Está en la página 1de 17

NANOCOMPOSITES,

POLYMERCLAY
Introduction
Polymer matrix nanocomposites are a fairly new class of engineered materials
which offer for a broad range of properties, an interesting and even radical alternative to more conventional lled polymers, yet at much lower ller loadings.
They can be dened as polymer matrix systems in which the dispersed inorganic
reinforcing phase has at least one of its dimensions in the nanometer range, which
is quite close to the scale of elementary phenomena at the molecular level. The
resulting unique combination of large interfacial area and small interparticle distance strongly inuences nanocomposite behavior.
Current status of research and industrial development of polymer nanocomposites clearly outlines the prominent position of clay nanocomposites and the
present review is mainly devoted to the latter materials.
From a general point of view, ller aspect ratio is a pertinent parameter to
distinguish between various types of nanocomposites. Table 1 summarizes typical dimensions of particles under concern. Spherical silica particles are an example of isotropic nanoparticles which either provide increased composite stiffness while retaining matrix transparency, or exhibit novel optical properties by
forming colloidal crystals. Although it is obviously critical to the optical behavior, it is generally observed that optimum mechanical properties are not achieved
in conjunction with the best state of dispersion (1). When only two dimensions
are in the nanometer range, ber-like structures, such as whiskers or carbon
nanotubes, with aspect ratios ranging between 50 and 1000 are dealt with. For
instance, cellulose whiskers extracted from tunicate shells have been shown to
dramatically improve composite stiffness in the case of poly(vinyl chloride) or
336
Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

337

Table 1. Typical Nanoller Dimensions


Material

Shape

Typical dimensions

Silica particles
Cellulose whiskers
Carbon nanotubes
Layered silicates

Spheres
Rigid rods
Flexible tubes (multiwall)
Flexible discs

Diameter: 30150 nm
Diameter: 15 nm; length: 1 m
Diameter: 30 nm; length: 1050 m
Diameter: 50500 nm; thickness: 1 nm

poly(styrene-co-butyl acrylate) matrix in the rubbery state. In the latter material


a 3 order of magnitude modulus improvement is achieved at only 6 mass% whisker
content. Percolation of a rigid whisker network is evoked to account for such a
property increase (2). Reaching a conductivity percolation threshold at very low
loadings, owing to the nanoscale dispersion of carbon nanotubes with large aspect
ratio, has also been of prime importance for designing conductive polymer blends
with electrostatic paintability, while limiting resin embrittlement (3). Preliminary
experiments with carbon nanotubepolymer composites also underline the potential of this nanoller for entering the eld of composite materials for structural
applications as well (4). Opportunities for functional and/or structural applications of carbon nanotubes will mainly depend on the capacity to reach large-scale
production at moderate cost and to monitor composite elaboration while retaining
nanotube integrity.
Among the variety of composites that display unique structure and behavior
at the nanometer level, as compared to classical micrometer scale particulate lled
materials, the use of layered silicates as a reinforcing phase is by far the most successful way of designing polymer nanocomposites with a broad range of markedly
modied properties. A report from Toyota Central Research Laboratories of the
development of a polyamide-6 (PA6)-based clay nanocomposite with remarkable
thermomechanical behavior, at low clay loadings relative to conventional ller
additives (below 5 mass% instead of 2030 mass%), triggered extensive research
efforts worldwide (5). In fact the benets were shown not only for strength and
stiffness, but also for thermal stability and barrier properties (6). Accordingly, the
following presentation focuses on these materials, starting with a brief description
of the layered silicates commonly used, followed by nanocomposite structural characterization and elaboration routes. Various properties of interest are reviewed
together with the currently emerging structureproperty relationship schemes.

Polymer-Layered Silicate Nanocomposites


Layered silicates, the more widely used in polymer nanocomposites, belong to the
same structural group, the 2:1 phyllosilicates, and more specically to the smectite
group. They comprise natural clay minerals such as montmorillonite, hectorite,
and saponite and also synthetic layered minerals, uorohectorite, laponite, or
magadiite. An idealized structure for montmorillonite is presented in Figure 1.
Elementary clay platelets consist of a 1-nm-thick layer made of two tetrahedral
sheets of silica fused to an edge-shared octahedral sheet of alumina or magnesia.
Isomorphic cation substitution results in an excess of negative charges within

338

NANOCOMPOSITES, POLYMERCLAY

Fig. 1. Structure of montmorillonite.

Al or Mg;

Vol. 3

OH;  O;

exchangeable cations.

the layer. Stacking of the layers leads to a regular van der Waals gap called
gallery or interlayer. Cations located in the galleries (eg Ca2+ , Na+ ) counterbalance the excess layer charges. They are usually hydrated. This negative surface
charge is quantied as the cation-exchange capacity (CEC), usually in the range
from 80 to 150 meq/100 g for smectites. The interlayer space can be penetrated
by organic cations or polar organic liquids as well. Exchange reactions with organic cations (aminoacids, alkylammonium ions, etc) enable to render the silicate
surface organophilic. A key parameter of the stacking is the basal spacing or
d-spacing, which ranges from 0.96 nm (ie, the layer thickness) in the fully collapsed state to about 2 nm, depending on the nature of the interlayer cation and
the amount of adsorbed water (7). Individual lamellae have a high aspect ratio,
with a diameter typically in the range 50500 nm. Primary crystals (also called
tactoids) consist of 810 lamellae, with usually disordered stacking. Their aggregation leads to a turbostratic structure. This organization is reected in the x-ray
diffraction (xrd) pattern where diffuse hk bands rather than sharp hkl reections
are observed. The basal reection (001) is of more interest since it is used to derive the d-spacing. It will be dependent on the amount and nature of intercalated
molecules in the galleries.
Characterization of Nanocomposite Microstructure. Mixing clay
with a polymer does not necessarily lead to a nanocomposite. Elaboration strategies are aimed at monitoring dispersion of the inorganic compound at the nanometer level, that is down to the elementary clay platelet. Figure 2 provides a
schematic illustration of the various microstructures readily achievable, namely
a conventional lled polymer with clay particles in the micrometer range, an

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

(a)

339

(b)

(c)

Fig. 2. Schematic representation of clay platelet dispersion as (a) primary crystals, (b)
intercalated, or (c) exfoliated structures.

intercalated nanocomposite in which extended polymer chains are intercalated


in the gap between silicate layers while the stacking order is retained (note that
in this case the host gallery height is much smaller than the radius of gyration of
the polymer chain), and a delaminated (or exfoliated) nanocomposite where clay
layers are individually dispersed in the host polymer matrix.
The two basic tools used to elucidate nanocomposite morphology are x-ray
diffraction (xrd) and transmission electron microscopy (tem). They provide complementary information on clay dispersion in the host matrix.
The basal plane reection (00l) diffraction peak yields a direct evaluation of
the d-spacing between the clay lamellae, as long as layer registry is preserved to
some extent. It allows therefore to follow the structural evolution from the pristine
clay stacking to any intercalated state. With changes in d values in the range 14
nm, such data are accessible at low diffraction angle, ie, for 1 < 2 < 9 . Follow up
of the increase in peak width at half maximum also reects the increase in the
degree of disorder in layer stacking during the intercalation process [see eg, (8)].
Figure 3 illustrates the change in d-spacing of a stearylammonium-modied
montmorillonite (C18 Mt) blended with a maleic anhydride-modied polypropylene

340

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

II
6.0 nm

6.3 nm

(d)
5.0 nm

Intensity, a.u.

(c)
3.4 nm

Intensity, a.u.

(d)

5.7 nm

(c)
3.5 nm

(b)
(b)
2.1 nm
2.1 nm

(a)
0.5

3.5
5
6.5
2 -Cu, deg

9.5

(a)
0.5

3.5
5
6.5
2 -Cu, deg

9.5

Fig. 3. Evolution of the d-spacing showing montmorillonite intercalation in PP-MA


oligomer (I), and no subsequent change upon further blending with PP (II). Reproduced
from Ref. 9.

oligomer (PP-MA). As the amount of the latter component is increased, the interlayer distance increases drastically (spectra Iad), indicating the efciency of the
intercalation process. Spectrum IIa reveals the lack of intercalation of C18 Mt by
pure PP. Spectra IIbd do not show any change in the state of intercalation assessed in Spectra I when the PP-MA intercalated organoclay is further dipersed
in a PP matrix. It is concluded that PP penetration in the galleries is not favored
in this particular case (9).
Whenever the xrd pattern becomes silent, meaning that the interlayer spacing is no longer accessible by conventional wide-angle x-ray diffraction, the need
still remains to have access to a mesoscale description of the spatial distribution
of the clay platelets. Small-angle x-ray scattering may provide some answers regarding average interparticle distance, and also about platelet orientation with
respect to process geometry (10).
Transmission electron microscopy has the great advantage to give a direct view of the microstructure. Typical micrographs reveal alternating dark and
bright bands refering to the silicate layers and polymer matrix respectively. The
illustrations of Figure 4 show exfoliated and intercalated-exfoliated structures in
the case of 1 and 5 mass% dimethylditallow ammonium-modied montmorillonite

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

500 nm

341

200 nm

Fig. 4. Transmission electron microscopy of water-aided melt-dispersed organoclay in


PA6. Courtesy of Dr M. van Es, DSM Reseach, Geleen.

100 nm

Fig. 5. Transmission electron microscopy of a melt-intercalated organoclay tactoid in a


PP matrix. Courtesy of Dr M. Bacia, UST Lille.

dispersed by melt extrusion in a PA6 matrix. Additionally tem enables imaging


of any mesoscale long-range ordering of the clay. It also reveals platelet exibility promoted by the high aspect ratio and nanometer thickness as illustrated in
Figure 5 for an intercalated organoclay tactoid in a PP matrix. An accurate description of polymer clay nanocomposites is therefore available from a combination
of electron microscopy and xrd techniques.

Elaboration of PolymerOrganoclay Nanocomposites


Various methods have been developed in order to prepare polymer-layered silicate
nanocomposites. These include in situ polymerization, polymer intercalation in

342

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

solution, emulsion polymerization in the presence of layered silicates, and melt


intercalation. A few examples are given below to illustrate the different strategies.
Review articles are available for more detailed information [see eg (1113), and
references therein].
In situ Polymerization. Early works in the 1960s (14) demonstrated the
feasability of intercalation polymerization of methyl methacrylate (MMA) after
insertionadsorption of the polar monomer between the lamellae of a sodium
montmorillonite. Polymerization was initiated by free-radical catalysts or with
irradiation. Renewed interest for the method in the last decade came from the
work on PA6clay hybrids at Toyota Central Research Laboratories (5), following a scheme adopted earlier by Unitika Co. (15). Sodium montmorillonite was
rst cation exchanged with -amino acids [H3 N+ (CH2 )n 1 COOH]. X-ray diffraction data show that the basal spacing is highly sensitive to the length of the
alkyl chain. For n > 11, the -amino acid chain lies slanted to the layer and provides optimum swelling behavior by -caprolactam. 12-Amino-l-lauric acid (n =
12) modied montmorillonite (12-Mt) was used for performing the in situ ringopening polymerization of -caprolactam. The carbonyl end groups of 12-Mt initiate polymerization and an increasing amount of polyamide-6 is bonded to the
clay as the 12-Mt clay content is increased as clearly established by nmr studies
(16). The resulting hybrid is either exfoliated at low 12-Mt content or increasingly intercalated beyond 10 wt% clay. In an alternative approach, the same team
successfully developed a so-called one-pot synthesis of PA6clay hybrid without
preliminary cation exchange of the montmorillonite. Ring-opening polymerization
of -caprolactam with 6-amino-l-caproic acid as an accelerator was performed in
a water dispersion of montmorillonite in the presence of an acid. Phosphoric acid
seems to be the best candidate to achieve true exfoliation in this particular process (17). 12-Amino-l-lauric acid was also successfully used both as a uorinated
silicate modier and as a monomer in order to prepare a polyamide-12-based
nanocomposite with exfoliatedintercalated structure (18). Some attempts to produce polystyrene (PS)-based nanocomposites through the in situ polymerization
route have been reported (19). Intercalated structures are obtained. Intercalative polymerization of -caprolactone has also been achieved in the presence of
-protonated amino acid-exchanged montmorillonite. Upon heating, the organic
acid groups initiate ring-opening polymerization of the monomer and the resulting
polymer is ionically bound to the silicate platelets with a good level of delamination as revealed by xrd (20). In situ polymerization has also been extended to
polyolens, polyesters, or polycarbonate in recent years.
The production of thermoset-based nanocomposites by the same method has
been investigated by many authors (2123). In the case of epoxyclay nanocomposites, the organophilic clay is rst swollen in the mixture of epoxy prepolymer
and curing agent. The gel state and nal structure are strongly dependent on the
nature of the onium ion, cross-linking amine, and curing conditions. In particular, larger chain length of the alkylammonium and ion-exchange with protonated
primary amines should be preferred. Polyurethane networks are equally good candidates for clay nanolayer reinforcement (22). The approach consists in solvating
polyol precursors in montmorillonite exchanged with long-chain onium ions and
further adding the diisocyanate curing agent. Intercalated tactoids are obtained
in the nal cured nanocomposite.

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

343

Intercalation in Solution. In the case of water-soluble polymers, it is possible to prepare solvent-cast nanocomposites by using water as a cosolvent. Pristine montmorillonite can be easily dispersed in water, owing to its hydrophilic character, and blending for instance with polymers such as poly(ethylene oxide) (PEO)
or poly(vinyl alcohol) (PVOH) is thus achievable (see eg (2426)). Intercalated
exfoliated structures are observed for PVOH-based nanocomposites. On the contrary, in the presence of PEO, intercalated silicate layers are organized as large
clay tactoids; this indicates that reaggregation of the initial silicate water suspension occurred during the lm casting process. A similar elaboration strategy
has also been developed, starting from an organoclay. In the examples of poly(caprolactone) and poly(l-lactide), chloroform was used as a cosolvent. In both situations no evidence of intercalation could be found but the clay tactoids displayed a
remarkable geometric arrangement, with their surfaces lying parallel to the cast
lm surface (27). These few illustrations are indicative of the high sensitivity of
the nal materials structure to the nature of the host matrix and to the interplay
of polymerpolymer and polymerclay interactions.
The above technique can be advantageously adapted to a situation where
the polymer of interest is not soluble in any solvent, as is the case for polyimides.
The polyimide precursor, ie, a poly(amic acid) solution in dimethylacetamide, is
prepared and blended to a dispersion of an organomodied montmorillonite in the
same solvent. Dimethylacetamide is then gradually removed, and upon heating
the poly(amic acid) lm at 300 C under nitrogen atmosphere the polyimideclay
hybrid is obtained. When an ammonium salt of dodecylamine is used in the ionexchange process, true exfoliation is achieved and furthermore the clay platelets
align parallel to the lm surface (28).
Emulsion Polymerization. Considering the high hydrophilic character
of sodium montmorillonite, it was anticipated that polymerization in an aqueous medium might provide an alternative route for polymerclay nanocomposite
preparation. The rst report dealing with such an approach concerned the emulsion polymerization of MMA dispersed in a water phase in the presence of Na+ montmorillonite. The structural analysis conrms intercalation of the PMMA in
the clay galleries. A PSclay nanocomposite has been elaborated according to the
same procedure, with the same resulting intercalated morphology. Intercalation
by the emulsion technique was also achieved for an epoxy system, again without
requiring any ion-exchange treatment (29).
Melt Intercalation. Although emulsion polymerization or in situ polymerization may be considered as viable routes for industrial-scale production, as exemplied for PA6 with the latter process, melt compounding remains the most
obvious route for cost-effective development of nanoreinforced polymers. The rst
industrial applications refer to the PA6clay hybrid in situ polymerization process
patented by Toyota (30), but growing interest in achieving clay nanodispersion by
melt compounding is observed worldwide (31). Patent literature is clearly indicative of the current trend with many references to polyamides and other polar
polymers such as polyesters or polyimides [see eg (32)]. Emphasis is put on the
nature of the surfactant and processing conditions. Nonpolar polymers such as
polyolens are also highly attractive candidates with research efforts primarily
driven by the automotive market. Polypropyleneclay hybrids have been prepared
by melt mixing an organoclay, maleic anhydride-modied PP oligomers (PP-MA),

344

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

and PP. The polar PP-MA intercalates in the clay galleries and the quality of clay
dispersion in the hybrids are clearly affected by the degree of miscibility of the
polar oligomers in PP (9,33).
Direct melt intercalation of PEO in pristine montmorillonite by static annealing of cold-pressed powder mixture slightly above the melting temperature of PEO
has been reported (34). The resulting d-spacing of the order of 1.8 nm indicates
that the PEO chains are constrained in a 0.8-nm interlayer space. Differential
scanning calorimetry studies reveal that the polymer is deprived of any thermodynamic transition. Neither the heat capacity jump characteristic of the glass
transition nor the melting endotherm is observed. Chain dynamics appear quite
peculiar in these systems. Thermally stimulated current results point at the essentially noncooperative nature of the motions of the conned chains. Intercalation
kinetics have also been followed by xrd monitoring of the basal spacing reection
in model systems consisting of monodisperse PS and organically modied uorohectorite. The most striking result is that the mobility of the polymer chains
in this conned environment seems larger than that in the bulk melt (34,35).
Spin-echo nmr experiments using deuterated PS suggest the coexistence of multiple environments, from solid-like to liquid-like, for intercalated chains. Molecular
dynamics simulations relate this complex dynamic behavior to strong density inhomogeneity normal to the surface [(36), and Refs. 45 to 49 therein].
An extensive investigation of nanocomposite preparation with the aid of a
swelling agent that is known to intercalate the clay provides some additional
interesting experimental ground on the phase behavior. The results underline
the inuence of the polymer/swelling agent miscibility (as assessed by a solubility parameter approach) on nanocomposite formation. The example of an epoxy
monomer as the swelling agent shows that either complete miscibility or strong
immiscibility are preferable (37).

Structure Development in PolymerClay Nanocomposites


Whatever the elaboration route is, understanding phase behavior of the resulting
nanocomposites is of prime importance to achieve reliable material development.
A lattice-based mean eld model has been developed in order to address the
problem of nanocomposite formation (38). By deriving the evolution of the free energy of the system with interlayer spacing, the model provides some basic predictions regarding the equilibrium states (ie, exfoliated, intercalated, or immiscible),
in relation to the enthalpic and entropic factors of the interacting constituents,
namely the silicate, the tethered surfactant chains, and the polymer. Polymer
connement results in entropy loss but the latter may be compensated in part by
the entropy gain induced through the increase in conformational freedom of the
tethered surfactant chain upon layer separation. As a consequence, melt intercalation is predicted to depend primarily on energetic (enthalpic) factors. Polar
polymers or polymers containing groups interacting with the silicate surface will
favor polymerclay hybrid formation (35,38).
A theoretical investigation of the phase behavior of model analogues
of polymerclay nanocomposites has been conducted (39). Combining a selfconsistent eld model with density functional theory, the investigation underlines

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

345

some trends regarding phase morphology and stability for these systems. In particular, the calculations point at the key inuence of the surfactant chain length.
Polymer-like values lead to enhanced miscibility of the clay platelets and polymer
matrix (exfoliated structure), even at moderate level of interactions between the
grafted chains and polymer melt (39). The equilibrium behavior of a mixture of
functionalized and nonfunctionalized chains (taken chemically identical) in the
presence of two innite planar surfaces has alternatively been considered using
scaling theory. Functionalized chains have a telechelic architecture; ie, the active
groups are located at each end of the chain. Qualitative phase diagrams are derived with prime consideration of the respective inuences of interaction energy
between the surface and the end group, and volume fraction of functionalized
chains. Provided the interaction energy is high enough, exfoliation occurs more
easily as is increased, whereas low values result in an immiscible mixture
(40).

Nanocomposites Behavior
Thermal Stability. Thermal stability improvement was already recognized in the pioneering work by Blumstein on PMMA intercalated in
montmorillonite. Intercalated PMMA degraded at a temperature 50 C higher
than that of conventional unlled PMMA. In recent years, thermogravimetric
analysis of various polymerclay systems have conrmed this observation even
for low nanoller loadings [see eg (13,41), and references therein]. A particularly
striking example is that of cross-linked poly(dimethylsiloxane) incorporating 10
mass% exfoliated organomontmorillonite (42) for which thermal stability under
nitrogen ow is enhanced by 140 C. Overall, restricted thermal motion in silicate
interlayers and hindering of the diffusion of decomposition products are certainly
key factors, but polymer structure and nature of the degradation mechanisms
and degradation conditions are equally important to account for the disparities
observed in literature.
Owing to what has been said previously on the role of the organic modier of the layered silicate in elaboration and phase behavior of the polymerclay
nanocomposites, understanding of the key structural factors that inuence its
thermal degradation is of prime importance. Major concern is of course toward
processing conditions and re-retardant behavior as well. Recent work (43) on
alkyl quaternary ammonium montmorillonite by a combination of thermogravimetric analysis, Fourier transform infrared spectroscopy, and mass spectrometry
points at the complex degradation behavior of the organic surfactant. The initial
degradation temperature is insensitive to chain architecture and length, or exchange ratio. Compared to the parent alkyl ammonium, the thermal stability of
a fraction of the surfactant is substantially lowered because of catalytic sites on
the layered silicates. Polymer processing generally implying temperatures higher
than 180 C, chemical degradation of the surfactant is expected to occur. Many
questions remain regarding the role of the decomposition products on the melt
intercalation mechanism and subsequent nanocomposite phase stability. In this
respect the potential of nmr techniques to follow the fate of the organic modier
seems quite promising (44).

346

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

Crystal Organization. Semicrystalline polymer nanocomposites present


a unique interplay between nanoscale morphology of crystal lamellae on one hand
and clay platelet organization on the other. Considering the importance of interfacial interactions and the conned chain environment, one may expect drastic
changes in crystal organization. The case of PA6clay hybrids is rather well documented. The preferred orientation of the silicate layers under melt-ow conditions, together with polymer connement, will affect crystallization behavior. For
instance, in injection-molded bars, close to the surface both the silicate layers
and polymer chain axes (and hence lamella thickness) are parallel to the surface,
whereas the chain axes rotate by 90 inside the bar and remain perpendicular
to the silicate layers (45). Clay also favors nucleation of the phase (16,44), contrary to bulk PA6 which predominantly crystallizes in the more stable form. The
elevated temperature (205 C) crystal morphology of PA6clay hybrids has been examined by performing simultaneous small- and wide-angle x-ray scattering (46).
These results clearly establish that the nanoscale correlations of the silicate layer
organization (4060 nm) affect polymer crystallization, resulting in a less-ordered
crystal phase. Evidence is also provided of the impact of polymer-layer interactions (tethered vs nontethered chains), the more defective lamellae pertaining to
the in situ polymerized (tethered) nanocomposites. Spherulitic structure is unable
to develop as in bulk polymers, which ought to inuence the nonelastic mechanical
response. Similar crystallization behavior is also observed in other semicrystalline
polymers such as poly(-caprolactone) or poly(ethylene terephtalate) nanocomposites (27,47). The observation of high melting temperature phases, though defective, might come from a reduced entropy of fusion Sf due to the conned crystallization environment.
Fire-Retardant Behavior. Controlling polymer ammability remains a
key issue in numerous applications of engineering plastics and commodity
polymers as well. The re-retardant additive approach provides cost-effective
solutions, but generally at the expenses of some physical and mechanical properties. There is also growing pressure for environmentally safe products and processes, including recyclability and use of halogen-free compounds. For these reasons, recognition of improved ammability properties in the case of polymerclay
nanocomposites has triggered the development of extensive research programs on
a large variety of materials (41,48,49).
Cone calorimetry is used to evaluate the ammability under re-like conditions. Relevant parameters such as the rate of heat release (HRR) and its peak
value, heat of combustion (Hc), smoke yield (specic extension area, SEA), and
carbon monoxide yield are obtained. Table 2 shows some typical data for layered silicate nanocomposites based on organically treated montmorillonite, with
polyamide 6, poly(propylene-graft-maleic anhydride), and polystyrene as the host
matrix. Nanocomposites under investigation have either delaminated (PA6) or
intercalateddelaminated structures. In all cases there is a substantial reduction in peak HRR value (5075%), whereas Hc and CO formation show little
variation.
Table 2 also compares the PSclay nanocomposite with a PSclay mix for
which intercalation does not occur. The kinetics of heat release are displayed
in Figure 6. In the case of the mix, particle dispersion is only achieved at the
primary particle micrometer level. The peak HRR value remains identical to

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

347

Table 2. Cone Calorimeter Dataa

Nylon-6
Nylon-6clay 2%
delaminated
Nylon-6clay 5%
delaminated
PS
PSclay mix 3%
immiscible
PSclay 3%
intercalated
PP-g-MA
PP-g-MAclay 5%
intercalated
a After

Residue
yield,
%

Peak
HRR,
kW/m2

Mean
HRR,
kW/m2

Mean
Hc,
MJ/kg

Total
heat
released,
MJ/m2

Mean
SEA,
m2 /kg

Mean
CO
yield,
kg/kg

1.0
3.0

1011
686

603
390

27
27

413
406

197
271

0.01
0.01

5.7

378

304

27

397

296

0.02

0
3.2

1118
1080

703
715

29
29

102
96

1464
1836

0.09
0.09

3.7

567

444

27

89

1727

0.08

0
8.0

2028
922

861
651

38
37

219
179

756
994

0.04
0.05

Ref. 48.
1400

Heat release rate, kW/m2

1200
1000
800
600
400
200
0
0

50

100

150
Time, s

200

250

300

Fig. 6. Kinetics of heat release rate for PS-based compounds. Reproduced from Ref. 48.
PS pure;
PS immiscible (3% silicate); PS intercalated (3% silicate).

that of pure PS while the intercalated PSclay nanocomposite shows a 50%


reduction.
The observed reduced ammability in the nanocomposites may not be attributed to an additional retention of carbon alone since the residue yields are
not markedly increased. Some key insights are provided by radiative gasication experiments which enable to follow pyrolysis either in a nitrogen or in a
nitrogenoxygen atmosphere. The example of PA6 nanocomposites is revealing

348

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

enough of enhanced char formation and reduced mass loss rate in comparison to
pure PA6. The current interpretation of these results is that the nanocomposite
ame-retardant mechanism occurs through the build up of a reinforced char layer,
which acts as an insulator, and a mass transport barrier so as to slow down the
escape of the decomposition products. Developments according to this concept include the use of a PA6clay nanocomposite in intumescent formulations as in the
case of ethylenevinyl acetate copolymers with ammonium polyphosphate (50).
Enhanced ame-retardant performance is related to the formation of a thermally
stable phosphocarbonaceous structure in the char, and the blend even shows a
slight improvement in mechanical behavior.
Processing conditions also strongly inuence the ame-retardant behavior.
For example, in the case of PS-based nanocomposites, extrusion above 180 C under
partially oxidative conditions yields an intercalated nanocomposite but with no
ammability improvement, whereas the melt-extruded system at 170 C under
nitrogen or vacuum exhibits ame-retardant efciency (41). The way thermal
degradation of the organic modier alters the ammability reduction mechanism
has yet to be understood.
Barrier Properties. Early work in the Toyota Research group acknowledged the great potential of polymerclay nanocomposites to reduce moisture
absorption and decrease water and gas permeability, even at low clay loadings
(6). A further advantage for packaging applications lies in the fact that lowering
of permeability is achieved while preserving transparency, owing to the suitable
dispersion of platelets smaller than the wavelength of visible light.
The example of polyimide clay lms illustrates the dramatic decrease of
permeability coefcients. Only 2 mass% montmorillonite loading reduces the permeability by more than 50% of the pure polymer value for water vapor, oxygen,
or helium. Notwithstanding possible changes in diffusion and/or solubility, it has
been postulated that the major role of the clay platelets consists in substantially
increasing the path length of the permeant, that is by creating a highly tortuous
path, due to the high aspect ratio of the clay. A simple theory derived by Nielsen
expresses the relative permeability as follows:
Pc /Po = 1/[1+(L/2W)Vf ]
in which V f is the volume fraction of plates, L is the plate length, and W is the
thickness. Pc and Po stand for the nanocomposite and polymer permeability respectively.
Using equivalent loadings of clay but varying the aspect ratio yields results in
fairly good agreement with the theoretical prediction (28). In the same way, a signicant reduction in water vapor permeability was observed in the case of a poly(caprolactone)organomontmorillonite nanocomposite, showing a vefold reduction at only 4.8 vol% clay whereas it is only halved at best with a 20 vol% conventionally lled silicate composite (20). The tortuosity model shows discrepancies between actual aspect ratios measured by tem and values deduced from curve tting.
Among the reasons for that are the deviations from the ideal dispersion of platelets
parallel to the lm surface, and possible aggregation of individual platelets.
Although tortuosity plays a role in barrier enhancement, some other factors
ought to be taken into account (see BARRIER POLYMERS). Recent work on MXD6

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

349

Table 3. Comparison of PA6 and PA6Clay Hybrids Mechanical


Propertiesa
Property

PA6 clay hybrid

PA6

Tensile strength, MPa


At 23 C
At 120 C
Tensile modulus, GPac
At 23 C
At 120 C
Flexural strength, MPab
At 23 C
At 120 C
Flexural modulus, GPac
At 23 C
At 120 C
Izod impact, J/md
Charpy impact, KJ/m2e
HDT (1.82 MPa), C

97
32
1.9
0.6

69
27
1.1
0.2

143
33

89
12

4.3
1.2
18
6.1
152

2.0
0.3
21
6.2
65

a After

Ref. 5.
convert MPa to psi, multiply by 145.
c To convert GPa to psi, multiply by 145,000.
d To convert J/m to ftlbf/in., divide by 53.38.
e To convert kJ/m2 to ftlbf/in.2 , divide by 2.4.
b To

nanocomposites is indicative of an oxygen permeation reduction in humid environment far beyond what is expected from the increase in path length alone (51).
In the light of the ndings regarding chain packing and dynamics in such conned environments (36), models for the prediction of barrier properties ought to
take into account the changes induced in terms of solubility and diffusion. A conceptual model has been proposed (52). So far no general predictive approach is
available.
Mechanical Behavior. Being able to improve strength and stiffness with
limited alteration of toughness is a goal readily achievable with polymerclay
nanocomposites (see MECHANICAL PROPERTIES; REINFORCEMENT).
Table 3 gathers some key data of the original work by the Toyota group (5),
which show the dramatic inuence of organomontmorillonite on mechanical properties of PA6clay hybrids at low mineral loading (4.7 mass%). The improvement
in strength is claimed to have little or no inuence on impact properties as evaluated from Izod or Charpy tests. Increase in modulus is paralleled by a substantial
rise in heat distortion temperature. A concept of constrained polymer region related to the ion-bonding strength of clay and PA6 is introduced to account for the
observed behavior.
Linear Elastic and Rubber Elastic Behavior. Although stiffening is quite
noticeable in the glassy regime of the amorphous phase, the most spectacular effect is seen in the rubber elastic regime phase, as already evoked in the case of
reinforcement by cellulose whiskers (2). The PA6clay hybrids example presented
in Table 3 is quite representative of the situation encountered with semicrystalline
thermoplastics, but elastomeric networks benet as well of clay layer dispersion
with a two- to threefold increase in modulus for polyurethane or epoxy networks

350

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

(22). In the meantime, improved elongation at break is observed, contrary to what


is seen in classical lled systems, presumably due in part to dangling chain formation in the network (see ELASTICITY, RUBBER-LIKE). The dynamic mechanical
loss peak related to the glass-transition mechanism is equally informative on the
extent of polymerclay interaction. This shows mainly in the reduction of the loss
peak area and/or in the evolution of the peak temperature with clay content and
elaboration conditions (5,53). Predictive modeling of both storage and loss modulus faces a complex challenge in order to account for the mechanical coupling
between the phases, the potential existence of an interphase, and/or a certain
degree of connectivity between the llers. In the latter situation, a percolation
approach should be useful (2,54). Otherwise models derived from HalpinTsai
equations seem quite promising for modulus prediction in relation to clay platelet
arrangement (55,56).
Plasticity and Rupture. The main drawback identied regarding the solidstate drawing behavior is certainly the limited elongation at break encountered
for most thermoplasticclay nanocomposites (9,33) in the vicinity or below the
glass-transition temperature. Nanovoiding and subsequent extensive brillation
of the polymer matrix is clearly evidenced from volume strain measurements
during drawing (53) and from in situ tem observations (57). Such enhancement of
nanoscale plasticity offers an opportunity for optimizing the stiffness/toughness
balance. However critical microvoids may develop from areas where load transfer
is no more achievable because of splitting of clay platelet aggregates. This points at
the most critical issue in nanocomposite development, ie, monitoring of elaboration
and processing conditions.
Key Role of Processing. Scarce work has been devoted to the inuence of
processing on microstructure and properties of polymerclay nanocomposites (58,
59). It is shown that twin-screw extrusion enables achieving a signicant degree
of dispersion of the clay platelets, provided residence time and degree of shearing
are optimized in conjunction with the nature of the organoclay. Thermal stability
of the organic modier is again at the heart of the problem.
In the same way as demonstration was made in the last decade of the importance of processing to design polymer blends, taking the full benets of the
interesting combination of properties displayed by polymer nanocomposites will
mainly rely on key developments in the eld of processing. Automotive and packaging markets are undoubtedly the driving force for it.

BIBLIOGRAPHY
1. J. M. Jethmalani and W. T. Ford, Chem. Mater. 8, 2138 (1996); Z. Pu and co-workers,
Chem. Mater. 9, 2442 (1997).
2. L. Chazeau and co-workers, J. Appl. Polym. Sci. 71, 1797 (1999); V. Favier and
co-workers, Polym. Eng. Sci. 37, 1732 (1997).
3. J. J. Scobbo, in SPE Antec 98 Proceedings, 1998, p. 2468.
4. D. Quian and co-workers, Appl. Phys. Lett. 76, 2868 (2000); L. Jin, C. Bower, and
O. Zhou, Appl. Phys. Lett. 73, 1197 (2000).
5. A. Usuki and co-workers, J. Mater. Res. 8, 1174 (1993); Y. Kojima and co-workers,
J. Mater. Res. 8, 1185 (1993).
6. A. Okada and co-workers, Polym. Prep. 28, 447 (1987).

Vol. 3

NANOCOMPOSITES, POLYMERCLAY

351

7. W. A. Deer, R. A. Howie, and J. Zussman, An Introduction to Rock-Forming Minerals,


2nd ed., Longman Scientic, Essex, U.K., 1992.
8. R. A. Vaia and co-workers, Chem. Mater. 8, 2628 (1996).
9. N. Hasegawa and co-workers, J. Appl. Polym. Sci. 67, 87 (1998).
10. K. Varlot and co-workers, J. Polym. Sci., Part B 39, 1360 (2001).
11. E. P. Giannelis, Adv. Mater. 8, 29 (1996).
12. P. C. Lebaron, Z. Wang, and T. J. Pinnavaia, Appl. Clay Sci. 15, 11 (1999).
13. M. Alexandre and P. Dubois, Mater. Sci. Eng. 28, 1 (2000).
14. A. Blumstein, J. Polym. Sci., Part A 3, 2653 (1965).
15. Jpn. Pat. JP-A- 51-109998 (1976), S. Fujiwara and T. Sakamoto (to Unitika).
16. R. D. Davis, W. L. Jarrett, and L. J. Mathias, Polym. Mater. Sci. Eng. 82, 272 (2000).
17. Y. Kojima and co-workers, J. Polym. Sci., Part A 31, 1755 (1993).
18. P. Reichert and co-workers, Acta Polym. 49, 116 (1998).
19. A. Akelah and A. Moet, J. Mater. Sci. 31, 3589 (1996); J. G. Doh and I. Cho, Polym.
Bull. 41, 511 (1998).
20. P. B. Messersmith and E. P. Giannelis, J. Polym. Sci., Part A 33, 1047 (1995).
21. P. B. Messersmith and E. P. Giannelis, Chem. Mater. 6, 1719 (1994).
22. T. Lan, P. D. Kaviratna, and T. J. Pinnavaia, J. Phys. Chem. Solids 57, 1005 (1996);
Z. Wang, T. Lan, and T. J. Pinnavaia, Chem. Mater. 8, 2200 (1996); Z. Wang and T. J.
Pinnavaia, Chem. Mater. 10, 1820 (1998); Z. Wang and T. J. Pinnavaia, Chem. Mater.
10, 3769 (1998).

23. C. Zilg, R. Mulhaupt,


and J. Finter, Macromol. Chem. Phys. 200, 661 (1999).
24. P. Aranda and E. Ruiz-Hitzky, Chem. Mater. 4, 1395 (1992); J. Wu and M. M. Lerner,
Chem. Mater. 5, 835 (1993).
25. R. A. Vaia and co-workers, J. Polym. Sci., Part B 35, 59 (1997).
26. N. Ogata, S. Kawakage, and T. Ogihara, J. Appl. Polym. Sci. 66, 573 (1997); K. E.
Strawhecker and E. Manias, Chem. Mater. 12, 2943 (2000).
27. N. Ogata and co-workers, J. Polym. Sci., Part B 35, 389 (1997); G. Jimenez and
co-workers, J. Appl. Polym. Sci. 64, 2211 (1997).
28. K. Yano and co-workers, J. Polym. Sci., Part A 31, 2493 (1993); K. Yano, A. Usuki, and
A. Okada, J. Polym. Sci., Part A 35, 2289 (1997).
29. D. C. Lee and L. W. Jang, J. Appl. Polym. Sci. 61, 1117 (1996); M. W. Noh and D. C.
Lee, Polym. Bull. 42, 619 (1999); D. C. Lee and L. W. Jang, J. Appl. Polym. Sci. 68,
1997 (1998).
30. U.S. Pat. 4,739,007 (Apr. 19, 1988), A. Okada and co-workers (to Toyota Co.).
31. L. Liu, Z. Qi, and X. Zhu, J. Appl. Polym. Sci. 71, 1133 (1999).
32. WO Pat. 93-04117 (Mar. 4, 1993), M. R. Maxeld and co-workers (to Allied-Signal Inc.);
WO Pat. 99-29767 (June 17, 1999), R. A. Korbee and co-workers (to DSM NV); WO
Pat. 99-32403 (July 1, 1999), R. B. Barbee and co-workers (to Eastman Chemical Co.);
WO Pat. 99-50340 (Oct. 7, 1999), H. Okamoto and co-workers (to Toyota Co.); WO Pat.
2000034180 (June 15, 2000), J. N. Gilmer and co-workers (to Eastman Chemical Co.).
33. M. Kawasumi and co-workers, Macromolecules 30, 6333 (1997).
34. R. A. Vaia and co-workers, J. Polym. Sci., Part B 35, 59 (1997); R. A. Vaia and
co-workers, Macromolecules 28, 8080 (1995).
35. E. P. Giannelis, R. Krishnamoorti, and E. Manias, Adv. Polym. Sci. 138, 107 (1999).
36. R. A. Vaia and E. P. Giannelis, MRS Bull. 26, 394 (2001).
37. H. Ishida, S. Campbell, and J. Blackwell, Chem. Mater. 12, 1260 (2000).
38. R. A. Vaia and E. P. Giannelis, Macromolecules 30, 7990 (1997); 8000 (1997).
39. V. V. Ginzburg, C. Singh, and A. C. Balazs, Macromolecules 33, 1089 (2000).
40. D. V. Kuznetsov and A. C. Balazs, J. Chem. Phys. 112, 4365 (2000).
41. J. W. Gilman and co-workers, Chem. Mater. 12, 1866 (2000).
42. S. D. Burnside and E. P. Giannelis, Chem. Mater. 7, 1597 (1995).

352
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.

NANOCOMPOSITES, POLYMERCLAY

Vol. 3

W. Xie and co-workers, Chem. Mater. 13, 2979 (2001).


D. L. Vanderhart, A. Asano, and J. W. Gilman, Macromolecules 34, 3819 (2001).
Y. Kojima and co-workers, J. Polym. Sci., Part B 33, 1039 (1995).
D. M. Lincoln and co-workers, Polymer 42, 1621 (2001).
Y. Ke, C. Long, and Z. Qi, J. Appl. Polym. Sci. 71, 1139 (1999).
J. W. Gilman, T. Kashiwagi, and J. Lichtenham, SAMPE J. 33, 40 (1997); J. W. Gilman,
Appl. Clay Sci. 15, 31 (1999).
D. Porter, E. Metcalfe, and M.-J. K. Thomas, Fire Mater. 24, 45 (2000).
S. Bourbigot and co-workers, Fire Mater. 24, 45 (2000).
T. Lan, Y. Liang, G. W. Beall, and K. Kamena, in Additives 1999 Conference Proceedings, San Francisco, Calif., 1999.
G. W. Beall, in T. J. Pinnavaia and G. W. Beall, eds., Polymer-Clay Nanocomposites,
(Wiley Series in Polymer Science), John Wiley & Sons, Inc., New York, 2001, p. 267.
J. M. Gloaguen and J. M. Lefebvre, Polymer 42, 5841 (2001).
E. Reynaud, C. Gauthier, and J. Perez, Rev. Metall.-CIT 169 (Feb. 1999).
M. van Es, F. Xiqiao, J. van Turnhout, and E. van der Giessen, in Additives 2000
Conference Proceedings, San Francisco, Calif., 2000.
D. A. Brune and J. Bicerano, Polymer 43, 369 (2002).
G. M. Kim and co-workers, Polymer 42, 1095 (2001).
H. R. Dennis and co-workers, Plast. Eng. 57, 1, 56, (2001).
J. W. Cho and D. R. Paul, Polymer 42, 1083 (2001); T. D. Fornes and co-workers, Polymer
42, 9929 (2001).

JEAN-MARC LEFEBVRE
Universite des Sciences et Technologies de Lille

NOVOLAKS.

See PHENOLIC RESINS.

NUCLEIC ACIDS.
NYLON.

See POLYNUCLEOTIDES.

See POLYAMIDES.

OLEFIN-SULFUR DIOXIDE POLYMERS.


ORGANOMETALLIC POLYMERS.
ORIENTED FILMS.

See POLYSULFONES.

See METAL CONTAINING POLYMERS.

See FILMS, ORIENTATION.

También podría gustarte