You are on page 1of 63

Materials Science and Engineering, 28 (2000) 1±63

Polymer-layered silicate nanocomposites: preparation, properties


and uses of a new class of materials
Michael Alexandre, Philippe Dubois*
Laboratory of Polymeric and Composite Materials, University of Mons-Hainaut, 20 Place du Parc, B-7000 Mons, Belgium

Accepted 20 March 2000

Abstract
This review aims at reporting on very recent developments in syntheses, properties and (future) applications
of polymer-layered silicate nanocomposites. This new type of materials, based on smectite clays usually rendered
hydrophobic through ionic exchange of the sodium interlayer cation with an onium cation, may be prepared via
various synthetic routes comprising exfoliation adsorption, in situ intercalative polymerization and melt
intercalation. The whole range of polymer matrices is covered, i.e. thermoplastics, thermosets and elastomers.
Two types of structure may be obtained, namely intercalated nanocomposites where the polymer chains are
sandwiched in between silicate layers and exfoliated nanocomposites where the separated, individual silicate layers
are more or less uniformly dispersed in the polymer matrix. This new family of materials exhibits enhanced
properties at very low filler level, usually inferior to 5 wt.%, such as increased Young's modulus and storage
modulus, increase in thermal stability and gas barrier properties and good flame retardancy. # 2000 Elsevier
Science S.A. All rights reserved.

Keywords: Layered silicate nanocomposites; Intercalative polymerization; Melt intercalation; Exfoliation±adsorption;


Mechanical properties; Thermal stability

1. Introduction

Manufacturers fill polymers with particles in order to improve the stiffness and the toughness of
the materials, to enhance their barrier properties, to enhance their resistance to fire and ignition or

Abbreviations: AFM, atomic force microscopy; AIBN, N,N0 -azobis(isobutyronitrile); ALA, aminolauric acid; APP,
ammonium polyphosphate; BDMA, benzyldimethylamine; BTFA, boron trifluoride monomethylamine; CEC, cation
exchange capacity; DGEBA, diglycidyl ether of bisphenol A; DMA, dynamic mechanical analysis; DSC, differential
scanning calorimetry; EDX, energy dispersive X-ray; EVA, ethylene vinyl acetate copolymer; FTIR, Fourier transform
infrared spectroscopy; HDPE, high density poly(ethylene); HPMC, hydroxypropylmethylcellulose; HRR, heat release rate;
MAO, methylaluminoxane; MMT, montmorillonite; NBR, nitrile rubber; NMA, nadic methyl anhydride; PAA,
poly(acrylic acid); PAN, poly(acrylonitrile); PANI, poly(aniline); PBD, poly(butadiene); PCL, poly(e-caprolactone);
PDDA, poly(dimethyldiallylammonium); PDMS, poly(dimethylsiloxane); PEO, poly(ethylene oxide); PFT, polymeriza-
tion-filling technique; PI, poly(imide); PLA, poly(lactide); PP, poly(propylene); PP-MA, maleic anhydride modified
poly(propylene); PP-OH, hydroxyl modified poly(propylene); PPV, poly(p-phenylenevinylene); PS, poly(styrene); PS3Br,
poly(3-bromostyrene); PVA, poly(vinyl acetate); PVCH, poly(vinylcyclohexane); PVOH, poly(vinyl alcohol); PVP,
poly(2-vinyl pyridine); PVPyr, poly(vinylpyrrolidone); PXDMS, poly(p-xylenylene dimethylsulfonium bromide); SBS,
symmetric(styrene±butadiene±styrene) block copolymer; SEC, size exclusion chromatography; TEM, transmission
electron microscopy; TEOS, tetraethylorthosilicate; THF, tetrahydrofuran; XRD, X-ray diffraction
*
Corresponding author. Tel.: ‡32-65-373481; fax: ‡32-65-373484.
E-mail address: philippe.dubois@umh.ac.be (P. Dubois)

0927-796X/00/$ ± see front matter # 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 9 2 7 - 7 9 6 X ( 0 0 ) 0 0 0 1 2 - 7
2 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 1
Example of layered host crystals susceptible to intercalation by a polymer
Chemical nature Examples
Element Graphite [8]
Metal chalcogenides (PbS)1.18(TiS2)2 [9], MoS2 [10]
Carbon oxides Graphite oxide [11,12]
Metal phosphates Zr(HPO4) [13]
Clays and layered silicates Montmorillonite, hectorite, saponite, fluoromica, fluorohectorite, vermiculite, kaolinite,
magadiite, . . .
Layered double hydroxides M6Al2(OH)16CO3nH2O; MˆMg [14], Zn [15]

simply to reduce cost. Addition of particulate fillers sometimes imparts drawbacks to the resulting
composites such as brittleness or opacity.
Nanocomposites are a new class of composites, that are particle-filled polymers for which at
least one dimension of the dispersed particles is in the nanometer range. One can distinguish three
types of nanocomposites, depending on how many dimensions of the dispersed particles are in the
nanometer range. When the three dimensions are in the order of nanometers, we are dealing with
isodimensional nanoparticles, such as spherical silica nanoparticles obtained by in situ sol±gel
methods [1,2] or by polymerization promoted directly from their surface [3], but also can include
semiconductor nanoclusters [4] and others [2]. When two dimensions are in the nanometer scale and
the third is larger, forming an elongated structure, we speak about nanotubes or whiskers as, for
example, carbon nanotubes [5] or cellulose whiskers [6,7] which are extensively studied as
reinforcing nanofillers yielding materials with exceptional properties. The third type of
nanocomposites is characterized by only one dimension in the nanometer range. In this case the
filler is present in the form of sheets of one to a few nanometer thick to hundreds to thousands
nanometers long. This family of composites can be gathered under the name of polymer-layered
crystal nanocomposites, and their study will constitute the main object of this contribution. These
materials are almost exclusively obtained by the intercalation of the polymer (or a monomer
subsequently polymerized) inside the galleries of layered host crystals. There is a wide variety of
both synthetic and natural crystalline fillers that are able, under specific conditions, to intercalate a
polymer. Table 1 presents a non-exhaustive list of possible layered host crystals.
Amongst all the potential nanocomposite precursors, those based on clay and layered silicates
have been more widely investigated probably because the starting clay materials are easily available
and because their intercalation chemistry has been studied for a long time [16,17]. Owing to the
nanometer-size particles obtained by dispersion, these nanocomposites exhibit markedly improved
mechanical, thermal, optical and physico-chemical properties when compared with the pure polymer
or conventional (microscale) composites as firstly demonstrated by Kojima and coworkers [18] for
nylon±clay nanocomposites. Improvements can include, for example, increased moduli, strength and
heat resistance, decreased gas permeability and flammability.
The aim of this report is to review the different techniques used to obtain polymer-layered
silicates nanocomposites and the improved properties that those materials can display.

2. Generalities

2.1. Structure of layered silicates

The layered silicates commonly used in nanocomposites belong to the structural family known
as the 2:1 phyllosilicates. Their crystal lattice consists of two-dimensional layers where a central
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 3

Fig. 1. Structure of 2:1 phyllosilicates (reproduced from [19] with permission).

octahedral sheet of alumina or magnesia is fused to two external silica tetrahedron by the tip so that
the oxygen ions of the octahedral sheet do also belong to the tetrahedral sheets. The layer thickness
is around 1 nm and the lateral dimensions of these layers may vary from 300 A Ê to several microns
and even larger depending on the particular silicate. These layers organize themselves to form stacks
with a regular van der Walls gap in between them called the interlayer or the gallery. Isomorphic
substitution within the layers (for example, Al3‡ replaced by Mg2‡ or by Fe2‡, or Mg2‡ replaced by
Li‡) generates negative charges that are counterbalanced by alkali or alkaline earth cations situated
in the interlayer. As the forces that hold the stacks together are relatively weak, the intercalation of
small molecules between the layers is easy [16]. In order to render these hydrophilic phyllosilicates
more organophilic, the hydrated cations of the interlayer can be exchanged with cationic surfactants
such as alkylammonium or alkylphosphonium (onium). The modified clay (or organoclay) being
organophilic, its surface energy is lowered and is more compatible with organic polymers. These
polymers may be able to intercalate within the galleries, under well defined experimental conditions
as will be reported about in Section 3.
Montmorillonite, hectorite and saponite are the most commonly used layered silicates. Their
structure is given in Fig. 1 [19] and their chemical formula are shown in Table 2.
This type of clay is characterized by a moderate negative surface charge (known as the cation
exchange capacity, CEC and expressed in meq/100 g). The charge of the layer is not locally constant
as it varies from layer to layer and must rather be considered as an average value over the whole
crystal. Proportionally, even if a small part of the charge balancing cations is located on the external
crystallite surface, the majority of these exchangeable cations is located inside the galleries. When
the hydrated cations are ion-exchanged with organic cations such as more bulky alkyammoniums, it
usually results in a larger interlayer spacing.

Table 2
Chemical structure of commonly used 2:1 phyllosilicatesa
2:1 Phyllosilicate General formula
Montmorillonite Mx(Al4ÿxMgx)Si8O20(OH)4
Hectorite Mx(Mg6ÿxLix)Si8O20(OH)4
Saponite MxMg6(Si8ÿxAlx)O20(OH)4
a
Mˆmonovalent cation; xˆdegree of isomorphous substitution (between 0.5 and 1.3).
4 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 2. Alkyl chain aggregation in layered silicates: (a) lateral monolayer; (b) lateral bilayer; (c) paraffin-type monolayer
and (d) paraffin-type bilayer (reproduced from [21] with permission).

In order to describe the structure of the 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 preferentially resides at the layer surface, leaving the organic tail radiating away from the
surface. In a given temperature range, two parameters then define the equilibrium layer spacing: the
cation exchange capacity of the layered silicate, driving the packing of the chains, and the chain
length of organic tail(s). According to X-ray diffraction (XRD) data, the organic chains have been
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 surface, forming mono or even
bimolecular tilted `paraffinic' arrangement [20] as shown in Fig. 2.
A more realistic description has been proposed by Vaia et al. [21], based on FTIR experiments.
By monitoring frequency shifts of the asymmetric CH2 stretching and bending vibrations, they found
that the intercalated chains exist in states with varying degrees of order. In general, as the interlayer
packing density or the chain length decreases (or the temperature increases), the intercalated chains
adopt a more disordered, liquid-like structure resulting from an increase in the gauche/trans
conformer ratio. 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. 3).
This interpretation has been recently confirmed by molecular dynamics simulations where a
strong layering behavior with a disordered liquid-like arrangement has been found, that can evolve
towards a more ordered arrangement by increasing the chain length [22]. As the chain length

Fig. 3. 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 environment (reproduced from [21] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 5

Fig. 4. Scheme of different types of composite arising from the interaction of layered silicates and polymers: (a) phase-
separated microcomposite; (b) intercalated nanocomposite and (c) exfoliated nanocomposite.

increases, the interlayer structure appears to evolve in a stepwise fashion, from a disordered to more
ordered monolayer then `jumping' to a more disordered pseudo-bilayer.

2.2. Nanocomposite structures

Depending on the nature of the components used (layered silicate, organic cation and polymer
matrix) and the method of preparation, three main types of composites may be obtained when a
layered clay is associated with a polymer (Fig. 4).
When the polymer is unable to intercalate between the silicate sheets, a phase separated
composite (Fig. 4a) is obtained, whose properties stay in the same range as traditional
microcomposites. Beyond this classical family of composites, two types of nanocomposites can
be recovered. Intercalated structure (Fig. 4b) in which a single (and sometimes more than one)
extended polymer chain is intercalated between the silicate layers resulting in a well ordered
multilayer morphology built up with alternating polymeric and inorganic layers. When the silicate
layers are completely and uniformly dispersed in a continuous polymer matrix, an exfoliated or
delaminated structure is obtained (Fig. 4c). Two complementary techniques are used in order to
characterize those structures. XRD is used to identify intercalated structures. In such nanocompo-
sites, the repetitive multilayer structure is well preserved, allowing the interlayer spacing to be
determined. The intercalation of the polymer chains usually increases the interlayer spacing, in
comparison with the spacing of the organoclay used (Fig. 5), leading to a shift of the diffraction peak
towards lower angle values (angle and layer spacing values being related through the Bragg's
relation: lˆ2d sin y, where l corresponds to the wave length of the X-ray radiation used in the
diffraction experiment, d the spacing between diffractional lattice planes and y is the measured
diffraction angle or glancing angle).
As far as exfoliated structure is concerned, no more diffraction peaks are visible in the XRD
diffractograms either because of a much too large spacing between the layers (i.e. exceeding 8 nm in
the case of ordered exfoliated structure) or because the nanocomposite does not present ordering
6 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 5. XRD patterns of: (a) phase separated microcomposite (organo-modified fluorohectorite in a HDPE matrix); (b)
intercalated nanocomposite (same organomodified fluorohectorite in a PS matrix) and (c) exfoliated nanocomposite (the
same organo-modified fluorohectorite in a silicone rubber matrix) (reproduced from [19] with permission).

anymore. In the latter case, transmission electronic spectroscopy (TEM) is used to characterize the
nanocomposite morphology. Fig. 6 shows the TEM micrographs obtained for an intercalated and an
exfoliated nanocomposite. Besides these 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 observed and one must rely on TEM observation to define the overall
structure.

Fig. 6. TEM micrographs of poly(styrene)-based nanocomposites: (a) intercalated nanocomposite (reproduced from [60]
with permission) and (b) exfoliated nanocomposite (reproduced from [61] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 7

3. Nanocomposite preparation

Several strategies have been considered to prepare polymer-layered silicate nanocomposites.


They include four main processes [23]:
 Exfoliation±adsorption: The layered silicate is exfoliated into single layers using a solvent in
which the polymer (or a 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. The polymer then adsorbs onto the delaminated
sheets and when the solvent is evaporated (or the mixture precipitated), the sheets reassemble,
sandwiching the polymer to form, in the best case, an ordered multilayer structure. Under this
process are also gathered the nanocomposites obtained through emulsion polymerization where
the layered silicate is dispersed in the aqueous phase.
 In situ intercalative polymerization: In this technique, the layered silicate is swollen within the
liquid monomer (or a monomer solution) so as the polymer formation can occur in between the
intercalated sheets. Polymerization can be initiated either by heat or radiation, by the diffusion of a
suitable initiator or by an organic initiator or catalyst fixed through cationic exchange inside the
interlayer before the swelling step by the monomer.
 Melt intercalation: The layered silicate is mixed with the polymer matrix in the molten state.
Under these conditions and if the layer surfaces are sufficiently compatible with the chosen
polymer, the polymer can crawl into the interlayer space and form either an intercalated or an
exfoliated nanocomposite. In this technique, no solvent is required.
 Template synthesis: This technique, where the silicates are formed in situ in an aqueous solution
containing the polymer and the silicate building blocks has been widely used for the synthesis of
double-layer hydroxide-based nanocomposites [14,15] but is far less developed for layered
silicates. In this technique, based on self-assembly forces, the polymer aids the nucleation and
growth of the inorganic host crystals and gets trapped within the layers as they grow. The
following sections review the four aforementioned preparation techniques, that will be illustrated
with representative examples.

3.1. Exfoliation±adsorption

3.1.1. Exfoliation±adsorption from polymers in solution


This technique has been widely used with water-soluble polymers to produce intercalated
nanocomposites [24,25] based on poly(vinyl alcohol) (PVOH) [26,27], poly(ethylene oxide) (PEO)
[27±31], poly(vinylpyrrolidone) (PVPyr) [32] or poly(acrylic acid) (PAA) [31]. When polymeric
aqueous solutions are added to dispersions of fully delaminated sodium layered silicates, the strong
interaction existing between the hydrosoluble macromolecules and the silicate layers often trigger
the reaggregation of the layers as it occurs for PVPyr [32] or PEO [27]. In the presence of PVOH, the
layers remain in colloidal distribution [27]. In the wet state or after mild drying (air drying), the
silicate layers are distributed and embedded in the so-obtained PVOH gel. This state actually
corresponds to a true nanocomposite hybrid material. However, more intense drying of the PVOH
gel in vaccuo causes part of the silicate layers to reaggregate and intercalated species are formed.
This is indicated by a basal spacing of 1.36 nm, corresponding to the intercalation of a polymeric
monolayer in between the silicate layers. In fact, sterical constraints from the PVOH matrix impede
reaggregation of all the silicate layers and some of them remain exfoliated [33]. Interestingly,
polymers intercalation using the so-called exfoliation±adsorption technique can also be performed in
8 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 7. X-ray diffractograms of modified montmorillonite, HDPE and nitrile copolymer composite systems (* Ð a high
degree of crystallinity of the HDPE is evident) (reproduced from [35] with permission).

organic solvents. PEO has been successfully intercalated in sodium montmorillonite and sodium
hectorite by dispersion in acetonitrile [34], allowing to stoichiometrically incorporate one or two
polymer chains in between the silicate layers and increasing the intersheet spacing from 0.98 to 1.36
and 1.71 nm, respectively. Study of the chain conformation using two-dimensional double-quantum
NMR on 13 C enriched PEO intercalated in sodium hectorite [10] reveals that the conformation of the
`±OC±CO±' bonds of PEO is 905% gauche, inducing constraints on the chain conformation in the
interlayer.
Jeon and coworkers [35] have investigated this technique in attempts to produce
nanocomposites with nitrile-based copolymer (Barex 210 E) and polyethylene-based polymer.
These nanocomposites were filled with sodium montmorillonite previously modified by a protonated
dodecylamine as the organic cation. Upon treatment, the interlayer spacing increased from 11.8 to
16.5 AÊ , attesting for effective cation exchange. In order to produce the nitrile copolymer-based
nanocomposite, the copolymer was dissolved in dimethylformamide in the presence of 15 wt.%
modified clay. After solvent evaporation in a vacuum oven at 808C for 24 h the film recovered was
characterized by both XRD and TEM. XRD reveals a broad diffraction peak that has been shifted
towards a higher interlayer spacing (21.5 A Ê , see Fig. 7). The large broadening of the peak may
indicate that partial exfoliation has occurred, as corroborated by TEM analysis (Fig. 8) where both
stacked (intercalated) and isolated (exfoliated) silicate layers can be observed.
High density polyethylene (HDPE)-based nanocomposite has been produced by using a similar
technique where the polyolefinic chains were dissolved in a mix of xylene and benzonitrile
(80:20 wt.%) with 20 wt.% modified clay dispersed within. The composite material was then
recovered by precipitation from tetrahydrofuran (THF) followed by several washings with THF. As
seen in Fig. 7, the small increase in the interlayer spacing could account for some intercalation even
if the TEM observation let only show small stacks of flake-like particles. Even if conducted under
similar experimental conditions, these two syntheses indicate that the exfoliation±adsorption
technique can provide quite different results much depending upon the polymer matrix. In other
words, it does mean that for every type of polymer, one has to find the right layered clay, organic
modifier and solvent(s).
Ogata et al. applied the exfoliation±adsorption method for the production of poly(lactide) (PLA)
[36] and poly(e-caprolactone) (PCL) biodegradable nanocomposites [37] using montmorillonite
modified with distearyldimethylammonium cations. The composites were prepared by dissolving
either PLA or PCL in hot chloroform in presence of a given amount of the modified clay, then
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 9

Fig. 8. TEM micrograph of the nitrile copolymer filled with 15 wt.% of organo-modified montmorillonite. Iˆindividual
silicate layer and Sˆstacked silicate layers (reproduced from [35] with permission).

vaporizing the solvent to obtain homogeneous films. However, under those conditions, it was found
that no intercalation took place in the presence whatever polyester. It is worth to point out that the
organo-modified clay rather formed a remarkable geometric structure in the filled polymers where
tactoids consisting of several silicate monolayers form a superstructure in the thickness direction of
the film. Such structural features have been found on one hand to substantially increase the Young's
modulus of the PLA-based composites (which is almost doubled with 5 wt.% of organo-modified
clay) and on the other hand, to enhance both storage and loss moduli determined by dynamic
mechanical analysis (DMA) carried out on the organoclay-filled PCL.

3.1.2. Exfoliation±adsorption from prepolymers in solution


Some polymeric materials such as poly(imides) or some conjugated polymers have the
particular property of being infusible and insoluble in organic solvents. Therefore, the only possible
route to produce nanocomposites with these types of polymers consists in using soluble polymeric
precursors that can be intercalated in the layered silicate and then thermally or chemically converted
in the desired polymer. This has been successfully achieved by using the exfoliation±adsorption
process.
The Toyota Research group has been the first to use this method to produce poly(imide) (PI)
nanocomposites [38]. The polyimide±montmorillonite nanocomposite has been synthesized by
mixing in dimethylacetamide a modified montmorillonite with the poly(imide) precursor, that is a
poly(amic acid) obtained from the step polymerization of 4,40 -diaminodiphenyl ether with
pyromellitic dianhydride. The organo-modified montmorillonite was prepared by previous
intercalation with dodecylammonium hydrochloride. After elimination of the solvent, an organoclay
filled poly(amic acid) film was recovered, which was thermally treated up to 3008C in order to
trigger the imidization reaction and to produce the poly(imide) nanocomposite. The XRD patterns of
these filled PI films do not show any diffraction peak typical of an intercalated morphology leading
the authors to conclude to the formation of an exfoliated structure and explaining the excellent gas
10 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 3
Nature, CEC and length of the layered silicates used in the synthesis of polyimide as presented in [39]
Layered silicate CEC (meq/100 g) Ê )a
Length of dispersed particles (A
‡
Hectorite Na 55 460
Saponite Na‡ 100 1650
Montmorillonite Na‡ 119 2180
Synthetic mica Na‡ 119 12300
a
Longer particle dimension as determined by TEM observation.

barrier properties of the resulting films (see Section 4.3). This experiment has been extended to other
layered silicates (hectorite, saponite and synthetic mica) with different aspect ratios [39] (Table 3).
X-ray diffractograms of the obtained polyimide-based nanocomposites again show no
noticeable peak indicating an exfoliated structure for the montmorillonite and the synthetic mica.
For both hectorite and saponite, a broad peak, centered on a value of 15 A Ê is observed, indicating that
for those layered silicates, polymer intercalation occurs probably together with some exfoliation. For
saponite, the measured interlayer spacing is even smaller than the value measured for the starting
organically modified clay (18 A Ê ) suggesting that the organic cation could have been expelled from
the clay interlayer during imidization reaction. The same phenomenon has been observed by Lan
et al. [40] when studying the effect of the chain length of the organic cation in the preparation of PI
nanocomposites by the same synthetic methodology. Starting from sodium montmorillonite with a
CEC of 92 meq/100 g and various protonated linear primary alkylamine, i.e. CH3±(CH2)nÿ1NH3‡
(where nˆ4, 8, 12, 16 and 18), they obtained, after imidization by curing at 3008C, composites
showing the same interlayer spacing of 13.2 A Ê , independently of the chain length of the intercalated
alkylammonium cation (Fig. 9).

Fig. 9. XRD patterns of polymer/CH3(CH2)nÿ1NH3‡ modified montmorillonite composites (clay loading: 10 wt.%): (A)
air-dried poly(amic acid) films and (B) poly(imide) films cured at 3008C (reproduced from [40] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 11

This value again corresponds to the intercalation of the PI chains adopting a flattened
conformation within the galleries, and the removal of the organic cation out of the interlayer. As the
same is achieved at much lower temperature (drying at 1008C), this structural change cannot be
related to the thermal degradation of the onium ion still stable in this temperature range. Further
experimental evidence for the cation eviction out of the clay interlayers comes from the fact that the
13.2 AÊ interlayer spacing does not change when the material is heated up to at 4508C, a temperature
where alkylammonium cations are usually degraded. No clear explanation for the expulsion of the
alkylammonium ions has been reached.
The presence of organically modified layered silicate such as montmorillonite previously
modified with protonated p-phenylene diamine has also shown to improve the kinetics of the
imidization reaction, allowing for a reduction of both the imidization temperature and reaction time
[41]. The activation energy of the imidization reaction (based on a first order kinetics), monitored by
FTIR spectroscopy, is shown to drop down by ca. 20% in presence of 7 wt.% of organoclay. A
reaction mechanism has been tentatively proposed, involving the modified silicate layers as active
partners in the imidization process (Fig. 10).
It has to be noted that a complete exfoliated structure has been observed for those nanomaterials
by both XRD and TEM.
Conjugated polymers are another family of polymers prone to be intercalated through this two-
step technique. Oriakhi et al. [42] have elegantly shown that the exfoliation±adsorption method could
be explored to prepare nanocomposites with poly(p-phenylenevinylene) (PPV) as the continuous
polymeric matrix. The polymer precursor to be intercalated was the poly(p-xylenylene
dimethylsulfonium bromide) (PXDMS). The PXMDS-montmorillonite layered nanocomposite was
accordingly prepared by reaction of a colloidal dispersion of Na-montmorillonite with an aqueous
solution of PXMDS at 08C. As the precursor bears two cationic sites, it readily intercalates between
the montmorillonite sheets by cationic exchange. The precursor is then chemically transformed into
PPV by a base-mediated elimination of dimethylsulfide and HBr. This is achieved by stirring the
crude product with 20% ethanolic NaOH solution at ambient temperature for 48 h. XRD patterns
before and after chemical conversion give interlayer spacing of 15.1 and 14.6 A Ê , respectively.
Compared to the initial interlayer spacing of the Na-montmorillonite (9.6 A Ê ), it indicates a gallery
Ê
expansion of, respectively, 5.5 and 5.0 A, consistent with the expected dimensions for a polymer

Fig. 10. A possible reaction mechanism for involving silicate layers in the imidization process (reproduced from [41] with
permission).
12 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

monolayer with the phenyl rings oriented perpendicular to the layered silicate surfaces. This gallery
expansion together with the absence of any XRD peak related to the interlayer spacing of Na-
montmorillonite attest for the formation of an intercalated nanocomposite. The presence of a weak
absorption bandpeak in the sp3 C±H region of the FTIR spectrum (2890 cmÿ1), however, suggests
that the elimination of the dimethylsulfonium groups is not quantitative and that the incorporated
polymer could be in fact a copolymer of PXMDS and PPV.

3.1.3. Exfoliation±adsorption by emulsion polymerization


Emulsion polymerization has been also studied in order to promote the intercalation of water
insoluble polymers within Na-montmorillonite that is well known to readily delaminate in water
[43±45]. Poly(methyl methacrylate) (PMMA) was first tested by this method [43]. The emulsion
polymerization was thus carried out in water in the presence of various amounts of the layered
silicate. The previously distilled methyl methacrylate monomer (MMA) was dispersed in the
aqueous phase with the aid of sodium lauryl sulfate as a surfactant. Polymerization was conducted at
708C for 12 h by using potassium persulfate as the free-radical initiator. The obtained latex is then
coagulated with an aluminum sulfate solution, filtered and dried under reduced pressure. The
obtained composites were extracted with hot toluene for 5 days by means of Soxhlet extraction.
Contents of intercalated polymer were determined for both extracted and non-extracted materials and
are given in Table 4 together with the molecular weights and polydispersities of the extracted
polymers.
These results demonstrate that part of the PMMA chains stay immobilized onto and/or inside
the layered silicates and cannot be extracted. This is further confirmed by FTIR of the extracted
composite that shows the absorption bands typical of PMMA chains. It can be observed that the
relative content of clay does not substantially modify the PMMA molecular weights (Mw), the value
of which is quite comparable to the Mw of PMMA polymerized in absence of clay (entry 1, Table 4).
Clearly, the presence of layered silicates does not seem to perturb the free-radical polymerization.
Intercalation is evidenced by XRD where an increase in the interlayer distance of about 5.5 A Ê is
observed for both PMMA 10, 20 and 30. This increase relatively well correlates with the thickness of
the polymer chain in its extended form. DSC data obtained for the extracted nanocomposites does
not show any glass transition, in accordance with what is usually observed for intercalated polymers.
Ion±dipole interactions are believed to be the driving force for the immobilization of the organic
polymer chains lying flat onto the layered silicate surface. The same methodology has been also
applied to produce montmorillonite intercalated with poly(styrene) (PS) [45]. The nanocomposite

Table 4
Montmorillonite feed ratios, PMMA contents in non-extracted and extracted composites, average molecular weights and
polydispersities of extracted PMMAs
Sample Feed ratio of PMMA content (wt.%)a Mn10ÿ3 Mw10ÿ3 Mw/Mn
MMA/clay (g/g) Non-extracted Extracted b (g/mol) (g/mol)

PMMA 100/0 ± ± 23 160 6.6


PMMA10 100/10 87.4 58.7 44 250 5.8
PMMA20 100/20 79.3 49.6 60 200 3.4
PMMA30 100/30 60.4 33.4 63 150 2.4
PMMA40 100/40 58.6 22.8 82 390 4.8
PMMA50 100/50 46.1 18.4 38 290 7.6
a
As determined by TGA.
b
Composite recovered after Soxhlet extraction in toluene for 5 days.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 13

Table 5
Montmorillonite feed ratios, PS contents in non-extracted and extracted composites and interlayer distances in PS-based
nanocomposites obtained by emulsion polymerization
Sample Feed ratio styrene/MMT (g/g) PS content (wt.%)a Ê)
Interlayer distance (A
b
Non-extracted Extracted
MMT 0/100 ± ± 9.8
PS5 95/5 88.7 45.6 15.5
PS10 90/10 82.8 33.9 14.6
PS20 80/20 75.4 28.8 13.8
PS30 70/30 65.9 21.7 12.4
a
As determined by TGA.
b
Composite recovered after Soxhlet extraction in toluene for 5 days.

syntheses were essentially comparable to MMA emulsion polymerization except that the Na-
montmorillonite was sonicated prior to polymerization. Results are presented in Table 5.
Here again, a true intercalated structure is formed with an interlayer spacing that changes with
the PS content. The interlayer distance slightly decreases at higher montmorillonite feed ratios.
Similarly to MMA polymerization, the molecular weight of the PS recovered fraction does not seem
to be affected by the presence of the dispersed clay. DSC thermograms also attest for the
intercalation of PS in an extended form as no more glass transition can be observed in the extracted
nanocomposites. However, in the case of filled PS, the stabilization of the PS chains in the interlayer
cannot be accounted for by ion±dipole interactions anymore. Rather, the authors propose the
cooperative formation of ion-induced dipole interactions. Note finally a report on the formation of
epoxy-montmorillonite intercalated composites by emulsion polymerization [44]. Again an increase
of the interlayer spacing of about 6 A Ê is observed. But contrary to observations achieved for both
PMMA and PS, the epoxy content in the extracted composites appears to increase with the
montmorillonite content, indicating that the layered silicates possibly participate in the
polymerization reaction.

3.2. In situ intercalative polymerization

3.2.1. Thermoplastic nanocomposites


Many interlamellar polymerization reactions were studied in the 1960s and the 1970s using
layered silicates (see [25,46] and references therein) but it is with the work initiated by the Toyota
research team [47,48] that the study of polymer-layered silicate nanocomposites came into vogue
about 10 years ago. They studied the ability of Na-montmorillonite organically modified by
protonated a,o-aminoacid (‡H3N±(CH2)nÿ1±COOH, with nˆ2, 3, 4, 5, 6, 8, 11, 12, 18) to be
swollen by the e-caprolactam monomer (melting temperatureˆ708C) at 1008C and subsequently to
initiate its ring opening polymerization to obtain nylon-6-based nanocomposites [48,49]. A clear
difference occurs in the swelling behavior between the montmorillonite with relatively short (n<11)
and longer alkyl chains as depicted in Table 6, indicating that a larger amount of monomer can be
intercalated for longer alkyl chains.
The 12-aminolauric acid (nˆ12) modified montmorillonite was chosen to develop the
intercalative ring opening polymerization of e-caprolactam [48]. In a typical synthesis, the modified
montmorillonite (12-Mont) was mixed with the monomer in a mortar. A small amount of 6-
aminocaproic acid was added as a polymerization accelerator when the relative amount of 12-Mont
used was smaller than 8 wt.% (relative to 12-Mont). The mixture was heated first at 1008C for
14 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 6
Basal spacing on organo-modified montmorillonite in the presence of e-caprolactam (e-CLa) at 1008C
‡
H3N±(CH2)nÿ1±COOH (n) Interlayer spacing of the Interlayer spacing when swollen
modified clay (AÊ) by (e-CLa) at 1008C (AÊ)
2 12.7 14.4
3 13.1 19.7
4 13.2 19.9
5 13.2 20.4
6 13.2 23.4
8 13.4 26.4
11 17.4 35.7
12 17.2 38.7
18 28.2 71.2

30 min then at 2508C for 6 h. The cooled and solidified product was crushed, washed with water at
808C, and then dried. Depending on the amount of 12-Mont introduced, either exfoliated (for less
then 15 wt.%) or delaminated structure (from 15 to 70 wt.%) were obtained as evidenced by XRD
and TEM measurements. Comparison of the titrated amount 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-Mont
surface are responsible for the polymerization initiation. Moreover, approximately all the `‡ NH3 '
end groups present in the matrix should be interacting with the montmorillonite anions. Finally, the
ratio of bonded to non-bonded polymer chains increased with the amount of incorporated
montmorillonite (from 32.3% of bonded chains for 1.5 wt.% montmorillonite to 92.3% of bonded
chains for 59.6 wt.% clay). Further works [50] have demonstrated that intercalative polymerization
of e-caprolactam could be realized without the necessity to render the montmorillonite surface
organophilic. Indeed, this monomer was able to directly intercalate the Na-montmorillonite in water,
in the presence of hydrochloric acid. This intercalation was proved by the increase in interlayer
spacing observed on the isolated montmorillonite/e-caprolactam product, going from 10 A Ê for the
neat Na-montmorillonite to 15.1 A Ê when e-caprolactam intercalates. At high temperature (2008C) in
the presence of excess e-caprolactam, the so modified clay can be swollen again, allowing for the
ring opening polymerization to proceed at 2608C when 6-aminocaproic acid is added as an
accelerator. The resulting composite does not present any diffraction peak characteristic of an
interlayer spacing in XRD and TEM observation agrees with a molecular dispersion of the silicate
sheets. In attempts to carry out the synthesis in one-pot [51], the system has proved to be sensitive to
the nature of the acid used to promote the intercalation of e-caprolactam. Table 7 gives the results

Table 7
Peak intensity (Im) and interlayer spacing (d) of nylon-6-based nanocomposites prepared in presence of different acid
derivatives by the one-pot technique
Acid Ima (cps) Ê)
d (A
Phosphoric acid 0 0
Hydrochloric acid 200 21.7
Isophtalic acid 255 20.2
Benzenesulfonic acid 280 19.3
Acetic acid 555 20.3
Trichloroacetic acid 585 21.3
No acid 1840 18.6
a
As defined in Fig. 11.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 15

Fig. 11. XRD intensity curve of injection molded nylon-6 nanocomposite as obtained by the one-pot intercalation
polymerization process in the presence of acetic acid (reproduced from [51] with permission).

obtained for different acids in relation to the intensity (Im) of the XRD intercalation peak that might
be present in the obtained nanocomposites (Fig. 11).
These results show that only phosphoric acid allows for the preparation of a truly exfoliated
nanocomposite, the other acids tending to promote the formation of partially exfoliated-partially
intercalated structures. No clear reason indicates why only phosphoric acid works. One can also
point out that an intercalated structure is obtained even if no acid is added on purpose. Actually, the
addition of 6-aminocaproic acid in each experiment as polymerization accelerator could have played
the same role.
Another polyamide, i.e. nylon-12, is reported to form nanocomposites using the in situ
intercalative polymerization. Indeed, Reichert et al. [52] have used 12-aminolauric acid (ALA) as
both the layered silicate modifier and the monomer. They first studied by XRD the effect of ALA on
the swelling behavior of a synthetic three-layer silicate (commercial SOMASIF ME100, a
fluorinated silicate obtained by heating talcum and Na2SiF6 at high temperature for several hours).
Increasing amounts of ALA dispersed in a constant HCl volume (20 mmol/l) were poured into a
water suspension of ME100. The swelling process in function of ALA concentration was monitored
and it can be separated in two regimes: a cation-exchange of inorganic cations by protonated ALA at
low ALA concentration giving an interlayer distance of 17 A Ê and a further diffusion of zwitterionic
12-aminolauric acid into the interlayer spacing when the ALA concentration exceeds the amount of
HCl in the medium (interlayer spacing over 20 A Ê , Figs. 12 and 13).
The swelling was found to be independent of both the swelling temperature, the layered silicate
concentration and the type of mineral acid used to protonate ALA (HCl, H2SO4, H3PO4). ALA was

Fig. 12. Interlayer distance of fluoro-modified talc (ME 100) in function of an increasing amount of aminolauric acid used
as the organic modifier (reproduced from [52] with permission).
16 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 13. Schematic representation of the swelling behavior of the fluoro-modified talc ME 100 in presence of aminolauric
acid (reproduced from [52] with permission).

then polymerized at high temperature (2808C) and under elevated pressure (up to 20 bar) during
9.5 h with both types of swollen clay. XRD, TEM, Scanning TEM coupled with energy dispersive
X-ray (EDX) as well as atomic force microscopy (AFM) have been used to characterize the resulting
composites. They all confirm the structure as being partially exfoliated and otherwise intercalated
nanocomposites.
By using the method developed by Usuki et al. [48] for the polymerization of e-caprolactam
(vide supra), Messersmith and Giannelis [53] have modified a Na-montmorillonite by the protonated
aminolauric acid and dispersed this modified clay in liquid e-caprolactone before polymerizing it at
high temperature. The poly(e-caprolactone) (PCL) nanocomposites were prepared by mixing up to
30 wt.% of the modified clay with dried and freshly distilled e-caprolactone at room temperature for
a couple of hours followed by the ring opening polymerization under stirring at 1708C for 48 h. XRD
patterns of the modified clay after contact with e-caprolactone at room temperature do not show any
significant increase in the layered spacing (13.6 A Ê ). The authors assumed that the monomer
intercalates in the gaps between the aminolauric acid chains so that no gallery expansion could be
seen. This is in contrast with what is usually observed in in situ intercalative polymerization where
the insertion of the monomer within the silicate gallery induces an increase in the interlayer spacing.
Another possibility may be that intercalation of the monomer occurs only during the heating step of
the solution. After polymerization, XRD patterns of the obtained composites did not show any
diffraction peak indicating the presence of intercalated structures, more likely indicating that
exfoliation occurred. This exfoliation induces some clear modifications in the onset of polymer
melting (Tm) of the obtained nanocomposites that appears at lower temperature when the weight
fraction of layered silicates increases (Fig. 14).
This phenomenon is attributed to the formation of smaller crystallites, probably due to the
physical barrier effect of the dispersed layers that limits the extent of crystallization of the PCL
chains. The authors propose a polymerization mechanism where the carboxylic acid of the attached
aminolauric acid undergoes a nucleophilic addition on the e-caprolactone carbonyl function, creating
by ring opening reaction a carboxylic anhydride bridge that links the first opened lactone monomer
to the silicate surface. The ring opening reaction is then proposed to occur via an `o-acyl' cleavage of
the monomer with the formation of an hydroxyl end group that propagates the lactone
polymerization. However, it is now commonly accepted [54] that carboxylic species are not
reactive enough to promote the ring-opening polymerization of e-caprolactone. At such a high
temperature (1708C), a mechanism known as the `active hydrogen polymerization' should rather take
place where traces of water could initiate the ring opening polymerization of the lactone [55]. This
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 17

Fig. 14. Polymer onset melting temperature as a function of organomodified montmorillonite for poly(e-caprolactone)
(PCL)-based nanocomposites. PCLA corresponds to the melting temperature for unfilled PCL (reproduced from [53] with
permission).

mechanism ruled out the grafting of the growing polyester chains onto the silicate layers through the
formation of a carboxylic anhydride link. If any links exist (as suggested by indirect evidence such as
the broadening of the solution 1 H-NMR spectra of the composites when the amount of layered
silicates increases or by the possibility to recover the PCL chains free of silicates after cation
exchange with Li‡), they would rather be formed by a post-esterification reaction at high
temperature between the terminal hydroxyl group of some PCL chains with the carboxylic acid
function of the fixed aminoacid surface agent. Note finally that e-caprolactone ring opening
polymerization could be also initiated by primary amino end-groups of some aminolauric acids that
did not react with the layered silicate surface. Indeed, Rozenberg has claimed that e-caprolactone
polymerization initiated with amines in the presence of a protic acid at a high temperature (1808C)
can proceed through a zwitterionic mechanism [56]. The resulting macromolecular zwitterions, i.e.
a-H3N‡, o-CO2ÿ PCL chains could then interact with the charged layered silicates, assuring some
efficient surface grafting.
The above poly(e-caprolactone)-based nanocomposite synthesis has been recently applied by
Chen et al. [57] to produce novel segmented polyurethane/clay nanocomposites articulated on
diphenylmethane diisocyanate, butanediol and preformed polycaprolactone diol. Even if the
mechanism proposed for the chemical link between the nanofiller surface and the polymer does
not appear appropriate (ammonium salts are not known to induce e-caprolactone ring-opening
polymerization), they succeeded in producing a material where the nanofiller acts as a
multifunctional chain extender inducing the formation of star-shaped segmented poly(urethane).
Messersmith and Giannelis [58] have also reported on the e-caprolactone polymerization inside
a Cr3‡-exchanged fluorohectorite at 1008C during 48 h. Interestingly enough, the monomer
intercalation has been this time evidenced by the interlayer spacing increase from 12.8 to 14.6 A Ê
upon the addition of the liquid monomer. After the polymerization reaction, an intercalated poly(e-
caprolactone) is obtained as proved by an interlayer spacing of 13.7 A Ê observed by XRD, due to the
dimensional change accompanying the polymerization of the cyclic monomer as sketched in Fig. 15.
The observed layer spacing of 13.7 A Ê correlates well with the sum of the thickness of the
Ê
silicate layer (9.6 A) and the known interchain distance (4.0 A Ê ) in the crystal structure of poly(e-
18 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 15. XRD patterns of e-caprolactone intercalated in Cr3‡ modified fluorohectorite (plain line) and the resulting poly(e-
caprolactone)-based nanocomposite (dashed line). Insets are schematic illustrations corresponding the intercalated
monomer (left) and intercalated polymer (right) (reproduced from [58] with permission).

caprolactone). Extraction of unintercalated polymer with acetone allows to analyze the intercalated
composite that contains 7.7 wt.% of the polymer. DSC analysis of this extracted nanocomposite does
not show any melting endotherm in the vicinity of 608C, that is the expected melting temperature of
the PCL. This effect is attributed to the polymer chains confinement within the silicate galleries that
prevents the formation of polymer crystallites.
In situ intercalative polymerization has also been largely studied for producing poly(styrene)-
based nanocomposites. Akelah and Moet [59] modified the interlayer of Na-montmorillonite and Ca-
montmorillonite by exchanging the inorganic cations with (vinylbenzyl)trimethyl ammonium
chloride, increasing the interlayer spacing by 5.4 AÊ . These modified clays were then dispersed and
swollen in various solvent and cosolvent mixtures such as acetonitrile, acetonitrile/toluene and
acetonitrile/THF. Styrene polymerizations were carried out in presence of N,N0 -azobis(isobutyroni-
trile) (AIBN) and carried out at 808C for 5 h. The composites were isolated by precipitation of the
colloidal suspension in methanol, filtered off and dried. By this technique, intercalated composites
were produced with interlayer spacings varying between 17.2 and 24.5 A Ê depending on the nature of
the solvent used. Even if the polymer is well intercalated, a drawback of this technique remains that
the macromolecule produced is not a pure PS but rather a copolymer between styrene and
(vinylbenzyl)trimethyl ammonium cations.
Do and Cho have developed a technique using more commonly modified montmorillonites [60].
The authors compared the ability of several tetraalkylammonium cations incorporated into Na-
montmorillonite through ion-exchange to promote the intercalation of poly(styrene) through the free
radical polymerization of styrene initiated by AIBN at 508C. Three tetraalkylammonium cations
were tested, all based on the following formula: (CH3)2N‡(hydrogenated tallow alkyl)R where
hydrogenated tallow alkyl corresponds to a mixture of mainly octadecyl chains together with small
amounts of lower linear homologues and R may be either another hydrogenated tallow alkyl (Ta), 2-
ethyl hexyl (Eh) or benzyl (Bz) group. These so-modified organo-montmorillonites were coded as
Ta-MMT, Eh-MMT and Bz-MMT, respectively. Layer spacings obtained for the three MMTs and
corresponding composites are presented in Table 8. An evaluation of the filler dispersibility within
styrene during the polymerization reaction is provided as well.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 19

Table 8
Interlayer spacings of organo-modified montmorillonites (X-MMT) and as obtained PS-based nanocomposites and the clay
dispersibility within the polymerization medium
Xa-MMT Ê)
Interlayer spacing (A 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 ‡
Ta-MMT 32.7 32.9 ‡
a
Organo-modifiers: (CH3)2N‡(hydrogenated tallow alkyl)R with RˆBz (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.

It turns out that the best intercalation occurs for Bz-MMT. This is probably due to a better
affinity between styrene and benzyl groups spread all along the layered montmorillonite surfaces as
further demonstrated by the perfect dispersibility of this organo-modified filler in styrene. If Eh-
MMT does also intercalate some PS, the interlayer spacing for Ta-MMT does not change a lot (only
a 2A Ê increase) but TGA analysis seems to indicate that intercalation does occur in this composite
too (see Section 4.2.1 for further information). According to authors' discussion, intercalation within
Ta-MMT should occur without an important increase in the interlayer spacing because such
hydrogenated tallow alkyl chains should be long enough (mainly C18 chains) to easily accommodate
PS. Even though this technique allows an extensive intercalation of PS chains through an adequate
choice of the alkylammonium cation neither exfoliation nor control over the molecular parameter of
the polymer (PS) produced have been observed.
Such a control has been, however, achieved by Weimer et al. [61] who modified a Na-
montmorillonite by anchoring an ammonium cation bearing a nitroxide moiety known for its ability
to mediate the controlled/`living' free radical polymerization of styrene. The intercalative
polymerization principle used is sketched in Fig. 16.
Styrene polymerization was carried out in bulk (in absence of solvent) at 1258C for 8 h yielding
homogeneous and transparent composites. The absence of diffraction peaks in the low angle area of
the XRD patterns together with the TEM observations of silicate layers randomly dispersed within
the PS matrix attest for the complete exfoliation of the layered silicate. The polymer chains were
desorbed from the silicate layers by refluxing the nanocomposite in THF in the presence of LiCl.
Remarkably, the PS number average molecular weight measured by size exclusion chromatography
(SEC) (Mnˆ21,500) was in perfect agreement with the theoretical Mn expected from the initial
monomer to nitroxide initiator molar ratio, assuming a living polymerization (Mn theo-
reticalˆ24,400). Furthermore, the molecular weight distribution was monomodal with a narrow
polydispersity of 1.3. It is worth pointing out that the tailoring of the PS molecular weight can be
achieved by varying the amount of fixed initiator. For doing so, variable quantities of an inactive
alkylammonium (trimethylbenzylammonium) was added together with the nitroxide-bearing
ammonium, during the ion-exchange reaction. Thus, this strategy allows to tune up with the
molecular weight of the recovered PS while keeping a narrow polydispersity and an exfoliated
structure. `Livingness' of the free-radical polymerization was definitively evidenced by a
quantitative resumption and chain extension experiment. Indeed, addition of a second styrene feed
to the reaction medium led to a Mn increase from 30,000 to 74,000 after complete monomer
consumption, well in agreement with the calculated value of 75,300. It has to be noted that this
20 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 16. Schematic representation of the montmorillonite modification with nitroxyl-based organic cation and its
subsequent use to produce PS-based exfoliated nanocomposite (reproduced from [61] with permission).

technique is the only known way to obtain complete layered silicate exfoliation in a PS matrix, still
unreachable trough exfoliation±adsorption (see Section 3.1) or melt intercalation processes (see
Section 3.3).
Polyolefins represent another important family of polymers that has been investigated in order
to produce nanocomposites through in situ intercalative polymerization process. Tudor et al. [62]
have demonstrated the ability of soluble metallocene catalysts to intercalate inside silicate layers and
to promote the coordination polymerization of propylene. Accordingly, a synthetic hectorite
(Laponite RD) was first treated with methylaluminoxane (MAO) in order to remove all the acidic
protons and to prepare the interlayer spacing to receive the transition metal catalyst. It has to be
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 21

noted that MAO is commonly used in association with metallocenes to produce coordination
catalysts active in olefin polymerization. During this first treatment step, no noticeable increase of
the layer spacing was observed (even if the diffraction peak broadened slightly) but both increase in
Al content and IR data showing complete disappearance of absorptions assignable to Si±OH groups
agree with the MAO reaction/adsorption inside the layered silicate galleries. Upon the addition of the
metallocene catalyst ([Zr(Z-C5H5)2Me(THF)]‡), a cation exchange occurs with Na‡ and the catalyst
is incorporated in the hectorite galleries as demonstrated by an increase in the interlayer spacing of
4.7 AÊ , consistent with the size of the catalyst species. Using a synthetic fluorinated mica-type silicate
that is deprived from any protons in the galleries, the catalyst was even incorporated directly within
the filler interlayer, without the need of MAO pretreatment. These two modified layered silicates
catalyzed with reasonably high activity the polymerization of propylene when contacted with an
excess of MAO, producing PP oligomers. Unfortunately, no composite characterization was provided
with, so as one cannot claim intercalated or exfoliated morphology.
In a very recent work [63,64], intercalated/exfoliated nanocomposites based on high density
polyethylene matrices have been synthesized by the so-called polymerization-filling technique (PFT)
[65]. This method consists in anchoring in a first step, a Ziegler±Natta type catalyst or any other
coordination catalysts, which include MAO activated metallocenes, onto a filler surface, and then in
situ polymerizing ethylene and/or a-olefins directly from the surface treated fillers. High
performance microcomposites, combining both high stiffness and toughness, have been produced
by PFT, as a result of homogeneous filler dispersion, strong filler/matrix interfacial adhesion and the
unique possibility to get highly filled ultrahigh molecular weight polyethylene. It was, therefore, of
prime interest to apply this technique to nanofillers. Layered silicates (montmorillonite and
hectorite) in aqueous clay colloidal suspension were made less hydrophilic through the elimination
of water by freeze drying. The obtained fluffy materials could then be nicely dispersed in a non-polar
solvent such as heptane or toluene. The clay dispersion was then surface treated with MAO and, after
solvent removal by evaporation, a high temperature treatment at 1508C was applied to modify the
layered silicate. After washing the modified clay with toluene in order to remove unreacted MAO,
the silicate layers were contacted with a metallocene precatalyst, i.e. (tert-butylamido) dimethyl
(tetramethyl-Z5-cyclopentadienyl) silane titanium dimethyl and aged for 1 h in order to form the
active catalyst species. The polymerization is then carried out by adding ethylene in the medium. It
has to be pointed out that contrary to the procedure reported by Tudor et al. (vide supra), no ion
exchange reaction was required since the added metallocene precatalyst was not a cationic one.
Rather, the strategy here relies upon the immobilization of the active species through electrostatic
interactions with surface anchored MAO as already performed with more conventional microfillers
(kaolin, silica) [66]. Some typical in situ intercalative polymerization experiments are gathered in
Table 9.
Polymerizations were conducted in highly diluted conditions (ca. 2 g filler per liter of
dried heptane) so that a low level of layered silicates was reached in the final composition (down
to 3 wt.%). When ethylene polymerization was carried out in the absence of molecular
hydrogen (thus without any transfer agent), layered silicate filled UHMWPE is produced which
is an extremely viscous material and is highly difficult to melt (Table 9, entries 1 and 2). Addition
of hydrogen to the polymerization medium allows molecular weight to be reduced down
(Mwˆ504,000 and Mnˆ77,000 as determined by SEC) with substantial improvement of the melt
processability (Table 9, entry 3). Examination of the TEM pictures reveals that layered silicate
exfoliation partially occurred. For instance, at high filler content (ca. 30 wt.% montmorillonite),
besides some stacked multilayered silicates, one can observe well exfoliated and dispersed nano-
sized sheets (Fig. 17).
22 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 9
Synthesis and composition of PE-based nanocomposite produced by in situ intercalative polymerization of ethylene
(P(C2H4)ˆ10 bar) in non-organo-modified layered silicatesa
Filler MAO Catalyst P(H2)b Filler loadingc HPDE Mn
(10ÿ3 mol) (10ÿ6 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
a
hˆhectorite and mˆmontmorillonite.
b
Hydrogen partial pressure at start.
c
Measured by thermogravimetric analysis (TGA).
d
Insoluble UHMWPE that cannot be eluted by SEC.

Heinemann et al. [67] have also investigated intercalative polymerization to produce


(co)polyolefin nanocomposites. They carried out the polymerizations in the presence of modified
layered silicates such as bentonite after sodium exchange with dimethyldistearylammonium or
dimethylbenzylstearylammonium cations. For the sake of comparison, non-organo-modified clays
were studied as well, including both synthetic hectorite and fluoromica. As a typical synthesis, the
filler silicate was dispersed in toluene (or methylene chloride) followed (when needed) by the

Fig. 17. TEM micrograph of Na-montmorillonite exfoliated in HDPE after in situ intercalative polymerization of ethylene.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 23

addition of 1-octene as a comonomer. The reactor was then saturated with ethylene and
polymerization was started by the addition of the catalyst. Ethylene was continuously introduced
during polymerization. After a given time, the polymerization was stopped by addition of 2-propanol
and the composite recovered by precipitation from acidified methanol, then filtered and dried.
Different catalysts were used, affording HDPE and ethylene±octene copolymers (zirconium-based
catalyst), and branched polyethylene (nickel and palladium-based catalysts). It appeared that the
modified bentonites had a dramatic negative effect on the polymerization activity of the zirconium-
based catalyst, while Ni- and Pd-based catalysts were much less affected by the nature of the clay.
This effect was attributed to the high sensitivity of the Zr-based active species towards any kind of
polar functionality, including the anionic silicate layers covered by the alkylammonium cations.
Nanocomposites were observed to be formed when polymerization was carried out by using organo-
modified clays. At the opposite, composites prepared either by in situ polymerization with non-
modified silicates or by melt-blending a preformed HDPE with a modified bentonite, only gave
microcomposites as demonstrated by TEM and XRD. The presence of n-alkyl branches along the
polyethylene chains (as a result of 1-octene copolymerization or migratory insertion ethylene
polymerization promoted by Ni and Pd-based catalysts) was reported to enhance compatibility
between the (co)polyolefin matrix and the dispersed layered silicates, improving the mechanical
properties of the resulting nanocomposite materials.
A very recent report discusses also about the synthesis of poly(ethylene terephtalate) (PET)
nanocomposites by using the in situ intercalative polymerization [68]. The montmorillonite,
modified by organic onium not given by the authors (interlayer distance: 11.2 A Ê ), is reported
to react with PET comonomers (ethylene glycol and terephtalic acid derivatives) to form an
intercalated nanocomposite with an interlayer distance going from 14 to 35 A Ê depending on the clay
content.

3.2.2. Thermoset nanocomposites


Next to all the aforementioned thermoplastic nanocomposites, in situ intercalative polymeriza-
tion has also been explored to create thermoset-based nanocomposites. This method has been widely
described for the production of both intercalated and exfoliated epoxy-based nanocomposites.
Messersmith and Giannelis [69] have analyzed the effect of different curing agents and curing
conditions in the formation of nanocomposites based on the diglycidyl ether of bisphenol A
(DGEBA) and a montmorillonite modified by bis(2-hydroxyethyl)methyl hydrogenated tallow alkyl
ammonium cation. They found that this modified clay dispersed readily in DGEBA when sonicated
for a short time period as determined by the increase in viscosity at relatively low shear rates and the
transition of the suspension from opaque to semitransparent. The increase in viscosity was attributed
to the formation of a so-called `house-of-cards' structure in which edge-to-edge and edge-to-face
interactions between dispersed layers form percolation structures. XRD patterns of uncured clay/
DGEBA also indicate that intercalation occurred during mixing and that this intercalation improves
going from room temperature to 908C (Fig. 18).
At room temperature a mix of intercalated and unintercalated clay species coexists as
demonstrated by the persistence of the shoulder in the diffraction peak at 2yˆ5.88 and the
appearance of the intercalation peak at 2yˆ2.58. At increasing temperature, the diffraction peak
Ê keeping a constant intensity, suggesting that little or no delamination
slightly shifts from 36 to 38 A
occurs at or below 1508C. The influence of the curing agent used in the polymerization process
appeared to be determinant in the resulting structure of the formed nanocomposite. When diamines
were used, only intercalated epoxy-clay structures could be obtained. One possibility for this
behavior may arise from the bridging of the silicate layers by the bifunctional amine molecules,
24 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 18. XRD patterns of organomodified montmorillonite, its uncured mixture with epoxy precursor (DGEBA) at room
temperature and after 1 h annealing at 908C (reproduced from [69] with permission).

which prevents further expansion of the layers from taking place. When other curing agents such as
nadic methyl anhydride (NMA), boron trifluoride monomethylamine (BTFA) or benzyldimethyla-
mine (BDMA) were added, delamination during heating of the reaction mixture occurred as attested
by XRD for BDMA (Fig. 19).
Addition of the curing agent induced first an increase of the interlayer from 36 to 39 A Ê,
indicating some partial intercalation. With further heating, disappearance of the interlayer spacing
reflection indicated that delamination occurred. XRD of completely cured nanocomposites showed
no more evidence of ordered reflections and TEM analysis displays silicate platelets separated from
the others by more than 100 A Ê . Moreover, study of the curing reactions tended to prove that the
particular alkylammonium used (that bears two hydroxyl functions) could play an active role,
especially when BDMA or NMA were added as the curing agents. Indeed, BDMA can catalyze the

Fig. 19. XRD patterns of organo-modified montmorillonite/DGEBA/BDMA mixture (4 wt.% filler) heated in situ to
various temperatures. The spectra are displaced vertically for clarity, with scan temperatures increasing from bottom to top
(reproduced from [69] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 25

reaction between 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. As far as NMA is concerned, the resin formation
may occur following a series of reactions beginning with the opening of the cyclic anhydride by the
hydroxyl group of the alkylammonium. At higher temperature, the formed carboxylic acid can add
onto the oxirane, liberating an hydroxyl group that reacts further with the oxirane groups to form the
epoxy network. This mechanism is accredited by the fact that a complete curing cannot occur by
heating DGEBA and NMA only. Furthermore, the DSC analysis shows two exotherms at 180 and
2478C that could correspond to the above two-step reaction.
Finally, one can note the synthesis of nanocomposites based on montmorillonite and unsaturated
polyester [70]. In this study, the montmorillonite was treated with a methacrylate±silane
coupling agent in order to render the filler hydrophobic and reactive. The unsaturated polyester
was polymerized by free radical polymerization with the modified montmorillonite dispersed
in it. The authors claim the formation of exfoliated structure confirmed by the absence of diffraction
peaks in the XRD pattern and a TEM micrograph showing a molecular dispersion of the clay
layers. This example is the first one presenting an alternative way of modifying layered silicates by
silane coupling reaction rather than interlayer ion-exchange yielding to the formation of a
nanocomposite.

3.2.3. Elastomeric nanocomposites


In a first study, Lan and Pinnavaia [71] have examined the formation of nanocomposites with a
rubber-epoxy matrix obtained from a DGEBA derivative (Epon 828) cured with a polyether diamine
(Jeffamine D2000) so as to reach subambient glass transition temperatures. Two montmorillonites
modified, respectively, by the protonated n-octylamine and the protonated n-octadecylamine have
been used in this study. It has been shown that depending on the alkyl chains length of modified
clays, an intercalated and partially exfoliated (n-octyl) or a totally exfoliated (n-octadecyl)
nanocomposite can be obtained. With the protonated octadecylamine-modified clay, heating the
reaction mixture at 758C triggered the epoxide and diamine to migrate into the clay galleries and to
form an intermediate with an interlayer spacing of 54 A Ê . Upon additional heating, further
polymerization occurs (catalyzed by the protonated primary amine) with deep penetration of the
components within the gallery, leading to the formation of the exfoliated structure. In the case of the
octyl derivative whose hydrophobicity is lower, the amount of intercalated epoxide and diamine is
insufficient to achieve complete exfoliation. Therefore, only a portion of the clay is delaminated, as
evidenced by the broadening and the decrease in the scattering intensity of the 15.2 A Ê reflection,
corresponding to the modified clay interlayer spacing. These authors have also studied other
parameters such as the nature of alkylammonium cations present in the gallery and the effect of the
cation-exchange capacity of the clay [72] when Epon 828 was cured with m-phenylene diamine. It
was demonstrated that when mixed with the epoxide and whatever the length of the protonated
primary alkylamine (with 8, 10, 12, 16 or 18 carbons), the modified clays adopt a structure where the
carbon chains are fully extended and oriented perpendicularly to the silicate plane incorporating
maximum one monolayer of epoxide molecules. In order to activate clay exfoliation, presence of
acidic species in the intergallery is necessary, as illustrated (Fig. 20) by the decrease in layer
exfoliation with decreasing BroÈnsted acidity of the exchange ion in the order CH3(CH2)17NH3‡
>CH3(CH2)17N(CH3)H2‡>CH3(CH2)17N(CH3)2H‡>CH3(CH2)17N(CH3)3‡.
The effect of cationic exchange capacity has been addressed using four different clays
(hectorite, CECˆ67 meq/100 g; montmorillonite, 90 meq/100 g; fluorohectorite, 122 meq/100 g;
vermiculite, 200 meq/100 g) modified by CH3(CH2)15NH3‡. XRD patterns of the diamine-cured
26 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 20. XRD patterns of diamine-cured epoxy/clay nanocomposites formed from montmorillonite clays (5 wt.%)
containing primary, secondary, tertiary and quaternary onium ions with n-C18 chain lengths (reproduced from [72] with
permission).

epoxy clay composites formed with 5 wt.% loading of the modified clays show that an exfoliated
structure is formed with hectorite and montmorillonite while with fluorohectorite, a partially
exfoliated/partially intercalated structure is obtained. For vermiculite, an intercalated structure is
obtained because the density of onium ions present is very high and limits the penetration of
monomers. As a result, only a little amount of epoxy resin is formed inside the intergallery that
cannot expand enough to undergo exfoliation.
The same kind of results were obtained by Zilg et al. [73] who cured DGEBA with
hexahydrophtalic acid anhydride in the presence of different clays (fluoromica, hectorite and
bentonite) modified by a wide variety of oniums. An original result comes from the inefficiency of
functionalized oniums (carboxylic acid or hydroxyl group) to improve exfoliation comparing to
simple protonated alkylamine.
Starting from another type of layered material, magadiite, which is a layered silicic acid, Wang
and Pinnavaia [74] have developed a different type of exfoliated structure where the layers were still
organized but spaced by around 80 A Ê of elastomeric epoxy matrix arising from the curing of Epon
828 (DGEBA derivative) by Jeffamine 2000 (diamine polyether). Such unusual structure can be
obtained by carefully controlling of the organo-modification of magadiite with a mixture of
alkylammonium cation and the corresponding alkylamine (dimethyloctadecylamine) in order to form
a paraffin-like environment between the silicic acid layers. It has to be noted that exfoliation of
magadiite layers within the epoxy matrix can only be achieved with this paraffin-like structure. If the
alkylamine is not used, a `lateral monolayer' is formed which is unable to swell in the monomer
slurry and does not lead to the formation of a nanocomposite.
Wang and Pinnavaia [75] have also synthesized intercalated nanocomposites based on
elastomeric polyurethanes. An organomontmorillonite modified with the protonated dodecylamine
or octadecylamine is swollen in a polyol such as ethylene glycol, poly(ethylene glycol), or Voranol
(glycerol propoxylate with molecular weight ranging from 700 to 3000), then cross-linked using a
commercial methylene diphenyl diisocyanate prepolymer (Rubinate). After curing at 508C for 12 h
an intercalated nanocomposite can be obtained with an interlayer spacing of 50.8 A Ê.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 27

Fig. 21. Variation of entropy change per area unit (hDSV/NAkB) in function of the change in gallery height for an arbitrary
polymer and a silicate functionalized with octadecylammonium groups calculated for the polymer chain (dashed point
line), the tethered chains (dashed line) and addition of the two contributions (plain line). h1ÿh0 is the change in gallery
height for a fully extended octadecyl chain (reproduced from [76] with permission).

3.3. Melt intercalation

3.3.1. Melt intercalation of modified montmorillonite

3.3.1.1. Theoretical concepts. The thermodynamics that drives the intercalation of a polymer inside
a modified layered silicate while the polymer is in the molten state has been approached through a
lattice-based mean field theory by Vaia and Giannelis [76]. They found that, in general, the outcome
of polymer intercalation is determined by an interplay of entropic and enthalpic factors. In fact,
although the confinement of the polymer chains inside the silicate galleries results in a decrease in
the overall entropy of the macromolecular chains, this entropic penalty may be compensated by the
increase in conformational freedom of the tethered alkyl surfactant chains as the inorganic layers
separate, due to the less confined environment as depicted in Fig. 21.
Since small increases in the gallery spacing do not influence strongly the total entropy change,
intercalation will rather be driven by the changes in total enthalpy. In this study, the enthalpy of
mixing has been classified in two components: apolar interactions generally unfavorable and arising
from interaction between polymer and surfactant aliphatic (apolar) chains, and polar interactions
which originate from the Lewis acid/Lewis base character of the layered polar silicates interacting
with the polymer chains. Indeed, since in most conventional organo-modified silicates, the tethered
surfactant chains are apolar, dispersion forces dominate polymer±surfactant interactions. On the
other hand, a favorable energy decrease is associated with the establishment of many favorable
polymer surface polar interactions. The enthalpy of mixing can thus be rendered favorable by
maximizing the magnitude and number of favorable polymer±surface interactions while minimizing
the magnitude and number of unfavorable apolar interactions between the polymer and the aliphatic
chains introduced along the modified layer surfaces. Fig. 22 shows the free energy per area (hDfV)
evolution in function of the gallery height (hÿh0) at 423 K for various esp,sa. The notation esp,sa
corresponds to espÿesa where esp…or a† is the pair-wise silicate surface (s)±polymer (p) chain or
surfactant (a) interaction energy per area of the s±p (or a) contacts relative to the initial s±s and p±p
(or a±a) interaction. For the simplicity of the model, the pair-wise interaction between the polymer
28 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 22. Variation of the free energy per unit area (hDfV) in function of the change in gallery height (hÿh0) calculated for
an arbitrary polymer interacting with a silicate layer functionalized with octadecyl ammonium groups. Curves are
calculated for various values of the difference in pair-wise interaction energies between the silicate layer and the polymer
chain and the silicate layer and the tethered chains esp,sa. The pair-wise interaction energy between the polymer chain and
the tethered chains (eap) is taken equals to 0. Free energy curves I, IIa, IIb and III correspond to esp,sa values of 0, ÿ4, ÿ8,
ÿ12 mJ/m2, respectively (reproduced from [76] with permission).

and the surfactant eap is taken equal to 0. This situation describes in fact melt intercalation of a series
of polyethylene-based polymers with esp,saˆ0 for pure HDPE and esp,sa<0 for polyethylene
copolymers in which a small ration of ethylene units are replaced with moieties exhibiting more
favorable interactions with the silicate.
The free energy curves may be grouped into three types. First, curves that are positive at all
gallery heights (type I, esp,saˆ0). In this case, polymer intercalation is unfavorable, and the polymer
and the organo-modified layered silicates are immiscible. The second type regroups the curves
displaying one minimum (type IIa, esp,saˆÿ4) or more than one minimum (type IIb, esp,saˆÿ8) and
corresponding, respectively, to well defined intercalated structure and ill-defined intercalated
structures or intermediate intercalated structures before complete layer exfoliation. Finally, the third
type of curves displays a continuous decrease in the free energy values with gallery height expansion
(type III, esp,saˆÿ12) indicating that polymer intercalation and complete layer separation is
favorable. This last type corresponds to complete polymer-silicate layer miscibility, characteristic of
exfoliation.
Balazs et al. [77,78] have considered the self-consistent field theory (SFC) in order to
investigate the factors promoting the penetration of polymers into layered silicates. They first varied
properties related to the nature of the tethered surfactant chains. They found out that an increase in
the surfactant length (approaching the length of the polymer chains) improves the layers separation
by the formation of a broad interface (or interphase) which allows the polymer from adopting more
conformational degrees of freedom. Accordingly, exfoliated or intercalated structures can be formed
even for slightly unfavorable interactions between the polymer and the modified surfaces. On the
opposite side, increase in the length of polymer chains tends to render the interlayer mixture
immiscible. These authors also reported on the effect of the surfactant density on the intercalation
process, showing that excessive density of tethered alkyl chains can impede the formation of
intercalated structures.
In order to model the macroscopic phase behavior of the polymer/clay mixture, Balazs et al.
[78,79] have adapted to disk-like particles the Onsager model for the equilibrium behavior of rigid
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 29

Fig. 23. Phase diagrams of disks dispersed in a polymer matrix for different disk aspect ratios. The disk aspect ratios (D/L)
were varied by changing the dimension of the diameter of the disks (D), keeping their thickness constant (Lˆ1)
(reproduced from [79] with permission).

rods set up dispersed in a polymer matrix. From this new model, one can calculate phase diagrams of
the polymer/clay composites in function of the Flory±Huggins interaction parameter. These
diagrams can differentiate immiscible and miscible regions further separated in isotropic or nematic
(relatively ordered) arrangement. In this study, the blend miscibility appears to be strongly negatively
influenced by an increase in the length of polymer chains. For very high polymer length, the particle/
polymer mixture gets immiscible even for negative values of the interaction parameter. The phase
diagram appears to be also strongly dependent upon the aspect ratio of the particle (i.e. the diameter
(D) to thickness (L) ratio of the assumed disk-shaped particle) as shown in Fig. 23.
An increase in the particle diameter favors the apparition of a nematic phase at a low volume
fraction implying the formation of relatively ordered structures for low clay particle content.
Finally, Ginzburg and Balazs [80] have developed an even more complex model, based on
perturbation-type density functional theory to describe the complete (isotropic, nematic, smectic A,
columnar and intercalated called `crystalline') phase diagram of an incompressible polymer disk
mixture. The phase diagram is shown to be strongly dependent upon the shape anisotropy of the filler
particles, the polymer chain length, and the strength of the interparticle interaction. For instance, an
increase in the interparticle interaction strength leads to the complete disappearance of the nematic
phase in favor to direct coexistence between isotropic and columnar or `crystalline' (intercalated)
structure.

3.3.1.2. A model polymer: polystyrene. Melt intercalation of polystyrene and derivatives has been
widely studied and has served to experimentally describe the aforementioned thermodynamics as
well as the kinetics and morphologies accordingly generated.
Vaia and Giannelis [81] have studied PS as the matrix for dispersing different types of clays. Li-
fluorohectorite (CECˆ150 meq/100 g), saponite (100 meq/100 g), and sodium montmorillonite
(80 meq/100 g) were accordingly modified using various ammonium cations: dioctadecyldimethy-
lammonium, octadecyltrimethylammonium, and a series of primary alkylammonium cations with
carbon chains of 6, 9±16 and 18 carbon atoms. The nanocomposites were synthesized by statically
annealing (without any mixing or shearing) a pelletized intimate mix of the modified silicate in PS
(Mwˆ30,000, Mw/Mnˆ1.06) under vacuum at 1708C, a temperature well above the PS glass
transition [82] (Table 10).
Comparison of the first three entries indicates that, for a given alkyl surfactant, increasing the
cation exchange capacity from 80 to 120 meq/100 g allows for PS intercalation to occur. At a low
CEC and for single alkyl chain built up cation (entries 1 and 2), no intercalation is observed. Under
these conditions, aliphatic alkyl chain of the organic cation adopts a pseudo monolayer arrangement
30 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 10
Characteristics of polystyrene melt intercalation within octadecylammonium modified claysa
Entry Clay (CEC in Ammonium Initial gallery Final gallery Net change
meq/100 g) cation height (nm) height (nm) (nm)
1 M (80) PODAb 0.75 0.75 0
2 S (100) PODAb 0.83 0.83 0
3 F (120) PODAb 1.33 2.16 0.83
4 F (120) QODAc 1.57 2.69 1.12
5 M (80) DODMDAd 1.43 2.25 0.82
6 S (100) DODMDAd 1.50 2.35 0.85
7 F (120) DODMDAd 2.85 2.85 0
a
Mˆmontmorillonite, Sˆsaponite and Fˆfluorohectorite.
b
PODA: primary octadecylammonium.
c
QODA: quaternized octadecylammonium.
d
DMDODA: dioctadecyldimethylammonium.

characterized by low gallery height. Interestingly enough, intercalation can take place at lower CECs
at the condition to modify the clay surface with ammonium cation bearing two long alkyl chains
(entries 5 and 6). However, excessive packing of the chains all along the layer surface (high CEC and
two long alkyl chains per cationic head, entry 7) leads to the formation of a non-intercalated
structure as predicted by the theory introduced by Vaia and Giannelis [76] (vide supra). The authors
also found out that polymer intercalation depends on the length of the exchange ammonium cation as
well as on the annealing temperature. At annealing temperature equal to or lower than 1608C and for
chain lengths inferior to 12 carbon atoms, no intercalation is detected, while for higher chain lengths
PS readily intercalates. At 1808C, intercalation occurs whatever the ammonium chain length be.
However, the full width at half-maximum of the XRD diffraction peak, that is known to witness for
the intercalation regularity, increases for chain lengths lower than or equal to 12 carbon atoms. Such
a peak broadening attests for a more disordered and irregular structure and arises only for organo-
modified clays with a pseudo bilayer chain arrangement. The authors interpret this behavior as an
extended propensity of these organoclays to separate under these conditions, exfoliation being
nevertheless kinetically limited by the static conditions used to produce the nanocomposites, i.e.
simple annealing without mixing or shearing. The effect of PS molecular weight was also studied.
Contrary to what was predicted by the theory, intercalation occurred whatever the Mw of the PS
chains. Only the period of time needed for the intercalation to proceed was different, going from 6 h
for a Mw of 30,000 to 24 h for 90,000 and 48 h for 400,000 at 1608C. Clearly, high molecular weight
PS chains decrease the kinetics of intercalation by decreasing the diffusivity of the polymer in the
interlayer. Finally, the authors studied the influence of the nature of the polymer matrix, creating
under the same experimental conditions different composites based on poly(vinylcyclohexane)
(PVCH), poly(3-bromostyrene) (PS3Br) and poly(vinylpyridine) (PVP). All these polymers appear
to be immiscible with the dioctadecyldimethylammonium-modified fluorohectorite but interestingly
formed intercalated structures with montmorillonite (thus with a lower CEC) modified by the same
ammonium cation with an exception for PVCH. These differences in intercalation behavior indicate
that the chemical nature of the polymer pendant group (polar or not) greatly affects the formation of
the hybrid composition. Moreover, in the case of intercalated PVP, a broadening of the diffraction
peak in XRD pattern indicated a less regular arrangement of the intercalated structure possibly due to
some propensity for exfoliation. As far as the dodecylammonium-modified fluorohectorite was
concerned, the broadening of the diffraction peak observed for PVP-based nanocomposite was even
more pronounced attesting for a very disordered structure. These results are totally in accordance
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 31

Fig. 24. Temporal series of XRD patterns for a organo-modified fluorohectorite/PS pellet annealed in situ at 1608C in
vacuum. p(0 0 1) and p(0 0 2) locate the basal reflections for the unintercalated fluorohectorite while i(0 0 1), i(0 0 2) and
i(0 0 3) correspond to the basal reflections for the intercalated nanocomposite that is forming with time (reproduced from
[83] with permission).

with the proposed thermodynamic model that anticipated a better intercalation and even exfoliation
when polar interactions between the layered silicates and the polymer chains are enhanced.
Vaia et al. [83] have also studied the kinetics of melt intercalation by following the time
evolution of XRD diffraction patterns for statically annealed polystyrene/octadecylammonium-
exchanged fluorohectorite. The change in intensity of the pristine and intercalated diffraction peaks
with time was used as a reflect of the kinetics of the polymer intercalation process (Fig. 24).
By integrating the intensity of both non-intercalated and intercalated peaks, the authors were
able to estimate the fraction of intercalated silicates as a function of the annealing time. Influence of
both annealing temperature and PS molecular weight were then determined. Higher annealing
temperatures, as well as lower molecular weights, increase the rate of PS intercalation. In order to
give a more concrete interpretation of the kinetic data, the morphology of the modified
montmorillonite has been examined by TEM. The authors found that the nominal silicate particle
(agglomerate of ca. 175 mm in diameter) consist of smaller oblong-shaped particles, coined as
primary particles forming agglomerates. Those primary particles of 1±10 mm length themselves
consist of a compact face-to-face stacking or low-angle intergrowth of individual silicate crystallites
(also known as tactoids). These crystallites are built up as a coherent stacking of individual silicate
layers.The layers are roughly circular, 0.05±0.5 mm in diameter, and ca. 1 nm thick. They are
separated by a van der Waals interlayer (or gallery), which contains the alkylammonium cations
(10 A Ê thick) in the pristine organosilicate. This overall morphology can be schematized as shown
in Fig. 25.
Based on this morphology and TEM observations of partially intercalated composites, it has
been demonstrated that the accessibility of the interlayer to the polymer chains depends on the
location and orientation of the primary particles within the agglomerates and on the location and
orientation of the crystallites within the primary particles, meaning that crystallites near the edge will
be more accessible to polymer chains than those near the center. Since the silicate layers are
impenetrable, the polymer must enter the gallery from the edges of the crystallites. The authors
observed that, under their experimental conditions, the polymer penetrates the agglomerate and
surrounds the primary particles before the occurrence of substantial intercalation. Therefore, they
assumed that the polymer penetration within the agglomerate was not a limiting step for
32 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 25. Schematic of the morphology of organo-modified fluorohectorite (reproduced from [83] with permission).

intercalation. Furthermore, they demonstrated that the rate of PS intercalation was dependent upon
the size of the primary particles, bigger primary particles being less intercalated than smaller ones,
for a given annealing time. This size dependence was explained by the lack of significative
difference in the mass transport in between and through the crystallites within a given primary
particle so as polymer intercalation could be described as a Fickian process with a single, apparent
diffusivity. The polymer diffusion into the primary particles can be sketched as a mass flow through
the curved lateral surface of a cylinder bearing impermeable flat circular faces of area equal to the
mean primary particle size. Accordingly, the activation energy of the melt intercalation was
calculated based on the experimental values of the effective diffusional rate obtained at different
temperature. Assuming an Arrhenius temperature dependence, the activation energy for the PS
(Mwˆ30,000 and Mw/Mnˆ1.06)/modified fluorohectorite composite is 16612 kJ/mol, comparable
to the activation energy measured for self-diffusion of PS (167 kJ/mol) [83]. This indicates that the
largest energy barrier to PS intercalation is comparable to that of PS chain motion within the
polymer melt. Since mass transport into the primary particle has been observed to be the limiting
step to complete intercalation, any dynamic mixing should decrease the intercalation time by
breaking down the primary particles and increasing the sample uniformity, that opens the way to the
preparation of intercalated PS within a few minutes.
In a further study, Vaia et al. [84] have synthesized, by the above reported static annealing
procedure, ordered and disordered intercalated nanocomposites by mixing, respectively, 30,000 or
400,000 Mw PS with octadecylammonium-exchanged fluorohectorite on one side and 30,000 Mw PS
or 55,000 Mw poly(3-bromostyrene) with dodecylammonium-exchanged fluorohectorite on the other
side. TEM characterization shows that the ordered intercalates exhibit regular microstructures
similar to the non-intercalated organo-modified fluorohectorite. In this case, polymer intercalation
occurs as a front which penetrates the primary particles from the external edges while for disordered
intercalated systems, a heterogeneous microstructure is observed, with pronounced interlayer
disorder with greater spacings towards the polymer-primary particle boundary. In these latter
nanocomposites, individual silicate layers are observed near the edges, whereas small coherent layer
packets separated by polymer-filled gaps are prevalent inside the primary particle. This study also
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 33

shows that intercalation could be favored by the presence of defects in the layer stacking that would
allow the polymer to penetrate more easily inside the interlayer spaces.

3.3.1.3. Nylon-6. While the preparation of nanocomposites based on nylon-6 matrix has been
widely described using the in situ intercalative polymerization (see Section 3.2.1), less attention
has been drawn to the nylon-6-based nanocomposites prepared by melt blending. Liu et al. [85]
have prepared nanocomposites based on a commercial nylon-6 melt blended with an octa-
decylammonium-exchanged montmorillonite (CECˆ100 meq/100 g) in a twin screw extruder. They
prepared composites with a filler content ranging from 1 to 18 wt.%. An intercalated structure was
observed to be formed by XRD for composites containing more than 10 wt.% of the organoclay, with
an interlayer spacing increasing from 15.5 AÊ for the pristine organoclay to 36.8 A
Ê for the intercalated
species. At filler content lower than 10 wt.%, no interlayer spacing could be detected through XRD
and the TEM micrographs allow for the observation of an exfoliated structure, indicating that
exfoliation in this case is highly dependent upon the filler content. XRD and DSC data also showed
that exfoliated structures strongly influenced the nature of the nylon-6 crystallization, favoring the
formation of g-crystals in addition to the crystals of the a-form observed in the native nylon-6
matrix. Moreover, DSC cooling scans showed that exfoliated layered silicates highly increased the
crystallization rate, having a strong heterophase nucleation effect.

3.3.1.4. Polypropylene. Polypropylene (PP) has also been tested for the preparation of
nanocomposites. However, no direct intercalation of PP in simply organically modified layered
silicates has been observed so far, PP being too much apolar to correctly interact with the modified
layers. In a first study, Kato et al. [86] described the melt intercalation of PP chains modified with
either maleic anhydride (PP-MA) or hydroxyl groups (PP-OH) in octadecylammonium-exchanged
montmorillonite (CEC: 119 meq/100 g). When PP-MA (Mwˆ30,000, acid valueˆ52 mg KOH/g) or
PP-OH (Mwˆ20,000, OH valueˆ54 mg KOH/g) was melt blended under shearing with a same given
amount of modified montmorillonite at 2008C for 15 min, intercalated nanocomposites were
recovered as demonstrated by the increase in the layer spacing from 21.7 A Ê for the initially organo-
modified montmorillonite to 38.2 and 44.0 A Ê for PP-MA and PP-OH based nanocomposites,
respectively. Interestingly enough, a PP-MA matrix with a lower maleic anhydride content (acid
valueˆ7 mg KOH/g for Mwˆ12,000) did not intercalate under the same conditions, showing that a
minimal functionalization of the PP chains has to be reached for intercalation to proceed. The
authors also examined the effect of polymer to clay ratio on the intercalation extent and showed that
intercalation increased when the PP-MA fraction was increased, going up to 72.2 A Ê for a PP-MA to
clay ratio of 3:1.
Intercalation of PP-MA in modified clay was used in order to prepare PP-based nanocomposites
[87,88]. In both studies, the three components (PP, PP-MA and modified clay) were melt blended in
a twin-screw extruder at 2108C in order to obtain composites filled with 5 wt.% clay. Formation of
an exfoliated structure was observed for: (1) relatively high PP-MA content (typically 22 wt.%), (2)
sufficient polar functionalization of PP-MA chains (acid valueˆ26 mg KOH/g for Mwˆ40,000).
However, the relative content in maleic anhydride cannot exceed a given value in order to keep some
miscibility between PP-MA and PP chains. Indeed, when too many carboxyl groups are spread along
the polyolefin chains (e.g. acid valueˆ52 mg KOH/g), no further increase in the interlayer spacing
was obtained in clay/PP/PP-MA blends, leading rather to the dispersion of PP-MA intercalated clay
in the PP matrix.
Another way to obtain nanocomposites from organo-modified clays and PP has been recently
proposed by Wolf et al. [89]. In this technique, the authors modified a commercially available
34 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Table 11
Interlayer spacing of various modified montmorillonite and the resulting composites obtained with EVA (10.76 mol% vinyl
acetate)
Code Cation Ê)
Interlayer spacing (A
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
Mont-3CNC21COOH (CH3)3N‡C21H42COOH 20.1 20.1

organoammonium-exchanged montmorillonite using an organic swelling agent (whose boiling point


is situated between 100 and 2008C, such as ethylene glycol, naphtha or heptane) in order to increase
the interlayer spacing. The swollen organo-modified clay was then compounded with PP in a twin-
screw extruder at 2508C. The swelling agent was volatized during extrusion process, leading to the
formation of a `nano' composite which did not present any crystalline reflection in the 20±40 A Ê
range of XRD patterns.

3.3.1.5. Ethylene-vinyl acetate copolymers. In a very recent work, ethylene-vinyl acetate


copolymers with various vinyl acetate contents (4.2, 7.1, 10.8 and 24.8 mol% vinyl acetate) have
been used as matrices for the preparation of nanocomposites [90]. The presence of polar groups
(ester group of the vinyl acetate moieties) all along the chains improves the ability of these
copolymers to intercalate in organo-modified montmorillonites. Several exchanging cations bearing
either simple alkyl chains or aliphatic chains terminated by a carboxylic group have been studied for
modifying montmorillonites and are described in Table 11.
Nanocomposites were only formed when EVA copolymers were melt blended at 1308C with
non-functionalized organo-montmorillonites such as montmorillonite exchanged with dimethyl-
dioctadecyl ammonium (Mont-2CN2C18). A partially intercalated±partially exfoliated structure was
observed by both the presence of peaks characteristic of the intercalation process in the XRD
patterns (see Table 11) and appearance of dispersed silicate layers in TEM micrographs (Fig. 26).
This intercalation±exfoliation morphology occurs even at low vinyl acetate content (4.2 mol%)
in the copolymer matrix while no intercalation is observed for HDPE, thus in the absence of polar
ester groups. It is moreover independent of the processing temperature. A set of experiments based
on the EVA matrix containing 10.8 wt.% of vinyl acetate has also allowed to determine that
concomitant intercalation and exfoliation occur whatever the filler content (from 1 to 50 wt.% of
Mont-2CN2C18) even if the exfoliation step tends to decrease when the organo-montmorillonite
amount increases. With the same EVA matrix, the use of ammonium cations functionalized with
carboxylic groups did not lead to the formation of an intercalated structure (see last two entries in
Table 11), indicating that the functionalization of the clay interlayer is detrimental to the
intercalation process. EVA copolymers appears to easily form nanocomposites even if totally
exfoliated structures have not been achieved yet.

3.3.1.6. Poly(styrene-b-butadiene) copolymer (SBS). Symmetric (styrene±butadiene±styrene) block


copolymer has been the first thermoplastic elastomer matrix investigated in the preparation of
nanocomposites. Laus et al. [91] have melt blended a commercial SBS (Mnˆ70,000, Mw/Mnˆ1.18,
and 30 wt.% in PS) with a dimethyldioctadecylammonium-exchanged montmorillonite in a
Brabender at 1208C for 10 min. Composites with 10, 20 and 30 wt.% of filler were compounded.
A comparative set of composites with the corresponding native Na-montmorillonite has also been
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 35

Fig. 26. TEM micrograph of organo-modified montmorillonite exfoliated in a ethylene-vinyl acetate copolymer matrix.

prepared. In addition, all the samples were further annealed at 1208C under nitrogen for time periods
ranging from 16 to 73 h in order to study the effect of the thermal treatment on their structural
characteristics. The authors have controlled that no chemical modification or degradation occurred
during compounding and thermal treatments by checking the molecular parameters of the SBS
copolymers. SBS recovered by complete extraction in CHCl3 of the so prepared composites
displayed unmodified molecular parameters compared to starting materials. While no change in
XRD patterns appears for the composites prepared with Na-montmorillonite (interlayer spacingˆ
12.7 AÊ for the filler and both three composites), a substantial change in the XRD patterns could be
observed for the composites based on the organo-montmorillonite. These composites showed both a
small increase in the interlayer spacing (from 41 A Ê for the organoclays to 46 AÊ in the composites)
and a broadening of the diffraction peak that increased with annealing time. These information tend
to indicate the formation of intercalated nanocomposites that is far from being complete during the
blending time. Analysis of the dynamic mechanical properties (see also Section 4.1.3) shows an
increase in the glass transition temperature (Tg) of the poly(styrene) outer blocks (PS) with the filler
content while the Tg corresponding to the poly(butadiene) block (PBD) is kept unchanged (Fig. 27).
Under very similar conditions, the presence of pristine Na-montmorillonite does not have any
effect on both PS and PBD blocks. This Tg increase for the PS blocks (which is also enhanced for
longer annealing times) can be interpreted by a selective intercalation of the styrene blocks into the
silicate galleries, which is more important at higher nanofiller content and longer annealing times.

3.3.1.7. Elastomers. Burnside and Giannelis [92] have described the two-step preparation of silicon
rubber-based nanocomposites. First, silanol-terminated poly(dimethylsiloxane) (PDMS, Mwˆ
18,000) was melt blended at room temperature with dimethylditallowammonium-exchanged
montmorillonite then the silanol end groups were cross-linked with tetraethylorthosilicate (TEOS)
in the presence of tin bis(2-ethylhexanoate) as catalyst at room temperature. However, in order to
36 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 27. Trends of the glass transition temperatures of the polystyrene (squares) and polybutadiene (circles) blocks
domains for symmetric (styrene±butadiene±styrene) block copolymer-based nanocomposites (open symbols) and
microcomposites (full symbols) as a function of the filler content (reproduced from [91] with permission).

obtain exfoliated nanocomposites (as determined by featureless XRD patterns), several conditions
were required such as mixing the modified clay and PDMS under sonication and addition of a small
quantity of water (typically corresponding to a monolayer coverage of the silicate surface). Nature of
both silicon matrix and clay modifier play an important role. For example, neither exfoliation nor
intercalation can occur if montmorillonite is modified with benzyl dimethyloctadecylammonium
cation or if a too large water amount is added. On another side, only intercalation was observed when
a PDMS-poly(diphenylsiloxane) random copolymer containing 14±18 mol% diphenylsiloxane units
was used. The results stress out again the key importance to get a right match between matrix and
organoclay in order to optimize the layered silicate exfoliation.
More recently, Wang et al. [93] have prepared a series of intercalated nanocomposites based on
the same type of silicon rubber. In their synthesis, they dispersed a hexadecyltrimethylammonium-
exchanged montmorillonite in a silanol-terminated PDMS (Mwˆ68,000). After heating the
dispersion at 908C for 8 h, TEOS and dibutyltin dilaurate were added and the materials were
further cured at room temperature for 12 h. Under these conditions, a silicone rubber-based
nanocomposite was obtained with an intercalated structure as determined by the increase of the
interlayer spacing from 20.2 A Ê (organoclay) to 37.1 A Ê for the composites. TEM observation
confirms the intercalation but also indicates that the filler is uniformly dispersed in the rubber matrix
as crystallites about 50 nm thick.
Okada and co-workers [94,95] obtained a nitrile rubber (NBR)-based nanocomposite in a two-
step synthesis. They first modified a Na-montmorillonite through cation exchange with an amino-
end-capped poly(butadiene-acrylonitrile) oligomer (Mwˆ3400) cationized by HCl in water. This
modified clay was then melt blended on a two-roll mill with NBR and usual additives for
vulcanization such as sulfur and ZnO were added in order to obtain vulcanized rubber sheets after
compression molding at 1608C for 15 min. Even if no direct and objective evidence of the `nano'
structure is given, a large number of properties (gas permeability, enhanced mechanical properties,
. . .) tends to demonstrate that the behavior of these NBR-based composites is in the range of what is
usually observed for nanocomposites.

3.3.2. Melt intercalation of unmodified montmorillonite


Poly(ethylene oxide) (PEO), is a polymer that readily intercalates by melt blending in pristine
Li or Na‡ montmorillonite. Vaia et al. [96] have prepared PEO-based intercalated nanocomposites
‡
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 37

Fig. 28. DSC traces for poly(ethylene oxide)/Na-montmorillonite mixtures heated to 808C for 0, 2 and 6 h (reproduced
from [96] with permission).

by `statically' annealing at 808C a pelletized mixture of PEO and Li‡ (or Na‡) montmorillonite
(CECˆ80 meq/100 g). Under these conditions, complete intercalation arises within 6 h, increasing
the interlayer spacing from 11.4 A Ê (for the pristine montmorillonite) to 17.7 AÊ for the intercalated
species. It has to be noted that this last value corresponds exactly to what has been obtained through
preparation by the exfoliation±adsorption method. The intercalation is further observed through
FTIR measurements where the C±H stretching vibration within the PEO chain at 2900 cmÿ1 is split
into a doublet at 2910 and 2878 cmÿ1 due to polymer host interactions. Finally, the disappearance of
the melting transition of PEO after a 6 h annealing at 808C (Fig. 28) indicates that the intercalation
of PEO leads to a confinement of the macromolecules that prevents the polymer from crystallizing.
In addition to the melting enthalpy decrease, the melting endotherm also shifts towards lower
temperatures. The authors interpret this phenomenon by the interaction of water molecules
(displaced from the interlayer by the PEO intercalation) with `external' PEO crystallites. Besides
studies dealing with the relaxation of confined chains through either the evolution of the intercalated
PEO glass transition properties [97] or conformation by NMR techniques [10], the PEO intercalation
in unmodified montmorillonite has been interestingly exploited to favor the co-intercalation of
poly(methyl methacrylate) [98]. This co-intercalation is claimed to be a potential way to enhance the
ionic conductivity of polymer-layered silicate nanocomposites (see also Section 4.4.1).
A theoretical approach to the intercalation of macromolecules in unmodified montmorillonite
has been described by Balazs et al. [77] using both numerical and analytical self-consistent field
theory. They showed that it is theoretically possible to promote the exfoliation of an unmodified
montmorillonite by melt blending it with a polymer mixture comprising the desired polymer matrix
and a small amount of end-functionalized polymer whose terminal function could strongly interact
with the silicate layers. This end-functionalized polymer could also be replaced by a diblock
copolymer, one sequence of which can intercalate (like PEO) while the other one is either of the
same nature than the polymer matrix or at least compatible with it. Under these conditions (but
without taking into account kinetics factors), exfoliation would occur at a fraction of end-
functionalized polymer (or diblock copolymers) as low as 5 mol%.
These theoretical results have found some experimental confirmations through the work of
Fisher et al. [99] which describes the intercalation and exfoliation of unmodified layered silicates
38 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

(but also double layered hydroxides) by diblock copolymers comprising one block that can
intercalates, e.g. PEO or poly(2-vinylpyridine), and a more hydrophobic block such as PS or PMMA.
They first prepared an intercalated nanocomposite by `statically' annealing the layered silicate
(hectorite, saponite or montmorillonite) with the diblock copolymer then, in a second step, partially
exfoliated this nanocomposite by melt blending with a bulk polymer, the nature of which
corresponds to the hydrophobic block of the copolymer. In the first step, intercalation is observed by
XRD analysis where an increase in the interlayer spacing analogous to what is observed for pure
PEO can be recorded. Moreover, depending on the relative length of the hydrophilic and
hydrophobic sequences, a partially exfoliated nanocomposite can be formed as claimed by the
authors for a diblock with a short PEO block (Mwˆ1000) and a longer PS block (Mwˆ3000). In the
second step, the dispersion by extrusion of these intercalated or partially exfoliated nanocomposites
in the corresponding hydrophobic polymer at least maintains the state of dispersion observed in the
first step. This process, using unmodified layered silicates and commercially available diblock
copolymers could be an alternative towards the need of carefully organo-modified clays.

3.4. Template synthesis

A last technique reported for preparing layered silicate-based nanocomposites implies the in situ
hydrothermal crystallization of the clay layers (hectorite) in an aqueous polymer gel medium where
the polymers often act as template for the layers formation. This method is particularly adapted
to water soluble polymers, and some attempts have been achieved with polymers such as
poly(vinylpyrrolidone) (PVPyr), hydroxypropylmethylcellulose (HPMC), poly(acrylonitrile) (PAN),
poly(dimethyldiallylammonium) (PDDA) and poly(aniline) (PANI) [100]. The typical method for in
situ hydrothermal crystallization of a polymer/hectorite nanocomposite consists in refluxing for 2
days a 2 wt.% gel of silica sol, magnesium hydroxide sol, lithium fluoride and the desired polymer in
water. XRD patterns attest for the formation of a polymer/hectorite intercalated nanocomposite. It is
worth pointing out that the interlayer spacing linearly depends upon the wt.% of polymer
incorporated as shown in Fig. 29 for PVPyr matrix.
Delaminated structures are suspected in case of PANI and PAN, unusually weak peaks
corresponding to the interlayer spacing are indeed observed in the XRD patterns. As far as PDDA is
concerned, the polymer loading cannot exceed 20 wt.%. It seems that polymer incorporation
within the growing layers are limited to the strict balance between negative charges of the clay layers
and cations borne by the polymer chains. It has to be noted that the size of the layers obtained by this

Fig. 29. Correlation between d spacing from XRD patterns and weight percent polymer in synthetic PVP-hectorite clay
nancomposites (reproduced from [100] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 39

template synthesis cannot compete with natural layered silicates for kinetic reasons and their average
length is limited, at best, to about one-third of their natural counterparts.

4. Properties

Layered silicate nanofillers have proved to trigger a tremendous properties improvement of the
polymers in which they are dispersed. Amongst those properties, unexpected large increase in
moduli (tensile or Young's modulus and flexural modulus) of nanocomposites at filler contents
sometimes as low as 1 wt.% has drawn a lot of attention. Thermal stability and fire retardancy
through char formation are other interesting and widely searched properties displayed by
nanocomposites. Those new materials have also been studied and applied for their superior barrier
properties against gas and vapor transmission. Finally, depending on the type of polymeric materials,
they can also display interesting properties in the frame of ionic conductivity or thermal expansion
control.

4.1. Mechanical properties

4.1.1. Effect on tensile properties

4.1.1.1. Young's modulus. The Young's modulus (or tensile modulus), expressing the stiffness of a
material at the start of a tensile test, has shown to be strongly improved when nanocomposites are
formed. Nylon-6 nanocomposites obtained through the intercalative ring opening polymerization of
e-caprolactam (see Section 3.2.1), leading to the formation of exfoliated nanocomposites, show a
drastic increase in the Young's modulus at rather low filler content (Table 12).
Actually, the material stiffness is substantially enhanced whatever the way of preparation:
polymerization within organo-modified montmorillonite (NCH) [18], polymerization within
protonated e-caprolactam swollen montmorillonite (L-NCH) [50], and polymerization within natural
montmorillonite, in the presence of e-caprolactam and an acid catalyst (one-pot-NCH) [51]. The
dependence of Young's modulus measured at 1208C for exfoliated nylon-6 nanocomposites with
various clay contents obtained by in situ intercalative polymerization of e-caprolactam in the

Table 12
Effect of nylon-6-based nanocomposite preparation on the Young's modulus related to the filler content and the average
molecular weight of the matrix
Sample preparation Filler content (wt.%) MW (103) Young's modulus (GPa)
Commercial nylon-6 0 13.0 1.11
NCCa 5 13.0 1.06
NCHb 4.7 16.3 1.87
c
L-NCH 5.3 19.7 2.04
d
One-pot-NCH 4.1 22.6 2.25
a
NCC: montmorillonite-based nylon microcomposite.
b
NCH: nanocomposite obtained by in situ intercalative polymerization of e-caprolactam in protonated aminodode-
canoic modified montmorillonite [18].
c
L-NCH: nanocomposite obtained by in situ intercalative polymerization of e-caprolactam in protonated e-caprolactam
modified montmorillonite [50].
d
One-pot-NCH: nanocomposite obtained by in situ intercalative polymerization of e-caprolactam with Na-
montmorillonite [51].
40 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 30. Dependence of tensile modulus E at 1208C on clay content for organo-modified montmorillonite and saponite-
based nanocomposites (reproduced from [18] with permission).

presence of protonated aminododecanoic acid-modified montmorillonite (average length: 1000 A Ê)


Ê
and saponite (500 A) [18] is shown in Fig. 30.
This dependence clearly indicates that the ability of dispersed silicate layers to increase the
Young's modulus of nylon-6-based nanocomposites can be directly related to the average length of
the layers, hence to the aspect ratio of the dispersed nanoparticles. Moreover, the difference in the
extent of exfoliation, as observed for nylon-6-based nanocomposites synthesized by in situ
intercalative polymerization of e-caprolactam using Na-montmorillonite and various acids [51],
strongly influences the measured Young's modulus values (Table 13).
Depending on the nature of the acid added to catalyze the polymerization, one can observe
variation of the XRD peak intensity (Im) that is inversely related to the amount of exfoliated layers in
the nanocomposite. For an increase in the Im values, a parallel decrease in the Young's modulus is
observed, indicating that exfoliated layers are the main factor responsible for the stiffness
improvement, while intercalated particles, having a less important aspect ratio, rather play a minor
role. All these observations are furthermore confirmed in Fig. 31 which presents the evolution of the
Young's modulus at room temperature of nylon-6 nanocomposites obtained by melt intercalation in
function of the filler weight content measured in this case at room temperature [85].
The preparation of nanocomposites by this technique has the advantage to use the same matrix
for each composite, thus with the same Mw and MWD nylon-6. Fig. 31 shows a constant and large
increase in the modulus up to ca. 10 wt.% of nanoclay, above this threshold the Young's modulus
seems to level off. This change exactly corresponds to the passage from totally exfoliated structure
(below 10 wt.%) to partially exfoliated±partially intercalated structure (for 10 wt.% and upper) as
determined by XRD and TEM analyses [85].
The same behavior can account for the evolution of Young's modulus in polypropylene
nanocomposites obtained by melt intercalation when the amount of maleic anhydride-modified PP

Table 13
XRD peak intensity (Im) and Young's modulus of various nylon-6-based nanocomposites obtained by a one-step in situ
intercalative polymerization of e-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
Benzenesulfonic acid 280 1.74
Acetic acid 555 1.63
Trichloroacetic acid 585 1.67
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 41

Fig. 31. Effect of clay content on tensile modulus, measured at room temperature, of organomodified montmorillonite/
nylon-6-based nanocomposite obtained by melt intercalation (reproduced from [85] with permission).

(PP-MA) added to increase intercalation (and to possibly favor exfoliation, see Section 3.3.1) is
varied [88]. Results are reported in Table 14 and compared to the corresponding microcomposite as
well as to simple PP-MA/PP polymer blends.
One can immediately see that increasing the amount of PP-MA (from sample PPCH 1/1 to
PPCH 1/3) not only improves intercalation or partial exfoliation, but increases also the modulus
value. Comparison of PP with the simple PP-MA/PP blends rules out any possible effect of some
matrix modification due to the presence of increasing amounts of PP-MA (entries 1±3, Table 14).
EVA-based nanocomposites obtained by melt intercalation within dimethyldioctadecylammo-
nium modified montmorillonite [90] also display a non-linear evolution of the relative Young's
modulus with the filler content (vol.%) (Fig. 32).
As XRD and TEM observations for each filler content indicate that both intercalation and
exfoliation occur, the non-linear increase in the relative tensile modulus may be explained by a
decrease in exfoliated particle fraction at higher filler content. Another possible explanation would
take into account a continuous variation of the mean aspect ratio of the primary particles, decreasing
when the filler content is increased. This explanation is supported by the way the experimental points
in Fig. 32 follow the theoretical curves calculated for various aspect ratios, i.e. f, from 12.5 to 20,
using the modified Guth model [101], initially studied to describe the evolution of Young's modulus
in true elastomeric matrices filled with carbon black particles that organize in high aspect ratio
structures. Actually, at very low filler content (volume fractionˆ0.05, that corresponds, to 1 wt.%),
experimental values for the relative tensile modulus range above the higher theoretical curve (drawn

Table 14
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
PPCC 6.9 0 830
PPCH 1/1 7.2 7.2 838
PPCH 1/2 7.2 14.4 964
PPCH 1/3 7.2 21.6 1010
a
PPˆpolypropylene; PP-MA x: polypropylene modified by maleic anhydride (xˆwt.% of PP-MA in the blend);
PPCCˆpolypropylene-based microcomposite; PPCH y/zˆpolypropylene-based nanocomposite (y/zˆweight ratio between
y parts of filler and z parts of PP-MA).
42 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 32. Evolution of the relative Young's modulus (Ey (nanocomposite)/Ey (unfilled EVA)) with the filler volume fraction
and confrontation of the experimental results to the modified Guth model for different aspect ratios. Modified Guth model:
Ey (composite)/Ey (matrix)ˆ1‡0.67Ffrv‡1.62Ff2 r2v where Ey is the Young's modulus, rv the filler volume fraction and Ff
is the aspect ratio of the filler particles.

for fˆ20). Then, at higher filler content, they level off and fit curves calculated for much lower
aspect ratios. This suggests that silicate nanolayers with very high shape factor are only predominant
at very low filler content.
In contrast to the above results, when simply intercalated structures (without any exfoliation)
are concerned, such as for PMMA [43] or PS [45] based nanocomposites, obtained by emulsion
polymerization in presence of water-swollen Na-montmorillonite, the increase in Young's modulus is
relatively weak, going, e.g. from 1.21 to 1.30 GPa for pure PMMA and PMMA containing 11.3 wt.%
intercalated montmorillonite, respectively. This again attests for the inefficiency of intercalated
structures to improve the stiffness of the so-obtained nanocomposites.
Exfoliation of layered materials such as magadiite in an elastomeric epoxy matrix [74] also
gives rise to a noticeable increase in the Young's modulus of the obtained composites as depicted in
Fig. 33.

Fig. 33. Comparison of the evolution of tensile modulus with filler content for nanocomposites based on an epoxy matrix
and various organomodified fillers. C18A-montmorilloniteˆmontmorillonite modified with octadecylammonium C18A-
magadiiteˆmagadiite modified with octadecylammonium C18A1M-magadiiteˆmagadiite modified with methyl-
octadecylammonium (reproduced from [74] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 43

Fig. 34. Tensile modulus vs. organoclay loadings for elastomeric polyurethane/clay nanocomposites (reproduced from
[75] with permission).

In Fig. 33 is shown the evolution of the tensile modulus with filler loading for three layered
materials: a montmorillonite modified with octadecylammonium cation (C18A-montmorillonite), a
magadiite modified with the same alkylammonium (C18A-magadiite) and a magadiite modified with
methyl-octadecylammonium cation (C18A1M-magadiite). This figure shows a much significant
increase of the tensile modulus for the montmorillonite-based nanocomposites for filler contents
higher than 4 wt.%. The authors explain this behavior by the difference in layer charge density for
magadiite and montmorillonite. Organomagadiites have a higher layer charge density and
consequently a higher alkylammonium content than organomontmorillonite. As the alkylammonium
ions interact with the epoxy resin while polymerizing, dangling chains are formed. More of these
chains are thus formed in presence of organomagadiites. These dangling chains are known to weaken
the polymer matrix by reducing the degree of network cross-linking, then compromising the
reinforcement effect of the silicate layer exfoliation.
In a pure elastomeric matrix, exfoliation does not appear to be a prerequisite to improve the
material stiffness. Indeed, it has been reported that cross-linked soft polyurethane-based
nanocomposites can be characterized by a two-fold increase of the tensile modulus upon filling
with 10 wt.% of organoclay (Fig. 34) [75].
Here, no exfoliation has been observed, rather the highly regular intercalation, with interlayer
spacing as high as 50.8 A Ê (meaning that more than one polymer chain is intercalated), would explain
the mechanical properties improvement.
A large increase in the tensile modulus for an exfoliated structure is also observed for thermoset
matrices [71,72]. Fig. 35 shows the evolution of modulus for various amine-cured epoxy-based
nanocomposites filled with 2 wt.% montmorillonite previously modified by alkylammonium cations
of different length [72].
While the montmorillonite modified with butylammonium only gives an intercalated structure
with a low tensile modulus, the other three nanocomposites with alkyl chains of 8, 12 and 16 carbons
are characterized by exfoliated structures as determined by TEM and XRD, and consecutively give
much higher modulus values.
However, Zilg et al. [73] have reported about rather weak stiffness improvements in the case of
anhydride-cured epoxy-based nanocomposites when true exfoliated structures were observed. For
these authors, the real key for the matrix stiffness improvement resides in the formation of
supramolecular assemblies obtained by the presence of dispersed anisotropic laminated
nanoparticles. They also describe a stiffening effect when the montmorillonite is modified by a
44 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 35. Dependence of tensile modulus of amine-cured epoxy/clay nanocomposites on onium ion carbon number at clay
loadings of 2 wt.%.

functionalized organic cation (carboxylic acid or hydroxyl groups) that can interact with the matrix
during curing.

4.1.1.2. Stress at break. In thermoplastic-based (intercalated or exfoliated) nanocomposites, the


stress at break, which expresses the ultimate strength that the material can bear before break, may
vary strongly depending on the nature of the interactions between the matrix and the filler as shown
in Table 15.
Even if caution has to be taken concerning the absolute tensile stress at break values (most of
the described nanocomposites were obtained by in situ polymerization so as properties may be
influenced by changes in the matrix molecular parameters as well), the differences observed are
usually sufficiently sizeable to draw some conclusions. Filled polymers such as exfoliated nylon-6-
Table 15
Tensile stress evolution for nanocomposites based on various thermoplastic matrices
Matrix Matrix tensile Nanofiller content Nanocompo Nanocomposite
stress (MPa) (wt.%) site type tensile stress (MPa)
Nylon-6 68.6 4.7 NCHa 97.2
b
68.6 5.3 L-NCH 97.3
68.6 4.1 One-pot-NCHc 102
PMMA 53.9 12.6 Intercalated 62.0
53.9 20.7 Intercalated 62.0
PP 31.4 5.0 Intercalatedd 29.5
32.6 4.8 Intercalatede 31.7
(‡exfoliated?)
PS 28.7 11.3 Intercalated 21.7
28.7 17.2 Intercalated 23.4
28.7 24.6 Intercalated 16.6
28.7 34.1 Intercalated 16.0
a
NCH: exfoliated nanocomposite prepared by in situ intercalative polymerization of e-caprolactam in protonated
aminododecanoic acid modified montmorillonite.
b
L-NCH: exfoliated nanocomposite prepared by in situ intercalative polymerization of e-caprolactam in
montmorillonite pre-intercalated with e-caprolactam.
c
One-pot-NCH: exfoliated nanocomposite prepared by in situ intercalative polymerization of e-caprolactam activated
by phosphoric acid in Na-montmorillonite.
d
PP added with PP-MA so as a PPCH 1/1 is reached (see Table 14).
e
PP added with PP-MA so as a PPCH 1/3 is reached (see Table 14).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 45

based nanocomposites prepared following different methods [18,50,51] or intercalated PMMA-based


nanocomposites [43] exhibit an increase in the stress at break, that is usually explained by the
presence of polar (PMMA) and even ionic interactions (nylon-6 grafted onto the layers) between the
polymer and silicate layers. This increase appears to be much more pronounced in case of nylon-6
which has both an exfoliated structure and ionic bonds with the silicate layers. As far as
polypropylene-based nanocomposites are concerned [88], no or only very slight tensile stress
enhancement are measured (Table 15). This behavior can be partially explained by the lack of
interfacial adhesion between apolar PP and polar layered silicates. Addition of maleic anhydride
modified polypropylene to the polypropylene matrix has, however, proved to be favorable to the
intercalation of the PP chains and maintains the ultimate stress at an acceptable level. In PS-
intercalated nanocomposites [45], ultimate tensile stress is even much decreased compared to PP
matrix and further drops down at higher filler content. This lack of properties is attributed by the
authors to the fact that only weak interactions exist at the poly(styrene)-clay interface contrary to the
previous compositions in which (stronger) polar interactions may take place, strengthening the filler
matrix interface.
Epoxy resins-based nanocomposites display a totally different behavior depending upon their
glass transition temperature, located above or below room temperature.
In high Tg epoxy thermosets [72,73], neither intercalated nor exfoliated nanosilicates lead to an
improvement of the tensile stress at break, they rather make the materials more brittle. This effect
appears to be generally more pronounced for intercalated structures than for exfoliated ones. In
contrast, nanocomposites based on both epoxy [71,74] and polyurethanes [75] elastomeric matrices
exhibit a sizeable increase in tensile stress at break upon the addition of small quantities of
nanofillers. This increase follows qualitatively those observed previously for the Young's modulus
measurements. The same increase is also observed for silicon rubber-based nanocomposites [93] as
shown in Fig. 36.
In this study, the increase in tensile stress at break in function of the volume content of filler for
a partially-intercalated partially-exfoliated nanocomposite is compared to a composite filled with
silica anisotropic nanoparticles (5±20 nm), coined as aerosilica. It appears thus that for low glass
transition temperature cured nanocomposite materials, the tensile strength increase does not rely
upon the aspect ratio of the dispersed particle but rather on the presence of nanoparticles dispersed in
the cross-linked soft matrix.

Fig. 36. Tensile strength of nanocomposites vs. volume content of filler: (square) silicone rubber/organo-modified
montmorillonite nanocomposites; (triangle) silicone rubber/aerosilica nanocomposites (reproduced from [93] with
permission).
46 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 37. Comparison of the strain at break values for an exfoliated epoxy/magadiite nanocomposite prepared from
magadiite modified with methyl-octadecylammonium ion (C18A1M), an intercalated nanocomposite prepared from
magadiite modified with trimethyloctadecylammonium ion (C18A3M) and a conventional composite prepared from
magadiite modified with octadecylammonium ion (C18A) (reproduced from [74] with permission).

4.1.1.3. Elongation at break. The effect of nanocomposite formation on the elongation at break has
not been widely investigated. When dispersed in thermoplastics such as for intercalated PMMA [43]
and PS [45] or intercalated±exfoliated PP, the elongation at break is reduced. In the last case, the
decrease is very important, dropping from 150 and 105% for a pure PP matrix and a 6.9 wt.% non-
intercalated clay microcomposite, respectively, down to 7.5% in the better case for a PP-based
nanocomposite filled with 5 wt.% silicate layers.
Interestingly enough, such a loss in ultimate elongation does not occur in elastomeric epoxy
[74] or polyol polyurethane matrices [75]. Rather, the addition of a nanoclay in cross-linked matrices
triggers an increase of the elongation at break as clearly depicted in Fig. 37.
When a conventional composite (magadiite exchanged with octadecylammonium cation,
C18A) is prepared, a drop in the elongation at break is observed as expected for such a
material while an intercalated nanocomposite (magadiite exchanged with trimethyloctade-
cylammonium cation, C18A3M) tends to slightly improve this property. But the exfoliated
nanocomposite as prepared with methyloctadecylammonium cation (C18A1M) displays a large
increase of the elongation at break. The improvement in elasticity may be attributed in part to the
plasticizing effect of the gallery oniums and to their contribution to the formation of dangling chains
but also probably to conformational effects at the clay-matrix interface. The combination of
improved stiffness (Young's modulus), toughness (stress at break) and elasticity (strain at break) is
quite exceptional and make elastomeric nanocomposites a new family of highly performant
materials.
Another matrix which exhibits both an increase in stress and elongation at break is poly(imide),
PI [102]. Indeed, when filled by montmorillonite exchanged with hexadecylammonium cation
(16CNH), these properties increase with the filler loading at least up to a 5 wt.% filler content. At
higher filler content, both properties experience a sharp drop towards values lower than those
recorded for the filler-free matrix (see Fig. 38).
This behavior is explained by the formation of non-exfoliated aggregates at higher filler content,
that makes these composites much more brittle.

4.1.2. Impact properties


Impact properties have been measured for nylon-6-based nanocomposites prepared either by in
situ intercalative polymerization of e-caprolactam using protonated aminododecanoic acid-
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 47

Fig. 38. Tensile strength and elongation at break in function of filler content for poly(imide)-based nanocomposites filled
with montmorillonite modified with hexadecylammonium ion (reproduced from [102] with permission).

exchanged montmorillonite [18] or by melt intercalation of nylon-6 in octadecylammonium-


exchanged montmorillonite [85]. Both methods lead to exfoliated nanocomposites especially when
the filler content does not exceed 10 wt.% (at higher filler level, melt-intercalation provides partially-
exfoliated±partially-intercalated materials). The formation of nylon-6-based nanocomposites does
not reduce too much the impact properties, whatever the exfoliation process used. In the case of in
situ intercalative polymerization, the Izod impact strength is reduced from 20.6 to 18.1 J/m when
4.7 wt.% of nanoclay is incorporated. Charpy impact testing shows 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. Fig. 39 shows that the decrease in the Izod impact strength of melt-intercalated
nylon-6-based nanocomposites is not too much pronounced over a relatively large range of filler
content.
This relatively good resistance to impact, together with a high Young's modulus, good flexural
modulus and a remarkable enhancement of the increase in the heat distortion temperature (i.e. a
measure of the material softening point) going from 658C for pure nylon-6 to more than 1508C for
nanocomposites, have allowed this material to replace glass fiber reinforced nylon or
poly(propylene) in the production of timing belt covers of automotive engines [95]. The belt cover,
obtained by injection-molding, shows good rigidity, excellent thermal stability and no wrap. It
moreover saves weight up to 25% due to the very small content of inorganic material in the final
composition.

Fig. 39. Effect of clay content on notched Izod impact strength of nylon-6/clay nanocomposites obtained trough melt
intercalation (reproduced from [85] with permission).
48 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 40. Temperature dependence of E0 and tan d for unfilled poly(styrene) and an intercalated nanocomposite (17.2 wt.%
of montmorillonite) (reproduced from [45] with permission).

4.1.3. Dynamic mechanical analysis


Dynamic mechanical analysis (DMA) measures the response of a given material to a cyclic
deformation (usually tension or three-point flexion type deformation) as a function of the
temperature. DMA results are expressed by three main parameters: (i) the storage modulus (E0 ),
corresponding to the elastic response to the deformation; (ii) the loss modulus (E00 ), corresponding to
the plastic response to the deformation and (iii) tan d, that is the (E0 /E00 ) ratio, useful for determining
the occurrence of molecular mobility transitions such as the glass transition temperature.
DMA analysis has been studied to track the temperature dependence of storage modulus upon
the formation of an intercalated PS nanocomposites [45]. Fig. 40 shows the temperature dependence
of E0 and tan d for pure PS and a nanocomposite intercalated with 17.2 wt.% of Na-montmorillonite
as synthesized by exfoliation±adsorption during emulsion polymerization (see Section 3.1.3).
No significant difference in E0 can be seen in the investigated temperature range, indicating that
intercalated nanocomposites do not strongly influence the elastic properties of the matrix. On the
other hand, the shift and broadening of the tan d peak towards higher temperatures for the
nanocomposite indicate an increase in the glass transition temperature together with some
broadening of this transition. This behavior has been ascribed to the restricted segmental motions at
the organic±inorganic interface neighborhood of intercalated compositions.
Intercalation of PS sequences plays nevertheless a much more important role in the dynamic
mechanical properties of symmetric (styrene±butadiene±styrene) block copolymers [91]. As
previously described (see Section 3.3.1) when SBS are melt blended with a montmorillonite
modified by dimethyldioctadecylammonium cation, a nanocomposite is formed, in which only the
PS blocks can intercalate within the layered silicates. This particular structure provides a material
with a sizeable improvement of the storage modulus at 258C as depicted in Fig. 41.
This figure compares the values of the storage modulus for two sets of samples for which the
filler content is varied from 0 to 30 wt.%. The first series (open squares) displays the values recorded
for nanocomposites filled with organo-modified clay and the second one (filled squares) shows the
results obtained for composites prepared by melt-blending the SBS matrix and Na-montmorillonite
under the same conditions (microcomposites). One can clearly see the large increase in elastic
modulus for nanocomposites while microcomposites do not present any improvement for this
property, whatever the filler content be.
Increase in storage modulus related to a better nanoclay dispersion is further demonstrated in
the case of polypropylene-based nanocomposites (see Table 14). The degree of nanofiller dispersion
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 49

Fig. 41. Trend of the storage modulus E0 at 258C for SBS-based nanocomposites (&&) and microcomposites (&&) as a
function of the filler level (reproduced from [91] with permission).

can be tuned up either by using different amounts of a given PP modified with maleic anhydride
(acid valueˆ52 mg KOH/g) [88] or can even be drawn towards the formation of exfoliated
nanocomposites by playing with the relative functionalization content all along of the modified PP
(using a PP-MA with an acid valueˆ26 mg KOH/g) [87]. Varying the relative amount of PP-MA
within the PP matrix highly modifies the temperature dependence of the storage modulus reduced to
the value of unfilled PP matrix as described in Fig. 42a). Below Tg (located around 138C for all the
composites) the relative modulus values of the nanocomposites are not so much enhanced (1.2±1.3)
compared to the pure matrix while above Tg, an important increase of the moduli is observed,
reaching a maximal value of 1.76 around 808C, in the case of the PPCH 1/3 composite. On the other
hand, the relative storage modulus of the microcomposite (PPCC) stays relatively low and quickly
reaches a plateau value around 1.26. One can further observe that the storage moduli show a sharp
decline above 1408C. The main reason for this decline resides in the fact that the softening point of
the PP-MA matrix is reached at this temperature, strongly reducing the elastic response of the
material.
Fig. 42b shows the relative storage modulus values in reference to the unfilled PP/PP-MA
blends properties instead of the pure PP matrix. The very large increase observed for the PPCH 1/3
composition can be directly related to the far better dispersion obtained for this nanocomposite (see
Section 3.3.1). When the dispersion is further improved by the use of a more compatible PP-MA
(code 1001; 26 mg KOH/g), higher storage moduli are reached (Fig. 43).
A two-fold increase in the relative moduli (E0 /E0 -matrix) is even measured for a nanocomposite
based on organo-modified montmorillonite nanoparticles (PPCH-C18-Mt/1001). Interestingly
enough, a maximum value of 2.4 is reached by substituting synthetic fluorinated mica (PPCH-
C18-Mc/1001) for previously studied montmorillonite. The behavior difference for the two types of
filler was not explained by the authors but it is more likely related to the respective aspect ratio of the
dispersed particles since mica layers are known to be much longer than montmorillonite layers.
The influence of dispersion and length of the layered particles is further demonstrated in case of
poly(imide)-based nanocomposites using various organoclays (hectorite, saponite, montmorillonite
and synthetic mica, see Table 3) [38]. In this study, exfoliated structures were obtained for mica and
montmorillonite clays while a partially-exfoliated±partially-intercalated structure was found for
saponite and a mainly intercalated morphology was attributed to the hectorite-based nanocomposite.
Fig. 44 shows the temperature dependence of the storage modulus for these nanocomposites filled
with 2 wt.% of clay and for the unfilled matrix.
50 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 42. Temperature dependence of the relative storage modulus (E0 ): (a) storage modulus of PPCHs nanocomposites and
PPCC microcomposite relative to the storage modulus of the PP matrix and (b) storage modulus of PPCH 1/1 and 1/3
relative to the corresponding PP-MA matrices. See Table 14 for code explanation (reproduced from [88] with permission).

At a given temperature, higher storage moduli results from the better nanofiller dispersion.
The huge difference between exfoliated montmorillonite and exfoliated mica-based nanocom-
posites may be again explained by the respective aspect ratio of the dispersed silicate layers, with
lengths of 0.218 and 1.23 mm, respectively, for montmorillonite and synthetic mica, as observed by
TEM.
Finally, DMA studies carried out on organoclays exfoliated within cross-linked matrices reveals
again a very marked improvement of the storage modulus, especially above Tg. For instance, for
epoxy matrix below Tg [69], a 58% increase in modulus results from the dispersion of 4 vol.%
montmorillonite with the formation of a well-ordered exfoliated nanocomposite (silicate layers
separated by approximately 100 A Ê ). At 408C, E0 equals 2.44 and 1.55 GPa for the nanocomposite
and the unfilled cross-linked matrix, respectively. But above Tg, e.g. at 1508C, the storage modulus
improvement reaches a 4.5 factor with E0 values of 11 and 50 MPa for the unfilled and filled epoxy,
respectively. Similar behavior is observed for room temperature elastomer such as nitrile rubber [95],
A three-fold increase of the storage elastic modulus is noted by the simple dispersion/exfoliation of
10 parts of organoclay per 100 parts of rubber, with a modulus as high as 8.8 MPa. This value
corresponds to what can be obtained with the same matrix filled with 40 parts of carbon black per
100 parts of rubber, thus reducing by a factor of four the amount of filler.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 51

Fig. 43. Relative dynamic storage moduli (E0 /E0 -matrix) of the PP-based nanocomposites based on maleic anhydride
modified PP (PP-MA/1001) to that of corresponding PP/1001 blends taken as the matrix in function of the temperature
(reproduced from [87] with permission).

In summary, the storage elastic modulus appears to be substantially enhanced at temperatures


above Tg for exfoliated nanocomposites filled with layered silicates of high aspect ratio. A possible
explanation for such an improvement could be the creation of a three-dimensional network of
interconnected long silicate layers, strengthening the material through mechanical percolation.

4.2. Thermal stability and flame retardant properties

Another highly interesting property exhibited by polymer-layered silicate nanocomposites


concerns their increased thermal stability but also their unique ability to promote flame retardancy at
quite low filling level through the formation of insulating and incombustible char.

4.2.1. Thermal stability


The thermal stability of a material is usually assessed by thermogravimetric analysis (TGA)
where the sample mass loss due to volatilization of degraded by-products is monitored in function of
a temperature ramp. When the heating is operated under an inert gas flow (nitrogen, helium, . . .), a
non-oxidative degradation occurs while the use of air or oxygen allows to follow the oxidative
degradation of the sample.
The first indication of thermal stability improvement in nanocomposites appears in the seminal
work by Blumstein [103] who studied the thermal stability of PMMA intercalated within

Fig. 44. Temperature dependence on storage elastic modulus for poly(imide)-based nanocomposites filled with 2 wt.% of
organo modified synthetic mica, montmorillonite, saponite and hectorite (reproduced from [39] with permission).
52 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

montmorillonite. In this work, the author shows that a 10 wt.% clay intercalated PMMA (produced
by free radical polymerization of the intercalated monomer) degrades at temperature 40±508C
superior to the degradation of the pure unfilled PMMA matrix. He also found that the thermal
stability of the PMMA extracted from the montmorillonite was higher than for a PMMA
conventionally produced in solution. The higher stability of PMMA synthesized by in situ
intercalative polymerization is more likely due to a decrease in the relative amount of PMMA end-
capped by carbon±carbon double bond, as a result of reduced propensity to disproportionation
reactions. Extracted PMMA chains were nevertheless less stable than when intercalated, and the
author proposed that the enhanced thermal stability of the PMMA-based nanocomposites was not
only due to difference in chemical structure, but also to restricted thermal motion of the
macromolecule in the silicate interlayer.
Since then and more particularly in the 1990s, several authors have drawn attention to the
thermal stabilization brought by the nanocomposites. Burnside and Giannelis [92] have measured by
TGA (under nitrogen flow) the thermal stability of cross-linked poly(dimethylsiloxane) in which
10 wt.% of organomontorillonite was exfoliated. When compared to unfilled cross-linked PDMS
(Fig. 45), the nanocomposite TGA trace shows a drastic shift of the weight loss towards higher
temperature, with a stabilization as high as 1408C at 50% weight loss.
The authors attributed the much better thermal stability to hindered out-diffusion of the volatile
decomposition products (mainly cyclic silicates), as a direct result of the decrease in permeability,
usually observed in exfoliated nanocomposites (see Section 4.3). More recently, another group [93]
has produced nanocomposites based on cross-linked PDMS using slightly different operating
conditions in order to produce mainly intercalated structures. In this case, the increase in thermal
stability for a nanocomposite intercalated with 8.1 vol.% of organomontmorillonite was limited to
about 608C at 50% of weight loss. The authors proposed other possible origins for the observed
thermal stability improvement (which was also reported, in the same study, for silica-based
nanocomposites) such as some inactivation of the centers active in silicone main chain
decomposition by interaction with the filler or by prevention of the unzipping degradation from
occurring through physical and chemical cross-linking points built up between polymer chains and
filler particles.
Increase in thermal stability has also been reported for intercalated nanocomposites prepared by
emulsion polymerization of methyl methacrylate [43], styrene [45] and epoxy precursors [44] in the

Fig. 45. TGA traces for PDMS (solid line) and PDMS nanocomposite (dashed line) containing 10 wt.% organo-modified
montmorillonite (reproduced from [92] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 53

Fig. 46. Decomposition onset temperature as a function of filler loading of PS-based nanocomposites obtained with
montmorillonite modified with: (filled square) dimethyl(hydrogenated tallow alkyl)benzyl ammonium ion (filled circle)
dimethyl di(hydrogenated tallow alkyl) ammonium ion (filled triangle) dimethyl(hydrogenated tallow alkyl) 2-ethylhexyl
ammonium ion (open square) Na-montmorillonite (reproduced from [60] with permission).

presence of water swollen Na-montmorillonite. In every case, a high temperature increase in the
decomposition onset was observed.
Doh and Cho [60] have measured by TGA under nitrogen atmosphere the onset of thermal
decomposition of intercalated PS-based nanocomposites produced by in situ polymerization of
styrene within various organo-modified montmorillonites (see Table 8). In Fig. 46 are collected these
decomposition onset temperatures of PS-based nanocomposites filled with increasing filler content
together with a Na-montmorillonite-based microcomposite for the sake of comparison.
It is clearly seen that a large increase in the onset of decomposition occurs for nanocomposites
at very low filler content and quickly levels off. The threshold is reached for a filler content as low as
0.3 wt.% when intercalating an organoclay (modified with a dimethylbenzyloctadecylammonium
cation) well compatible with PS. In contrast, Na-montmorillonite does not modify a lot the
decomposition onset of the PS matrix. This is another widely searched characteristic feature of
nanocomposites in which the thermal properties improvement arises at very low filler content, often
making the obtained material cheaper, lighter and easier to process than more conventional
microcomposites.
Another key factor that may determine the extent of the thermal stabilization in nanocomposites
could also arise from the actual nature of the thermal degradation mechanism, often different from a
polymer to another. For example, when poly(imide) exfoliated nanocomposites [41] are thermally
degraded under nitrogen, their thermal stability is only enhanced by about 258C (at 50% of weight
loss) which is much less than the 1408C jump observed in exfoliated PDMS nanocomposites.
Without any doubt, the chemical nature of the studied polymeric material and its degradation
mechanism play here an important role.
Finally, the experimental conditions of the material degradation have proved to highly influence
the history and mechanism of the degradation as well. The thermal stability of EVA partially-
intercalated±partially-exfoliated nanocomposites have been investigated through TGA under helium
(thermodegradation) and under air flow (thermooxidative degradation) [90]. EVA is known to
degrade in two consecutive steps, the first one, identical for both oxidative and non-oxidative
degradations, consists in the loss of acetic acid and occurs between 350 and 4008C. The second step
involves the thermal degradation of the so obtained unsaturated backbone either by radical scission
(non-oxidative route) or by thermal combustion (oxidative route). Fig. 47a presents the TGA traces
54 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 47. Influence of the purge gas on the TGA and DTG traces EVA-based nanocomposite (5 wt.% of organo-modified
montmorillonite, long dashes), microcomposite (5 wt.% of Na-montmorillonite small dashes) and the unfilled matrix (solid
line): (a) under helium flow and (b) under air flow.

and their derivatives (DTG) for, respectively, the unfilled EVA matrix with a vinyl acetate content of
27 wt.%, a 5 wt.% Na-montmorillonite/EVA microcomposite and a 5 wt.% modified montmor-
illonite/EVA nanocomposite as measured under helium at 208C/min. Fig. 47b presents the same
TGA and DTG traces but recorded under air flow.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 55

Table 16
Maximal temperature at the main degradation peak (DTG) as measured under air flow at 208C/min for EVA and EVA-
based nanocomposites with different nanoclay contents
Filler content (wt.%) Degradation temperature maximum (8C)
0 452.0
1 453.4
2.5 489.2
5 493.5
10 472.0
15 454.0

One can note first that in both cases, the Na-montmorillonite (microcomposite) does not
influence the thermal degradation of the matrix. The most striking observation comes out by
comparing the second degradation step under helium and under air flow. While, under helium, the
nanocomposite experienced only a very slight loss in thermal stability (48C), a sizeable increase of
more than 408C at the maximum of the DTG curves is measured under air. An explanation for this
behavior may be found in the formation of char that clearly appears under oxidative degradation.
Char may act as a physical barrier between the polymer medium and the superficial zone where
flame combustion occurs.
Comparison of different nanofiller contents in EVA-based nanocomposites on the degradation
peak in the DTG curve under oxidative degradation, has been also very informative. Results are
collected in Table 16.
Optimal thermal stabilization is obtained at a filler content of ca. 2.5±5 wt.%. Below this value
no thermal stabilization is observed while increasing too much the amount of nanofiller also
decreases the thermal stability. Such a behavior could account from the change in relative proportion
of exfoliated and intercalated species with the filler content. At low filler content, exfoliation
dominates but the amount of exfoliated particles is not high enough to promote the thermal stability
through char formation. When increasing the filler content, relatively more exfoliated particles are
formed, char forms more easily and increases the thermal stability of the nanocomposite until 2.5±
5 wt.% of nanofiller is reached. At higher levels, equilibrium between exfoliation and intercalation is
drawn towards intercalation and, even if char is still formed in high quantity, the morphology of the
nanocomposite does not allow for maintaining a good thermal stability. This explanation, likely valid
for EVA nanocomposites may not be applicable to other polymer matrices as demonstrated in a study
on poly(etherimide) nanocomposites [104] where intercalated nanolayers were found to be better
thermal stabilizers than the exfoliated ones.

4.2.2. Flame retardancy


The flame retardant properties of nanocomposites have been very recently reviewed in detail by
Gilman [105]. The main bench-scale method used to measure important parameters in the flame
retardant behavior of a material (heat release rate, peak of heat release rate, heat of combustion, . . .)
is Cone calorimetry. In a typical experiment, the sample is exposed to a given heat flux (often taken
as 35 kW/m2) and the heat release rate (HRR) as well as the mass loss rate are recorded as a function
of time. It is worth noting that reduction of the peak HRR is the most clear-cut evidence for the
efficiency of a flame retardant. Moreover, gas and soot production are also measured. As a typical
example, Fig. 48 shows the HRR plot obtained for nylon-6 and a nylon-6 exfoliated nanocomposite
(5 wt.% of exfoliated montmorillonite).
A 63% reduction in the peak HRR is clearly observed for the nanocomposite. Cone calorimetry
experiments have been carried out on other nanocomposites such as exfoliated nylon-12 (2 wt.%
56 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Fig. 48. Comparison of the heat release rate (HRR) plots for nylon-6 and nylon-6 layered silicate nanocomposite (5 wt.%)
at 35 kW/m2 heat flux, showing a 63 wt.% reduction in HRR for the nanocomposite (reproduced from [105] with
permission).

organoclay), exfoliated poly(methylmethacrylate-co-dodecylmethacrylate) [106], intercalated PS


(3 wt.%) or intercalated PP (2 wt.%) and for each material, a significant decrease in the peak HRR is
recorded while the heat of combustion, smoke and the carbon monoxide yields (other important
properties in flammability concern) are usually not increased. These data tend to demonstrate that
the improvement in flame retardancy does not occur by a process in the gas phase but rather by a
modification of the combustion process in the condensed phase. Experiments carried out in a
radiative gasification apparatus [107] have allowed to determine that the flame retardant effect of
nanocomposites mainly arises from the formation of char layers obtained through the collapse of the
exfoliated and/or intercalated structures. This multilayered silicate structure may act as an excellent
insulator and mass transport barrier, slowing down the escape of the volatile decomposition products
as observed in nylon-6 but also in thermoset nanocomposites [108]. Whatever the nature of the
matrix (thermoplastics or thermosets) and whatever the structure of the nanocomposite (exfoliated or
intercalated), always the same interlayer spacing (13 A Ê ) was found for the recovered chars as
analyzed by XRD, implying the formation by combustion of a residue of the same nature.
Nylon-6 nanocomposite filled with 2 wt.% nanoclay has also been used as an additive to replace
pentaerythritol (in order to avoid exudation and water solubility) in an intumescent flame retardant
formulation using ammonium polyphosphate, APP [109]. This new formulation has shown very
good fire retardant properties, increasing the low oxygen index (LOI) values by 5% for the best APP/
nylon-6 nanocomposite composition and highly decreasing the HRR values.

4.3. Gas barrier properties

The high aspect ratio characteristic of silicate nanolayers in exfoliated nanocomposites has been
found to highly reduce the gas permeability in films prepared from such nanomaterials.
The permeability to carbon dioxide has been measured for the partially-exfoliated poly(imide)-
based nanocomposites prepared by Lan et al. [40]. Interestingly, the relative permeability values, i.e.
Pc/Pp where Pc and Pp stand for the composite and the unfilled polymer permeability, respectively,
have been plotted in function of the filler volume fraction (Fig. 49).
The curve fitting has been achieved by using a theoretical expression allowing the prediction of
gas permeability in function of the length and width of the filler particles as well as their volume
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 57

Fig. 49. CO2 permeability of polyimide clay composites prepared by curing CH3(CH2)17NH3‡ montmorillonite-poly(amic
acid) films at 3008C. Curves A and C are calculated for filler with an aspect ratio of 20 and 2000, respectively, Curve B was
generated by least squares fitting of the permeability equation to the experimental data. The inset illustrates a possible self-
similar aggregation mechanism for the clay plates (reproduced from [40] with permission).

fraction within the PI matrix. The best fitting is actually obtained for an aspect ratio of 192.
Surprisingly, this value is much smaller than what could be awaited for a truly exfoliated system
structure, meaning an aspect ratio of ca. 2000 for montmorillonite single layers [40]. Explication for
such a low aspect ratio is illustrated by the inset shown in Fig. 49, where can be seen a sketch of the
proposed nanocomposite structure. It consists of the so called `self-similar clay aggregation
mechanism' in which face±face associated and elongated layers are skipped in staircase-like fashion.
These self-similar structures can, therefore, exhibit enhanced aspect ratio.
The effect on water permeability of both partially and totally exfoliated poly(imide)-based
nanocomposites has been reported by Yano et al. [39], using organoclay with different layer lengths
(see Table 3). Fig. 50 presents the clay length dependence of the relative permeability coefficient for
poly(imide) filled with 2 wt.% of organoclay, either exfoliated montmorillonite and synthetic mica,
or intercalated clay tactoids (hectorite and saponite).
A relatively good agreement between the experimental values and the corresponding theoretical
curve can be achieved, as the length of the clay increases, the relative permeability decreases
drastically. In other words, it does mean that the best gas barrier properties will be obtained by fully
exfoliated rather long layered silicates.

Fig. 50. Clay length dependence on the relative permeability coefficient for poly(imide)/clay nanocomposites (reproduced
from [39] with permission).
58 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

Permeability to water vapor has also been investigated for exfoliated nanocomposite based on
poly(e-caprolactone) (PCL), synthesized by in situ intercalative polymerization of the lactone
monomer inside organo-modified montmorillonite [53]. In order to produce nanocomposite films
with different filler contents, PCL nanocomposite containing 15 wt.% (6 vol.%) of exfoliated
montmorillonite was separately prepared and mixed in variable composition with a commercial
preformed PCL by co-dissolution in toluene followed by solvent casting. Again, a dramatic drop in
the relative permeability of the polymer coefficient is triggered by dispersing increasing amounts of
layered nanofiller. A relative permeability of 0.2 (unfilled PCL permeabilityˆ1) is measured for a
filler loading of 4.8 vol.%. Fitting of the Pc/Pp versus filler volume fraction dependency lead the
authors to conclude to an aspect ratio of 70, again well below the value expected for exfoliated
montmorillonite. The authors explained this apparent discrepancy by the fact that the fitting model
was originally developed for particles totally oriented parallel to the film plane. However, under the
experimental conditions used for the film formation, the silicate layers could be not so well aligned
flat along the film surface, with even possible filler aggregation. Finally, it is worth pointing out that
Bayer has recently commercialized a new grade of plastic films for food packaging, Durethan1
LPDU 601, based on nylon-6 exfoliated nanocomposites with improved gas barrier properties
(oxygen transmission rate divided by a factor of two compared with the pure nylon-6), improved
transparency and gloss and increased tensile modulus [110].

4.4. Miscellaneous

4.4.1. Ionic conductivity


Nanocomposites have been also considered by Vaia et al. [96] to tune ionic conductivity of
PEO. An intercalated nanocomposite obtained by melt intercalation of poly(ethylene oxide)
(40 wt.%) into Li-montmorillonite (60 wt.%) has shown to enhance the stability of the ionic
conductivity at lower temperature (see Fig. 51) when compared to more conventional PEO/LiBF4
mixture.
This improvement is explained by the fact that PEO is not able to crystallize when intercalated,
hence eliminating the presence of crystallites, non-conductive in nature. The conductivity of PEO/
Li-montmorillonite nanocomposite is 1.610ÿ6 S/cm at 308C and exhibits a weak temperature
dependence with an activation energy of 2.8 kcal/mol. The higher ionic conductivity at ambient

Fig. 51. Arrhenius plots of ionic conductivity for LiBF4/PEO and PEO Li‡ montmorillonite intercalated nanocomposite
(reproduced from [96] with permission).
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 59

temperature compared to conventional LiBF4/PEO electrolytes combined with a single ionic


conductor character makes those nanocomposites new promising electrolyte materials.

4.4.2. Thermal expansion coefficient


Due to the high aspect ratio of the exfoliated silicate layers, the thermal expansion coefficient of
poly(imide)-based nanocomposites prepared by Yang et al. [102] with hexadecylammonium cation
exchange montmorillonite can be strongly reduced, going from 3.610ÿ5 Kÿ1 to values as low as
1.5510ÿ5 Kÿ1 when 10 wt.% of nanofiller is dispersed. A noticeable decrease of 45%
(1.9610ÿ5 Kÿ1) was already observed with only 1 wt.% of nanoclay.

4.4.3. Other properties


Finally, nanocomposites have been used in highly technical domains such as the improvement
of ablative properties in aeronautics [111], the potentiality to use polyaniline-based nanocomposite
as electrorheological sensitive additive [112] or the combination of dispersed layered silicates in a
liquid crystal medium for the production of stable electro-optical devices exhibiting a bistable and
reversible electro-optical effect between a light scattering opaque state and a transparent state [113].

5. Conclusions

The large array of improved thermo-mechanical properties attained at very filler content
(5 wt.% or less) together with the ease of production through simple processes such as melt
intercalation, directly applicable by extrusion or injection molding make layered silicate-based
nanocomposites a very promising new class of materials. They are already commercially available
and applied in car and food packaging industries. Undoubtedly, the unique combination of their key
properties and potentially low production costs paves the way to much broader range of applications.
Furthermore, the quite low filler level required to display sizeable properties enhancement makes
them competitive with other materials. Their incineration produces ceramic chars in low yield and
the very limited filler content makes them compatible with recycling process. Nevertheless, a lot of
research remains to be carried out in order to fully understand factors such as exfoliation versus
intercalation driving forces yielding the different nanocomposite structures. Moreover, a much better
understanding of the actual structure/properties relationships has to be fulfilled in some important
area such as fire retardancy or physico-mechanical properties. Finally, it is worth pointing out that
new types of layered nanoparticles have recently been reported and their ability to form
nanocomposites with enhanced properties has been proposed. For example, superconductive
nanofillers [114] and magnetic particles [115] have shown to be capable of exfoliation under
controlled conditions and one can foresee at short-term the development of nanocomposite materials
articulated onto such multifunctional nanofillers.

Acknowledgements

Financial support from `ReÂgion Wallonne' and European Community (FEDER and FSE) in the
frame of `PoÃle d'Excellence Materia Nova: Objectif 1' is greatly acknowledged.

References

[1] J.E. Mark, Ceramic reinforced polymers and polymer-modified ceramics, Polym. Eng. Sci. 36 (1996) 2905±2920.
[2] E. Reynaud, C. Gauthier, J. Perez, Nanophases in polymers, Rev. Metall./Cah. Inf. Tech. 96 (1999) 169±176.
60 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

[3] T. von Werne, T.E. Patten, Preparation of structurally well defined polymer±nanoparticle hybrids with controlled/
living radical polymerization, J. Am. Chem. Soc. 121 (1999) 7409±7410.
[4] N. Herron, D.L. Thorn, Nanoparticles. Uses and relationships to molecular clusters, Adv. Mater. 10 (1998)
1173±1184.
[5] P. Calvert, Potential applications of nanotubes, in: T.W. Ebbesen (Ed.), Carbon Nanotubes, CRC Press, Boca Raton,
FL, 1997, pp. 277±292.
[6] V. Favier, G.R. Canova, S.C. Shrivastava, J.Y. Cavaille, Mechanical percolation in cellulose whiskers
nanocomposites, Polym. Eng. Sci. 37 (1997) 1732±1739.
[7] L. Chazeau, J.Y. Cavaille, G. Canova, R. Dendievel, B. Boutherin, Viscoelastic properties of plasticized PVC
reinforced with cellulose whiskers, J. Appl. Polym. Sci. 71 (1999) 1797±1808.
[8] H. Shioyama, Polymerization of isoprene and styrene in the interlayer spacing of graphite, Carbon 35 (1997)
1664±1665.
[9] L. Hernan, J. Morales, J. Santos, Synthesis and characterization of poly(ethylene oxide) nanocomposites of misfit
layer chalcogenides, J. Solid State Chem. 141 (1998) 327±329.
[10] D.J. Harris, T.J. Bonagamba, K. Schmidt-Rohr, Conformation of poly(ethylene oxide) intercalated in clay and MoS2
studied by two-dimensional double-quantum NMR, Macromolecules 32 (1999) 6718±6724.
[11] Y. Matsuo, K. Tahara, Y. Sugie, Synthesis of poly(ethylene oxide)-intercalated graphite oxide, Carbon 34 (1996)
672±674.
[12] Y. Matsuo, K. Tahara, Y. Sugie, Structure and thermal properties of poly(ethylene oxide)-intercalated graphite
oxide, Carbon 35 (1997) 113±120.
[13] Y. Ding, D.J. Jones, P. Maireles-Torres, Two-dimensional nanocomposites: alternating inorganic±organic polymer
layers in zirconium phosphate, Chem. Mater. 7 (1995) 562±571.
[14] O.C. Wilson Jr., T. Olorunyolemi, A. Jaworski, L. Borum, D. Young, A. Siriwat, E. Dickens, C. Oriakhi, M. Lerner,
Surface and interfacial properties of polymer-intercalated layered double hydroxide nanocomposites, Appl. Clay
Sci. 15 (1999) 265±279.
[15] C.O. Oriakhi, I.V. Farr, M.M. Lerner, Thermal characterization of poly(styrene sulfonate)/layered double hydroxide
nanocomposites, Clays and Clay Minerals 45 (1997) 194±202.
[16] B.K.G. Theng, The Chemistry of Clay-Organic Reactions, Wiley, New York, 1974.
[17] M. Ogawa, K. Kuroda, Preparation of inorganic±organic nanocomposites through intercalation of organoammonium
ions into layered silicates, Bull. Chem. Soc. Jpn. 70 (1997) 2593±2618.
[18] Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, Y. Fukushima, T. Karauchi, O. Kamigaito, Mechanical properties of
nylon-6±clay hybrid, J. Mater. Res. 6 (1993) 1185±1189.
[19] E.P. Giannelis, R. Krishnamoorti, E. Manias, Polymer±silica nanocomposites: model systems for confined polymers
and polymer brushes, Adv. Polym. Sci. 118 (1999) 108±147.
[20] G. Lagaly, Interaction of alkylamines with different types of layered compounds, Solid State Ionics 22 (1986) 43±51.
[21] R.A. Vaia, R.K. Teukolsky, E.P. Giannelis, Interlayer structure and molecular environment of alkylammonium
layered silicates, Chem. Mater. 6 (1994) 1017±1022.
[22] E. Hackett, E. Manias, E.P. Giannelis, Molecular dynamics simulations of organically modified layered silicates,
J. Chem. Phys. 108 (1998) 7410±7415.
[23] C. Oriakhi, Nano sandwiches, Chem. Br. 34 (1998) 59±62.
[24] M. Lerner, C. Oriakhi, in: A. Goldstein (Ed.), Handbook of Nanophase Materials, Marcel Dekker, New York, 1997,
p. 199.
[25] G. Lagaly, Introduction: from clay mineral±polymer interactions to clay mineral±polymer nanocomposites, Appl.
Clay Sci. 15 (1999) 1±9.
[26] D.J. Greenland, Adsorption of polyvinylalcohols by montmorillonite, J. Colloid Sci 18 (1963) 647±664.
[27] N. Ogata, S. Kawakage, T. Ogihara, Poly(vinyl alcohol)±clay and poly(ethylene oxide)±clay blend prepared using
water as solvent, J. Appl. Polym. Sci. 66 (1997) 573±581.
[28] R.L. Parfitt, D.J. Greenland, Adsorption of poly(ethylene glycols) on montmorillonites, Clay Mineral 8 (1970)
305±323.
[29] X. Zhao, K. Urano, S. Ogasawara, Adsorption of polyethylene glycol from aqueous solutions on montmorillonite
clays, Colloid Polym. Sci 267 (1989) 899±906.
[30] E. Ruiz-Hitzky, P. Aranda, B. Casal, J.C. GalvaÂn, Nanocomposite materials with controlled ion mobility, Adv.
Mater. 7 (1995).
[31] J. Billingham, C. Breen, J. Yarwood, Adsorption of polyamine, polyacrylic acid and polyethylene glycol on
montmorillonite: an in situ study using ATR-FTIR, Vibr. Spectrosc. 14 (1997) 19±34.
[32] R. Levy, C.W. Francis, Interlayer adsorption of polyvinylpyrrolidone on montmorillonite, J. Colloid Interface Sci.
50 (1975) 442±450.
[33] G. Lagaly, Smectic clays as ionic macromolecules, in: A.D. Wilson, H.J. Prosser (Eds.), Development in Ionic
Polymers, Elsevier, London, 1986, pp. 77±140.
[34] J. Wu, M.M. Lerner, Structural, thermal, and electrical characterization of layered nanocomposites derived from
sodium-montmorillonite and polyethers, Chem. Mater. 5 (1993) 835±838.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 61

[35] H.G. Jeon, H.-T. Jung, S.W. Lee, S.D. Hudson, Morphology of polymer silicate nanocomposites. High density
polyethylene and a nitrile, Polym. Bull. 41 (1998) 107±113.
[36] N. Ogata, G. Jimenez, H. Kawai, T. Ogihara, Structure and thermal/mechanical properties of poly(L-lactide)±clay
blend, J. Polym. Sci.: Part B: Polym. Phys. 35 (1997) 389±396.
[37] G. Jimenez, N. Ogata, H. Kawai, T. Ogihara, Structure and thermal/mechanical properties of poly(e-caprolactone)±
clay blend, J. Appl. Polym. Sci 64 (1997) 2211±2220.
[38] K. Yano, A. Usuki, A. Okada, T. Kurauchi, O. Kamigaito, Synthesis and properties of polyimide±clay hybrid,
J. Polym. Sci.: Part A: Polym. Chem. 31 (1993) 2493±2498.
[39] K. Yano, A. Usuki, A. Okada, Synthesis and properties of polyimide±clay hybrid films, J. Polym. Sci. A: Polym.
Chem. 35 (1997) 2289±2294.
[40] T. Lan, P.D. Kaviratna, T.J. Pinnavaia, On the nature of polyimide±clay hybrid composites, Chem. Mater. 6 (1994)
573±575.
[41] H.-L. Tyan, Y.-C. Liu, K.-H. Wei, Enhancement of imidization of poly(amic acid) through forming poly(amic acid)/
organoclay nanocomposites, Polymer 40 (1999) 4877±4886.
[42] C.O. Oriakhi, X. Zhang, M.M. Lerner, Synthesis and luminescence properties of a poly(p-phenylenevinylene)/
montmorillonite layered nanocomposite, Appl. Clay Sci. 15 (1999) 109±118.
[43] D.C. Lee, L.W. Jang, Preparation and characterization of PMMA±clay hybrid composite by emulsion
polymerization, J. Appl. Polym. Sci. 61 (1996) 1117±1122.
[44] D.C. Lee, L.W. Jang, Characterization of epoxy±clay hybrid composite prepared by emulsion polymerization,
J. Appl. Polym. Sci. 68 (1998) 1997±2005.
[45] M.W. Noh, D.C. Lee, Synthesis and characterization of PS±clay nanocomposite by emulsion polymerization,
Polym. Bull. 42 (1999) 619±626.
[46] M.P. Eastman, E. Bain, T.L. Porter, K. Manygoats, R. Whitehorse, R.A. Parnell, M.E. Hagerman, The formation of
poly(methyl-methacrylate) on transition metal-exchanged hectorite, Appl. Clay Sci. 15 (1999) 173±185.
[47] Y. Fukushima, A. Okada, M. Kawasumi, T. Kurauchi, O. Kamigaito, Swelling behavior of montmorillonite by
poly-6-amide, Clay Mineral, 23 (1988) 27±34.
[48] A. Usuki, Y. Kojima, M. Kawasumi, A. Okada, Y. Fukushima, T. Kurauchi, O. Kamigaito, Synthesis of nylon-6±
clay hybrid, J. Mater. Res. 8 (1993) 1179±1183.
[49] A. Usuki, M. Kawasumi, Y. Kojima, A. Okada, T. Krauchi, O. Kamigaito, Swelling behavior of montmorillonite
cation exchanged for o-amino acid by e-caprolactam, J. Mater. Res. 8 (1993) 1174±1178.
[50] Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, T. Kurauchi, O. Kamigaito, Synthesis of nylon-6-clay hybrid by
montmorillonite intercalated with e-caprolactam, J. Polym. Sci. Part A: Polym. Chem. 31 (1993) 983±986.
[51] Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, T. Kurauchi, O. Kamigaito, One-pot synthesis of nylon-6±clay
hybrid, J. Polym. Sci Part A: Polym. Chem. 31 (1993) 1755±1758.
[52] P. Reichert, J. Kressler, R. Thomann, R. MuÈlhaupt, G. StoÈppelmann, Nanocomposites based on a synthetic layer
silicate and polyamide-12, Acta Polym. 49 (1998) 116±123.
[53] P.B. Messersmith, E.P. Giannelis, Synthesis and barrier properties of poly(e-caprolactone)-layered silicate
nanocomposites, J. Polym. Sci.: Part A Polym. Chem. 33 (1995) 1047±1057.
[54] D. Mecerreyes, R. JeÂroÃme, P. Dubois, Novel macromolecular architectures based on aliphatic polyesters: relevance
of the coordination-insertion ring opening polymerization, Adv. Polym. Sci. 147 (1999) 1±59.
[55] A. LoÈfgren, A.-C. Albertsson, P. Dubois, R. JeÂroÃme, Recent advances in ring-opening polymerization of lactones
and related compounds, J. Macromol. Sci.-Rev. Macromol. Chem. Phys. C35 (1995) 379±418.
[56] B.A. Rozenberg, Pure Appl. Chem. 53 (1981) 1715±1722.
[57] T.K. Chen, Y.I. Tien, K.H. Wei, Synthesis and characterization of novel segmented polyurethane/clay
nanocomposite via poly(e-caprolactone)/clay, J. Polym. Sci.: Part A Polym. Chem. 37 (1999) 2225±2233.
[58] P.B. Messersmith, E.P. Giannelis, Polymer-layered silicate nanocomposites: in situ intercalative polymerization of
e-caprolactone in layered silicates, Chem. Mater. 5 (1993) 1064±1066.
[59] A. Akelah, A. Moet, Polymer±clay nanocomposites: free-radical grafting of polystyrene on to organophilic
montmorillonite interlayers, J. Mater. Sci. 31 (1996) 3589±3596.
[60] J.G. Doh, I. Cho, Synthesis and properties of polystyrene-organoammonium montmorillonite hybrid, Polym. Bull.
41 (1998) 511±517.
[61] M.W. Weimer, H. Chen, E.P. Giannelis, D.Y. Sogah, Direct synthesis of dispersed nanocomposites by in situ living
free radical polymerization using a silicate-anchored initiator, J. Am. Chem. Soc. 121 (1999) 1615±1616.
[62] J. Tudor, L. Willington, D. O'Hare, B. Royan, Intercalation of catalytically active metal complexes in phyllosilicates
and their application as propene polymerization catalysts, Chem. Commun. (1996) 2031±2032.
[63] M. Alexandre, P. Dubois, R. JeÂroÃme, M. Garcia-Marti, T. Sun, J.M. Garces, D.M. Millar, A. Kuperman, Polyolefin
nanocomposites, WO Patent WO9947598A1 (1999).
[64] M. Alexandre, P. Dubois, J.M. Garces, T. Sun, R. JeÂroÃme, in preparation.
[65] P. Dubois, M. Alexandre, F. Hindryckx, R. JeÂroÃme, Homogeneous polyolefin-based composites, J. Macromol. Sci.:
Rev. Macromol. Chem. Phys. C38 (1998) 511±565.
[66] M. Alexandre, E. Martin, P. Dubois, M. Garcia-Marti, R. JeÂroÃme, submitted for publication.
62 M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63

[67] J. Heinemann, P. Reichert, R. Thomann, R. MuÈlhaupt, Polyolefin nanocomposites formed by melt compounding and
transition metal catalyzed ethene homo- and copolymerization in the presence of layered silicates, Macromol. Rapid
Commun. 20 (1999) 423±430.
[68] Y.C. Ke, C.F. Long, Z.N. Qi, Crystallization, properties, and crystal and nanoscale morphology of PET±clay
nanocomposites, J. Appl. Polym. Sci. 71 (1999) 1139±1146.
[69] P.B. Messersmith, E.P. Giannelis, Synthesis and characterization of layered silicate-epoxy nanocomposites, Chem.
Mater. 6 (1994) 1719±1725.
[70] X. Kornmann, L.A. Berglund, J. Sterte, nanocomposite based on montmorillonite and unsaturated polyester, Polym.
Eng. Sci. 38 (1998) 1351±1358.
[71] T. Lan, T.J. Pinnavaia, Clay-reinforced epoxy nanocomposites, Chem. Mater. 6 (1994) 2216±2219.
[72] T. Lan, P.D. Kaviratna, T.J. Pinnavaia, Mechanism of clay tactoid exfoliation in epoxy±clay nanocomposites, Chem.
Mater. 7 (1995) 2144±2150.
[73] C. Zilg, R. MuÈlhaupt, J. Finter, Morphology and toughness/stiffness balance of nanocomposites based upon
anhydride-cured epoxy resins and layered silicates, Macromol. Chem. Phys. 200 (1999) 661±670.
[74] Z. Wang, T.J. Pinnavaia, Hybrid organic±inorganic nanocomposites: exfoliation of magadiite nanolayers in an
elastomeric epoxy polymer, Chem. Mater. 10 (1998) 1820±1826.
[75] Z. Wang, T.J. Pinnavaia, Nanolayer reinforcement of elastomeric polyurethane, Chem. Mater. 10 (1998) 3769±3771.
[76] R.A. Vaia, E.P. Giannelis, Lattice of polymer melt intercalation in organically-modified layered silicates,
Macromolecules 30 (1997) 7990±7999.
[77] A.C. Balazs, C. Singh, E. Zhulina, Modeling the interactions between polymers and clay surfaces through self-
consistent field theory, Macromolecules 31 (1998) 8370±8381.
[78] A.C. Balazs, C. Singh, E. Zhulina, Y. Lyatskaya, Modeling the phase behavior of polymer/clay nanocomposites,
Acc. Chem. Res. 8 (1999) 651±657.
[79] Y. Lyatskaya, A.C. Balazs, Modeling the phase behavior of polymer±clay composites, Macromolecules 31 (1998)
6676±6680.
[80] V.V. Ginzburg, A.C. Balazs, Calculating phase diagram of polymer±platelet mixtures using density functional
theory: implication for polymer/clay composites, Macromolecules 32 (1999) 5681±5688.
[81] R.A. Vaia, E.P. Giannelis, Polymer melt intercalation in organically-modified layered silicates: model predictions
and experiment, Macromolecules 30 (1997) 8000±8009.
[82] R.A. Vaia, H. Ishii, E.P. Giannelis, Synthesis and properties of two-dimensional nanostructures by direct
intercalation of polymer melts in layered silicates, Chem. Mater. 5 (1993) 1694±1696.
[83] R.A. Vaia, K.D. Jandt, E.J. Kramer, E.P. Giannelis, Kinetics of polymer melt intercalation, Macromolecules 28
(1995) 8080±8085.
[84] R.A. Vaia, K.D. Jandt, E.J. Kramer, E.P. Giannelis, Microstructural evolution of melt intercalated polymer-
organically modified layered silicates nanocomposites, Chem. Mater. 8 (1996) 2628±2635.
[85] L.M. Liu, Z.N. Qi, X.G. Zhu, Studies on nylon-6 clay nanocomposites by melt-intercalation process, J. Appl.
Polym. Sci. 71 (1999) 1133±1138.
[86] M. Kato, A. Usuki, A. Okada, Synthesis of polypropylene oligomer-clay intercalation compounds, J. Appl. Polym.
Sci 66 (1997) 1781±1785.
[87] M. Kawasumi, N. Hasegawa, M. Kato, A. Usuki, A. Okada, Preparation and mechanical properties of
polypropylene±clay hybrids, Macromolecules 30 (1997) 6333±6338.
[88] N. Hasegawa, M. Kawasumi, M. Kato, A. Usuki, A. Okada, Preparation and mechanical properties of
polypropylene±clay hybrids using a maleic anhydride-modified polypropylene oligomer, J. Appl. Polym. Sci 67
(1998) 87±92.
[89] D. Wolf, A. Fuchs, U. Wagenknecht, B. Kretzschmar, D. Jehnichen, L. HaÈussler, Nanocomposites of polyolefin clay
hybrids, in: Proceedings of the Eurofiller'99, Lyon-Villeurbanne, 6±9 September 1999.
[90] M. Alexandre, G. Beyer, C. Henrist, R. Cloots, A. Rulmont, P. Dubois, in preparation.
[91] M. Laus, O. Francesangeli, F. Sandrolini, New hybrid nanocomposites based on an organophilic clay and
poly(styrene-b-butadiene) copolymers, J. Mater. Res. 12 (1997) 3134±3139.
[92] S.D. Burnside, E.P. Giannelis, Synthesis and properties of new poly(dimethylsiloxane) nanocomposites, Chem.
Mater. 7 (1995) 1597±1600.
[93] S.J. Wang, C.F. Long, X.Y. Wang, Q. Li, Z.N. Qi, Synthesis and properties of silicone rubber organomontmorillonite
hybrid nanocomposites, J. Appl. Polym. Sci. 69 (1998) 1557±1561.
[94] A. Okada, K. Fukumori, A. Usuki, Y. Kojima, T. Kurauchi, O. Kamigaito, Rubber±clay hybrid Ð synthesis and
properties, Polym. Prep. 32 (1991) 540±541.
[95] A. Okada, A. Usuki, The chemistry of polymer±clay hybrids, Mater. Sci. Eng. C3 (1995) 109±115.
[96] R.A. Vaia, S. Vasudevan, W. Krawiec, L.G. Scanlon, E.P. Giannelis, New polymer electrolyte nanocomposites: melt
intercalation of poly(ethylene oxide) in mica-type silicates, Adv. Mater. 7 (1995) 154±156.
[97] R.A. Vaia, B.B. Sauer, O.K. Tse, E.P. Giannelis, Relaxations of confined chains in polymer nanocomposites: glass
transition properties of poly(ethylene oxide) intercalated in montmorillonite, J. Polym. Sci.: Part B Polym. Phys. 35
(1997) 59±67.
M. Alexandre, P. Dubois / Materials Science and Engineering 28 (2000) 1±63 63

[98] W. Chen, Q. Xu, R.Z. Yuan, Modification of poly(ethylene oxide) with polymethylmethacrylate in polymer-layered
silicate nanocomposites, J. Mater. Sci. Lett. 18 (1999) 711±713.
[99] H.R. Fischer, L.H. Gielgens, T.P.M. Koster, Nanocomposites from polymers and layered materials, Acta Polym. 50
(1999) 122±126.
[100] K.A. Carrado, L.Q. Xu, In-situ synthesis of polymer±clay nanocomposites from silicate gels, Chem. Mater. 10
(1998) 1440±1445.
[101] L. Mullins, N.R. Tobin, J. Appl. Polym. Sci. 9 (1965) 2993±3005.
[102] Y. Yang, Z.-K. Zhu, J. Yin, X.-Y. Wang, Z.-E. Qi, Preparation and properties of hybrids of organo-soluble polyimide
and montmorillonite with various chemical surface modifications methods, Polymer 40 (1999) 4407±4414.
[103] A. Blumstein, Polymerization of adsorbed monolayers: II. Thermal degradation of the inserted polymers, J. Polym.
Sci. A3 (1965) 2665±2673.
[104] J. Lee, T. Takekoshi, E. Giannelis, Fire retardant polyetherimide nanocomposites, Mater. Res. Soc. Symp. Proc. 457
(1997) 513±518.
[105] J.W. Gilman, Flammability and thermal stability studies of polymer layered-silicate (clay) nanocomposites, Appl.
Clay Sci. 15 (1999) 31±49.
[106] F. Dietsche, R. MuÈlhaupt, Thermal properties and flammability of acrylic nanocomposites based upon organophilic
layered silicates, Polym. Bull. 43 (1999) 395±402.
[107] J.W. Gilman, T. Kashiwagi, S. Lomakin, E.P. Giannelis, E. Manias, J.D. Lichtenhan, P. Jones, Nanocomposites:
radiative gasification and vinyl polymer flammability, in: Proceedings of the 6th European Meeting on Fire
Retardancy of Polymeric Materials (FRPM'97), University of Lille, France, 24±26 September 1997, pp. 203±221.
[108] J.W. Gilman, T. Kashiwagi, J.E.T. Brown, S. Lomakin, Flammability studies of polymer layered silicate
nanocomposites, SAMPE J. 43 (1998) 1053±1066.
[109] F. Dabrowski, M. Le Bras, S. Bourbigot, J. Gilman, T. Kashiwagi, PA-6 montmorillonite nanocomposite in
intumescent fire retarded EVA, in: Proceedings of the Eurofillers '99, Lyon-Villeurbanne, France, 6±9 September
1999.
[110] C. Scherer, PA Film grade with improved barrier properties for flexible food packaging applications, in: Proceedings
of the New plastics'99, London, 2±4 February 1999.
[111] R.A. Vaia, G. Price, P.N. Ruth, H.T. Nguyen, J. Lichtenhan, Polymer/layered silicate nanocomposites as high
performance ablative materials, Appl. Clay Sci. 15 (1999) 67±92.
[112] J.W. Kim, S.G. Kim, H.J. Choi, M.S. Jhon, Synthesis and electrorheological properties of polyaniline±Na‡-
montmorillonite suspensions, Macromol. Rapid Commun. 20 (1999) 450±452.
[113] M. Kawasumi, N. Hasegawa, A. Usuki, A. Okada, Liquid crystal/clay mineral composites, Appl. Clay Sci. 15
(1999) 93±108.
[114] J.-H. Choy, S.-J. Kwon, S.-J. Hwang, Y.-I. Kim, W. Lee, Intercalation route to nano hybrids: inorganic/organic Ð
high Tc cuprate hybrid materials, J. Mater. Chem. 9 (1999) 129±135.
[115] V. Laget, C. Hornick, P. Rabu, M. Drillon, Hybrid organic±inorganic layered compound prepared by anion exchange
reaction: correlation between structure and magnetic properties, J. Mater. Chem. 9 (1999) 169±174.

You might also like