11% in
PHRR, 5% in THR and 10% in smoke creation compared
to that offered by the incorporation of 2 wt% clay in the PA6
network (34%, 63%, and 32%, separately). By analysing the
residue, it was concluded that amid the burning of coated
nanocomposites, a carbonaceous-silicate and silica like layer
was created, the volatilisation of active fragments was
reduced by it and absorbed and dissipated heat more effectively than the char of the uncoated nanocomposite or the
polysiloxane coating alone (covered PA6). Huo et al. [99]
explored the likelihood of a silicon-based coating to present
synergism with layered silicates in the flame retardancy of
poly (ethylene terephthalate) (PET). Rather than depositing
on the surface of polymer, a highly crosslinked PBSiO containing phenyl was synthesised to coat a commercial clay
15
(Nanomer I.34TCN, Nanocor, USA), which was subsequently melt blended with PET. Over conventional polysiloxanes, PBSiO presents the merit of superior performance at
high temperatures and oxidising conditions. The decay of
PBSiO coatings leads to the development of borosilicate
glassy structures, which keep the substrate away from degradation. In combination with the organoclay, PBSiO was
found to build up a defensive borosilicate-carbonaceous
char on the surface of PET, improving significantly its performance towards fire. The addition of 5 wt% PBSiO and
2.5 wt% OMMT was adequate to get a reduction of 59% in
pHRR and of 51% in smoke yield. The cone calorimetric
information of samples containing uncoated clay unfortunately is not detailed in the paper, to completely welcome the
contribution of the two components. However, the higher
decomposition temperatures recorded for PET/PBSiO/
OMMT formulations with respect to PET/OMMT give
a sign of synergistic activity between PBSiO and clay particles. Rather than coating the silicates, Zhu et al. [100] prepared silicon methoxide-altered clay by means of treating
pristine MMT with 3[-(trimethoxysilyl) propyl] octadecyldimethyl-ammonium chloride. The clay was utilised to blend
(intercalated) PS nanocomposites of enhanced fire retardancy and thermal stability with respect to the virgin polymer. But the improvement was near to that of PS by regular
alkyl ammonium modified MMTs. Introducing layered silicates flame-retardant elements is well known to add brittleness in the polymer matrix, therefore restricting the use of
the resulting material in applications where the elongation
ability is essential.
With an aim to overcome the above drawback, Dong et al.
[101] prepared a novel compound, namely the silicone elastomeric nanoparticles (S-ENP) and used it to incorporate in
PA6 for heat resistance, good processability, high toughness
and stiffness. Other novel elastomeric flame retardant
(S-ENPC) was synthesised by spray drying S-ENP latex
mixed with the slurry of an unmodified MMT. Meltblending S-ENPC with PA6 proved that S-ENP and silicates
co-operate to further improve the flame retardancy and
mechanical properties; compared with PA6/S-ENP, the elongation at break increased by 30%, the PHRR of PA6/S-ENPC
decreased by 8%, while and the flexural modulus by 19%. The
values of mean effective heat combustion, mean CO yield
and mean CO2 yield were almost same for the neat polymer
and the composites, suggesting that the improvement in
flammability was due to changes in the condensed phase
decomposition process and not the gas-phase. A barrier of
two layers made of silica and silica/clay, respectively, was
created on the surface of PA6/S-ENPC, which provides
superior performance towards fire than the single layer
formed on PA6/S-ENP. By applying the same procedure,
Wang et al. [102] fabricated poly (vinyl chloride) (PVC)
nanocomposites containing sodium MMT and nitrile rubber
nanoscale particles. Contrary to the previous case, nanoparticles significantly deteriorated the flame retardancy of the
polymer matrix, acting as defective sites which loosened the
compact char formed superficially on PVC (typical charforming polymer) during combustion. Ammonium sulfamate (NH2SO3NH4, symbolised as AS), which belongs to
the family of sulphur-containing compounds, joint with Di
comprise another flame-retardant system whose ability to
exert synergistic effect with clay particles was examined by
Lewin et al. [103]. The system acts through the development
16
R. GIRI ET AL.
of a three-dimensional network in the condensed phase that
carbonises upon ignition and generates a carbonaceous char
by migrating to the polymer surface. Referring to their publication on PA6 with MC [94], the experiments were performed by the authors in which the PA6-AS + Di samples
were treated with pristine MMT or the organically modified
MMT (Cloisite 25A, Southern Clay Products, USA). The
introduction of 2.0 wt% AS and 0.7 wt% Di in PA6 matrix
yielded materials of 35.7 vol. % LOI and of UL-94 V-0
classification that preserved with the co-addition of 1wt%
OMMT. Increasing clay content reduced the LOI values and
UL-94 ratings, because of the catalytic effect of clay on matrix
degradation and the partial absorption of AS + Di in the clay
galleries that prevented their contribution to flame retardancy. Adding 5.0 wt% PVP moderately retained the original
values of LOI as it seemed to enter the interlayer replacing
AS + Di and thus neutralising the catalysis. In comparison
with the perfect matrix, the TTI of PA6/AS + Di were higher,
moving downwards within the sight of OMMT. Then again,
the PHRR was increased with AS + Di addition, but reduced
to a value below that of the original PA6 when 3 wt% OMMT
was added. The hostility did not appear to exist when pristine
MMT was utilised (ideal at 3.0 wt %). The poor dispersion of
pristine clay, barely influencing mechanical properties, prevented extended associations with the flame retardant and
the polymer for the advantage of fire resistance; a uniform
char without floccules was formed upon combustion leading
to decrease of 30% in PHRR values, a slight increment of
TTI, a LOI of 35.4 vol. % and a V-0 classification.
2.2.5. Nanocomposites with new types of nanoparticles
Without a doubt, the usage of layered silicates represents
right now the most utilised course of nanotechnology to
improve polymers with flame retardancy. This study mainly
focuses on natural aluminosilicates; thus only a portion of
the known structures of these nanofillers with different sorts
of nanoparticles is examined in this segment.
In the previous 10 years, rich writing work has been
composed and committed on the utilisation of carbon nanotubes (CNTs) which can be utilised as superior nanofillers.
CNTs comprises of a novel group of carbon materials which
are developed of carbon molecules orchestrated in pentagons
and hexagons (graphite structure) framing chambers. Three
primary strategies are normally utilised for their creation,
including circular segment discharge, chemical vapour testimony and laser removal, albeit most extreme research movement is spent on more economic techniques. Amid
amalgamation, contaminations as formless carbon, nontubular structures and impetus particles are likewise
acquired; in this manner the sanitisation of CNTs is required
and performed more often than not through oxidation or
acid refluxing strategies. Normally, CNTs comprise both of
one (single-wall carbon nanotubes, SWCNTs) or progressively (multi-wall carbon nanotubes, MWCNTs) concentric
round and hollow shells of graphitic sheets. Due to the
nearness of their symmetric structure, these confine like
carbon chambers gangs remarkable mechanical, thermal
and electric properties. In the handling of polymer nanocomposites, homogeneous dispersion and distribution of
nanotubes inside the polymer matrix and upgraded CNTmatrix wetting are basic issues (not effectively tended to)
with the goal that proficient fortification can be accomplished [104–106].
The subject of blending clays with CNTs has been considered in many papers, the greatest piece of them created by
Beyer. His first production [107] regarding this matter managed
the synthesis of flame-retardant nanocomposites by means of
melt-blending EVA with MWCNTs or/and a commercially
accessible OMMT (Nanofil 15, Süd-Chemie, Germany).
MWCNTs were integrated by catalytic decomposition of acetylene on Co–Fe/Al(OH)3 impetus and then by purging by
means of disintegration of the impetus bolster in bubbling
concentrated sodium hydroxide and disintegration of the
impetuses in concentrated hydrochloric acid. The organised
nanocomposites were thought about on premise of fire execution utilising cone calorimetric tests under 35kW/m2 heat flux.
The acquired results are referred to in (Table 5), certainly
illustrates diminish in PHRR by including separately both the
nanofillers (2.5 or 5 phr), more articulated for MWCNTs. Extra
reduction in the PHRR bend (Figure 13) was considered for the
specimen which contained 2.5 phr of MWCNTs and 2.5 phr of
clay, giving confirmation for the synergism between the two
portions. The two fillers responded in the consolidate phase in
the midst of combustion and framing char that was more
resistant to cracking by virtue of MWCNTs as a result of their
long aspect proportion. Concentrating again on (Figure 13),
plainly regardless of the nearness of MWCNT, the TTI diminished along with the nearness of silicates in view of the thermal
degradation of the clay surfactant occurring over the traverse of
a fire. The thermal protection, flame retardancy and pliable
properties of EVA copolymer soften blended with clay particles
and MWCNTs was contemplated in a writing of Peeterbroeck
et al. [108], in which Beyer was one of the co-creators. The
impact of crude and cleansed MWCNTs on the properties of
nanocomposites was researched.
Thermogravimetric scans with a heating slope of 20 K/
min, under air atmosphere, demonstrated an unmistakable
synergistic effect between MWCNTs (FUNDP, Belgium) and
Cloisite 30B (Southern Clay Products, USA). Most extreme
delay in degradation temperature of polymer matrix was
conveyed by the blend of 3 wt% of Cloisite 30B and 1 wt%
of refined nanotubes, which moved by 56 ◦C towards higher
temperature. Exactly when included freely, neither of the two
nanofillers caused such a change, regardless of the substance.
Filtered nanotubes were demonstrated to be more useful due
to the weakening of MWCNTs with synergist bolster by
virtue of the crude nanotubes; MWCNTs were polluted by
30 wt% alumina. Mechanical tests demonstrate that both the
nanofillers opened up stiffness with similar effectiveness
independent of the substance proportion of clay/MWCNT,
while saving the great ductility of the slick polymer. The cone
calorimetric data, which were represented in this writing
alluded to tests containing Nanofil 15 (Süd-Chemie,
Germany) as clay filler; these results were likewise presented
and deciphered in the past papers of Beyer [107]. Better
understanding about the consolidated impact of MWCNTs
Table 5. Fire properties of EVA nanocomposite containing OMMT and/or
MWCNTs (reproduced from Ref [107].).
Sample
1
2
3
4
5
6
OMMT (phr)
2.5
5
2.5
MWCNTs (phr)
2.5
5
2.5
TTI (s)
84
85
83
70
67
71
pHRR (kW/m2)
580
520
405
530
470
370
MATERIALS RESEARCH INNOVATIONS
17
Figure 13. HRR vs. time: EVA + 5.0 phr OMMT (A); EVA + 5.0 phr MWCNT (B); EVA + 2.5 phr OMMT+ 2.5 phr MWCNTs (C) (heat flux = 35 kW/m2) (reproduced from
Ref [107].).
Figure 14. Morphology of chars produced from clay/EVA, clay/MCWNT/EVA and MCWNT/EVA nanocomposites after cone calorimeter test (top row) and the natural
burning (bottom row) (reproduced from Ref [109].).
and layered silicates on the fire execution of EVA nanocomposites was attempted by Gao et al. in his investigation as
shown in Figure 14 [109].
The legitimacy of this speculation was demonstrated by
subjecting small plate examples of the three composites to
broad oxidation; all the more particularly, the specimens
were burnt in a mute furnace at 600°C for 20 min. It was
seen in the cone calorimeter that the higher reactivity char of
MWCNT-overhauled composite was burnt out while the
char shaped on the example doped with clay held its honesty.
The set up synergism amongst nanotubes and silicates in
upgrading polymer flame retardancy was furthermore
avowed by Beyer [110,111] when certifiable things were
created. A link creating extruder was utilised to plan two
insulating wires having flame retardancy based on ATHcontaining EVA/PE-clay nanocomposites, either with or
without MWCNTs. The total filler loading was kept predictable in the two cases; for the generation of the second wire,
50 wt% of the organoclay was substituted by a comparable
measure of MWCNTs. Conducting small-scale fire tests,
where the insulations were presented to a Bunsen flame
(IEC 60,332–1), showed that no dripping and equivalent
charred length was watched for the two formulations.
Likewise, the char was fortified by the guide of nanotubes.
The updated char structure achieved prevalent fire properties, as assessed by cone calorimetric examinations.
Concentrating on PHRR, which is believed to be the primary
reason for flame, it was found that the 1:1 blend of clay and
nanotubes procured diminish in PHRR (more vital than
including clay alone) to levels that allow the age of flame
retardant connect compounds. As opposed to simultaneously including the nanotubes and layered silicates in the
matrix.
Tang et al. [112] explored the possibility of enhancing the
fire protection of PLSN by means of the in situ advancement
of CNTs amid combustion. The viability of this idea was
approved by blending PP/clay nanocomposites with a nickel
impetus (nickel bolstered on silica – alumina); the combusted residue of the formulation comprised of fibre-like
charring particles intermixed with MMT layers. Likewise,
18
R. GIRI ET AL.
the measure of the residue was hoisted in correlation with
that of the specimens containing either the clay (Cloisite
15A, Southern Clay Products, USA) or the impetus.
Clearly, the catalysed carbonisation of the polymer was
expanded within the sight of clay, prompting a subjective
too quantitative improvement in the development of char. In
the limited environment made by layered silicates the
decomposition results of the polymer could not escape
effortlessly, along these lines they stayed for additional time
in contact with the impetus, experiencing dehydrogenation
and aromatisation to yield char with CNTs, since nickel is an
impetus for the arrangement of CNT. Resulting in the protective layer developed on the surface of an example doped
with 5 wt% clay + 5 wt% impetuses which was superior to
that of PP with 10 wt% clay as far as PHRR decrease. In any
case, no progressions in THR and TTI were watched, inferring that the applicable gas-phase mechanism is missing.
A few questions on how the carbon nanotubes and layered
silicates synergise to build the flame retardancy of polymers
are replied in the truly far-reaching writing of Ma et al. [113]
by learning about ABS copolymer. The three parts were melt
blended in a mixer at 190°C for 10 min at a screw speed of
60 rpm to make nanocomposites of blended intercalated/
exfoliated morphology. Without a doubt, some MWCNTs
were thought to have been embedded in the clay galleries for
the advantage of clay dispersion. The fire properties were
assessed by cone calorimetry and demonstrated that clay
went with MWCNTs diminished the PHRR and the entire
combustion procedure of ABS was deferred, more viably
than MWCNTs or clay alone. A network structure which
firmly affected flame retardancy by constituting a barrier to
oxygen, heat and flammable gasses was shaped by presentation of nanoparticles in polymer matrix. Rheological estimations demonstrated that the conjunction of nanotubes and
silicates, blocking moreover the development of macromolecules, prompted a more permeated network. The propelled
fire protection of ABS/MWCNTs/clay was additionally
related with the char shaped on its surface amid combustion,
which was thicker and denser, introducing less cracks than
that of the surface of ABS/clay or ABS/MWCNTs. TEM
pictures uncovered that some MWCNTs kept running crosswise over between silicates, inferring a solid cooperation
between them. The microstructure of char was examined
utilising XRD examination and Raman spectroscopy; it was
discovered that when the clay was embedded in the nanotubes-containing formulation, expanded the level of graphitisation of char henceforth ensured it against thermal
oxidation. The level of graphitisation mirrors the change
degree of a carbon material from turbostratic (cluttered) to
graphitic (idealise) structure. Accordingly, the confined
environment induced by silicates provoked the rearrangement of carbon crystallites, eliminating any dislocations and
defects. Likewise, Al2O3, which comprises of one of the parts
of MMT, is a famous impetus of graphitisation. In any case,
the mix of MWCNTs and clay was discovered lacking breeze
through the UL94 test for ABS.
Metal oxides like titanium oxide (TiO2) and iron oxide
(Fe2O3) are another developing group of nanoparticles with
a recognised beneficial outcome on the flame retardancy and
thermal security of PS [114] and PMMA [115]. Laachachi
et al. [116] examined whether metal oxides could likewise
contain promising synergists to organoclays, with the goal
that materials of higher fire protection could be made. So the
Table 6. Fire properties of PMMA nanocomposite containing metal oxides
(reproduced from Ref [116]).
OMMT
Sample (wt%)
1
2
10
3
5
4
5
Fe2
TiO2 O3
(wt (wt TTI
%) %) (s)
- 69
- 74
5
- 86
5 53
Total
burning
(s)
250
480
630
590
pHRR THR SEA
Char
(kW/ (MJ/ (m2/ residue
2
2
m ) m ) kg)
(%)
620 110 430
0
320 110 810
6
360 100 530
8
350 100 600
10
creators arranged PMMA – clay nanocomposites melt exacerbated with TiO2 or Fe2O3. TGA bends of the readied tests
demonstrated that the incorporation of metal oxides changed the beginning degradation temperature of PMMA – clay
nanocomposite to higher esteems. A synergistic impact
between layered silicates and oxide nanoparticles on upgrading the fire protection of PMMA was seen through cone
calorimetric tests; hoisted TTI, lessened THR and PHRR,
diminish in released smoke and impressive ascent in total
burning time were apparent at higher degree in the nanocomposites doped with oxides than in the specimens containing just a single of the two parts (Table 6).
The addition of oxide nanoparticles improved the covering of the material’s surface when the polymer ablation took
place and the decomposition of the organic modifier disorganised the deposition of clay particles on the sample’s
exterior. Better results were observed in the case of
clay–TiO2 formulations. With the high thermal stability,
TiO2 allowed its function as a heat sink, limiting the thermal
conduction in polymer bulk. Moreover, the increase in melt
viscosity of the polymer hampered the emission of volatiles.
Finally, convection forces generated during combustion
probably promoted particles migration to the surface of the
nanocomposites reinforcing the char layer.
3. Fire retardants based on nanocomposites
Recently nanoparticles have attracted many considerations
in materials science since they regularly show properties that
are quite different from those of their counterpart polymer
microcomposites containing the matrices of same inorganic
components. The nanofiller’s surface areas are definitely
expanded with the goal that polymer nanocomposites appear
large scale/small scale/nano-interfaces. Incorporation of
CNTs will not only enhance the mechanical properties but
along with that the functionalities also, for example, thermal,
flammable and electrical properties of composites. CNTs are
amongst the most representative nano-materials which are
used to give remarkable properties to polymers. Innovation
for the production of CNTs on large scale has been produced, reducing the cost of CNTs to ~$100/kg in 2013. Thus,
some CNT that are based on nanocomposites have started
showing up. For instance, Evonik Industries are creating
moulding PA12 CNT-containing compounds for fuel lines
[117]. The fundamental advantage of this material includes
avoiding the ignition caused by electrostatic charges. Fire
hazard can be considerably decreased by producing percolation systems of CNTs in polymers. Adding CNTs to polymers likewise changes their flammability. Nanofiller-based
flame retardants indicate high flame-retardant efficiencies.
Including just a little sum (i.e., 5%) of nanofiller can diminish
the pinnacle warm discharge rates (PHRRs) of polymers and
therefore lessen the speed at which flames spread all through
MATERIALS RESEARCH INNOVATIONS
them. Further, the little measure of nanofiller does not
decrease polymer processability and can enhance the
mechanical properties of polymers. Be that as it may, including no one but nanofiller cannot deliver self-extinguishing
(V-0, −1, and −2) polymers, which are required for most fireretardant items. The nanofillers ought to be joined with other
customary flame retardants to give a superior adjust of
flammability/mechanical properties.
3.1. Fire-retardant mechanism for clay-based
nanocomposites
Montmorillonite is the most commonly used clay since it is
regularly present, can be obtained at high virtue and negligible cost, and shows particularly rich intercalation chemistry, which implies that it can be easily organically adjusted.
The typical clay surface is hydrophilic, so the clay effortlessly
disperses in aqueous solutions, however not in polymers.
Common clays are routinely adjusted using regular cations,
for instance, alkylammonium and alkylphosphonium
cations, framing hydrophobic organomodified clays that
can be promptly dispersed in polymers. Clay-based nanocomposites are regularly gathered into three classifications
since clay properties are exceptional: (1) immiscible (likewise
called as microcomposites), (2) intercalated, and (3) exfoliated (additionally called delaminated). Exfoliated nanocomposites are normally wanted in light of the fact that
they demonstrate enhanced mechanical properties [118].
Clay-based nanocomposite loaded with 5% clay is starting
at now used as a commercial flame retardant by virtue of its
upgraded mechanical properties and flame retardancy. The
fire-retardancy systems for clay and carbon-based nanocomposites are for all intents and purposes indistinguishable.
One fire-retardancy component is the abatement in PHRR
on account of the advancement of a defensive surface barrier/insulation layer comprising of clay platelets collected
with a little measure of carbonaceous char [119,120]. The
clay platelets gather on the surface that the clay staying at
first glance from polymer decay and clay migration was
pushed by different rising bubbles of degradation things.
19
The surface quality appears to choose the flame-retardant
proficiency.
Another instrument proposed by Wilkie et al. is that the
paramagnetic iron in the matrix traps radicals and redesigns
thermal solidness. Actually, including only 0.1 wt% iron
containing clay diminished the polystyrene (PS) PHRR by
60% [121]. This impact was not watched for carbon-based
nanocomposites and on the grounds that their contact with
the polymer is negligible [122]. (Figure 15) indicates HRR
blends for polymer and clay-based nanocomposites [119].
Including clay can diminish the PHRR and generally
lessens the ignition time (IT); nonetheless, it cannot change
the total heat release rate (THRR). Carbon-based nanocomposites exhibit comparative results in light of the fact that the
fire-retardancy components for clay-and carbon-based
nanocomposites; i.e., the barrier/insulation impacts, are
comparable. Subsequently, the key factor in choosing clayand carbon-based-nanocomposite flame retardancy is the
advancement of a surface system layer. The barrier/insulation impact relies upon the external heat flux. Schartel et al.
investigated the relations among HRR and external heat flux,
as showed up in (Figure 16) [123]. THRR does not depend
upon external heat flux for the PP and nanocomposites. The
PP PHRR, on the other hand, increments with expanding
external heat flux, regardless of the way that the nanocomposite PHRR does not change. The fire-retardant proficiency
emphatically depends upon irradiance with the ultimate
objective that the nanocomposite fire retardancy decreases
with diminishing irradiance. The results gained through
extrapolation to small irradiances identify with flammability
circumstances, for instance, LOI and the UL 94 test. These
results clear up why including nanofillers cannot improve
LOI and UL characterisation. In addition, they suggest that
including nanofiller is successful for polymers showing high
HRRs. The flame-retardancy adequacy of clay-based nanocomposites depends upon the sort of matrix [20,62,67,120–
133]. The IT of polymers, for instance, PP, polyethylene (PE),
PS, ethylene-vinyl acidic acid determination (EVA), and
PMMA ordinarily lessens, when nanofiller is included, in
view of the way that the clay itself is possibly catalytic.
Strikingly, the IT of the PA6 nanocomposite expanded
Figure 15. HRR plots for pure PP, PP/C18, PP/Na-MMT, PP/H-MMT and PP/OMMT (reproduced from Ref [119].
20
R. GIRI ET AL.
Figure 16. PHRR and THR plotted against the external eat flux for PP (PP-g-MA-I) and nanocomposite (PPC20A-I) (reproduced from Ref [123].).
when the nanoclay was included. Besides, rate at which
PHRR diminishes depends whereupon polymer matrix is
used. PA6 and PS both lessening PHRR ~40–75% [6,
127,128]. PMMA, of course, just decreases PHRR 10–30%
[20,131,132].
Wilkie et al. exhibited that polymer nanocomposite, for
instance, PA6 and PS, which generally diminish PHRR, indicate essential intermolecular reactions and that the degradation pathway changes by consolidating nanoclay; however,
the PMMA did not show any adjustment in the polymer
degradation pathway or any significant HRR diminish
[133]. Despite the way that adding the nanoclay to the
PMMA does not basically impact PMMA flammability, adding nanosilica to high-consistency PMMA lessens PHRR by
50% in light of the fact that silica covers the entire surface
[134]. Including nanoclay into PMMA is possibly convincing
when the nanoclay covers the entire PMMA surface. The
nanoclay was less viable in improving PMMA fire retardancy
may be because of the fact that the low thickness of the
PMMA kept the nanoclay from covering the entire PMMA
surface. Adjusting the clay surfaces is the most key parameter
for upgrading the fire retardancy of clay-based nanocomposite. Microcomposites are procured instead of nanocomposite when unmodified clays are fused to polymers. The
flammability of the microcomposites is for the most part
essentially like or now and again more loathsome than
those of the neat polymers. Organomodifying clays convey
intercalated or exfoliated nanocomposites. Additionally, the
char substance and cone-calorimetry experiments of organomodified nanocomposites depend upon the organomodifier content: expanded organomodifier content prompts
a more articulated catalytic impact and more escalated char
development [124].
Figure 15 shows the effect of surface adjustment on the
HRR conduct of PP. Consolidating organomodified montmorillonite (OMMT) diminished the PHRR the most inferable from the nano-dispersed clay and the catalytic
properties of the organomodifier. The second most imperative factor in upgrading nanocomposite fire retardancy is
clay loading. Not at all like CNT loading, expanding clay
loading enhances nanocomposite fire retardancy and there is
no perfect clay loading in the range 15 wt% [67,128]. It is
difficult to frame a crack-free clay-network layer. In this
manner, the major flame-retardancy mechanism is through
the arrangement of barrier against the heat source as
opposed to gasses. Photos of deposites procured from
degraded PS/OMMT tests containing distinctive OMMT
substance are shown in Figure 17. The residues from the
degraded PS/OMMT tests containing 6-and 15-wt% OMMT
demonstrate cracks. Thicker floccules can be gained by
including more clay. The advancement of thick floccules
can basically lessen HRR. Clay-based nanocomposite flame
Figure 17. Digital photos showing the residue morphology of different PS/OMMT composites after degraded at 400 ºC for 3 h (reproduced from Ref [128].).
MATERIALS RESEARCH INNOVATIONS
retardancy could be additionally improved if polymer-clay
nanocomposites could be tuned to shape steadier crack-free
networks in the midst of consuming.
The effect of nano morphology on flame retardancy has
just been analysed in the writing [127,135,136]. Most scientists have finished up that polymer/clay nanocomposites
ought to in any event show PHRR decrease if nanomorphology is accomplished through exfoliation and intercalation.
The distinction in nanomorphologies does not altogether
influence polymer/clay nanocomposite flame retardancy.
Nanocomposites can be obtained by organomodifying
clays, and is easily achieved through melt-compounding.
However, organomodified clay surfaces degrade at high temperatures, rendering organomodification problematic in
melt-compounding and decreasing nanocomposite flame
retardancy [62].
3.2. Fire-retardancy mechanism of carbon-based
nanocomposites
Nanocomposites can be classified into three classes according to the quantity of dimensions of the nanofillers (100 nm)
scattered in polymers: (1) lamellar, (2) nanotubular and (3)
spherical polymer nanocomposites. Carbon-based nanomaterials demonstrating such morphologies are in this way
named graphene, carbon nanotubes (CNTs), and carbon
black (CB), separately. Graphene is the completely exfoliated
structure of graphite (single layer). The technique for delivering graphene was built up as of late, so graphene has
pulled in critical research intrigue [137–141].
CNTs are mostly used as fillers to improve the mechanical, electrical, and flame retardancy properties of nanocomposites. The flame retardancy mechanism of CNTs was
shown by Kashiwagi et al. [122,142,143,]. Since CNTs are
the well-established material, we present mainly their fireretardancy mechanism and some of their disadvantages here.
Figure 18 demonstrates the cone-calorimetry results for
single-wall nanotube and poly (methyl methacrylate)
(PMMA) composites. Ignition time (IT), PHRR and THRR
21
are the essential parameters in cone calorimetry to characterise material flammability. PHRR is the most indispensable
parameter used to delineate flammability and is normal as
the fundamental source of the fire. Including CNTs can
diminish the PHRR; that is, the combustion heat power. In
any case, it cannot change the THRR on the grounds that
CNTs do not act in the vapour phase, inferring that the
measure of fuel gas required for combustion is not changed
by including CNTs. CNTs accelerate flame ignition (i.e., they
diminish the IT). Most polymer/CNT composites demonstrate these propensities. Kashiwagi et al. watched the residues (Figure 19) after the cone calorimetry tests. The rate of
PHRR decrease was little for the nanocomposite containing
0.2 wt% filler, as showed up in (Figure 18). Many black
discrete spots had formed amid the test. Then again, the
surfaces of the nanocomposites containing 0.5 wt% CNTs
were totally secured with uniform, crack free, opening free
CNT network layers, which basically decreased the PHRRs.
The key motivation behind nanocomposite fire retardancy is
the course of action of a uniform CNT layer. CNTcontaining nanocomposites absorb more radiation than
polymers amid fires; hence, nanocomposite temperature
increments are faster than polymer ones. The materials
reduce as the CNTs absorb a ton of radiation. Polymers
begin to burn when they are heated to temperatures at
which thermal degradation begins. The degradation things
are superheated and nucleated to form bubbles. The bubbles
burst at heated surfaces, discharging their substance as fuel
vapour into the gas phase. There are a couple of possible
mechanisms through which CNTs accumulate at material
surfaces: the energy of different rising bubbles in the midst
of combustion pushes the CNTs to the material surface or
the power of the polymer dying down from the material
surface in the midst of pyrolysis, abandoning the CNTs.
The fire-retardancy mechanism of the CNT or char layer
is depicted in Figure 20. Kashiwagi et al. demonstrated that
almost 50% of the occurrence flux was lost through emission
from the hot nanotube surface layer and that the rest of the
flux was traded to the nanotube-network layer and the virgin
Figure 18. Effects of SWNT concentration on mass loss rate of PMMA/AWNT in a nitrogen atmosphere (reproduced from Ref [142].).
22
R. GIRI ET AL.
Figure 19. Residues of PMMA/SWNT after the gasification tests in a nitrogen atmosphere a PMMA, b PMMA/SWNT (0.2%), c PMMA/SWNT (0.5%), d PMMA/SWNT
(1%) (reproduced from Ref [142].
Figure 20. Fire-retardant mechanism for nanoparticles.
example [142]. The nanotube-network layer emits radiation
from the material surface and goes about as a barrier against
the decomposed gas provided from the bulk polymer and
against oxygen diffusing from the air into the material, which
accelerates polymer decomposition. The nanotube-network
layer must be smooth, crack free, and opening free with the
objective that it might go about as a viable gas barrier [144].
Surface-layer cracks deteriorate nanocomposite flame retardancy in the midst of combustion. Rheological properties
appear to rule the making of smooth CNT networks or char
layers for all carbon-based nanocomposites [141,144,145].
Figure 21 demonstrates rheological properties ordinary of
PP/CB nanocomposites [145]. Perfect PP demonstrates regular low-frequency G′−ω scaling, where ω speaks to the
oscillatory frequency. Conversely, the low-frequency G′ scaling vanishes, and G′ turns out to be almost steady at low
frequency for the nanocomposites containing 5 wt% CB,
implying that the nanocomposite changes from a liquid to
a solid, which goes with the arrangement of a mechanically
stable network structure. It is realised that nanocomposites
carrying on like fluids cannot convey smooth CNT network
layer on the material surface. The bubbles bursting at the
surface bothers the course of action of accumulation layer.
Nanocomposites acting like solids contain bubbles that stay
little in the high-thickness layer and transport to the material
surface, which tends not to bother the plan of the accumulation layer [142].
Reliably dispersed nanocomposites demonstrate rheological properties like those of authentic solids. Thus, the
dispersion of carbon-based nanofillers chooses the idea of
the surface layer framed in the midst of combustion and in
this way impacts the nanocomposite flame retardancy [146].
Picking appropriate CNTs is fundamental. Barus et al. contemplated the properties of three sorts of CNTs and found
that the dispersivity of the CNTs themselves impacts the
thermal degradation of nanocomposites [147]. The CNT
stack is likewise fundamental in choosing fire retardancy,
and really, the perfect CNT loading diminishes PHRR, as
appeared in (Figure 18). Including more CNTs, once the
uniform surface layer has shaped disintegrates flame retardancy since it energises the agglomeration of CNTs and
upgrades thermal conductivity. The CNT aspect proportion
impacts fire retardancy, and higher aspect ratios prompt
more imperative abatements in PHRR [148], demonstrating
that a system for compounding thermoplastics to disperse
CNTs and to abandon them longer is satisfactory. From
a handy perspective, a twin-screw extruder can be utilised
to first compound thermoplastics, which will then be injection moulded all together to develop a method for large-scale
manufacturing thermoplastic-based items.
MATERIALS RESEARCH INNOVATIONS
23
Figure 21. The rheological properties of neat PP and PP/CB nanocomposites (reproduced from Ref [145].).
Figure 22. TGA (a) Effects of concentration of MWNT in PP on heat release rate of PP/MWNT nanocomposite at 50 kW/m2. (b) SEM picture of MWNT dispersion in
the PP/MWNT (4%) nanocomposite after solvent removal of PP. (c) Collected residues after the gasification experiment at 50 kW/m2 in nitrogen about PP/MWNT
(1%). (reproduced from Ref [122].).
The relations amongst handling and flame-retardancy
viability should be talked about to apply CNTs in commercial items, for example, flame retardants. Actually, dispersing
CNTs through a twin-screw extruder is more terrible than
the nanocomposite prepared by utilising a closed kneader.
Besides, injection moulding can break down material
strands. How these influences of nanocomposite flammability should be talked about. Pötschke et al. explored the
relations between CNT dispersion and handling conditions
[149–151]. They concentrated on the electrical conductivities
of nanocomposites and directed extensive investigations.
Successful compounding techniques are vital for growing
low-CNT-loading fire-retardant nanocomposites. Other carbon-based materials, for example, CB and graphene have
additionally as of late been explored as flame retardants.
Dittrich et al. demonstrated that graphene was the best
carbon-based fire-retardant material [141]. Moreover, Wen
et al. discovered new fire-retardancy mechanism for
CB [145].
They demonstrated that peroxy radicals, the principle
factor affecting the thermal degradation of polypropylene
(PP), could be successfully gotten in CB at raised temperatures to outline a gelled-ball crosslinked composition. The
PHRR was decreased 75% and the LOI enhanced from 18%
to 27.6% by joining CB (to trap the peroxy radicals) and CNT
(to make the networked layer) [152]. Surprisingly, adding CB
and CNTs reduced the THRR; other nanocomposite systems
do not demonstrate this inclination. The new fire-retardancy
24
R. GIRI ET AL.
Figure 23. (a) Cone calorimeter experiments, heat release rate versus time curves: comparison of the effect of the 3wt% different nanofillers (Na-MMT, MWNTs and
graphene). (b) Morphologies of the final chars after cone calorimeter tests of PVA-Graphene 3wt% composite. (Reproduced from Ref [153].).
mechanism for CB can possibly additionally enhance the
flame retardancy of carbon-based nanocomposites.
Kashiwagi et al. [122] examined the flammability of a CNT
polymer nanocomposite material and a carbon dark (CB)
polymer composite. CNTs have alternate morphology as
compared with nanosilicate or graphene; silicate or silicate
clay has two-dimensional plate-type morphology, while
CNTs have one-dimensional wire like morphology.
However, CNT composites have comparative mechanisms,
CNT additionally creates residual form, have CNT structural
cone. As shown at (Figure 22) CNT polymer nanocomposite
indicates low HRR graph compared with novel polypropylene. One per cent CNT/PP composite demonstrates the least
heat release rate; the explanation behind the most minimal
PHRR is expected to the balance between the impact of
thermal conductivity and the protecting execution of external radiant flux (and heat feedback from the flame) contingent upon the concentration of MWNT in the sample.
A nanotube network layer comprising of carbon nanotubes
is formed, and it covers the whole sample surface with no
noteworthy crack formation amid burning.
Graphene has a chemical structure similar to CNTs, composed of sp2 carbon atoms arranged in a honeycomb structure,
and the same morphology as nanosilicate clay’s twodimensional sheet. When MMT, CNT, and graphene are compared with the same 3 wt% of nanometric polymer composite
Figure 24. (a) Heat release rate versus time chart for unfilled PP and ternary PP nanocomposite (10 wt. % Sep + 2 wt. % MWNT). SEM images of (b) PP + CNT (c) PP
+ Sep and (d) PP ternary system post-cone calorimeter test. (Reproduced from Ref [154].).
MATERIALS RESEARCH INNOVATIONS
25
for flammability using a cone calorimeter, they have different
behaviours, as shown in (Figure 23). Compared to pure PVA,
the PHRR of PVA-G3 is reduced by 49% and is lower than
those of PVA-MMT and PVA-MWNTs. To explain this unexpected fire behaviour, morphologies of the final chars after cone
calorimeter tests of the PVA composites were based on graphene nanosheets by using SEM. The SEM images (Figure 23
(b)) for the residues of PVA-G3 after cone calorimeter tests
showed that many quadrate carbonaceous particles (1–3μm)
joined each other and formed the compact, dense and uniform
char. The graphene can promote the formation of compact char
layers in condensed phase during combustion of the polymer
matrix. Furthermore, the char structure effectively prevents the
inside thermal decomposition products into the flame zone and
that of the O2 into the underlying polymer matrix [153].
These nanometric composites indicated great flammability compared with novel polymer resin. Also, they can have
a synergetic impact on flammability, when they are utilised
together in a polymer matrix as additives. Sepiolite nanoclay
(Sep) has an extraordinary morphology in the class of nanoclay, and CNT makes a network, forming a tight char. Clay
and CNT both make char layers. When they are utilised
together, the clay and CNT composite makes a much higher
density network frame, when compared with other CNT/PP
or clay/PP composite cone calorimeter deposit, appeared in
(Figure 24). In this manner, the PHRR can be powerfully
reduced from a novel PP of 1,933 kW/m2 to PP/CNT/Sep
composite of 355 kW/m2 for the 10wt% Sepiolite + 2wt%
MWNT ternary nanocomposites system, these outcomes
began from a higher density char residue. The utilisation of
sepiolite nanoclay in combination with multi-walled CNTs
demonstrated that the PHRR was fundamentally decreased
by 82% compared with the neat polymer in the cone calorimeter as shown in (Figure 24) [154]. Marosfoil et al. likewise
contemplated the flame retardancy performance of CNT
filled PP and figured out how to reduce the PHRR from
2,755 kW/m2 to 760kW/m2 for PP and PP/CNT, respectively
[154,155]. In this way, utilising diverse morphological nanometric additives can prompt synergetic flammability
execution.
mechanism of activity is achieved so new innovative leaps
forward could be spearheaded. In that regard, the principle
target of this investigation was to audit the collaborations
between polymer nanocomposites and conventional flame
retardants, as dissected in open writing, planning to give
a useful device to those directing exploration in this field.
Attempting an inside and out investigation of these frameworks uncovered that, in spite of the current advance, some
essential inquiries on a few aspects still stay unanswered. One
inquiry, for instance, is whether the responsive refers to living
on silicates are idle amid fire or respond with the flame
retardant impacting its viability. Then again, there exists no
hypothetical foundation to portray quantitatively the correlation\between the diminishment in pHRR and dispersion of
silicates. Polymer-nanocomposite retardancy towards supplanting halogenated flame retardants with safe, ecoaccommodating polymer nanocomposite. Nanocomposites
should be joined with legitimate flame retardants to accomplish a similar flame retardancy that halogenated mixes
appear.
The principle mechanism of nanofiller-based flameretardant materials is that the nanofillers act as barriers
against gas flow and oxygen dissemination at the condensed
phase. In this manner, delivering solid, thick, crack-free
nanocoating surface layers amid combustion is the key factor
in creating compelling polymer nanocomposite flame retardants. Nanofiller dispersion and distribution are essential
elements adding to flame retardancy synergistic impacts,
which are hopeless when nanofillers are incorporated into
other flame retardants since poor nanofiller distribution dissipates surface layers rendering them unfit to go about as
barrier layers. Nanofiller/flame-retardant cooperations likewise add to nanofiller flame retardant – proficiency. The
representative nanometric materials, including nanoclay,
carbon nanotube and graphene, referred to in this chapter
can be used as flame retardancy additives. In addition, an
organic intumescent system was summarised regarding its
chemical composition and mechanism. Occasionally nanometric additives, for achieving high flammability, have priority for disparity in polymer composites.
4. Conclusions
Author Contributions
The flame retardancy of polymer/layered silicate nanocomposites, assessed in this section, is the result of a few mechanisms;
the level to which every last one of them impacts material’s
execution is exceedingly connected with the fire situation
under core interest. The development of a shallow carbonaceous-silicate charred layer is for the most part considered as
the primary method of ‘nanocomposites’ flame-retardant
activity in creating fires. The likelihood of formulating selfextinguishing nanocomposites is given by conventional flame
retardants which can enhance fire execution without truly
disintegrating mechanical properties, when utilised at unassuming loadings. All conceivable methods for embedding
flame retardants into polymer/layered silicate nanocomposites
(PLSN) are discovered: direct blending, chemical connection
to the clay cation, and incorporation in the polymer backbone
to form flame-retardant copolymer nanocomposites. In many
cases, the results bring up that flame retardants serve to
advance the fire properties of nanocomposites. The extensive
scale commercialisation, in any case, of flame-retardant PLSN
will be expert just when more profound learning on their
R.G., L.N., and M.R. all have partly contributed to writing the
manuscript.
Disclosure statement
The authors declare no conflict of interest.
Funding
This research received no external funding.
Notes on contributors
Dr. Radhashyam Giri is presently working as an Assistant Professor &
H.O.D of the Plastics Engineering Department CIPET: IPT
Bhubaneswar, Odisha India. He received Master of Technology (M.
Tech) degree in Plastics Engineering and Technology in 2007 from Biju
Pattanaik University of Technology (BPUT) Rourkela, Orissa, India.
Dr. Giri received Doctor of Philosophy (Ph.D) from Rubber
Technology Centre, Indian Institute of Technology, Kharagpur, West
Bengal, India in 2012. Dr. Giri has nineteen international peer reviewed
26
R. GIRI ET AL.
journal papers, three book chapters, one book published, 43 international oral conference papers on his credit and five more journal papers
are in the pipeline. Apart from this Dr. Giri has so many achievements
towards academic & research activity: He received R & D Institution
runner up 5th National Awards for Technology Innovation in
Petrochemical and downstream Plastic Processing Industry from Late
Shri Ananth Kumar, Hon'ble former Minister for Chemicals &
Fertilizers, Government of India on 21st February, 2015 at Hotel Lalit
Ashok, Bengaluru. He is Editorial Board member for the esteemed
journal, International Journal of Engineering Development and
Research (IJEDR).
Dr. Lalatendu Nayak is a Senior Manager in Specialty division at
Phillips Carbon Black’s Global Research and Development Centre,
Palej, Gujarat, India. He has completed his PhD (Polymer) from
Indian Institute of Technology, Kharagpur, West Bengal, India,
MTech (Polymer) from Central Institute of Plastics Engineering and
Technology, Bhubaneswar, Odisha, India, and M.Sc. (Organic
Chemistry) from Fakir Mohon University, Odisha, India. He has published more than 20 research articles in different international journal
and published two book chapters. He is an active reviewer in different
international journals. He has more than 10 years’ research experience
in the field of conductive polymer composites and nanocomposites,
synthesis and surface treatment of nanomaterials. His research interest
includes surface modification of fillers for ink, paint and coating applications, synthesis of nanomaterials for energy storage applications,
conductive polymer nanocomposites for electrostatic discharge (ESD),
electromagnetic interference shielding (EMIS), and stealth technology
applications
Dr. Mostafizur Rahaman is an Assistant Professor at the Department of
Chemistry at the College of Science, King Saud University, Riyadh
11451, Saudi Arabia. He obtained his M. Sc. (Chemistry) from T. M.
Bhagalpur University, India and Ph. D. (Chemical/Polymer Chemistry)
from the Indian Institute of Technology Kharagpur, India. He completed his M. Tech. in Plastics Engineering at the Central Institute of
Plastics Engineering and Technology (CIPET), Bhubaneswar, Orissa,
India. He has published 60+ articles in international journals, 9 book
chapters, and 15 research articles in international conference proceedings. He has also published 1 patent and 3 books. Dr. Rahaman has 11
years of teaching and 12 years of research experience. He has completed
eight research projects and attended/presented at various international
conferences/seminars. He has been an active reviewer for various international journals and member of journal advisory boards. An expert in
handling sophisticated instruments. His research interests include polymer nanotechnology/nanocomposites; polymer membrane, polymer
thin film; polymer-based sensors; catalytic synthesis of polymers and
conducting polymers; polymer-based coating for corrosion protection;
polymer fuel cell and solar energy and bio-polymers for biomedical
applications.
References
[1] Troitzsch J. International plastics flammability handbook: principles-regulations-testing and approval. Munich, Germany:
Hanser; 1983.
[2] Brossas J. Fire retardance in polymers: an introductory lecture.
Polym Degrad Stab. 1989;23(4):313–325.
[3] Wilkie CA, Morgan AB. Fire retardancy of polymeric materials.
Florida, US: CRC press; 2009.
[4] Manzi-Nshuti C, Songtipya P, Manias E, et al. Polymer nanocomposites using zinc aluminum and magnesium aluminum
oleate layered double hydroxides: effects of the polymeric compatibilizer and of composition on the thermal and fire properties
of PP/LDH nanocomposites. Polym Degrad Stab. 2009;94
(11):2042–2054.
[5] Irvine D, McCluskey J, Robinson I. Fire hazards and some
common polymers. Polym Degrad Stab. 2000;67(3):383–396.
[6] Wei M, Murphy D, Barry C, et al. Halogen-free flame retardants
for wire and cable applications. Rubber Chem Technol. 2010;83
(3):282–302.
[7] Betts K FLAME RETARDANT TRAVELS THE GLOBE. AMER
CHEMICAL SOC 1155 16TH ST, NW, WASHINGTON, DC
20036 USA; 2010.
[8] Wang N Flame retardancy of polymer nanocomposites based on
layered aluminum phosphate and computational study of intercalation of amines into α-zirconium phosphate and adsorption
of a model organic pollutant. 2011.
[9] Wang L, He X, Wilkie CA. The utility of nanocomposites in fire
retardancy. Materials. 2010;3(9):4580–4606.
[10] Wang D, Wilkie CA. A stibonium-modified clay and its polystyrene nanocomposite. Polym Degrad Stab. 2003;82
(2):309–315.
[11] Habibi S, Rashidi A, Shahvaziyan M, et al. Effect of clay on
degradation of poly (ethylene terephthalate)/montmorillonite
Nanocomposites. Asian J Chem. 2010;22(9):7087.
[12] Sun L, Boo WJ, Liu J, et al. Effect of nanoplatelets on the
rheological behavior of epoxy monomers. Macromol Mater
Eng. 2009;294(2):103–113.
[13] Lu H, Song L, Hu Y. A review on flame retardant technology in
China. Part II: flame retardant polymeric nanocomposites and
coatings. Polym Adv Technol. 2011;22(4):379–394.
[14] Wang D-Y, Das A, Leuteritz A, et al. Thermal degradation
behaviors of a novel nanocomposite based on polypropylene
and Co–Al layered double hydroxide. Polym Degrad Stab.
2011;96(3):285–290.
[15] Becker CM, Gabbardo AD, Wypych F, et al. Mechanical and
flame-retardant properties of epoxy/Mg–Al LDH composites.
Compos Part A Appl Sci Manuf. 2011;42(2):196–202.
[16] Zhang J, Jiang DD, Wilkie CA. Thermal and flame properties of
polyethylene and polypropylene nanocomposites based on an
oligomerically-modified clay. Polym Degrad Stab. 2006;91
(2):298–304.
[17] Satyanarayana KG. Clay surfaces: fundamentals and applications. Vol. 1. Boston, US: Elsevier; 2004.
[18] Nyambo C, Wilkie CA. Layered double hydroxides intercalated
with borate anions: fire and thermal properties in ethylene vinyl
acetate copolymer. Polym Degrad Stab. 2009;94(4):506–512.
[19] Zhang J, Jiang DD, Wilkie CA. Fire properties of styrenic polymer–clay nanocomposites based on an oligomerically-modified
clay. Polym Degrad Stab. 2006;91(2):358–366.
[20] Zhu J, Start P, Mauritz KA, et al. Thermal stability and flame
retardancy of poly (methyl methacrylate)-clay nanocomposites.
Polym Degrad Stab. 2002;77(2):253–258.
[21] Landrock AH Handbook of plastics flammability and combustion toxicology: principles, materials, testing, safety, and smoke
inhalation effects. NOYES DATA CORP, MILL ROAD AT
GRAND AVE, PARK RIDGE, NJ 07656, USA 1983. 1983.
[22] Morgan AB, Gilman JW. An overview of flame retardancy of
polymeric materials: application, technology, and future
directions. Fire Mater. 2013;37(4):259–279.
[23] Cullis CF, Hirschler M. The combustion of organic polymers.
Vol. 5. USA: Oxford University Press; 1981.
[24] Lyon RE. Pyrolysis kinetics of char forming polymers. Polym
Degrad Stab. 1998;61(2):201–210.
[25] Glassman I, Yetter RA, Glumac NG. Combustion. Academic
press; Elsevier, Boston, US; 2014.
[26] Knauss D, McGrath J, Kashiwagi T, editors. In Fire and
Polymers II: materials and Tests for Hazard Prevention.
Nelson, GL, Ed. ACS Symposium Series, An American
Chemical Society Publication, US, 1995.
[27] Cote AE. Fire protection handbook. Vol. 2. Massachusetts, US:
NationalFireProtectionAssoc; 2008.
[28] Scudamore M, Briggs P, Prager F. Cone calorimetry—a review of
tests carried out on plastics for the association of plastic manufacturers in Europe. Fire Mater. 1991;15(2):65–84.
[29] Frazer AH High temperature resistant polymers. 1968.
[30] Lin S-H, Li F, Cheng SZ, et al. Organo-Soluble Polyimides:
synthesis and Polymerization of 2, 2 ‘-Bis (trifluoromethyl)-4, 4
‘, 5, 5 ‘-Biphenyltetracarboxylic Dianhydride. Macromolecules.
1998;31(7):2080–2086.
[31] Mallakpour SE, Hajipour AR, Mahdavian AR, et al. Synthesis
and characterization of novel optically active and flame-retardant heterocyclic polyimides. J Appl Polym Sci. 2000;76
(2):240–248.
[32] Chang J-H, Chen MJ, Farris RJ. Effect of heat treatment on the
thermal and mechanical properties of a precursor polymer:
polyhydroxyamide. Polymer. 1998;39(23):5649–5654.
MATERIALS RESEARCH INNOVATIONS
[33] Sirkecioglu O, Tunca AA, Talinli N, et al. Ladder type polymers
from dihydroxyaromatic compounds and dialdehydes. Die
Angew Makromol Chemie. 1999;271(1):8–10.
[34] Gajiwala H, Zand R. Synthesis and characterization of thermally
stable polymers containing phenazine. Polymer. 2000;41
(6):2009–2015.
[35] Barbosa VF, MacKenzie KJ, Thaumaturgo C. Synthesis and
characterisation of materials based on inorganic polymers of
alumina and silica: sodium polysialate polymers. Int J Inorg
Mater. 2000;2(4):309–317.
[36] Kumar D, Gupta AD, Khullar M. Heat-resistant thermosetting
polymers based on a novel tetrakisaminophenoxycyclotriphosphazene. J Polym Sci A Polym Chem. 1993;31(11):2739–2745.
[37] Lyons JW. chemistry and uses of fire retardants. 1970.
[38] Council NR. Fire and smoke: understanding the hazards.
National Academies Press; 1986.
[39] Bolger R FLAME RETARDANT MINERALS: BROMINE ISSUE
SMOULDERS ON. R Bolger, Ind Minerals(1996). 1996 (340).
[40] Bachtiar EV, Kurkowiak K, Yan L, et al. Thermal stability, fire
performance, and mechanical properties of natural fibre
fabric-reinforced polymer composites with different fire
retardants. Polymers. 2019;11(4):699.
[41] Taniguchi S, Sakuma Y, Yoshii T Flame retardant additives
based on alumina trihydrate and ethylene polymer compositions, containing same, having improved flame retardant
properties. Google Patents; 1984.
[42] Hornsby PR. The application of magnesium hydroxide as a fire
retardant and smoke-suppressing additive for polymers. Fire
Mater. 1994;18(5):269–276.
[43] Hornsby PR, editor. The application of hydrated mineral fillers
as fire retardant and smoke suppressing additives for polymers.
Macromolecular Symposia. New Jersey, US: Wiley Online
Library; 1996.
[44] O’Driscoll M. CAUSTIC MAGNESIA MARKETS. PLAYING
WITH FIRE. Ind Miner. 1994;318:23–45.
[45] Hilado CJ. Flammability handbook for plastics. Florida, US:
CRC Press; 1998.
[46] Lee F, Nicholson P, Green J. New reactive phosphorus
flame-retardant for rigid polyurethane application1. J Fire
Retardant Chem. 1982;9(3):194–205.
[47] Allen CW. The use of phosphazenes as fire resistant materials.
J Fire Sci. 1993;11(4):320–328.
[48] Green J. Phosphorus-bromine flame retardant synergy in
a polycarbonate/polyethylene terephthalate blend. J Fire Sci.
1994;12(3):257–267.
[49] Inoue K, Nakamura H, Ariyoshi S, et al. Heat-resistant polymers
prepared from [(4ʹ-(2-vinyl)-4-biphenylyl) oxy] pentachlorocyclotriphosphazene. Macromolecules. 1989;22(12):4466–4469.
[50] Montaudo G, Scamporrino E, Vitalini D. The effect of ammonium polyphosphate on the mechanism of thermal degradation
of polyureas. J Polym Sci. 1983;21(11):3321–3331.
[51] Horacek H, Grabner R. Advantages of flame retardants based on
nitrogen compounds. Polym Degrad Stab. 1996;54
(2–3):205–215.
[52] Horacek H, Grabner W, editors Nitrogen based flame retardants
for nitrogen containing polymers. In: Makromolekulare
Chemie. Macromolecular Symposia. Wiley Online Library;
1993.
[53] Iji M, Serizawa S. Silicone derivatives as new flame retardants for
aromatic thermoplastics used in electronic devices? Polym Adv
Technol. 1998;9(10-11):593–600.
[54] Muiambo HF, Focke WW, Asante JK, editors. Flame retardant
properties of polymer composites of urea complex of magnesium and vermiculite. AIP Conference Proceedings; 2019: AIP
Publishing.
[55] Kalali EN, Zhang L, Shabestari ME, et al. Flame-retardant wood
polymer composites (WPCs) as potential fire safe bio-based
materials for building products: preparation, flammability and
mechanical properties. Fire Saf J. 2019;107:210–216.
[56] Suoware TO, Edelugo SO, Ezema IC. Impact of hybrid flame
retardant on the flammability and thermomechanical properties
of wood sawdust polymer composite panel. Fire Mater. 2019;43
(4):335–343.
[57] Kambour R, Klopfer H, Smith S. Limiting oxygen indices of
silicone block polymer. J Appl Polym Sci. 1981;26(3):847–859.
27
[58] Zaikov GE, Lomakin SM. New aspects of ecologically friendly
polymer flame retardant systems. Polym Degrad Stab. 1996;54
(2–3):223–233.
[59] Allen N, Edge M, Corrales T, et al. Entrapment of stabilisers in
silica: I. Controlled release of additives during polypropylene
degradation. Polym Degrad Stab. 1997;56(2):125–139.
[60] Chao TC, Sarmah SK, Boisvert RP, et al. Non-burning silicone
resin composite materials. Materials and process affordabilityKeys to the future. 1998:1029–1041.
[61] Bolf A, Lichtenhan J, editors. Thermoplastic elastomer blends
prepared from polycarbonsilane and polysilastyrene preceramic
polymers. Meeting of the Division of Polymer Chemistry. 1994.
[62] Gilman JW, Jackson CL, Morgan AB, et al. Flammability properties of polymer− layered-silicate nanocomposites. Polypropylene
and polystyrene nanocomposites. Chem Mater. 2000;12
(7):1866–1873.
[63] Giannelis EP. Polymer layered silicate nanocomposites. Adv
Mater. 1996;8(1):29–35.
[64] West R, Mark J, Allcock H, et al. Inorganic Polymers. New
Jersey: Pren-tice Hall; 1992. p. 141–185.
[65] Grand AF, Wilkie CA. Fire retardancy of polymeric materials.
Florida, US: CRC Press; 2000.
[66] Troitzsch J. Plastics flammability handbook: principles, regulations, testing, and approval. Munich, Germany: Hanser Verlag;
2004.
[67] Zhao C, Qin H, Gong F, et al. Mechanical, thermal and flammability properties of polyethylene/clay nanocomposites. Polym
Degrad Stab. 2005;87(1):183–189.
[68] Camino G, Costa L, Di Cortemiglia ML. Overview of fire retardant mechanisms. Polym Degrad Stab. 1991;33(2):131–154.
[69] Skinner G. Flame retardancy: the approaches available. In:
Plastics Additives: Dordrecht, Netherlands: springer. 1998. p.
260–267.
[70] Hirschler MM. Chemical aspects of thermal decomposition of
polymeric materials. In: Fire retardancy of polymeric materials.
CRC Press, Florida, US 2000..
[71] Pettigrew A. Flame retardants, halogenated. In: Kirk-Othmer
Encyclopedia of Chemical Technology. Wiley Online Library,
New Jersey, US, 2000.
[72] Levchik SV, Weil ED, Levchik SV, et al. Combustion and fire
retardancy of aliphatic nylons. Poly Int. 2000;49(10):1033–1073.
[73] Zanetti M, Camino G, Canavese D, et al. Fire retardant halogen−
antimony− clay synergism in polypropylene layered silicate
nanocomposites. Chem Mater. 2002;14(1):189–193.
[74]. Hu Y, Song L. Nanocomposites with halogen and nonintumescent phosphorus flame retardant additives. Flame Retardant
Polymer Nanocomposites. Morgan AB, Wilkie CA. Eds 2007;
p. 191–234
[75] Lee J, Nam J, Lee D, et al. Flame retardancy of polypropylene/
montmorillonite nanocomposites with halogenated flame
retardants. Polymer-Korea. 2003;27(6):569–575.
[76] Hu Y, Wang S, Ling Z, et al. Preparation and combustion
properties of flame retardant nylon 6/montmorillonite
nanocomposite. Macromol Mater Eng. 2003;288(3):272–276.
[77] Wang S, Hu Y, Zong R, et al. Preparation and characterization of
flame retardant ABS/montmorillonite nanocomposite. Appl
Clay Sci. 2004;25(1–2):49–55.
[78] Ma H, Fang Z, Tong L. Preferential melt intercalation of clay in
ABS/brominated epoxy resin–antimony oxide (BER–AO) nanocomposites and its synergistic effect on thermal degradation and
combustion behavior. Polym Degrad Stab. 2006;91(9):1972–1979.
[79] Weil ED, Patel NG, Said M, et al. Oxygen index: correlations to
other fire tests. Fire Mater. 1992;16(4):159–167.
[80] Wang D, Echols K, Wilkie CA. Cone calorimetric and thermogravimetric analysis evaluation of halogen-containing polymer
nanocomposites. Fire Mater. 2005;29(5):283–294.
[81] Chigwada G, Jash P, Jiang DD, et al. Synergy between nanocomposite formation and low levels of bromine on fire retardancy in
polystyrenes. Polym Degrad Stab. 2005;88(3):382–393.
[82] Si M, Zaitsev V, Goldman M, et al. Self-extinguishing polymer/
organoclay nanocomposites. Polym Degrad Stab. 2007;92
(1):86–93.
[83] Levchik SV. Introduction to flame retardancy and polymer
flammability. Flame Retardant Polym Nanocomposites.
2007;1–29.
28
R. GIRI ET AL.
[84] Laoutid F, Bonnaud L, Alexandre M, et al. New prospects in
flame retardant polymer materials: from fundamentals to
nanocomposites. Mater Sci Eng R Rep. 2009;63(3):100–125.
[85] Song P, Tong L, Fang Z. Polypropylene/clay nanocomposites
prepared by in situ grafting-melt intercalation with a novel
cointercalating monomer. J Appl Polym Sci. 2008;110
(1):616–623.
[86] Zhang S, Horrocks AR. A review of flame retardant polypropylene fibres. Prog Polym Sci. 2003;28(11):1517–1538.
[87] Hao X, Gai G, Liu J, et al. Flame retardancy and antidripping
effect of OMT/PA nanocomposites. Mater Chem Phys. 2006;96
(1):34–41.
[88] Chigwada G, Wilkie CA. Synergy between conventional phosphorus fire retardants and organically-modified clays can lead to
fire retardancy of styrenics. Polym Degrad Stab. 2003;81
(3):551–557.
[89] Zheng X, Wilkie CA. Flame retardancy of polystyrene nanocomposites based on an oligomeric organically-modified clay containing phosphate. Polym Degrad Stab. 2003;81(3):539–550.
[90] Kim J, Lee K, Lee K, et al. Studies on the thermal stabilization
enhancement of ABS; synergistic effect of triphenyl phosphate
nanocomposite, epoxy resin, and silane coupling agent mixtures.
Polym Degrad Stab. 2003;79(2):201–207.
[91] Gijsman P, Steenbakkers R, Fürst C, et al. Differences in the
flame retardant mechanism of melamine cyanurate in polyamide
6 and polyamide 66. Polym Degrad Stab. 2002;78(2):219–224.
[92] Weil ED, Levchik S. Current practice and recent commercial
developments in flame retardancy of polyamides. J Fire Sci.
2004;22(3):251–264.
[93] Kiliaris P, Papaspyrides CD, Pfaendner R. Polyamide 6 filled
with melamine cyanurate and layered silicates: evaluation of
flame retardancy and physical properties. Macromol Mater
Eng. 2008;293(9):740–751.
[94] Zhang J, Lewin M, Pearce E, et al. Flame retarding polyamide 6
with melamine cyanurate and layered silicates. ?Polym Adv
Technol. 2008;19(7):928–936.
[95] Gianelli W, Camino G, Tabuani D, et al. Fire behaviour of
polyester–clay nanocomposites. Fire Mater. 2006;30(5):333–341.
[96] Wilkie CA An introduction to the use of fillers and nanocomposites in fire retardancy. Fire retardancy of polymers: new
applications of mineral fillers. 2005:1–15.
[97] Butts M, Cella J, Wood CD, et al. Silicones. In: Encyclopedia of
polymer science and technology. 2002.
[98] Quede A, Mutel B, Supiot P, et al. Characterization of organosilicon films synthesized by N2-PACVD. Application to fire
retardant properties of coated polymers. Surf Coat Technol.
2004;180:265–270.
[99] Huo Y, Fan Q, Dembsey NA, et al. Improvements on flame
retardant properties of PET/montmorillonite nanocomposite
caused by polyborosiloxane. MRS Online Proc Lib Arch.
Cambridge University Press. 2007;1007: S12–33.
[100] Zhu J, Start P, Mauritz KA, et al. Silicon-methoxide-modified
clays and their polystyrene nanocomposites. J Polym Sci
A Polym Chem. 2002;40(10):1498–1503.
[101] Dong W, Zhang X, Liu Y, et al. Flame retardant nanocomposites
of polyamide 6/clay/silicone rubber with high toughness and
good flowability. Polymer. 2006;47(19):6874–6879.
[102] Wang Q, Zhang X, Jin Y, et al. Preparation and Properties of
PVC Ternary Nanocomposites Containing Elastomeric
Nanoscale Particles and Exfoliated Sodium-Montmorillonite.
Macromol Mater Eng. 2006;291(6):655–660.
[103] Lewin M, Zhang J, Pearce E, et al. Flammability of polyamide 6
using the sulfamate system and organo-layered silicate. ?Polym
Adv Technol. 2007;18(9):737–745.
[104] Thostenson ET, Ren Z, Chou T-W. Advances in the science and
technology of carbon nanotubes and their composites: a review.
Compos Sci Technol. 2001;61(13):1899–1912.
[105] Dubois P, Alexandre M. Performant clay/carbon nanotube polymer nanocomposites. Adv Eng Mater. 2006;8(3):147–154.
[106] Fischer JE. Carbon nanotubes: structure and properties.
Nanotubes Nanofibers. 2006;1:1.
[107] Beyer G. Carbon nanotubes as flame retardants for polymers.
Fire Mater. 2002;26(6):291–293.
[108] Peeterbroeck S, Alexandre M, Nagy J, et al. Polymer-layered
silicate–carbon nanotube nanocomposites: unique nanofiller
synergistic
effect.
Compos
Sci
Technol.
2004;64
(15):2317–2323.
[109] Gao F, Beyer G, Yuan Q. A mechanistic study of fire retardancy
of carbon nanotube/ethylene vinyl acetate copolymers and their
clay composites. Polym Degrad Stab. 2005;89(3):559–564.
[110] Beyer G. Flame retardancy of nanocomposites based on organoclays and carbon nanotubes with aluminium trihydrate. ?Polym
Adv Technol. 2006;17(4):218–225.
[111] Beyer G. Filler blend of carbon nanotubes and organoclays with
improved char as a new flame retardant system for polymers and
cable applications. Fire Mater. 2005;29(2):61–69.
[112] Tang T, Chen X, Chen H, et al. Catalyzing carbonization of
polypropylene itself by supported nickel catalyst during combustion of polypropylene/clay nanocomposite for improving fire
retardancy. Chem Mater. 2005;17(11):2799–2802.
[113] Ma H, Tong L, Xu Z, et al. Synergistic effect of carbon nanotube
and clay for improving the flame retardancy of ABS resin.
Nanotechnology. 2007;18(37):375602.
[114] Kuljanin J, MarinovićCincović M, Zec S, et al. Influence of Fe 2
O 3-filler on the thermal properties of polystyrene. J Mater Sci
Lett. 2003;22(3):235–237.
[115] Laachachi A, Cochez M, Ferriol M, et al. Influence of TiO2 and
Fe2O3 fillers on the thermal properties of poly (methyl
methacrylate)(PMMA). Mater Lett. 2005;59(1):36–39.
[116] Laachachi A, Leroy E, Cochez M, et al. Use of oxide nanoparticles and organoclays to improve thermal stability and fire retardancy of poly (methyl methacrylate). Polym Degrad Stab.
2005;89(2):344–352.
[117] Arao Y, editor. Flame retardancy of polymer nanocomposite.
Flame Retardants. Springer; 2015.
D,
Robeson
LM.
Polymer
nanotechnology:
[118] Paul
nanocomposites. Polymer. 2008;49(15):3187–3204.
[119] Qin H, Zhang S, Zhao C, et al. Flame retardant mechanism of
polymer/clay nanocomposites based on polypropylene. Polymer.
2005;46(19):8386–8395.
[120] Kashiwagi T, Harris JRH, Zhang X, et al. Flame retardant
mechanism of polyamide 6–clay nanocomposites. Polymer.
2004;45(3):881–891.
[121] Zhu J, Uhl FM, Morgan AB, et al. Studies on the mechanism by
which the formation of nanocomposites enhances thermal
stability. Chem Mater. 2001;13(12):4649–4654.
[122] Kashiwagi T, Grulke E, Hilding J, et al. Thermal and flammability properties of polypropylene/carbon nanotube
nanocomposites. Polymer. 2004;45(12):4227–4239.
[123] Bartholmai M, Schartel B. Layered silicate polymer nanocomposites: new approach or illusion for fire retardancy? Investigations
of the potentials and the tasks using a model system. ?Polym Adv
Technol. 2004;15(7):355–364.
[124] Song R, Wang Z, Meng X, et al. Influences of catalysis and
dispersion of organically modified montmorillonite on flame
retardancy of polypropylene nanocomposites. J Appl Polym
Sci. 2007;106(5):3488–3494.
[125] Hu Y, Tang Y, Song L. Poly (propylene)/clay nanocomposites
and their application in flame retardancy. ?Polym Adv Technol.
2006;17(4):235–245.
[126] Fina A, Cuttica F, Camino G. Ignition of polypropylene/montmorillonite nanocomposites. Polym Degrad Stab. 2012;97
(12):2619–2626.
[127] Zhang J, Bai M, Wang Y, et al. Featured structures of fire residue
of high-impact polystyrene/organically modified montmorillonite nanocomposites during burning. Fire Mater. 2012;36
(8):661–670.
[128] Liu J, Fu M, Jing M, et al. Flame retardancy and charring
behavior of polystyrene-organic montmorillonite nanocomposites. ?Polym Adv Technol. 2013;24(3):273–281.
[129] Wang S, Hu Y, Zhongkai Q, et al. Preparation and flammability
properties of polyethylene/clay nanocomposites by melt
MATERIALS RESEARCH INNOVATIONS
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]
[139]
[140]
[141]
[142]
[143]
intercalation method from Na+ montmorillonite. Mater Lett.
2003;57(18):2675–2678.
Zhang J, Wilkie CA. Preparation and flammability properties of
polyethylene–clay nanocomposites. Polym Degrad Stab. 2003;80
(1):163–169.
Jash P, Wilkie CA. Effects of surfactants on the thermal and fire
properties of poly (methyl methacrylate)/clay nanocomposites.
Polym Degrad Stab. 2005;88(3):401–406.
Sahoo PK, Samal R. Fire retardancy and biodegradability of poly
(methyl methacrylate)/montmorillonite nanocomposite. Polym
Degrad Stab. 2007;92(9):1700–1707.
Jang BN, Costache M, Wilkie CA. The relationship between
thermal degradation behavior of polymer and the fire retardancy
of
polymer/clay
nanocomposites.
Polymer.
2005;46
(24):10678–10687.
Kashiwagi T, Shields JR, Harris JRH, et al. Flame-retardant
mechanism of silica: effects of resin molecular weight. J Appl
Polym Sci. 2003;87(9):1541–1553.
Samyn F, Bourbigot S, Jama C, et al. Fire retardancy of polymer
clay nanocomposites: is there an influence of the
nanomorphology?. Polym Degrad Stab. 2008;93(11):2019–2024.
Szustakiewicz K, Kiersnowski A, Gazińska M, et al. Flammability,
structure and mechanical properties of PP/OMMT
nanocomposites. Polym Degrad Stab. 2011;96(3):291–294.
Stankovich S, Dikin DA, Dommett GH, et al. Graphene-based
composite materials. Nature. 2006;442(7100):282.
Dikin DA, Stankovich S, Zimney EJ, et al. Preparation and
characterization of graphene oxide paper. Nature. 2007;448
(7152):457.
Potts JR, Dreyer DR, Bielawski CW, et al. Graphene-based
polymer nanocomposites. Polymer. 2011;52(1):5–25.
Kim H, Abdala AA, Macosko CW. Graphene/polymer
nanocomposites. Macromolecules. 2010;43(16):6515–6530.
Dittrich B, Wartig K-A, Hofmann D, et al. Flame retardancy
through carbon nanomaterials: carbon black, multiwall nanotubes, expanded graphite, multi-layer graphene and graphene in
polypropylene. Polym Degrad Stab. 2013;98(8):1495–1505.
Kashiwagi T, Du F, Winey KI, et al. Flammability properties of
polymer nanocomposites with single-walled carbon nanotubes:
effects of nanotube dispersion and concentration. Polymer.
2005;46(2):471–481.
Kashiwagi T, Du F, Douglas JF, et al. Nanoparticle networks
reduce the flammability of polymer nanocomposites. Nat Mater.
2005;4(12):928.
29
[144] Kashiwagi T, Mu M, Winey K, et al. Relation between the
viscoelastic and flammability properties of polymer
nanocomposites. Polymer. 2008;49(20):4358–4368.
[145] Wen X, Wang Y, Gong J, et al. Thermal and flammability
properties of polypropylene/carbon black nanocomposites.
Polym Degrad Stab. 2012;97(5):793–801.
[146] Wang X, Kalali EN, Wan J-T, et al. Carbon-family materials for
flame retardant polymeric materials. Prog Polym Sci.
2017;69:22–46.
[147] Barus S, Zanetti M, Bracco P, et al. Influence of MWCNT
morphology on dispersion and thermal properties of polyethylene nanocomposites. Polym Degrad Stab. 2010;95
(5):756–762.
[148] Cipiriano BH, Kashiwagi T, Raghavan SR, et al. Effects of
aspect ratio of MWNT on the flammability properties of
polymer nanocomposites. Polymer. 2007;48(20):6086–6096.
[149] Villmow T, Kretzschmar B, Pötschke P. Influence of screw configuration, residence time, and specific mechanical energy in
twin-screw extrusion of polycaprolactone/multi-walled carbon
nanotube composites. Compos Sci Technol. 2010;70
(14):2045–2055.
[150] Alig I, Pötschke P, Lellinger D, et al. Establishment, morphology
and properties of carbon nanotube networks in polymer melts.
Polymer. 2012;53(1):4–28.
[151] Kasaliwal GR, Pegel S, Göldel A, et al. Analysis of agglomerate dispersion mechanisms of multiwalled carbon nanotubes during melt mixing in polycarbonate. Polymer.
2010;51(12):2708–2720.
[152] Liu H, Zhang L, Guo Y, et al. Reduction of graphene oxide to
highly conductive graphene by Lawesson’s reagent and its electrical applications [10.1039/C3TC00067B]. J Mater Chem C.
2013;1(18):3104–3109.
[153] Huang G, Gao J, Wang X, et al. How can graphene reduce the
flammability of polymer nanocomposites? Mater Lett. 2012;66
(1):187–189.
[154] Hapuarachchi TD, Peijs T, Bilotti E. Thermal degradation
and flammability behavior of polypropylene/clay/carbon
nanotube composite systems. ?Polym Adv Technol. 2013;24
(3):331–338.
[155] Marosfoi B, Garas S, Bodzay B, et al. Flame retardancy study on
magnesium hydroxide associated with clays of different morphology in polypropylene matrix? Polym Adv Technol. 2008;19
(6):693–700.