You are on page 1of 303

Accepted Manuscript

Title: A review on polymer–layered silicate nanocomposites

Authors: S. Pavlidou, C.D. Papaspyrides

PII: S0079-6700(08)00070-1
DOI: doi:10.1016/j.progpolymsci.2008.07.008
Reference: JPPS 553

To appear in: Progress in Polymer Science

Received date: 25-6-2007


Revised date: 11-7-2008
Accepted date: 25-7-2008

Please cite this article as: Pavlidou S, Papaspyrides CD, A review on


polymer–layered silicate nanocomposites, Progress in Polymer Science (2008),
doi:10.1016/j.progpolymsci.2008.07.008

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
A review on polymer–layered silicate nanocomposites

t
S. Pavlidou1, C. D. Papaspyrides 2(*)

ip
cr
1 CLOTEFI, Clothing, Textile and Fibre Technological Development,

4 El. Venizelou Str., Kallithea, Athens 176 76, Greece

us
2
Laboratory of Polymer Technology, Department of Chemical Engineering,

National Technical University of Athens, Zographou, Athens 157 80, Greece

(*)
To whom correspondence should
anbe addressed (e-mail address:
M
kp@softlab.ece.ntua.gr

____________________________________________________________________
d

ABSTRACT
te

This review reports recent advances in the field of polymer-layered silicate


p

nanocomposites. These materials have attracted both academic and industrial attention
ce

because they exhibit dramatic improvement in properties at very low filler contents.

Herein, the structure, preparation and properties of polymer-layered silicate


Ac

nanocomposites are discussed in general, and detailed examples are also drawn from

the scientific literature.

Keywords: nanocomposites, polymers, layered silicates, clays

1
Page 1 of 302
Contents

1. Introduction

2. Milestones in the research on polymer-layered silicate nanocomposites

3. Layered silicates

t
3.1. Structure and characteristics of layered silicates

ip
3.2. Organic modification of layered silicates

cr
4. Nanocomposites structures and characterization
4.1. Nanocomposite structures

us
4.2. Structural characterization of nanocomposites

5. Preparation of nanocomposites

an
5.1. Introduction
5.2. Intercalation of polymer from solution
5.3. In-situ intercalative polymerization
M
5.3.1. In-situ intercalative polymerization of thermoplastic polymers
5.3.2. In-situ intercalative polymerization of thermosetting polymers
5.4. Polymer melt intercalation
d

5.4.1. Introduction and advantages of the technique


te

5.4.2. Factors affecting polymer melt intercalation


5.4.3. Compatibility issues in the case of non-polar polymers
p

5.4.4. Degradation problems encountered during melt intercalation


ce

6. Nanocomposite Properties
6.1 Mechanical properties
Ac

6.1.1. The reinforcing mechanism of layered silicates


6.1.2. Modulus and strength
6.1.3. Toughness and strain
6.1.4. Comparison and synergistic effects of clays and conventional
reinforcements
6.2. Dynamic mechanical properties
6.3. Barrier properties
6.4. Thermal stability
6.5. Flame retardance

2
Page 2 of 302
6.5.1. Flame retardance in polymer-layered silicate nanocomposites
6.5.2. Synergism between nanocomposites and flame retardants
6.6. Heat distortion temperature
6.7. Rheological properties
6.8. Crystallinity

t
6.9. Biodegradation

ip
6.10. Photo-degradation
6.11. Optical clarity

cr
7. Nanocomposites: advantages and applications

us
8. Summary

an
M
d
p te
ce
Ac

3
Page 3 of 302
Nomenclature

AIBN N,N′-azobis(isobutyronitrile)
CEC cation exchange capacity
CX-clay X is the number of carbon atoms in clay organic modifier
d spacing between diffractional lattice planes

t
D diffusivity

ip
DETDA diethyl toluene diamine
DGEBA diglycidyl ether of bisphenol A

cr
DM dioctadecyldimethyl ammonium chloride
DMA dynamic mechanical analysis
DSC differential scanning calorimetry

us
DTGA differential thermogravimetric analysis
E′ storage modulus under bending mode
E′′ loss modulus under bending mode
EVA ethyl-vinyl-acetate copolymer

an
FTIR Fourier transform infrared spectroscopy
fwhm full width at half-maximum
G′ storage modulus under tensile mode
G′′ loss modulus under tensile mode
M
GPC gas-permeation chromatography
HDPE high density polyethylene
HDT heat distortion temperature
HRR heat release rate
d

MMT montmorillonite
NMR nuclear magnetic resonance
te

o-clay organo-modified clay


OMLS organo-modified layered silicate, organosilicate, or
organoclay
p

P permeability
PA polyamide
ce

PAA poly(acrylic acid)


PBO polybenzoxale
PCL polycaprolactone
Ac

PCN polymer-clay nanocomposite


PDMS polydimethylsiloxane
PE polyethylene
PEI poly(erther imide)
PEO poly(ethylene oxide)
PET poly(ethylene terephthalate)
PHB poly(3-hydroxybutyrate)
PHRR peak heat release rate
PI polyimide
PA polylactide
PLS polymer-layered silicate nanocomposite
PMMA polymethyl methacrylate
PP: polypropylene
PP-MA or PP-g-MA maleic anhydride-grafted polypropylene

4
Page 4 of 302
PS polystyrene
PSF polysulphone
PU polyurethane
PVA poly(vinyl acetate)
PVC poly(vinyl chloride)
PVE poly(vinyl ethylene)
PVOH poly(vinyl alcohol)
PVP poly(vinyl pyrrilidone)

t
S solubility

ip
SBS poly( styrene-butadiene-styrene)
SEA specific extinction area
Tan δ G′/G′′

cr
Tc crystallization temperature
TEM transmission electron microscopy

us
Tg glass transition temperature
TGA thermogravimetric analysis
TGAP triglycidyl p-aminophenol
TGDDM tetrafunctional tetraglycidyldiaminodiphenylmethane

an
TPO thermoplastic olefin
UP unsaturated polyester
WAXD or WAXS wide angle X-ray diffraction or scattering
XRD X-ray diffraction
M
d
p te
ce
Ac

5
Page 5 of 302
1. Introduction

Traditionally, polymeric materials have been filled with synthetic or natural

inorganic compounds in order to improve their properties, or simply to reduce cost.

t
Conventional fillers are materials in the form of particles (e.g. calcium carbonate),

ip
fibers (e.g. glass fibers) or plate-shaped particles (e.g. mica). However, although

cr
conventionally filled or reinforced polymeric materials are widely used in various

fields, it is often reported that the addition of these fillers imparts drawbacks to the

us
resulting materials, such as weight increase, brittleness and opacity [1-

an
5].Nanocomposites, on the other hand, are a new class of composites, for which at

least one dimension of the dispersed particles is in the nanometer range. Depending
M
on how many dimensions are in the nanometer range, one can distinguish

isodimensional nanoparticles when the three dimensions are on the order of


d

nanometers, nanotubes or whiskers when two dimensions are on the nanometer scale
te

and the third is larger, thus forming an elongated structure, and, finally, layered

crystals or clays, present in the form of sheets of one to a few nanometers thick and
p

hundreds to thousands nanometers in extent [1, 4].Among all the potential


ce

nanocomposite precursors, those based on clay and layered silicates have been most

widely investigated, probably because the starting clay materials are easily available
Ac

and because their intercalation chemistry has been studied for a long time [6].

Polymer-layered silicate nanocomposites, which are the subject of the present

contribution, are prepared by incorporating finely dispersed layered silicate materials

in a polymer matrix [2]. However, the nanolayers are not easily dispersed in most

polymers due to their preferred face-to-face stacking in agglomerated tactoids.

Dispersion of the tactoids into discrete monolayers is further hindered by the intrinsic

6
Page 6 of 302
incompatibility of hydrophilic layered silicates and hydrophobic engineering plastics.

Therefore, layered silicates first need to be organically modified to produce polymer-

compatible clay (organoclay). In fact, it has been well-demonstrated that the

replacement of the inorganic exchange cations in the cavities or “galleries” of the

t
native clay silicate structure by alkylammonium surfactants can compatibilize the

ip
surface chemistry of the clay and a hydrophobic polymer matrix [7].

cr
Thereafter, different approaches can be applied to incorporate the ion-exchanged

layered silicates in polymer hostsby in situ polymerization, solution intercalation or

us
simple melt mixing. In any case, nanoparticles are added to the matrix or matrix

an
precursors as 1-100 µm powders, containing associated nanoparticles. Engineering the

correct interfacial chemistry between nanoparticles and the polymer host, as described
M
previously, is critical but not sufficient to transform the micron-scale compositional

heterogeneity of the initial powder into nanoscale homogenization of nanoparticles


d

within a polymeric nanocomposite [8]. Therefore, appropriate conditions have to be


te

established during the nanocomposite preparation stage.

The resulting polymer-layered silicates hybrids possess unique properties –


p

typically not shared by their more conventional microscopic counterparts – which are
ce

attributed to their nanometer size features and the extraordinarily high surface area of

the dispersed clay [1, 4]. In fact, it is well established that dramatic improvements in
Ac

physical properties, such as tensile strength and modulus, heat distortion temperature

(HDT) and gas permeability, can be achieved by adding just a small fraction of clay to

a polymer matrix, without impairing the optical homogeneity of the material. Most

notable are the unexpected properties obtained from the addition of stiff filler to a

polymer matrix, e.g. the often reported retention (or even improvement) of the impact

strength. Since the weight fraction of the inorganic additive is typically below 10 %,

7
Page 7 of 302
the materials are also lighter than most conventional composites [2, 9-12]. These

unique properties make the nanocomposites ideal materials for products ranging from

high-barrier packaging for food and electronics to strong, heat-resistant automotive

components [11]. Additionally, polymer-layered silicate nanocomposites have been

t
proposed as model systems to examine polymer structure and dynamics in confined

ip
environments [12-13].

cr
However, despite the recent progress in polymer nanocomposite technology, there

are many fundamental questions that have not been answered. For example, how do

us
changes in polymer crystalline structure induced by the clay affect overall composite

an
properties? How does one tailor organoclay chemistry to achieve high degrees of

exfoliation reproducibility for a given polymer system? How do process parameters


M
and fabrication affect composite properties? Further research is needed that addresses

such issues [14]. The objective of this work is to review recent scientific and
d

technological advances in the field of polymer-layered silicate nanocomposite


te

materials and to develop a better understanding of how superior nanocomposites are

formed.
p
ce

2. Milestones in the research on polymer-layered silicate nanocomposites


Ac

The incorporation of layered silicates into polymer matrices has been known for

over 50 years [15]. In fact, one of the earliest systematic studies of the interaction

between a clay mineral and a macromolecule dates back to 1949, when Bower [16]

described the absorption of DNA by montmorillonite. Even in the absence of X-ray

diffraction (XRD) evidence, this finding implied insertion of the macromolecule in

the lamellar structure of the silicate. In 1950, Carter et al. [17] developed organoclays

with several organic onium bases to reinforce latex-based elastomers and in 1958, E.

8
Page 8 of 302
A. Hauser [18] was granted a patent for “clay complexes with conjugated unsaturated

aliphatic compounds of four to five carbon atoms”. Uskov [19], in 1960, found that

the softening point of poly(methyl methacrylate) derived by polymerization of methyl

methacrylate was raised by montmorillonite modified with octadecylammonium,

t
while in the following year Blumstein [20] obtained a polymer inserted in the

ip
structure of a montmorillonite by polymerizing a previously inserted vinyl monomer.

cr
Two years later, Greenland [21] used a poly(vinyl alcohol)/montmorillonite system to

show that a polymer could be directly inserted in a clay in an aqueous solution. The

us
same year, the incorporation of organoclay into a thermoplastic polyolefin matrix was

an
disclosed by Nahin and Backlund [22] of Union Oil Co. They obtained organoclay

composites with strong solvent resistance and high tensile strength by irradiation-
M
induced cross linking. However, they did not focus on the intercalation characteristics

of the organoclay or the potential properties of the composites. In 1975, Tanihara &
d

Nakagawa [23] reached a similar result by intercalating polyacrylamide and


te

poly(ethylene oxide) from an aqueous solution. In 1976 Fujiwara and Sakamoto [24]

of the Unichika Co. described the first organoclay hybrid polyamide nanocomposite.
p

However, it was not until Toyota researchers began a detailed examination of


ce

polymer-layered silicate composites that nanocomposites became more widely studied

in academic, government and industrial laboratories [25-28]. The Toyota research


Ac

group disclosed improved methods for producing nylon 6 clay nanocomposites using

in situ polymerization similar to the Unichika process. They reported that these

polymer-clay nanocomposites exhibit superior strength, modulus, heat distortion

temperature, water and gas barrier properties, with comparable impact strength as neat

nylon 6. They also reported on various other types of polymer-clay hybrid

9
Page 9 of 302
nanocomposites based on epoxy resin and polystyrene, acrylic polymer, rubber, and

polyimides formed using a similar approach.

On the other hand, work by Giannelis et al. [29-30] revealed that intercalation of

polymer chains into the galleries of an organoclay can occur spontaneously on heating

t
a mixture of polymer and silicate clay powder above the polymer glass transition or

ip
melt temperature. Once sufficient polymer mobility is achieved, chains diffuse into

cr
the host silicate clay galleries, thereby producing an expanded polymer-silicate

structure.

us
To summarize: although the intercalation chemistry of polymers when mixed with

an
appropriately modified layered silicates and synthetic layered silicates has long been

known, two major findings have stimulated the revival of interest in polymer-layered
M
silicate nanocomposite materials. First, the report from the Toyota research group on a

nylon 6/montmorillonite nanocomposite, in which very small amounts of layered


d

silicate loadings resulted in pronounced improvements of thermal and mechanical


te

properties; and second, the observation by Giannelis and his co-workers that it is

possible to melt-mix polymers with layered silicates, without the use of organic
p

solvents.
ce

Since then, the high promise for industrial applications has motivated vigorous

research, and today efforts are being conducted globally, using almost all types of
Ac

polymer matrices. In fact, nanocomposites have been demonstrated with many

thermoplastic and thermosetting polymers of different polarities including, among

others, polystyrene, polycaprolactone, polypropylene, poly(ethylene oxide), epoxy

resin, polysiloxane and polyurethane [7, 14-15, 31-33]. It must be noted, however,

that so far most of these materials have been produced only on the laboratory scale;

10
Page 10 of 302
and until a short time ago, research tended to center on proof of exfoliation of the clay

[32].

3. Layered silicates

t
ip
3.1. Structure and characteristics of layered silicates

cr
Layered silicates used in the synthesis of nanocomposites are natural or synthetic

us
minerals, consisting of very thin layers that are usually bound together with counter-

an
ions. Their basic building blocks are tetrahedral sheets in which silicon is surrounded

by 4 oxygen atoms, and octahedral sheets in which a metal like aluminum is


M
surrounded by 8 oxygen atoms. Therefore, in 1:1 layered structures (e.g. in kaolinite)

a tetrahedral sheet is fused with an octahedral sheet, whereby the oxygen atoms are
d

shared [34].
te

On the other hand, the crystal lattice of 2:1 layered silicates (or 2:1

phyllosilicates), consists of two-dimensional layers where a central octahedral sheet


p

of alumina is fused to two external silica tetrahedra by the tip, so that the oxygen ions
ce

of the octahedral sheet also belong to the tetrahedral sheets, as shown in Fig. 1. The

layer thickness is around 1 nm and the lateral dimensions may vary from 300 Å to
Ac

several microns, and even larger, depending on the particulate silicate, the source of

the clay and the method of preparation (e.g. clays prepared by milling typically have

lateral platelet dimensions of approximately 0.1-1.0 µm). Therefore, the aspect ratio

of these layers (ratio length/thickness) is particularly high, with values greater than

1000 [1, 35-37].

11
Page 11 of 302
The basic 2:1 structure with silicon in the tetrahedral sheets and aluminum in the

octahedral sheet, without any substitution of atoms, is called pyrophillite. Since the

layers do not expand in water, pyrophyllite has only an external surface area and

essentially no internal one. When silicon in the tetrahedral sheet is substituted by

t
aluminum, the resulting structure is called mica. Due to this substitution the mineral is

ip
characterized by a negative surface charge, which is balanced by interlayer potassium

cr
cations. However, because the size of the potassium ions matches the hexagonal hole

created by the Si/Al tetrahedral layer, it is able to fit very tightly between the layers.

us
Consequently, the interlayers collapse and the layers are held together by the

an
electrostatic attraction between the negatively charged tetrahedral layer and the

potassium cations. Therefore, micas do not swell in water and, like pyrophyllite, have
M
no internal surface area [38]. On the other hand, if in the original pyrophyllite

structure the trivalent Al-cation in the octahedral layer is partially substituted by the
d

divalent Mg-cation, the structure of montmorillonite is formed, which is the best-


te

known member of a group of clay minerals, called “smectites” or “smectite clays”. In

this case the overall negative charge is balanced by sodium and calcium ions, which
p

exist hydrated in the interlayer [39]. A particular feature of the resulting structure is
ce

that, since these ions do not fit in the tetrahedral layer, as in mica, and the layers are

held together by relatively weak forces, water and other polar molecules can enter
Ac

between the unit layers, causing the lattice to expand [33].

Along with montmorillonite, hectorite and saponite are the layered silicates that

are most commonly used in nanocomposite materials. Their chemical formula is given

in Table 1 [1].

12
Page 12 of 302
The reason why these materials have received a great deal of attention recently, as

reinforcing materials for polymers, is their potentially high aspect ratio and the unique

intercalation/exfoliation characteristicsthat will be discussed later [15].

In general, it is well established that structural perfection is more and more nearly

t
reached as the reinforcing elements become smaller and that the ultimate properties of

ip
reinforcing composite elements may be expected if their dimensions reach atomic or

cr
molecular levels. For example, carbon nanotubes display the so far highest known

values of elastic modulus (ca. 1.7 TPa!). Similarly, individual clay sheets, being only

us
1 nm thick, display a perfect crystalline structure. However, the smaller the

an
reinforcing elements are, the larger is their internal surface and hence their tendency

to agglomerate rather than to disperse homogeneously in a matrix [2]. In fact, the


M
silicate layers have the tendency to organize themselves to form stacks with a regular

van der Waals gap between them, called an “interlayer” or “gallery” [1, 35-36]. The
d

interlayer dimension is determined by the crystal structure of the silicate (for


te

dehydrated Na-montmorillonite this dimension is approximately 1 nm) [37].

Analysis of layered silicates has shown that there are several levels of
p

organization within the clay minerals. The smallest particles, primary particles, are on
ce

the order of 10 nm and are composed of stacks of parallel lamellae. Micro-aggregates

are formed by lateral joining of several primary particles, and aggregates are
Ac

composed of several primary particles and micro-aggregates [40].

3.2. Organic modification of layered silicates

Since, in their pristine state layered silicates are only miscible with hydrophilic

polymers, such as poly(ethylene oxide) and poly(vinyl alcohol), in order to render

13
Page 13 of 302
them miscible with other polymers, one must exchange the alkali counter-ions with

cationic-organic surfactant, as shown in Fig. 2. Alkylammonium ions are mostly used,

although other “onium” salts can be used, such as sulfonium and phosphonium [1, 39,

41]. This can be readily achieved through ion-exchange reactions that render the clay

t
organophilic [42]. In order to obtain the exchange of the onium ions with the cations

ip
in the galleries, water swelling of the silicate is needed. For this reason alkali cations

cr
are preferred in the galleries because 2-valent and higher valent cations prevent

swelling by water. Indeed, the hydrate formation of monovalent intergallery cations

us
is the driving force for water swelling. Natural clays may contain divalent cations

an
such as calcium and require exchange procedures with sodium prior to further

treatment with onium salts [41]. The alkali cations, as they are not structural, can be
M
easily replaced by other positively charged atoms or molecules, and thus are called

exchangeable cations [43].


d

The organic cations lower the surface energy of the silicate surface and improve
te

wetting with the polymer matrix [4, 42]. Moreover, the long organic chains of such

surfactants, with positively charged ends, are tethered to the surface of the negatively
p

charged silicate layers, resulting in an increase of the gallery height [44]. It then
ce

becomes possible for organic species (i.e. polymers or prepolymers) to diffuse

between the layers and eventually separate them [42, 45]. Sometimes, the
Ac

alkylammonium cations may even provide functional groups that can react with the

polymer or initiate polymerization of monomers [33]. The microchemical

environment in the galleries is, therefore, appropriate to the intercalation of polymer

molecules [46]. Conclusively, the surface modification both increases the basal

spacing of clays and serves as a compatibilizer between the hydrophilic clay and the

hydrophobic polymer [45].

14
Page 14 of 302
The excess negative charge of layered silicates and their capability to exchange

ions can be quantified by a specific property known as the cation-exchange capacity

(CEC) and expressed in meq/g [1, 39]. This property is highly dependent on the

nature of the isomorphous substitutions in the tetrahedral and octahedral layers and

t
therefore on the nature of the soil where the clay was formed. This explains, for

ip
example, why montmorillonites from different origins show differences in CEC,

cr
ranging from approximately 0.9 to 1.2 meq/g [39, 42]. The charge of the layer is not

locally constant, as it varies from layer to layer, and must rather be considered as an

us
average value over the whole crystal. Proportionally, even if a small part of the charge

an
balancing cations is located on the external crystallite surface, the majority of these

exchangeable cations are located inside the galleries [1].


M
Depending on the functionality, packing density, and length of the organic

modifiers, the organo-modified layered silicates (OMLSs, organosilicates or


d

organoclays) may be engineered to optimize their compatibility with a given polymer


te

[43, 47]. It is worth noticing that, on the basis of the CEC of the clay, the content of

the surfactant is usually about 35-45 wt % [48]. Actually, one way to measure the clay
p

CEC is by determining the amount of alkylammonium salt retained by the


ce

organoclays. That is, dried clays that have been subjected to organo-modification

along with a sample of the corresponding untreated clay are ignited at 1000 °C. From
Ac

the differences in the loss on ignition of the sample and blank and the molecular

weight of the alkylammonium salt, the milliequivalents of the organic substance

retained by the clays are calculated and those values are taken as their CEC.

Alternatively, chemical analysis of the clay can also be applied for CEC

determination [42].

15
Page 15 of 302
In general, the longer the surfactant chain length, and the higher the charge

density of the clay, the further apart the clay layers will be forced. This is expected

since both of these parameters contribute to increasing the volume occupied by the

intragallery surfactant [7]. For example, Wang et al. prepared organoclays with

t
different alkylammonium chain lengths and also used an organophilic clay, Cloisite

ip
20A, which has two long alkyl chains. They found that the interlayer spacing

cr
increases with the increase in size of alkylamine chain length. The interlayer spacings

of C12M, C16M, C18M (with 12, 16 and 18 carbon atoms in the alkylammonium

us
chain) and 20A were 1.36, 1.79, 1.85 and 2.47 nm respectively [49].

an
However, the interlayer distance also depends on the way the onium ion chains

organize themselves in the organoclay. In order to describe the structure of the


M
interlayer in organoclays, one has to know that, as the negative charge originates in

the silicate layer, the cationic head group of the alkylammonium molecule
d

preferentially resides at the layer surface, leaving the organic tail radiating away from
te

the surface [1, 41].

Initially, the orientation of surfactant chains was deduced from infrared and X-ray
p

diffraction (XRD) measurements, according to which the organic chains have been
ce

long thought to lie either parallel to the silicate layer, forming mono or bilayers or,

depending on the packing density and the chain length, to radiate away from the
Ac

surface, forming mono or even a bimolecular tilted “paraffinic” arrangement, as

shown in Fig. 3 [1, 7].

A more realistic description has been proposed by Vaia et al., based on FTIR

experiments. By monitoring frequency shifts of the asymmetric CH2 stretching and

bending vibrations, they showed that alkyl chains can vary from liquid-like to solid-

like, with the liquid-like structure dominating as the interlayer density or chain length

16
Page 16 of 302
decreases, or as the temperature increases. When the available surface area per

molecule is within a certain range, the chains are not completely disordered but retain

some orientational order similar to that in the liquid crystalline state (Fig. 4). As the

chain length increases, the interlayer structure appears to evolve in a stepwise fashion,

t
from a disordered to more ordered monolayer, then “jumping” to a more disordered

ip
pseudo-bilayer. In addition, an NMR study reported by Wang et al. indicated the

cr
coexistence of ordered trans and disordered gauche conformations [50].

Fornes et al. conducted WAXS scans for different organoclays and for pristine

us
sodium montmorillonite, and plotted basal spacing values obtained versus the mass of

an
organic component per unit mass of inorganic MMT for each organoclay, as shown in

Fig. 5. They further analyzed the data by expressing the mass of organic material per
M
unit volume of gallery, or gallery density, as:

mass organic 1 (mass organic / mass MMT )


ρ gallery = =
d

gallery volume (d - d0 )  area / side 


 
 mass MMT 
te

where d - d0 is the gallery height, as indicated by Fig. 5. The slope of the linear
p

relation between gallery height and the second quantity in the above equation gives
ce

the density of the organic material in the gallery, ρgallery. Forcing the fit through an

intercept d0 = 9.6 Å, which is the basal spacing of pristine sodium montmorillonite,


Ac

produces a calculated density of 1.07 g/cm3. This range of densities encompasses

what might be expected for organic liquids or solids. Therefore, the densities

calculated in that work agree with conclusions made by Vaia et al. about the

conformation and structure of the organic interlayer of similar organoclays,

suggesting that organoclay galleries exhibit molecular environments ranging from

solid- to liquid-like [51].

17
Page 17 of 302
Summarizing this section, there are two particular characteristics of layered

silicates that are exploited in polymer-layered silicate nanocomposites. The first is the

ability of the silicate particles to disperse into individual layers. Since dispersing a

layered silicate can be pictured like opening a book, an aspect ratio as high as 1000

t
for fully dispersed individual layers can be obtained (contrast that to an aspect ratio of

ip
about 10 for undispersed or poorly dispersed particles). The second characteristic is

cr
the ability to fine-tune their surface chemistry through ion exchange reactions with

organic and inorganic cations. These two characteristics are, of course, interrelated

us
since the degree of dispersion in a given matrix that, in turn, determines aspect ratio,

an
depends on the interlayer cation [4, 40].
M
4. Nanocomposite structures and characterization

4.1. Nanocomposite structures


d
te

Any physical mixture of a polymer and silicate (or inorganic material in general)
p

does not necessarily form a nanocomposite. The situation is analogous to polymer


ce

blends. In most cases, separation into discrete phases normally takes place. In

immiscible systems, the poor physical attraction between the organic and the
Ac

inorganic components leads to relatively poor mechanical properties. Furthermore,

particle agglomeration tends to reduce strength and produce weaker materials [4].

Thus, when the polymer is unable to intercalate between the silicate sheets, a phase-

separated composite is obtained, whose properties are in the same range as for

traditional microcomposites [1, 35].

Beyond this traditional class of polymer-filler composites, two types of

nanocomposites can be obtained, depending on the preparation method and the nature

18
Page 18 of 302
of the components used, including polymer matrix, layered silicate and organic cation

[1, 35]. These two types of polymer-layered silicate nanocomposites are depicted in

Fig. 6 [50].

Intercalated structures are formed when a single (or sometimes more) extended

t
polymer chain is intercalated between the silicate layers. The result is a well ordered

ip
multilayer structure of alternating polymeric and inorganic layers, with a repeat

cr
distance between them. Intercalation causes less than 20-30 Ǻ separation between the

platelets [1, 4, 33, 35, 44, 52].

us
On the other hand, exfoliated or delaminated structures are obtained when the clay

an
layers are well separated from one another and individually dispersed in the

continuous polymer matrix [1, 4, 33, 44]. In this case, the polymer separates the clay
M
platelets by 80-100 Ǻ or more [52]. That is, the interlayer expansion is comparable to

the radius of gyration of the polymer rather than that of an extended chain, as in the
d

case of intercalated hybrids [4].


te

The exfoliation or delamination configuration is of particular interest because it

maximizes the polymer-clay interactions making the entire surface of layers available
p

for the polymer. This should lead to the most significant changes in mechanical and
ce

physical properties [35]. In fact, it is generally accepted that exfoliated systems give

better mechanical properties than intercalated ones [5, 33]. The complete dispersion
Ac

of clay nanolayers in a polymer optimizes the number of available reinforcing

elements for carrying an applied load and deflecting cracks. The coupling between the

tremendous surface area of the clay and the polymer matrix facilitates stress transfer

to the reinforcement phase, allowing for mechanical property improvements [35, 53].

However, it is not easy to achieve complete exfoliation of clays and, indeed with

few exceptions, the majority of the polymer nanocomposites reported in the literature

19
Page 19 of 302
were found to have intercalated or mixed intercalated-exfoliated nanostructures [33].

This is because the silicate layers are highly anisotropic, with lateral dimensions

ranging from 100 to 1000 nm, and even when separated by large distances (i.e. when

delaminated) cannot be placed completely randomly in the sea of polymer.

t
Furthermore, the majority of the polymer chains in the hybrids are tethered to the

ip
surface of the silicate layers. Thus, it can be expected that there are domains in these

cr
materials, even above the melting temperature of the constituent polymers, wherein

some long-range order is preserved and the silicate layers are oriented in some

us
preferred direction. This long-range order and domain structure is likely to become

an
better defined at the higher silicate contents, where the geometrically imposed mean

distance between the layers becomes less than the lateral dimensions of the silicate
M
layers, thus forcing some preferential orientation between the layers. However, there

might be considerable polydispersity effects in terms of the orientation and the


d

distance between the silicate layers. Many such randomly oriented grains make up the
te

entire sample leading to the presence of disordered material. Thus, in general the

material possesses a layered structure, with grains wherein the silicate layers are
p

oriented in a preferred direction leading to the presence of grain boundaries and


ce

concomitant defects [54].

At this point, it is worth mentioning six interrelated structural characteristics,


Ac

distinguishing polymer-silicate nanocomposites from conventional filled systems.

These characteristics, which are attributed to the nanoscopic dimensions and the

extreme aspect ratios of layered silicates, are [8]:

• Low percolation threshold (ca. 0,1-2 vol %)

• Particle-particle correlation (orientation and position) arising at low volume

fractions

20
Page 20 of 302
• Large number of particles per particle volume (106-108 particles/µm3)

• Extensive interfacial area per volume of particles (103-104 m2/ml)

• Short distances between particles (10-50 nm at φ ~ 1-8 vol %)

• Comparable size scales among the rigid nanoparticle inclusions, distance between

t
particles, and the relaxation volume of polymer chains.

ip
cr
4.2. Nanocomposite Structural Characterization

us
Two complementary techniques are generally used to characterize the structures

an
of nanocomposites: X-ray diffraction (XRD) and transmission electron microscopy

(TEM) [1, 35, 55-56].


M
Due to its ease of use and availability, X-ray diffraction is most commonly used to

probe the nanocomposite structure and occasionally to study the kinetics of polymer
d

melt intercalation [55]. This technique allows the determination of the spaces between
te

structural layers of the silicate utilizing Bragg’s law: sinθ = nλ / 2d, where λ

corresponds to the wave length of the X-ray radiation used in the diffraction
p

experiment, d the spacing between diffractional lattice planes and θ is the measured
ce

diffraction angle or glancing angle [1, 56]. By monitoring the position, shape and

intensity of the basal reflections from the distributed silicate layers, the
Ac

nanocomposite structure may be identified [55].

For immiscible polymer/OMLS mixtures, the structure of the silicate is not

affected, and thus, the characteristics of the OMLS basal reflections do not change.

On the other hand, in comparison with the spacing of the organoclay used, the

intercalation of the polymer chains increases the interlayer spacing, leading to a shift

of the diffraction peak towards lower angle, according to Bragg’s law. In such

21
Page 21 of 302
intercalated nanocomposites, the repetitive multilayer structure is well preserved,

allowing the interlayer spacing to be determined (Fig. 7). In contrast, the extensive

layer separation associated with exfoliated structures disrupts the coherent layer

stacking and results in a featureless diffraction pattern. Thus, for exfoliated structures

t
no more diffraction peaks are visible in the XRD diffractograms either because of a

ip
much too large spacing between the layers (i.e. exceeding 8 nm in the case of ordered

cr
exfoliated structure) or because the nanocomposite does not present ordering [1, 35,

47].

us
The influence of polymer intercalation on the order of the OMLS layers may be

an
monitored by changes in the full-width-at-half-maximum (fwhm) and intensity of the

basal reflections. An increase in the degree of coherent layer stacking (i.e. a more
M
ordered system) results in a relative decrease in the fwhm of the basal reflections upon

hybrid formation. On the other hand, a decrease in the degree of coherent layer
d

stacking (i.e. a more disordered system) results in peak broadening and intensity loss
te

[47].

However, although XRD offers a conventional method to determine the interlayer


p

spacing of the silicate layers in the original layered silicates and the intercalated
ce

nanocomposites (within 1-4 nm), little can be said about the spatial distribution of the

silicate layers or any structural inhomogeneities in nanocomposites. Additionally,


Ac

some layered silicates initially do not exhibit well-defined basal reflections. Thus,

peak broadening and intensity decreases are very difficult to study systematically.

Therefore, conclusions concerning the mechanism of nanocomposite formation and

structure based solely on XRD patterns are only tentative. On the other hand, TEM

allows a qualitative understanding of the internal structure and can directly provide

22
Page 22 of 302
information in real space, in a localized area, on morphology and defect structures

[57-58].

Since the silicate layers are composed of heavier elements (Al, Si, O) than the

interlayer and surrounding matrix (C, H, N), they appear darker in bright-field images.

t
Therefore, when nanocomposites are formed, the intersections of the silicate sheets

ip
are seen as dark lines which are the cross sections of the silicate layers, measuring 1

cr
nm thick. However, special care must be exercised to guarantee a representative

cross-section of the sample [56-57]. Figure 8 shows the TEM micrograps obtained for

us
an intercalated and an exfoliated nanocomposite. As already mentioned, besides these

an
two well defined structures other intermediate organizations can exist presenting both

intercalation and exfoliation. In this case, a broadening of the diffraction peak is often
M
observed and one must rely on TEM observation to define the overall structure [1].
13
Characterization of polymer/layered silicate nanocomposites by C solid-state
d

nuclear magnetic resonance (13C NMR) has also been proposed. VanderHart et al.
te

first used this technique as a tool for gaining greater insight about the morphology,

surface chemistry, and to a very limited extent, the dynamics of exfoliated polymer
p

clay nanocomposites. They were especially interested in developing NMR methods to


ce

quantify the level of clay exfoliation, a very important facet of nanocomposite

characterization [59]. The main objective in solid-state NMR measurement is to


Ac

connect the measured longitudinal relaxations, T1s, of proton (and 13C nuclei) with the

quality of clay dispersion [55].

The surfaces of naturally occurring layered silicates such as MMT are mainly

made of tetrahedral silica, while the central plane of the layers contains octahedrally

coordinated Al3+ with frequent non-stoichiometric substitutions, where an Al3+ is

replaced by Mg2+ and, somewhat less frequently, by Fe3+. The concentration of the

23
Page 23 of 302
later ion is very important because Fe3+ is strongly paramagnetic in this distorted

octahedral environment. Typical concentrations of Fe3+ in naturally occurring clays

produce nearest-neighbor Fe-Fe distances of about 1.0-1.4 nm, and at such distances,

the spin-exchange interaction between the unpaired electrons on different Fe atoms is

t
expected to produce magnetic fluctuations in the vicinity of the Larmor frequencies

ip
13
for protons or C nuclei. The spectral density of these fluctuations is important

cr
because the T1H of protons (and 13C nuclei) within about 1nm of the clay surface can

be directly shortened. For protons, if that mechanism is efficient, relaxation will also

us
propagate into the bulk of the polymer by spin diffusion. Thus the paramagnetically

an
induced relaxation will influence the overall measured T1H to an extent that will

depend both on the Fe concentration in the clay layer and, more importantly, on the
M
average distances between clay layers. The latter dependence suggests a potential

relationship between measured T1H values and the quality of the clay dispersion. If the
d

clay particles are stacked and poorly dispersed in the polymer matrix, the average
te

distances between polymer/clay interfaces are greater, and the average paramagnetic

contribution to T1H is weaker. VanderHart et al. also employed the same arguments in
p

order to understand the stability of a particular OMLS under different processing


ce

conditions [55, 60-61].

Some authors also used Fourier transform infrared spectroscopy (FTIR) to


Ac

elucidate the structure of the nanocomposites [62-63]. FTIR may be able to identify

differences between the bonding in a mixture and the bonding in a related

nanocomposite, but as these variations are minute, even when intercalation has taken

place, at present FTIR is an unreliable method of characterization in most cases [56].

Finally, differential scanning calorimetry (DSC) provides further information

concerning intercalation. The many interactions the intercalated chains of the polymer

24
Page 24 of 302
form with the host species greatly reduce their rotational and translational mobility.

The situation is similar to that in a reticulated polymer, where restrictions on its

mobility increase its glass transition temperature (Tg). A similar increase is anticipated

to occur in a nanocomposite due to elevation of the energy threshold needed for the

t
transition. This effect is readily detected by DSC. Fig. 9 presents DSC traces of

ip
polystyrene (PS), a PS/OMLS mixture and an intercalated PS/OMLS nanocomposite.

cr
The PS and PS/OMLS mixture curves clearly display the characteristic peak due to

glass transition of the polymer. The presence of this peak in the mixture is evidence of

us
the absence of interactions between the organic and the inorganic phases. The

an
transition is absent in the nanocomposite curve and in fact occurs at temperatures

higher than those shown in Fig. 9. In addition to being an interesting analytical datum,
M
the considerable increase in Tg is an important property of these materials that enables

them to be employed at higher temperatures compared with the original polymer and
d

thus extends their fields of application [41]. To date,the aforementioned subsidiary


te

methods have only been used to confirm the evidence from the primary methods .

However, building a clearer picture of the changes that occur when a nanocomposite
p

forms is important, as it not only helps to characterize the material, but in principle
ce

could indicate novel methods of synthesis [56].

Concerning the evaluation of other nanocomposites structural characteristics, it


Ac

should be noted that the amount of clay present in a sample may be estimated, as for

conventional composites, i.e. by placing pre-dried nanocomposite pellets in a furnace

at ca. 900 °C for approximately 45 min. The resulting ash is then weighed and

corrected for loss of structural water [64].

However, unlike conventional fiber composites, the determination of filler aspect

ratio for layered silicate nanocomposites is not straightforward. Good estimates

25
Page 25 of 302
require a thorough analysis of TEM photomicrographs at different magnifications.

Figure 10 depicts various complications of calculating an aspect ratio from TEM

photomicrographs that arise from variations in both length/diameter, and thickness.

Clay platelets intrinsically have a distribution of lateral dimensions. The recovery,

t
refinement, chemical treatment, and post-treatment of these clays may contribute to

ip
the variation in filler geometry. Furthermore, extrusion of these clays with polymer

cr
and any additional melt processing steps that follow, e.g.. injection molding, will

amplify the range of particle shapes and sizes, particularly when the layered silicate is

us
not completely exfoliated, as illustrated in Fig. 10. Finally, microtoming of the

an
nanocomposite sample into thin sections for TEM analysis will also result in an

apparent distribution of observed particle sizes even if all disk-like platelets were the
M
same size [64].

It becomes obvious from this section that a major issue when synthesizing
d

polymer-layered silicate nanocomposites is the characterization of the product. In fact,


te

many of the studies conducted so far in this field are solely dedicated to structural

characterization of the nanocomposites, without reporting properties of the products.


p
ce

5. Preparation of nanocomposites
Ac

5.1. Introduction

At present there are four principal methods for producing polymer-layered silicate

nanocomposites: (1) in situ template synthesis, (2) intercalation of polymer or pre-

polymer from solution, (3) in situ intercalative polymerization and (4) melt

intercalation [1, 14, 35, 44, 52, 65].

26
Page 26 of 302
Template synthesis (sol-gel technology). In this technique, the clay minerals are

synthesized within the polymer matrix, using an aqueous solution (or gel) containing

the polymer and the silicate building blocks. As precursors for the clay silica sol,

magnesium hydroxide sol and lithium fluoride are used. During the process, the

t
polymer aids the nucleation and growth of the inorganic host crystals and gets trapped

ip
within the layers as they grow. Although theoretically this method has the potential of

cr
promoting the dispersion of the silicate layers in a one-step process, without needing

the presence of the onium ion, it presents serious disadvantages. First of all, the

us
synthesis of clay minerals generally requires high temperatures, which decompose the

an
polymers. An exception is the synthesis of hectorite-type clay minerals which can be

performed under relatively mild conditions. Another problem is the aggregation


M
tendency of the growing silicate layers. Therefore, this technique, although widely

used for the synthesis of double-layer hydroxide-based nanocomposites, is far less


d

developed for layered silicates and will not be considered in the following discussion.
te

However, it should be mentioned that several workers have successfully applied it for

the preparation of nanocomposite materials. For example, Carrado et al. and Carrado
p

and Xu synthesized hectorites from gels consisting of silica, magnesium hydroxide,


ce

lithium fluoride and polymers like poly(vinyl alcohol), polyaniline and

polyacrylonitrile. Evidently, some silicate layers aggregated, but most of them


Ac

remained uniformly distributed in the polymer matrix [1, 3, 41]

Intercalation of polymer or pre-polymer from solution. Following this technique,

the layered silicate is exfoliated into single layers using a solvent in which the

polymer (or prepolymer in case of insoluble polymers, such as polyimide) is soluble.

It is well known that such layered silicates, owing to the weak forces that stack the

layers together can be easily dispersed in an adequate solvent. After the organoclay

27
Page 27 of 302
has swollen in the solvent, the polymer is added to the solution and intercalates

between the clay layers. The final step consists of removing the solvent, either by

vaporization, usually under vacuum, or by precipitation. Upon solvent removal the

sheets reassemble, sandwiching the polymer to form, a nanocomposite structure.

t
Under this process are also gathered the nanocomposites obtained through emulsion

ip
polymerization where the layered silicate is dispersed in the aqueous phase. The

cr
major advantage of this method is that intercalated nanocomposites can be

synthesized that are based on polymers with low or even no polarity. However, the

us
solvent approach is difficult to apply in industry owing to problems associated with

an
the use of large quantities of solvents [1, 35].

In situ intercalative polymerization. In situ-polymerization was the first method


M
used to synthesize polymer-clay nanocomposites based on polyamide 6. In this

technique, the modified layered silicate is swollen by a liquid monomer or a monomer


d

solution. The monomer migrates into the galleries of the layered silicate, so that the
te

polymerization reaction can occur between the intercalated sheets. The reaction can

be initiated either by heat or radiation, by the diffusion of a suitable initiator or by an


p

organic initiator or catalyst fixed through cationic exchange inside the interlayer
ce

before the swelling step by the monomer. Polymerization produces long-chain

polymers within the clay galleries. Under conditions in which intra- and extra-gallery
Ac

polymerization rates are properly balanced, the clay layers are delaminated and the

resulting material possesses a disordered structure [1, 35, 37].

Melt intercalation: This technique consists of blending the layered silicate with

the polymer matrix in the molten state. Under such conditions – if the layer surfaces

are sufficiently compatible with the chosen polymer – the polymer can crawl into the

28
Page 28 of 302
interlayer space and form either an intercalated or an exfoliated nanocomposite [1, 35,

37].

Among the aforementioned methods, in situ polymerization and melt intercalation

are considered as commercially attractive approaches for preparing polymer/clay

t
nanocomposites. Melt intercalation, in particular, is especially of practical interest,

ip
since it presents significant advantages that will be discussed in the corresponding

cr
paragraph.

us
5.2. Intercalation of Polymer from Solution

an
Intercalation of a polymer from a solution is a two-stage process in which the
M
polymer replaces an appropriate, previously intercalated solvent, as shown in Fig. 11.

Such a replacement requires a negative variation in the Gibbs free energy. It is


d

thought that the diminished entropy due to the confinement of the polymer is
te

compensated by an increase due to desorption of intercalated solvent molecules. In

other words, the entropy gained by desorption of solvent molecules is the driving
p

force for polymer intercalation from solution [66-71].


ce

Even though this technique has been mostly used with water soluble polymers,

such as PEO, PVE, PVP and PAA [3, 21, 72-77], intercalation from non-aqueous
Ac

solutions has also been reported [78-81]. For example, HDPE-based nanocomposites

have been prepared by dissolving HDPE in a mixture of xylene and benzonitrile with

dispersed OMLS. The nanocomposite was then recovered by precipitation from THF

[79]. PS/OMLS exfoliated nanocomposites have also been prepared by the solution

intercalation technique, by mixing pure PS and organophilic clay with adsorbed cetyl

pyridium chloride [82]. Similarly, several studies have focused on the preparation of

29
Page 29 of 302
PLA-layered silicate nanocomposites using intercalation from solution. The first

attempts by Ogata [78], involved dissolving the polymer in hot chloroform in the

presence of organo-modified MMT. However, TEM and WAXD analyses revealed

that only microcomposites were formed and that an intercalated morphology was not

t
achieved. In a later study, Krikorian et al. [83] prepared PLA nanocomposites using

ip
dichloromethane as the polymer solvent and as the OMLS dispersion medium. The

cr
authors obtained intercalated or exfoliated nanocomposites, depending on the type of

OMLS used. That is, exfoliated nanocomposites were formed when diols were present

us
in the organic modifier of the clay, due to the favorable enthalpic interaction between

an
these diols and the C=O bonds in the PLA backbone. Chang et al. [84] reported the

preparation of PLA-based nanocomposites with different kinds of OMLS via solution


M
intercalation using N,N′-dimethylacetamide (DMA).

In the case of polymeric materials that are infusible and insoluble even in organic
d

solvents, the only possible route to produce nanocomposites with this method is to use
te

polymeric precursors that can be intercalated in the layered silicate and then thermally

or chemically converted to the desired polymer [1, 85].


p

Summarizing the above: although a number of nanocomposites have been


ce

produced by intercalation from solution (representative examples are presented in

Table 2), it is important to note that, in using this method, intercalation only occurs
Ac

for certain polymer/clay/solvent systems, meaning that for a given polymer one has to

find the right clay, organic modifier and solvents [1, 50]. Moreover, from the

industrial point of view, this method may involve the copious use of organic solvents,

which is usually environmentally unfriendly and economically prohibitive [50].

5.3. In-situ intercalative polymerization

30
Page 30 of 302
5.3.1. In-situ intercalative polymerization of thermoplastic polymers

The Toyota research group first reported the ability of α,ω-amino acid (COOH-

(CH2)n-1-NH2+, with n=2, 3, 4, 5, 6, 8, 11, 12, 18) modified Na+-MMT to be swollen

t
by ε-caprolactam monomer at 100 °C and subsequently initiatering opening

ip
polymerization to obtain PA6/MMT nanocomposites [25]. The number of carbon

cr
atoms in the α,ω amino acid was found to have a strong effect on the swelling

behavior as reported in Fig. 12, indicating that the extent of intercalation of ε-

us
caprolactam monomer is high when the number of carbon atoms in the ω-amino acid

an
is large [93]. Moreover, it was found from a comparison of different types of

inorganic silicates that clays having higher CEC lead to more efficient exfoliation of
M
the silicate platelets [51]. Figure 13 represents the conceptual view of the swelling

behavior of α,ω-amino acid modified Na+-MMT by ε-caprolactam [93]. In a typical


d

synthesis, 12-aminolauric acid-modified MMT (12-MMT) was mixed with ε-


te

caprolactam and the mixture was heated at 250–270 °C for 48 h to polymerize ε-

caprolactam, using 12-MMT as a catalyst (when the relative amount of 12-MMT in


p

the mixture was less than 8 wt %, 6-aminocaproic acid was added as a polymerization
ce

accelerator and the heating profile was slightly modified). Depending on the amount

of 12-MMT introduced, either exfoliated (for less then 15 wt %) or intercalated


Ac

structures (from 15 to 70 wt %) were obtained, as evidenced by XRD and TEM.

Comparison of the titrated amounts of COOH and NH2 end groups present in the

synthesized nanocomposites with given values, such as the CEC of the

montmorillonite used (119 meq/100 g), have led to the conclusion that the COOH end

groups present along the 12-MMT surface are responsible for the polymerization

initiation [25].

31
Page 31 of 302
Further work demonstrated that intercalative polymerization of ε–caprolactam

could be realized without modifying the MMT surface. Indeed, this monomer was

able to directly intercalate the Na+-MMT in water in the presence of hydrochloric

acid, as proved by the increase in interlayer spacing from 10 Å to 15.1 Å. At high

t
temperature (200 °C), in the presence of excess ε-caprolactam, the clay so modified

ip
can be swollen again, allowing the ring opening polymerization to proceed at 260 °C

cr
when 6-aminocaproic acid is added as an accelerator. The resulting composite does

not present a diffraction peak in XRD, and TEM observation agrees with a molecular

us
dispersion of the silicate sheets [94].

an
In attempts to carry out the whole synthesis in one pot, the system proved to be

sensitive to the nature of the acid used to promote the intercalation of ε-caprolactam.
M
Table 3 gives results obtained for different acids in relation to the intensity (Im) of the

XRD intercalation peak that might be present in the nanocomposites obtained (Fig.
d

14). These results show that, for unclear reasons, only phosphoric acid allows for the
te

preparation of a truly exfoliated nanocomposite, the other acids tending to promote

the formation of partially exfoliated-partially intercalated structures. One can also


p

point out that an intercalated structure is obtained even if no acid is added [95].
ce

Another polyamide, nylon 12, has also been reported to form nanocomposites via

in situ intercalative polymerization. Reichert et al. [96] used 12-aminolauric acid


Ac

(ALA) as both the layered silicate modifier and the monomer. They first studied by

XRD the dependence of the clay swelling process on ALA concentration in HCl, and

found that it can be separated in two regimes: a cation-exchange of inorganic cations

by protonated ALA at low ALA concentration and a further diffusion of zwitterionic

12-aminolauric acid into the interlayer space, when the ALA concentration exceeds

the amount of HCl in the medium (Figs. 15 and 16). The swelling was found to be

32
Page 32 of 302
independent of the swelling temperature, the layered silicate concentration and the

type of acid used to protonate ALA (HCl, H2SO4, H3PO4). ALA was then

polymerized at high temperature (280 °C) and under elevated pressure (ca, 20 bar)

with both types of swollen clay. XRD and TEM, coupled with energy dispersive X-

t
ray (EDX), as well as atomic force microscopy (AFM), confirmed that the resulting

ip
structures were partially exfoliated and otherwise intercalated nanocomposites.

cr
However, although in situ polymerization was successfully applied for the

preparation of PA6 and PA12/clay nanocomposites, few publications focused on the

us
preparation of polyamide from diamine and diacid. In one of these studies, Wu et al.

an
[97] investigated the preparation of PA1012 nanocomposite by polycondensation

polymerization. A dispersion of organoclay in absolute alcohol was added to 1,10


M
diaminodecane in absolute alcohol. Then, this mixture was added to an absolute

alcohol solution of 1,10 decanedicarboxylic acid under vigorous stirring, resulting in


d

the immediate precipitation of a diaminodecane-decanedicarboxylic acid salt. The salt


te

was recrystallized from a mixture of alcohol and water and was obtained as a white

powder. It was then added with a slight excess of diaminodecane to a U-shaped glass
p

tube which was purged with nitrogen before the reaction. The tube was immersed in
ce

an oil bath and the temperature was quickly raised to 200 °C to start the reaction.

After maintaining the autoclave for 2 h at 200 °C, the temperature was increased to
Ac

215 °C and held for 1.5 h under these conditions. The glass tube was flushed with

nitrogen each time to remove the water produced in polycondensation. In the last step

a vacuum (< 0.1 atm) was applied to remove the water and residual monomer, the

temperature was increased to 225 °C, and the reaction was continued for another 2 h.

The glass tube was then cooled to room temperature and the resulting

33
Page 33 of 302
PA1012/organoclay hybrid was obtained as a white solid. The absence of peaks in the

XRD pattern indicated the exfoliation of the clay platelets in the PA1012 matrix.

The in-situ polymerization technique has also been applied for the preparation of

nanocomposites based on thermoplastic polymers other than polyamides, including

t
polymethyl methacrylate (PMMA) [98-99], polystyrene (PS) [100], polybenzoxale

ip
(PBO) [101], polyolefins (PP and PE) [102-105], and polyethylene terephthalate

cr
(PET) [106].

For example, in a study discussing the synthesis of PET nanocomposites using in

us
situ polymerization, the organo-modified montmorillonite is reported to react with

an
PET comonomers (ethylene glycol and terephthalic acid derivatives) to form an

intercalated nanocomposite [106].


M
However, according to Lee et al. [107] endeavors to prepare PET nanocomposites

using direct condensation polymerization of diol and diacid result in formation of


d

oligomers with significantly low molecular weight, due to ineffective control of


te

stoichiometry; and thus, a large increase in intragallery distance is hard to obtain. On

the other hand, attempts to prepare PET nanocomposites through melt intercalation
p

have resulted in limited intercalation of guest molecules, presumably due to the high
ce

viscosity of PET polymer. Therefore, the authors proposed and successfully applied

the ring-opening polymerization of ethylene terephthalate cyclic oligomers (ETCs)


Ac

with organically modified MMT, as an alternative approach to the preparation of

PET-based nanocomposites. Because of low molecular weight and cyclic molecular

architecture, cyclic oligomers of PET have much lower solution and melt viscosities

than the corresponding polymer; so it may be expected that, when clay intercalation is

intended by mixing with cyclic oligomer instead of linear polymer, easier diffusion

and a higher degree of intercalation or exfoliation would be obtained. In addition,

34
Page 34 of 302
problems such as precise control of stoichiometry and high vacuum requirements,

strictly required in preparation of PET by conventional condensation polymerization

of difunctional monomers, can be effectively avoided through ring opening reaction

of cyclic oligomers (Fig. 17).

t
HDPE nanocomposites have been synthesized by the so-called polymerization-

ip
filling technique (PFT), which involves anchoring in a first step a Ziegler-Natta type

cr
or any other coordination catalyst onto a filler surface and then in situ polymerizing

ethylene directly from the surface-treated fillers. In order to apply this technique to

us
nanofillers, layered silicates in aqueous colloidal suspension were made less

an
hydrophilic through elimination of water by freeze drying. The fluffy materials

obtained could then be nicely dispersed in non-polar solvents such as heptane or


M
toluene. The clay dispersion was then surface-treated with MAO, and after solvent

removal by evaporation, a high temperature treatment at 150 °C was applied to


d

modify the layered silicate. After removal of unreacted MAO, the silicate layers were
te

contacted with a metallocene pre-catalyst, i.e. (tert-butylamido)dimethyl (tetramethyl-

Z5-cyclopentadienyl) silane titanium dimethyl, for 1 h to form the active catalyst


p

species. The polymerization was carried out by adding ethylene to the medium. It has
ce

to be pointed out that no ion exchange reaction was required. Rather, this strategy

relied on the immobilization of the active species through electrostatic interactions


Ac

with surface anchored MAO. Some typical in situ intercalation experiments are

gathered in Table 4. When ethylene polymerization was carried out in the absence of

molecular hydrogen (thus without any transfer agent), layered silicate/UHMWPE was

produced, which is an extremely viscous material, very difficult to melt (Table 4,

entries 1 and 2). Addition of hydrogen to the polymerization medium allows

molecular weight to be reduced with substantial improvement of melt processability

35
Page 35 of 302
(Table 4, entry 3). Examination of the TEM pictures reveals partial exfoliation, even

at high filler content (ca. 30 wt % montmorillonite) [108-110].

Similarly, Bergman et al. [103] used a palladium-based complex and synthetic

fluorohectorite as the polymerization catalyst and inorganic component respectively

t
for the synthesis of polyethylene nanocomposites. They first intercalated palladium

ip
catalyst into the galleries of modified fluorohectorite (C14N-2) and exposed the dry

cr
powder to ethylene gas. Over a 2 h period, they observed monomer consumption and

a dramatic increase in the size of the silicate-catalyst composite, while after 12 h a

us
large mass of colorless, rubbery polymer formed. Analysis of the toluene-extracted

an
polyethylene by GPC revealed a high molecular weight. Moreover, the complete

absence of diffraction peaks in the XRD patterns strongly suggests the formation of
M
an exfoliated nanocomposite.

Jin et al. [105] also applied the in situ exfoliation method during ethylene
d

polymerization by fixing a Ti-based Ziegler-Natta catalyst at the inner surface of


te

MMT. They used organic salts with hydroxyl groups for the modification of MMT

(MMT-OH) since hydroxyl groups in intercalation agents offer facile reactive sites for
p

anchoring catalysts between silicate layers. Figure 18 represents the TiCl4 fixation
ce

mechanism between silicate layers of MMT-OH. The polymerization of ethylene was

conducted by injecting ethylene into the catalyst slurry (30-50°C, 4 bars). After
Ac

predetermined reaction times, polymerization was quenched with a dilute HCl

solution in methanol. The polymer was precipitated in methanol, separated by

filtration and dried in vacuum. The powder was then used as the master batch and

mixed with HDPE by melt extrusion. WAXD analysis again revealed exfoliation of

MMT layers in both the powdery reaction product and the HDPE matrix.

36
Page 36 of 302
In a study by Shin et al. [111] montmorillonite was intercalated with

triisobutylaluminum and ω-undecyleny-alcohol. The intercalation process allows the

transition metal catalyst and the activator methylaluminoxane to enter the clay

galleries and polymerize ethylene, while formation of polyethylene inside the

t
galleries leads to the exfoliation of layered silicates. The authors noted that two types

ip
of polymerization are possible: (a) the homopolymerization of ethylene, and (b) the

cr
copolymerization of ethylene and the vinyl ends of alcohol modifier connected to the

surface during intercalation, which produces polyethylene chains, chemically

us
connected to the silicate surface.

an
Heinemann et al. [112] carried out ethylene polymerizations in the presence of

layered silicates, such as organo-modified bentonite and unmodified hectorite and


M
fluoromica, using 1-octene as comonomer. Different catalysts were used, affording

HDPE and ethylene-octene copolymers (zirconium-based catalyst) and branched


d

polyethylene (nickel and palladium-based catalysts). It appeared that the modified


te

bentonites had a dramatic negative effect on the polymerization activity of the

zirconium-based catalyst due to its high sensitivity towards any kind of polar
p

functionality, while Ni-and Pd-based catalysts were much less affected by the nature
ce

of the clay. However, nanocomposites formed only in the case of organo-modified

clays, while in situ polymerization with unmodified silicates gave microcomposite


Ac

structures.

PP/clay nanocomposites have also been prepared by in situ intercalative

polymerization. Tudor et al. [104] treated a synthetic hectorite with

methylaluminoxane (MAO) to remove the acidic protons and to prepare the interlayer

spacing to receive the transition metal catalyst. Details of the preparation route are

shown in Fig. 19. Using a synthetic fluorinated mica-type layered silicate deprived of

37
Page 37 of 302
protons in the galleries, the catalyst was intercalated directly within the silicate layers,

without the need of MAO treatment. Sun and Garces [102] also reported the

preparation of PPCN by in situ polymerization with metallocene/clay catalysts.

Following the in situ intercalative polymerization technique, Do and Cho [113]

t
compared the ability of several tetraalkylammonium cations incorporated into Na+-

ip
MMT to promote the intercalation of PS through the free radical polymerization of

cr
styrene, initiated by AIBN at 50 °C. Three tetraalkylammonium cations were tested,

all based on the following formula: (CH3)2N.(hydrogenated tallow alkyl)R, where

us
hydrogenated tallow alkyl corresponds to a mixture of mainly octadecyl chains

an
together with small amounts of lower linear homologs and R may be either another

hydrogenated tallow alkyl (Ta), 2-ethyl hexyl (Eh) or benzyl (Bz) group. The so-
M
modified o-MMTs were coded as Ta-MMT, Eh-MMT and Bz-MMT, respectively.

Layer spacings obtained for the three MMTs and corresponding composites are
d

presented in Table 5 and reveal that the best intercalation occurs for Bz-MMT. This is
te

probably due to a better affinity between styrene and benzyl groups spread all along

the layered MMT surfaces. It is worth mentioning that even though the interlayer
p

spacing for Ta-MMT does not change much, the authors assumed that intercalation
ce

occurs in this case also; however this is not evident from the d-spacings because such

hydrogenated tallow alkyl chains should be long enough (mainly C18 chains) to easily
Ac

accommodate the PS. Even though this technique allows for extensive intercalation of

PS chains through the choice of an appropriate alkylammonium cation, neither

exfoliation nor control over the molecular weight of PS produced has been observed.

Akelah and Moet [100] modified MMT with vinylbenzyltrimethyl ammonium

cation, and they dispersed and swelled the modified clay in various solvent and co-

solvent mixtures, such as acetonitrile or acetonitrile/toluene under N2 . Styrene

38
Page 38 of 302
polymerizations were carried out in presence of AIBN at 80 °C for 5 h. In this way,

intercalated PS/MMT nanocomposites were produced, with the extent of intercalation

depending upon the nature of the solvent used. However, although PS is well

intercalated, one drawback of this procedure is that the macromolecule produced is

t
not a pure PS, but rather a copolymer between styrene and surfactant.

ip
A similar approach has been applied by Weimer et al. [114], who modified Na+-

cr
MMT by anchoring an ammonium cation bearing a nitroxide moiety known for its

ability to mediate the free radical polymerization of styrene in bulk. The absence of

us
WAXD peaks together with TEM observations of silicate layers randomly dispersed

an
within the PS matrix attest to the complete exfoliation of the layered silicate.

In another study, Yei et al. [115] prepared PS/clay nanocomposites through


M
emulsion polymerization, by suspending the clay in styrene monomer. The clay was

treated with either cetyl-pyridinium chloride (CPC) or the inclusion complex of CPC
d

in α-cyclodextrin (CPC/α-CD). This was the first study reporting cyclodextrins (CDs)
te

as surfactants for intercalation in clay. CDs comprise a series of α-1,4-linked cyclic

oligosaccharides with shapes resembling hollow truncated cones. The cavities of CDs
p

are hydrophobic; thus CDs have the ability to include hydrophobic molecules within
ce

their cavities. For the preparation of an inclusion complex, CPC was mixed with a

saturated solution of α-CD in water and and the complex was obtained as a white
Ac

crystalline precipitate. The modified clays were also obtained as white precipitate

after adding a solution of the surfactant (CPC or CPC/α-CD inclusion complex) in a

Na+-MMT suspension. Emulsion polymerization was performed as follows: A

surfactant solution (CPC or CPC/α-CD inclusion complex) was added to an aqueous

suspension of clay. After stirring for 4 h, KOH and SDS were added to the solution

and the temperature was raised to 50 °C. Styrene monomer and K2S2O8 were added

39
Page 39 of 302
slowly and polymerization was performed at 50 °C for 8 h. After cooling, 2,5 %

aqueous aluminum sulfate was added to the polymerized emulsion, followed by dilute

HCl with stirring. Finally, acetone was added to break the emulsion. XRD patterns

indicate that the CPC surfactant is intercalated successfully into the galleries of the

t
clay, increasing their spacing from 1.43 to 2.27 nm. Furthermore, the d-spacing

ip
caused by the inclusion complex is 5.12 nm, i.e. substantially higher than that caused

cr
by CPC alone. This was explained by the authors considering that the linear, aliphatic

chain within the CPC/α-CD cannot bend within the galleries of the clay. On the other

us
hand, no peak was detected in the XRD patterns for polymer/clay nanocomposites

an
prepared from the CPC and CPC/α-CD treated clays which implies that they all

possess exfoliated structures.


M
In another study, exfoliated PLA/clay nanocomposites were prepared by in situ

coordination-insertion polymerization method [116]. The authors used two different


d

kinds of OMLS (C30B and C25A) for the preparation of nanocomposites. In a typical
te

synthetic procedure, the clay was thoroughly dried and placed in the polymerization

vial. L,L-lactide solution in dried THF was then transferred to the vial under nitrogen
p

and the solvent was eliminated under reduced pressure. Polymerizations were
ce

conducted in bulk at 120 °C for 48 h, after 1 h of clay swelling in the monomer melt.

When C30B was used, the polymerization was co-initiated AlEt3, while Sn(Oct)2 was
Ac

used to catalyze the polymerization of L,L-lactide in the presence of C25A. The clay

C30B led to fully exfoliated structure, whereas C25A based nanocomposites exhibited

an intercalated morphology.

Messersmith and Giannelis [117] reported the first preparation of PCL

nanocomposites, in 1993, by in situ intercalative polymerization, using Cr3+

exchanged fluorohectorite (FH). In a typical synthesis, a mixture of Cr3+ FH and CL

40
Page 40 of 302
was heated at 100 °C for 48 h. Upon cooling to room temperature the reaction mixture

solidified. Intercalation of the CL monomer was revealed by XRD, which showed an

increase in the silicate d-spacing from 1.28 to 1.46 nm. The d001 spacing observed

prior to polymerization was found to be consistent with the orientation of the CL ring

t
perpendicular to the silicate layers. XRD analysis of the nanocomposites after

ip
polymerization indicates a reduction in the silicate d-spacing from 1.46 to 1.37 nm, as

cr
presented in Fig. 20. The decrease in the d-spacing is consistent with the dimensional

change accompanying polymerization of CL monomer.

us
The same researchers also modified a Na+-MMT by the protonated aminolauric

an
acid and dispersed this modified clay in liquid ε-caprolactone before polymerizing it

at high temperature. The PCL nanocomposites were prepared by mixing up to 30 wt


M
% of the modified clay with ε-caprolactone at room temperature for two h, followed

by the ring opening polymerization at 170 °C for 48 h. Interestingly enough, XRD


d

patterns of the modified clay after contact with ε-caprolactone at room temperature do
te

not show any significant increase in the layer spacing (13.6 Å). The authors assumed

that the monomer intercalates in the gaps between the aminolauric acid chains so that
p

no gallery expansion could be seen, in contrast to what is usually observed in in situ


ce

intercalative polymerization, where the monomer insertion within the silicate gallery

induces an increase in the interlayer spacing. Another possibility may be that


Ac

intercalation of the monomer occurs only during the heating step of the solution. After

polymerization, XRD patterns of the obtained composites did not show a diffraction

peak, indicating that exfoliation occurred [118].

Pantoustier et al. [119-120] compared PCL nanocomposites prepared by

intercalative polymerization using both pristine MMT and ω-aminododecanoic acid

modified MMT. The polymerization of CL with pristine and modified MMT gives

41
Page 41 of 302
PCL with a molar mass of 4800 g/mol and 7800 g/mol, respectively, and a narrow

distribution. For comparison, the authors also conducted the same experiment without

MMT, but found no CL polymerization. These results demonstrate the ability of

MMT to catalyze and control CL polymerization, at least in terms of a molecular

t
weight distribution that remains remarkably narrow. Even though the polymerization

ip
mechanism of ε-caprolactone in the presence of clay generally remains unclear and

cr
different assumptions have been made [121-123], the authors assumed that CL is

activated through interaction with the acidic site on the clay surface.

us
On the other hand, Gorrasi et al. [6] suggested that the low molecular weight of

an
PCL synthesized by in situ polymerization in the presence of layered silicate may be

responsible for the high brittleness of the resulting nanocomposites. Therefore, they
M
proposed the preparation of blends of high molecular weight PCL with different

amounts of o-MMT/PCL nanocomposite prepared via in-situ polymerization and


d

containing a high amount (30 wt %) of clay. More specifically, they added a


te

suspension of an intercalated PCL nanocomposite (prepared by intercalative

polymerization of ε-CL and having a basal spacing of 24 Å) in chloroform to a


p

chloroform solution of PCL and after stiirng for 15 h, they obtained the final
ce

composite by precipitation into hexane. The XRD patterns of this product showed a

peak corresponding to a basal spacing of 18 Å, suggesting that the PCL is indeed


Ac

intercalated within the silicate layers of the clay in the blend that, however, underwent

partial restructuring during the blend formation.

At this point, it is worth mentioning that, even though in situ intercalative

polymerization has proved successful in the preparation of various polymer-layered

silicate nanocomposites, important drawbacks of this technique have also been

pointed out: (1) it is a time-consuming preparation route (the polymerization reaction

42
Page 42 of 302
may take more than 24 h); (2) exfoliation is not always thermodynamically stable; and

the platelets may re-aggregate during subsequent processing steps; and (3) the process

is available only to the resin manufacturer who is able to dedicate a production line

for this purpose [124].

t
ip
5.3.2. In-situ intercalative polymerization of thermosetting polymers

cr
Despite the aforementioned disadvantages of in situ intercalative polymerization,

us
this is the only viable technique for the preparation of thermoset-based

nanocomposites, since such nanocomposites obviously cannot be synthesized by melt

an
intercalation, which is the other commercially important preparation method [42, 125-

127].
M
In this case, the exfoliation ability of the organoclays is determined by their

nature, including the catalytic effect on the curing reaction, the miscibility with the
d

curing agent etc. Since there is a curing competition between intragallery and
te

extragallery resin, as long as the intragallery polymerization occurs at a rate


p

comparable to the extragallery polymerization, the curing heat produced is enough to


ce

overcome the attractive forces between the silicate layers and an exfoliated

nanocomposite structure can be formed. In contrast, if the extragallery polymerization


Ac

is more rapid than the intragallery diffusion and polymerization or if intragallery

polymerization is retarded, the extragallery resin will gel before the intragallery resin

produces enough curing heat to drive the clay to exfoliate; consequently, exfoliation

will not be reached. It can be inferred, therefore, that factors promoting the curing

reaction of intragallery resin will facilitate the exfoliation of the clay. Such factors

include the catalytic effect of organoclay on the curing reaction, the good penetrating

ability of curing agent to clay, the long alkyl-chain of the organo-cation, meaning a

43
Page 43 of 302
greater amount of intragallery resin preload and a completed organization of the clay,

and meaning weaker attractive forces between the silicate layers [33, 128].

In fact, a number of research groups have studied the effect of various parameters

on the exfoliation of clays in epoxy resins. Pioneering studies by Pinnavaia [129] on

t
MMT/epoxy systems established the initial conceptual methodology. Interfacial

ip
modifiers, such as primary ammonium alkyls are intercalated between the MMT

cr
layers, not only to compatibilize the inorganic aluminosilicate and organic resin, but

also to accelerate the crosslinking reaction between the layers through acid catalysis.

us
That is, as the curing agent is mixed into the clay/epoxy mixture, it is thought that the

an
modifiers introduced into the galleries of the clay sheets would promote the reaction

between the epoxy in the gallery with the curing agent. This would make the
M
intragallery curing reaction faster than the extragallery reaction, thus facilitating the

expansion of the clay sheets and helping to achieve exfoliation [130].


d

For the diglycidyl ether of bisphenol A (DGEBA), crosslinked with m-phenylene


te

diamine (MPDA) and containing 5 % modified MMT, Lan et al. [131] found that the

clays with primary and secondary onium ions formed exfoliated nanocomposites,
p

whereas those with tertiary and quaternary onium ions retained an intercalated
ce

structure. It was argued that acidic alkyl ammonium ions tend to favor exfoliation by

catalyzing homopolymerization of DGEBA molecules inside the clay galleries. It was


Ac

also found that the length of alkyl chains of modified MMT may determine whether

an intercalated and partially exfoliated or a totally exfoliated nanocomposite will be

obtained. The same kind of study was also conducted by Zilg et al. [132] who cured

DGEBA with hexahydrophthalic acid anhydride in the presence of different types of

clays, again modified with a wide variety of surfactants.

44
Page 44 of 302
Another parameter of the clay that greatly affects the outcome of nanocomposite

processing is the cation exchange capacity (CEC), which determines the amount of

surfactant ions present between the clay layers and therefore controls the space

available for diffusion of epoxy molecules during mixing with the organoclay. It has

t
been established that the highest CEC provides the minimum space. In this context,

ip
the swelling phase is of critical importance to the final nanocomposite structure. An

cr
MMT with a low CEC is exfoliated already during swelling in the epoxy resin prior to

curing. A possible mechanism explaining this phenomenon is homopolymerization of

us
the epoxy resin during the swelling phase, causing diffusion of new epoxy molecules

an
into the clay galleries. The large amount of space available between the layers favors

the diffusion. On the other hand, the duration of swelling of the clay with high CEC is
M
shown to be critical for the synthesis of an exfoliated nanocomposite [42].

Other researchers investigated the effect of the polymer resin. For example,
d

Becker et al. [133] prepared nanocomposites of three different epoxy resins:


te

diglycidyl ether of bisphenol A (DGEBA), triglycidyl p-amino phenol (TGAP) and

tetrafunctional tetraglycidyldiamino diphenylmethane (TGDDM), using a mixture of


p

two diethyltoluene diamine (DETDA) isomers as the hardener and a commercially


ce

available octadecyl ammonium ion modified MMT as the clay. All epoxy resin

systems intercalated the organically modified layered silicate and increased the d-
Ac

spacing from 23 Å up to 80 Å. Similarly, Hackman and Hollaway [134] noted that the

epoxy resin component of the nanocomposite has little effect on the exfoliation of the

clay layers; although it is the basic unit, the curing agent controls the rate of cure.

Lower viscosity resins lead to faster pre-intercalation, but they do not seem to offer

any significant long-term advantage.

45
Page 45 of 302
Interestingly, much research has focused on the effect of curing agent on the

intercalation/exfoliation of clays in epoxy resins. Messersmith and Giannelis [135]

analyzed the effect of different curing agents on the formation of nanocomposites

based on DGEBA and a montmorillonite modified by bis(2-hydroxyethyl)methyl

t
hydrogenated tallow alkyl ammonium cation. They found that when primary and

ip
secondary amines, such as methylene dianiline, were used, only intercalated epoxy-

cr
clay structures could be obtained. This was attributed to either the bridging of the

silicate layers by the bifunctional amine molecules, which prevents further expansion

us
of the layers from taking place, or to the strong polarity of the N-H groups in the

an
primary and secondary amines that causes a re-aggregation of dispersed silicate

layers. When other curing agents, such as nadic methyl anhydride (NMA), boron
M
trifluoride monomethylamine (BTFA) or benzyldimethylamine (BDMA) were added,

delamination during heating of the reaction mixture occurred. Addition of the curing
d

agent induced first an increase of the interlayer spacing from 36 to 39 Å, indicating


te

some partial intercalation. With further heating, disappearance of the interlayer

spacing reflection indicated that delamination had occurred. Study of the curing
p

reactions tended to prove that the particular alkylammonium used (that bears two
ce

hydroxyl functions) could play an active role, especially when BDMA or NMA were

added as the curing agents. For example, BDMA can catalyze the reaction between
Ac

the hydroxyl groups of the alkylammonium and the oxirane of the monomer,

producing a new hydroxyl that subsequently reacts with free DGEBA via a similar

base-catalyzed oxirane ring opening to build up the epoxy network.

Recently, Le Pluart et al. [136] investigated the influence of curing agent and clay

organophilic treatment on the reactivity and cure behavior of epoxy networks and on

the morphology of the final composites. They used two different curing agents: an

46
Page 46 of 302
aliphatic diamine with a polyoxypropylene backbone (D2000) and 4,4′-

methylenebis[3-chloro-2,6-diethylaniline] (MCDEA) as well as two montmorillonites

modified with different alkyl ammonium ions having the same chain length. The first

MMT is Tixogel MP250, a benzyl dimethyl tallow alkyl ammonium MMT, while the

t
second is OPTC18, an octadecylammonium ion modified MMT. The authors noticed

ip
that gel time was decreased in the presence of Tixogel for both the DGEBA/D2000

cr
and the DGEBA/MCDEA systems. However, for unclear reasons the authors could

us
not determine a gel time in the presence of OPTC18. Concerning the influence of

network formation on organoclay dispersion, the d-spacings of the clays at the

an
beginning and the end of polymerization are presented in Table 6. In the case of

Tixogen, both reactive mixtures have swelled the organoclay already at the beginning
M
of the reaction, since the d-spacing increased from 20.2 up to 34 Å. During the

polymerizations the d-spacing remains the same, whatever the curing agent used,
d

demonstrating that the polymerization of the epoxy/amine systems does not modify
te

the dispersion of Tixogen. The observations are very different in the case of OPTC18.

Again the initial state of dispersion is good, but the d-spacing further increases during
p

isothermal reaction of both reactive mixtures. This demonstrates the possibility of


ce

improving the quality of dispersion at the nanometer scale during the polymerization

of the network. The authors noted that such improvement is not only due to the kind
Ac

of organoclay used nor to the epoxy system chosen, but rather is linked with

interactions and chemical affinities between the organoclay and the network

precursors, as well as with reaction and diffusion kinetics of the reactive systems.

In another study, Chin et al. [33] confirmed the formation of exfoliated

nanocomposites when DGEBA was auto-polymerized with MMTs. Exfoliated

nanocomposites were also observed with DGEBA cured with MPDA of less than

47
Page 47 of 302
equimolar concentration. However, as the curing agent concentration increased, the

extragallery crosslinking dominated, resulting in intercalated nanocomposites.

Among the various parameters, the effect of processing has also been investigated.

Lu et al. [128] intercalated an organically modified MMT by epoxy resin by both

t
direct or solution mixing. XRD patterns indicated that intercalation was realized

ip
irrespective of the mixing method applied. It was also concluded that prolonging the

cr
stirring time above a certain level or using solution intercalation would not further

improve the intercalation. An important finding was that the intercalated hybrids were

us
quite stable for storage. In order to prepare nanocomposites, a curing agent was added

an
into the hybrid and mixed thoroughly. The results indicated that if the organoclays can

be exfoliated at all, the exfoliation will be finished before the gel point of the epoxy.
M
Koerner et al. [137] studied the impact of shear during epoxy nanocomposite

processing and demonstrated that with proper mechanical processing conditions,


d

uniform dispersion and a high degree of exfoliation is possible in systems that


te

typically only show intercalated morphologies after traditional cure cycles.

Conceptually, this is achieved by maximizing epoxy viscosity by halting cure before


p

gelation and by compounding at sub-ambient temperatures near the resin’s glass


ce

transition. High shear forces, due to the high viscosity of the system, facilitate

homogenization of the layered silicate nanocomposite.


Ac

The so-called “high pressure mixing” (HPM) method, involving clay processing

in solvent (acetone) under high pressure and also against solid obstacles, has been

applied by Liu et al. [130] for the preparation of epoxy nanocomposites. Using this

method, nanoclays can be dispersed in acetone and an epoxy solution to form a stable

suspension, in which the basal spacing of the nanoclay is increased. Using TGDDM

as the matrix, 4,4′-diaminodiphenyl sulphone (DDS) as the hardener and an octadecyl

48
Page 48 of 302
amine modified MMT, nanocomposites of up to 7.5 wt % clay loading were

successfully synthesized with the HPM method. TEM images show that the

agglomerates of nanoclays were broken down to form small particles consisting of

several clay platelets.

t
Similarly, the “slurry compounding” approach has been developed for epoxy/clay

ip
nanocomposite preparation using sodium MMT [138]. The most significant feature of

cr
this technique is that very little (< 5 wt %) organic modifier is required to facilitate

the exfoliation and dispersion of the clay, reducing the cost of the nanocomposites. To

us
further reduce the cost of polymer/clay nanocomposites Wang et al. prepared

an
epoxy/crude clay nanocomposites using the “slurry-compounding” technique. A clay-

acetone slurry was mixed with required quantities of epoxy resin at 50 °C. Acetone
M
was then evaporated and a stoichiometric quantity of the curing agent was added. The

mixture was degassed and dried, leading to the formation of ordered exfoliated
d

nanocomposites.
te

As for epoxy-based nanocomposites, the synthesis of unsaturated polyester

(UP)/layered silicate nanocomposites involves two steps: first the mixing process,
p

wherein the UP linear chains are mixed with the curing agent and layered silicate and
ce

second the curing process, during which the crosslinking reaction takes place by

decomposing the initiators [139].


Ac

The synthesis of nanocomposites based on montmorillonite and UP has been

reported by different research groups. In a typical example, Bharadwaj et al. [140]

described the preparation of crosslinked polyester/clay nanocomposites by dispersing

organically modified MMT in pre-promoted polyester resin and subsequently

crosslinking the system using methyl ethyl ketone peroxide (MEKP) catalyst at

49
Page 49 of 302
several clay concentrations. The formation of exfoliated nanocomposites was

confirmed by XRD and TEM.

In another study, montmorillonite was treated with methacrylate-silane coupling

agent in order to rend the filler hydrophilic and reactive. Then, UP was polymerized

t
by the free radical polymerization with the modified montmorillonite dispersed in it.

ip
The authors claimed the formation of exfoliated structure, based on XRD and TEM

cr
findings [127].

Furthermore, Suh et al. [139] reported on the formation mechanism of UP

us
nanocomposites. They used two different kinds of MMT: a dodecyl ammonium

an
bromide MMT and Cloisite 20A, containing dimethyl dehydrogenated tallow

ammonium as an organic modifier, and used two different ways of mixing in order to
M
prepare UP-based nanocomposites, i.e. simultaneous mixing or sequential mixing,

where in the first step, pre-intercalates of the UP and MMT nanocomposites (i.e. the
d

mixtures of UP and modified MMT) are prepared; and then the styrene monomer,
te

acting as the curing agent, is added with varying mixing times from 15 to 180 min.

Finally, all mixtures were cured at 80 ° C for 3 h and post-cured at 120 °C for 4 h.
p

The structures of UP/MMT nanocomposites were investigated by XRD and TEM,


ce

whereas in order to investigate the formation mechanism of UP/MMT

nanocomposites, the authors used DMTA, solution rheometry and melt rheometry.
Ac

The results led the authors to suggest the following mechanism of UP/silicate

nanocomposite formation: The styrene monomer moves more easily than uncured UP

chains. This may generate higher styrene monomer concentration in the MMT gallery

than in any other part in a simultaneous mixing system. If polymerization occurs in

these conditions, the total crosslinking density of the sample decreases, because of the

low concentration of styrene in uncured UP linear chains. In the sequential mixing

50
Page 50 of 302
method, the styrene monomer diffuses into the gallery of MMT intercalated with UP

as time goes on. Therefore, it is thought that crosslinking density and Tg of UP/silicate

nanocomposite increase to some extent. Hence, the styrene monomers are more easily

dispersed inside and outside the silicate layers as mixing time increases. Therefore,

t
the crosslinking reaction takes place homogeneously inside and outside the silicate

ip
layers, and crosslinking density reaches the degree of crosslinking density of the

cr
cured pure UP.

A number of researchers have synthesized polyurethane (PU)-based

us
nanocomposites by the in situ polymerization method. The first examples of

an
elastomeric PU/clay nanocomposites with improved properties compared to the

pristine polymer were reported by Wang and Pinnavaia [141]. Conventional PU


M
microcomposites are usually formulated by premixing the inorganic component with

the polyol and then curing the mixture with the diisocyanate. The approach of Wang
d

and Pinnavaia to form polyurethane nanocomposites, therefore, focused on the


te

solvation of the organoclay by polyols. Interestingly, they found that montmorillonites

exchanged with long chain onium ions (carbon number>12) are easily solvated by
p

several polyols that are commonly used in polyurethane chemistry.


ce

More recently, Yao et al. [142] reported the preparation of a novel kind of

PU/MMT nanocomposite using a mixture of modified 4,4′-diphenyl methylate


Ac

diisocyanate (MMDI), modified polyether polyol (MPP) and Na+-MMT. In a typical

synthetic route, a known amount of Na+-MMT was blended with a known amount of

MMDI and cured at 78 °C for 168 h. As measured through XRD, the gallery spacing

of the layered clay is 1.1 nm and increases to 1.6 nm for the PU/clay (21.5 %)

nanocomposites, indicating that the PU chains were intercalated between the layers of

the clay.

51
Page 51 of 302
On the other hand, Mulhaupt et al. [143] prepared PU nanocomposites from

modified reactive fluoromica clay. The dried organophilic mica was dispersed by

means of a high shear mixer in trihydroxy-terminated oligo-propylene-oxide. Stable

and transparent polyol dispersions were obtained and then cured with

t
diisocyanatophenylmethane and accelerated with 0.6 wt % N,N-dimethyl-benzyl-

ip
amine.

cr
Berta et al. [144] synthesized elastomer polyurethane nanocomposites showing

several degrees of dispersion, using polyols with different molecular weight and

us
functionality (see Table 7) as well as methylenediphenylene diisocyanate (MDI). For

an
the preparation of nanocomposites, the desired weights of polyol and organoclay were

mixed and then butanediol (chain extender), catalyst and MDI were added in
M
appropriate amounts. The system was cured at 120 °C for 3 h and 80 °C for 24 h. The

synthesis of the PU/clay nanocomposites is schematically illustrated in Fig. 21. Table


d

8 summarizes the d-spacings determined by SAXS for the PU/clay nanocomposites. It


te

is clear that all the PU/clay materials are nanocomposites, as the 18 Å d-spacing,

associated with the base organoclay has disappeared. No peaks associated with clay
p

gallery spacings were observed for NC-II and NC-IV, indicating that in these
ce

nanocomposites there is no longer sufficient ordering of the clay platelets to produce a

scattering peak. The authors postulated that this is due to the higher equivalent Mw
Ac

(Mn/functionality) of the polyols in these nanocomposites. According to the authors,

this is of particular importance, since it means that the dimensions of the polyol

molecule control the gallery spacing in the initial step of dispersing the clay in polyol.

Finally, Gao et al. [145] prepared PU nanocomposites, as well as their foams by in

situ polymerization and batch foaming with different modified MMTs. MMT was

modified by dibutyldimethoxytin (DBDMT) and, thus, organophilic montmorillonite

52
Page 52 of 302
with a catalytic function, denoted as MMT-Tin, was obtained. For the synthesis of PU

nanocomposites and foams, the authors used a polymeric aromatic isocyanate based

on diphenylmethane 4,4′-diisocyanate (MDI) and two trifunctional polyester polyols.

For PU nanocomposites, clay was first mixed with one monomer and then the second

t
monomer was added. Catalyst was always added with polyol. After polymerization

ip
under ambient conditions, the hybrid was post-cured at 100 °C for 4 h. For reactive

cr
foaming of PU, a surfactant as well as pentane (blowing agent) were also used. The

us
mixture of all ingredients was mixed and foaming occurred in a closed plastic

container with fixed volume at ambient temperature, followed by curing at 100 °C. By

an
XRD analysis it was found that the basal spacing of the organoclay increased

compared to MMT (d001=1.16 nm), because the gallery of MMT was expanded by
M
molecular chains of the modifier. However, the d-spacing decreased from 1.77 to 1.43

nm when MMT-OH was further modified by DBDMT – an effect attributed to


d

conjugation of hydroxyl groups from different layers. The authors studied the effect
te

of the clay-monomer mixing sequence on clay dispersion and found that the two-step

process offered better clay dispersion than the one-step approach, wherein all
p

ingredients were mixed simultaneously. Especially, premixing the functional clays


ce

with isocyanate provides better clay dispersion, which was attributed to the reaction

between the isocyanate monomers and the hydroxyl groups on alkyl chains of MMT-
Ac

OH, causing an increase of gallery spacing of clay. In the XRD spectra no diffraction

peak was observed for PU nanocomposite with MMT-Tin. In fact, TEM analysis

revealed that nanocomposites containing both MMT-OH and MMT-Tin exhibited

good clay dispersion, however MMT-Tin showed better exfoliation and more uniform

dispersion, probably due to the intragallery catalysis of organotin. Turning to the

reactive foaming of PU nanocomposites, Fig. 22 shows SEM images of the freeze

53
Page 53 of 302
fractured surface of PU foams. It is clearly observed that the neat PU foam has fewer

cells and a larger cell size than PU nanocomposite foams with 5 wt % organoclay,

whereas there is little difference between MMT-Tin/PU and MMT-OH/PU

nanocomposite foams in terms of cell size and density. The appearance of a shoulder

t
at very low angle in the XRD patterns of nanocomposite foams implies that clay

ip
orientation and dispersion is somewhat affected by the foaming process. However, as

cr
the authors noted, the detailed mechanism of how nanoparticles influence cell

morphology needs further investigation.

us
A list of representative thermoset and thermoplastic – based nanocomposites

an
prepared through in situ polymerization is given in Table 9.
M
5.4. Polymer Melt Intercalation
d
te

5.4.1. Introduction and advantages of the technique

For most technologically important polymers, both in-situ polymerization and


p

intercalation from solution are limited because neither a suitable monomer nor a
ce

compatible polymer-silicate solvent system is always available. Moreover, they are


Ac

not always compatible with current polymer processing techniques. These

disadvantages drive the researchers to the direct melt intercalation method, which is

the most versatile and environmentally benign among all the methods of preparing

polymer-clay nanocomposites (PCNs) [4, 146].

As already mentioned, nanocomposite synthesis via polymer melt intercalation

involves annealing, usually under shear, of a mixture of polymer and layered silicate

above the softening point of the polymer. During annealing, polymer chains diffuse

54
Page 54 of 302
from the bulk polymer melt into the galleries between the silicate layers, as shown in

Fig. 23 [13, 15, 47].

The advantages of forming nanocomposites by melt processing are quite

appealing, rendering this technique a promising new approach that would greatly

t
expand the commercial opportunities for nanocomposites technology [14-15, 46-47].

ip
If technically possible, melt compounding would be significantly more economical

cr
and simpler than in situ polymerization. It minimizes capital costs because of its

compatibility with existing processes. That is, melt processing allows nanocomposites

us
to be formulated directly using ordinary compounding devices such as extruders or

an
mixers, without the necessary involvement of resin production. Therefore, it shifts

nanocomposite production downstream, giving end-use manufacturers many degrees


M
of freedom with regard to final product specifications (e.g. selection of polymer

grade, choice of organoclay, level of reinforcement, etc). At the same time, melt
d

processing is environmentally sound since no solvents are required [14-15]; and it


te

enhances the specificity for the intercalation of polymer, by eliminating the competing

host-solvent and polymer-solvent interactions [29].


p

Thus, the majority of thermoplastic polymers, including PA [14, 36, 51, 147], PET
ce

[148] (and recycled PET [149]), EVA [81, 150], thermoplastic polyurethane [88],

polyolefins [111, 151], PLA [152-154], PCL [155-156], etc, have been used to study
Ac

nanocomposite formation by melt intercalation.

Before discussing in detail the factors affecting clay delamination during melt

blending as well as the degradation issues involved in this technique, it is worth

describing a slightly modified approach aiming to facilitate exfoliation and consisting

of melt processing the polymer with a pre-intercalated clay or slurry.

55
Page 55 of 302
Following such an approach, Liu and Wu [146] prepared a co-intercalated clay, by

absorbing epoxide compound between the silicate layers. They expected strong

interaction between PA66 and this new kind of modified clay, since Ishida had

successfully prepared PA6, PA12 and other polymer nanocomposites using similarly

t
modified clay. The preparation of the new kind of co-intercalated organophilic clay

ip
used was as follows. Na+-MMT was dispersed in hot water using a homogenizer.

cr
Then, hexadecyltrimethylammonium bromide, dissolved in hot water, was poured into

the Na+-MMT-water solution with vigorous stirring for 30 min to yield a white

us
precipitate, which was collected and washed with hot water. After thorough drying in

an
a vacuum oven, the precipitate was ground into a product termed PrEMMT. PrEMMT

and epoxy resin GY 240 were mixed in a Haake Reocorder 40 mixer for 1 h; thus the
M
co-intercalated clay, termed EMMT was obtained. Subsequently, a twin-screw

extruder was used for the preparation of nanocomposites. In X-ray diffraction (XRD)
d

patterns, Na-MMT showed a characteristic diffraction peak corresponding to the


te

(001) plane at 1.24 nm. PrE-MMT showed a 1.96 nm basal spacing in the XRD

pattern, while the basal spacing of E-MMT was 3.77 nm. The obviously larger layer
p

distance of E-MMT demonstrates the advantage of co-intercalated organophilic clay.


ce

As the authors pointed out, the alkylammonium ion exchange enables conversion of

the hydrophilic interior clay surface to hydrophobic and increases the layer distance as
Ac

well. This is the condition of PrEMMT. In this organophilic environment, epoxy resin

then diffuses into the clay galleries to further increase the layer distance. In addition,

the co-intercalated clay also brings the active functional group into the PA66 system.

Therefore, a better dispersion effect can be expected for EMMT. In fact, using the

aforementioned procedure, Liu and Wu obtained well exfoliated nanocomposites

when clay loading was less than 7 wt %.

56
Page 56 of 302
Also in the case of polyolefins, the use of a swelling agent (a monomer or polymer

known to intercalate/exfoliate smectite clay) next to the surfactant placed at the clay

surfaces after an ion-exchange reaction has allowed the preparation of

nanocomposites. Present in small amounts, the swelling agent serves to swell the clay

t
layers, allowing the organic matrix to be virtually any polymer [2]. As an example,

ip
Wolf et al. [157] modified a commercially available organo-ammonium-exchanged

cr
montmorillonite, using an organic swelling agent (with boiling point between 100 °C

and 200 °C, such as ethylene glycol, naphtha or heptane) in order to increase the

us
interlayer spacing. The swollen organo-modified clay was then compounded with PP

an
in a twin-screw extruder at 250 °C. The swelling agent was volatilized during

extrusion, leading to the formation of nanocomposites. Similarly, Liu and Wu [158]


M
reported the preparation of PP nanocomposites via melt-compounding, using a type of

co-intercalated organophilic clay. One of the co-intercalation monomers is


d

unsaturated, so it could tether on the PP backbone through a grafting reaction. The co-
te

intercalated organophilic clay (EM-MMT) was prepared as follows.

Hexadecylammonium modified MMT (C16-MMT) was mixed with epoxypropyl


p

methacrylate in a Haake mixer for 1 h. Before mixing with clay, the initiator for the
ce

grafting reaction, dibenzoyl peroxide (BPO) and a donor agent were dissolved in

epoxypropyl methacrylate. The nanocomposites were prepared using a twin-screw


Ac

extruder with a screw speed of 180 rpm operating at 200 °C. WAXD patterns and

TEM observations established that the larger interlayer spacing and the strong

interaction caused by grafting can improve the dispersion effect of silicate layers in

the PP matrix.

In another study, Zheng et al. [159] used an oligomerically-modified clay,

prepared by ion-exchange with the oligomer prepared from maleic anhydride (MA),

57
Page 57 of 302
styrene (ST) and vinylbenzyltrimethylammonium chloride (VBTACl) terpolymer,

herein called MAST, to prepare PS/clay nanocomposites by melt blending. The

synthetic route for the formation of the terpolymer MAST is depicted in Fig. 24.

Thereafter, a portion of MAST oligomer, dissolved in acetone was added drop-wise to

t
a dispersion of clay in distilled water and acetone. A precipitate (MAST hectorite

ip
clay) formed immediately. Nanocomposites were subsequently prepared by melt

cr
blending in a Brabender Plasticorder at 60 rpm and 190 °C for 15 min. XRD

measurements indicated a mixed intercalated/delaminated structure for the MAST

us
modified clay, whereas no peaks were observed for the PS/MAST. By combining

an
XRD and TEM analyses the authors concluded that the hybrids formed were

characterized by a mixed immiscible/intercalated/delaminated structure.


M
Recently, Hasegawa et al. [160] reported a novel compounding process for the

preparation of PA6/MMT nanocomposites, using a Na+-MMT water slurry as an


d

alternative for organically modified MMT. In this process, the Na+-MMT slurry was
te

blended with PA6 using an extruder, followed by removal of the water. WAXD

patterns and TEM observations clearly indicate the exfoliation of MMT layers in the
p

PA6 matrix but the final properties of PA6/Na+- MMT nanocomposites were nearly
ce

equal to those of conventional PA6/MMT nanocomposites prepared by dry

compounding with MMT. Figure 25 shows schematically dispersion of the Na+-MMT


Ac

silicate layers of the clay slurry into the PA6 matrix during compounding by an

extruder. According to this study, the exfoliation of silicate layers into the PA6 matrix

occurs as follows: (a) the clay slurry is first pumped into the melting matrix under

vigorous shear; (b and c) the clay slurry reduces to finer drops during blending and, at

the same time, the water of the slurry drops begins to evaporate in contact with the

PA6 melt; (d) the evaporated water is removed under vacuum, and silicate layers are

58
Page 58 of 302
dispersed into the PA6 melt as monolayer or as a few layers. The dispersion of silicate

layers in this process is quite different from that of conventional compounding

process using organophilic clay, where polymer chains first intercalate into the

stacked silicate galleries and then exfoliate into the matrix. In this process, the

t
exfoliated silicate layers are directly fixed in the polymer matrix without aggregation

ip
of the silicate layers.

cr
5.4.2. Factors affecting polymer melt intercalation

us
Predicting whether or not a polymer-silicate nanocomposite will form through

an
melt compounding is not straightforward, as a wide variety of factors influence the

outcome. These include energy changes, arising from the confinement of the polymer
M
within the silicate, the expansion of the spaces between the layers of the silicate, and

those associated with intermolecular interaction among silicate surface, tethered chain
d

and polymer. With a view to improving predictability, attempts have been made to
te

model the behavior of hybrids that form as a result of direct melt intercalation with
p

organically modified clays and to assess the parameters required to favor intercalation
ce

[56]. Some of the questions that need to be addressed are, for example, why do certain

polymer-silicate systems favor intercalated hybrids, others delamination, and yet


Ac

others are immiscible, leading to micro composites? How does packing density and

chain length of the alkylammonium chains in the organosilicate layer, charge of the

silicate or specific groups on the polymer (or the alkyl chains) affect hybrid formation

and miscibility? How does temperature or shear affect processing? And finally, how

does the type of bonding at the polymer/silicate interface (i.e. hydrogen, dipole-

dipole, van der Waals or covalent in which the alkylammonium chains become part of

the polymer chain) affect the properties of the hybrid [4].

59
Page 59 of 302
Thermodynamic aspects. To address some of these questions Giannelis et al.

focused on the thermodynamics governing nanocomposite formation. To that end,

they developed a mean-field, lattice-based thermodynamic model. Assuming the

t
configurations and interactions of the various constituents are independent, the free

ip
energy change of hybrid formation can be separated into independent enthalpic and

cr
entropic terms. The entropic term is the sum of the configurational changes associated

with the polymer and the silicate (including the alkylammonium chains in

us
organosilicates). Configurational changes of the silicate are determined using a

an
modified Flory-Huggins lattice model in which the occupation of the lattice is

weighted to simulate the preferred orientations of the alkylammonium cations in the


M
presence of two impenetrable surfaces (silicate layers). The confinement of the

intercalated polymer chains is similarly approximated using a self-consistent field


d

treatment of a random-flight polymer with excluded volume between two surfaces.


te

For the enthalpic term a modified mean-field, site-fraction approach, where the

number of contacts per lattice site is replaced by an interaction area per lattice site, is
p

used. This modified approach allows one to express the interaction parameter as
ce

energy per area and may be approximated by interfacial or surface energies [1, 4].
Ac

In general, the conclusions of the mean-field model developed, which are widely

accepted by other researchers, may be summarized as follows. Since the spacing (or

“gallery”) between the sheets is on the order of 1 nm, which is smaller than the radius

of gyration of typical polymers, there is obviously a large entropic barrier that inhibits

the polymer from penetrating this gap and intermixing with the clay [9, 11, 147]. In

this case, of course, unlike the solution intercalation method, the decreased entropy

due to the confinement of the polymer is not compensated by an increase due to

60
Page 60 of 302
desorption of solvent molecules. However, this entropy loss, associated with the

confinement of a polymer melt is not prohibitive to nanocomposites formation,

because an entropy gain, associated with layer separation and greater conformational

energy of the aliphatic chains of the alkylammonium cations, balances the entropy

t
loss of polymeric intercalation, resulting in a net entropy change near zero. Thus,

ip
from the theoretical model, the outcome of hybrid formation via polymer melt

cr
intercalation depends on energetic factors which may be determined from the surface

energies of the polymer and OMLS. Thus, whether a mixture of polymer and OMLS

us
produces an exfoliated or intercalated nanocomposite or a conventional

an
microcomposite depends critically upon the characteristics of the polymer and the

OMLS, including the nature of the polymer as well as the type, packing density and
M
size of the organic modifiers on the silicate surface [1, 2, 4, 41, 47, 147, 161].

It is worth mentioning that even when the surfactant chains are miscible with the
d

polymer matrix, a complete layer separation depends on the establishment of very


te

favorable polymer-surface interactions to overcome the penalty of polymer

confinement. If this is not the case, good dispersion of the particles may be achieved
p

by the help of strong shear forces during the preparation and processing of the
ce

nanocomposite materials; the system, however, remains thermodynamically unstable.

This can be observed, e.g. in the case of PP or PE nanocomposite materials prepared


Ac

by melt mixing of the polymers with surface modified clay using high shear forces. If

such a mixture is heated (e.g. during processing) to temperatures above the melting

temperature, a (partial) re-agglomeration of the particles takes place [2] immediately.

The effect of layered silicates and their organic modification. In order to achieve

clay exfoliation, the interlayer structure of the OMLS should be optimized to

61
Page 61 of 302
maximize the configurational freedom of the functionalizing chains upon layer

separation and to maximize potential interaction sites at the interlayer surface [47].

Therefore, as already mentioned in Section 3.2, the organic modification of the clay is

a very important factor affecting the resulting structures. In this respect, the type of

t
surfactant, the chain length and the packing density may play an important role.

ip
In a detailed study, Fornes et al. [51] used various amine compounds to exchange

cr
the sodium ion of native montmorillonite clay. The selection of amines shown in Fig.

us
26a permitted the authors to make six structural comparisons, as presented in Fig.

26b. Among different variables, three surfactant structural issues were found to

an
significantly affect nylon 6 nanocomposite morphology and properties: decreasing the

number of long alkyl tails from two tallows to one, use of methyl rather than hydroxy-
M
ethyl groups, and use of an equivalent amount of surfactant with the montmorillonite,

as opposed to adding an excess, led to greater extents of silicate platelet exfoliation.


d

However, the authors emphasized that these effects may be specific to nylon 6
te

matrices.
p

In fact, Hotta and Paul [162] prepared PE/clay nanocomposites by melt


ce

compounding various combinations of LLDPE-g-MA, LLDPE and two organoclays

in a co-rotating twin screw extruder with a barrel temperature set at 200 °C and a
Ac

screw speed of 280 rpm. Nanocomposites based on the organoclay having two alkyl

chains are superior to the nanocomposites based on the organoclay having one alkyl

chain, in terms of clay dispersion. This was attributed to the relatively better affinity

of LLDPE for the alkyl chains than for the silicate surface. Therefore, it is reasonable

that, in this case, increasing the number of alkyl chains should lead to better

dispersion of the organoclay. As deduced from these findings, the conditions favoring

exfoliation may be quite different depending on the specific system investigated.

62
Page 62 of 302
The effect of clay organo-modification on the morphology of EVA-based

nanocomposites has also been investigated. In a relevant study EVA was melt mixed

with two clays: Cloisite Na+ and Cloisite 30B (modified with methyl-tallow-bis-2-

hydroxyethyl ammonium ions). Mixing was performed in a Brabender Laboratory

t
Mixer at 160 °C. Even though, quite surprisingly, XRD measurements revealed a

ip
decrease in clay interlayer spacing after blending with EVA, for both

cr
montmorillonites, the authors claimed that nanocomposites were formed in the case of

Cloisite 30B, on the basis of TEM observations. However, microstructures were

us
obtained in the case of Cloisite Na+ [163].

an
Zhang et al. [164] also synthesized EVA/clay nanocomposites through a melt

blending method, using different EVAs and octadecyltrimethyl ammonium,


M
dioctadecyldimethyl ammonium, and tricetadecymethyl ammonium ion exchange

MMTs. Again, EVA chains intercalated into the organomodified MMT sheets, but in
d

the case of sodium montmorillonite (Na+-MMT) there was no such intercalation.


te

Moreover, differences in the intercalation behavior have also been observed

between clays modified with different surfactants. For example, two different EVAs
p

containing 12 and 19 % vinyl acetate, abbreviated as EVA-12 and EVA-19,


ce

respectively, were processed into their nanocomposites with synthetic clay

fluorohectorite (FH) organomodified by octadecyl ammonium ion (ODA) and


Ac

aminododecanoic acid. These materials were melt-blended at 120 °C in a twin screw

microcompounder to obtain nanocomposites. Octadecyl ammonium ion intercalated

FH (FH-ODA) formed a delaminated nanostructure, whereas ammonium dodecanoic

acid intercalated FH formed a microcomposite. This means that the octadecyl

ammonium ion is more compatible than ammoniumdodecanoic acid when intercalated

in FH during the synthesis of EVA/FH nanocomposites [165].

63
Page 63 of 302
In another study by Alexandre et al. [166], several exchanging cations bearing

either simple alkyl chains or aliphatic chains terminated by a carboxylic group were

studied for modifying montmorillonites as described in Table 10. Nanocomposites

were only formed when EVA copolymers were melt mixed with unfunctionalized

t
organo-montmorillonites, such as montmorillonite exchanged with

ip
dimethyldioctadecyl ammonium (Mont-2CNC18), whereas with the same EVA

cr
matrix, the use of ammonium cations functionalized with carboxylic groups did not

lead to the formation of an intercalated structure, indicating that functionalization of

us
the clay interlayer is detrimental to the intercalation process.

an
In order to determine the effect of the nature of the clay and clay organic modifier

on nanocomposite morphology and properties, Peeterbroeck et al. [167] used EVA


M
copolymer containing 27 wt % VA and various commercial organoclays, presented in

Table 11. EVA and clay (5 wt %) were compounded in a two-roll mill for 12 min at
d

140 °C. The results of XRD analysis are reported in Table 12. Evidently, while all the
te

tested organomodified clays result in both intercalated and exfoliated structures, the

nanocomposites based on Cloisite 30B display the highest exfoliation and clay
p

stacking destruction, characterized by the absence of a characteristic XRD peak. This


ce

better filler dispersion might arise from interactions between the acetate functions of

EVA and the hydroxyl-bearing ammonium cations that modify Cloisite 30B. It is also
Ac

worth noticing that, whatever the clay nature, dispersion of unmodified clays (Cloisite

Na, Nanofil 757 or Somasif ME100) in EVA is characterized by the formation of

microcomposites, since no significant increase in the basal spacings recorded for these

materials can be observed. When comparing Cloisite 20A, Nanofil 15 and Somasif

MAE, characterized by clays of various origins but modified by the same ammonium

64
Page 64 of 302
cation, one can remark that the final interlayer spacings are very close, independent of

the interlayer spacing of the clays used or the amount of modifier.

Concerning the conformation of surfactant chains, it is generally accepted that at

low interlayer packing densities of the organic modifier, the chains adopt a disordered

t
monolayer arrangement. As the packing density increases, the chains adopt more

ip
extended conformations (and thus larger initial gallery heights), ultimately resulting in

cr
a solid-like paraffinic arrangement of the chains [47].

According to Vaia and Giannelis, the optimal structure appears to exhibit a chain

us
arrangement slightly greater than a pseudo-bilayer. That is, there is an optimum

an
interlayer structure favoring hybrid formation that is intermediate between a

disordered monolayer and a solid-like paraffinic arrangement of aliphatic chains. The


M
difference between primary and quaternary ammonium head groups did not appear to

be a predominant factor, at least for the polystyrene intercalation, which they


d

examined [47].
te

Ginzburg et al. calculated phase diagrams and showed that as the length of the

grafted chains and/or their density is increased, the miscibility between the clay sheets
p

and the polymer is improved and the resulting mixture can exhibit exfoliated structure
ce

for a range of clay volume fractions. According to their work, for short surfactant

molecules, the polymer is unable to penetrate the gallery between the clay surfaces,
Ac

and the equilibrium morphology becomes immiscible (two-phase) for most values of

the Flory-Huggins parameter and the clay volume fraction. Only in the limit of large

negative χ (strong attraction between grafted chains and polymer melt) can such a

composite become exfoliated [9].

Next to these studies, Balazs and coworkers proposed a theoretical model that

uses numerical self-consistent field calculations to study the effect on morphological

65
Page 65 of 302
behavior of varying the surfactant–matrix enthalpic interaction, surfactant coverage,

and surfactant length. The model indicated that longer surfactants (clay organic

modifiers) promote intercalation of matrix molecules in the modified clay galleries by

providing a reduction in the entropic penalty to the intercalating polymer. However,

t
very high degrees of surfactant coverage on the silicate surface make intercalation and

ip
exfoliation unfavorable [168].

cr
This last conclusion is consistent with that from the experimental work of Kurian

et al. These authors initially prepared organically modified clays by exchanging most

us
cation exchange positions in Na+-MMT with custom-made quaternary ammonium ion

an
terminated PS surfactants of five different molecular weights. They showed that high

levels of modification resulted in dense polymer brushes on the clay surfaces,


M
preventing intercalation of PS homopolymer molecules. Phase-separated

morphologies were observed in all cases, regardless of surfactant molecular weight.


d

This was explained within the framework of well-established theories of dewetting


te

from dense polymer brushes [169].

Thus, in a later study the authors described a new scheme termed mixed coverage,
p

wherein the silicate is modified with a mixture of PS-based surfactants of different


ce

lengths, to create a silicate surface grafted with a bimodal brush. Texturing of the

silicate surface was expected to allow intercalation of matrix molecules in the


Ac

galleries by providing a favorable entropic gain to the intercalating polymer through

interaction with the longer grafted molecules, and reduced enthalpic interaction with

the clay surface. However, XRD data indicated that the samples remained in a phase-

separated morphology, with no indication of intercalation or exfoliation of the layered

silicate in the PS matrix. To explain this result, the authors suggested that, despite the

bimodal nature of the brushes in this experiment, it is still possible that the brush is

66
Page 66 of 302
still effectively dense (only 17 % of the exchangeable cations were replaced by the

longer surfactant and the remaining 83 % by the short surfactant), making the long

surfactant fraction ineffective in fostering wetting and intercalation. Figure 27 shows

a schematic of this scenario. In Fig. 27a, a long surfactant is illustrated at low

t
coverage over a surface that is repulsive to the polymer matrix (PS homopolymer). In

ip
this case, the surfactant will stretch away from the surface until the energy required to

cr
stretch is no longer less than that of the repulsive enthalpic interactions. This is the

scenario for typical surfactant-modified clay at very low coverage. In the bimodal

us
brush, the shorter surfactant at full coverage (83%) effectively replaces the silicate

an
surface. As illustrated in Fig. 27b, the surface now can be thought of as being

comprised of the dense brush. When a small fraction of those surfactants are replaced
M
with longer surfactants, the situation may become something like that illustrated in

Fig. 27c. The length of the longer surfactant is still important, but the conformation of
d

the longer surfactant must also be considered. If the mixed surfactants used to create
te

the textured surface do not differ much in their lengths, one has effectively a short

surfactant at high coverage. As the molecular weight difference between the two
p

surfactants increases, the surfactant present in lesser amount becomes effectively


ce

longer, but still it is possible that it does not actually assist in the wetting and

intercalation processes, because there is now no enthalpic repulsion causing the


Ac

surfactant to stretch away from the silicate surface. This would result in the scenario

illustrated in Fig. 27c, where the longer surfactant now remains close to the brush of

chemically similar molecules and does not aid intercalation by interacting with matrix

polymer molecules [169]. This work clearly demonstrates that when tailoring the

silicate characteristics that favor intercalation or exfoliation, one has to

67
Page 67 of 302
simultaneously consider different factors, e.g. the surfactant chain length and packing

density.

Summarizing the above, the simple organic modification of the clay through ion-

exchange reactions is not always enough to achieve nanocomposite formation. This is

t
because an ideal compatibilization agent between two intrinsically incompatible

ip
components should have (combined in one molecule) parts which mix

cr
thermodynamically stable and easily with both components [2]. In fact, Balazs et al.

likened the situation found in nanocomposites to the behavior of fiber-reinforced

us
composites. The clay substrate represents the surface of a fiber and the “surfactants”

an
represent a coating, which is applied to enhance the adhesion between the fiber and

polymer matrix [11]. Unfortunately, surfactants fulfill only partly this requirement.
M
That is, the ionic part interacts certainly in a favorable way with the charged surface

of the sheet-like clay particles; however, the long alkyl tail displays only a limited
d

compatibility with the polymer chains. Second, as stated above, polymer-surface


te

interactions are also important for a complete and stable dispersion. Therefore, better

compatibilization should be expected from “macro-surfactants” like block or graft


p

copolymers combining blocks which can interact with the solid particle surface and
ce

with the matrix polymer – thus meeting most of the requirements listed above (Fig.

28) [2]. For example, the clay compatible block could be PEO. The PEO block acts as
Ac

a complexing group for the sodium ions located between the sheets of the inorganic

clay. Therefore, an ion-exchange reaction may not even be necessary. Alternatively,

the block copolymer may also be used as a “co-surfactant” with the surfactant on the

surface of organo-clays, thus fine-tuning the thermodynamic characteristics of the

clay surface with respect to the polymer matrix material. Toyota, for example,

developed such a process for the incorporation of MMT sheets into PP. In this

68
Page 68 of 302
procedure, a first ion-exchange step with double tailed ammonium cations is followed

by a further incorporation of end-group modified (grafted) oligomeric polypropylene

between the organically modified sheets and a subsequent mixing with the matrix

polymer. This process relies on the compatibilizing action of the functionalized

t
oligomers [2].

ip
Finally, it is important to note that, apart from the clay chemistry, the amount of

cr
clay incorporated in the polymer matrix also plays a determining role. In fact, it is

often reported that while low clay loadings favor exfoliation, higher amounts of clay

us
(usually above 10 wt %) allow only intercalation of polymer chains in the layered

an
silicate galleries to occur. Further dispersion of clay platelets is hindered, most

probably due to percolation phenomena [2, 170].


M
The effect of the polymer matrix. Polymer matrix parameters may also determine
d

the outcome of melt blending a polymer and a layered silicate.


te

In this context, the effect of the matrix molecular weight has been considered.

Early work by Vaia and Giannelis showed that, for statically annealed PS samples, the
p

final hybrid structure is independent of the molecular weight of the polymer. Only the
ce

time needed for intercalation to proceed was different, going from 6 h for Mw of

30000 to 24 h for 90000 and 48 h for 400000 at 160 °C. Clearly, high Mw PS
Ac

decreases the kinetics of intercalation by decreasing the diffusivity of the polymer in

the interlayer. However, the authors noted that additional experimental work

examining the intercalation behavior of a broader range of polymer molecular weights

coupled with dynamic blending of constituents to enhance equilibrium mixing would

be required to further explore this issue [47].

69
Page 69 of 302
In another study, Fornes et al. [14] prepared nanocomposites based on three

different molecular weight grades of nylon 6 (low, medium and high) using a co-

rotating twin screw extruder. WAXD and TEM results collectively reveal a mixed

structure for the LMW based nanocomposites, having regions of intercalated and

t
exfoliated clay platelets. Qualitative TEM observations were supported by a

ip
quantitative analysis of high magnification TEM images. The average number of

cr
platelets per stacks was shown to decrease with increasing molecular weight, thereby

revealing larger extents of clay platelet exfoliation for the nanocomposites in the order

us
HMW>MMW>LMW composites. As a result, tensile tests revealed superior

an
performance for the higher molecular weight nylon 6 composites, particularly those

based on HMW, as will be discussed in Section 6.1.2.


M
However, probably the most critical condition for the formation of intercalated

and especially exfoliated hybrids via polymer melt intercalation, is the presence of
d

polar type interactions (i.e. other than Van der Waals forces). Therefore, polar
te

polymers containing groups capable of associative type interactions, such as Lewis

acid-base interactions or hydrogen bonding, favor the intercalation of macromolecular


p

chains into the silicate galleries [47, 49], while in the case of apolar polymer matrices,
ce

clay delamination typically requires the use of compatibilizers, as discussed in the

following paragraph.
Ac

A good example demonstrating the importance of polar interactions is the

formation of EVA-based nanocomposites via melt intercalation. It is well established

that the presence of polar groups (ester groups of the vinyl acetate moieties) all along

the chains improves the ability of these polymers to intercalate in organo-modified

montmorillonites [1]. Therefore, several studies have focused on the effect of the

vinyl acetate content on the dispersion of clay nanoplatelets. In general, it has been

70
Page 70 of 302
observed that the higher the vinyl acetate content the larger is the basal spacing

increase of the clay, inducing the formation of intercalated to exfoliated

nanostructures.

In a representative study, Chaudhary et al. [171] prepared EVA-based

t
nanocomposites via melt intercalation using an intermeshing counter-rotating twin

ip
screw extruder. They used EVA copolymers with 9, 18 and 28 % VA (vinyl acetate)

cr
and two organomodified clays: Cloisite 15A (C15A), which is more suitable for the

less polar EVA9 due to long aliphatic chains in C15A, and Cloisite 30B (C30B),

us
suitable for the more polar polymers, like EVA18 or EVA28. They prepared

an
composites with filler level varying from 2.5 to 7.5 wt %, and subsequently

characterized the structures obtained by WAXD and TEM. The results indicated that
M
only intercalation occurred in the case of the less polar EVA9, while the clay was

exfoliated in the more polar EVA18 and EVA28. Therefore, the authors concluded
d

that an increase in the content of polar VA groups in EVA 18 and EVA28 as


te

compared to EVA9, which lowered the thermodynamic energy barrier for clay-

polymer interaction, possibly allowed a relatively higher number of polymer chains to


p

migrate and stabilize within the clay platelet and form partially exfoliated and/or
ce

disordered intercalated states.

It is worth mentioning that despite the presence of polar groups in EVAs, polymer
Ac

compatibilizers have been used in the preparation of EVA-based nanocomposites. For

example, Li and Ha [172] selected a maleic anhydride grafted EVA containing 18 mol

% VA to process its nanocomposites with organomodified Cloisite through melt

blending at 175 °C and found that the dispersion of Cloisite in the maleic anhydride

grafted EVA was much better than in the simple EVA matrix.

71
Page 71 of 302
The effect of melt intercalation processing conditions. Melt processing conditions

play a key role in achieving high levels of exfoliation. Indeed, nanocomposites have

been formed using a variety of shear devices (e.g. extruders, mixers, ultrasonicators,

etc), among which twin screw extruders have proven to be most effective for the

t
exfoliation and dispersion of silicate layers [14].

ip
The screws in twin screw extruders intermesh so that the relative motion of the

cr
flight of one screw inside the channel of the other acts as a paddle that pushes the

us
material from screw to screw and from flight to flight. Two different patterns for

intermeshing twin-screw extruders are possible. In the co-rotating pattern the screws

an
rotate in the same direction and the material is passed from one screw to another and

follows a path over and under the screws. This gives high contact with the extruder
M
barrel, which improves the efficiency of heating. The path also ensures that most of

the resin will be subjected to the same amount of shear as it passes between the
d

screws and the barrel. The self-wiping nature of the co-rotating screws is much more
te

complete than in the counter-rotating system, thus in the co-rotating case there is less
p

likelihood that material will become stagnant. In the counter-rotating pattern, on the
ce

other hand, the screws rotate counter to each other and the material is brought to the

junction of the two screws, building up in what is called the material bank on the top
Ac

of the junction. This buildup of material is conveyed along the length of the screw by

the screw flights. As the material passes between the screws, high shear is created, but

shear elsewhere is very low. Since only a small amount of material passes between

the screws, total shear is lower than in single-screw extruders and in co-rotating twin-

screw extruders. Therefore, co-rotating systems are more effective than either

counter-rotating or single-screw extruders [173]. On the other hand, although it is

often stated that twin screw extruders favor intercalation when compared to single

72
Page 72 of 302
screw systems, this may not be the case for counter-rotating extrusion systems, for the

aforementioned reason.

Focusing on the effect of the extrusion system on the degree of intercalation, Cho

and Paul [15] prepared nylon 6/o-MMT nanocomposites using either an intermeshing

t
co-rotating twin screw extruder or a single screw extruder. They found that for the

ip
composite prepared by single screw extrusion, full exfoliation is not achieved, which

cr
was attributed to insufficient amount of shear and short residence time. On the other

hand, by mixing in the twin screw extruder, the organoclay is uniformly dispersed

us
into nylon 6 and the individual layers are aligned along the flow axis.

an
However, other researchers have reported on nanocomposite preparation using

single-screw extruders. For example, McNally et al. [36] successfully prepared


M
PA12/clay nanocomposites using conventional single-screw melt blending.

Moreover, it is important to notice that, even for a given extruder, processing


d

conditions may determine the outcome. More specifically, increasing the mean
te

residence time in the extruder generally improves the delamination and dispersion.

However, there appears to be an optimum extent of back mixing and an optimum


p

shear intensity; excessive shear intensity or backmixing apparently causes poorer


ce

delamination and dispersion [52]. Often, special screw designs, including provisions

for additional mixing, or static mixers at the end of the screw are used to enhance
Ac

mixing and thus the silicate dispersion [173].

Li et al. [174] developed an ultrasonic extrusion technology, which organically

combines extruder and ultrasound power. The authors claimed that the introduction of

ultrasonic irradiation in extrusion processing of polymer can improve the

processibility of polymer materials, and also reduce the size and size distribution of

dispersed particles in polymer blends. They used the ultrasonic oscillation extrusion

73
Page 73 of 302
system developed to prepare polymer/MMT nanocomposites. The system, consisting

of an extruder and a cylinder die connected to a generator of ultrasonic oscillations in

the direction parallel to the flow of the polymer melt, was found to improve the

exfoliation of the clay, though only for specific matrices.

t
It should be noted at this point that, apart of the various extrusion systems, internal

ip
mixers (i.e. batch mixing devices where mixing occurs in a closed chamber)may also

cr
be successfully used for the preparation of exfoliated nanocomposites, as

demonstrated, for example, in the case of PEI matrix [46, 52]. However, these devices

us
appear to be much less popular in nanocomposite preparation.

an
Another factor affecting the resulting structures in the case of crystalline polymer

matrices, is the crystallization temperature. For example, Okamoto et al. [175, 176]
M
observed through X-ray analyses that the intergallery spacing of PP-MA based

nanocomposites increases with the crystallization temperature Tc for any amount of


d

clay content in the nanocomposites. The microstructure of the nanocomposites,


te

observed directly by TEM, showed that the clay particles are well dispersed at low Tc

and that segregation of silicate layers occurs at high Tc. This implies that, by
p

controlling intercalation through crystallization at a suitable temperature, one can


ce

control the fine structure, the morphology – and thus the properties of crystalline

polymer/clay nanocomposites.
Ac

Conclusively, a number of factors affect the outcome of melt intercalation. In this

context, Dennis et al. [52] presented a simplified scheme to underline the conditions

under which clay exfoliation into a polymer occurs during melt blending. The

proposed scheme (Fig. 29a) is based on the relationship between the compatibility of

the chemistry of the clay treatment and the matrix and the process conditions used to

make a nanocomposite. It distinguishes three cases. The first case is chemistry-

74
Page 74 of 302
dependent. When the clay chemical treatment and the resin are compatible, almost

any set of processing conditions can be used to form an exfoliated nanocomposite. In

case 2, clay chemical treatment and polymer are marginally compatible. In this

situation, the process conditions can be optimized to give an exfoliated

t
nanocomposite. That is, the organoclay chemical treatment and the matrix are

ip
compatible enough that processing conditions can be tailored to optimize

cr
delamination and dispersion. Finally, in case 3, there is no apparent compatibility of

the clay chemical treatment and the polymer. Processing conditions can be optimized

us
to give intercalants or tactoids that are minimized in size, but even partial exfoliation

an
does not occur.

These authors also presented possible clay delamination paths (Fig. 29b) to
M
demonstrate that increasing shear intensity is not enough to achieve exfoliation. In

pathway 1, stacks of platelets are decreased in height by sliding platelets apart from
d

each other, a pathway that requires shear intensity. Pathway 2 shows polymer chains
te

entering the clay galleries pushing the ends of platelets apart. This pathway does not

require high shear intensity, but involves diffusion of polymer into the clay galleries
p

(driven by either physical or chemical affinity of the polymer for the organoclay
ce

surface) and is, thus, facilitated by residence time in the mixing device. As more

polymer enters and goes further in between clay platelets, especially near the edge of
Ac

the clay galleries, the platelets appear to peel from the edge, since they are able to

bend [14, 52].

5.4.3. Compatibility issues in non-polar polymers

In contrast to polar polymers, like polyamides, that can effectively exfoliate

organically modified clays using conventional melt processing techniques, for

75
Page 75 of 302
nonpolar polymers, such as the most widely used polyolefins, PE or PP, synthesis of

well exfoliated nanocomposites appears to be more difficult, because these polymers

are so hydrophobic and lack suitable interactions with the clay surface, even after it

has been organically modified [49, 162, 177-179]. However, the development of

t
polyolefin/clay nanocomposites is a field of rapidly growing industrial relevance due

ip
to their promise of improved performance in packaging and engineering applications

cr
[111]. Therefore, ways to resolve the difference in polarity between polyolefins and

clays, in order to prepare nanocomposites by conventional melt compounding, have

us
been proposed.

an
More specifically, initial attempts to create nonpolar polymer/clay

nanocomposites by simple melt mixing were based on the introduction of a modified


M
oligomer to mediate the polarity between the clay surface and the polymer [2, 49,

177]. The most promising strategy at the present time is to add a small amount of a
d

maleic anhydride grafted polyolefin that is miscible with the base polyolefin. It is
te

believed that the polar character of the anhydride has an affinity for the clay materials,

such that the maleated polyolefin can serve as a “compatibilizer” between the matrix
p

and the filler [162, 180, 181].


ce

In fact, Zhai et al. [179] showed that two kinds of hybrids are formed by melt

mixing: an o-MMT with PE and with PE-g-MA, respectively. For the PE/o-MMT
Ac

system the intercalate effect is limited and the dispersion of clay is unsatisfactory.

However, for the PE-g-MA/o-MMT nanocomposites, MMT was exfoliated in the

matrix, as testified by both XRD and TEM.

Wang et al. [49] prepared several types of nanocomposites with different

compositions of the organically modified clay and maleated polyethylene by melt

compounding at 140 °C using a Brabender mixer operating at 60 rpm for 20 min.

76
Page 76 of 302
They found that the alkylammonium chain length may change the degree of

interaction between clay and polyethylene and that the original basal reflection peak

of the clay disappeared completely above a certain grafting level of MA, about 0.1 wt

%.

t
Quite interestingly, Zhang and Wilkie [182] obtained PE-organoclay

ip
nanocomposites by adding maleic anhydride directly as a compatibilizer during the

cr
melt blending. As the authors suggested, the maleic anhydride probably reacted with

PE during the high temperature blending in air, leading up to the formation of a graft

us
copolymer in which maleic anhydride units are attached to the PE chains.

an
In another study, Tang et al. [183] described the preparation of PP-based

nanocomposites through a successful combination of clay modification and


M
intercalation in one step. The authors mixed pristine MMT with the surfactant (C16)

and PP with or without PP-MA using a high speed mixer . The mixed powder was
d

then processed in a twin screw extruder and nanocomposites were obtained. The
te

results of this study showed that the structure of PP-clay nanocomposites is sensitive

to the compatibilizer and the surfactant, since their increasing concentrations will
p

reduce the free energy of the system, which is favorable for thermodynamic stability.
ce

The dispersion mechanism proposed is the following. At first, some surfactant chains

diffuse into the interlayer under physical absorption and shear, rendering the clay
Ac

organophilic. However, some surfactant remains in the polymer matrix, which may

enhance the compatibility when the matrix is intercalated into the interlayer. In fact,

the authors suggested that there is some interaction between the polymer matrix and

the surfactant, just as in the interaction between the surfactant and the silicates. On the

other hand, some PP-MA may be intercalated into the interlayer of MMT, after the

surfactant makes the clay sufficiently organophilic. At the same time, PP-MA may act

77
Page 77 of 302
as a high molecular weight surfactant. The interlayer spacing of the clay increases

and, if the miscibility of PP-MA with PP is good enough to allow dispersion at the

molecular level, the exfoliation of intercalated MMT should take place.

In addition to maleic anhydride and maleic anhydride grafted PE, EVA has also

t
been used as a compatibilizer to prepare PE-based nanocomposites. For example,

ip
Zanetti and Costa [177] prepared several types of composites with different PE/EVA

cr
ratios and 5 wt % organoclay by melt compounding at 150 °C using a Brabender

internal mixer with a screw speed of 60 rpm for 10 min. The polymer EVA contained

us
19 wt % VA. No interaction was obtained by compounding the PE with the clay in

an
absence of a compatibilizer. However, 1 wt % EVA was enough to intercalate all the

organoclay. Further increasing the amount of EVA above 10 wt % caused a decrease


M
in the degree of coherent layer stacking (i.e. a more disordered system).

Zhao et al. [184] investigated chlorosilane-modified montmorillonites and their


d

results showed that intercalated PE nanocomposites were obtained by melt


te

intercalation using common alkylammonium intercalated clay, which was pretreated

with chlorosilane. In a later work, the authors used directly a reactive intercalating
p

agent (N-γ-trimethoxyl-silanepropyl) octadecyldimethylammonium chloride


ce

(abbreviated JSAc) to modify the montmorillonite clay, so that the chemical reaction

with hydroxyl groups at the edge of the clay layers and the interlayer ion exchange
Ac

were carried out in one step. PE/clay nanocomposites were then directly prepared by

melt intercalating PE and the above mentioned clay in a twin screw extruder at 180°C

and 200 rpm, whereas only microcomposites were formed when using common

intercalating agent.

It is also worth mentioning the work of Preston et al. [178], who prepared

nanocomposites using the following matrices: two EVAs with different vinyl acetate

78
Page 78 of 302
contents, poly(ethylene-co-methyl acrylate) (EMA), poly(ethylene-co-methyl

acrylate-co-acrylic acid) (EMAAA), and a blend of LDPE with maleated ethylene

copolymer (PE-g-MA). Structures for each of these materials are given in Table 13.

The organoclay they used was organically modified bentonite clay. Composites were

t
prepared by melt mixing in a twin screw extruder operating at 380 rpm with a screw

ip
configured for intensive mixing. Through XRD measurements, the authors concluded

cr
that no interaction was likely between the LDPE and the silicate, whereas

intercalation of the organoclay occurred in the presence of the four polar polymers.

us
As in the case of PE, it is difficult to get exfoliated and homogenous dispersion of

an
the silicate layer at the nanometer level in polypropylene, due to its low polarity.

Consequently, PP also is usually modified with polar oligomers prior to introduction


M
of modified clay, in order to achieve nanometric dispersion of the clay [161].

One typical example is the PP/clay nanocomposite system described by Toyota.


d

The Toyota research group prepared PP/clay nanocomposites by direct melt


te

compounding of PP with organo-modified MMT, in the presence of a maleic

anhydride grafted PP (PP-g-MA). They added three times as much PP-g-MA as the
p

clay by weight to prepare well mixed PP/clay nanocomposites, and pointed out that
ce

the miscibility between maleated oligomer and matrix polymer played a key role in

composite properties [2, 49, 177, 185]. In fact, it has been suggested that the relative
Ac

content in maleic anhydride cannot exceed a given value, in order to retain some

miscibility between PP-MA and PP chains. When too many carboxyl groups were

spread along the polyolefin chains, no further increase in the interlayer spacing was

obtained in PP/PP-g-MA/clay blends, leading rather to the dispersion of PP-g-MA

intercalated clay in the PP matrix [1].

79
Page 79 of 302
Similarly, Hasegawa et al. [186] reported the preparation of exfoliated PP-based

nanocomposite by melt blending PP-g-MA and organically modified MMT at 200 °C,

using a twin screw extruder. Figure 30 shows a schematic representation of the clay

dispersion process in PP-MA-based nanocomposites. According to the authors, the

t
driving force of exfoliation originates from the strong hydrogen bonding between the

ip
MA groups and the polar clay surface.

cr
Kato et al. [188] prepared PP-based nanocomposites by the melt intercalation of

PP chains modified by either maleic anhydride (PP-g-MA) or hydroxyl groups (PP-

us
OH) in o-MMT. For both matrices, intercalated nanocomposites were recovered after

an
melt blending at 200 °C for 15 min. However, a PP-g-MA matrix with a lower maleic

anhydride content did not intercalate under the same conditions, indicating that a
M
minimal functionalization of PP chains has to be reached for intercalation to proceed.

The authors also noticed that intercalation increased with the polymer-to-clay ratio,
d

i.e. when the PP-g-MA fraction was increased.


te

Using the same method, Okamoto et al. [189] prepared PP/MMT nanocomposites.

The authors mixed PP-g-MA (0,2 wt % MA) and different amounts (2.4 and 7.5 wt
p

%) of C18-MMT in a twin screw extruder at 200 °C and obtained nearly exfoliated


ce

structures formed when 2 wt % clay was added. However, addition of 4 and 7.5 wt %

clay led to disordered intercalated nanocomposites and ordered intercalated structures,


Ac

respectively.

Lopez et al. [161] used two different polar coupling agents, diethyl maleate

grafted PP (PP-g-DEM) and commercial maleic anhydride grafted PP (PP-g-MA) to

prepare PP-based nanocomposites. DEM was chosen as the compatibilizing agent,

because of its high thermal stability, high boiling point, and good compatibilization

with polyolefins, compared to other compatibilizing agents. Furthermore, the low

80
Page 80 of 302
homopolymerization behavior of DEM allows better control of the functionalization

reaction. Maleic anhydride was used as reference, since it has been widely used as

compatibilizer for this kind of system. The PP/clay hybrids were prepared by melt

compounding with two different clays, commercial modified montmorillonite and

t
sodium bentonite modified with octadecylammonium ions. The results showed that

ip
although the commercial clay outperforms the octadecylammonium treated bentonite,

cr
differences in mechanical properties when using different clays are smaller if DEM is

used instead of MAH. This is a consequence of the very low degree of

us
compatibilization between the polymer matrix and the clay. In fact, this study proves

an
that clay dispersion and interfacial adhesion are greatly affected by the kind of matrix

modification. DEM has a lower polarity compared to MAH, providing a less effective
M
interaction with the polar components of the clay. The authors therefore, concluded

that clay and matrix modification are synergistic factors which need to be properly
d

modulated in order to obtain the desired final properties in this kind of nonpolar
te

polymer-based nanocomposite.

Finally, as in the case of PE and PP matrices, compatibilization is a critical issue


p

also for other polymers, such as PS. Therefore, Wang and Wilkie [190] prepared
ce

PS/clay nanocomposites by in situ reactive blending with both the organically

modified clays and the pristine inorganic clay in the presence of maleic anhydride,
Ac

and found that maleic anhydride increases the possibility of nanocomposite formation.

Also, Hasegawa et al. [191] produced partially exfoliated PS/clay nanocomposites by

compounding in a twin-screw extruder organically modified MMT with an blend of

PS and ≥ 50% of another compatibilizer, namely poly(styrene-co-methyl vinyl

oxazoline).

81
Page 81 of 302
5.4.4. Degradation problems encountered during melt intercalation

Despite the aforementioned advantages of polymer melt intercalation, this

technique may involve polymer degradation problems since, when preparing

clay/polymer nanocomposites using melt blending, a certain temperature is needed in

t
ip
the processing. Also, apart from the polymer matrix degradation, if the processing

temperature to make the PLS is beyond the thermal stability of the organic treatment

cr
on the OMLS, some decomposition will take place. The onset temperature of

us
decomposition of the organic modifier is, therefore, important in the process to make

a polymer/clay nanocomposite, since polymer processing is normally done above 150

an
°C. Moreover, in addition to common detrimental aspects of degradation, the resulting

products may play a major but yet to be determined role in the formation of exfoliated
M
nanostructures [43].

Therefore, the degradation issues encountered during melt intercalation have been
d

addressed in several studies. For example, Finnigan et al. [88] prepared TPU
te

nanocomposites by both twin screw extrusion and solvent casting, in order to compare
p

the outcomes of these methods. The authors employed two TPUs: a soft elastomer
ce

(SPU) and a hard elastomer (HPU) consisting of the same soft and hard segments, but

in different relative amounts. As demonstrated by WAXD analysis, both processing


Ac

routes led to delaminated structures, which illustrates that if there is a good driving

force for intercalation between the polymer and organosilicate, the need for an

optimized processing route is diminished. Although melt compounding offered

slightly better silicate dispersion than solvent casting, the authors suggested that

solvent casting must be the preferred processing route for these materials, owing to

elimination of PU and surfactant degradation. In fact, as shown in Table 14, it was

found that the number average molecular weight (Mn) of PU significantly decreases

82
Page 82 of 302
during melt compounding and, to a smaller extent, during solvent casting (due to an

ultrasonic probe that was applied). In this particular case, the authors identified as

additional causes of thermal degradation the small size of the extruder (and thus the

large surface to volume ratio) as well as the absence of additives to reduce

t
degradation.

ip
Xie et al. [43] focused on the effect of organic modifiers on the thermal

cr
decomposition of OMLSs and found that, while different long alkyl substituents have

no effect or very little effect, the thermal degradation of organically modified

us
montmorillonite is quite different compared to that of pure montmorillonite. The

an
DTGA thermal curve shown in Fig. 31 for the OMLS was considered in four parts: (a)

the free water region below 200 °C; (b) the region where organic substances evolve in
M
the temperature range 200-500 °C; (c) the structural water region in the temperature

range 500-800 °C; (d) a region between 800 and 1000 °C where organic carbon reacts
d

in some yet unknown way. In OMLS sample the free water disappears by 40 °C.
te

There is no interlayer water in OMLS as the quaternary ammonium salt has been

exchanged for the hydrated sodium cation. The most distinguishing difference
p

between the sodium montmorillonite and the organically modified montmorillonite is


ce

in the temperature range 200-500 °C, as the organic constituent in the organo-clay

starts to decompose somewhat above 200 °C. Another distinguishing difference


Ac

between sodium montmorillonite and the organically modified montmorillonite is in

the temperature range from 800 to 1000 °C. Sodium montmorillonite is very stable

when the temperature is higher than 800 °C, however, the OMLS continues to lose

weight and a larger amount of CO2 is released at temperatures over 800 °C.

Davis et al. [192] found that during melt blending MMT/PA6 nanocomposites in a

twin-screw extruder at 240 °C, a particular quaternary alkyl ammonium treatment of

83
Page 83 of 302
the montmorillonite clay degraded it to an extent correlated with extruder residence

time. To address this issue, they conducted an investigation on the processing

degradation of PA6/MMT nanocomposites and clay organic modifier. The results led

them to the following conclusions.

t
1. PA6 nanocomposites significantly degraded during processing at 300 °C.

ip
Within experimental uncertainty, drying at 120 °C rather than 80 °C prior to

cr
processing had little effect on the degree of degradation. Virgin PA6 did not degrade

under identical processing conditions.

us
2. Thermal decomposition of PA6 nanocomposite may have resulted from

an
hydrolytic peptide scission. The catalytic activity of MMT was not investigated in this

particular study; however, on the basis of previous research, it appeared that MMT
M
could be involved in PA6 thermal degradation.

3. Heating at 120 °C for 4 h thoroughly dried virgin PA6; but drying at 80 °C


d

resulted in no water loss. The amount of volatile water in PA6 nanocomposites was
te

greater than was observed in virgin PA-6. Longer drying times and higher

temperatures resulted in drier PA6 nanocomposites.


p

4. MMT and water are responsible for the degradation of PA6 nanocomposites.
ce

5. When PA6 was processed at 300 °C some water was present, however, little

degradation was observed. This means that: (a) water itself may not be sufficient to
Ac

cause degradation, (b) water escapes from PA-6 faster than from the nanocomposite.

(c) clay and water are a special catalyst combination and/or (d) clay is a source of

high-temperature reactive water or hydroxyls.

In addition to the aforementioned studies, which focus mainly on the degradation

mechanism, others are exploring ways to overcome or limit degradation during

polymer melt intercalation. In general, when a material is subjected to extrusion,

84
Page 84 of 302
degradation is detected as discoloration and lowered physical or mechanical

properties. A strong odor may also indicate degradation. If the degradation is general,

that is, the entire extrudate is affected, as shown by discoloration throughout, although

darker streaks may also be present, the most likely cause is that the heat is too high for

t
the speed of extrusion. The obvious solutions are to reduce the heat or to increase the

ip
extrusion speed. Some combination of these two variables are likely feasible since the

cr
speed of the extruder affects mechanical heating of the material [173]. However, this

needs to be done carefully since, as mentioned above, processing conditions may

us
affect the morphology of the resulting material.

an
On the other hand, a number of researchers have developed and applied clays

exhibiting high thermal stability. In this context, Chang et al. [193] developed a
M
thermally stable montmorillonite through an ion exchange reaction between Na+-

MMT and dodecyl triphenyl phosphonium chloride (C12PPh-Cl-) (Fig. 32). Gilman et
d

al. [194] described the preparation of PA6-based nanocomposites of MMT modified


te

with trialkylimidazolium cations to obtain high stability OMLS at high processing

temperatures.
p

A surprising result reported in another study was that poly(3-hydroxybutyrate)


ce

(PHB) nanocomposites prepared via melt intercalation showed severe degradation as

testified by GPC, when an organically modified MMT was used, whereas no


Ac

degradation was found with nanocomposites based on organically modified

fluoromica. Even though there is no explanation on how organically modified

fluoromica acted to protect the system, the authors suggest that the presence of Al

Lewis acid sites, which catalyze the hydrolysis of ester linkages at high temperature,

may be one reason[195].

85
Page 85 of 302
Finally, it is worth noting at this point that, despite the aforementioned

degradation problems encountered during melt intercalation, very few authors have

used stabilization systems in the preparation of polymeric nanocomposites.

In Table 15 several PLS nanocomposites prepared via melt intercalation are

t
presented as typical examples.

ip
cr
6. Nanocomposite properties

us
6.1. Mechanical properties

an
6.1.1 The Reinforcing Mechanism of Layered Silicates

The first mechanism that has been put forward to explain the reinforcing action of
M
layered silicates is one also valid for conventional reinforcements, such as fibers,

which is schematically depicted in Fig. 33. That is, rigid fillers are naturally resistant
d

to straining due to their high moduli. Therefore, when a relatively softer matrix is
te

reinforced with such fillers, the polymer, particularly that adjacent to the filler
p

particles, becomes highly restrained mechanically. This enables a significant portion


ce

of an applied load to be carried by the filler, assuming, of course, that the bonding

between the two phases is adequate [64]. From this mechanism it becomes obvious
Ac

that the larger the surface of the filler in contact with the polymer, the greater the

reinforcing effect will be. This could partly explain why layered silicates, having an

extremely high specific surface area (on the order of 800 m2/g) impart dramatic

improvements of modulus even when present in very small amounts in a polymer. In

fact, the low silicate loading required in nanocomposites to effect significant property

improvements, is probably their most distinguishing characteristic. In most

conventionally filled polymer systems, the modulus increases linearly with the filler

86
Page 86 of 302
volume fraction, whereas for nanocomposites much lower filler concentrations

increase the modulus sharply and to a much larger extent [55].

However, some authors have argued that the dramatic improvement of modulus

for such extremely low clay concentrations (i.e. 2-5 wt %) cannot be attributed simply

t
to the introduction of the higher modulus inorganic filler layers. A proposed

ip
theoretical approach assumes a layer of affected polymer on the filler surface, with a

cr
much higher modulus than the bulk equivalent polymer. This affected polymer can be

thought of as a region of the polymer matrix that is physisorbed on the silicate

us
surface, and is thus stiffened through its affinity for and adhesion to the filler surface.

an
Obviously, for such high aspect ratio fillers as the layered silicate layers, the surface

area exposed to the polymer is huge and, therefore, the significant increases in the
M
modulus with very low filler content are not surprising. Furthermore, beyond the

percolation limit, the additional silicate layers are incorporated in polymer regions
d

that are already affected by other silicate layers, and thus it is expected that the
te

enhancement of modulus will become much less dramatic [198].

In order to prove the effect of degree of exfoliation on nanocomposite mechanical


p

properties, Fornes et al [64] used an analytical approach to elucidate how incomplete


ce

exfoliation influences nanocomposite stiffness. They expressed the modulus of a

simple clay stack in the direction parallel to its platelets, by using the rule of mixtures
Ac

E stack = φ MMT E MMT + φ gallery E gallery

where φMMT is the volume fraction of silicate layers in the stack, EMMT is the modulus

of MMT, φgallery is the volume fraction of gallery space and Egallery is the modulus of

87
Page 87 of 302
the material in the gallery, which is expected to be much less than EMMT. The volume

fraction occupied by gallery space, φgallery can be expressed in terms of X-ray d-

spacings, as

t
ip
cr
where n is the number of platelets per stack, d001 is the repeat spacing between silicate

particles, and tplatelet is the thickness of a silicate platelet. Obviously, when the number

us
of platelets in a stack is equal to one, the system represents an individual exfoliated

an
platelet. Table 16 shows how the number of platelets in a stack affects the

reinforcement factor(RF in an unexchanged, non-expandable clay (d001 = 0.96 nm) as


M
well as in an intercalated or organically modified clay (d001 = 1.8 nm). Increasing n in

both stacking scenarios leads to lower reinforcement efficiencies, especially for the
d

intercalated clay. Interestingly, the largest drop in reinforcement is experienced when


te

going from one to two platelets per stack. Overall, the trends in Table 16 show the

high sensitivity of nanocomposite stiffness to the level of exfoliation. Fornes et al.


p

concluded that stacks of platelets reduce stiffness of nanocomposites through lower


ce

effective filler moduli and reduced aspect ratio, the effects shown separately in Fig.

34.
Ac

6.1.2. Modulus and strength

In general, the addition of an organically modified layered silicate in a polymer

matrix results in significant improvements of Young’s modulus, as can be seen in

Table 17 for a number of different materials. For example, Gorrasi et al. [156]

reported an increase from 216 to 390 MPa for a PCL nanocomposite containing 10 wt

88
Page 88 of 302
% ammonium-treated montmorillonite, while in another study [201], Young’s

modulus was increased from 120 to 445 MPa with addition of 8 wt % ammonium

treated clay in PCL. Similarly, in the case of nylon 6 nanocomposites obtained

through the intercalative ring opening polymerization of ε-caprolactam, a large

t
increase in the Young’s modulus at rather low filler content has been reported,

ip
whatever the method of preparation: polymerization within organo-modified

cr
montmorillonite, polymerization within protonated ε-caprolactam swollen

montmorillonite or polymerization within natural montmorillonite in the presence of

us
ε-caprolactam and an acid catalyst [45].

an
However, exceptions to this general trend have been reported. As shown in Fig.

35, in crosslinked polyester/OMLS nanocomposites, the modulus decreases with


M
increasing clay content; in fact, the drop for the 2.5 wt % nanocomposite was greater

than expected. To explain this phenomenon, it was proposed that the intercalation and
d

exfoliation of the clay in the polyester resin serve to effectively decrease the number
te

of crosslinks from a topological perspective. The origin of the greater drop in

properties of the 2.5 wt % nanocomposites may be traced to the morphology; i.e. it


p

was observed that the sample showed exfoliation on a global scale compared to the
ce

nanocomposite containing 10 wt % clay, indicating that the crosslinking density is

inversely proportional to the degree of exfoliation [140].


Ac

Apart from the modulus, the addition of OMLS in a polymer matrix usually also

increases the tensile strength compared to that of the neat polymer material. For

example, Shelley et al. [32] reported a 175 % improvement in yield stress

accompanied by a 200 % increase in tensile modulus for a nylon 6 nanocomposite

containing 5 wt % clay. However, it should be emphasized that the effect of

nanocomposite formation on tensile strength is not as clear as in the case of the

89
Page 89 of 302
modulus since reductions of tensile strength upon nanocomposite formation have also

been reported. Such examples are included in Table 18, which lists the tensile

strengths of a number of nanocomposite materials and comparisons them with the

corresponding values for the neat polymers.

t
Most polymer-clay nanocomposite studies report tensile properties, such as

ip
modulus, as a function of clay content [31], as in Fig.36. This plot of Young’s

cr
modulus of nylon 6 nanocomposite vs. filler weight content, shows a constant large

rate of increase of modulus up to ca. 10 wt % of nanoclay, whereas above this

us
threshold the aforementioned leveling-off of Young’s modulus is observed. This

an
change corresponds to the passage from totally exfoliated structure (below 10 wt %)

to partially exfoliated – partially intercalated structure (for 10 wt % clay and above),


M
as determined by XRD and TEM [1, 55].

In another study, Liu and Wu [146] studied the mechanical performance of PA66
d

nanocomposites prepared via melt intercalation, using epoxy co-intercalated clay. The
te

tensile strength increases rapidly from 78 MPa for PA66 up to 98 MPa for PA66CN5,

but the increasing amplitude decreases when the clay content is above 5 wt %. A
p

similar phenomenon is observed in the dependence of tensile modulus of PA66CN on


ce

clay content. The smaller increase in amplitude observed with a clay loading above 5

wt % was again attributed to the inevitable aggregation of the layers at high clay
Ac

content.

Another example, Fig. 37, shows both the tensile modulus and the yield strength

of neat PA12 and the nanocomposites, which increased steadily with increasing

organoclay loading. Compared to the virgin polymer, the tensile modulus of

PA12/clay systems was improved by about 40 % upon adding only 5 wt % of clay,

while limited improvement of the tensile strength was observed by incorporating clay

90
Page 90 of 302
in the matrix. Again, it was suggested that there is an optimum clay concentration for

nanocomposite tensile strength improvement. With further increase in clay loading a

moderate decrease of tensile strength was observed, suggesting that the relative

amount of intercalation/exfoliation of the clay morphology gradually increases with

t
increasing clay content, since the tensile strength is usually sensitive to the degree of

ip
dispersion [147].

cr
Similarly, other factors that influence the degree of exfoliation, apart from the clay

content, also have an impact on nanocomposite modulus and strength.

us
This explains the variations observed in moduli of PA6 nanocomposites prepared

an
by intercalative ring opening polymerization of ε-caprolactam, with different kinds of

acids to catalyze the polymerization (Table 19). The WAXD peak intensity Im, which
M
is inversely related to the amount of exfoliated layers in the nanocomposite, also

depends on the nature of the acid used to catalyze the polymerization process. For an
d

increase in Im , a parallel decrease in Young’s modulus is observed, indicating that


te

exfoliated layers are the main factor responsible for the improvement in stiffness,

while intercalated particles, having a smaller aspect ratio, play a rather minor role [1,
p

55].
ce

Cho and Paul [15] studied the effect of mixing device and processing parameters

on the mechanical properties of polyamide nanocomposites. In the case of composites


Ac

formed by single-screw extrusion, the exfoliation of the clay platelets is not extensive.

Even after a second pass through this extruder, undispersed tactoids are still easily

observed with naked eye. However, the tensile strength and modulus were slightly

improved by the second pass. On the other hand, nylon 6 nanocomposites with good

properties can be obtained over a broad range of processing conditions in the twin

screw extruder. The final nanocomposite properties are almost independent of the

91
Page 91 of 302
barrel temperature over the range of typical nylon 6 processing, but they are slightly

improved by increasing the screw speed or by a second pass through the extruder.

Therefore, processing conditions need to be optimized to allow greater exfoliation of

the clay platelets and, thus, greater improvement in mechanical properties.

t
The effect of PA6 molecular weight and MMT content on nanocomposite tensile

ip
modulus is shown in Fig. 38. The addition of organoclay leads to a substantial

cr
improvement in stiffness for the composites based on each of the three PA6 samples

examined, i.e. LMW, MMW and HMW (low, medium and high molecular weight

us
respectively). Interestingly, the stiffness increases with increasing matrix molecular

an
weight at any given concentration, even though the moduli of the neat PA6s are all

quite similar. Similar trends with respect to the level of organoclay content and
M
molecular weight are evident in the yield strength results (Fig. 39). Yield strength

increases with MMT content; however, while the HMW and MMW-based
d

nanocomposites show a steady increase in strength with clay content, the LMW-based
te

nanocomposites show a less pronounced effect. These differences with respect to

molecular weight are attributed to the better exfoliation achieved for the higher
p

molecular weight matrices [14].


ce

Other factors that may play a crucial role in improvement of nanocomposite

mechanical properties include the organic modification of the clay and the addition of
Ac

compatibilizers to the polymer matrix. As a representative example, Young’s modulus

values of PP/PP-MA nanocomposites are listed in Table 20 and compared with the

corresponding microcomposite as well as simple PP-MA/PP polymer blends. It is

readily observed that increasing the amount of PP-MA increases the modulus, while

comparison of PP with the simple PP-MA/PP blends rules out any possible effect of

matrix modification due to the presence of increasing amounts of PP-MA [1].

92
Page 92 of 302
In another study, Hotta and Paul [162] performed tensile tests on various PE and

PE-MA nanocomposites based on organoclays with one or two alkyl tails. The

increase in modulus with addition of MMT is much stronger for the organoclay with

two alkyl tails than for the one with a single tail, as would be expected on the basis of

t
the much better dispersion of clay platelets for the surfactant with two alkyl tails .

ip
Similar trends were observed also for nanocompositeyield strength. Interestingly, the

cr
authors noted that there is no advantage in adding PE-MA for building modulus or

strength at low MMT content (≤ 2.5 wt %), in spite of the morphological differences

us
seen. On the contrary, there is a clear advantage in adding PE-MA at higher MMT

an
contents. Even though the benefit for modulus is not as great as might be expected, in

the absence of PE-MA, the yield strength actually decreases on addition of MMT
M
beyond 2.5 wt %.

Table 21 lists the strength and modulus values for PE-based nanocomposites, in
d

which the initial montmorillonite was modified by two intercalating agents: the
te

commonly used dioctadecyldimethyl ammonium chloride (DM) and the reactive N-γ-

trimethoxysilanepropyl octadecyldimethyl ammonium chloride (JS). Both strength


p

and modulus are higher in the case of the reactive intercalating agent, owing to to the
ce

better dispersion of the organoclay [184].

The effect of clay organic modification on nanocomposite mechanical properties


Ac

is also demonstrated in Fig. 40, which presents the ultimate strength of PU-

nanocomposites with different contents of two organically treated montmorillonites:

MO-MMT, treated with a thermally stable, aromatic amine modifier containing active

groups, and C16-MMT, treated with a quaternary alkyl ammonium salt. As can be seen

the ultimate strength increased dramatically with clay content and reached a

maximum at 5 wt % MMT, where the ultimate strength of the nanocomposites

93
Page 93 of 302
increased by about 450 % for C16-MMT and 600 % for MO-MMT, compared with

that of pure PU, indicating that the improved mechanical strength depends on the

characteristics of the modifier [202].

The extent of improvement of nanocomposite mechanical properties will also

t
depend directly upon the average length of the dispersed clay particles, since this

ip
determines their aspect ratio and, hence, their surface area [55, 203].

cr
At this point we note that several authors have also pointed out factors that have

an adverse effect on nanocomposite modulus and/or strength and need to be taken into

us
consideration when preparing nanocomposite materials.

an
Quite interestingly, Gopakumar et al. [151] found that the exfoliation of 5 wt %

and 10 wt % clay in PE-MA increased Young’s modulus by 30 % and 53 %,


M
respectively, whereas the tensile stress at yield showed only a marginal increase, up to

a maximum of 15 % for the 10 wt % clay composition. The authors noted that the
d

greatly enhanced interfacial area derived from exfoliation of the clay improves the
te

mechanical reinforcement potential of the filler. However, given that the mechanical

properties of a filled system depend on two principal factors, i.e. crystallinity of the
p

polymer matrix and the extent of filler reinforcement, the degree of crystallinity must
ce

also be considered.

In another study dealing with the effect of matrix variations on mechanical


Ac

properties of nanocomposites, Chaudhary et al. [171] studied the tensile properties of

nanocomposites based on EVAs with various VA contents and two alternative

organoclays. Since in EVA with increasing VA content the crystallinity of the

polymer decreases (and will lower the stiffness), while the polarity increases (and will

increase the intercalation), the authors suggested that in their system, the stiffness and

toughness responses would reflect an interplay of two factors: (a) an increase in the

94
Page 94 of 302
“rigid” amorphous phase due to polymer-clay intercalation and (b) an increase in the

“mobile” amorphous phase due to the increasing VA content. Experimental results

showed that the influence of increasing clay concentration on the tensile behavior of

EVA matrices was significant only with a low or moderately polar EVA matrix (9 %

t
and 18 % VA). Thus, a linear proportionality was found between clay concentration

ip
and tensile modulus for EVA-9 and EVA-18, a relation not observed with EVA-28. In

cr
fact, it is very difficult to compare the extent of the improvement of the mechanical

properties of different EVA/clay nanocomposites reported so far, because EVAs of

us
different vinyl acetate contents have been processed into the nanocomposites with

an
different clays and different modifying agents by different methods [81].

In the case of high Tg thermosets, it has been suggested that neither intercalated
M
nor exfoliated nanosilicates lead to an improvement of the tensile stress at break, but

rather make the materials more brittle. This effect appears to be generally more
d

pronounced for intercalated structures than for exfoliated ones [1].


te

The results of tensile tests conducted by Hackman and Hollaway [134] on epoxy

nanocomposites conventionally prepared under low-shear (stirring for 5 h at 90 °C)


p

are highlighted in Table 22. The tensile modulus increased by 9.0 and 19.9 % with 5
ce

and 10 wt % clay loading respectively. However, the ultimate tensile stress decreased

with increasing clay content, although the variation was large. The authors attributed
Ac

this phenomenon to the fact that large clay particles act as impurities and increase

stress concentrations. Flexural tests were also conducted and the results are outlined

in Table 23. As can be seen, the flexural modulus and ultimate flexural stress

increased by 19.6 and 7.7 % respectively for specimens containing 10 wt % clay. For

nanocomposites processed under high shear, the tensile modulus and ultimate tensile

stress increased by 18.7 and 9.3 % respectively when 5 wt % clay loading was

95
Page 95 of 302
applied. In this case, the improvement in ultimate tensile strength was attributed to the

smaller particles not generating stress concentrations leading to premature failure.

A summary of the tensile properties of soft (SPU) and hard (HPU) polyurethane

elastomers and of the corresponding nanocomposites, prepared by either solvent

t
casting or melt compounding, is provided in Table 24. As can be seen, upon silicate

ip
addition large improvements in stiffness were observed, which however were

cr
accompanied by a decrease in tensile strength and elongation [88]. Similar trends

have been reported by Tortora et al. [204]. Both exfoliated and intercalated PU/o-

us
MMT nanocomposites showed an improvement in the elastic modulus upon

an
increasing the clay content, but a decrease in the stress and strain at break .

In general, it has been argued that in the presence of polar or ionic interactions
M
between the polymer and the silicate layers, the stress at break is usually increased,

whereas when there is lack of interfacial adhesion, no or very slight tensile strength
d

enhancement is recorded [1]. Pegoretti et al. [149] found that the yield strength was
te

not reduced by the addition of clay to recycled PET and considered this to be a sign

of good interfacial adhesion; however, in the same study, a slight decrease of stress at
p

break and a dramatic reduction of strain were reported. On the other hand, in PS
ce

intercalated nanocomposites the ultimate tensile stress was found to decrease

compared to that of the PS matrix and dropped further at higher filler content. This
Ac

lack of strength was attributed to the fact that only weak interactions exist at the

PS/clay interface, contrary to other compositions in which polar interactions may

prevail, strengthening the matrix interface [205].

It should be noted that several authors have reported inability to measure

nanocomposite yield stress, because the materials often become brittle and fail before

reaching the yield point. Such remarks were made by Gorrasi et al. [6], who

96
Page 96 of 302
conducted tensile tests on PCL nanocomposite, containing 30 wt % clay, and also on

blends of this nanocomposite with HMW PCL. For the blend containing 15 wt % clay

only the elastic modulus could be evaluated since the sample did not reach the yield

point, while lower clay concentrations in the blend led to better mechanical properties

t
in terms of flexibility and drawabillity. For the initial nanocomposite, however, it was

ip
not even possible to draw the sample because of its brittleness.

cr
An interesting study was performed by Chang et al. [193] who prepared PET-

based nanocomposites through in situ intercalative polymerization, and subsequently

us
produced nano-hybrid fibers by extrusion through the die of a capillary rheometer.

an
The hot extrudates were stretched through the die of a capillary rheometer at 270 °C

and immediately drawn to various draw ratios (DR). As is evident from Table 25, the
M
tensile properties of the fibers formed increased with increasing amount of organoclay

at DR=1. When the organoclay was increased from 0 to 3 wt % in hybrids at DR=1,


d

the strength linearly improved from 46 to 71 MPa, and the modulus from 2,21 to 4.10
te

GPa. On the other hand, it is quite interesting to note the effect of DR on the tensile

strength and modulus of PET and PET nanocomposite fibers. As shown in Table 25,
p

for pure PET, the strength and modulus increased from 46 to 51 MPa and 2.21 to 2.39
ce

GPa, respectively, as the DR was increased from 1 to 16. However, the ultimate

strength and modulus of the hybrid fibers decreased markedly with increasing DR. An
Ac

increase in the mechanical tensile strength with increasing DR is very common for

engineering plastics and is usually observed in flexible coil-like polymers. However,

nanocomposite fibers did not follow this trend. Chang et al. suggested that higher

stretching of the fiber leads to debonding and creation of voids in the hybrid, which

reduce the tensile mechanical properties. This study clearly illustrates that

97
Page 97 of 302
nanocomposite materials may have a different response to mechanical loads than the

corresponding neat polymer matrices.

Finally, even though nanocomposite researchers are generally interested in the

tensile properties of the final materials, there are a few reports concerning the flexural

t
properties of PLS nanocomposites [206, 207]. Some results obtained by bending tests

ip
on nanocomposite materials are presented in Tables 26-27.

cr
6.1.3. Toughness and strain

us
The brittle behavior often exhibited by nanocomposites probably originates from

an
the formation of microvoids due to debonding of clay platelets from the polymer

matrix upon failure. This has been testified through careful inspection of fracture
M
surfaces and is also correlated to observations by in situ deformation experiments

using TEM. [147, 181]. In fact, the observation of nanocomposite fracture surfaces is
d

quite interesting. Figure 41(A) shows a typical fracture morphology in virgin nylon 12
te

and a ductile fracture as evidenced by plastic deformation. Figures 41(B) and (C)

show fracture surfaces of the nanocomposites containing 1 and 5 wt % clay,


p

respectively. No distinct clay agglomerates are observed by scanning electron


ce

microscopy (SEM) even at high magnification, as shown in Fig. 41(D). For 1 wt %

clay addition (Fig. 41(B)), the fracture surface became smoother compared with that
Ac

of neat PA12; an even more brittle feature for clay concentration of 5 wt % was

observed in Fig. 41(C). Careful inspection of the fracture surface at higher

magnification of nanocomposite with 5 wt % clay (Fig. 41(D)) verifies the formation

of microvoids due to the debonding of clay platelets from the matrix. Usually,

microvoids are formed around the large inhomogeneities, which become evident

98
Page 98 of 302
especially at high clay loadings. These microvoids will coalesce with formation of

larger cracks causing embrittlement, ultimately resulting in reduced toughness [147].

In the case of nylon 12 nanocomposites, Fig. 42 shows that the Izod impact

strength monotonically decreases as the clay concentration increases. The toughness

t
(representing the energy absorption during the fracture process) decreases by about 25

ip
% with 5 wt % of clay. Similar observations of reduction in impact strength are also

cr
reported in nylon 6/clay nanocomposites and PE-based nanocomposites, indicating

that the incorporation of clay into semicrystalline thermoplastics usually results in

us
toughness reduction, i.e. the aforementioned embrittlement effect from clay addition

an
[147].

On the other hand, some studies report little or no change of toughness upon clay
M
intercalation/exfoliation. For example, while the tensile strength and modulus of PP

nanocomposites increased rapidly with increasing clay content from 0 to 5 wt %, the


d

notched Izod impact strength was constant, within experimental error, in the clay
te

content range between 0 and 7 wt % [158]. Another study reports the impact

properties for exfoliated nylon 6-based nanocomposites prepared either by in situ


p

intercalative polymerization or by melt intercalation. In that study marginal reductions


ce

in impact properties are reported, whatever the exfoliation process used. In the case of

in situ intercalative polymerization, the Izod impact strength is reduced from 20.6 to
Ac

18.1 J/m when 4,7 wt % clay is incorporated. Charpy impact tests show similar

reduction in the impact strength, with a drop from 6.21 kJ/m2 for the filler free matrix,

down to 6.06 kJ/m2 for the 4.7 wt % nanocomposite. Figure 43 shows that the

decrease in the Izod impact strength of melt-intercalated nylon 6 nanocomposites is

not very pronounced over a relatively large range of filler content [170].

99
Page 99 of 302
Furthermore, toughness improvements upon clay dispersion have also been

reported, – a remarkable result, considering that conventional polymer-clay

composites, containing aggregated nanolayer tactoids ordinarily improve rigidity but

sacrifice toughness and elongation [7].

t
As an example, Liu and Wu [146] observed a toughening effect in PA66CN. The

ip
notched Izod impact strength increased from 96 J/m to 146 J/m upon 5 wt % clay

cr
addition, and remained higher than that of PA66, even with higher clay content.

Such results are particularily surprising, considering that from length-scale

us
arguments it is known that toughening occurs over a specific size range; effective

an
toughening necessitates a filler size greater than 0.1 µm and may not be energetically

favorable at nano-length scales. Also, the sizes of the nanoparticles are generally too
M
small to provide toughening through a crack-bridging mechanism and cannot

effectively enhance crack-trajectory tortuosity. Therefore, the extremely reduced scale


d

of a fully exfoliated nanocomposite does not lend itself to a toughening application.


te

However, in an intercalated system there is considerable interaction between silicate

layers that might alleviate this concern [45].


p

For example, Zerda and Lesser [45] showed that the gross yielding behavior of a
ce

glassy thermoset was substantially modified upon the formation of intercalated

nanocomposites, with void formation within clay aggregates leading to the evolution
Ac

of a visible shear-banded zone in compression samples. The fracture behavior appears

to be most dramatically improved in the intercalated system. The fracture energy of

the composites was increased by 100 % at clay concentration of 5 wt %. By

investigating the surface roughness and crack propagation under subcritical loading, it

has been hypothesized that the creation of additional surface area on crack

propagation is the primary means for toughening intercalated systems. The

100
Page 100 of 302
morphology of the system plays an important role in the toughening mechanism

because the spacing of regions of intercalated clay is important to toughening. It is

believed, therefore, that the intercalated morphology can afford some property

improvements that are unavailable to the fully exfoliated systems.

t
Concerning the fracture behavior of EVA-based nanocomposites, Peeterbroeck et

ip
al. [167] concluded that it is independent of the origin of the clay, while it appears to

cr
be related to the nature of the clay organo-modifier and the state of nanocomposite

dispersion. On the other hand, Kim et al. [44] attributed the enhanced toughness they

us
observed for intercalated PA12 nanocomposites to the fact that some amount of

an
applied energy is dissipated by splitting, sliding or opening of the separated bundles in

the stacked layers. Also, Le Pluart et al. [136] reported that the incorporation of a
M
benzyl dimethyl tallow alkyl ammonium montmorillonite in rubbery and glassy epoxy

matrices leads to promising improvement of mechanical properties. They obtained an


d

interesting stiffness/ toughness balance for very low filler contents and without
te

reducing the Tg of the matrix, which is particularly interesting considering how the

brittleness of epoxies limits their use in technological areas where their high Tg is
p

often highly appreciated.


ce

Quite interestingly, Fornes et al. [14] investigated how the matrix molecular

weight as well as the extension rate during tensile tests affect the ductility of PA6-
Ac

based nanocomposites. Figure 44 presents the relationship between MMT content and

elongation at break for PA6 matrices of different molecular weights, for two different

rates of extension. As shown in Fig. 44a, the virgin polyamides are very ductile at a

test rate of 0.51 cm/min. With increasing clay content, the ductility gradually

decreases, however, the HMW and MMW based composites attain reasonable levels

of ductility at MMT concentrations as high as 3.5 wt %, while the elongation at break

101
Page 101 of 302
for the LMW based nanocomposites decreases rapidly at low MMT content (around 1

wt %). Even though the opposite result could have been anticipated, considering that

high molecular weight matrices favor clay exfoliaton, the authors attribute the larger

reduction of elongation at break in the LMW-based systems to the presence of stacked

t
silicate layers, as seen in TEM photographs. On the other hand, the higher testing rate

ip
of 5.1 cm/min yields similar trends, as shown in Fig. 44b, but the absolute level of

cr
elongation at break is significantly lower. Interestingly, the strain at break for LMW

composites is relatively independent of the rate of extension, similar to what has been

us
observed in glass fiber reinforced composites. Even at the highest clay content, the

an
HMW composite exhibits ductile fracture, whereas the LMW and MMW based

nanocomposites fracture in a brittle manner at the highest clay content.


M
As in the case of toughness, contradictory results have also been presented

concerning the effect of nanocomposite formation on the elongation at break, as can


d

be seen in Table 28. Even though in most cases this property deteriorates when a
te

layered silicate is dispersed into a polymer matrix, nanocomposites have been

reported exhibiting similar or even higher elongations at break than the neat matrices.
p

For example, in the case of EVA-12/MMT nanocomposites, a significant increase


ce

of both the strength and the elongation has been reported with the introduction of the

organoclay into the EVA-12 matrix. However, this enhancement is a maximum when
Ac

the clay concentration is only 2 wt %. Further increase of clay content causes

reduction in mechanical properties, probably due to aggregation of clay layers, as

already discussed [209]. Thellen et al. [152] conducted tensile tests on PLA–based

nanocomposite blown films and recorded improvements up to 40 % for both the

modulus and the elongation. Yao et al. [142] also reported improvements in strain at

break. Data of tensile strength and strain-at-break vs. clay content are shown in Table

102
Page 102 of 302
29. The authors suggested that the improved elasticity is due in part to the plasticizing

effect of gallery onium ions, which contribute to dangling chain formation in the

matrix. Accordingly, Chen and Evans [210] observed a dramatic improvement in

tensile elongation at break in the presence of clay. At low clay loadings, test pieces

t
underwent yielding during tension, similar to pristine PCL, but with dramatic

ip
increases in ductility, quite the opposite of the usual effect of adding a particulate

cr
filler to a polymer. When the clay loading was high, typically higher than 20 wt % ,

the composites became brittle and did not reach the yield point.

us
Finally, it is worth summarizing the work of Hong et al. [185] on PP-based

an
RTPO/clay nanocomposites, prepared by using PP-MA as a compatibilizer. PP-based

RTPO (or in reactor made TPO) is a blend of PP and poly(ethylene-co-propylene)


M
(EPR), produced by the bulk polymerization of propylene, followed by gas-phase

copolymerization of ethylene and propylene driven by the TiCl4/MgCl2 based catalyst


d

system. Such materials, like the conventional blends of PP/EPR prepared by


te

mechanical blending, exhibit improved flexibility and toughness compared to neat PP.

Moreover, because the rubber phase can be dispersed uniformly and reach a high
p

degree of dispersion in these in situ blends, it is possible to achieve more intimate


ce

interaction between the matrix and the rubber phase. The compositions and tensile

properties of polypropylene-based RTPO/clay nanocomposites are reported in Table


Ac

30. As can be seen, the tensile moduli of the nanocomposites became higher as the

clay content increases. On the other hand, the elongation at break decreases as the

clay content increases, but the value of nanocomposites containing 10 wt % clay is

437 %, which is much higher than that of PP/clay nanocomposites reported elsewhere.

As the authors claim, these elongational properties of PP based RTPO/clay

nanocomposites are unique and promising for many applications. In fact, for reasons

103
Page 103 of 302
of comparison, Hong et al. also prepared and tested nanocomposites using PP/EPR

mechanical blend matrix, modified with PP-MA. For these materials, the elongation

at break values were about 50 %, which are much lower than those of RTPO clay

nanocomposites and is not suitable for industrial application. The authors attributed

t
this discrepancy to the difference of dispersion homogeneity and domain size of

ip
ethylene copolymer between RTPO and PP/EPR mechanical blends.

cr
6.1.4. Comparison and Synergistic Effects of Clays and Conventional Reinforcements

us
Typically, layered silicates are incorporated in polymeric materials as the sole

an
reinforcing element. However, several studies have investigated the potential

synergistic effects of clays and conventional reinforcements, such as glass fibers.


M
In this context, Wu et al. [52] studied the effect of adding glass fibers to PA6 and

PA6-based nanocomposites containing 3 wt % montmorillonite. They found that the


d

tensile strength of PA6/clay containing 30 wt % glass fibers is 11 % higher than that


te

of PA6 containing 30 wt % glass fibers, while the tensile modulus of the

nanocomposites increases by 42 %. Bending strength and bending modulus of neat


p

PA6/clay are similar to PA6 reinforced with 20 % glass fibers. However, the notched
ce

Izod impact strength of the nanocomposite is lower than that of neat polyamide 6, and

is further decreased with the addition of fibers.


Ac

In another study, typical stress-strain diagrams for PA6 and composites containing

5 wt % of fillers are compared, as shown in Fig. 45 (at 5.08 cm/min) and Fig. 46 (at

0.5 cm/min). A summary of the mechanical properties of these materials is shown in

Table 31. As can be seen from the table , regardless of the type of filler, the strength

and modulus are substantially increased relative to the neat PA6, without significant

variation in toughness or impact strength, as measured by the standard Izod test.

104
Page 104 of 302
Furthermore, nanocomposites show superior mechanical properties, especially

modulus, as compared with conventional PA6 composites formed by compounding

with glass fibers or untreated clay. The elongation at break for the nanocomposites is

more or less the same as that of the neat PA6, whereas, values for the conventional

t
composites are dramatically decreased. Also, the elongation at break for the

ip
nanocomposites is greatly affected by the crosshead speed, as is the case for neat PA6.

cr
On the other hand, rate of extension has little effect on the elongation of glass fiber

composites. It is noteworthy that the composite of PA6 with untreated Na+-MMT

us
shows higher strength and modulus than neat PA6, which is quite contrary to the

an
results from other investigators, who claim that untreated clay composites with PA6

are inferior to neat polymer in terms of some mechanical properties. Interestingly, a


M
synergistic effect on the tensile strength and modulus is again observed when the

exfoliated nanocomposite is used as the matrix for a glass fiber reinforced composite.
d

As shown in Table 31, the modulus of the nanocomposite with 5 wt % loading of the
te

organoclay is improved about 38 % relative to neat PA6 and the glass fiber composite

shows a 22 % improvement. When glass fibers are added to the nanocomposite, the
p

modulus is 81 % higher than that of PA6. This exceeds what is expected on the basis
ce

of simple additivity. Stiffness and strength are dramatically improved as the amount

of organoclay increases. On the other hand, the impact strength and elongation at
Ac

break remain at the levels of neat PA6 up to about 5 wt % of the organoclay, and

decrease thereafter [15].

6.2. Dynamic Mechanical PropertiesDynamic mechanical analysis (DMA) measures

the response of a material to a cyclic deformation (usually tension or three-point

bending type deformation) as a function of the temperature. DMA results are

105
Page 105 of 302
expressed by three main parameters: (i) the storage modulus (E΄ or G΄), corresponding

to the elastic response to the deformation; (ii) the loss modulus (E΄΄ or G΄΄),

corresponding to the plastic response to the deformation and (iii) tan δ, that is, the E΄/

E΄΄ (or G΄/ G΄΄) ratio, useful for determining the occurrence of molecular mobility

t
transitions such as the glass transition temperature [1].

ip
Indicatively, the temperature dependence of G΄, G΄΄ and tan δ of a nylon 6 matrix

cr
and various nanocomposites is presented in Fig. 47. In the case of nanocomposites,

us
the main conclusion derived from dynamic mechanical studies is that the storage

modulus increases upon dispersion of a layered silicate in a polymer. This increase is

an
generally larger above the glass transition temperature, and for exfoliated PLS

nanocomposite structures is probably due to the creation of a three-dimensional


M
network of interconnected long silicate layers, strengthening the material through

mechanical percolation [1]. Above the glass transition temperature, when materials
d

become soft, the reinforcement effect of the clay particles becomes more prominent,
te

due to the restricted movement of the polymer chains. This results in the observed

enhancement of G΄ [55]. For example, an epoxy-based nanocomposite, containing 4


p

vol. % silicates, showed a 60 % increase in G΄ in the glassy region, compared to the


ce

unfilled epoxy, while the equivalent increase in the rubbery region was 450 % [135].

Similar results have also been reported in the case of PP- [189], PCL- [80], SBS-
Ac

[211], PA- [64, 212], PLA- [83, 153, 208, 213], and epoxy-based nanocomposites

[135]

Enhancement of the loss modulus, G΄΄, has also been reported for nanocomposite

materials, however this aspect of dynamic mechanical performance is far less

discussed in the literature.

106
Page 106 of 302
Finally, the tan δ values are affected in different ways by nanocomposite

formation, depending on the polymer matrix. For example, in PS based

nanocomposites, a shift of tan δ to higher temperatures has been observed,

accompanied by a broadening of this transition [205], while the opposite effect was

t
reported in the case of PP-based nanocomposites [189]. Some authors observed a

ip
decrease of tan δ peaks, and considered this indicative of a glass transition

cr
suppression by the presence of the clay. However, Fornes and Paul [64] pointed out

that this conclusion is a misinterpretation, since the low values for the

us
nanocomposites are simply the result of dividing the relatively constant loss modulus

an
values in the Tg region, by larger and larger storage modulus values.

Quite surprisingly, DMA showed that above Tg, the moduli for the pure PU and
M
the PU/o-MMT nanocomposites show no obvious difference, while below Tg, addition

of o-MMT strongly influences the modulus values. Interestingly, the authors found
d

that E′ and E′′ of the PU/o-MMT decrease in comparison with values for the PU, for
te

unclear reasons. On the other hand, significant enhancements of E′ and E′′ were seen

for the the nanocomposite prepared using a particular modified clay [202]. In the case
p

of PLA-based nanocomposites, it was observed that PLACNs with a very small


ce

amount of o-PCL as a compatibilizer exhibited a very large enhancement of


Ac

mechanical properties compared to that of PLACN with comparable clay loading

[153]. Krikorian and Pochan [83] also studied the dynamic mechanical properties of

neat PLA and nanocomposites prepared with OMLS. These authors found that at high

temperatures the reinforcement effect of OMLS weakens, and suggested that this

indicates a weakening of the thermomechanical stability of these materials at high

temperature.

107
Page 107 of 302
6.3. Barrier Properties

Generally, polymer/layered silicate nanocomposites are characterized by very

strong enhancements of their barrier properties. Polymers ranging from epoxies and

t
ip
good sealants (like siloxanes) to semi-permeable (e.g. polyureas) and highly

hydrophilic (e.g. PVA) are all improved up to an order of magnitude by low clay

cr
loadings [31].

us
The dramatic improvement of barrier properties can be explained by the concept

of tortuous paths. That is, when impermeable nanoparticles are incorporated into a

an
polymer, the permeating molecules are forced to wiggle around them in a random

walk, and hence diffuse by a tortuous pathway, as shown in Fig. 48 [4, 7, 55, 206,
M
214-216].

The tortuosity factor is defined as the ratio of the actual distance, d΄, that the
d

penetrant must travel to the shortest distance d that it would travel in the absence of
te

barriers. It is expressed in terms of the length L, the width W and the volume fraction
p

of the sheets φs as:


ce

d' L
τ= = 1+ φs
d 2W
Ac

It becomes obvious from this expression that a sheet-like morphology is particularly

efficient at maximizing the path length, due to the large length-to-width ratio, as

compared to other filler shapes [1, 50, 55].

According to the model proposed by Nielsen, the effect of tortuosity on the

permeability may, in turn, be expressed as:

PPCN 1 - φs
=
Pp τ

108
Page 108 of 302
where PPCN and PP represent the permeability of the nanocomposite and the pure

polymer, respectively and φs is the clay content [50, 55, 217].

Although the above equations were developed to model the diffusion of small

t
molecules in conventional composites, they have also been used in reproducing

ip
experimental results for the relative permeability in PLS nanocomposites.

cr
Discrepancies between the experimental data and the theoretical line may be

attributed either to inadequacies of the model or to incomplete orientation of the

us
particles within the nanocomposite film plane [50, 162]. In fact, the key assumption of

an
the Nielsen model is that the sheets are placed in an arrangement such that the

direction of diffusion is normal to the direction of the sheets. Clearly, this


M
arrangement results in the highest tortuosity, and any deviation from it would, in fact,

lead to deterioration of the barrier properties [50, 55].


d

Moreover, the tortuous path theory, including the Nielsen equation as well as
te

other phenomenological relations (e.g. the Cussler [218] formula, the Barrel [219]

formula and the power law equation [220]), is grounded on the assumption that the
p

presence of nanoparticles does not affect the diffusivity of the polymer matrix.
ce

However, experimental observations demonstrate that molecular mobility in a

polymer matrix, which is intimately connected to the mass transport properties,


Ac

diminished by clay incorporation. This reduction should be accompanied by a

decrease in diffusivity of small molecules, which is not considered in the concept of

tortuous paths.

Messersmith and Giannelis [118] studied the permeability of liquids and gases in

nanocomposites and they observed that water permeability in PCL nanocomposites is

dramatically reduced compared to the unfilled polymer. They also noted how the

109
Page 109 of 302
decrease in permeability is much more pronounced in the nanocomposites compared

to conventionally filled polymers with much higher filler content.

Liu and Wu [146], recorded the water absorption curves of PA66 and

corresponding nanocomposites. They found that with increasing clay content, the

t
water absorption at saturation decreases rapidly from 7.6 % for PA66 to 5.2 % for the

ip
nanocomposite containing 5 wt % clay. They attributed this reduction to the presence

cr
of immobilized polymer in the amorphous phase. However, above 5 wt % clay

content, the decrease in the saturation amount of water is not so obvious, probably

us
because of aggregation of silicate layers. Also, the diffusion coefficient values

an
decrease greatly with increasing clay loading, but after 5 wt % clay content, the

amplitude of the decrease is obviously slower. In addition to the immobilized phase


M
explanation, and the increased average diffusion path length, since an epoxy-co-

intercalated clay was used in their study, the authors assumed that the epoxy groups
d

between silicate layers have a strong interaction with amino and amide groups of the
te

PA66 matrix, to some extent preventing them from forming hydrogen bonds with

water.
p

Significant reductions in diffusivity and maximum water uptake were also


ce

reported by Liu et al. [130] in epoxy-based nanocomposites. Here, however, the

decreased maximum water uptake was attributed to the low maximum water uptake of
Ac

the nanoclays (ca. 2.8 wt %) compared to the epoxy resin system (ca. 7.5 wt %).

Drozdov et al. [221] conducted moisture diffusion tests on vinylester resin-MMT

clay nanocomposites and demonstrated that the clay content affects in a similar way

the diffusion coefficient and the constants expressing the elastoplastic behavior,

indicating that moisture diffusion and elastoplasticity may reflect the same

phenomena at the microlevel, associated with molecular mobility of the polymeric

110
Page 110 of 302
matrix. Moreover, their experimental data demonstrated that moisture diffusion in the

neat polymer resin is nearly Fickian, but is transformed to an anomalous mode of

transport of the penetrant molecules with an increase in clay concentration. The

authors explained the anomalous moisture uptake by immobilization of water

t
molecules on the surfaces of the hydrophilic MMT clay layers. In fact, they pointed

ip
out that, after a nanocomposite plate is immersed in water, three processes occur in

cr
the nanocomposite: (1) sorption of water molecules on the sample surfaces, (2)

diffusion of water into the plate, and (3) adsorption of water molecules on the

us
hydrophilic surfaces of clay layers, where these molecules become immobilized.

an
Many studies reported in the literature have focused on nanocomposite barrier

properties against gases and vapors. As an example, Tortora et al. [204] measured the
M
transport properties of PU/o-MMT nanocomposites (prepared using a PCL

nanocomposite “master-batch”) using water vapor as hydrophilic permeant and


d

dichloromethane as hydrophobic one. For both vapors, the sorption behavior changed
te

in the presence of the clay, as can be seen for example in Fig. 49, where the

equilibrium concentration, Ceq (g/100 g), of water vapor is represented as a function


p

of the vapor activity for all nanocomposites and for the o-MMT. The sorption curve
ce

of water vapor for o-MMT follows the Langmuir sorption isotherm, in which the

sorption of solvent molecules occurs at specific sites; therefore, when all the sites are
Ac

saturated, a plateau is reached. On the other hand, the sorption of neat PU shows a

linear dependence of equilibrium concentration on activity, while nanocomposites

show a dual sorption shape, that is a downward concavity, an inflection point and an

upward curvature. The prevailing mechanism in the first zone is the sorption of

solvent molecules on specific sites, due to interacting groups. Tortora et al. inferred

that this type of sorption is due to the presence of clay in the polymers. At higher

111
Page 111 of 302
activities, the plasticization of the polymeric matrix determines a more than linear

increase of vapor concentration and a transition in the curve is observed, from a dual

type to a Flory-Huggins behavior. From the calculated values of the sorption

parameters, defined as: S=d(Ceq)/dp, and the zero-concentration diffusion coefficients

t
for water sorption and dichloromethane vapor, the authors concluded that the sorption

ip
did not drastically change on increasing the clay content, whereas the zero-

cr
concentration diffusion coefficient D0 strongly decreased with increasing inorganic

content. The permeability calculated as the product SD0, was largely dominated by the

us
diffusion parameter; it showed a remarkable decrease up to 20 wt % of clay and a

an
levelling off at higher contents.

Ke and Yongping [222] tested the O2 permeability of intercalated PET


M
nanocomposites. As demonstrated in Fig. 50, a small amount of clay effectively

reduced the permeability of the PET film. When the content of o-MMT reached 3 wt
d

% the permeation of O2 was reduced to half that of the pure PET film. Further
te

examples of barrier property improvement for PET nanocomposites designated for

packaging applications are given in Table 32 [216].


p

Ogasawara et al. [223] reported on improved helium gas barrier properties of


ce

epoxy/MMT nanocomposites, compared to the pure resin. The estimated diffusivity,

D, solubility, S, and permeability P are shown in Figs. 51-53, as functions of


Ac

montmorillonite weight fraction. Dispersing MMT particles in the epoxy decreased

the diffusion coefficient D. For example, the diffusion coefficient of the

nanocomposite with 6 wt % clay was approximately one-tenth that of the base

polymer. On the other hand, the solubility increased with montmorillonite dispersion

and permeability remained almost constant due to the balance of diffusivity and

solubility.

112
Page 112 of 302
On the other hand, Sinha Ray et al. [208] found that the O2 gas permeability of

PLA nanocomposites with 4, 5 and 7 wt % clay, was reduced to 88, 85 and 81 %

compared to that of neat PLA.

Similarly, Chang et al. [84] reported the oxygen gas permeability of PLA

t
nanocomposites prepared with three different kinds of OMLS using a melt

ip
intercalation technique. Table 33 summarizes the results for O2 gas permeability. The

cr
results show that O2 gas permeability of nanocomposites systematically decreased

with increasing clay content; and when the clay loading was as much as 10 wt %, the

us
permeability for nanocomposite decreased to half the PLA permeability, regardless of

an
the nature of the OMLS employed for the nanocomposite preparation.

Finally, Figure 54 presents water vapor and oxygen permeation results for PLA-
M
based nanocomposite films. In all cases the nanocomposite films were better oxygen

barriers than the pure PLA films (Fig. 54a), exhibiting a 15-48 % reduction in oxygen
d

permeation rate. It is worth noticing that the oxygen permeation rate of the
te

nanocomposite films was essentially independent of screw speed and feed rate,

whereas the oxygen barrier properties of the neat PLA homopolymer were quite
p

sensitive to processing, suggesting that the need to optimize processing parameters is


ce

more critical when working with the PLA homopolymer than with the PLA/OMLS

nanocomposites. The PLA/OMLS nanocomposites also exhibited much improved


Ac

barrier properties to water vapor relative to the neat films (Fig. 54b). In general, clay

incorporation decreased the permeation rate of the resulting films to water vapor by

about 40-50 %, again independently of processing [152].

Summarizing: although a decrease of diffusivity is a well-established result of

nanocomposite formation, contradictory results are reported concerning the saturation

uptake values of various solvents or gases. Increases of the saturation uptake level are

113
Page 113 of 302
usually attributed to clustering phenomena. It is worth noticing, however, that in

nanocomposites the coexistence of phases with different permeabilities can cause

complex transport phenomena. On the one hand, the organophilic clay gives rise to

superficial adsorption and to specific interactions with the solvents. In turn, the

t
polymer phase can be considered, in most cases, as a two-phase, crystalline-

ip
amorphous system, the crystalline regions being generally impermeable to penetrant

cr
molecules. The presence of the silicate layers may be expected to cause a decrease in

permeability, due to the more tortuous path for the diffusing molecules that must

us
bypass impenetrable platelets [133]. Simultaneously, the influence of changes in

an
matrix crystallinity and chain mobility, induced by the presence of the filler, should

always be taken into consideration [10].


M
6.4. Thermal Stability
d
te

The thermal stability of polymeric materials is usually studied by

thermogravimetric analysis (TGA). The weight loss due to the formation of volatile
p

products after degradation at high temperature is monitored as a function of


ce

temperature (and/or time). When heating occurs under an inert gas flow, a non-

oxidative degradation occurs, while the use of air or oxygen allows oxidative
Ac

degradation of the samples [50, 55].

Generally, the incorporation of clay into the polymer matrix was found to enhance

thermal stability by acting as a superior insulator and mass transport barrier to the

volatile products generated during decomposition, as well as by assisting in the

formation of char after thermal decomposition [50, 55, 133, 224].

114
Page 114 of 302
Vyazovkin et al. [225] compared the thermal degradation of a PS nanocomposite

with that of the virgin polymer under nitrogen and air. As seen in Fig. 55 in both

nitrogen and air the decomposition temperature of nanocomposites increased by 30-40

°C. The authors also observed that the virgin polymer degrades without forming any

t
residue, whereas the nanocomposite (as expected) leaves some residue.

ip
Zanetti et al. [226] reported TGA curves of a nanocomposite PE/EVA/o-MMT

cr
and the corresponding matrix PE/EVA. Under nitrogen, these samples do not show

great differences of stability. However, in air, the PE/EVA blend is subject to a

us
marked weight loss above 350 °C, to form a 5 wt % residue at 450 °C, which is

an
completely oxidized to volatile products between 470 and 550 °C. The

nanocomposite, on the other hand, displays a different pattern. The presence of 5 wt


M
% o-MMT is enough to change the polymer’s thermo-oxidative behavior and between

350 - 480 °C the amount of residue is higher to that observed in a nitrogen flow.
d

According to the authors, the organoclay shields the polymer from the action of
te

oxygen, dramatically increasing the thermal stability under oxidative conditions.

Bandyopadhyay et al. [227] reported the first improved thermal stability of


p

biodegradable nanocomposites that combined PLA and organically modified


ce

fluorohectorite or montmorillonite. They showed that the PLA intercalated between

the galleries of FH or MMT clay resisted the thermal degradation under conditions
Ac

that would otherwise completely degrade pure PLA. This conclusion has been verified

by a number of researchers in subsequent studies. Thellen et al. [152] presented TGA

curves for the neat polymer and corresponding nanocomposites (Fig. 56) and reported

that the onset of thermal degradation was approximately 9 °C higher for the

nanocomposite than for the neat PLA.

115
Page 115 of 302
The thermal stability of PCL-based nanocomposites has also been studied by

TGA. Generally, the degradation of PCL fits a two-step mechanism. First, random

chain scission through pyrolysis of the ester groups, with the release of CO2, H2O and

hexanoic acid, and in the second step, ε-caprolactone (cyclic monomer) formation as a

t
result of an unzipping depolymerization process. It has been reported that the thermal

ip
stability of PCL/o-MMT nanocomposites systematically increases with increasing

cr
clay, up to a loading of 5 wt % [1, 228].

On the other hand, contradictory results are found in the literature concerning the

us
thermal degradation of PA6-based nanocomposites. Pramoda et al. observed that the

an
degradation onset temperature is 12 °C higher for PA6 with 2.5 % clay loading than

that of virgin PA6 and that the onset temperature for the higher clay loading remained
M
unchanged. Also, Dabrowski et al. [229] showed that protective barriers are formed

during thermal degradation of polyamide 6/clay nanocomposite , which slow down


d

the rate of degradation via a diffusion process (hindering the escape of of volatiles).
te

However, TGA experiments by other workers did not show significant changes in the

onset of degradation. For example, according to Jang et al. [230] the mass loss
p

behavior of PA6/clay nanocomposites is not significantly different from that of virgin


ce

PA6. Irrespective of formulation, the temperatures at 50 % mass loss were 471-476

°C, which was within the error range of the TGA instrument used. Moreover, other
Ac

researchers found that PA6 nanocomposites have somewhat lower stability than neat

nylon 6, and attributed such observations to the degradation effect of the quaternary

alkylammonium treatment on the montmorillonite [15]. For example, Davis et al.

[192] studied the thermal stability of PA6 and PA6 nanocomposites, injection molded
13
at 300 °C, by C NMR. They found that PA6 does not degrade at processing

temperature, whereas there is significant decrease in molecular weight in

116
Page 116 of 302
nanocomposites under the same conditions. The authors noted that the degradation

might depend upon water in the nanocomposites, which may cause hydrolytic

cleavage. The thermal degradation mechanism of PA6, proposed by Levchik et al., is

shown in Fig. 57 [231].

t
In fact, despite the general improvement of thermal stability, decreases in the

ip
thermal stability of polymers upon nanocomposite formation have also been reported,

cr
and various mechanisms have been put forward to explain the results. It has been

argued, for example, that after the early stages of thermal decomposition the stacked

us
silicate layers could hold accumulated heat, acting as a heat source to accelerate the

an
decomposition process, in conjunction with the heat flow supplied by the outside heat

source [55]. Also, the alkylammonium cations in the organoclay could suffer
M
decomposition following the Hofmann elimination reaction, and the product could

catalyze the degradation of polymer matrices. Moreover, the clay itself can also
d

catalyze the degradation of polymer matrices. Thus, it becomes obvious that the
te

organoclay may have two opposing functions in thermal stability of nanocomposites:

a barrier effect, which should improve the thermal stability and a catalytic effect on
p

the degradation of the polymer matrix, which should decrease the thermal stability
ce

[184].

For example, Zhao et al. [184] investigated the thermal stability of PE-based
Ac

nanocomposites in a nitrogen atmosphere and recorded the TGA and DTGA

(derivative curve) profiles presented in Fig. 58. As can be seen from this figure, at the

initial stage of degradation (before 400 °C), the nanocomposites degrade faster than

the pure matrix; this was attributed to the Hofmann elimination reaction and the clay

catalyzed degradation. On the other hand, above 400 °C nanocomposites appear to be

more stable than pure PE. The onset temperatures for nanocomposites are all higher

117
Page 117 of 302
than that of pure PE, but decrease with increasing clay loading. Thus, the authors

suggested that when a low clay fraction is added to the polymer, the clay disperses

well and the barrier effect is predominant, but with increasing loading, the catalyzing

effect rapidly increases and becomes dominant, so that the thermal stability of the

t
nanocomposite decreases.

ip
Other researchers have studied the effect of clay concentration on the thermal

cr
stability of EVA-based nanocomposites. It has been found, for example, that the

thermal stability of EVA-12 (12 wt % VA) increases with the introduction of o-MMT.

us
With an increase in the o-MMT loading over 2 wt %, however, the hybrids show a

an
decreasing trend in their initial thermal decomposition temperature. These findings

were attributed to the fact that, at low filler contents homogenous exfoliation and
M
random dispersion of clay is achieved on a nanometer level, whereas the higher filler

loading destabilizes the matrix, because of the aggregation of silicate layers [232]. In
d

fact, EVA/clay nanocomposites containing more vinyl acetate maintain this thermal
te

stability improvement up to 6-8 wt % filler loading, probably because of the degree of

dispersion of clay in the polymer matrix [209].


p

Similarly, Paul et al. [233] observed an increase in thermal stability of PLA


ce

nanocomposites with increasing clay content, with a maximum at a clay loading of 5

wt %. With further increase of filler content, a decrease in thermal stability was


Ac

observed – an effect explained by the relative extent of exfoliation as a function of the

amount of OMLS.

Phang et al. [147] found that the thermal stability of PA12 is significantly

enhanced as the clay concentration increases; for instance, by about 20 °C with

incorporation of only 5 wt % clay into the matrix. For PA12/organoclay

nanocomposites with lower clay contents (<2 wt %) only limited improvement in

118
Page 118 of 302
thermal stability was observed. This was attributed to the fact that the exfoliated

structure formed at low clay content is probably not enough to trigger the thermal and

gas barrier properties effectively in the matrix, because of the short tortuous path

formed. This result is consistent with reports claiming that optimum thermal stability

t
is usually achieved for clay loadings between 2.5 and 5 wt %. For lower clay

ip
fractions, exfoliated morphology dominates, but this low clay content does not enable

cr
the barrier effect. At much higher clay content, on the other hand, particle

agglomeration occurs. Therefore, intercalated and exfoliated structures usually

us
coexist, which again does not allow maximization of the tortuous path.

an
As shown in Table 34, PET nanocomposites exhibit improved thermal stability,

with the initial thermal degradation temperature TDi increasing with the amount of
M
organoclay. A maximum increase of 16 °C was recorded in the case of 3 wt % clay.

Also, the weight of residue at 600 °C increased, ranging from 1 to 21 %, with clay
d

loading from 0 to 3 wt %[193].


te

An optimum clay loading for thermal stability enhancement was also reported for

PS-based nanocomposites. When dimethylbenzyloctadecyl-ammonium cation was


p

used for MMT modification, the threshold was reached at a surprisingly low MMT
ce

content of only 0.3 wt % [35].

Berta et al. [144] did TGA under nitrogen and air for PU and the corresponding
Ac

nanocomposite. Under nitrogen, the nanocomposite shows the same TGA and DTGA

profiles as PU, but displaced by 10 °C, a change attributed to the barrier effect. On the

contrary, when tested under air, the nanocomposites showed very different behavior

than pure PU. For pure PU, most of the polymer volatilizes in the first step of PU

decomposition and the second step is diminished in the TGA/DTGA curves in air

compared to those obtained in nitrogen; thus only a relatively small shoulder appears

119
Page 119 of 302
at 350 °C in air. However, in the nanocomposites the lower temperature degradation

is significantly suppressed and the second DTGA peak around 395 °C is clearly

present. This behavior was again attributed to greatly retarded thermal oxidation due

to the shielding of the material from oxygen.

t
The thermal properties of EVA-based nanocomposites have been widely studied,

ip
mainly by means of TGA. It has been well established by different research groups

cr
that EVAs exhibit two-step decomposition. The first step, which is identical in both

oxidative and non-oxidative conditions, occurs from 350-400 °C and corresponds to

us
the deacetylation reaction, with production of gaseous acetic acid and formation of

an
carbon-carbon double bonds along the polymer backbone. The second step, between

400 and 500 °C, involves thermal decomposition of the unsaturated backbone, either
M
by further radical scission (non-oxidative decomposition) or by thermal combustion

(oxidative decomposition) [35, 81, 167].


d

Riva et al. [234] observed accelerated acetic acid loss of EVA-clay


te

nanocomposites and speculated that this process can be catalyzed by acidic sites of

the nano-dispersed clay. For the second step of EVA degradation, Maurin et al. [235]
p

found as products 1-butene, carbon dioxide, ethylene, methane and carbon monoxide,
ce

while McGrattan [236] identified hydrocarbons ranging from C8 to C26, grouped in a

series of α,ω-dienes, 1-alkenes and n-alkenes. By thoroughly studying both the first
Ac

and second steps, Costache et al. [237] investigated the possibility that the presence of

nano-dispersed clay can change the degradation pathway of EVA. In the early stages

of the EVA degradation, the loss of acetic acid seems to be catalyzed by the hydroxyl

groups present on the edges of the clay lamellae. The TGA-FTIR and GC-MS results

of thermal degradation of EVA in the presence and in the absence of clay show that

even though the two processes are very similar, subtle changes occur, leading to the

120
Page 120 of 302
formation of products that differ both in quantity and identity. It has been suggested

that these products form as a result of radical recombination reactions that occur

because the degrading polymer is contained within the clay layers for sufficient time

to permit the reactions (Fig. 59). In cases where there are multiple degradation

t
pathways, the presence of the clay can promote one of these at the expense of another,

ip
and thus lead to different products and hence a different rate of volatilization.

cr
In another study, the thermal decomposition of EVA nanocomposites has been

investigated using TGA in helium and in air. In helium, EVA nanocomposites exhibit

us
a negligible reduction in thermal stability compared to virgin EVA or EVA

an
microcomposite. In contrast, when decomposed under air, the same nanocomposites

exhibit a rather large increase in thermal stability, as the maximum of the second
M
degradation peak is shifted 40 °C to higher temperature, while the maximum of the

first decomposition peak remains unchanged (see Table 35). The explanation for the
d

improved thermal stability is the formation of protective char under oxidative


te

conditions [238].

Finally, several studies have focused on the effect of clay organic modification on
p

thermal stability of nanocomposites. More specifically, it has been suggested that the
ce

thermal stability of the nanocomposites is directly related to the stability of the OMLS

used for their preparation [84].


Ac

In a typical example, Xiong et al. [202] studied the thermal stability of PU and

nanocomposites prepared using MMT modified either by a quaternary alkyl

ammonium salt (C16-MMT) or by an aromatic modifier (MO-MMT). The TGA curves

obtained (Fig. 60) show that the degradation rates of the nanocomposites were slightly

slower compared to that of PU, indicating an improvement of thermal stability.

Comparing the quaternary alkyl ammonium salt and the aromatic amine modifier they

121
Page 121 of 302
used, the authors found the degradation onset temperature of PU/C16-MMT at 316.3

°C, thus lower than that of the PU/MO-MMT (331.6 °C), showing the latter to have

the higher thermal stability. The main reason is that the aromatic chain of the

modifier in MO-MMT has higher thermal stability than the alkyl chain of the modifier

t
in C16-MMT. The authors also noted that the aromatic amine modifier can react with

ip
the pre-polyurethane matrix used, further strengthening the interaction between

cr
inorganic and organic phases.

In another interesting study, Yei et al. [115] used TGA to characterize PS-based

us
nanocomposites prepared from clay treated with cetylpyridinium chloride (CPC) and

an
CPC/α-cyclodextrin (CPC/α-CD) inclusion complex, as already described in Section

5.3.1. Figure 61 shows the TGA curves of pure CPC and the CPC/α-CD inclusion
M
complex. As is clearly evident, the CPC/α-CD inclusion complex decomposes at

higher temperature (284 °C) than the pure CPC (220 °C). Thus, the formation of an
d

inclusion complex between CPD and α-CD improved the thermal stability of the CPC
te

surfactant; the presence of the α-CD protects CPC from earlier decomposition. Table

36 summarizes the TGA results for the nanocomposites and the pure polymer. Both
p

nanocomposites display higher decomposition temperatures than the virgin PS, with
ce

the CPC/α-CD intercalated clay nanocomposite being the most thermally stable of the

three samples.
Ac

As deduced from the previous examples, even though contradictory results are

sometimes found in the literature concerning the thermal stability of polymeric

nanocomposites, the opportunity of achieving a significant improvement in thermal

stability through low filler content is particularly attractive because end-products can

be made cheaper, lighter and easier to process [35].

122
Page 122 of 302
6.5. Flame retardance

6.5.1. Flame retardance of polymer-layered silicate nanocomposites

Polymers are being used in more and more applications where flame retardant

t
behavior is of critical importance. Traditionally, flame retardancy has been achieved

ip
either by using intrinsically flame retardant polymers, such as fluoropolymers or

cr
PVC, or by incorporating flame retardants (FRs), such as aluminium trihydrate,

magnesium hydroxide, organic brominated compounds or intumescent systems.

us
However, such FRs exhibit significant disadvantages. For example, aluminium

an
trihydrate and magnesium hydroxide need to be applied at very high loadings to be

effective, resulting in high density and lack of flexibility of the end products, as well
M
as low mechanical properties and problems in compounding. On the other hand,

concerns over the environmental impact have made halogen containing materials a
d

less popular option in many countries. Moreover, the addition of many FRs increases
te

the production of soot and carbon monoxide during combustion. Finally, intumescent

systems are relatively expensive and electrical requirements can restrict their
p

application [35, 56].


ce

Among test methods developed for the evaluation of flame retardant properties,

the one most commonly used by researchers is cone calorimetry (Fig. 62), as it
Ac

provides valuable information and may even indicate the flame retardancy

mechanism. The measuring principle in this test, which is is standardized as ASTM E

1354 and ISO 5660, is that of oxygen consumption. This states that there is a constant

relationship between the mass of oxygen consumed from the air and the amount of

heat released during polymer combustion. In a typical cone calorimeter experiment

the sample is exposed to a defined heat flux, usually 35 or 50 kW/m2; properties such

123
Page 123 of 302
as heat release rate (HRR), peak of heat release (PHRR), time to ignition (TTI), total

heat released (THR), mass loss rate (MLR), mean CO yield and mean specific

extinction area can be simultaneously measured [35, 41]. The HRR is considered to

be the most important variable characterizing a fire. A high HRR causes fast ignition

t
and flame spread, while the PHRR represents the point in a fire where heat is likely to

ip
propagate further, or ignite adjacent objects [35, 239].

cr
As can be seen from Table 37, where cone calorimeter data are presented for

various nanocomposites and the corresponding neat polymers, the incorporation of

us
layered silicates results in significant reductions of PHRR and average HRR.

an
Moreover, it has been demonstrated that the reduction in PHRR is proportional to the

fraction of clay, and it also depends on its aspect ratio and surface charge density.
M
In the literature it is reported that the primary parameter responsible for the lower

HRR of nanocomposites is the MLR during combustion, which, in turn, is also


d

significantly reduced compared to the values observed for the pure polymer. This
te

difference comes into effect shortly after the initial combustion, when the

nanocomposite has had time to form a char on its surface [56, 177]. In fact, it is
p

generally considered that the formation of a thermal insulating and low permeability
ce

char on the outer surface of a nanocomposite during combustion is the key to

understanding improvement in flame retardant properties [35, 56]. More specifically,


Ac

heat transfer from an external source or from a flame promotes thermal decomposition

of the organoclay and the polymer. This results in accumulation and reassembly of

clay platelets on the surface of the burning material [226]. Therefore, the

carbonaceous char formed superficially during combustion is rich in silicates and can

be viewed as a sort of ceramic char-layered silicate nanocomposite [31, 178, 226]. In

fact, XRD and TEM examination of such residues has revealed intercalated structures

124
Page 124 of 302
[241]. Actually, it is exactly its own nanocomposite structure that allows the residue

formed to act as a protective barrier by reducing the heat and mass transfer between

the flame and the polymer. That is, the char insulates the underlying polymer from

heat and also slows oxygen uptake and the escape of volatile gases produced

t
bypolymer degradation. Both these actions interfere with the combustion cycle by

ip
reducing the amount of fuel available for burning [31, 35, 56, 178, 241, 242].

cr
This mechanism has been put forward in most studies reporting on the flame

retardant properties of nanocomposites. For example, combustion experiments on

us
EVA-based nanocomposites showed that the HRR and MLR were reduced by 70-80

an
% in a nanocomposite with low silicate loadings (2–5%) presumably because of a

refractory char-clay surface layer formed by the reassembling of the clay layers and
M
catalyzed charring of the polymer [177].

Duquesne et al. [163] also reported on the fire retardancy of EVA-based


d

nanocomposites, using two montmorillonites: Cloisite Na+ and Cloisite 30B. While
te

the PHRR was clearly reduced when either clay was added to the polymer (relative

decreases of 25 % for Cloisite Na+ and 50 % for Cloisite 30B), the authors noted that
p

the TTI was also reduced and that the THR was similar for pure polymer and the clay-
ce

containing polymer. Comparison of the residues after the cone calorimeter experiment

(Fig. 63) demonstrates the different behavior of the three materials. The pure polymer
Ac

does not give a residue, whereas the EVA/Na+-5 gives a powdery grey “ash,” and the

EVA/30B-5 system gives a fragile , carbonaceous residue around 3 mm thick.

In the case of PE/clay nanocomposites significant reductions of PHRR were

reported, whereas the TTI results were somewhat more complicated. For low clay

contents, TTI was increased, due to the barrier effect of the clay. However, with

increasing clay content a reduction of TTI was attributed to catalytic effects. Again, a

125
Page 125 of 302
large reduction of flammability in nanocomposites compared to pure PE was

attributed to the formation of char [184].

Table 38 summarizes the results of the cone calorimetry tests performed on PU

and PU-based nanocomposites. The data show that the average HRR and PHRR were

t
greatly reduced for the nanocomposites as compared with the reference materials.

ip
However the TTI was similar or slightly reduced for the nanocomposites and the

cr
initial rate of weight loss/heat release was higher, probably due to early

decomposition of the organic modifier of the clay and subsequent clay-catalyzed

us
polymer degradation. As shown by the “fire performance index” (PHRR/TTI), the

an
nanocomposites perform much better than pure PUs. However, they showed higher

rates of smoke emission than their pure PU counterparts, presumably because of a


M
greater degree of secondary condensed phase chemistry in an oxygen-depleted

environment, leading to the formation and evolution of aromatic and carbonaceous


d

species. Finally, Berta et al. [144] conducted vertical burn tests, where a flame was
te

applied twice for 10 s to the base of the test specimens and the burning behavior was

observed. Pure PU materials dripped heavily upon ignition, whereas for the PU
p

nanocomposites dripping was strongly suppressed or eliminated.


ce

HRR plots of PP based RTPO neat resin, microcomposites and nanocomposites

are shown in Fig. 64. The PHRR of the nanocomposite containing 10 wt % clay is 37
Ac

% lower than that of neat resin, while the microcomposite and the neat resin show

very similar behavior . At the end of combustion, the neat resin leaves no residue and

the microcomposite leaves only a little powder, while the nanocomposite leaves a

consistent charlike residue. Generally, PP thermally degrades to volatile products

above 250 °C through a radical chain process, propagated by carbon centered radicals

created by carbon-carbon bond scission. In neat resin, thermal degradation continues

126
Page 126 of 302
during the combustion, and all intermediate degradation products volatilize

completely. On the other hand, there may be physicochemical adsorption of the

intermediate degradation products on clay surfaces in clay nanocomposites. This

adsorption may delay volatilization of the degradation products and promote

t
accumulation of the incomplete degradation products on the clay [185].

ip
Zanetti and Costa [177] also found that the nanodispersed morphology in PE/EVA

cr
nanocomposites leads to a substantial decrease of combustion rate of the polymer

matrix, whereas the microcomposite shows a smaller decrease of the rate of

us
combustion.

an
However, a quite interesting observation is that, although the clay must be

nanodispersed to affect the flammability of nanocomposites, it does not need to be


M
completely delaminated. In other words, it has been argued thatthe flame retardancy

obtained in nanocomposites when merely intercalation has occurred is at least as good


d

as when complete exfoliation is achieved. In fact, excellent performance has been


te

observed when the clay layers have remained separated by only approximately 3 nm,

which is considered to be in the “intercalation” zone [55, 135, 231, 243].


p

At this point, it is worth summarizing a very interesting study by Kashiwagi et al.


ce

[242], aiming to further elucidate the nanocomposite flame retardancy mechanism.

The authors performed cone calorimeter tests on PA6 and nanocomposites containing
Ac

2 % and 5 % clay. They noticed that the MLR (burning) curve (Fig. 65) of each

sample is proportional to the HRR curve (Fig. 66). Thus, the specific heat of

combustion (Hc), obtained from the HRR divided by the MLR remains unchanged for

the three samples, implying that the observed reduction in HRR (and MLR) tends to

be due to chemical and physical processes in the condensed phase, rather than in the

gas phase. To prove this conjecture, the authors exposed samples to the same external

127
Page 127 of 302
flux as in the cone calorimeter, but in nitrogen, to avoid any gas phase effects. Despite

the quantitative difference in MLR between the two cases, the overall differences

among the three samples were very similar between the burning case and the

gasification case. Given that the burning behavior depends on processes in both the

t
gas and the condensed phase, while the gasification behavior depends only on those in

ip
the condensed phase, the authors suggested that the observed improvement in the

cr
flammability resistance of nanocomposites is due to chemical and physical processes

in the condensed phase. During gasification, the nanocomposites appeared to be more

us
viscous than the neat PA6 sample and dark floccules appeared on their surfaces, grew

an
with time (Fig. 67) and were left at the bottom of the container at the end of the test.

Similar carbonaceous floccules were also observed in the residues of the burnt
M
samples tested in the cone calorimeter. Further analysis showed that up to 80 % of the

protective floccules consist of clay particles, while the remaining 20 % consist of


d

thermally stable organic components with possible graphitic structure. The authors
te

attributed the accumulation of clay particles on the burning/gasifying sample surface

to two possible mechanisms. One is the recession of the polymer resin from the
p

surface by pyrolysis with dewetted clay particles left behind, and the other is the
ce

transportation of clay particles pushed by numerous rising bubbles of degradation

products.
Ac

The development of the aforementioned radiative gasification technique (Fig. 68)

is expected to increase the understanding of the condensed phase decomposition

processes of pyrolysis, since the mechanism of nanocomposite flame retardancy can

be looked at in more detail through such techniques. Withapparatus similar to the to

cone calorimeter, this method allows pyrolysis in a nitrogen atmosphere at heat fluxes

similar to those found in fires [35, 56, 242].

128
Page 128 of 302
Summarizing: nanocomposites could offer significant advantages in the area of

polymers flame retardancy and these advantages become even more evident when

nanocomposites are compared with conventional FRs. First of all, only very low

concentrations of silicate are necessary in nanocomposites, resulting in commercial

t
advantages such as low density, lower cost and ease of preparation. Moreover, these

ip
materials are an environmentally friendly alternative to some types of fire retardants,

cr
as they contain no halogens, phosphates or aromatics other than those that may be

present in the polymer matrix; and they do not produce increases in the carbon

us
monoxide and soot levels during combustion like those associated with conventional

an
FRs. Also, while traditional fillers very often severely degrade the physical properties

of the polymer or discolor it, an important feature of nanocomposites is the


M
simultaneous improvement in many physical properties, without any change in the

polymer color [3, 35, 56, 178]. Finally, unlike some FRs, silicates provide physical
d

integrity to the material burning in configurations (e.g. vertical upward combustion)


te

in which dripping of flaming material could occur, which constitutes an additional

hazard of fire propagation to surrounding materials [231].


p

However, even though flame retardancy is one of the most promising properties of
ce

nanocomposites, and despite the fact that these materials have shown to perform very

well in laboratory tests concerning their resistance to fire, much more research needs
Ac

to be carried out to establish their applications and limitations. The potential synergy

between traditional flame retardants and nanocomposites, which is discussed in the

following paragraph, could further extend the uses of nanocomposites in this field

[56].

6.5.2. Synergism between nanocomposites and flame retardants

129
Page 129 of 302
From the point of view of fire retardancy, the most significant advantage offered

by nanocomposite formation is the reduction in PHRR (from cone calorimeter

measurements). However, compared to the virgin polymer, the TTI is usually lowered

while the THR remains the same. This means that the nanocomposite ignites faster

t
than the virgin polymer and that, like the virgin polymer, the nanocomposite is

ip
eventually completely combusted [239, 244, 245]. Moreover, a major limitation of

cr
nanocomposites is that they only work in the condensed phase and do nothing to

inhibit the flame in the gas phase [246]. Therefore, even though these materials can

us
be considered to be flame retarded by definition, they perform poorly when tested

an
under industrially significant or regulatory fire safety tests, such as UL-94V. In fact, it

is believed that the excess quaternary ammonium surfactants used to disperse the
M
clays, also increase the probability of early ignition [239, 246]. Considering also the

disadvantages of traditional FRs, described in the previous paragraph, it becomes


d

obvious that it is necessary to develop novel synergistic flame retardant systems with
te

high efficiency and acceptable environmental impact.

In fact, as will be demonstrated in the following examples, polymer


p

nanocomposites work in a synergistic manner with other FRs. Since a portion of the
ce

FR is replaced by the nanocomposite in a less than 1:1 replacement ratio, products

with improved fire performance and a better balance of properties can be obtained
Ac

[239].

Despite the existing regulatory concerns in the area of halogenated FRs, polymer

nanocomposites combined with such additives are reported in the literature. For

example, Zanetti et al. [247] described a PP-based nanocomposite incorporating

decabromodiphenyl ether (DBE) and antimony oxide (AO) as the FR system. The

cone calorimeter results obtained are presented in Table 39. It is worth noticing that

130
Page 130 of 302
while the reference system ( PP-MA + DBE + AO) showed two peaks in HRR (one at

90 sec followed by a second larger one at 170 sec), the addition of 5 wt % organoclay

eliminated the secondary peak, indicating a more uniform flammability behavior. The

average HRR was lowered by 58 % and the PHRR by 70 %. Also, the nanocomposite

t
significantly increased the burn time for the sample when compared to the flame

ip
retardant control sample.

cr
Si et al. [246] demonstrated that self-extinguishing PMMA-based

nanocomposites, which can pass the stringent UL 94 V0 standard, can be successfully

us
prepared by combining organoclays with halogenated FRs. The authors mixed

an
PMMA, a highly combustible polymer, with DBE, AO and o-MMT. Both neat

PMMA and the control flame retarded material failed the UL 94 V0 standard, while
M
the sample where both FR and clay were added self-extinguished in 1 s after the

application of the flame without dripping. Optical images of the samples are shown in
d

Fig. 69, where it can be seen that the PMMA sample containing both FR and clay is
te

blackened by the smoke of the flame, but remaines unchanged in shape. It was also

noticed that the samples with all three components, clay, DBE and AO show a lower
p

PHRR and average MLR than those with only clay or the FRs.
ce

Chigwada et al. [244] conducted a study aiming to explore if any improvements in

the TTI and THR, while maintaining the reduction in PHRR, can be achieved by the
Ac

incorporation of a small amount of bromine chemically attached to the clay cation.

The authors first prepared an organically-modified clay using ammonium salts with

added oligomeric material consisting of vinylbenzyl chloride, styrene and

dibromostyrene. Then they prepared intercalated nanocomposites by bulk

polymerization. They found that the dibromostyrene enhances the flame retardancy of

PS nanocomposites as compared with both the virgin polymer and nanocomposites

131
Page 131 of 302
prepared from non-halogen containing, organically-modified clays. More specifically,

with the bromine-containing organoclay, nanocomposites were obtained exhibiting

reduced PHRR, THR and improved thermal stability. The reduction in PHRR is due

to nanocomposite formation and not to the presence of bromine, which, however, is

t
important in the reduction of the THR, i.e. the presence of bromine, even at less than

ip
4 %, prevents the PS burning.

cr
Another interesting study was carried out by Tang et al. [248]. They did clay

modification and intercalation simultaneously during melt mixing. More specifically,

us
they melt compounded PP, Na-MMT and hexadecyltrimethylammonium bromide to

an
yield a nanocomposite, suggesting that clay modification occurred in the molten

polymer matrix. Therefore, the sodium bromide produced by the ion-exchange


M
reaction remained in the nanocomposite. Subsequently, the authors combined the

nanocomposites obtained with an intumescent system consisting of pentaerythritol


d

and ammonium polyphosphate, with melamine phosphate sometimes added. Even


te

though synergism was observed between the nanocomposites and the intumescent

FRs, the interpretation of the data is complicated by the presence of NaBr, which acts
p

as a vapor phase FR. In fact, it remains unclear whether the improved flammability
ce

properties reported are due to the nanocomposite or to NaBr. In any case, as Morgan

[239] noted in his review on nanocomposite-FR synergism, this novel method of


Ac

nanocomposite formation should be investigated further, especially if the NaBr can be

removed, to produce a truly non-halogenated flame retardant formulation.

Tang et al. [183] also observed that when MMT and intumescent FRs are

combined in a polymer matrix, their synergistic effect is related to their composition

ratio. It was anticipated that the reassembling of silicate layers on the surface of the

nanocomposite during combustion may hinder NH3 from swelling, leading to a

132
Page 132 of 302
negative effect on the fire retarding properties. NH3 is the main gaseous product; it

volatilizes and makes the mixture of the carbonaceous residue and

phosphocarbonaceous materials swell, leading to the formation of the intumescent

residue char. Therefore, when the mass of MMT is increased, the negative effect may

t
exceed the positive effect on the fire retardingproperty, –which explains why the

ip
synergistic effect was no longer demonstrated at high MMT loadings.

cr
In another study dealing with the synergism of nanocomposites and intumescent

systems, the polyol (carbonization agent) was replaced by PA6. More specifically, a

us
PA6 nanocomposite was compounded with EVA and ammonium polyphosphate (acid

an
source). The results showed that the PHRR and THR were decreased by the PA6

nanocomposite, while the mechanical performance was significantly improved [249].


M
Another important group of FRs which exhibit synergism with nanocomposites

are the phosphorous-based FRs. For example, Chigwada et al. [245] observed
d

reduction in the PHRR, THR and MLR, when phosphorous-containing FRs were
te

incorporated in vinylester nanocomposite. The authors reported that these reductions

were directly proportional to the amount of phosphate added.


p

Zheng and Wilkie [250] investigated the synergistic action of phosphates and
ce

nanocomposites by incorporating the phosphate in the organic treatment of the clay.

They treated Na-MMT with a styrenic oligomer that contained ammonium pendant
Ac

groups and co-polymerized vinylphenyl phosphate. Then, PS nanocomposites were

prepared, either by melt intercalation or intercalation from solution. TGA/FT-IR

showed that the phosphate is released during thermal decomposition, and acts in the

gas phase as an FR. The cone calorimeter data, on the other hand, showed that in the

presence of the phosphate, the average HRR and PHRR can be signicantly reduced

while the time to PHRR is increased up to ca. 40 sec.

133
Page 133 of 302
Quite similarily, Kim et al. [251] intercalated triphenylphosphate (TPP) in the

galleries of an organomodified montmorillonite (Cloisite 30B). The clay obtained was

subsequently melt mixed with ABS yielding nanocomposite structures. The key

advantage of intercalating the phosphate in the organoclay is that the normally volatile

t
TPP is protected during melt compounding, allowing for a grater range of melt

ip
processing without TPP loss.

cr
On the other hand, Hussain et al. [252] introduced organophosphinate into the

backbone of an epoxy resin, which was subsequently crosslinked in the presence of o-

us
MMT. Quite surprisingly, flame retardant antagonism was observed with the PHRR

an
and THR increasing for the nanocomposite + FR system, as compared with the epoxy

+ FR control sample. No clear reason was given in this study; but in a later review
M
article, Morgan [239] indicated that the observed antagonism could have been due to

the lack of uniform clay dispersion.


d

Another example where flame retardant antagonism may be observed is the case
te

of FR systems that achieve flame retardancy by dripping away from the flame. As

Morgan [239] noticed, since nanocomposites inhibit polymer flow and dripping
p

during flaming combustion, such FRs will not work with a polymer nanocomposite
ce

system.

In general, however, the synergism reported between clays and conventional FRs
Ac

provides an alternative for improving fire performance without seriously deteriorating

appearance and mechanical behavior, and even opens the possibility of formulating

self-extinguishing polymer-based materials.

6.6. Heat Distortion Temperature

134
Page 134 of 302
Heat distortion temperature or heat deflection temperature (HDT) is the

temperature at which a polymer sample deforms under a specified load. Thus, it is an

index of a polymeric material’s heat resistance towards applied load and is assessed

by the procedure given in ASTM D-648 [55].

t
In general, improvements of HDT are reported by nanocomposite formation.

ip
Usually, a significant increase is achieved for clay contents of approximately 5 wt %,

cr
and then HDT values level off for higher clay loadings [94, 208, 253].

For example, Kojima et al. [94] showed that the HDT of pure PA6 increases up to

us
90 °C after nanocomposite preparation with OMLS, and they reported the clay

an
content dependence of HDT of PA6/MMT nanocomposites. More specifically, they

showed that there is a marked increase in HDT from 65 °C for neat PA6 to 152 °C for
M
4,7 wt % nanocomposite. Beyond that MMT loading, the HDT of the nanocomposites

levels off. Moreover, the HDT values of various PA6 nanocomposites prepared with
d

clay lamellae of different lengths showed that the HDT also depends on the aspect
te

ratio of dispersed clay particles.

Similarly, nanodispersion of MMT in a PP matrix has been found to promote a


p

higher HDT (Table 40). In the case of PP/f-MMT there is a marked increase of the
ce

HDT, from 109 °C for the neat polymer to 152 °C for a 6 wt % of clay, after which

the HDT of the nanocomposites levels off. When the same neat PP polymer is filled
Ac

with alkylammonium modified MMT, the HDT is also increased but to a smaller

extent, reflecting the lower exfoliation level of the inorganic fillers [254].

It should be emphasized that the increase of HDT due to clay dispersion is very

important, not only from application or industrial point of view, but also because it is

very difficult to achieve similar HDT enhancements by chemical modification or

reinforcement by conventional fillers [31, 50, 55].

135
Page 135 of 302
6.7. Rheological properties

The measurement of the rheological properties of any polymeric material is

t
crucial to gain fundamental understanding of the processability of that material and is

ip
usually conducted by either dynamic oscillatory shear or steady shear measurements

cr
[50, 55]. In the case of polymer-layered silicate nanocomposites, the study of

rheological properties is instructive for two reasons: First, these properties are

us
indicative of melt processing behavior in unit operations, such as injection molding.

an
Second, since the rheological properties of particle-filled materials are sensitive to the

structure, particle size, shape and surface characteristics of the dispersed phase,
M
rheology potentially offers a means to assess the state of dispersion in

nanocomposites, directly in the melt state. Thus, rheology can be envisaged as a tool
d

that is complementary to traditional methods of materials characterization [15, 37,


te

50].

It is generally established that when nanocomposites are formed, the viscosity at


p

low shear rates increases with filler concentration [15]. Very often, solid-like behavior
ce

is observed, which is attributed to the physical jamming or percolation of the

randomly distributed silicate layers, at surprisingly low volume fraction, due to their
Ac

anisotropy [55]. On the other hand, at high shear rates, shear thinning behavior is

usually observed [15]. It has been suggested that this is the result of the alignment of

silicate layers towards the direction of flow at high shear rates. Such observations

support the percolation argument used in the case of nanocomposite rheological

behavior under low shear [55]. Typical shear viscosity curves as a function of shear

are presented in Fig. 70 for PBS-based nanocomposites with various clay loadings.

136
Page 136 of 302
Shear thinning behavior at high shear rates has also been observed for a PA6

nanocomposite. In fact, this behavior is similar to that of neat PA6 and the composites

with the same amount of glass fiber or unmodified MMT. It is of great interest to note

that the absolute value of the melt viscosity of the nanocomposite is significantly

t
lower than that of neat nylon 6 or the other composites – which implies good melt

ip
processability over a wide range of processing conditions. One possible reason for

cr
reduction of melt viscosity in the nanocomposite is slip between the PA6 matrix and

the exfoliated organoclay platelets during high shear flow. Another possibility is a

us
reduced molecular weight of the nylon 6 due to degradation (e.g. hydrolysis) in the

an
presence of clay [15]. Similarly, capillary data revealed that nanocomposites based on

medium and low molecular weight PA6 exhibited lower viscosities than the
M
corresponding pure polymers [14].
d

6.8. Crystallinity
te

Crystal formation includes two stages, namely nucleation and crystal growth.
p

However, although it is well-established that nanometer sized clay platelets are


ce

effective nucleating agents, different effects have been reported on the linear growth

rate and the overall crystallization rate, depending on the type of polymer [97].
Ac

For example, Maiti et al. [175] found that although clay particles act as nucleating

agents for the crystallization of a PP-MA matrix, the linear growth rate and the overall

crystallization rate are not significantly influenced by the presence of clay.

Similarily, Di Maio et al. [255] studied the isothermal crystallization of PCL/clay

nanocomposites and noticed that the dispersed clay platelets act as nucleating agents

in the PCL matrix, remarkably reducing the crystallization half-time t1/2. By DSC

137
Page 137 of 302
analysis after isothermal crystallization (Fig. 71), the authors observed a reduction of

the melting temperature with the increase of clay content, indicating a reduced degree

of crystals perfection and degree of crystallinity. This was attributed to the

confinement of chains and segments in the presence of clay, hindering the segmental

t
rearrangement during crystallization and restricting the formation of perfect crystals

ip
in the polymer matrix.

cr
Ke and Yongping [222] conducted DSC analysis on intercalated PET/oMMT

nanocomposites. They found a reduction of Tg in the composite compared to the pure

us
matrix, which they attributed to the plasticizing effect of oMMT. However, they

an
noticed that by increasing the oMMT content the Tg is increased, as shown in Fig. 72.

Further, they observed that the cold crystallizing point of pure PET is 150 °C, while
M
for the nanocomposite it decreases to 130 °C. This result shows that adding oMMT

into PET is favorable to its crystallization.


d

For the effect of nanocomposite preparation on EVA crystallinity, contradictory


te

results have been reported. For example, Chaudhary et al. [171] found that the

crystallization process for EVA-28 is significantly affected by the presence of clays,


p

while Gelfer et al. [256] argued that the effect of clay on the crystalline phase in
ce

EVA-3 is insignificant. They observed the same crystalline structure (orthorhombic)

and similar melting behavior, as well as transition temperatures in pure EVA-3 and
Ac

nanocomposites containing 10 wt % clay. Gelfer et al. considered this to be an

anticipated result, since it is known that in ethylene-based copolymers side chains

(short branches) larger than the methyl group are excluded from the crystalline

domain and are aggregated into the amorphous phase. Consequently, the amorphous

domains in EVA are enriched by the polar VA units. It was hypothesized that the

organoclay particles dispersed in the EVA matrix have higher affinity with the polar

138
Page 138 of 302
VA units; hence they are predominantly confined to the amorphous phase where the

VA content is high, without significantly affecting the crystalline domain formed by

the non-polar ethylene segments.

While further work needs to be done to elucidate crystallization phenomena in

t
nanocomposites based on different polymers, the effect of layered silicates on

ip
polyamide crystallinity has been far more investigated. In fact, a number of

cr
researchers have focused on the effect of clay addition on the crystal morphology of

polyamides, as well as on the melting, glass transition, crystallization temperature and

us
crystallization rate. In semi-crystalline PA6, two phases generally exist at room

an
temperature. The α-phase and the γ-phase are monoclinic, but the latter can be

represented as a pseudo-orthotropic lattice.


M
In general, it has been observed that upon preparation of nanocomposites the

crystallography of the PA6 matrix changes. Maiti and Okamoto [257] found that in
d

the presence of clay particles, PA6 crystallizes faster and exclusively in the γ-phase
te

(Fig. 73). The nucleation and growth processes of PA6 in the nanocomposites are

presented in Fig. 74, as deduced from from direct observation by TEM.


p

Similarly, Lincoln et al. [12] showed through simultaneous SAXS and WAXS
ce

analyses at elevated temperature that the presence of silicate layers provides a

confined environment for crystallization and disrupts crystallite formation, ultimately


Ac

resulting in the predominance of the γ-phase in quenched samples. They noted that the

more ductile γ-phase that is formed may contribute greatly to the observed changes in

properties of the material, specifically the toughness. Evidence has also been found

that the layers affect not only the formation of the lamellae, but also of spherulites,

which may also be the root of some of the property enhancements. Moreover, certain

processes involved in nanocomposite preparation affect the global orientation of the

139
Page 139 of 302
crystal phase. That is, X-ray diffraction of PA/MMT nanocomposites has indicated

that the chain axes are normal to the silicate layers in the interior of an injection

molded component, while they are parallel to the layers in the near surface region of

molded and extruded components.

t
Varlot et al. [5] suggested that the overall crystallinity of the polyamide matrix

ip
remained unchanged with the addition of montmorillonite, even though the proportion

cr
of the γ-phase was higher.

Concerning the glass transition, a decrease in Tg has been found for PA6/clay

us
nanocomposites; this was attributed to an increase in the amorphous volume fraction

an
due to the surface/volume ratio of the crystallites, the number of nuclei and the

perfection of the crystal lamellae [139]. However, DSC measurements conducted by


M
other researchers, indicate that the presence of layered silicate filler does not affect the

Tg of the PA6 matrix, which occurs at approximately 53 °C. The melting peak of the
d

composites is at a slightly lower temperature than that of neat PA6, due to a slight
te

reduction of crystallite size in the presence of fillers. All the fillers cause an increase

in crystallization temperature relative to neat PA6. To various degrees, fillers may act
p

as nucleating agents, causing a higher crystallization rate than that of the neat PA6
ce

[15]. The same phenomena have also been observed in PA66 based nanocomposites

[146].
Ac

Finally, Wu et al. [97] reported DSC curves for isothermal crystallization of

PA1012 and PA1012/clay nanocomposites (Fig. 75). The crystallization exothermic

peak of the PA1012/clay nanocomposites is narrower than that of PA1012, indicating

that the nanocomposites have a higher crystallization rate. The heat of crystallization

of the nanocomposite is less than that of the PA1012, implying that the clay platelets

decrease the polymer crystallinity. The higher crystallization rate recorded for

140
Page 140 of 302
PA1012 nanocomposites was attributed to the nucleating action of clay platelets. On

the other hand, the authors suggested that the confinement of the chains and segment

movement will hinder segmental rearrangement during crystallization and restrict the

formation of perfect crystals in the PA1012/clay nanocomposite, thus resulting in

t
decreased crystallinity.

ip
cr
6.9. Biodegradation

us
An interesting aspect of nanocomposite technology is the enhancement in

an
biodegradability, often reported after nanocomposite formation. Tetto et al. [258] first

presented results on the biodegradability of nanocomposites based on PCL, reporting


M
that the PCL/OMLS nanocomposites showed improvedbiodegradability compared

withpure PCL.
d

Biodegradability improvements have also been reported in the case of PLA-based


te

nanocomposites. As an example, Sinha Ray et al. [208] presented photographs of the

recovered samples of neat PLA and PLACN4 from compost with time (Fig. 76(a)).
p

The decreased Mw and residual weight Rw percentage of the initial test samples with
ce

time is also shown in Fig. 76(b). Obviously, the biodegradability of neat PLA is

significantly enhanced after nanocomposite preparation. For a month, both the extent
Ac

of Mw loss and weight loss are are almost the same for neat PLA and PLACN4.5

However, after one month, a sharp change occurs in weight loss of PLACN4, and

within two months, it is completely degraded by compost. The degradation of PLA in

compost is a complex process involving: water absorption, ester cleavage and

formation of oligomer fragments, solubilization of oligomer fragments, and finally

diffusion of soluble oligomers. Therefore, any factor that increases the hydrolysis

141
Page 141 of 302
tendency of neat PLA, ultimately controls the degradation. However, from Fig. 76(b)

it is evident that the hydrolysis tendencies of PLA and PLACN4 are almost the same.

Sinha Ray et al. suggest that the presence of terminal hydroxylated edge groups of

the silicate layers may be one of the factors responsible for this behavior. In PLACN4,

t
the stacked and intercalated silicate layers are homogeneously dispersed in the PLA

ip
matrix, and these hydroxy groups initiate heterogeneous hydrolysis of the PLA matrix

cr
after absorbing water from the compost. This process takes some time to start. For this

reason, according to these authors, the weight loss and degree of hydrolysis of PLA

us
and PLACN4 is almost same up to one month (Fig. 76(b)). However, after one month

an
there is a sharp weight loss in the case of PLACN4 as compared with that of PLA.

This means that one month is a critical time to start heterogeneous hydrolysis, and due
M
to this type of hydrolysis the matrix degrades into very small fragments and is

eliminated with the compost. This assumption was confirmed by conducting the same
d

type of experiment with PLACN prepared by using dimethyl dioctadecylammonium


te

salt modified synthetic mica which has no terminal hydroxylated edge group; in that

case the degradation tendency was almost the same as that of PLA.
p

However, contradictory results concernng the effect of clay dispersion on polymer


ce

biodegradability are also found in the literature. In fact, Lee et al. [259] prepared

nanocomposites based on aliphatic unsaturated polyester and found a decrease in


Ac

biodegradability under composting with intercalation. They assumed that due to the

high aspect ratio and better dispersion of clay in the matrix a more tortuous path

formed for penetration of microorganisms into the bulk and hindered their diffusion.

Similarly, Maiti et al. [195] attributed the suppressed biodegradation tendency, which

they observed in the case of PHB nanocomposites, to an improvement of the barrier

properties of the matrix after nanocomposite formation, even though they did not

142
Page 142 of 302
report on the permeability. This explanation contradicts the results of Sinha Ray et al.

[260], who found no relation between biodegradability and barrier properties in

PLA/OMLS nanocomposites

From the aforementioned contradictory results, it becomes obvious that the

t
increase or the decrease in nanocomposite biodegradation is still under discussion and

ip
no conclusion can be driven about mechanisms on the basis of the current literature

cr
[231].

us
6.10. Photo-degradation

an
The degradability of nanocomposites under UV light is a serious problem that
M
may limit their applications. In the few studies addressing this issue, it has been found

that nanocomposites exhibit lower stability than the corresponding neat polymers
d

[231].
te

For example, Huaili et al. [261] studied the photo-oxidative degradation of

PE/MMT nanocomposites compared with neat polyethylene. Since it is well-


p

established that the degradation of hydrocarbon chains leads to the formation of


ce

hydroxyl and keto groups, they studied the extent of photo degradation by FTIR

observations. As is shown in Fig. 77, the degradation of PE/o-MMT nanocomposite


Ac

was greater than that of pure PE polymer after 200 h irradiation. Fig. 78 shows FTIR

spectra in the carbonyl region upon UV irradiation. Obviously, there is a considerable

increase in the intensity of the carbonyl region with an increase of irradiation time in

PE/o-MMT, which means that the material is undergoing degradation. In the pure PE,

on the other hand, the intensity in the carbonyl region was significantly less, which

indicates less degradation.

143
Page 143 of 302
Morlat [262] and Mailhot et al. [263] studied the effect of compatibilizers on

photo-degradation and its kinetics by comparing PP nanocomposites with neat

polymer. The increase in the absorbance at 3200-3600 cm-1 and 1600-1800 cm-1 was

rapid in nanocomposites in comparison with neat polymer. It was observed that the

t
induction period decreased from 8 to 4 h by using PP-g-MA as compatibilizer and a

ip
two-phase degradation mechanism was observed. In the first stage (up to 40 h) there

cr
was no evidence for hydroxyl band formation in the IR spectra, which implies the

absence of degradation on polymer backbone, whereas in the second stage a dramatic

us
increase in the rate of photo-oxidation was found. The degradation products were the

an
same in the composite and the neat polymer.

According to Patterson et al. [264], the incorporation of layered silicates in


M
polycarbonate appeared to increase the rate at which chain scission occurs.

Furthermore, these carbonate scissions produced a yellowing of the polycarbonate,


d

which could inhibit its use in applications where optical clarity is important.
te

However, in another study, the effect of accelerated weathering of polycarbonate

nanocomposites was investigated. The silicate content used ranged from 0 to 3.5
p

wt.%. A UV-accelerated weathering tester programmed to cycle for 8 h of UV


ce

radiation and 4 h of dark condensation was selected for the exposure study. The

materials were characterized by UV/vis spectroscopy and FTIR spectroscopy and it


Ac

was concluded that the degradation of the nanocomposite was less than that of the

neat polymer [265].

In any case, although no explanation has so far been given for the differences in

photo-degradation stability between the nanocomposites and the pure polymers, it has

been suggested that the best way to increase the outdoor durability would be to

144
Page 144 of 302
develop nanocomposites by modification of the clay rather than functionalization of

thermoplastics [231].

6.11. Optical clarity

t
ip
Microsized particles used as reinforcing agents scatter light, thus reducing light

cr
transmittance and optical clarity [266]. On the other hand, layered silicate platelets,

albeit their micron lateral size, are just 1 nm thick. Thus, when single layers are

us
dispersed in a polymer matrix, the resulting nanocomposite is optically clear in the

an
visible region, whereas, there is a loss of intensity in the UV region (for λ< 250 nm)

mostly due to scattering by the MMT particles. There is no marked decrease in the
M
clarity due to the nano-dispersed fillers (e.g. one has to load 20 wt % of C18-MMT in

3 mm thick film of PP before there develops haze observable by the eye). This is a
d

general behavior in UV/vis transmittance for thick films (3-5 mm) of polymer/MMT
te

nanocomposites based on PVA, PP, and several epoxies [31].


p

7. Nanocomposites: advantages and applications


ce

As described above, polymer nanocomposite materials often exhibit properties


Ac

superior to conventional composites, such as strength, stiffness, thermal and oxidative

stability, barrier properties, as well as flame retardant behavior [267]. These improved

properties are generally attained at lower filler content in comparison with

conventionally filled systems. Therefore, polymer-layered silicate nanocomposites are

far lighter in weight than a conventional composite, which makes them quite

competitive for specific applications [55].

145
Page 145 of 302
Moreover, for systems with favorable thermodynamics of mixing, the organoclay

can be incorporated in the final stages of polymer processing (e.g. extrusion, injection

or compression molding) to obtain nanocomposite materials. Thus, polymer

nanocomposites are amenable to most of the common processing techniques in

t
today’s industrial practice – which will lower the barriers to their commercialization

ip
[31].

cr
Another unique aspect of nanocomposites is the lack of property trade-offs.

Traditionally, blend or composite formulations require trade-offs between desired

us
performance, mechanical properties, cost and processability. However, polymer

an
nanocomposite technology provides a route around these traditional limitations, and

offers, for the first time, the opportunity to design materials without the compromises
M
typically found in conventionally filled polymers [267-268].

The aforementioned attractive characteristics of polymer-layered silicate


d

nanocomposites already suggest a variety of possible industrial applications:


te

automotive (gas tanks, bumpers, interior and exterior panels), construction (building

sections and structural panels), aerospace (flame retardant panels and high
p

performance components), food packaging, textiles, etc. [269].


ce

It is for this reason that many companies have taken a strong interest and have

invested in developing nanoclays (Table 41) and polymer nanocomposites (Table 42)
Ac

[270].

Among these, the first commercial product of clay-based polymer nanocomposites

was the timing-belt cover made from PA6 nanocomposites by Toyota Motors in the

early 1990s. This timing-belt cover exhibited good rigidity, excellent thermal stability

and no wrap. It also saved weight by up to 25 % [270]. Later, General Motors and

partners Basell, Southern Clay Products and Blackhawk Automotive Plastics

146
Page 146 of 302
announced external automotive body parts (step-assist) made from thermoplastic,

olefin layered silicate nanocomposites. A TPO nanocomposite with as little as 2.5 %

layered silicate is as stiff and much lighter than parts with 10 times the amount of

conventional talc filler. Thus, the weight savings can reach 20 %, depending on the

t
part and the material that is being replaced by the TPO nanocomposite [268]. William

ip
Windscheif, Basell’s Global Business Vice President for Advanced Polyolefins,

cr
called this application “a small step, but a giant one for nanocomposites” and added

that it heralds a broader shift to nano-PP in automotive technology [271].

us
Consultant Kenneth Sinclair, head of STA Research in Sndiomish, Wash.,

an
estimated that in the near future the application of PP-based nanocomposites in the

automotive industry will expand, with nano-PP mostly cannibilizing existing PP


M
applications. However, replacement of metals and engineering thermoplastics will

follow. In this context, a long term goal for Dow Plastics is in-reactor compounding
d

of nano-PP by using nano-clays as the catalyst support for in-situ polymerization of


te

PP homopolymer. Dow’s effort is focused on highly loaded (up to 10 % clay) nano

PPs for semi-structural automotive uses [271].


p

Meanwhile, a role for nanocomposites in polycarbonate automotive glazing is


ce

being explored by Exatec of Wixon, Mich., a joint venture of Bayer and GE Plastics,

that is dedicated to PC auto-glazing development, considered for the exterior coating


Ac

needed to achieve weatherability and abrasion resistance without reducing clarity. A

Bayer coating containing nanoparticles is one of several promising approaches being

pursued [271].

It is worth noticing that the weight advantage of polymer nanocomposites could

have a significant impact on environmental protection and material recycling. It is

predicted that widespread use of polymer nanocomposites would save 1.5 billion liters

147
Page 147 of 302
of gasoline over the life of one year’s production of vehicles and reduce related CO2

emissions by more than 5 billion kilograms [270].

Next to the automotive industry, polymer nanocomposites are expected to find

wide applications as barrier materials. In fact, the excellent barrier properties of clay-

t
based polymer nanocomposites could result in considerable enhancement of shelf-life

ip
for many types of packaged food. Meanwhile, the optical transparency of polymer

cr
nanocomposite film is generally similar to their pristine counterparts, which is

impossible with conventional polymer composites. Therefore, the above property

us
advantages would make them widely acceptable in packaging industries as wrapping

an
films and beverage containers. For example, Bayer has developed a new grade of

plastic films for food packaging, which are made from PA6 exfoliated
M
nanocomposites [270]. Honeywell is also developing nano-PA materials that can beat

the cost of high-barrier plastics or even glass. Nanocor’s Imperm compound


d

supplements the inherent gas barrier of amorphous MDX6 nylon from Mitsubishi Gas
te

Chemical with the addition of nanoclay. Used as the core of a three-layer PET bottle,

Imperm reportedly ensured a 28.5 week beer shelf life. Imperm is said to adhere to
p

PET without tie layers, while sufficient clarity is retained to meet requirements for the
ce

amber bottle [271].

Meanwhile, Ube America is developing nanocomposite barriers for automotive


Ac

fuel systems, using up to 5 % nanoclay in PA6 and PA6/66 blends [271].

Also quite interesting are potential applications of nanocomposites based on

biodegradable polymers. Such polymers have become indispensable in a wide range

of applications. However, despite their attractive degradation characteristics and

significant demand for such materials, the lack of structural and functional stability

prevents currently available biodegradable polymers from having widespread

148
Page 148 of 302
commercial impact [267]. Therefore, biodegradable polymer-based nanocomposites

appear to have a very bright future for a wide range of applications as high-

performance biodegradable materials [55].

Layered silicate nanoparticles, when distributed within the matrix of a fiber-

t
reinforced polymer (FRP), can retard the diffusion of environmental moisture and

ip
other chemicals to the fiber-matrix interface where their presence can result in

cr
delamination and fiber weakening. Thus, the use of nanoparticles helps to preserve the

integrity of FRPs and to prolong the service life of composites when these are used in

us
outdoor applications such as bridges and utility poles [272].

an
Other potential nanocomposite applications include nano-pigments that are

believed to be an environment-friendly alternative to toxic cadmium and palladium


M
pigments, as well as controlled drug delivery systems, etc [270].

However, despite the current optimism surrounding polymer-nanocomposite


d

materials, the mechanical performance benchmarks set by highly loaded short-fiber


te

composites are still beyond the capabilities of low concentration nanocomposites.

Manufacturing costs also remain a significant factor restricting the growth of


p

polymer-nanocomposite applications [200]. In fact, some early application


ce

development programs have lapsed for cost reasons. Such casualties include the

timing-belt cover based on nylon 6 nanocomposite [271].


Ac

According to Silberglitt, there are two possible paths or trends – a high-growth

path, under which nanocomposite materials are pervasively applied throughout

society and a low-growth path, under which the use of nanocomposites leads to

incremental improvements in specific technology areas (Fig. 79) [269].

149
Page 149 of 302
8. Summary

Polymer-layered silicate nanocomposites, although known for many years, have

attracted recent attention due to the report of the Toyota research group on the

improved properties of PA6 nanocomposites and also due to the observation by

t
ip
Giannelis and co-workers that their preparation is possible by simple melt-mixing of

cr
the polymer with the layered silicate. Other preparation routes include intercalation of

polymer or prepolymer from solution, in situ intercalative polymerization and

us
template synthesis. In most cases, layered silicates first need to be modified with

cationic-organic surfactants, in order to become miscible with polymeric matrices.

an
Then, whether a nanocomposite will form or not, and whether this will be intercalated

or exfoliated, depends on a variety of factors. These include the type of polymer,


M
layered silicate and organic modifier, the preparation technique and processing

conditions. In general, nanocomposite materials, particularly those with exfoliated


d

structures present significant improvements of modulus and strength, whereas


te

contradictory results are reported concerning their elongation and toughness.


p

Improvements of storage and loss moduli are also reported by many authors. Other
ce

interesting characteristics of this class of materials include improved barrier

properties, thermal stability and flame retardance. Despite some contradictory results
Ac

reported in the literature and presented here, concerning certain aspects of polymer-

layered silicate nanocomposite technology, we hope this review will be a useful tool

for those conducting research in this field.

150
Page 150 of 302
REFERENCES

[1] Alexandre M, Dubois P. Polymer-layered silicate nanocomposites:


preparation, properties and uses of a new class of materials. Mat Sci Eng R
2000;28:1-63.

t
ip
[2] Fischer H. Polymer nanocomposites: from fundamental research to specific
applications. Mat Sci Eng C 2003;23:763-772.

cr
[3] Lagaly G. Introduction: from clay mineral-polymer interactions to clay
mineral-polymer nanocomposites. Appl Clay Sci 1999;15:1-9.

us
[4] Giannelis EP. Polymer layered silicate nanocomposites. Adv Mater 1996;8:29-
35.
[5] Varlot K, Reynaud E, Kloppfer MH, Vigier G, Varlet J. Clay-reinforced

an
polyamide: preferential orientation of the montmorillonite sheets and the
polyamide crystalline lamellae. J Polym Sci Pol Phys 2001;39:1360-1370.
M
[6] Gorrasi G, Tortora M, Vittoria V, Galli G, Chiellini E. Transport and
mechanical properties of blends of poly(ε -caprolactone) and a modified
montmorillonite-poly(ε -caprolactone) nanocomposite. J Polym Sci Pol Phys
d

2002;40:1118-1124.
te

[7] LeBaron PC, Wang Z, Pinnavaia TJ. Polymer-layered silicate


nanocomposites: an overview. Appl Clay Sci 1999;15:11-29.
p

[8] Vaia RA, Wagner HD. Framework for Nanocomposites. Materials Today
ce

2004;7:32-37.
[9] Ginzburg VV, Singh C, Balazs AC. Theoretical phase diagrams of
polymer/clay composites: the role of grafted organic modifiers.
Ac

Macromolecules 2000;33:1089-1099.
[10] Osman MA, Mittal V, Lusti HR. The aspect ratio and gas permeation in
polymer-layered silicate nanocomposites. Macromol Rapid Comm
2004;25:1145-1149.
[11] Balazs AC, Singh C, Zhulina E, Lyatskaya Y. Modeling the phase behavior of
polymer/clay nanocomposites. Acc Chem Res 1999;32:651-657.
[12] Lincoln DM, Vaia RA, Wang Z-G, Hsiao BS. Secondary structure and
elevated temperature crystallite morphology of nylon 6/layered silicate
nanocomposites. Polymer 2001;42:1621-1631.

151
Page 151 of 302
[13] Vaia RA, Giannelis EP. Liquid crystal polymer nanocomposites: direct
intercalation of thermotropic liquid crystalline polymers into layered silicates.
Polymer 2001;42:1281-1285.
[14] Fornes TD, Yoon PJ, Keskkula H, Paul DR. Nylon 6 nanocomposites: the
effect of matrix molecular weight. Polymer 2001;42:9929-9940.

t
[15] Cho JW, Paul DR. Nylon 6 nanocomposites by melt compounding. Polymer

ip
2001;42:1083-1094.
[16] Bower CA. Studies on the form and availability of organic soil phosphorous.

cr
Iowa Agricoltural Experiment Station Research Bulletin 1949;362:39-42.
[17] Carter LW, Hendricks JG, Bolley DS. Elastomer reinforced with modified

us
clay. (assigned to National Lead Co) United States Patent No 2531396 1950.
[18] Hauser EA, Kollman RC. Clay complexes with conjugated unsaturated

an
aliphatic compounds. US Patent No 2951087 1960.
[19] Uskov IA. Filled polymers. III. Polymerization of methyl methacrylate during
dispersion of sodium bentonite. Vysokomol Soed 1960;2:926-930.
M
[20] Blumstein A. Etude des polymerisations en couche adsorbee. Bull Chim Soc
1961;899-906.
[21] Greenland DJ. Adsorption of poly(vinyl alcohols) by montmorillonite. J
d

Colloid Sci 1963;18:647-664.


te

[22] Nahin PG, Backlund PS. Organoclay-polyolefin compositions. (assigned to


Union Oil Co) United States Patent No 3084117 1963.
p

[23] Tanihara K, Nakagawa M. Flocculation treatment of waste waters containing


ce

montmorillonite. IV. Interlamellar complex formation between various ion


forms of montmorillonite and poly(ethylene oxide) or polyacrylamide. Nippon
Kagaku Kaishi 1975;5:782-789.
Ac

[24] Fujiwara S, Sakamoto T. Method for manufacturing a clay/polyamide


composite. (assigned to Unichika K. K., Japan) Japanese Kokai Patent
Application No 109998 1976.
[25] Usuki A, Kojima Y, Kawasumi M, Okada A, Fukushima Y, Kurauchi T,
Kamigatio O. Synthesis of nylon 6-clay hybrid. J Mater Res 1993;8:1179-
1184.
[26] Usuki A, Koiwai A, Kojima Y, Kawasumi M, Okada A, Kurauchi T,
Kamigatio O. Interaction of nylon 6-clay surface and mechanical properties of
nylon 6-clay hybrid. J Appl Polym Sci 1995;55:119-123.

152
Page 152 of 302
[27] Okada A, Usuki A. The chemistry of polymer-clay hybrids. Mat Sci Eng C –
Bio S 1995;3:109-115.
[28] Okada A, Fukushima Y, Kawasumi M, Inagaki S, Usuki A, Sugiyama S,
Kurauchi T, Kamigatio O. Composite material and its preparation. (assigned
to Toyota Motor Co., Japan) United States Patent No 4739007 1987

t
[29] Vaia RA, Ishii H, Giannelis EP. Synthesis and properties of two-dimensional

ip
nanostructures by direct intercalation of polymer melts in layered silicates.
Chem Mater 1993;5:1694-1696.

cr
[30] Mehrotra V, Giannelis EP. Conducting molecular multilayers: intercalation of
conjugated polymers in layered media. Materials Research Society

us
Symposium Proceedings 1990;171:39-44.
[31] Manias E. Origins of the materials properties enhancements in polymer/clay

an
nanocomposites. http://raman.plmsc.psu.edu/~manias/pdfs/nano2001b.pdf,
2001.
[32] Shelley JS, Mather PT, DeVries KL. Reinforcement and environmental
M
degradation of nylon 6/clay nanocomposites. Polymer 2002;42:5849-5858.
[33] Chin I-J, Thurn-Albrecht T, Kim H-C, Russell TP, Wang J. On exfoliation of
montmorillonite in epoxy. Polymer 2001;42:5947-5952.
d

[34] Miranda-Trevino JC, Coles CA. Kaolinite properties, structure and influence
te

of metal retention on pH. Appl Clay Sci 2003;23:133-139.


[35] Beyer G. Nanocomposites: a new class of flame retardants for polymers.
p

Plastics Additives & Compounding 2002;4(10):22-27.


ce

[36] McNally T, Murphy WR, Lew CY, Turner RJ, Brennan GP. Polyamide-12
layered silicate nanocomposites by melt compounding. Polymer
2003;44:2761-2772.
Ac

[37] Solomon MJ, Almusallam AS, Seefeldt KF, Somwangthanaroj A, Varadan P.


Rheology of polypropylene/clay hybrid materials. Macromolecules
2001;34:1864-1872.
[38] Dixon JB. Roles of clays in soils. Appl Clay Sci 1991;5:489-503.
[39] Manias E, Touny A, Wu L, Strawhecker K, Lu B, Chung TC. Polypropylene/
montmorillonite nanocomposites. Review of the synthetic routes and materials
properties. Chem Mater 2001;13:3516-3523.
[40] Ishida H, Campbell S, Blackwell J. General approach to nanocomposite
preparation. Chem Mater 2000;12:1260-1267.

153
Page 153 of 302
[41] Zanetti M, Lomakin S, Camino G. Polymer layered silicate nanocomposites.
Macromol Mater Eng 2000;279:1-9.
[42] Kornmann X, Lindberg H, Berglund LA. Synthesis of epoxy-clay
nanocomposites: influence of the nature of the clay on structure. Polymer
2001;42:1303-1310.

t
[43] Xie W, Gao Z, Liu K, Pan W-P, Vaia R, Hunter D, Singh A. Thermal

ip
characterization of organically modified montmorillonite. Thermochim Acta
2001;367-368:339-350.

cr
[44] Kim C-M, Lee D-H, Hoffmann B, Kressler J, Stoppelmann G. Influence of
nanofillers on the deformation process in layered silicate/polyamide 12

us
nanocomposites. Polymer 2001;42:1095-1100.
[45] Zerda AS, Lesser AJ. Intercalated clay nanocomposites: morphology,

an
mechanics and fracture behavior. J Polym Sci Pol Phys 2001;39:1137-1146.
[46] Huang J-C, Zhu Z-K, Yin J, Qian X-F, Sun Y-Y. Poly(etherimide)/
montmorillonite nanocomposites prepared by melt intercalation: morphology,
M
solvent resistance properties and thermal properties. Polymer 2001;42:873-
877.
[47] Vaia RA, Giannelis EP. Polymer melt intercalation in organically-modified
d

layered silicates: model predictions and experiment. Macromolecules


te

1997;30:8000-8009.
[48] Yu ZZ, Yang M, Zhang Q, Zhao C, Mai YW. Dispersion and distribution of
p

organically modified montmorillonite in nylon-66 matrix. J Polym Sci Pol


ce

Phys 2003;41:1234-1243.
[49] Wang KH, Choi MH, Koo CM, Choi CM, Chung IJ. Synthesis and
characterization of maleated polyethylene/clay nanocomposites. Polymer
Ac

2001;42:9819-9826.
[50] Ray SS, Bousima M. Biodegradable polymers and their layered silicate
nanocomposites: In greening the 21st century materials world. Prog Mater Sci
2005;50:962-1079.
[51] Fornes TD, Yoon PJ, Hunter DL, Keskkula H, Paul DR. Effect of organoclay
structure on nylon 6 nanocomposite morphology and properties. Polymer
2002;43:5915-5933.

154
Page 154 of 302
[52] Dennis HR, Hunter DL, Chang D, Kim S, White JL, Cho JW, Paul DR. Effect
of melt processing conditions on the extent of exfoliation in organoclay-based
nanocomposites. Polymer 2001;42:9513-9522.
[53] Wu S-H, Wang F-Y, Ma C-CM, Chang W-C, Kuo C-T, Kuan H-C, Chen W-J.
Mechanical, thermal and morphological properties of glass fiber and carbon

t
fiber reinforced polyamide 6 and polyamide 6/clay nanocomposites. Mater

ip
Lett 2001;49:327-333.
[54] Krishnamoorti R, Giannelis EP. Rheology of end-tethered polymer layered

cr
silicate nanocomposites. Macromolecules 1997;30:4097-4102.
[55] Ray SS, Okamoto M. Polymer – layered silicate nanocomposite: a review

us
from preparation to processing. Prog Polym Sci 2003;28:1539-1641.
[56] Porter D, Metcalfe E, Thomas MJK. Nanocomposite fire retardants – a review.

an
Fire Mater 2000;24:45-52.
[57] Ma J, Xu J, Ren J-H, Yu Z-Z, Mai Y-W. A new approach to
polymer/montmorillonite nanocomposites. Polymer 2003;44:4619-4624.
M
[58] Morgan AB, Gilman JW. Characterization of polymer-layered silicate (clay)
nanocomposites by transmission electron microscopy and X-ray diffraction: a
comparative study. J Appl Polym Sci 2003;87:1329-1338
d

[59] VanderHart DL, Asano A, Gilman JW. NMR measurements related to clay
te

dispersion quality and organic-modifier stability in nylon 6/clay


nanocomposites. Macromolecules 2001;34:3819-3822.
p

[60] VanderHart DL, Asano A, Gilman JW. Solid-state NMR investigation of


ce

paramagnetic nylon 6 clay nanocomposites. 2. Measurement of clay


dispersion, crystal stratification, and stability of organic modifiers. Chem
Mater 2001;13:3796-3809.
Ac

[61] VanderHart DL, Asano A, Gilman JW. Solid-state NMR investigation of


paramagnetic nylon 6 clay nanocomposites. 1. Crystallinity, morphology, and
the direct influence of Fe3+ on nuclear spins. Chem Mater 2001;13:3781-3795.
[62] Loo LS, Gleason KK. Fourier transforms infrared investigation of the
deformation behavior of montmorillonite in nylon 6/nanoclay nanocomposite.
Macromolecules 2003;36:2587-2590.
[63] Wu HD, Tseng CR, Chang FC. Chain conformation and crystallization
behavior of the syndiotactic polystyrene nanocomposites studied using Fourier
transform infrared analysis. Macromolecules 2001;34:2992-2999.

155
Page 155 of 302
[64] Fornes TD, Paul DR. Modelling properties of nylon 6/clay nanocomposites
using composite theories. Polymer 2003;44:4993-5013.
[65] Kim SW, Jo WH, Lee MS, Ko MB, Jho JY. Preparation of clay-dispersed
poly(styrene-co-acrylonitrile) nanocomposites using poly(ε -caprolactone) as a
compatibilizer. Polymer 2001;42:9837-9842.

t
[66] Vaia RA, Giannelis EP. Lattice of polymer melt intercalation in organically

ip
modified layered silicates. Macromolecules 1997;30:7990-7999.
[67] Arada P, Ruiz-Hitzky E. Polymer-salt intercalation complexes in layer

cr
silicates. Adv Mater 1990;2:545-547.
[68] Arada P, Ruiz-Hitzky E. Poly(ethylene oxide)-silicate intercalation materials.

us
Chem Mater 1992;4:1395-1403.
[69] Tunney JJ, Detellier C. Poly(ethylene glycol)-kaolinite intercalates. Chem

an
Mater 1996;8:927-935.
[70] Fischer HR, Gielgens LH, Koster TPM. Nanocomposites from polymers and
layered minerals. Acta Polymerica 1999;50:122-126.
M
[71] Theng BKG. Formation and Properties of Clay-Polymer Complexes.
Amsterdam: Elsevier, 1979.
[72] Ogata N, Kawakage S, Ogihara T. Poly(vinyl alcohol)/clay and poly(ethylene
d

oxide)/clay blend prepared using water as solvent. J Appl Polym Sci


te

1997;66:573-581.
[73] Parfitt RL, Greenland DJ. Adsorption of poly(ethylene glycols) on
p

montmorillonites. Clay Miner 1970;8:305-315.


ce

[74] Ruiz-Hitzky E, Aranda P, Casal B, Galvan JC. Nanocomposite materials with


controlled ion mobility. Adv Mater 1995;7:180-184.
[75] Levy R, Francis CW. Interlayer adsorption of polyvinylpyrrolidone on
Ac

montmorillonite. J Colloid Interf Sci 1975;50:442-450.


[76] Billingham J, Breen C, Yarwood J. Adsorption of polyamine, polyacrylic acid
and polyethylene glycol on montmorillonite: an in situ study using ATR-
FTIR. Vib Spectrosc 1997;14:19-34.
[77] Zhao X, Urano K, Ogasawara S. Adsorption of poly(ethylene vinyl alcohol)
from aqueous solution on montmorillonite clays. Colloid Polym Sci
1989;267:899-906.

156
Page 156 of 302
[78] Ogata N, Jimenez G, Kawai H, Ogihara T. Structure and thermal/mechanical
properties of poly(L-lactide)–clay blend. J Polym Sci Pol Phys 1997;35:389-
396.
[79] Jeon HG, Jung HT, Lee SW, Hudson SD. Morphology of polymer silicate
nanocomposites. High density polyethylene and a nitrile. Polym Bull

t
1998;41:107-113.

ip
[80] Jimenez G, Ogata N, Kawai H, Ogihara T. Structure and thermal/mechanical
properties of poly(ε-caprolactone)- clay blend. J Appl Polym Sci

cr
1997;64:2211-2220.
[81] Srivastava SK, Pramanik M, Acharya H. Ethylene/vinyl acetate

us
copolymer/clay nanocomposites. J Polym Sci Pol Phys 2006;44:471-480.
[82] Tseng C-R, Wu J-Y, Lee H-Y, Chang F-C. Preparation and crystallization

an
behavior of syndiotactic polystyrene–clay nanocomposites. Polymer
2001;42:10063-10070.
[83] Krikorian V, Pochan D. Poly(L-lactide acid)/layered silicate nanocomposite:
M
fabrication, characterization, and properties. Chem Mater 2003;15:4317-4324.
[84] Chang J-H, Uk-An Y, Sur GS. Poly(lactic acid) nanocomposites with various
organoclays. I. thermomechanical properties, morphology, and gas
d

permeability. J Polym Sci Pol Phys 2003;41:94-103.


te

[85] Yano K, Usuki A, Okada A, Kurauchi T, Kamigaito O. Synthesis and


properties of polyimide/clay hybrid. J Polym Sci Pol Chem 1993;31:2493-
p

2498.
ce

[86] Manias E, Chen E, Krishnamoorti R, Genzer J, Kramer EJ, Giannelis EP.


Intercalation kinetics of long polymers in 2 nm confinements. Macromolecules
2000;33:7955-7966.
Ac

[87] Strawhecker KE, Manias E. Structure and properties of poly(vinyl


alcohol)/Na+-montmorillonite nanocomposites. Chem Mater 2000;12:2943-
2949.
[88] Finnigan B, Martin D, Halley P, Truss R, Campell K. Morphology and
properties of thermoplastic polyurethane nanocomposites incorporating
hydrophilic layered silicates. Polymer 2004;45:2249-2260.
[89] Ginsburg VV, Balazs AC. Calculating phase diagrams for nanocomposites:
the effect of adding end-functionalized chains to polymer/clay mixture. Adv
Mater 2000;12:1805-1809.

157
Page 157 of 302
[90] Choi HJ, Kim SG, Hyun YH, Jhon MS. Preparation and rheological
characteristics of solvent-cast poly(ethylene oxide)/montmorillonite
nanocomposites. Macromol Rapid Comm 2001;22:320-325.
[91] Hackett E, Manias E, Giannelis EP. Molecular dynamics simulations of
organically modified layered silicates. J Chem Phys 1998;108:7410-7415.

t
[92] Sur GS, Sun HL, Lyu SG, Mark JE. Synthesis, structure, mechanical

ip
properties, and thermal stability of some polysulfone/organoclay
nanocomposites. Polymer 2001;42:9783-9789.

cr
[93] Usuki A, Kawasumi M, Kojima Y, Okada A, Kurauchi T, Kamigaito O.
Swelling behavior of montmorillonite cation exchanged for ω-amino acid by

us
ε -caprolactam. J Mater Res 1993;8:1174-1178.
[94] Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T, Kamigaito O.

an
Synthesis of nylon-6 hybrid by montmorillonite intercalated with ε -
caprolactam. J Polym Sci Pol Chem 1993;31:983-986.
[95] Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T, Kamigaito O. One-
M
pot synthesis of nylon-6-clay hybrid. J Polym Sci Pol Chem 1993;31:1755-
1758.
[96] Reichert P, Kressler J, Thomann R, Mulhaupt R, Stoppelmann G.
d

Nanocomposites based on a synthetic layer silicate and polyamide-12. Acta


te

Polymerica 1998;49:116-123.
[97] Wu Z, Zhou C, Qi R, Zhang H. Synthesis and characterization of nylon
p

1012/clay nanocomposite. J Appl Polym Sci 2002;83:2403-2410.


ce

[98] Okamoto M, Morita S, Taguchi H, Kim YH, Kotaka T, Tateyama H. Synthesis


and structure of smectic clay/ poly(methyl methacrylate) and clay polystyrene
nanocomposites via in situ intercalative polymerization. Polymer
Ac

2000;41:3887-3890.
[99] Okamoto M, Morita S, Kotaka T. Dispersed structure and ionic conductivity
of smectic clay/polymer nanocomposites. Polymer 2001;42:2685-2688.
[100] Akelah A, Moet M. Polymer–clay nanocomposites: free radical grafting of
polystyrene on to organophilic montmorillonite interlayers. J Mater Sci
1996;31:3589-3596.
[101] Hsu SLC, Chang KC. Synthesis and properties of polybenzoxazole– clay
nanocomposites. Polymer 2002;43:4097-4101.

158
Page 158 of 302
[102] Sun T, Garces JM. High-performance polypropylene–clay nanocomposites by
in-situ polymerization with metallocene/ clay catalysts. Adv Mater
2002;14:128-130.
[103] Bergman JS, Chen H, Giannelis EP, Thomas MG, Coates GW. Synthesis and
characterization of polyolefin-silicate nanocomposites: a catalyst intercalation

t
and in situ polymerization approach. J Chem Soc Chem Commun

ip
1999;21:2179-2180.
[104] Tudor J, Willington L, O’Hare D, Royan B. Intercalation of catalytically

cr
active metal complexes in phyllosilicates and their application as propene
polymerization catalyst. Chem Commun 1996;17:2031-2032.

us
[105] Jin Y-H, Park H-J, Im S-S, Kwak S-Y, Kwak S. Polyethylene/clay
nanocomposite by in situ exfoliation of montmorillonite during Ziegler–Natta

an
polymerization of ethylene. Macromol Rapid Comm 2002;23:135-140.
[106] Ke YC, Long CF, Qi ZN. Crystallization, properties, and crystal and nanoscale
morphology of PET-clay nanocomposites. J Appl Polym Sci 1999;71:1139-
M
1146.
[107] Lee S-S, Ma YT, Rhee H-W, Kim J. Exfoliation of layered silicate by ring
opening reaction of cyclic oligomers in PET-clay nanocomposites. Polymer
d

2005;46:2201-2210.
te

[108] Alexandre M, Dubois P, Jerome R, Garcia-Marti M, Sun T, Garces JM, Millar


DM, Kuperman A. Polyolefin nanocomposites. WO:947598A1, 1999.
p

[109] Alexandre M, Dubois P, Garces JM, Sun T, Jerome R. Polyethylene-layered


ce

silicate nanocomposites prepared by the polymerization-filling technique:


synthesis and mechanical properties. Polymer 2002;43:2123-2132.
[110] Dubois P, Alexandre M, Hindryckx F, Jerome R. Homogeneous polyolefin-
Ac

based composites. J Macromol Sci Rev Macromol Chem Phys 1998;C38:511-


565.
[111] Shin S-YA, Simon LC, Soares JBP, Scholz G. Polyethylene-clay hybrid
nanocomposites: in situ polymerization using bifunctional organic modifiers.
Polymer 2003;44:5317-5321.
[112] Heinemann J, Reichert P, Thomann R, Mulhaupt R. Polyolefin
nanocomposites formed by melt compounding and transition metal catalyzed
ethene homo- and copolymerization in the presence of layered silicates.
Macromol Rapid Comm 1999;20:423-430.

159
Page 159 of 302
[113] Doh JG, Cho I. Synthesis and properties of polystyrene-organoammonium
montmorillonite hybrid. Polym Bull 1998;41:511-518.
[114] Weimer MW, Chen H, Giannelis EP, Sogah DY. Direct synthesis of dispersed
nanocomposites by in situ living free radical polymerization using a silicate-
anchored initiator. J Am Chem Soc 1999;121:1615-1616.

t
[115] Yei D-R, Kuo S-W, Fu H-K, Chang F-C. Enhanced thermal properties of PS

ip
nanocomposites formed from montmorillonite treated with a
surfactant/cyclodextrin inclusion complex. Polymer 2005;46:741-750.

cr
[116] Paul MA, Alexandre M, Degee P, Calberg C, Jerome R, Dubois P. Exfoliated
polylactide/clay nanocomposites by in situ coordination–insertion

us
polymerization. Macromol Rapid Comm 2003;24:561-566.
[117] Messersmith PB, Giannelis EP. Polymer-layered silicate nanocomposites: in-

an
situ intercalative polymerization of ε -caprolactone in layered silicates. Chem
Mater 1993;5:1064-1066.
[118] Messersmith PB, Giannelis EP. Synthesis and barrier properties of poly(ε-
M
caprolactone)-layered silicate nanocomposites. J Polym Sci Pol Chem
1995;33:1047-1057.
[119] Pantoustier N, Alexandre M, Degee P, Calberg C, Jerome R, Henrist C, Cloots
d

R, Rulmont A, Dubois P. Poly(ε -caprolactone) layered silicate


te

nanocomposites: effect of clay surface modifiers on the melt intercalation


process. e-Polymer 2001;9:1-9.
p

[120] Pantoustier N, Lepoittevin B, Alexandre M, Kubies D, Calberg C, Jerome R,


ce

Dubois P. Biodegradable polyester layered silicate nanocomposites based on


poly(ε -caprolactone). Polym Eng Sci 2002;42:1928-1937.
[121] Rozenberg BA. Kinetics and mechanism of ε-caprolactone anionic
Ac

polymerization under the influence of amines. Pure Appl Chem 1981;53:1715-


1722.
[122] Mecerreyes D, Jerome R, Dubois P. Novel macromolecular architectures
based on aliphatic polyesters: relevance of the coordination-insertion ring
opening polymerization. Adv Polym Sci 1999;147:1-59.
[123] Lofgren A, Albertsson A-C, Dubois P, Jerome R. Recent advances in ring-
opening polymerization of lactones and related compounds. J Macromol Sci
Rev Macromol Chem Phys C 1995;35:379-418.

160
Page 160 of 302
[124] Sepehr M, Utracki LA, Zheng X, Wilkie CA. Polystyrenes with macro-
intercalated organoclay. Part I. Compounding and characterization. Polymer
2005;46:11557-11568.
[125] Lan T, Pinnavaia TJ. Clay-reinforced epoxy nanocomposites. Chem Mater
1994;6:2216-2219.

t
[126] Becker O, Varley R, Simon G. Morphology, thermal relaxations and

ip
mechanical properties of layered silicate nanocomposites based upon high-
functionality epoxy resins. Polymer 2002;43:4365-4373.

cr
[127] Kornmann X, Berglund LA, Sterte J. Nanocomposite based on
montmorillonite and unsaturated polyester. Polym Eng Sci 1998;38:1351-

us
1358.
[128] Jiankun L, Yucai K, Zongneng Q, Xiao-Su Y. Study on intercalation and

an
exfoliation behavior of organoclays in epoxy resin. J Polym Sci Pol Phys
2001;39:115-120.
[129] Lan T, Kaviratna PD, Pinnavaia TJ. Synthesis, characterization and
M
mechanical properties of epoxy–clay nanocomposites. Polym Mater Sci Eng
1994;71:527-528.
[130] Liu W, Hoa SV, Pugh M. Fracture toughness and water uptake of high-
d

performance epoxy/nanoclays nanocomposites. Compos Sci Technol


te

2005;65:2364-2373.
[131] Lan T, Kaviratna PD, Pinnavaia TJ. Mechanism of clay tactoid exfoliation in
p

epoxy-clay nanocomposites. Chem Mater 1995;7:2144-2150.


ce

[132] Zilg C, Mulhaupt R, Finter J. Morphology and toughness/ stiffness balance of


nanocomposites based upon anhydride cured epoxy resins and layered
silicates. Macromol Chem Phys 1999;200:661-670.
Ac

[133] Becker O, Varley RJ, Simon GP. Thermal stability and water uptake of high
performance epoxy layered silicate nanocomposites. Eur Polym J
2004;40:187-195.
[134] Hackman I, Hollaway L. Epoxy-layered silicate nanocomposites in civil
engineering. Compos Part A-Appl S 2006;37:1161-1170.
[135] Messersmith PB, Giannelis EP. Synthesis and characterization of layered
silicate-epoxy nanocomposites. Chem Mater 1994;6:1719-1725.

161
Page 161 of 302
[136] LePluart L, Duchet J, Sautereau H. Epoxy/montmorillonite nanocomposites:
influence of organophilic treatment on reactivity, morphology and fracture
properties. Polymer 2005;46:12267-12278.
[137] Koerner H, Misra D, Tan A, Drummy L, Mirau P, Vaia R. Montmorillonite-
thermoset nanocomposites via cryo-compounding. Polymer 2006;47:3426-

t
3435.

ip
[138] Wang K, Chen L, Kotaki M, He C. Preparation, microstructure and thermal
mechanical properties of epoxy/crude clay nanocomposites. Compos Part A-

cr
Appl S 2007;38:192-197.
[139] Suh D, Lim Y, Park O. The property and formation mechanism of unsaturated

us
polyester-layered silicate nanocomposite depending on the fabrication
methods. Polymer 2000;41:8557-8563.

an
[140] Bharadwaj RK, Mehrabi AR, Hamilton C, Trujillo C, Murga MF, Chavira A.
Structure– property relationships in cross-linked polyester–clay
nanocomposites. Polymer 2002;43:3699-3705.
M
[141] Wang Z, Pinnavaia TJ. Nanolayer Reinforcement of Elastomeric
Polyurethane. Chem Mater 1998;10:3769-3771.
[142] Yao KJ, Song M, Hourston DJ, Luo DZ. Polymer/layered clay
d

nanocomposites: 2 polyurethane nanocomposites. Polymer 2002;43:1017-


te

1020.
[143] Zilg C, Thomann R, Mulhaupt R, Finter J. Polyurethane nanocomposites
p

containing laminated anisotropic nanoparticles derived from organophilic


ce

layered silicates. Adv Mater 1999;11:49-52.


[144] Berta M, Lindsay C, Pans G, Camino G. Effect of chemical structure on
combustion and thermal behaviour of polyurethane elastomer layered silicate
Ac

nanocomposites. Polym Degrad Stabil 2006;91:1179-1191.


[145] Gao X, Lee J, Widya T, Macosko C. Polyurethane/clay nanocomposites
foams: processing, structure and properties. Polymer 2005;46:775-783.
[146] Liu X, Wu Q. Polyamide 66/clay nanocomposites via melt intercalation.
Macromol Mater Eng 2002;287:180-186.
[147] Phang IY, Liu T, Mohamed A, Pramoda KP, Chen L, Shen L, Chow SY, He
C, Lu X, Hu X. Morphology, thermal and mechanical properties of nylon
12/organoclay nanocomposites prepared by melt compounding. Polym Int
2005;54:456-464.

162
Page 162 of 302
[148] Davis CH, Mathias LJ, Gilman JW, Schiraldi DA, Shields JR, Trulove P,
Sutto TE, Delong HC. Effects of melt-processing conditions on the quality of
poly(ethylene terephthalate) montmorillonite clay nanocomposites. J Polym
Sci Pol Phys 2002;40:2661-2666.
[149] Pegoretti A, Kolarik J, Peronic C, Migliaresi C. Recycled poly(ethylene

t
terephthalate) layered silicate nanocomposites: morphology and tensile

ip
mechanical properties. Polymer 2004;45:2751-2759.
[150] Tang Y, Hu Y, Wang J, Zong R, Gui Z, Chen Z, Zhuang Y, Fan W. Influence

cr
of organophilic clay and preparation methods on EVA/montmorillonite
nanocomposites. J Appl Polym Sci 2004;91:2416-2421.

us
[151] Gopakumar TG, Lee JA, Kontopoulou M, Parent JS. Influence of clay
exfoliation on the physical properties of montmorillonite/ polyethylene

an
composites. Polymer 2002;43:5483-5491.
[152] Thellen C, Orroth C, Froio D, Ziegler D, Lucciarini J, Farrell R, D’ Souza NA,
Ratto JA. Influence of montmorillonite layered silicate on plasticized poly(L-
M
lactide) blown films. Polymer 2005;46:11716-11727.
[153] Ray SS, Maiti P, Okamoto M, Yamada K, Ueda K. New polylactide/layered
silicate nanocomposites. 1. Preparation, characterization and properties.
d

Macromolecules 2002;35:3104-3110.
te

[154] Pluta M. Morphology and properties of polylactide modified by thermal


treatment, filling with layered silicates and plasticization. Polymer
p

2004;45:8239-8251.
ce

[155] Di Y, Iannace S, Maio ED, Nicolais L. Nanocomposites by melt intercalation


based on polycaprolactone and organoclay. J Polym Sci Pol Phys
2003;41:670-678.
Ac

[156] Gorrasi G, Tortora M, Vittoria V, Pollet E, Lepoittevin B, Alexandre M,


Dubois P. Vapor barrier properties of polycaprolactone montmorillonite
nanocomposites: effect of clay dispersion. Polymer 2003;44:2271-2279.
[157] Wolf D, Fuchs A, Wagenknecht U, Kretzschmar B, Jehnichen D, Haussler L.
Nanocomposites of polyolefin clay hybrids. in: Proceedings of Eurofiller '99,
Lyon-Villeurbanne, France, 1999, p. 6-9.
[158] Liu X, Wu Q. PP/clay nanocomposites prepared by grafting melt intercalation.
Polymer 2001;42:10013-10019.

163
Page 163 of 302
[159] Zheng X, Jiang DD, Wilkie CA. Polystyrene nanocomposites based on an
oligomerically-modified clay containing maleic anhydride. Polym Degrad
Stabil 2006;91:108-113.
[160] Hasegawa N, Okamoto H, Kato M, Usuki A, Sato N. Nylon 6–
montmorillonite nanocomposites prepared by compounding nylon 6 with Na-

t
montmorillonite slurry. Polymer 2003;44:2933-2937.

ip
[161] Garcia-Lopez D, Picazo O, Merino JC, Pastor JM. Polypropylene-clay
nanocomposites: effect of compatibilizing agents on clay dispersion. Eur

cr
Polym J 2003;39:945-950.
[162] Hotta S, Paul DR. Nanocomposites formed from linear low density

us
polyethylene and organoclays. Polymer 2004;45:7639-7654.
[163] Duquesne S, Jama C, Le Bras M, Delobel R, Recourt P, Gloaguen JM.

an
Elaboration of EVA-nanoclay systems – characterization, thermal behavior
and fire performance. Compos Sci Technol 2003;63:1141-1148.
[164] Zhang W, Chen D, Zhao Q, Fang Y. Effects of different kinds of clay and
M
different vinyl acetate content on the morphology and properties of EVA/clay
nanocomposites. Polymer 2003;44:7953-7961.
[165] Zanetti M, Camino G, Thomann R, Mulhaupt R. Synthesis and thermal
d

behavior of layered silicate-EVA nanocomposites. Polymer 2001;42:4501-


te

4507.
[166] Alexandre M, Beyer G, Henrist C, Cloots R, Rulmont A, Jérôme R, Dubois P.
p

Preparation and properties of layered silicate nanocomposites based on


ce

ethylene vinyl acetate copolymers. Macromol Rapid Comm 2001;22:643-646.


[167] Peeterbroeck S, Alexandre M, Jerome R, Dubois Ph. Poly(ethylene-co-vinyl
acetate)/clay nanocomposites: Effect of clay nature and organic modifiers on
Ac

morphology, mechanical and thermal properties. Polym Degrad Stabil


2005;90:288-294.
[168] Balazs AC, Singh C, Zhulina. Modeling the interactions between polymers
and clay surfaces through self-consistent field theory. Macromolecules
1998;31:8370-8381.
[169] Kurian M, Dasgupta A, Beyer FL, Galvin ME. Investigation of the effects of
silicate modification on polymer-layered silicate nanocomposite morphology.
J Polym Sci Pol Phys 2004;42:4075-4083.

164
Page 164 of 302
[170] Liu LM, Qi ZN, Zhu XG. Studies on nylon-6 clay nanocomposites by melt-
intercalation process. J Appl Polym Sci 1999;71:1133-1138.
[171] Chaudhary DS, Prasad R, Gupta RK, Bhattacharya SN. Clay intercalation and
influence on crystallinity of EVA-based clay nanocomposites. Thermochim
Acta 2005;433:187-195.

t
[172] Li X, Ha CS. Nanostructure of EVA/organoclay nanocomposites: Effects of

ip
kinds of organoclays and grafting of maleic anhydride onto EVA. J Appl
Polym Sci 2003;87:1901-1909.

cr
[173] Strong AB. Plastics-Materials and Processing. Prentice Hall, Inc., 2nd edition,
2000.

us
[174] Li J, Zhao L, Guo S. Ultrasonic preparation of polymer/layered silicate
nanocomposites during extrusion. Polym Bull 2005;55:217-223.

an
[175] Maiti P, Nam PH, Okamoto M, Kotaka T, Hasegawa N, Usuki A. Influence of
crystallization on intercalation, morphology, and mechanical properties of
propylene/clay nanocomposites. Macromolecules 2002;35:2042-2049.
M
[176] Maiti P, Nam PH, Okamoto M, Kotaka T, Hasegawa N, Usuki A. The effect
of crystallization on the structure and morphology of polypropylene/clay
nanocomposites. Polym Eng Sci 2002;42:1864-1871.
d

[177] Zanetti M, Costa L. Preparation and combustion behavior of polymer/layered


te

silicate nanocomposites based upon PE and EVA. Polymer 2004;45:4367-


4373.
p

[178] Preston CML, Amarasinghe G, Hopewell JL, Shanks RA, Mathys Z.


ce

Evaluation of polar ethylene copolymers as fire retardant nanocomposite


matrices. Polym Degrad Stabil 2004;84:533-544.
[179] Zhai H, Xu W, Guo H, Zhou Z, Shen S, Song Q. Preparation and
Ac

characterization of PE and PE-g-MAH/montmorillonite nanocomposites. Eur


Polym J 2004;40:2539-2545.
[180] Lu H, Hu Y, Xiao J, Kong Q, Chen Z, Fan W. The influence of irradiation on
morphology evolution and flammability properties of maleated
polyethylene/clay nanocomposite. Mater Lett 2005;59(6):648-651.
[181] Morawiec J, Pawlak A, Slouf M, Galeski A, Piorkowska E, Kransnikowa N.
Preparation and properties of compatibilized LDPE/organo-modified
montmorillonite nanocomposites. Eur Polym J 2005;41:1115-1122.

165
Page 165 of 302
[182] Zhang J, Wilkie CA. Preparation and flammability properties of polyethylene-
clay nanocomposites. Polym Degrad Stabil 2003;80:163-169.
[183] Tang Y, Hu Y, Li B, Liu L, Wang Z, Chen Z, Fan W. Polypropylene /
montmorillonite nanocomposites and intumescent, flame-retardant
montmorillonite synergism in polypropylene nanocomposites. J Polym Sci Pol

t
Chem 2004;42:6163-6173.

ip
[184] Zhao C, Qin H, Gong F, Feng M, Zhang S, Yang M. Mechanical, thermal and
flammability properties of polyethylene/clay nanocomposites. Polym Degrad

cr
Stabil 2005;87:183-189.
[185] Hong HC, Lee YB, Bae JW, Jho YJ, Nam BU, Nam GJ, Lee KJ. Tensile and

us
flammability properties of polypropylene-based RTPO/clay nanocomposites
for cable insulating material. J Appl Polym Sci 2005;97:2375-2381.

an
[186] Hasegawa N, Kawasumi M, Kato M, Usuki A, Okada A. Preparation and
mechanical properties of polypropylene– clay hybrids using a maleic
anhydride-modified polypropylene oligomer. J Appl Polym Sci 1998;67:87-
M
92.
[187] Hasegawa N, Okamoto H, Kawasumi M, Kato M, Tsukigase A, Usuki A.
Polyolefin–clay hybrids based on modified polyolefins and organoclay.
d

Macromol Mater Eng 2000;280/281:76–79.


te

[188] Kato M, Usuki A, Okada A. Synthesis of polypropylene oligomer-clay


intercalation compounds. J Appl Polym Sci 1997;66:1781-1785.
p

[189] Nam PH, Maiti P, Okamoto M, Kotaka T, Hasegawa N, Usuki A. A


ce

hierarchical structure and properties of intercalated polypropylene/clay


nanocomposites. Polymer 2001;42:9633-9640.
[190] Wang D, Wilkie CA. In-situ reactive blending to prepare polystyrene-clay and
Ac

polypropylene-clay nanocomposites. Polym Degrad Stabil 2003;80:171-182.


[191] Hasegawa N, Okamoto H, Kawasumi M, Usuki A. Preparation and
mechanical properties of polystyrene-clay hybrids. J Appl Polym Sci
1999;74:3359-3364.
[192] Davis RD, Gilman JW, VanderHart DL. Processing degradation of polyamide
6/ montmorillonite clay nanocomposites and clay organic modifier. Polym
Degrad Stabil 2003;79:111-121.

166
Page 166 of 302
[193] Chang J-H, Kim SJ, Joo YL, Im S. Poly(ethylene terephthalate)
nanocomposites by in situ interlayer polymerization: the thermo-mechanical
properties and morphology of the hybrid fiber. Polymer 2004;45:919-926.
[194] Gilman JW, Awad WH, Davis RD, Shields J, Harris Jr RH, Davis C, Morgan
AB, Sutto TE, Callahan J, Trulove PC, DeLong HC. Polymer/layered silicate

t
nanocomposites from thermally stable trialkylimidazolium-treated

ip
montmorillonite. Chem Mater 2002;14:3776-3785.
[195] Maiti P, Batt CA, Giannelis EP. Renewable plastics: synthesis and properties

cr
of PHB nanocomposites. J Macromol Sci Rev 2003;88:58-59.
[196] Vaia RA, Vasudevan S, Krawiec W, Scanlon LG, Giannelis EP. New polymer

us
electrolyte nanocomposites: melt intercalation of poly(ethylene oxide) in
mica-type silicates. Adv Mater 1995;7:154-156.

an
[197] Kawasumi M, Hasegawa N, Kato M, Usuki A, Okada A. Preparation and
mechanical properties of polypropylene–clay hybrids. Macromolecules
1997;30:6333-6338.
M
[198] Shia D, Hui CY, Burnside SD, Giannelis EP. An interface model for the
prediction of Young’s modulus of layered silicate–elastomer nanocomposites.
Polym Composite 1998;19:608-617.
d

[199] Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T, Kamigaito O.


te

Sorption of water in nylon 6–clay hybrid. J Appl Polym Sci 1993;49:1259-


1264.
p

[200] Goettler LA. Overview of Property Development in Layered Silicate Polymer


ce

Nanocomposites. Ann Tech Confr Soc Plast Eng 2005;1980-1982.


[201] Tortora M, Vittoria V, Galli G, Ritrovati S, Chiellini E. Transport Properties
of Modified Montmorillonite-Poly( -caprolactone) Nanocomposites.
Ac

Macromol Mater Eng 2002;287:243-249.


[202] Xiong J, Liu Y, Yang X, Wang X. Thermal and mechanical properties of
polyurethane/montmorillonite nanocomposites based on a novel reactive
modifier. Polym Degrad Stabil 2004;86:549-555.
[203] Kojima Y, Usuki A, Kawasumi M, Okada A, Fukushima Y, Karauchi T,
Kamigaito O. Mechanical properties of nylon-6/clay hybrid. J Mater Res
1993;6:1185-1189.

167
Page 167 of 302
[204] Tortora M, Gorrasi G, Vittoria V, Galli G, Ritrovati S, Chiellini E. Structural
characterization and transport properties of organically modified
montmorillonite/ polyurethane nanocomposites. Polymer 2002;43:6147-6157.
[205] Noh MW, Lee DC. Synthesis and characterization of PS-clay nanocomposite
by emulsion polymerization. Polym Bull 1999;42:619-626.

t
[206] Ray SS, Yamada K, Okamoto M, Ogami A, Ueda K. New polylactide/layered

ip
silicate nanocomposites. 3. High performance biodegradable materials. Chem
Mater 2003;15:1456-1465.

cr
[207] Ray SS, Okamoto M. Biodegradable polylactide and its nanocomposites:
opening a new dimension for plastics and composites. Macromol Rapid

us
Comm 2003;24:815-840.
[208] Ray SS, Yamada K, Okamoto M, Ueda K. New polylactide/layered silicate

an
nanocomposites. 2. Concurrent improvements of material properties,
biodegradability and melt rheology. Polymer 2003;44:857-866.
[209] Pramanik M, Srivastava SK, Samantaray BK, Bhowmick AK. EVA/clay
M
nanocomposite by solution blending: effect of aluminosilicate layers on
mechanical and thermal properties. Macromol Res 2003;11:260-266.
[210] Chen B, Evans JRG. Poly(ε -caprolactone)-clay nanocomposites: structure and
d

mechanical properties. Macromolecules 2006;39:747-754.


te

[211] Laus M, Francesangeli O, Sandrolini F. New hybrid nanocomposites based on


an organophilic clay and poly(styrene-b-butadiene) copolymers. J Mater Res
p

1997;12:3134-3139.
ce

[212] Okamoto M, Nam PH, Maiti P, Kotaka T, Nakayama T, Takada M, Ohshima


M, Usuki A, Hasegawa N, Okamoto H. Biaxial Flow-Induced Alignment of
Silicate Layers in Polypropylene/Clay Nanocomposite Foam. Nano Lett
Ac

2001;1(9):503-505.
[213] Nielsen LE. Mechanical properties of polymer and composites, vol. 2. New
York: Marcel Dekker, 1981.
[214] Fredrickson GH, Bicerano J. Barrier properties of oriented disk composites. J
Chem Phys 1999;110:2181-2188.
[215] Burnside SD, Giannelis EP. Synthesis and properties of new
poly(dimethylsiloxane) nanocomposites. Chem Mater 1995;7:1597-1600.
[216] Lange J, Wyser Y. Recent innovations in barrier technologies for plastic
packaging – a review. Packag Technol Sci 2003;16:149-158.

168
Page 168 of 302
[217] Nielsen LE. Models for the permeability of filled polymer systems. J
Macromol Sci 1967;A1:929-942.
[218] Cussler EL, Hughes SE, Ward WJ, Aris R. Barrier membranes. J Membrane
Sci 1988;38:161-174.
[219] Petricova E, Knight R, Schadler LS, Twardowski TE. Nylon 11/silica

t
nanocomposite coatings applied by the HVOF process. II. Mechanical and

ip
barrier properties. J Appl Polym Sci 2000;78:2272-2289.
[220] Fukuda M, Kuwajima S. Molecular-dynamics simulation of moisture diffusion

cr
in polyethylene beyond 10 ns duration. J Chem Phys 1997;107:2149-2159.
[221] Drozdov AD, Christiansen J de C, Gupta RK, Shah AP. Model for anomalous

us
moisture diffusion through a polymer-clay nanocomposite. J Polym Sci Pol
Phys 2003;41:476-492.

an
[222] Ke Z, Yongping B. Improve gas barrier property of PET film with
montmorillonite by in situ interlayer polymerization. Mater Lett
2005;59:3348-3351.
M
[223] Ogasawara T, Ishida Y, Ishikawa T, Aoki T, Ogura T. Helium gas
permeability of montmorillonite/epoxy nanocomposites. Compos Part A-Appl
S 2006;37:2236-2240.
d

[224] Zhu J, Uhl FM, Morgan AB, Wilkie CA. Studies on the mechanism by which
te

the formation of nanocomposites enhances thermal stability. Chem Mater


2001;13:4649-4654.
p

[225] Vyazovkin S, Dranka I, Fan X, Advincula R. Kinetics of the thermal and


ce

thermo-oxidative degradation of a polystyrene-clay nanocomposite. Macromol


Rapid Comm 2004;25:498-503.
[226] Zanetti M, Bracco P, Costa L. Thermal degradation behavior of PE/clay
Ac

nanocomposites. Polym Degrad Stabil 2004;85:657-665.


[227] Bandyopadhyay S, Chen R, Giannelis EP. Biodegradable organic-inorganic
hybrids based on poly(L-lactide). J Macromol Sci Rev 1999;81:159-160
[228] Lepoittevin B, Pantoustier N, Devalckenaere M, Alexandre M, Kubies D,
Calderg C. Poly(ε -caprolactone)/clay nanocomposites by in situ intercalative
polymerization catalyzed by dibutyltindimethoxide. Macromolecules
2002;35:8385-8390.

169
Page 169 of 302
[229] Dabrowski F, Bourbigot S, Delobel R, LeBras M. Kinetic modeling of the
thermal degradation of polyamide 6 nanocomposite. Eur Polym J
2000;36:273-284.
[230] Jang BN, Wilkie CA. The effect of clay on the thermal degradation of
polyamide 6 in polyamide 6/clay nanocomposites. Polymer 2005;46:3264-

t
3274.

ip
[231] Pandey JK, Reddy KR, Kumar AP, Singh RP. An overview of the
degradability of polymer nanocomposites. Polym Degrad Stabil 2005;88:234-

cr
250.
[232] Dutta S, Bhowmick AK, Mukunda PG, Chaki TK. Thermal degradation

us
studies of electron beam cured ethylene-vinyl acetate copolymer. Polym
Degrad Stabil 1995;50:75-82.

an
[233] Paul M-A, Alexandre M, Degee P, Henrist C, Rulmont A, Dubois P. New
nanocomposite materials based on plasticized poly(L-lactide) and organo-
modified montmorillonites: thermal and morphological study. Polymer
M
2003;44:443-450.
[234] Riva A, Zanetti M, Braglia M, Camino G, Falqui L. Thermal degradation and
rheological behaviour of EVA/montmorillonite nanocomposites. Polym
d

Degrad Stabil 2002;77:299-304.


te

[235] Maurin MB, Dittert LW, Hussain AA. Thermogravimetric analysis of


ethylene-vinyl acetate copolymers with Fourier transform infrared analysis of
p

the pyrolysis products. Thermochim Acta 1991;186:97-102.


ce

[236] McGrattan BJ. Examining the decomposition of ethylene-vinyl acetate


copolymers using TG/GC/IR. Appl Spectrosc 1994;48:1472-1476.
[237] Costache MC, Jiang DD, Wilkie CA. Thermal degradation of ethylene-vinyl
Ac

acetate copolymer nanocomposites. Polymer 2005;63:6947-6958.


[238] Beyer G. Flame retardant properties of EVA-nanocomposites and
improvements by combination of nanofillers with aluminium trihydrate. Fire
Mater 2002;25:193-197.
[239] Morgan AB. Flame retarded polymer layered silicate nanocomposites: a
review of commercial and open literature systems. Polym Advan Technol
2006;17:206-217.
[240] Gilman JW, Jackson CL, Morgan AB, Harris Jr R, Manias E, Giannelis EP,
Wuthenow M, Hilton D, Phillips SH. Flammability properties of polymer-

170
Page 170 of 302
layered silicate nanocomposites. Propylene and polystyrene nanocomposites.
Chem Mater 2000;12:1866-1873.
[241] Gilman JW, Kashiwagi TCL, Giannelis EP, Manias E, Lomakin S, Lichtenhan
JD, Jones P. Radiative gasification and vinyl polymer flammability. In: Le
Bras M, Camino G, Bourbigot S, Delobel R, eds. Fire Retardancy of

t
Polymers. Cambridge: The Royal Society of Chemistry 1998, p. 203-221.

ip
[242] Kashiwagi T, Harris RH Jr, Zhang X, Briber RM, Cipriano BH, Ragharan SR,
Awad WH, Shields JR. Flame retardant mechanism of polyamide 6-clay

cr
nanocomposites. Polymer 2004;45:881-891.
[243] Lee J, Giannelis EP. Synthesis and characterization of unsaturated polyester

us
and phenolic resin nanocomposites. Polymer Preprints 1997;38:688-689.
[244] Chigwada G, Jash P, Jiang DD, Wilkie CA. Synergy between nanocomposite

an
formation and low levels of bromine on fire retardancy in polystyrenes. Polym
Degrad Stabil 2005;88:382-393.
[245] Chigwada G, Jash P, Jiang DD, Wilkie CA. Fire retardancy of vinyl ester
M
nanocomposites: Synergy with phosphorus-based fire retardants. Polym
Degrad Stabil 2005;89:85-100.
[246] Si M, Zaitsev V, Goldman M, Frenkel A, Peiffer DG, Weil E, Sokolov JC,
d

Rafailovich MH. Self-extinguishing polymer/organoclay nanocomposites.


te

Polym Degrad Stabil 2007;92:86-93.


[247] Zanetti M, Camino G, Canavese D, Morgan AB, Lamelas FJ, Wilkie CA. Fire
p

retardant halogen-antimony-clay synergism in polypropylene layered silicate


ce

nanocomposites. Chem Mater 2002;14:189-193.


[248] Tang Y, Hu Y, Wang S, Gui Z, Chen Z, Fan W. Intumescent flame retardant-
montmorillonite synergism in polypropylene-layered silicate nanocomposites.
Ac

Polym Int 2003;52:1396-1400.


[249] Dabrowski F, Le Bras M, Cartier L, Bourbigot S. The use of clay in an EVA-
based intumescent formulation. Comparison with the intumescent formulation
using polyamide-6 clay nanocomposite as carbonisation agent. J Fire Sci
2001;19:219-241.
[250] Zheng X, Wilkie CA. Flame retardancy of polystyrene nanocomposites based
on an oligomeric organically-modified clay containing phosphate. Polym
Degrad Stabil 2003;81:539-550.

171
Page 171 of 302
[251] Kim J, Lee K, Lee K, Bae J, Yang J, Hong S. Studies on the thermal
stabilization enhancement of ABS; synergistic effect of triphenyl phosphate
nanocomposite, epoxy resin, and silane coupling agent mixtures. Polym
Degrad Stabil 2003;79:201-207.
[252] Hussain M, Varley RJ, Mathys Z, Cheng YB, Simon GP. Effect of organo-

t
phosphorus and nano-clay materials on the thermal and fire performance of

ip
epoxy resins. J Appl Polym Sci 2004;91:1233-1253.
[253] Nam PH, Maiti P, Okamoto M, Kotaka T, Nakayama T, Takada M, Ohshima

cr
M, Usuki A, Hasegawa N, Okamoto H. Foam processing and cellular structure
of polypropylene/clay nanocomposites. Polym Eng Sci 2001;42:1907-1918.

us
[254] Karian HG. Handbook of Polypropylene and Polypropylene Composites. New
York: Marcel Dekker, 1999.

an
[255] Di Maio E, Iannace S, Sorrentini L, Nicolais L. Isothermal crystallization in
PCL/clay nanocomposites investigated with thermal and rheometric methods.
Polymer 2004;45:8893-8900.
M
[256] Gelfer MY, Burger C, Chu B, Hsiao BS, Drozdov AD, Si M, Rafailovich M,
Sauer B, Gilman JW. Relationships between structure and rheology in model
nanocomposites of ethylene-vinyl-based copolymers and organoclays.
d

Macromolecules 2005;38:3765-3775.
te

[257] Maiti P, Okamoto M. Crystallization controlled by silicate surfaces in nylon


6–clay nanocomposites. Macromol Mater Eng 2003;288:440-445.
p

[258] Tetto JA, Steeves DM, Welsh EA, Powell BE. Biodegradable poly(ε -
ce

caprolactone)/clay nanocomposites. Ann Tech Confr Soc Plast Eng


1999;1628–1632.
[259] Lee SR, Park HM, Lim H, Kang T, Li X, Cho WJ, Chang S. Microstructure,
Ac

tensile properties, and biodegradability of aliphatic polyester/clay


nanocomposites. Polymer 2002;43:2495-2500.
[260] Ray SS, Yamada K, Okamoto M, Fujimoto Y, Ogami A, Ueda K. New
polylactide/layered silicate nanocomposites: 5. Designing of materials with
desired properties. Polymer 2003;44:6633-6646.
[261] Huaili Q, Chungui Z, Shimin Z, Guangming C, Mingshu Y. Photo-oxidative
degradation of polyethylene / montmorillonite nanocomposite. Polym Degrad
Stabil 2003;81:497-500.

172
Page 172 of 302
[262] Morlat S, Mailhot B, Gonzalez D, Gardett J. Photo-oxidation of
polypropylene/montmorillonite nanocomposites. Chem Mater 2004;16:377-
383.
[263] Mailhot B, Morlat S, Gardett J, Boucard S, Duchet J, Gerard J.
Photodegradation of polypropylene nanocomposites. Polym Degrad Stabil

t
2003;82:163-167.

ip
[264] Patterson PH, Sloan JM, Hsieh AJ. Photo degradation mechanisms of layered
silicate/polycarbonate nanocomposites. Ann Tech Confr Soc Plast Eng 2002,

cr
p. 3936-3938
[265] Sloan JM, Patterson P, Hsieh A. Mechanisms of photo degradation for layered

us
silicate-polycarbonate nanocomposites. Polym Mater Sci Eng 2003;88:354-
355.

an
[266] Schmidt G, Malwitz MM. Properties of polymer-nanoparticle composites.
Curr Opin Colloid Interface Sci 2003;8:103-108.
[267] Schmidt D, Shah D, Giannelis EP. New advances in polymer/layered silicate
M
nanocomposites. Curr Opin Solid State Matr 2002;6:205-212.
[268] Kim JH, Koo CM, Choi YS, Wang KH, Chung IJ. Preparation and
characterization of polypropylene/layered silicate nanocomposites using an
d

antioxidant. Polymer 2004;45:7719-7727.


te

[269] Gacitua WE, Ballerini AA, Zhang J. Polymer nanocomposites: synthetic and
natural fillers. A Review. Maderas Ciencia y tecnología 2005;7:159-178.
p

[270] Zeng QH, Yu AB, Lu GQ, Paul DR. Clay-based polymer nanocomposites:
ce

research and commercial development. J Nanosci Nanotechno 2005;5:1574-


1592.
[271] Leaversuch R. Nanocomposites broaden roles in automotive, barrier
Ac

packaging. http://www.ptonline.com/articles/200110fa3.html, 2001.


[272] Rana HT, Gupta RK, GangaRao HVS, Sridhar LN. Measurement of moisture
diffusivity through layered-silicate nanocomposites. Mater Interfaces
Electrochemical Phenomena 2005;51:3249-3256.

173
Page 173 of 302
TABLE 1. Chemical structure of commonly used 2:1 phyllosilicatesa [1]
2:1 Phyllosilicates General formula
Montmorillonite Mx(Al4-xMgx)Si8O20(OH)4
Hectorite Mx(Mg6-xLix)Si8O20(OH)4
Saponite MxMg6(Si8-xAlx)O20(OH)4
a M=monovalent cation; x=degree of isomorphous substitution (between 0.5 and 1,3).

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 174 of 302


TABLE 2. PLS nanocomposites prepared by intercalation from solution
Nanocomposite Solvent(s) Ref.
PVOH/ Na+-MMT water [86]
PVA/ Na+-MMT water [87]
TPU/OMLS Toluene/DMAc [88]
PEO/ Na+-MMT or Na+-hectorite acetonitrile [89]
PEO/MMT chloroform [90]

t
ip
PLA/OMLS dichloromethane [83]
PLA/OMLS DMAc [84]
HDPE/ Protonated dodecylamine modified MMT Xylene/benzonitrile (80/20 wt %) [91]

cr
PSF/OMLS DMAC [92]
PI/ Dodecylammonium modified MMT DMAC [85]

us
an
M
d
p te
ce
Ac

Page 175 of 302


TABLE 3. Peak intensity (I m) and interlayer spacing (d) of nylon-6-based nanocomposites
prepared in presence of different acid derivatives by the one-pot technique [95]
Acid I ma (cps) d (A)
Phosphoric acid 0 0

t
Hydrochloric acid 200 21.7
Isophtalic acid 255 20.2

ip
Benzenesulfonic acid 280 19.3
Acetic acid 555 20.3
Trichloroacetic acid 585 21.3

cr
No acid 1840 18.6

us
an
M
d
p te
ce
Ac

Page 176 of 302


TABLE 4. Synthesis and composition of PE-based nanocomposite produced by in situ
intercalative polymerization of ethylene (P(C2H4)=10 bar) in non-organo-modified layered
silicatesa [1]

Filler MAO Catalyst P(H2)b Filler loadingc HDPE Mn


(10-3 mol) (10-6 mol) (bar) (wt %) (g/mol)
h 33.00 15.6 0 4.2 -d
m 27.20 12.5 0 3.3 -d
h 23.75 16.2 0.3 3.4 77.000

t
a h=hectorite and m=montmorillonite

ip
b Hydrogen partial pressure at start
c Measured by thermogravimetric analysis (TGA)
d Insoluble UHMWPE that cannot be eluted by SEC

cr
us
an
M
d
p te
ce
Ac

Page 177 of 302


TABLE 5. Interlayer spacings of organo-modified montmorillonites (X-MMT) and as-obtained PS-
based nanocomposites and the clay dispersibility within the polymerization medium [113]
Xa-MMT Interlayer spacing (Å) Dispersibilityb
In X-MMT In PS/X-MMT
Sodium MMT 11.8 14.2 -
BZ-MMT 19.1 34.0 ++
Eh-MMT 20.4 28.5 +

t
Ta-MMT 32.7 32.9 +

ip
a Organo-modifiers: (CH3)2N+ (hydrogenated tallow alkyl)R with R=Bz (benzyl), Eh (2-ethylhexyl),

or Ta (hydrogenated tallow alkyl),b It was judged by the appearance of the montmorillonite


dispersion in styrene monomer (++)fully dispersible, (+) partly dispersible, (-) non-dispersible

cr
us
an
M
d
p te
ce
Ac

Page 178 of 302


TABLE 6. d-spacing of Tixogel and OPTC18 at 5 phr in epoxy/amine systems at the beginning
and at the end of polymerization
Epoxy system + 5 phr Tixogel Epoxy system + 5 phr OPTC18
DGEBA/MCDEA DGEBA/D2000 DGEBA/MCDEA DGEBA/D2000
d-spacing when reaction
33.2 33.7 33.0 54.5
starts (Å)
d-spacing at the end of

t
33.9 35.0 70 110
the reaction (Å)

ip
cr
us
an
M
d
p te
ce
Ac

Page 179 of 302


TABLE 7. Polyurethane characteristics
PU PU Polyol
reference nanocomposite Name Calculated Functionalitya Wt % EOb
Mn end
I NC-I Acclaim 2220 2000 2 15
II NC-II Acclaim 4220 4000 2 15
III NC-III Daltocel F435 4000 2.5 17

t
IV NC-IV Arcol 1374 6000 2.3 15

ip
a Approximated
b EO=ethylene oxide

cr
us
an
M
d
p te
ce
Ac

Page 180 of 302


TABLE 8. d-spacings for organoclay and PU/clay nanocomposites
Material Clay d-spacing
Cloisite 30B (organoclay) 18
NC-I 65
NC-II None
NC-III 102
NC-IV none

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 181 of 302


TABLE 9. PLS nanocomposites prepared by in situ polymerization
Nanocomposite Monomer Polymerization conditions Ref.
PA6/12-aminolauric acid Ring opening, 250-270 °C,
ε-caprolactam [25]
modified MMT 48 h
ω-amino
280 °C, 20 bar, 9.5 h (after
PA12/Somasif ME 100 dodecanoic acid [44]
an initial swelling stage)
(ADA)

t
PCL/ aminolauric acid modified

ip
ε-caprolactone Ring opening, 170 °C, 48 h [117
MMT
PCL/Cr3+ exchanged
ε-caprolactone 100 °, 48 h [118]
fluorohectorite

cr
PE/MMT with fixed Ti-based
ethylene 30-50 °C, 4 bars [105]
Ziegler-Natta catalyst
PMMA/OMLS MMA Free radical, 80 °C, 5 h [98, 99]

us
PS/OMLS styrene Free radical, 100 °C, 16 h [98, 99]
PS/vinylbenzyl-trimethyl- Free radical, 80 °C, 5h,
styrene [100]
ammonium-MMT AIBN

an
Ethylene glycol,
PET/o-MMT terephthalic acid [106]
derivatives
Crosslinking, diamine
Epoxy/o-MMT DGEBA [125]
curing agent
M
Epoxy/ Octadecylammonium DGEBA, TGAP,
DETDA curing agent [126]
modified MMT TGDDM
UP/ MMT treated with a
Unsaturated
mathacrylate-silane coupling Free radical crosslinking [127]
d
polyester
agent
p te
ce
Ac

Page 182 of 302


TABLE 10. Interlayer spacing of various modified montmorillonite and the resulting composites
obtained with EVA (10.76 mol % vinyl acetate)
Code Cation Interlayer spacing
In modified clay In EVA composite
Mont-Na Na+ 12.2 12.6
Mont-2CN2C18 (CH3)2N+(C18H37)2 31.9 40.3
Mont-NC11COOH H3N+C11H22COOH 16.3 16.7

t
Mont-3CNC21COOH (CH3)3N+C21H42COOH 20.1 20.1

ip
cr
us
an
M
d
p te
ce
Ac

Page 183 of 302


TABLE 11. Characteristics of the various studied clays

Filler Clay type Interlayer cations Ammonium Interlayer


contenta (wt %) distanceb (Å)
Cloisite Na montmorillonite Na+ 0 12.1
Cloisite 20A montmorillonite (CH3)2N+ 29.2 22.1
(hydrogenated
tallow)2
Cloisite 25A montmorillonite (CH3)2N+ 26.9 20.7

t
(hydrogenated

ip
tallow)(2-ethylhexyl)
Cloisite 30B montmorillonite (CH3)2N+ 20.3 18.5
(tallow)(CH2CH2OH)2

cr
Nanofil 757 montmorillonite Na+ 0 12.2
Nanofil 15 montmorillonite (CH3)2 N+ 28.9 ~29 (broad)
(hydrogenated

us
tallow)2
Somasif ME100 fluoromica Na+ 0 12.2
Somasif MAE fluoromica (CH3)2N+ 40.8 31.1

an
(hydrogenated
tallow)2
Tallow: Linear alkyl chains (C18 (65%), C16 (30%), C14 (5%) with ~ 40% mono-unsaturated chains
a Determined by thermogravimetric analysis under helium atmosphere
b Determined by X-ray diffraction on as-received clays
M
d
p te
ce
Ac

Page 184 of 302


TABLE 12. Interlayer spacing variation as obtained from diffraction peaks measured by XRD on
clays and the resulting melt-blended EVA composites
Filler Filler interlayer Composite interlayer Interlayer distance
distance (Å) distance(Å) variation (Å)
Cloisite Na 12.1 12.2 0.1

Cloisite 20A 22.4 38.7 16.3

Cloisite 25A 20.7 36.8 16.1

Cloisite 30B 18.5 - -

t
12.2 12.2 0

ip
Nanofil 757
Nanofil 15 Ca 29 (broad) 40.2 Ca 11

Somasif ME100 12.2 12.3 0.1

cr
Somasif MAE 31.1 40.4 9.3

us
an
M
d
p te
ce
Ac

Page 185 of 302


TABLE 13. Structures of the polar polymers used as composite matrices
Polymer Structure

LDPE

EMA (y=0.215)

t
ip
EVA -18, -28
(x=0.18, 0.28)

cr
EMAAA
(y=0.18, z=0.06)

us
PE-g-MAH
(y≈0.02)

an
M
d
p te
ce
Ac

Page 186 of 302


TABLE 14. Molecular weight of host polymers
SPU HPU
Mn PDI Mn PDI
As-received PU 216.000 1.8 85.000 2.0
Solvent cast 121.000 1.7 82.000 2.1
Melt compounded 66.000 2.0 59.000 1.9

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 187 of 302


TABLE 15. PLS nanocomposites prepared by melt intercalation
Nanocomposite Mixing device and coditions Ref.
Co-rotating twin screw extruder,
PA6/[(HE)2M1R1] modified MMT [14]
240 °C, 280 rpm
PA6/Na+-MMT water slurry Extrusion and drying [160]
Statically heating at 165 °C,

t
PS/alkylammonium modified MMT [29]
vacuum, 25 h

ip
PEI/hexadecylamine modified MMT Internal mixer , 370 °C, 30 min [46]

PEO/Li+ or Na+-MMT Statically annealing, 80 °C, 6 h [196]

cr
PLA/C18-MMT Twin screw extruder, 190 °C [153]
Twin screw extruder, PP-MA
PP/stearylammonium modified clay [197]
compatibilizer

us
PP-MA/C18-MMT Twin screw extruder, 200 °C [186]
PP/o-MMT modified using an organic swelling
Twin screw extruder, 250 °C [166]
agent (Tb=100-200 °C)

an
EVA/dimethyl-dioctadecyl ammonium modified
130 °C [166]
MMT
PET/1,2-dimethyl-3-N-alkyl imidazolium salt Co-rotating mini twin screw
[148]
modified MMT extruder, 285 °C
M
d
p te
ce
Ac

Page 188 of 302


TABLE 16. Predicted reinforcing factors per number of platelets per stack
Reinforcing factor (RF) Halpin-Tsai
No of platelets Mori-Tanaka theory
equation
per stack
d001=0.96 nm d001=1.8 nm d001=0.96 nm d001=1.8 nm
1 49.2 49.2 34.8 34.8
2 39.7 27.5 23.8 16.6
3 33.5 21.1 18.3 11.7

t
4 29.0 17.5 14.9 9.3

ip
5 25.6 15.1 12.7 7.8
10 16.3 9.3 7.5 4.6

cr
us
an
M
d
p te
ce
Ac

Page 189 of 302


TABLE 17. Young modulus of various PLS nanocomposites
Nanocomposite Clay content (wt %) Young modulus (GPa) Ref
PA6/MMT (in situ polymerization) 0 1.11 [199]
4.7 1.87
5.3 2.04
PA6(LMW)/MMT (melt intercalation) 0 2.82 [14]
3.2 3.65

t
6.4 4.92

ip
PA6(MMW)/MMT (melt intercalation) 0 2.71 [14]
3.1 3.66
7.1 5.61

cr
PA6(HMW)/MMT (melt intercalation) 0 2.75 [14]
3.2 3.92
7.2 5.70
PP(7.2 % PP-g-MA)/OMLS 0 0.714 [186]

us
7.2 0.838
PP(21.6 % PP-g-MA)/OMLS 0 0.760 [186
7.2 1.010
EVA/Cloisite Na 0 0.0122 [167]

an
3 0.0135
EVA/Cloisite 20A 0 0.0122 [167]
3 0.0249
EVA/Cloisite 25A 0 0.0122 [167]
3 0.0220
M
EVA/Cloisite 30B 0 0.0122 [167]
3 0.0228
EVA/Nanofil 757 0 0.0122 [167]
3 0.0116
EVA/Nanofil 15 0 0.0122 [167]
d

3 0.0240
EVA/Somasif ME100 0 0.0122 [167]
te

3 0.0124
EVA/Somasif MAE 0 0.0122 [167]
3 0.021
p

Soft PU/30B (solution intercalation) 0 0.0075 [88]


3 0.0138
7 0.024
ce

Soft PU/30B (melt intercalation) 0 0.0072 [88]


3 0.0114
7 0.0193
Hard PU/30B (solution intercalation) 0 0.050 [88]
Ac

3 0.086
7 0.134
Hard PU/30B (melt intercalation) 0 0.061 [88]
3 0.081
7 0.119
HDPE/o-MMT 0 1.020 [200]
0.9 1.060
1.8 1.250
2.8 1.380
4.0 1.360

Page 190 of 302


TABLE 18. Tensile strength of various PLS nanocomposites
Nanocomposite Clay content (wt %) Tensile strength (MPa) Ref
PA6/MMT (in situ polymerization) 0 68.6 [199]
4.7 97.2
5.3 97.3
PA6(LMW)/MMT (melt intercalation) 0 69.2 [14]
3.2 78.9

t
6.4 83.6

ip
PA6(MMW)/MMT (melt intercalation) 0 70.2 [14]
3.1 86.6
7.1 95.2

cr
PA6(HMW)/MMT (melt intercalation) 0 69.7 [14]
3.2 84.9
7.2 97.6
PMMA/OMLS 0 53.9 [1]

us
12.6 62.0
PS/OMLS 0 28.7 [1]
17.2 23.4
24.6 16.6

an
EVA 0 28.4 [167]
EVA/Cloisite Na 3 25.9
EVA/Cloisite 20A 3 25.8
EVA/Cloisite 25A 3 26.2
EVA/Cloisite 30B 3 30.7
M
EVA/Nanofil 757 3 27.6
EVA/Nanofil 15 3 26.7
EVA/Somasif ME100 3 24.5
EVA/Somasif MAE 3 25.1
Soft PU/30B (solution intercalation) 0 45 [88]
d

3 31
7 21
te

Hard PU/30B (solution intercalation) 0 58 [88]


3 44
7 34
PU/MMT 0 5.9 [142]
p

5 6.2
10 6.5
ce

21.5 8.3
PE/JS 0 22 [184]
5 25
10 27
15 28
Ac

PE/DM 0 22 [184]
5 21
10 23
15 24
HDPE/o-MMT 0 27 [200]
0.9 26
1.8 26
2.8 26
4.0 25

Page 191 of 302


TABLE 19. XRD peak intensity (Im) and Young’s modulus of various nylon-6-based
nanocomposites obtained by a one-step in situ intercalative polymerization of ε-caprolactam
with Na-montmorillonite in the presence of different acids
Acid Im (cps) Young’s modulus (GPa)
Phosphoric acid 0 2,25
Hydrochloric acid 200 2,05
Isophtalic acid 255 1,74

t
Benzenesulfonic acid 280 1,74

ip
Acetic acid 555 1,63
Trichloroacetic acid 585 1,67

cr
us
an
M
d
p te
ce
Ac

Page 192 of 302


TABLE 20. Influence of maleic anhydride-modified polypropylene content on the stiffness of PP
matrices and PP/clay nanocompositesa
Sample Filler content (wt %) PP-MA content (wt %) Young’s modulus
(MPa)
PP 0 0 780
PP/PP-MA 7 0 7.2 714
PP/PP-MA 22 0 21.6 760

t
PPCC 6.9 0 830

ip
PPCH 1/1 7.2 7.2 838
PPCH 1/2 7.2 14.4 964
PPCH 1/3 7.2 21.6 1010

cr
a PP=polypropylene; PP-MA x: polypropylene modified by maleic anhydride )x=wt % of PP-MA

in the blend); PPCC= polypropylene-based microcomposite, PPCH y/z = = polypropylene-


based nanocomposite (y/z=weight ratio between y parts of filler and z parts of PP-MA)

us
an
M
d
p te
ce
Ac

Page 193 of 302


TABLE 21. Mechanical properties of PE and PE/clay composites
Sample Tensile strength Flexural strength Flexural modulus Izod impact
(MPa) (MPa) (MPa) strength (J/m)
PE 22 26 710 20
PE/JS5 25 28 780 16
PE/JS10 27 33 1050 16
PE/JS15 28 38 1330 12

t
PE/DM5 21 26 750 22

ip
PE/DM10 23 31 980 16
PE/DM15 24 30 1030 14

cr
us
an
M
d
p te
ce
Ac

Page 194 of 302


TABLE 22. Tensile test results of Polypox H205 nanocomposites processed with low shear
Specimen Average Average tensile Relative ultimate Relative tensile
ultimate tensile modulus stress modulus
stress (N/mm2) (N/mm2)
0 % I30E 59.0 ± 0.6 2565 ± 11 1.000 1.000
5 % I30E 54.3 ± 9.2 2796 ± 31 0.920 1.090
10 % I30E 53.6 ± 6.5 3075 ± 56 0.909 1.199

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 195 of 302


TABLE 23. Flexural test results of Polypox H205 nanocomposites processed with low shear
Specimen Average flexural Average Relative flexural Relative flexural
modulus ultimate flexural modulus stress
(N/mm2) stress (N/mm2)
0 % I30E 2755 ± 83 95.0 ± 1.8 1.000 1.000
5 % I30E 2966 ± 90 97.3 ± 2.0 1.076 1.034
7.5 % I30E 3101 ± 85 102.2 ± 3.2 1.126 1.055

t
10 % I30E 3294 ± 76 102.4 ± 3.1 1.196 1.077

ip
cr
us
an
M
d
p te
ce
Ac

Page 196 of 302


TABLE 24. Summary of tensile properties of thermoplastic PU-based nanocomposites
30B content Solvent Cast Melt compounded
(wt %) Young’s Tensile Fail strain Young’s Tensile Fail strain
modulus strength (%) modulus strength (%)
(MPa) (MPa) (MPa) (MPa)
SPU 0 7.5 45 1136 7.2 21 1445
3 13.8 31 1109 11.4 22 1163

t
7 24 21 1030 19.3 7 568

ip
HPU 0 50 58 898 61 44 776
3 86 44 808 81 20 283
7 134 34 704 119 15 100

cr
us
an
M
d
p te
ce
Ac

Page 197 of 302


TABLE 25. Tensile properties of PET hybrid fibers
Organoclay (wt %) DRa Ult. Strength (MPa) Modulus (GPa) E.B.b
0 (pure PET) 1 46 2.21 3
3 47 2.24 3
10 51 2.28 3
16 51 2.39 2
1 1 58 2.88 3

t
3 56 2.80 3

ip
10 50 2.63 3
16 48 2.47 3
2 1 68 3.31 3

cr
3 55 2.63 3
10 54 2.51 3
16 51 2.29 3
3 1 71 4.10 3

us
3 68 3.40 3
10 62 3.12 2
16 55 3.08 3
a Draw ratio

an
b Elongation percent at break
M
d
p te
ce
Ac

Page 198 of 302


TABLE 26. Bending modulus of various PLS nanocomposites
Nanocomposite Clay Content (wt %) Bending Modulus (GPa) Ref.
PLA/OMLS 0 4.8 [208]
4 5.5
5 5.6
7 5.8
PE/JS 0 0.71 [184]

t
5 0.78

ip
10 1.050
15 1.330
PE/DM 0 0.71 [184]

cr
5 0.75
10 0.98
15 1.030

us
an
M
d
p te
ce
Ac

Page 199 of 302


TABLE 27. Bending strength of various PLS nanocomposites
Nanocomposite Clay Content (wt %) Bending Strength (MPa) Ref.
PLA/OMLS 0 86 [208]
4 134
5 122
7 105
PE/JS 0 26 [184]
5 28
10 33

t
15 38

ip
PE/DM 0 26 [184]
5 26
10 31
15 30

cr
us
an
M
d
p te
ce
Ac

Page 200 of 302


TABLE 28. Elongation at break of various PLS nanocomposites
Nanocomposite Clay content (wt %) Elongation at break (%) Ref
PA6(LMW)/MMT 0 28 [14]
3.2 11
6.4 4.8
PA6(MMW)/MMT 0 101 [14]
3.1 18
7.1 5.0
PA6(HMW)/MMT (melt intercalation) 0 129 [14]

t
3.2 27

ip
7.2 6.1
EVA 0 1406 [167]
EVA/Cloisite Na 3 1403

cr
EVA/Cloisite 20A 3 1231
EVA/Cloisite 25A 3 1259
EVA/Cloisite 30B 3 1266
EVA/Nanofil 757 3 1358

us
EVA/Nanofil 15 3 1291
EVA/Somasif ME100 3 1312
EVA/Somasif MAE 3 1270
Soft PU/30B (solution intercalation) 0 1136 [88]
3 1109

an
7 1030
Soft PU/30B (melt intercalation) 0 1445 [88]
3 1163
7 568
Hard PU/30B (solution intercalation) 0 898 [88]
M
3 808
7 704
Hard PU/30B (melt intercalation) 0 776 [88]
3 283
d
7 100
PU/MMT 0 950 [142]
5 1020
te

10 1065
21.5 1160
HDPE/o-MMT 0 36 [200]
p

0.9 25
1.8 20
2.8 14
ce

4.0 15
Ac

Page 201 of 302


TABLE 29. Tensile strength and strain-at-break vs. the loading of layered clay
Weight fraction Tensile strength Strain-at-break
of the clay (%) (MPa) (%)
0 5.9 950
5 6.2 1020
10 6.5 1065
21.5 8.3 1160

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 202 of 302


TABLE 30. Compounding formulations and tensile properties of PP-based RTPO/PpgMA/clay
nanocomposites.
Sample PP- PpgMA Clay Tensile Tensile Elongati Tensile
Based (wt %) (wt %) strength strength on at modulus
RTPO (wt at yield at break break (MPa)
%) (MPa) (MPa) (%)
RTPO 100 0 0 5.1 20.6 1390 46.0

t
RTPO NC3 88 9 3 6.4 16.6 980 71.2

ip
RTPO NC5 80 15 5 8.1 16.8 859 78.3
RTPO NC10 60 30 10 14.2 16.6 437 251

cr
us
an
M
d
p te
ce
Ac

Page 203 of 302


TABLE 31. Mechanical properties of polyamide 6 composites
Polyamide Clay Izod impact Modulus Yield Elongation at break (%)
composites content strength (GPa) strength
(%) (J/m) (Mpa)
Crosshead Crosshead
speed 0,51 speed 5,08
cm/min cm/min

t
Polyamide 6 0 38 ± 4 2,66 ± 0,2 64,2 ± 0,8 200 ± 30 40 ± 8

ip
PA6/glass fiber 5 53 ± 8 3,26 ± 0,1 72,6 ± 0,8 18 ± 1,3 14 ± 4
PA6/MMT 5 40 ± 2 3,01 ± 0,1 75,4 ± 0,3 22 ± 6,0 14 ± 3

cr
PA6/organoclay 3,16 38 ± 3 3,66 ± 0,1 83,4 ± 0,7 126 ± 25 38 ± 19
PA6/organoclay/
8 44 ± 3 4,82 ± 0,1 95,0 ± 0,9 8 ± 0,5 7±4
glass fiber

us
an
M
d
p te
ce
Ac

Page 204 of 302


TABLE 32. Relative barrier performance of newly developed rigid packaging based on PET

Container composition (supplier) Relative oxygen transmission rate at 23 °C 50 % RH


PET 1
PET nanocomposite (Tetrapak) < 0.3
PET/PA nanocomposite (Eastman) 0.4-0.7

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 205 of 302


TABLE 33. Oxygen gas permeabilities of the PLA/OMLS hybrid films
Clay (wt %) O2 gas (cc/m2/day)
C16MMT DTA-MMT C25A
0 777 777 777
4 449 455 -
6 340 353 430
10 327 330 340

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 206 of 302


TABLE 34. Thermal properties of PET hybrid fibers (DRa=1)

Organoclay (wt %) ninhb Tm (°C) ∆Hmc (J/g) TDid (°C) wtR600e


0 (pure PET) 1.02 245 32 370 1
1 1.26 247 32 375 8
2 0.98 245 33 384 15
3 1.23 246 32 386 21

t
ip
a Draw ratio, b Inherent viscosities measured at 30 °C by using 0.1 g/100 ml solutions in a
phenol/1,1,2,2-tetrachloroethane (w/w) mixture, c Enthalpy change of fusion, d Initial weight
reduction onset temperature, e Weight percent of residue at 600 °C

cr
us
an
M
d
p te
ce
Ac

Page 207 of 302


TABLE 35. Maximum temperature at the main degradation peak as measured under air at 20
°C/min for EVA and EVA-nanocomposites with different organoclay contents.
Organoclay content (wt %) Maximum temperature
at the main degradation peak (°C)
0 452.0
1 453.4
2.5 489.2

t
5 493.5

ip
10 472.0
15 454.0

cr
us
an
M
d
p te
ce
Ac

Page 208 of 302


TABLE 36. Results of TGA and DSC data for the polystyrene nanocomposites
Sample Clay Tg T0.05 T0.5 Char at 600
content (°C)a (°C) b (°C) c °C
(wt %) (%)
PS 0 100 390 424 0
CPC/Clay/PS 3 102 408 424 2.9
a-CD/CPC/Clay/PS 3 106 423 452 5.8

t
ip
a Glass transition temperature (Tg)
b 5 % Degradation temperature (T0.05)
c 50 % Degradation temperature (T0.5)

cr
us
an
M
d
p te
ce
Ac

Page 209 of 302


TABLE 37. Cone calorimeter data of various polymers and nanocomposites with OMLS
Sample PHRR (kW/m2) HRRave (kW/m2) SEAave (m2/kg) Ref
PA6 1010 603 197 [240]
PA6/2% OMLS 686 390 271
PA6/5% OMLS 378 304 296
PS 1120 703 1460 [240]
PS silicate mix 3 % 1080 715 1840

t
PS nanocomposite 3 % 567 444 1730

ip
PSw/DBDPO/Sb2O3 30 %) 491 318 2580
PS 1230 1315 [159]
PS/MAST-Hect 1 % 1011 1336

cr
PS/MAST-Hect 3 % 894 1332
PS/MAST-Hect 5 % 728 1327
PPgMA 1525 536 704 [240]
PPgMA nanocomposite 2 % 450 322 1028

us
PPgMA nanocomposite 4 % 381 275 968
EVA/Na+ 5 % 1200 [163]
EVA/Cloisite 30B 3 % 860
EVA/Cloisite 30B 5 % 780

an
EVA/Cloisite 30B 10 % 630
EVA 2303 430 [237]
EVA/30B 1174 670
EVA/30BHect 1289 593
EVA/30BMag 2010 476
M
EVA/MMT 1959 517
PU (I) 2561 741 176 [144]
PU (I)/o-MMT 918 344 305
PU (II) 2254 637 235 [144]
PU (II)/o-MMT 641 363 412
d

PU (III) 2647 768 165 [144]


PU (III)/o-MMT 848 444 172
te

PU (IV) 2664 775 235 [144]


PU (IV)/o-MMT 797 435 412
PE 1470 510 [184]
PE/JS 2 % 670 440
p

PE/JS 5 % 320 450


PE/JS 10 % 540 380
ce

PE/JS 15 % 390 280


Ac

Page 210 of 302


TABLE 38. Cone calorimetry results for PU nanocomposites and reference materials
PU TTI (s) PHRR (kW/m2) PI (PHRR/TTI) HRRave (kW/m2) SEAave (m2/kg)
PU NC-PU PU NC-PU PU NC-PU PU NC-PU PU NC-PU
I 29 29 2561 918 88 32 741 344 176 305
II 35 33 2254 641 64 19 637 363 235 412
III 26 17 2647 848 102 50 768 444 165 172
IV 22 25 2664 797 102 50 775 435 235 412

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 211 of 302


TABLE 39. Cone calorimeter data (35 kW/m2) for PP nanocomposites [247]
Formulation PHRR PHRR, Average
(kW/m2) time (s) HRR
PP-MA + 22 wt % DBE + 5 wt % Sb2O3 300, 650 90, 170 254
PP-MA + 22 wt % DBE + 5 wt % Sb2O3 + 5 wt % organoclay 200 85 107
PP-MA + organoclay 350 85 188

t
PP-MA 600 180 279

ip
cr
us
an
M
d
p te
ce
Ac

Page 212 of 302


TABLE 40. HDT of PP/MMT nanocomposites and the respective unfilled PP

Organically modified MMT (wt %) HDT


PP-f-MMT PP/alkyl-MMT
0 109 109
3 144 130a
6 152 141b
9 153
a C18-MMT filler, extruder processed

t
b 2C18-MMT filler, twin head mixer

ip
cr
us
an
M
d
p te
ce
Ac

Page 213 of 302


Table 41. Selected commercial nanoclays
Product Characteristics Applications Producer
Cloisite Organophilic Additives to enhance Southern Clay
flexural and tensile Products
modulus, barrier
properties and flame
retardancy of

t
thermoplastics

ip
Nanomers Microfine powder Nylon, epoxy, Nanocor
unsaturated polyester,
engineering

cr
resins

Masterbatches Pellet Thermoplastic olefin PolyOne


and urethane, Corporation,

us
styrene-ethylene Clariant
butylene-styrene, Corporation,
ethylene vinyl acetate RTP Company

Bentone With a broad range of polarity Additives to enhance Elementis

an
mechanical, flame Specialties
retardancy and
barrier properties of
thermosets and
thermoplastics
M
Nanofil Improve the mechanical, Thermoplastics and Sud-Chemie
thermal and barrier properties thermosets
d
Planomers Additive, enhance mechanical Electric and electronic, TNO
barrier properties, thermal medical and
stability and flame resistance healthcare,
te

adhesive, building and


construction materials

PlanoColors Nanopigments, e.g., blue, red, Decorative coloring, TNO


p

green, yellow, high UV-stability UV-stable coloring,


heavy metal
ce

free coloring
PlanoCoatings Additive, excellent transparency Transparent packaging TNO
and improved barrier properties materials,
protective coatings,
transparent barrier
Ac

coatings

Page 214 of 302


Table 42. Selected commercial polymer nanocomposites
Product Characteristics Applications Producer

Nylon Improved modulus, Automotive parts Bayer, Honeywell


nanocomposites strength, heat (e.g., timing-belt Polymer, RTP
distort temperature, cover, engine Company, Toyota
barrier cover, barrier fuel Motors, Ube,
properties line), packaging Unitika
(e.g., cosmetics,

t
food, medical,

ip
electronics), barrier
film
Polyolefin Stiffer, stronger, less Step-assist for GMC Basell, Blackhawk

cr
nanocomposites brittle, Safari and Automotive
lighter, more easily chevrolet Astro Plastics, General
recycled, vans, heavy-duty Motors, Gitto
improved flame electrical enclosure Global Corporation,

us
retardancy Southern
Clay Products
M9™ High barrier properties Juice or beer Mitsubishi Gas
bottles, multi-layer Chemical

an
films, containers Company
Durethan KU2-2601 Doubling of stiffness, Barrier films, paper Bayer
(nylon 6) high gloss and coating
clarity, reduced oxygen
transmission rate,
M
improved
barrier properties
Aegis™ NC Doubling of stiffness, Medium barrier Honeywell Polymer
(nylon 6/barrier higher heat bottles and films
nylon) distort temperature,
d

improved clarity
Aegis™ OX Highly reduced oxygen High barrier beer Honeywell Polymer
te

transmission bottles
rate, improved clarity
SET™ Improved stiffness, Catheter shafts and Foster Corporation
p

nanocomposite permeability, fire balloons, tubing,


nylon 12 retardancy, film and barriers,
transparency and flexible
ce

recycling devices
Forte™ Improved temperature Automotive, Noble Polymer
nanocomposite resistance and furniture, appliance
stiffness, very good
Ac

impact properties

Page 215 of 302


FIGURES CAPTIONS

FIGURE 1. The structure of a 2:1 layered silicate [35]. Reproduced from Beyer by permission
of Elsevier Science Ltd. UK.

FIGURE 2. Schematic picture of an ion-exchange reaction. The inorganic, relatively small


(sodium) ions are exchanged against more voluminous organic onium cations. This ion-

t
exchange reaction has two consequences: firstly, the gap between the single sheets is

ip
widened, enabling polymer chains to move in between them and secondly, the surface
properties of each single sheet are changed from being hydrophilic to hydrophobic [2].

cr
Reproduced from Fischer by permission of Elsevier Science Ltd. UK.

FIGURE 3. Alkyl chain aggregation in layered silicates: (a) lateral monolayer; (b) lateral

us
bilayer; (c) paraffin-type monolayer and (d) paraffin-type bilayer [1]. Reproduced from
Alexandre and Dubois by permission of Elsevier Science Ltd. UK.

an
FIGURE 4. Alkyl chain aggregation models: (a) short alkyl chains: isolated molecules, lateral
monolayer; (b) intermediate chain lengths: in-plane disorder and interdigitation to form quasi
bilayers and (c) longer chain length: increased interlayer order, liquid crystalline-type
M
environment [1]. Reproduced from Alexandre and Dubois by permission of Elsevier Science
Ltd. UK.
d

FIGURE 5. WAXS results for organoclays: mass of organic per unit mass of montmorillonite
[51]. Reproduced from Fornes, Yoon, Hunter, Keskkula and Paul by permission of Elsevier
te

Science Ltd. UK.


p

FIGURE 6. Schematic illustration of two different types of thermodynamically achievable


polymer/layered silicate nanocomposites [50]. Reproduced from Ray and Bousima by
ce

permission of Elsevier Science Ltd. UK.

FIGURE 7. Typical XRD patterns from polymer/layered silicates: (a) PE + organoclay → no


Ac

formation of a nanocomposite, (b) PS + organoclay → intercalated nanocomposite, (c)


Siloxane + organoclay → delaminated nanocomposite [35]. Reproduced from Beyer by
permission of Elsevier Science Ltd. UK.

FIGURE 8. TEM micrographs of poly(styrene)-based nanocomposites: (a) intercalated


nanocomposite and (b) exfoliated nanocomposite [1]. Reproduced from Alexandre and
Dubois by permission of Elsevier Science Ltd. UK.

FIGURE 9. DSC traces of pure PS, a physical mixture of PS/OLS, and PS-intercalated OLS
[41]. Reproduced from Zanetti, Lomakin and Camino by permission of Wiley-VCH,
Germany.

Page 216 of 302


FIGURE 10. Examples of complications in the determination of the aspect ratio of layered
silicate fillers within polymer nanocomposites [64]. Reproduced from Fornes and Paul by
permission of Elsevier Science Ltd. UK.

FIGURE 11. Schematic representation of PLS obtained by intercalation from solution

FIGURE 12. XRD patterns of ω-amino acid [NH2(CH2)n-1COOH] modified Na+-MMT [93].

t
Reproduced from Usuki, Kawasumi, Kojima, Okada, Kurauchi and Kamigaito by permission

ip
of Materials Research Society, USA.

cr
FIGURE 13. Swelling behavior of ω-amino acid modified MMT by ε-caprolactam [93].
Reproduced from Usuki, Kawasumi, Kojima, Okada, Kurauchi and Kamigaito by permission
of Materials Research Society, USA.

us
FIGURE 14. XRD intensity curve of injection molded nylon-6 nanocomposite as obtained by
the one-pot intercalation polymerization process in the presence of acetic acid [95].

an
Reproduced from Kojima, Usuki, Kawasumi, Okada, Kurauchi and Kamigaito by permission
of John Wiley & Sons, Inc.
M
FIGURE 15. Interlayer distance of fluoro-modified talc (ME 100) in function of an increasing
amount of aminolauric acid used as the organic modifier [96]. Reproduced from Reichert,
Kressler, Thomann, Mulhaupt and Stoppelmann by permission of Wiley-VCH, Germany.
d

FIGURE 16. Schematic representation of the swelling behavior of the fluoro-modified talc
te

ME 100 in presence of aminolauric acid [96]. Reproduced from Reichert, Kressler, Thomann,
Mulhaupt and Stoppelmann by permission of Wiley-VCH, Germany.
p

FIGURE 17. Schematic representation of nanocomposite formation by ring-opening reaction


ce

of cyclic oligomers in between silicate layers [107]. Reproduced from Lee, Ma, Rhee and
Kim J by permission of Elsevier Science Ltd. UK.

FIGURE 18. Mechanistic representation of the fixing of TiCl4 between the silicate layers of
Ac

MMT–OH [105]. Reproduced from Jin, Park, Im, Kwak and Kwak by permission of Wiley-
VCH, Germany.

FIGURE 19. Schematic illustration of the modification and ion exchange of laponite with
[Zr(n-C5H5)2Me(thf)]BPh4 and propene polymerization [104]. Reproduced from Tudor,
Willington, O’Hare and Royan by permission of The Royal Society of Chemistry, UK.

FIGURE 20. Powder XRD of the composite before (solid line) and after (dashed line)
polymerization. Insets are schematic illustrations (not drawn to scale) corresponding to the
intercalated monomer (left) and intercalated polymer (right) [117]. Reproduced from
Messersmith and Giannelis by permission of the American Chemical Society, USA.

Page 217 of 302


FIGURE 21. Preparation of PU/clay nanocomposites [144]. Reproduced from Berta, Lindsay,
Pans and Camino by permission of Elsevier Science Ltd. UK.

FIGURE 22. SEM Micrographs of PU foams at cross-sections parallel to foam rising


direction [145]. Reproduced from Gao, Lee, Widya and Macosko by permission of Elsevier
Science Ltd. UK.

t
FIGURE 23.The ‘melt intercalation’ process [35]. Reproduced from Beyer by permission of

ip
Elsevier Science Ltd. UK.

cr
FIGURE 24. The synthetic route for the formation of the terpolymer MAST [159].
Reproduced from Zheng, Jiang and Wilkie by permission of Elsevier Science Ltd. UK.

us
FIGURE 25. Schematic figures depicting dispersion of the Na+-MMT slurry into nylon-6
during compounding [160]. Reproduced from Hasegawa, Okamoto, Kato, Usuki and Sato by
permission of Elsevier Science Ltd. UK.

an
FIGURE 26. (a) Molecular structure and nomenclature of amine salts used to organically
modify sodium montmorillonite by ion exchange. The symbols M: methyl, T: tallow
M
(predominantly composed of chains with 18 carbons (~65%)), HT: hydrogenated tallow, HE:
2-hydroxy-ethyl, R: rapeseed (consisting largely of chains with 22 carbons (~45%)), C*: coco
(product made from coconut oil, consisting predominantly of chains with 12 carbons (~48%)),
d

and H: hydrogen designate the substituents on the nitrogen. (b) Organoclays used to evaluate
the effect of structural variations of the amine cations on nanocomposite morphology and
te

properties [51]. Reproduced from Fornes, Yoon, Hunter, Keskkula and Paul by permission of
Elsevier Science Ltd. UK.
p

FIGURE 27. Illustration, for a bimodal surfactant brush, of one possible outcome on longer
ce

surfactant conformation of the densely packed brush of a shorter surfactant [169]. Reproduced
from Kurian, Dasgupta, Beyer and Galvin by permission of John Wiley & Sons Inc.
Ac

FIGURE 28. Schematic picture of the action of block copolymer for an exfoliation of clay
sheets within a polymeric matrix [2]. Reproduced from Fischer by permission of Elsevier
Science Ltd. UK.

FIGURE 29. Proposed mechanism of how the organoclay particles disperse into polymers
during melt processing. Part (a) shows three cases involving the interplay between chemistry
and process conditions in the mixing device. Part (b) illustrates schematically how platelets
peel apart under the action of shear [52]. Reproduced from Dennis, Hunter, Chang, Kim,
White, Cho and Paul DR by permission of Elsevier Science Ltd. UK.

Page 218 of 302


FIGURE 30. Schematic illustration of OMLS dispersion process in PP-g-MA matrix [187].
Reproduced from Hasegawa, Okamoto, Kawasumi, Kato, Tsukigase and Usuki by permission
of Wiley-VCH, Germany.

FIGURE 31. Comparison of DTGA curves of various OMLS [43]. Reproduced from Xie,
Gao, Liu, Pan, Vaia, Hunter and Singh by permission of Elsevier Science Ltd. UK.

t
FIGURE 32. Synthesis of the thermally stable organoclay [193]. Reproduced from Chang,

ip
Kim, Joo and Im by permission of Elsevier Science Ltd. UK.

cr
FIGURE 33. Reinforcement mechanism in composite materials

FIGURE 34. The effect of the number of platelets per stack on the (a) modulus and (b) aspect

us
ratio for the simplified arrangement of platelets [64]. Reproduced from Fornes and Paul
permission of Elsevier Science Ltd. UK.

FIGURE 35. Tensile modulus vs clay concentration for crosslinked polyester nanocomposites

an
[140]. Reproduced from Bharadawaj, Mehrabi, Hamilton, Trujillo, Murga, Fan, Chavira and
Thompson by permission of Elsevier Science Ltd, UK.
M
FIGURE 36. Effect of clay content on tensile modulus, measured at room temperature, of
organomodified montmorillonite/nylon-6-based nanocomposite obtained by melt intercalation
[170]. Reproduced from Liu, Qi and Zhu by permission of John Wiley & Sons, Inc.
d

FIGURE 37. Tensile modulus and yield strength of PA12/clay nanocomposites as a function
te

of clay concentration [147]. Reproduced from Phang, Liu, Mohamed, Pramoda, Chen, Shen,
Chow, He, Lu and Hu by permission of John Wiley & Sons Ltd on behalf of the SCI.
p

FIGURE 38. Effect of MMT content on tensile modulus for LMW, MMW, and HMW based
ce

nanocomposites [14]. Reproduced from Fornes, Yoon, Keskkula and Paul by permission of
Elsevier Science Ltd. UK.
Ac

FIGURE 39. Effect of MMT content on yield strength for LMW, MMW, and HMW based
nanocomposites [14]. Reproduced from Fornes, Yoon, Keskkula and Paul by permission of
Elsevier Science Ltd. UK.

FIGURE 40. Effect of organic-MMT loading on the tensile strength of (a) PU/MO-MMT and
(b) PU/C16-MMT [202]. Reproduced from Xiong, Liu, Yang and Wang by permission of
Elsevier Science Ltd. UK.

FIGURE 41. SEM images showing fracture surfaces after impact tests: (A) neat PA12; (B)
and (C) PA12 nanocomposites containing 1 and 5 wt% clay, respectively; (D) high

Page 219 of 302


magnification of (C) [147]. Reproduced from Phang, Liu, Mohamed, Pramoda, Chen, Shen,
Chow, He, Lu and Hu by permission of John Wiley & Sons Ltd on behalf of the SCI.

FIGURE 42. Izod impact strength of PA12/clay nanocomposites as a function of clay


concentration [147]. Reproduced from Phang, Liu, Mohamed, Pramoda, Chen, Shen, Chow,
He, Lu and Hu by permission of John Wiley & Sons Ltd on behalf of the SCI.

t
FIGURE 43. Effect of clay content on notched Izod impact strength of nylon-6/clay

ip
nanocomposites obtained trough melt intercalation [170]. Reproduced from Liu, Qi and Zhu
by permission of John Wiley & Sons, Inc.

cr
FIGURE 44. Effect of MMT content on elongation at break for LMW, MMW, and HMW
based nanocomposites at a crosshead speed of (a) 0.51 cm/min and (b) 5.1 cm/min [14].

us
Reproduced from Fornes, Yoon, Keskkula and Paul by permission of Elsevier Science Ltd.
UK.

an
FIGURE 45. Stress-strain curves for nylon 6 and 95/05 composites at a crosshead speed of
5.08 cm/min [15]. Reproduced from Cho and Paul by permission of Elsevier Science Ltd.
UK.
M
FIGURE 46. Stress-strain curves for nylon 6 and 95/05 composites at a crosshead speed of
0,51 cm/min [15]. Reproduced from Cho and Paul by permission of Elsevier Science Ltd.
d

UK.
te

FIGURE 47. Temperature dependence of G΄; G΄΄ and tanδ for N6 matrix and various N6CNs
[55]. Reproduced from Ray and Okamoto by permission of Elsevier Science Ltd. UK.
p

FIGURE 48. Proposed model for the torturous zigzag diffusion path in an exfoliated
ce

polymer–clay nanocomposite when used as a gas barrier [85]. Reproduced from Yano, Usuki,
Okada, Kurauchi and Kamigaito by permission of John Wiley & Sons Inc.

FIGURE 49. Equilibrium concentration of water vapor, Ceq (g/100 g), as a function of
Ac

activity a=p/p0 for samples OMont (), NPU0 (●), NPU4 (), NPU20 (∆), NPU40 (◊) [204].
Reproduced from Tortora, Gorrasi, Vittoria, Galli, Ritrovati and Chiellini by permission of
Elsevier Science Ltd. UK.

FIGURE 50. The O2 permeability of PET/oMMT [222]. Reproduced from Ke and Yongping
by permission of Elsevier Science Ltd. UK.

FIGURE 51. Effect of montmorillonite weight fraction on diffusivity D of


montmorillonite/epoxy nanocomposites [223]. Reproduced from Ogasawara, Ishida,
Ishikawa, Aoki and Ogura by permission of Elsevier Science Ltd. UK.

Page 220 of 302


FIGURE 52. Effect of montmorillonite weight fraction on solubility, S, of
montmorillonite/epoxy nanocomposites. The numerical curve based on rule-of-mixture is
superimposed [223]. Reproduced from Ogasawara, Ishida, Ishikawa, Aoki and Ogura by
permission of Elsevier Science Ltd. UK.

FIGURE 53. Effect of montmorillonite weight fraction on permeability, P, of

t
montmorillonite/ epoxy nanocomposites [223]. Reproduced from Ogasawara, Ishida,

ip
Ishikawa, Aoki and Ogura by permission of Elsevier Science Ltd. UK.

FIGURE 54. MOCON permeation data (a) oxygen and (b) water vapour transmission [152].

cr
Reproduced from Thellen, Orroth, Froio, Ziegler, Lucciarini, Farrell, D’ Souza and Ratto by
permission of Elsevier Science Ltd. UK.

us
FIGURE 55. TGA curves of the degradation of PS100 and nPS90 at a heating rate 5 °C min-1
in air and nitrogen [225]. Reproduced from Vyazovkin, Dranka, Fan and Advincula by

an
permission of Wiley-VCH, Germany.

FIGURE 56. TGA curves of neat PLA and nano-PLA/MLS [152]. Reproduced from Thellen,
Orroth, Froio, Ziegler, Lucciarini, Farrell, D’ Souza and Ratto by permission of Elsevier
M
Science Ltd. UK.

FIGURE 57. Dominant PA-6 thermal degradation products in the absence of a nucleophile
d

(A) and in the presence of a nucleophile, such as water (B) [192]. Reproduced from Davis,
Gilman and VanderHart by permission of Elsevier Science Ltd. UK.
te

FIGURE 58. TGA and DTGA curves of PE and PE/clay nanocomposites in nitrogen
p

atmosphere [184]. Reproduced from Zhao, Qin, Gong, Feng, Zhang and Yang by permission
of Elsevier Science Ltd. UK.
ce

FIGURE 59. Possible radical recombination reactions for EVA nanocomposites [237].
Reproduced from Costache, Jiang and Wilkie by permission of Elsevier Science Ltd. UK.
Ac

FIGURE 60. TGA curves of (a) pure PU, (b) PU/3 wt% C16-MMT and (c) PU/3 wt% MO-
MMT [202]. Reproduced from Xiong, Liu, Yang and Wang by permission of Elsevier
Science Ltd. UK.

FIGURE 61. TGA curves of (a) pure CPC and (b) the CPC/a-CD inclusion complex [115].
Reproduced from Yei, Kuo, Fu and Chang by permission of Elsevier Science Ltd. UK.

FIGURE 62. Schematic view of the cone calorimeter [35]. Reproduced from Beyer by
permission of Elsevier Science Ltd. UK.

Page 221 of 302


FIGURE 63. Residues after cone calorimeter experiment for EVA, EVA/30B-5 and
EVA/Na+-5 [163]. Reproduced from Duquesne, Jama, Le Bras, Delobel, Recourt and
Gloaguen by permission of Elsevier Science Ltd. UK.

FIGURE 64. HRR measured with a cone calorimeter (heat flux 35 kW/m2) for various PP-
based RTPO materials: (a) PP-based RTPO (neat resin), (b) PP-based RTPO/clay

t
microcomposite, containing 10 wt % clay, (c) PP-based RTPO/PPgMA/clay nanocomposite,

ip
containing 10 wt % clay, 30 wt % PPgMA [185]. Reproduced from Hong, Lee, Bae, Jho,
Nam, Nam and Lee by permission of John Wiley & Sons, Inc.

cr
FIGURE 65. Effects of clay content on mass burning rate of PA6 (8 mm thick) at 50 kW/m2
[242]. Reproduced from Kashiwagi, Harris, Zhang, Briber, Cipriano, Ragharan, Awad and

us
Shields by permission of Elsevier Science Ltd. UK.

FIGURE 66. Effects of clay content on HRR of PA6 (8 mm thick) at 50 kW/m2 [242].

an
Reproduced from Kashiwagi, Harris, Zhang, Briber, Cipriano, Ragharan, Awad and Shields
by permission of Elsevier Science Ltd. UK.

FIGURE 67. Selected video images at 100, 200, and 400 s in nitrogen at 50 kW/m2 [242].
M
Reproduced from Kashiwagi, Harris, Zhang, Briber, Cipriano, Ragharan, Awad and Shields
by permission of Elsevier Science Ltd. UK.
d

FIGURE 68. A schematic view of the “radiative gasification” apparatus [35]. Reproduced
from Beyer by permission of Elsevier Science Ltd. UK.
te

FIGURE 69. Optical images of PMMA composites after the combustion of UL 94 V0 test: (a)
p

PMMA/DB/AO (75/20/5) and (b) PMMA/DB/AO/Cloisite 20A (70/20/5/5)


[246].Reproduced from Si, Zaitsev, Goldman, Frenkel, Peiffer, Weil, Sokolov and
ce

Rafailovich by permission of Elsevier Science Ltd. UK.

FIGURE 70. Shear viscosity as a function of shear rates [55]. Reproduced from Ray and
Ac

Okamoto by permission of Elsevier Science Ltd. UK.

FIGURE 71. DSC thermograms of PCL and of selected PCL/clay nanocomposites after
isothermal crystallization at 45 °C. Number refers to clay weight concentration; 0* is the DSC
thermogram of non-isothermal crystallized (-10 °C/min) pure PCL [255]. Reproduced from
Di Maio, Iannace, Sorrentini and Nicolais by permission of Elsevier Science Ltd. UK.

FIGURE 72. The relationship between OMMT content and Tg of composite [222].
Reproduced from Ke and Yongping by permission of Elsevier Science Ltd. UK.

FIGURE 73. Typical TEM images of PA6CN3.7 crystallized at (a) 170 and 210 °C. The
enlarged part shown (a) is form the indicated lamella in the original image. The black strip

Page 222 of 302


inside the white part is clay. Figure (b) shows the typical shish-kebab type of structure [257].
Reproduced from Maiti and Okamoto by permission of Wiley-VCH, Germany.

FIGURE 74. Schematic view of the nucleation and growth mechanism in nylon-6
nanocomposite [257]. Reproduced from Maiti and Okamoto by permission of Wiley-VCH,
Germany.

t
FIGURE 75. The heat flow versus the time during isothermal crystallization of nylon 1012

ip
(spectrum a) and nylon 1012/clay nanocomposite (spectrum b) by DSC [97]. Reproduced
from Wu, Zhou, Qi and Zhang by permission of John Wiley & Sons, Inc.

cr
FIGURE 76. (a) Real pictures of biodegradability of neat PLA and PLACN4 recovered from
compost with time. Initial shape of the crystallized samples was 3 x 10 x 0.1 cm3. (b) Time

us
dependence of residual weight, Rw and of matrix, Mw of PLA and PLACN4 under compost
[208]. Reproduced from Ray, Yamada, Okamoto and Ueda by permission of Elsevier Science

an
Ltd. UK.

FIGURE 77. (a) FT-IR spectra of PE/OMMT nanocomposites before and after 200 h
irradiation. (b). FT-IR spectra of pure PE before and after 200 h UV irradiation [261].
M
Reproduced from Huaili, Chungui, Shimin, Guangming and Mingshu by permission of
Elsevier Science Ltd. UK.
d

FIGURE 78. FT-IR spectra of PE/nanocomposite at carbonyl region during photodegradation


[261]. Reproduced from Huaili, Chungui, Shimin, Guangming and Mingshu by permission of
te

Elsevier Science Ltd. UK.


p

FIGURE 79. Possible growth paths for nanocomposites applications.


ce
Ac

Page 223 of 302


ed
pt
ce
Ac

Page 224 of 302


ce
Ac
Page 225 of 302
pt
ce
Ac
Page 226 of 302
c
Ac
Page 227 of 302
e
pt
ce
Ac

Page 228 of 302


pt
ce
Ac
Page 229 of 302
ed
pt
ce
Ac

Page 230 of 302


a
M
ed
pt
ce
Ac

Page 231 of 302


e
pt
ce
Ac

Page 232 of 302


e
pt
ce
Ac
Page 233 of 302
ed
pt
ce
Ac
Page 234 of 302
ed
pt
ce
Ac

Page 235 of 302


c
Ac
Page 236 of 302
ce
Ac
Page 237 of 302
ep
c
Ac
Page 238 of 302
p
ce
Ac Page 239 of 302
Ac
ce
pt
ed
M
an
us

Page 240 of 302


cr
i
Ac Page 241 of 302
ec
Ac
Page 242 of 302
e
pt
ce
Ac

Page 243 of 302


ce
Ac
Page 244 of 302
M
ed
pt
ce
Ac

Page 245 of 302


c
Ac Page 246 of 302
ce
Ac Page 247 of 302
ed
pt
ce
Ac

Page 248 of 302


ed
pt
ce
Ac

Page 249 of 302


ce
Ac Page 250 of 302
pt
ce
Ac
Page 251 of 302
a
M
ed
pt
ce
Ac

Page 252 of 302


pt
ce
Ac
Page 253 of 302
a
M
ed
pt
ce
Ac

Page 254 of 302


pt
ce
Ac
Page 255 of 302
us
an
M
ed
pt
ce
Ac

Page 256 of 302


te
ep
c
Ac
Page 257 of 302
e
pt
ce
Ac

Page 258 of 302


ce
Ac
Page 259 of 302
ed
pt
ce
Ac

Page 260 of 302


p
ce
Ac
Page 261 of 302
p
ce
Ac
Page 262 of 302
e
pt
ce
Ac

Page 263 of 302


a
M
ed
pt
ce
Ac

Page 264 of 302


ed
pt
ce
Ac

Page 265 of 302


c
Ac
Page 266 of 302
ce
Ac Page 267 of 302
t
ep
c
Ac

Page 268 of 302


pt
ce
Ac

Page 269 of 302


M
ed
pt
ce
Ac

Page 270 of 302


c
Ac
Page 271 of 302
pt
ce
Ac
Page 272 of 302
p
ce
Ac
Page 273 of 302
pt
ce
Ac
Page 274 of 302
t
ep
c
Ac
Page 275 of 302
pt
ce
Ac
Page 276 of 302
ed
pt
ce
Ac

Page 277 of 302


te
ep
c
Ac

Page 278 of 302


pt
ce
Ac
Page 279 of 302
a
M
ed
pt
ce
Ac

Page 280 of 302


M
ed
pt
ce
Ac

Page 281 of 302


t
ep
c
Ac
Page 282 of 302
e
pt
ce
Ac

Page 283 of 302


e
pt
ce
Ac

Page 284 of 302


ed
pt
ce
Ac

Page 285 of 302


ce
Ac Page 286 of 302
e
pt
ce
Ac

Page 287 of 302


e
pt
ce
Ac

Page 288 of 302


pt
ce
Ac

Page 289 of 302


M
ed
pt
ce
Ac

Page 290 of 302


e
pt
ce
Ac

Page 291 of 302


ed
pt
ce
Ac

Page 292 of 302


p
ce
Ac
Page 293 of 302
ep
c
Ac
Page 294 of 302
ce
Ac
Page 295 of 302
M
ed
pt
ce
Ac

Page 296 of 302


p
ce
Ac
Page 297 of 302
pt
ce
Ac
Page 298 of 302
u
an
M
ed
pt
ce
Ac

Page 299 of 302


an
M
de
pt
ce
Ac

Page 300 of 302


ed
pt
ce
Ac

Page 301 of 302


e
pt
ce
Ac
Page 302 of 302

You might also like