You are on page 1of 32

J. Ind. Eng. Chem., Vol. 12, No.

6, (2006) 811-842
REVIEW

Rheology of Polymer/Layered Silicate Nanocomposites


Suprakas Sinha Ray†
Macromolecular Nanotechnology Research Group, Nanoscience Research Centre, CSIR Materials Science and
Manufacturing, PO Box 395, Pretoria 0001, Republic of South Africa
Received November 14, 2006

Abstract: Over the last few years, polymer/layered silicate (PLS) nanocomposites have been an area of intense
academic and industrial research. No matter the measure-journal articles, patents, and industry research and
development (R&D) funding-efforts in PLS-nanocomposite have been exponentially growing worldwide for
the last ten years. Aside from the understanding in fabrication, characterization, and improved mechanical and
other properties, the measurement of rheological properties of PLS-nanocomposites under molten state is
crucial to gain fundamental understanding of the processability of these materials. In the case of PLS-
nanocomposites, the measurements of rheological properties are also helpful to find out the strength of
polymer-layered silicate interactions and the structure-property relationship in nanocomposites. This is
because rheological behaviors are strongly influenced by their nanoscale structure and interfacial
characteristics. In this article, recent advances in melt-state rheological properties of PLS-nanocomposites are
highlighted. To begin this a very brief description of structure and dynamics of PLS-nanocomposites are also
described.

Keywords: polymer/layered silicate nanocomposite, structure, thermodynamics, melt rheology

Introduction [1], metal oxide nanoparticles [2,3], inorganic nanotubes


1) [4], expandable graphite [5-7], layered titanate [8], cellu-
The word composite is generally used to define any ma- lose nanowhiskers [9], polyhedral oligomeric silses-
terial made of more than one component. Depending on quioxanes (POSS) [10], carbon nanotubes [11-17], etc.
the matrix nature, composite materials can be broadly Polymer nanocomposites based on layered silicates have
divided into three categories such as polymeric, metallic, attracted great interest in today’s materials research, as it
and ceramic. In most of the commercially available com- is possible to achieve impressive enhancements of prop-
posite materials the structural unit is on the micrometer erties when compared with virgin polymer [1,18-20].
‐6
(10 m) length scale and is used mostly to improve the These improvements can include high moduli [1,18], in-
mechanical properties of matrix material. Now one can creased strength and heat resistance [1,18], decreased gas
change the dimension of the structural building block to permeability [21-23] and flammability [24-27], and in-
‐9
the nanometer (10 m) scale, the resulting composite ma- creased degradability of biodegradable polymers [28-30].
terial is called a Nanocomposite. The nanometer is one On the other hand, there has been considerable interest in
billionth of meter and ∼10,000 times finer than a human theory and simulations addressing the preparation and
hair. Figure 1 compares the morphology of classical- and properties of these materials [31], and they are also con-
nano-filler filled composite. sidered to be unique model systems to study the structure
Here we are interested on polymer nanocomposites that and dynamics of polymers in confined environments
means nanocomposite made with polymer and nano- [32].
scale filler. Over the last few years, various types of The main reason for these improved properties in
nano-fillers are in development such as layered silicates nanocomposites is the interfacial interaction between the
matrix and layered silicate, as opposed to conventional

To whom all correspondence should be addressed. composites [1]. Layered silicates have layer thickness on
(e-mail: RSuprakas@csir.co.za)
812 Suprakas Sinha Ray

Figure 1. Schematic comparison between microcomposite (classical filler filled) and “Nanocomposite. Reproduced with permission
from Elsevier Science Ltd., UK.

ratio, and a repeat distance of few nanometers, and (3)


exfoliated nanocomposites, in which the individual
silicate layers are separated in the polymer matrix by
average distances that totally depend on the layered
silicate loading.
This article highlights recent advances in melt-state
rheological properties of PLS nanocomposites. To begin
this a brief description on structure and properties of
layered silicates, nanocomposites preparation techniques
and thermodynamics are described. Finally, attention is
drawn to how rheological behavior of PLS nanocom-
posites can be understood directly from their structure.

Structure and Properties of Layered Silicates


Figure 2. Schematic illustration of three different types of ther-
modynamically achievable polymer/layered silicate compo- The commonly used layered silicates for the preparation
sites. of PLS nanocomposites belong to the same general
family of 2 : 1 layered- or phyllosilicates [33]. Their
the order of 1 nm, and very high aspect ratios (e.g. 10∼ crystal structure consists of layers made up of two
1,000). A few weight percent of layered silicate particles tetrahedrally coordinated silicon atoms fused to an
that are properly dispersed throughout the matrix can edge-shared octahedral sheet of either aluminum or
thus create a much larger surface area for polymer filler magnesium hydroxide. The layer thickness is around 1
interactions than do conventional composites. On the nm, and the lateral dimensions of these layers may vary
basis of the strength of the polymer/layered silicate from 30 nm to several microns or larger, depending on
interfacial interaction, three structurally different types of the particular layered silicate. Stacking of the layers leads
composites are achievable (see Figure 2): (1) phase- to a regular van der Waal’s gap between the layers called
separated composite, when polymer matrix has no the interlayer or gallery. Isomorphic substitution within
+3 +2 +2
interaction with layered silicate, (2) intercalated nanoco- the layers (for example, Al replaced by Mg or Fe , or
+2 +1
posites, where insertion of polymer chains into the Mg replaced by Li ) generates negative charges that
silicate structure occurs in a crystallographically regular are counterbalanced by alkali and alkaline earth cations
fashion, regardless of the polymer-to-layered silicate situated inside the galleries. This type of layered silicate
Rheology of Polymer/Layered Silicate Nanocomposites 813

Two particular characteristics of layered silicates are


generally considered for PLS nanocomposites fabri-
cation. The first is the ability of the silicate particles to
disperse into individual layers inside the polymer matrix.
The second characteristic is the ability to fine-tune their
surface chemistry through ion exchange reactions with
organic and inorganic cations. These two characteristics
are, of course, interrelated, since the degree of dispersion
of layered silicate in a particular polymer matrix depends
on the interlayer cation. Pristine layered silicates usually
+ +
contain hydrated Na or K ions [1]. Obviously, in this
pristine state, layered silicates are only miscible with hy-
drophilic polymers such as poly(ethylene oxide) (PEO)
[36] or poly(vinyl alcohol) (PVA) [37]. To render lay-
ered silicates miscible with various polymer matrices,
Figure 3. Structure of 2:1 phyllosilicates. Reproduced from ref. one must convert the normally hydrophilic silicate sur-
[1] by permission of Elsevier Science Ltd., UK. face to an organophilic one, making the intercalation of
many polymers possible. Generally, this can be done by
is characterized by a moderate surface charge known as ion-exchange reactions with cationic surfactants includ-
the cation exchange capacity (CEC), and generally ing primary, secondary, tertiary, and quaternary alky-
expressed as mequiv/100 gm. This charge is not locally lammonium or alkylphosphonium cations. Alkylam-
constant, but varies from layer to layer, and must be monium or alkylphosphonium cations in the organo-
considered as an average value over the whole crystal. silicates lower the surface energy of the inorganic host
Layered silicates have two types of structure: tetrahed- and improve the wetting characteristics of the polymer
rally substituted and octahedrally substituted. In the case matrix, and result in a larger interlayer spacing.
of tetrahedrally substituted layered silicates, the negative Additionally, the alkylammonium or alkylphosphonium
charge is located on the surface of silicate layers and, cations can provide functional groups that can react with
hence, the polymer matrices can interact more readily the polymer matrix, or in some cases initiate the poly-
with these than with octahedrally substituted material. merization of monomers to improve the strength of the
Generally, layered silicate minerals are divided into interface between the inorganic and the polymer matrix
three major groups: (a) the kaolinite group, (b) the semc- [1].
tite group and (c) the illite or the mica-clay group.
Among the three major groups, smectite types or more
precisely montmorillonite, saponite and hectorite are the Preparative Methods
most commonly used layered silicates in the filed of pol-
ymer nanocomposite technology. Again, among mont- Intercalation of polymer chains inside layered silicate
morillonite, saponite, hectorite, montmorillonite (MMT) galleries has proven to be a successful approach to fab-
is the most commonly used layered silicates for the fab- ricate PLS-nanocomposites. Thermoplastic polymers
rication of PLS nanocomposites, because it is highly based PLS-nanocomposites may be prepared by one of
abundant and inexpensive. MMT is the name given to the the three following routes: (i) in-situ intercalative poly-
layered silicate found near montmorillonite in France, merization of monomer, (ii) intercalation of polymer or
where MMT was first identified by Knight in 1896. The pre-polymer from solution, and (iii) melt-intercalation.
chemical formula of MMT is Mx(Al4‐xMgx)Si8O20(OH)4. The second method may be processed using emulsion,
The CEC (generally 90∼110 meq/100 g) and particle whereas when the good solvent is not available, re-
length of MMT (100∼150 nm) depends on the source. searchers have to choose either in-situ polymerization
The specific surface area of MMT is equal to 750∼800 method (i) or direct melt blending (iii).
m2/g and the modulus of each MMT sheet is around 170 The direct melt-intercalation process involves annealing
GPa [34]. The interlayer thickness of hydrated MMT is a mixture of the polymer and layered silicate (organically
equal to 1.45 nm and the average density ρ= 2.385 modified or not) above the softening point of the poly-
g/mL. Drying MMT at 150 °C reduces the gallery height mer, statically or under shear. While annealing, the poly-
to 0.28 nm which corresponds to a water monolayer and mer chains diffuse from the bulk polymer melt into the
hence the interlayer spacing decreases to 0.94 nm and the galleries between the silicate layers. A range of nano-
average density increases to 3.138 g/mL [35]. The unit composites with structures from intercalated to exfoliate
structure of MMT is presented in Figure 3. can be obtained, depending on the degree of penetration
814 Suprakas Sinha Ray

of the polymer chains into the silicate galleries. This sent author observes that (a) an optimal interlayer struc-
process was found to be completely reversible in the pol- ture on the OMLS, with respect to the number per unit
ystyrene (PS) based nanocomposite because PS has al- area and size of surfactant chains, is most favorable for
most no interaction with the organically modified layered nanocomposite formation, and (b) polymer intercalation
silicate (OMLS) surface, but when polymer chains have depends on the existence of polar interactions between
some favorable interaction with silicate surface, this the OMLS and the polymer matrix.
process found to be irreversible. However, for a large va- In order to understand the thermodynamic issue asso-
riety of nanocomposites, this method requires often the ciated with nanocomposite formation during melt-inter-
use of extrusion process in order to get a homogeneous calation, Vaia and coworkers [38,39] applied a mean‐
dispersion of intercalated silicate layers in the polymer field statistical lattice model, reporting that calculations
matrix. based on the mean field theory agree well with ex-
perimental results. Although there is entropy losses asso-
ciated with the confinement of a polymer melt during
Thermodynamics and Molecular Modelling of nanocomposite formation, this process is allowed be-
PLS Nanocomposite Formation cause there is an entropy gain associated with the layer
separation, resulting in a net entropy change near to zero.
The solution intercalation method is based on a solvent Thus, from the theoretical model, the outcome of nano-
system in which the polymer or pre-polymer is soluble composite formation via polymer melt intercalation de-
and the silicate layers are swellable. The layered silicate pends primarily on energetic factors, which may be de-
is first swollen in a solvent, such as water, chloroform, or termined from the surface energies of the polymer and
toluene. When the polymer and layered silicate solutions OMLS.
are mixed, the polymer chains intercalate and displace Based on the Vaia and coworkers [38,39] study and the
the solvent within the interlayer of the silicate. Upon sol- construction of product maps, general guidelines may be
vent removal, the intercalated structure remains, resulting established for selecting potentially compatible poly-
in polymer/layered silicate nanocomposite. mer/OMLS systems. Initially, the interlayer structure of
For the overall process, in which polymer is exchanged the OMLS should be optimized in order to maximize the
with the previously intercalated solvent in the gallery, a configurational freedom of the functionalizing chains
negative variation in the Gibbs free energy is required. upon layer separation, and to maximize potential inter-
The driving force for the polymer intercalation into lay- action sites at the interlayer surface. For these systems,
ered silicate from solution is the entropy gained by de- the optimal structures exhibit a slightly more extensive
sorption of solvent molecules, which compensates for the chain arrangement than with a pseudo-bilayer. Polymers
decreased entropy of the confined, intercalated chains. containing polar groups capable of associative inter-
Using this method, intercalation only occurs for certain actions, such as Lewis-acid/base interactions or hydrogen
polymer/solvent pairs. This method is good for the inter- bonding, lead to intercalation. The greater the polar-
calation of polymers with little or no polarity into layered izability or hydrophilicity of the polymer, the shorter the
structures, and facilitates production of thin films with functional groups in the OMLS should be in order to
polymer-oriented clay intercalated layers. minimize unfavorable interactions between the aliphatic
In recent years, the melt intercalation technique has be- chains and the polymer.
come the standard for the preparation of PLS-nano- One of the great advantages of the Vaia mean-field stat-
composites. During polymer intercalation from solution, istical lattice model is the ability to determine analyti-
a relatively large number of solvent molecules have to be cally the effect of various aspects of the nanocomposite
desorbed from the host to accommodate the incoming formation. According to this model, the variation of the
polymer chains. The desorbed solvent molecules gain free energy of mixing and subsequent dependence on en-
one translational degree of freedom, and the resulting en- thalpic and entropic factor, suggest the formation of three
tropic gain compensates for the decrease in conforma- possible structures-phase separated, intercalated and
tional entropy of the confined polymer chains. Therefore, exfoliated. Although, Vaia model able to address some of
there are many advantages to direct melt intercalation the fundamental and qualitative thermodynamic issues
over solution intercalation. For example, direct melt in- associated with the nanocomposite formation, however,
tercalation is highly specific for the polymer, leading to some of the assumptions such as the separation of con-
new hybrids that were previously inaccessible. figurational terms and intermolecular interaction and the
So far, experimental results indicate that the outcome of further separation of the entropic behavior of the con-
polymer chains intercalation into two dimensional sili- stituents, somewhat limit the usefulness of the model.
cate galleries depends critically on layered silicate sur- Not only that, this model is based on nanocomposites
face functionalization and constituent interactions. Pre- where polymer chains are completely tethered with the
Rheology of Polymer/Layered Silicate Nanocomposites 815

silicate surface, which is not the case for most of the pol- trices always lead to the formation of phase separated
ymer nanocomposites. structure. The SCF calculations and phase diagrams lead
To overcome the limitation of Vaia model, Balazs and to the same conclusion.
coworkers [40] proposed a model based on a self-con- To overcome this problem, Balazs and coworkers then
sistent field (SCF) calculation, such as the Fleer and proposed the scheme of using a mixture of functionalized
Scheutjens theory [41]. In the Fleer and Scheutjens theo- and non-functionalized polymers for the melts [42].
ry, the phase behavior of polymer systems is modeled by While the stickers at the chain ends are highly attracted
combining Markov chain statistics with a mean field to the surface, the remainder of polymer does not react
approximation. These calculations involve a planar lat- with the substrate. Thus, as the polymer chain penetrate
tice where lattice spacing represents the length of a stat- the sheets, the majority of the chain is not likely to glue
istical segment within a polymer chain. Details regarding the surfaces together.
this theory can be found in ref. [41]. Using this method Balazs and coworkers [43,44] also proposed a simple
Balazs and coworkers tried to calculate the interactions model that describes the nematic ordering in the poly-
between two surfactant-coated surfaces and a polymer mer-layered silicate systems. Because of the very high
melt. They considered two planar surfaces that lie paral- degree of anisotropy (a typical layered silicate platelet is
lel to each other in the xy-plane and investigated the ef- ∼1 nm in thickness and 100∼200 nm in diameter), lay-
fect of increasing the separation between the surfaces in ered silicate particles experience strong orientational or-
the z-direction. The two surfaces are effectively im- dering at low volume fractions and can form liquid crys-
mersed within a polymer melt. As the separation between talline phases such as nematic, smectic, or columnar, in
the surfaces is increased, polymer from the surrounding addition to traditional liquid and solid phases (see Figure
bath penetrates the gap between these walls. Each surface 4). Starting from the Onsager free energy functional for
is covered with monodispersed end-grafted chains, i.e. the nematic ordering of rigid rods, they developed a
surfactants. Now if χsurf represents the Flory-Huggins in- modified expression to combine the disk orientational
teraction parameter between the polymers and the under- and positional entropy, steric excluded-volume effects,
lying solid substrate and χssurf represents the Flory‐ translational entropy of the polymer, and finally the
Huggins interaction between the surfactant and surface. Flory-Huggins enthalpic interaction. The resulting iso-
Therefore, χsurf -χssurf = 0. It should be noted here; in thropic-nematic phase diagram correctly represents many
their calculation Balazs and coworkers did not consider important features, such as the role of shape anisotropy
electrostatic interaction. in depressing the ordering transition and the increase in
Their calculations show that on increasing the attraction the size of the immiscibility region with increases in the
between the polymers and the modified surfaces are polymer chain length. Unlike most of the phenomeno-
qualitatively similar to observation as made by Vaia and logical theories of polymer-liquid crystal systems [45-
coworkers [38,39]. However, Balazs and coworkers 48], in the Onsager-type model the features of the phase
found that the actual phase behavior and morphology of diagram are directly derived from the geometric charac-
the mixer can be affected by the kinetics of the polymers teristics of the anisotropic component.
penetrate the gap between the plates. At the beginning, Balazs and coworkers first modified and expanded
the polymer chains have to penetrate the space between Onsager theory by including nematic, smectic and
the silicate layers from an outer edge and then diffuse to- columnar crystalline phases. To do calculation, they also
ward the centre of the gallery. Now if we consider the adopted Somoza-Tarazona formalism of density functio-
case where χsurf < 0 and thus, the polymer and surface nal theory (DFT) and then incorporated expressions that
experience an attractive interaction. In this case as the describe the entropy of mixing between the different
polymer diffuses through the energetically favorable gal- components and the enthalphic interaction between the
lery, it maximizes contact with the two confining layers. platelets [34,49]. The resulting free energy function can
As a results, the polymer ‘glues’ the two surfaces togeth- be minimized with respect to both the orientational and
er as it moves through the interlayer. This ‘fused’ con- positional single-particle distribution function of the
dition could represent a kinetically trapped state and con- platelets, and thus, potentially, all phases and coexistence
sequently, increasing the attraction between the polymer regions can be determined. The resulting phase diagram
and layered silicate sheets would only lead to interca- was shown to exhibit a strong dependence on the shape
lated, rather than exfoliated structures. On the other hand, anisotropy of the layered silicate particles, the polymer
in the case where χsurf > 0, the polymer can separate the chain length, and the strength of the interparticle
sheets, as the chain tries to retain its coil-like con- interaction. In particular, an increase in the shape
formation and gain entropy. However, recent literatures anisotropy for oblate ellipsoidal filler particles leads to
revealed that the melt mixing of organically modified the broadening of the nematic phase at the expense of the
layered silicates and almost no attractive polymer ma- isotropic region. The increase in the polymer chain
816 Suprakas Sinha Ray

Figure 4. Possible mesophases of oblate uniaxial particles dispersed in a polymer: (a) isotropic, (b) nematic, (c) smectic A, (d) col-
umnar, and (e) crystal. The nematic director n in ordered phases is aligned along Z axis; the disks lie in the XY plane. Dashed lines
show smectic layers (c) and columns (d). Reprinted from ref. [44] with permission from American Chemical Society, USA.

length leads to the formation of the crystalline and/or plastics. The dominance of interfacial regions resulting
liquid crystalline mesophases and promotes segregation from the nanoscopic phase dimensions implies that the
between polymerrich regions and filler particles. Finally, behavior of PLS nanocomposites cannot be understood
an increase in the strength of the interparticle potential by simple scaling arguments that begin with the behavior
leads to the complete elimination of the nematic phase of conventional polymer composites. Since an interface
and to the direct coexistence between isotropic and limits the number of conformations that polymer mole-
crystal or columnar phases. The only limitation of this cules can adopt, the free energy of the polymer at the in-
model is that this model cannot determine the topology terfacial region is fundamentally different from that of
of the phase diagram and the nature of the ordered phases polymer molecules far away from the interface i.e. bulk.
for intermediate and high volume fractions of colloidal The influence of interface always depends on the funda-
particles. mental length scale of the adjacent matrix. In the case of
The huge interfacial area and the nanoscopic di- polymer molecules, this is of the order of the radius of
mensions between nanoelements differentiate PLS-nano- gyration of a polymer chain, Rg and this is equal to 5∼10
composites from conventional composites and filled nm. For this reason, in nanocomposites with a very few
Rheology of Polymer/Layered Silicate Nanocomposites 817

volume percent of dispersed nanofillers in polymer ma- vertify at temperatures well above the bulk transition
trix, the entire matrix may be considered to be a nano- temperature [65]. Because of the strong interactions be-
scopically confined interfacial polymer. The restrictions tween the confined molecules and the atoms or mole-
in chain conformations will alter molecular mobility, re- cules of the confining medium, the mobility of the mole-
laxation behavior, free volume, and thermal transition cules in the confined environment is greatly reduced
such as the glass transition temperature. compared to the bulk [51,68-71]. Molecules in films of
Recent molecular dynamics simulation studies have nanometer thickness organize in layers parallel to the
shown that the dynamics of the polymer chains undergo surface. However, the confining medium induces two di-
radical changes at the interfacial region. For example, mensional orders in these layers [67]. On the other hand,
both the chain mobility and the chain relaxation times in certain circumstances, confinement phenomenon may
can be slowed by three orders of magnitude near phys- also have opposite effect of enhancing molecular mobi-
isorbing surfaces. Not only that, extensive surface forces lities in a supercooled thin film, relative to the bulk.
apparatus experiments report how these novel dynamics Lee and coworkers [72] have presented an investigation
of nanoconfined polymers are manifested through vis- of the molecular mechanism of polymer melt inter-
cosity increases, two values orders of magnitude higher calation using molecular dynamics simulations. They
than the bulk values solid-like responses to imposed tried to find out the motion of polymer chains from a
shear, and confinement induced ‘sluggish’ dynamics that bulk melt into a confined volume. In their model they
suggest the existence of a ‘pinned’, ‘immobilized’ layer represented macromolecules by bead-spring chains, leave
adjacent to the confining mica surface. a reservoir of bulk melt to enter a slit of rectangular cross
The kinetics of intercalation of polymer chains into the section and fixed dimension. They adopted a coarse-
silicate galleries to form layered nanocomposites has grained description of polymers, because such a picture
been studied by Vaia and coworkers [50]. They have in- has been demonstrated to provide a useful description of
vestigated the kinetics of the intercalation of PS above its melt dynamics over longer time scales than would be ac-
entanglement molecular weight into octadecylammonium cessible with an atomistic model [73,74]. They also con-
exchanged fluorohectorite. XRD reveals the average lay- sidered a slit of fixed dimension to understand the trans-
er spacing in the unintercalated silicate to be 2.13 nm. port of a polymer melt from the bulk into a confined vol-
During intercalation, this spacing increases to 3.13 nm. ume of fixed dimension. However, they did not consider
Measurement of this layer spacing as a function of time the presence of surfactant molecules in the slit and the
during intercalation yields a time-dependent fraction of swelling of the slit during intercalation.
intercalated silicate that is directly comparable to the They found that the intercalation process can be approx-
time dependent number of beads in the slit, χ(t) and it is imately characterized by an effective diffusion co-
well fit by the prediction of a continuum diffusion model. efficient that is twice as large as the equilibrium self-con-
In their model Vaia and coworkers considered the dif- sistence in the bulk melt. Increasing the polymer-silicate
fusion of polymer chains into an empty cylinder with a interactions is found to induce spontaneous intercalation,
permeable wall and impermeable caps. The diffusion co- but for a high-polymer silicate affinity, the amount of in-
efficient determined by fitting χ(t) to this model, effec- tercalated material at a given time is reduced compared
tive diffusion coefficient, Deff, is large compared to the to the case of a weaker polymer-silicate attraction. The
equilibrium self-diffusion coefficient in bulk polystyrene crossover from polymer-silicate miscibility to interca-
of the appropriate molecular weight. These results sug- lated structures with increasing polymer-surface affinity
gested that the process of intercalation of polymer chains has already been mentioned by Vaia and coworkers [50].
into the two dimensional silicate galleries is limited by However, their study suggested that an important role
the transport of the polymer chains into the primary par- may be played by the relaxations of polymer bridges that
ticles of the silicate and not specifically by transport of connect the two silicate surfaces. The number of these
polymer chains inside the silicate galleries. bridges, as well as their dynamical properties, will be
Now to understand the detail formation kinetics and controlled by chain length [75].
physical properties of PLS-nanocomposites we need a Manias and coworkers [76] extended Vaia and cow-
clear molecular picture of the structure and dynamics of orkers studies and presented a systematic study of the ki-
confined polymers. Researchers generally used the surfa- netics of polymers entering 2 nm wide galleries of mica-
ces force apparatus [51-55] and computation studies [56- type silicates as a function of polymer molecular weight
67] to understand the behavior of confined fluids. and polymer-surface interactions. They varied the poly-
Confinement of a fluid on length scales comparable to mer-surface interactions in two ways: either through
the molecular size has been demonstrated to dramatically changing the surface modification, i.e. by varying the or-
alter its structure and dynamical pictures [51-67]. For ex- ganic coverage or through attaching strongly interacting
ample, confined fluids have been shown to solidify or sites-sticky groups-along the polymer chain. The poly-
818 Suprakas Sinha Ray

mer-surface affinity is the one of the most crucial param- must also increase the driving force for intercalation; evi-
eters during nanocomposite preparation, since it controls dently such increases depress the friction coefficient ζ
the polymer-surface monomeric friction coefficient and much more strongly.
thus determines the motion of the polymer next to a solid To understand the atomistic details of the structure of
surface. The interaction between the polymer and silicate the confined-intercalated-polystyrene chains inside the
surface can be controlled in two novel and well-con- two dimensional silicate gallery, Manias and coworkers
trolled ways: (i) keeping the polymer the same and mod- also carried out molecular dynamic simulation. Details
ifying the silicate surface, via controlled surface covering regarding simulation can be found in ref. [76]. They used
by surfactants of varying length at the same grafting den- the rotational-isomeric-state (RIS) model to create initial
sity, and (ii) keeping the organically modified surface the polymer conformations of PS oligomers. Conformations
same and modifying the polymer friction coefficient, by that fit in the interlayer gallery were chosen, and the PS
attaching along the polymer chain a controlled amount of chains were equilibrated by an off-lattice Monte Carlo
groups that interact strongly with the silicate surface. scheme that employed small random displacements of
They [76] used the PS as polymer and the same octade- the backbone atoms and orientational biased Monte
cylammonium modified fluorohectorite to study the in- Carlo rotations of the phenyl rings; at the same time, the
tercalation process. While Vaia and coworkers [50] made surfactants were equilibrated by a configurational biased
all XRD measurements on similar hybrids during in-situ scheme in coexistence with the polymer chains. After
annealing; Manias and coworkers [76] used ex-situ sam- equilibrium, MD simulations were used to obtain the
ple for XRD and other measurements. The ex-situ meth- structure and density profiles of the intercalated poly-
od has several advantages over the in-situ method: First, mer/surfactant films. The numbers of polymer chains and
one can able to anneal the sample under vacuum, thus re- alkylammonium surfactants were chosen so as to match
ducing any polymer chain degradation. Second, one can the densities found in the experimental studies. The re-
heat and subsequently investigate the same side/surface sults suggested that the confined film adopts a layered
of the pellet sample by XRD, a procedure that provides a structure normal to the solid surfaces, with the polar phe-
much more accurate control of the annealing temperature nyls dominating the organic materials adsorbed on the
than in-situ case, where pellet is heated in one side and is walls, and the aliphatic groups predominantly in the cen-
1 29
studied by XRD on the opposite side. They made several ter of the pore. H‐ Si cross-polarized nuclear magnetic
comparative studies by using concurrent in-situ small-an- resonance (NMR) measurements revealed a coexistence
gle neutron scattering (SANS) and intermediate-angle of ultra-fast and solid-like slow segmental dynamics
neutron scattering (IANS) to monitor the changes in di- throughout a wide temperature range, below and above
mensions of the polymer, i.e. Rg, and also to follow the Tg, for both the styrene phenyl and the backbone groups.
changes in the single chain scattering function during The mobile moieties concentrate at the center of the slit
intercalation. pore, especially for the higher temperatures. This leads to
Results of concurrent SANS and IANS studies shown a strong density inhomogeneity in the direction normal to
that during intercalation the silicate gallery expansion di- the surface. Fast dynamics occur in the lower density re-
rectly reflects the motion inside the 2 nm slit pore of the gions, whereas slower dynamics occur in high-segment‐
polymer chains. For the same polymer and the same an- density regions close to the surface. This heterogeneous
nealing temperature, they found the experimentally mobility combined with an observed persistence of mo-
measured effective diffusion coefficient depends strongly bility below the bulk glass transition temperature has im-
on the surfactant used for the modification of fluorohec- plications to nanocomposite properties.
torite. For several different polymer molecular weights
and annealing temperatures they observed that Deff in-
creases markedly with longer surfactant lengths and Rheology
much more than is expected just from the enhancement
of polymer mobility resulting from the polymer dilution Let us begin with‐what is rheology? Rheology is the
by small hydrocarbon oligomers. Longer surfactants re- study of the flow of matter. The name derives from the
sult in less silicate surface area exposed to the polymer, greek word “rheo” which means “to flow”. A solid will
thus effectively reduced the density attraction sites. On usually respond to a stress force by deforming and
the other hand, introducing controlled amount of groups storing energy elastically. A liquid, however, will flow
along the polymer chain that interact strongly with the and dissipate energy continuously in viscous losses. A
silicate surface, resulting in a strong decrease of Deff. Newtonian liquid has a linear relationship between the
Therefore, increasing either the density or the strength of shear rate and stress. Complex fluids are interesting
these attractive sites leads to much slower intercalation because they exhibit both an elastic and viscous
kinetics. However, an increase in site density or strength responses and exhibit a non‐linear relationship between
Rheology of Polymer/Layered Silicate Nanocomposites 819

properties not only in inhomogeneous materials but also


at high frequencies that are inaccessible using mechani-
cal rheometer. This subject is known as ‘microrheology’
[77].
Now consider a similar system where nanoscopic-sized
particles are dispersed in polymer network, enjoying rap-
id development in recent years, known as ‘nanorheol-
ogy’. Here we are going to discuss about the melt-rheo-
logical behavior of polymer materials where nanometer-
size highly anisotropic particles are dispersed, known as
polymer nanocomposites. In this section author survey
some of the current development in the melt-state rheo-
logical behavior of polymer nanocomposites based on
layered silicates.
The measurement of rheological behavior of any poly-
meric material under molten state is crucial to gain fun-
damental understanding of its processability. In the case
of PLS-nanocomposites, the melt-rheological properties
are also helpful to find out the degree of polymer-filler
Figure 5. Schematic diagram describing the end‐tethered
interactions and the structure-property relationship in
nanocomposites. The layered silicates are highly anisotropic
with a thickness of 1nm and lateral dimensions (length and nanocomposites. This is because rheological material
width) ranging from ∼100 nm to a few microns. The polymer functions are strongly influenced by the structure and the
chains are tethered to the surface via ionic interactions between interfacial properties.
the silicate layer and the polymer-end. Reproduced from ref
[78] by permission of American Chemical Society. Linear Dynamic Rheological Properties
Small deformation amplitude oscillatory shear measure-
the shear rate and strain. ments of polymeric materials are generally performed by
One of the properties often dealt within rheology is applying a time dependent strain
viscosity which measures how thick a fluid is. Some
materials are intermediate between solids and fluids and γ(t) = γ0 sin(ωt) (1)
the viscosity is not enough to characterize them. A solid
material can be described by its elasticity or resilience: and measuring the resultant shear stress
when it is deformed it will store the energy and fight
back. Imagine a spring that regains its original shape σ(t) = γ0 [G'sin(ωt) + G"cos(ωt) ] (2)
after being deformed. The other extreme is a fluid which
stores no energy while deformed and just flows. A where G' and G" are the storage and loss moduli, re-
viscoelastic material is intermediate and stores some spectively and ω is the frequency.
energy and flows a little when deformed. Consider a Krishnamoorti and coworkers [78] first described the
system where a micron-sized probe particle is doped flow behavior of delaminated nanocomposites based on
inside a rheologically complex fluid such as polymer poly(ε-caprolactone) (PCL) and nylon 6 (N6). Both
network and Brownian fluctuations are recorded. Now nanocomposites were prepared by in-situ polymerization
question is how to predict the mechanical properties of method in the presence of organically modified MMT,
the medium from the observed Brownian motion of the which engendered direct grafting of macromolecular
particle. If the fluid is Newtonian and the probe molecule chains on the MMT surface-a hairy layered silicate
is spherical, one can easily determine the viscosity of the platelets structure (see Figure 5). The authors coded these
medium from the measured diffusion coefficient by nanocomposites as ‘end-tethered polymer layered silicate
using the standard Stokes-Einstein (SE) relation. How- nanocomposites’. Figure 6 represents the rheological
ever, a polymer network is generally viscoelastic, with a master curves based on linear viscoelastic measurements
complex shear modulus having both elastic i.e. solid-like for a series of end-tethered PCL-MMT nanocomposites.
and viscos i.e. liquid-like components of similar magni- The flow behavior of PCL nanocomposites differed
tudes over a large range of frequencies. Now question is- significantly from that of the corresponding neat
is it possible to extend the SE relation to determine shear matrices, whereas the thermorheological properties of the
complex modulus for such a system? An affirmative nanocomposites were entirely determined by the
answer would allow one to measure the local rheological behavior of the matrices. The slopes of log G'(ω) and
820 Suprakas Sinha Ray

Figure 6. Storage modulus (G') for PCL‐based silicate nan-


ocomposites. Silicate loadings are indicated in the figure.
Master-curves were obtained by application of time-temper-
ature superposition and shifted to T0 = 55 oC. Reproduced from
ref. [78] by permission of American Chemical Society.

G"(ω) versus log aTω are much smaller than 2 and 1,


respectively. Values of 2 and 1 are expected for non-
crosslinked polymer melts, but large deviations occur,
especially in the presence of a very large amount of
layered silicate loading, which may be due to the
formation of a network structure in the molten state.
However, nanocomposites based on the in-situ poly-
merization technique exhibit a fairly broad molar mass
distribution in the polymer matrix, which hides the Figure 7. Time-temperature superposed linear storage modulus
relevant structural information and impedes the inter- (a, top) and loss modulus (b, bottom) for the series of
pretation of the results. The similar behavior was also 2C18M-based PSPI18 intercalated hybrids. As expected, the
observed in the case of end-tethered N6-MMT nan- moduli increase with increasing silicate loading at all
frequencies. At high frequencies, the qualitative behavior of the
ocomposites.
storage and loss moduli are essentially unaffected. However, at
Ren and Krishnamoorti [79] then showed significant low frequencies the frequency dependence of the moduli grad-
modification of flow behavior of nanocomposites when ually changes from liquidlike to solidlike for nanocomposites
polymer chains are not tethered on the silicate surface. with 6.7 and 9.5 wt% silicate. Reproduced from ref. [79] by
The nanocmposites of polystyrene-1, 4-polyisoprene (7 permission of American Chemical Society.
mol% 3, 4 and 93 mol% 1, 4) diblock copolymer (PSPI
18) were prepared with (0.7, 2.1, 3.5, 6.7, and 9.5 wt%) intercalation of PS into the silicate layers may be due to
dimethyl dioctadecyl ammonium modified MMT by the slight Lewis base character imparted by the phenyl
solution mixing in toluene. The homogeneous solution ring in PS, leading to the favorable interactions with the
was dried extensively at room temperature and 2C18-MMT layers.
o
subsequently annealed at 100 C in a vacuum oven for ∼ All rheological measurements were conducted in the
12 h to remove any remaining solvent and to facilitate temperature range of 80∼105 °C. The time-temperature
complete polymer intercalation between the silicate master curves for neat PSPI and various nanocomposites
layers. The XRD patterns of PSPI18/2C18-MMT and are presented in Figure 7. It is clear from the figures that
PS/2C18-MMT clearly indicates the formation of viscoelastic behavior of PSPI significantly modified after
intercalated structures, whereas the composite prepared nanocomposite formation. At high frequency region
with 1, 4-polyisoprene (PI) showed no change in gallery (where frequency > 10 rad.s), viscoelastic properties of
height. The intercalation of 2C18-MMT by PS and not PSPI were unaffected by the presence of silicate particle,
by PI is consistent with the results of previous while at the low frequency region (region where fre-
experimental work and theories of Vaia and coworkers quency < 10 rad.s), both G' and G" for the nano-
[37,38] and of Balazs and coworkers [39,40]. The composites exhibited a significant diminished frequency
Rheology of Polymer/Layered Silicate Nanocomposites 821

dependence and this behavior was more prominent with foliated hybrids.
higher loading of silicate layers, similar behavior as ob- To verify above percolation theory, Krishnamoorti and
served in the case of end-tethered nanocomposite. coworkers estimated the percolation threshold for these
Similar behavior has also been noted for carbon black hybrids on the basis of stack size of the intercalated sili-
[80], calcium carbonate [81], and talc [82] filled cate layers. They assumed a hypothetical hydrodynamic
polymers. However, in those cases the particle volume sphere surrounding each tactoid and the percolation of
fractions were substantially higher, 5∼10 wt% for which these hydrodynamic spheres to signify the onset of in-
strong frequency independent behavior is seen for PLS‐ complete relaxation and the presence of pseudo-solid-
nanocomposites. like behavior in the intercalated nanocomposites. To find
Time-temperature superposition required simultaneous out the relation between the number of silicate layers per
horizontal and vertical shifting of the shear moduli: bT G' tactoid, nper and the layered silicate weight fraction at
and bT G" versus aTω, where aT is the horizontal shift percolation, wsil, per, as
factor, bT is the vertical shift factor and ω is the
frequency. While aT for all nanocomposites followed the        
    
same WLF (Williama-Landel-Ferry) [83] dependence but              

bT did not. The authors attributed this could be the block (3)
architecture of the polymer examined, and the relative
proximity of the order-disorder temperature to the tem- where Rh is the radius of the hydrodynamic volume (in
perature of the experimental measurements or may be this case equivalent to the radius of the disk-like layered
due to formation of different types of structural ori- silicates), øper is the percolation volume fraction, hsil is
entation in the presence of organocaly. the thickness of the silicate layers (generally 1 nm), ρorg
To understand the viscoelastic behavior observed at low and ρsil are the densities of the organic component and
frequencies region in the dynamic measurements for the layered silicate respectively. By considering the uniform
nanocomposites, they conducted stress relaxation mea- disks diameter (2Rh) of 0.5 µm and the wsil, per to be
surements. The stress relaxation data of these nano- 0.067, they calculated the average tactoid size to be
composites in the terminal region showed a solid-like around 30 layers. This calculated value was well matched
behavior. Like G', the effect was particularly pronounced with the value calculated from XRD pattern and TEM
at ≥ 6.7 wt% silicate loading. The authors attributed this measurement. Furthermore, this tactoid size implies that
behavior is due to the presence of stacks of intercalated the effective anisotropy associated with the filler is 5∼
silicate layers and those are randomly oriented in the pol- 10. It is this anisotropy, along with the random relative
ymer matrix, forming three dimension network structure. arrangement of the tactoids that leads to the observation
These stacked intercalated silicate layers have only trans- of the percolation phenomenon at extremely low loading
lational motion. A large-amplitude oscillatory shear is of the silicate.
able to orient these structures and reduce the solid like Krishnamoorti and coworkers also used above equation
behavior. Another explanation may be the formation of to estimate the percolation threshold for the exfoliated
PI micro-domains. Results of XRD revealed that only nanocomposites by assuming the same geometrical pa-
PS block diffused into the interlayer galleries, leaving the rameters as considered previously. They suggested that
PI block out of the silicate galleries. These PI blocks can in the case of exfoliated nanocomposites, there exist con-
form phase-separated micro-domain structures, which ac- siderable local orientational order which has enormous
tually hinder the orientation of stack intercalated silicate consequences for the physical and mechanical properties
layers under small-amplitude shear. of the hybrids. This orientational ordering along with the
Another explanation is the physical jamming of the dis- end-tethering of the polymer chains on the layered sili-
persed stack intercalated silicate layers owing to their cate surface in fact responsible for the non-terminal inter-
highly anisotropic nature. Transmission electron micro- mediate frequency response in the PCL and N6-based
scopy image revealed the formation of intercalated struc- nanocomposites.
ture for the PSPI-based nanocomposites. On the basis of Over the last few years, the linear dynamic and steady
this mesoscopic structure and at low silicate loadings, au- rheological properties of several nanocomposites have
thors suggested that, beyond a critical volume fraction, been examined for a wide range of polymer matrices
the tactoids and the individual layers are incapable of including N6 with various matrix molecular weights
freely rotating and when subjected to shear are prevented [84], polyamides [85-91], polystyrene [92-96], poly-
from relaxing completely. This incomplete relaxation propylene [97-102], polyethylene [103], ethylene vinyl
due to the physical jamming or percolation of the nano- copolymers [104-106], polycarbonate [107-109], poly
scopic fillers leads to the presence of the pseudo-solid- (ethylene oxide) [110-112], poly(ε-caprolactone) [113,
like behavior observed in both the intercalated and ex- 114], poly(ethylene terephthalate) [115], polylactide
822 Suprakas Sinha Ray

Figure 8. Complex viscosity versus frequency from a dynamic parallel plate rheometer (solid points) and steady shear viscosity ver-
sus shear rate from a capillary rheometer (open points) at 240 °C for (a) pure HMW and its (HE)2M1R1 organoclay nanocomposite,
(b) pure MMW and its (HE)2M1R1 organoclay nanocomposite. The nanocomposites contain ∼3.0 wt% MMT. Reproduced from ref.
[84] by permission of Elsevier Science Ltd., UK.

[116-118], poly(butylene succinate) [119-121], poly under molten state revealed that an increase in shear vis-
[(butylene succinate)-co-adipate] [122,123], synthetic cosity and storage and loss moduli of nanocomposites
biodegradable aliphatic polyester (BAP) [124,125], with clay content. Like other properties, they also fond
liquid crystal polymer [126,127], PLS-nanocomposites that rheological properties of nanocomposites directly re-
based on immiscible and miscible polymers blend [128, lated to the degree of dispersion of silicate layers in the
129], etc. polymer matrix and also the level of interfacial inter-
Choi and coworers described the melt-state rheological actions between the layered silicate surface and polymer
behavior of a series of PEO/organoclay nanocomposites chains. They concluded that linear rheological properties
[111,112]. To understand both the effect of degree of dis- are strongly correlated with the mesoscopic structure and
persion of silicate layers and the interaction between the it was postulated that the molecular weight and inter-
organoclay surface and polymer matrix, they used three action strength would affect the mesoscopic structure and
types of surfactants modified montmorillonite. First, they the rheological properties of these nanocomposites.
selected two different organoclays, which have different Fornes and coworkers [84] have conducted dynamic
modifier concentration but the same alkylamonium salt. and steady shear capillary experiments over a large range
Second, they differentiated the types of alkylamonium of frequencies and shear rates of pure N6 with different
salts having the same organic modifier. All nanocompo- molecular weights, and their nanocomposites with
sites were prepared via solvent casting method using OMLS. Figure 8 shows bilogarithmic plots of the
chloroform as a co-solvent. XRD patterns and TEM ob- complex viscosity, |η*| vs. ω at 240 oC for pure N6 and
servations revealed the formation of intercalated (HE)2M1R1 nanocomposites based on (a) high molecular
nanocomposites. Rheological properties measurements weight (HMW), (b) medium molecular weight (MMW),
Rheology of Polymer/Layered Silicate Nanocomposites 823

three systems follows the order HMW > MMW > LMW,
and hence the resulting shear stresses exerted by the pure
polymers also follow the same order. Therefore, during
melt mixing the level of stress exerted on the OMLS by
the LMW polyamide is significantly lower than those de-
veloped in the presence of HMW or MMW polyamides.
As a result, the break-up of layered silicate particles is
much easier in the case of HMW polyamides, and lay-
ered silicate particle dispersion is ultimately improved.
The role of polymer molecular weight is believed to stem
from an increase in the melt viscosity, facilitating the
degradation of the taller stacks into shorter ones. The fi-
nal step in exfoliation involves peeling away the platelets
of the stacks one by one, and this takes time and requires
the strong matrix-OMLS interaction to cause sponta-
neous wetting.
Aubry and coworkers were studied the dynamaic and
steady flow properties of a polyamide-12 (PA12) melt
layered silicate nanocomposites as a function of the sili-
Figure 9. Influence of frequency on shear storage modulus for cate volume fraction, ø [88]. They found above a vol-
pure LMW, and for LMW, MMW and HMW based nano- ume fraction threshold øp is equal to 1.5 %, and below a
composites. Reproduced from ref. [84] by permission of critical strain γc, the storage (G') and loss moduli (G")
Elsevier Science Ltd., UK. were shown to exhibit a low-frequency plateau and the
flow curve was shown to exhibit a stress, τy. The study
and (c) low molecular weight (LMW), obtained using the of G', γc, and τy as a function of ø showed the energy
parallel plate oscillating rheometer. Figure 8 also shows needed for removing connectivity on a mesoscale did not
a bilogarithmic plot of the steady-state shear viscosity η depend on the silicate loading. These original properties
versus shear rate  , obtained using a capillary rheometer. were attributed to the existence, in the quiescent state, of
Inspection of these figures reveals a significant differ- mesoscopic domains composed of correlated silicate
ence between the nanocomposites, particularly at low layers. Moreover, the steady shear response of all sam-
frequencies. The HMW-based nanocomposites show ples at solid volume fractions above øp showed the ex-
very strong non-Newtonian behavior, and this is more istence of a critical shear rate is equal to 1/s, separating a
pronounced at low frequencies. On the other hand, the behavior governed by the networked domains from a be-
magnitude of the non-Newtonian behavior gradually de- havior dominated by the polymer matrix.
creases with decreasing molecular weight of the matrix, Gelfer and coworkers have investigated the particular
and with LMW it behaves like pure polymer. This trend relationship among the molecular structure of polymer
is more clearly observed in the plot of G' vs. ω, due to chains, morphology, and rheology of nanocomposites
the extreme sensitivity of G' towards dispersed morphol- prepared by melt-blending of Cloisite organoclays and
ogy in the molten state (see Figure 9). The difference in ethylene-co-vinyl acetate (EVA) and ethylene-co-methyl
the terminal zone behavior may be due to different ex- acrylate (EMA) random copolymers [104,106]. TEM ob-
tents of exfoliation of the clay particles in the three types servation confirmed the mixed intercalated-exfoliated
of matrices. morphology in these nanocomposite systems. The melt-
At the other extreme, the steady shear capillary rheol- state rheological properties were very similar in EVA-
ogy shows a trend with respect to the matrix molecular and EMA-based nanocomposites. These materials ex-
weight. The HMW and MMW-based nanocomposites hibited a pseudo-solid like rheological behavior, which
exhibit lower viscosities compared to that of their corre- became more pronounced upon heating, especially at
sponding matrices, whereas the viscosities of LMW- temperature above 180∼200 °C. Authors attributed this
based nanocomposites are higher than the pure matrix. behavior is due to the physical gelation due to the for-
This behavior is also due the higher degree of exfoliation mation of 3D tactoid network in the polymer matrix.
in the case of HMW and MMW-based nanocomposites They found that this gelation behavior is directly related
compared to the LMW-based nanocomposite. to the overall content of organoclay and the extent of
Finally they considered the differences in the melt vis- miscibility between the organoclay and polymer matrix.
cosity among the three systems. Over the range of fre- The SAXS data indicated that the silicate gallery
quencies and shear rates tested, the melt viscosity of the spacing (d001), intercalated by EVA and EMA chains,
824 Suprakas Sinha Ray

Figure 10. G' vs. reduced frequency for PS100 matrix (black
line), CPS (A), and C-PEA (W). Silicate loading: 5 wt%.
Reproduced from ref. [89] by permission of Wiley-VCH,
Germany.

decreases with increasing temperature. At temperature


above 200 °C, the desorption of surfactant in organoclay
decreases the compability between the clay surface and
the polymer matrix, which finally exhibited a LCST
(lower critical solution temperature) type behavior Figure 11. Reduced frequency dependence of storage modulus,
between organoclays and polymer, enhancing the state of G' and loss modulus, G" of PLA and various PLACNs.
gelation. This behavior manifested itself by a reverse Reprinted from ref. [118] by permission of Wiley-VCH,
temperature dependence of viscoelastic properties and Germany.
strong deviation of rheological behavior from the time‐
temperature superposition principle. dynamic oscillatory shear measurements of PLA nano-
Hoffmann and coworkers [89] recently conducted rheo- composites with intercalated structure. Melt rheological
logical measurements of nanocomposites of polystyrene measurements were conducted on Rheometric Dynamic
(PS100) with synthetic mica (ME100) modified with Analyzer (RDAII) instrument using 25 mm diameter
amine-terminated polystyrene (AT-PS8) and 2-phenyl- parallel plates with a sample thickness of ca. 1.5 mm and
ethylamine (PEA) respectively, in order to verify the in the temperature range of 175∼205 °C. To avoid
presence of a particle network formation via inter-par- nonlinear response, the strain amplitude was fixed to 5 %
ticle interaction and self-assembly. Figure 10 represents to obtain reasonable signal intensities even at elevated
G'(ω) vs. aTω for PS100 and its corresponding nano- temperature or low w. For each nanacomposite investiga-
composites prepared with AT-PS8 modified ME100 (C- ted, the limits of linear viscoelasticity were determined
PS) and PEA modified ME100 (C-PEA). The rheological by performing strain sweeps at a series of fixed
responses of PS100 and C-PEA are the same, whereas frequencies. The master curves were generated using the
the rheological response of C-PS is completely different. principle of time-temperature superposition and shifted
At the lowest frequencies, which correspond to the to a reference temperature (Tref) of 175 °C, which was
marked region III, G'(ω) strongly increases, and the chosen as the most representative of a typical processing
slope approaches to zero. Such behavior is an indication temperature of PLA.
of network formation involving the assembly of in- The master curves for G' and G" of pure PLA and
dividual plates composed of silicate layers. In the regime various nanocomposites with different weight percenta-
of intermediate frequencies (region II, see Figure 10) the ges of C18 MMT loading are presented in Figure 11. At
G'(ω) value is lower in comparison to that of PS100. high frequencies (aT.ω > 10), the viscoelastic behavior
This might be due to dilution of the amine-terminated PS of all nanocomposites was the same. On the other hand,
being below the entanglement molecular weight. On the at low frequencies (aTω < 10), both moduli exhibited
other hand, in high frequency region (region I, Figure 10) weak frequency dependence with increasing C18MMT
the rheological behavior is the same for all systems. content, with a gradual change of behavior from liquid-
Sinha Ray and coworkers [116-118] have conducted the like [G' ∝ω2 and G" ∝ω] to solid-like ( G' ≥ G") with
Rheology of Polymer/Layered Silicate Nanocomposites 825

Figure 12. Flow activation energy as a function of C18- Figure 13. Reduced frequency dependence of complex vis-
MMTcontent. Reprinted from ref. [118] by permission of cosity (|η*|) of PLA and various PLACNs. Reprinted from
Wiley-VCH, Germany. ref. [118] by permission of Wiley-VCH, Germany.

increasing C18MMT content. lastic behavior was almost identical to that obtained for
The slope of G' and G" in the terminal region of the neat PBS. At high frequencies (aTω< 5), both moduli
master curves of PLA matrix was 1.85 and 1 exhibited weak frequency dependence with increasing
respectively. On the other hand, the slopes of G' and G" clay content, which means that there are gradual changes
were considerably lower for all PLACNs compared to of behavior from liquid-like to solid-like with increasing
those of pure PLA. In fact, for PLACNs with high clay content.
C18MMT content, G' becomes nearly independent at low Lepoittevin and coworkers [113] reported the melt rheo-
aTω and exceeds G", characteristic of materials exhibit- logical properties of PCL-based nanocomposites pre-
ing a pseudo-solid-like behavior. pared by melt intercalation method. G' and G" of unfilled
+
Figure 12 represents the C18MMT content dependent PCL and PCL filled with 3 wt% of MMA-Na , MMT‐
(wt%) flow activation energy (Ea) of pure PLA and vari- Alk or MMT-(OH)2 were measured at 80 ºC in the fre-
‐2
ous nanocomposites obtained from an Arrhenius fit of quency range from 16 to 10 Hz. The rheological behav-
master curves [118]. There was a significant increase of ior of the PCL filled with 3 wt% of MMT-Alk and
Ea for PLA/C18 MMT3 compared to that of pure PLA MMT-(OH)2 was significantly different compared to un-
followed by a much slower increase with C18MMT filled PCL and PCL/MMT-Na composites. The fre-
content. This behavior may be due to the dispersion of quency dependence of G' and G" was, however, per-
intercalated and stacked C18MMT silicate layers in the turbed by organically modified MMT. The effect was
PLA matrix. dramatic in the slope of G' which drops from 2 to 0.14
The dynamic complex viscosity (|η*|) master curves and 0.24 for MMT-(OH)2 and MMT‐Alk, respectively.
for the pure PLA and nanocomposites, based on linear The dependence of G' and G" on frequency with the
dynamic oscillatory shear measurements, are presented in filler content is presented in Figure 15 for the MMT-
‐1
Figure 13 [118]. At low aTω region ( < 10 rad.s ), pure (OH)2 clay. When the clay content exceeded 1 wt%, not
PLA exhibited almost Newtonian behavior while all only the classical power laws for the frequency depend-
nanocomposites showed very strong shear-thinning ence of G' and G" were deeply modified, particularly in
tendency. On the other hand, Mw and PDI of pure PLA the case of G' but the moduli increased dramatically at
and various nanocomposites were almost the same, thus low frequency. This is a characteristic of pseudo-solid-
the high viscosity of PLACNs were explained by the like behavior due to the formation of a network percolat-
flow restrictions of polymer chains in the molten state ing clay lamellae. The same behavior was also observed
due to the presence of MMT particles. in the case of PLA or PBS-based nanocomposites [118,
Figure 14 represents the master curves for G' and G" of 120,121].
neat PBS and various PBSCNs prepared with two Sinha Ray and his coworker recently described the
different types of OMLS [120]. At all frequencies, both melt-state rheological properties of poly[(butylene
G' and G" of nanocomposites increased monotonically succinate)-co-adipate] (PBSA) nanocomposites [122].
®
with increasing OMLS loading with the exception of Cloisite 30B (C30B) organoclay was used for the
PBS/C18MMT1 and PBS/qC16SAP1 for which viscoe- preparation of nanocomposites. Nanocomposites were
826 Suprakas Sinha Ray

Figure 14. (a) Reduced frequency dependence of the storage modulus,G'(ω), and the loss modulus, G"(ω), of PBS and various
PBSCNs prepared with C18-mmt clay. Tref ) 120 °C. (b) Reduced frequency dependence of the storage modulus, G'(ω), and the
loss modulus, G"(ω), of PBS and various PBSCNs prepared with qC16‐sap. Tref ) 120 °C. Reproduced from ref. [120] with permis-
sion of American Chemical Society, USA.
o
prepared at 135 C in a batch mixer. Nanocomposites were conducted. Results of linear relaxation behaviors of
(PBSACNs) prepared with three different amount of neat PBSA and PBSACNs are presented in Figure 17.
C30B of 3, 6, and 9 wt% were correspondingly For any fixed time after imposition of strain, the relaxa-
abbreviated as PBSACN3, PBSACN6, and PBSACN9. tion modulus, G(t) increases with increasing C30B load-
The structure of the nanocomposites was studied using ing, similar to that observed in the dynamic oscillatory
XRD and TEM that revealed a coexistence of exfoliated shear measurements.
and intercalated silicate layers dispersed in the PBSA Similar to the behavior observed at high frequencies in
matrix, regardless of the silicate loading. dynamic mode (Figure 16a and b), the stress relaxation
The angular ω dependence of storage, G'(ω) and loss modulus of all PBSACNs is quite of the same magnitude
modulus, G"(ω) of neat PBSA and PBSACNs with three as that of the neat PBSA at short times, with only a slight
different wt% of C30B loading are presented in Figure increase with C30B content (Figure 17). However, at
16, parts a and b, respectively. At all frequencies, both long time periods, the pure PBSA and PBSACN3 relaxes
G'(ω) and G"(ω) for PBSACN6 and PBSACN9 in- like a viscoelastic liquid, while the PBSACNs with 6 and
crease with increasing C30B content with the exception 9 wt% of C30B loading behave like a pseudo-solid-like
of PBSACN3 for which the viscoelastic response is al- material.
most identical to that observed for neat PBSA. At high- Therefore, on the basis of both the dynamic oscillatory
ω region, the viscoelastic behavior of all PBSACNs is shear (Figure 16a and b) and linear stress relaxation
quite the same, with only a small systematic increase in measurements (see Figure 17), it is clear that C30B lay-
G'(ω) with the C30B loading, indicating that the ob- ered silicate has a profound effect on the long relaxation
served chain relaxation modes are almost unaffected by time of the nanocomposites. For PBSACNs with C30B
the presence of the layered silicate particles (Figure 16a loading in excess of 3 wt% , the liquid-like behavior ob-
and b). However, at the low-ω region, both dynamic served in the case of pure PBSA gradually changes to
moduli exhibit weak frequency dependence with C30B pseudo-solid-like behavior.
loading. For the 9 wt% of C30B loading, G'(ω) exceeds A reasonable explanation for this low frequency
G"(ω) and becomes nearly independent of ω. viscoelastic behavior of PBSACNs is that the moderate
To verify the interesting viscoelastic behavior observed interactions of the PBSA backbone with C30B surface
at low frequency region in the dynamic oscillatory shear lead to the high degree of confinement of polymer chains
measurements for the nanocomposites with high wt% inside the silicate layers. As a result anisotropic silicate
C30B loading, the linear stress relaxation measurements layers are fully dispersed in the PBSA matrix, as
Rheology of Polymer/Layered Silicate Nanocomposites 827

Figure 15. Storage modulus (G0) and loss modulus (G00) for Figure 16. Frequency dependence of (a) dynamic storage, G'
unfilled PCL and PCL modified by 3 wt% of MMT-Na, (ω) and (b) loss, G"(ω) moduli of neat PBSA and various
MMT-Alk, MMT-(OH)2 at 80 C [160]. Reproduced from ref. PBSACNs. Reprinted from ref. [122] by permission of Wiley‐
[113] by permission of Elsevier Ltd., UK. VCH, Germany.

observed in XRD patterns and TEM images [122]. percolating network that imparts the blend with solid-like
Because of their highly anisotropic nature, the dispersed behavior. Such a behavior can also be seen on the
layered silicate particles would exhibit local correlations dynamic complex viscosity, η*(ω), shown in Figure
that finally cause the formation of micro-domains or 18. Little effect of C30B addition is observed at high
mesoscopic structure [116,124,125]. Like in liquid frequencies, where the relaxation mechanism is mainly
crystalline and ordered block copolymer systems [130- dominated by that of the PBSA matrix, whereas at low
133] the presence of these micro-domains or mesoscopic frequencies, the relaxation is that of particle-particle
structure causes only a slight increase of the low-ω interactions inside the percolating network of the silicate
elastic modulus that varies in low power-law fashion. layers.
This is the case for low C30B concentration (∼3 %). In a recent literature Huang and Han reported the melt-
However, beyond this concentration, exfoliated and or state rheological behavior of PLS nanocomposites based
disordered intercalated silicate layers form a network- on thermotropic liquid-crystalline polymers (TLCP)
type structure rendering the system highly elastic as [126]. In their study they used two different types of
revealed by the low frequency plateau. Such a plateau is TLCPs such as TLCP having pendent pyridyl group
similar to the one observed in rubber-toughened (PyHQ12) and TLCP having pendent phenylsulphonyl
polymers [134,135] where rubber particles form a group (PSHQ12). They also used two different types of
828 Suprakas Sinha Ray

Table 1. Characteristics of Organically Modified Layered Silicates (OMLS) used in this Research
Modifier Concentration a
Organoclay Type Code Chemical Structure of Organic Modifier δ of Organic Modifierb, c
(meq/ 100 g)

CH 3

Cloisite 15 A C15 A 125 H3C N+ HT 16.94

HT

Cloisite 93 A C93 A 90 H3C N+ HT 17.96

HT

CH 2CH 2OH

Cloisite 30 B C30 B 90 H3C N+ HT 21.08

CH 2CH 2OH
a
HT is hydrogenated tallow (∼65 % C18; ∼30 % C16; ∼5 % C14)
b
δ is the solubility parameter calculated by group contribution method of Fedors
c
δ value of neat PBSA is 23.8

Authors presented dramatic differences in linear dy-


namic viscoelastic properties between PyHQ12/C30B
and PSHQ12/C30B nanocomposites. They attributed that
the presence of strong interaction between PyHQ12 poly-
mer chains with C30B surface through hydrogen bond is
main responsible factor for significant improvement in
rheological properties in case of PyHQ12/C30B nano-
composite. Authors also observed a considerable amount
of loss in the degree in liquid crystallinity of PyHQ12
matrix in PyHQ12/C30B nanocomposite, although it has
intercalated structure. This is due to the strong interaction
between polymer backbone and clay surface, which fi-
nally lead to the highly restricted movement of the poly-
mer chains in nanocomposite. The same type of con-
clusion has been made by this group of researchers in
case of PC/C30B nanocomposite, but they did not con-
sider the degradation of matrix or C30B at the processing
Figure 17. Stress relaxation behavior of Neat PBSA and vari- temperature (260 oC) [109].
ous PBSACNs. Reprinted from ref. [122] by permission of To find out the direct relation between the structure and
Wiley-VCH, Germany. melt-state rheological properties in case of nanocompo-
sites, Sinha Ray and coworkers prepared a series of PLS
OMLS such as C30B and C20A for the preparation of nanocomposites based on PBSA [123]. Three different
nanocomposites. Nanocomposites were prepared by types of commercially available organically modified
solution blending method using pyridine as a co-solvent. montmorillonite (OMMT) were used for the preparation
XRD patterns and TEM revealed the high degree of of nanocomposites. By varying the hydrophobicity, and
dispersion of silicate layers in PyHQ12/C30B nanocom- consequently the favorable interactions of the surfactant
posites, whereas silicate layers formed aggregates in with the polymer matrix, a spectrum of structures
PyHQ12/C20A and PSHQ12/C30B nanocomposites. including phase-separated, intercalated, disordered inter-
Rheology of Polymer/Layered Silicate Nanocomposites 829

is presented in Figure 20. As expected, G'(ω) of the


nanocomposites increases with increasing the degree of
dispersion of the silicate layers in the PBSA matrix over
all frequency range. At high frequencies, only a slight in-
crease in modulus with respect to that of the pure PBSA
is observed. G'(ω) of the matrix and the nanocomposites
decreases linearly with lowering of frequency. However,
at very low frequencies a significant increase in G'(ω) is
observed in the case of PBSA/C30B nanocomposite, in-
dicating that C30B silicate layers have strong effect on
the chain relaxation of PBSA. Upon increase in the fa-
vorable specific interactions between the surfactant and
PBSA matrix, G'(ω) becomes a weak function of fre-
quency and exhibits a pseudo-plateau in the terminal
zone of the pure PBSA. This gradual change of behavior
from liquid-like to solid-like is mainly attributed to the
extent of dispersion and distribution of the clay lamellae
Figure 18. Frequency dependence of dynamic complex vis- that form three dimensional percolating networks [116].
cosity (η*) of neat PBSA and various PBSACNs. Reprinted This of course depends on the extent of enthalpic
from ref. [122] by permission of Wiley-VCH, Germany. interactions between the PBSA backbone and the surface
of OMMTs, which finally leads to the confinement of
calated, and intercalated-exfoliated were obtained (Fig- polymer chains inside the silicate galleries. Because of
ure 19). The chemical structure of the specific surfactant their highly anisotropic nature, the fully dispersed silicate
and the cation exchange capacity of each of the OMMT layers would exhibit local correlations, which finally lead
used in this research are presented in Table 1. to the formation of three dimensional mesoscopic
The small amplitude oscillatory shear storage modulus, structures [123] and the tendency of formation of this
G'(ω), of the neat PBSA and the three nanocomposites mesoscopic structure gradually decreases with decreasing

Figure 19. TEM bright field images of (a) PBSA/6CNa hybrid, (b) PBSA/6C15A nanocomposite, (c) PBSA/6C93A nanocomposite,
and (d) PBSA/6C30B nanocomposite. Reproduced from ref. [123] by permission of Elsevier Science Ltd., UK.
830 Suprakas Sinha Ray

Figure 20. Frequency dependence of dynamic storage modulus, Figure 21. Frequency dependence of dynamic complex vis-
(G'(ω)), of neat PBSA and various nanocomposites. Repro- cosity (η*) of neat PBSA and various nanocomposites.
duced from ref. [123] by permission of Elsevier Science Ltd., Reproduced from ref. [123] by permission of Elsevier Science
UK. Ltd., UK.

the degree of dispersion of intercalated silicate layers in Furthermore, the elastic modulus of PPCH8 at low fre-
the polymer matrix. Among the three OMMTs used for quency decreases with annealing time, but that of PPCH9
the preparation of PBSA nanocomposites, C30B has the remains more or less unchanged. Thus at high frequency,
higher degree of interactions with PBSA matrix and as a the response of the PPCH9 hybrid is dominated by the
result intercalated silicate layers are relatively well matrix, while at lower frequencies its solid-like response
dispersed in the PBSA matrix (see Figure 19). Conse- is strongly influenced by the presence of clay. On the
quently, silicate layers are highly correlated in the other hand, the low frequency response of PPCH8 is not
PBSA/C30B nanocomposite. With decreasing the degree dominated by the presence of clay, which highlights the
of dispersion of intercalated layers in the PBSA matrix, important role played by PP-MA in the formation of the
the tendency of formation of this type of correlated hybrids.
structure is also decreased and thus the increment in The above results show that the solid-like rheological
modulus systematically decreases from PBSA/C30B behavior of nanocomposites at lower frequencies is com-
nanocomposite to PBSA/C15A nanocomposite. The pletely independent on the fine structure of the nano-
solidlike behavior can also be evidenced by the sharp composites i.e., whether it is end-tethered or stacked in-
increase in viscosity in the low frequency region for tercalated, but it is depends primarily upon the amount of
nanocomposites, where the pure PBSA shows rather a clay loading in the nanocomposites. Galgali and cow-
Newtonian behavior (see Figure 21). orkers [97] have also shown that the typical rheological
Parts a and b of Figure 22 show the time dependent response in nanocomposites arises from frictional inter-
evolution of the storage and loss moduli for PPCH8 (PP actions between the silicate layers and not due to the im-
= 91 wt%, MMT = 9 wt%) and PPCH9 (PP = 82 wt%, mobilization of confined polymer chains between the sil-
MA = 9 wt% and MMT = 9 wt%) samples, subjected to icate layers. They have also shown a dramatic decrease
small strain oscillatory shear tests every 10 min during in the creep compliance for the PPCH prepared with MA
their total annealing time of 3 h at 200 oC [97]. At high modified PP and 6 wt% MMT. In contrast, for PPCH
frequencies the elastic moduli of PPCH8 and PPCH9 are prepared with MA modified PP and 9 wt% of MMT,
comparable at similar annealing times, and both decrease they show a dramatic three-order of magnitude drop in
with annealing due to possible degradation of the PP- the zero shear viscosity beyond the apparent yield stress,
matrix. At low frequencies the elastic modulus of PPCH suggesting that the solid‐like behavior in the quiescent
9 is qualitatively different than that of PPCH 8. The state is a result of the percolated structure of the layered-
low-frequency elastic modulus of as-extruded PPCH9 is silicate.
always higher than that of PPCH8, indicating that most In order to characterize the viscoelastic properties of the
of the micro-structural development in PPCH9 has so-called network formation, Aubry and coworkers [88]
already occurred during the extrusion process, while only measured G' as a function of the layered silicates volume
subtle micro-structural changes occurred during annealing. fraction, ø. Parts a and b of Figure 23 respectively
Rheology of Polymer/Layered Silicate Nanocomposites 831

Figure 22. Zero shear viscosity as a function of annealing time Figure 23. (a) Storage modulus G' as a function of frequency at
for several PPCH samples in the presence (denoted by filled (●) 0 %, (O) 0.5 %, (■) 1 %, (□) 1.5 %, (∆) 2.5 %, (▲) 5
symbols) and absence (denoted by open symbols) of PP-MA. %, and (▼) 10 %. (b) Loss modulus G" as a function of fre-
(b) Zero-shear viscosity as a function of clay content for sam- quency at (●) 0 %, (O) 0.5 %, (■) 1 %, (□) 1.5 %, (∆) 2.5
ples with and without compatibilizer. Reproduced from ref. %, (▲) 5 %, and (▼) 10 %. Reproduced from ref. [88] by per-
[97] by permission of American Chemical Society. mission of The Society of Rheological, USA.

represents G' and G" of PA12‐based nanocomposites. This yield energy, Ey can be defined as follows:
The ø dependence of G' is presented in Figure 24, which

is rather well fitted by a quadratic equation: 
   
    ′

(6)
G' ∝ø (4)
where τis the shear stress.
According to them the appearance of nonlinearities in the Combining the equations (5) and (6), the yield energy,
viscoelastic response of nanocomposites at a strain γc Ey is found to be independent of the layered silicate vol-
(see Figure 25) is due to partial rupture of connectivity ume fraction, ø:
within the system and it is found that γc is inversely re-
lated to the volume fraction of layered silicate added   ∝ø0 (7)
(ø), i.e.
Now it will be very interesting if we compa0re this
γc ∝ø‐1 (5) yield energy data with the “cohesive” energy (Ec) data
for aqueous clay suspensions. Ramsay in 1986 [130,131]
The amount of elastic energy stored in the nano- and Sohm and Tadors in 1989 [132] respectively
composite at γc is the energy needed (termed as yield determined the ø dependence of Ec for laponite
1.8
energy, Ey) to break some crucial connections between dispersion (Ec ∝ø ) and for MMT suspensions (Ec ∝
3.1
silicate entities and allowing the system to yield, but not ø ) in the gel structure. However, the concept of
to disturb, as shown by the relatively smooth decrease of cohesive energy, although defined by the same equation
both viscoelastic moduli above γc (see Figure 25). as (6), still differs from the concept of yield energy
832 Suprakas Sinha Ray

Figure 24. Plateau storage modulus G' as a function of nano-


filler volume fraction. The solid line corresponds to Eq. (4).
The square of the correlation coefficient r2 is 0.95. Reproduced
from ref. [88] by permission of The Society of Rheological,
USA.

discussed here.
Indeed the term cohesive energy is appropriate for sys-
tems which extensively disrupt at a critical strain γc due
to rupture of all elastic links, as observed in clay gels
[130], whereas the term yield energy is more appropriate
for systems which yield above γc due to the rupture of
some crucial elastic links, as generally observed in the
case of layered silicate-based nanocomposites as dis-
cussed here. The difference in scalling behavior of the
elastic energy stored at γc means that the structure of the
Figure 25. (a) Storage modulus G' as a function of strain γ at
percolated network in layered silicate nanocomposites (●) 0 %, (O) 0.5 %, (■) 1 %, (□) 1.5 %, (∆) 2.5 %, (▲) 5
differs significantly from that of the gel in concentrated %, and (▼) 10 %. (b) Loss modulus G"(ω) as a function of
aqueous swelling layered silicate suspensions. Recently, strain γ at (●) 0 %, (O) 0.5 %, (■) 1 %, (□) 1.5 %, (∆) 2.5
Okamoto and coworkers completely disagree with these %, (▲) 5 %, and (▼) 10 %. Reproduced from ref. [88] by per-
interpretations. In the former structure, adding clay par- mission of The Society of Rheological, USA.
ticles seems to have no significant effect on the average
connectivity of the network as Ey ∝ø0, whereas increas- sample at a given frequency, G' and G" have the same or-
ing the layered silicate loading in the latter increases the der of magnitude showing that dissipative phenomena
number of network links as Ec depends strongly on ø. are as important as elastic phenomena in the behavior of
The peculiar feature of the scaling behavior of the Ey of nanocomposites. This point is all the more important to
the percolated network guggests that the entities respon- be emphasized since G" results are scarcely presented
sible for the yield properties of the nanocomposites are and discussed in the leature dealing with the rheology of
not individual layered silicate particles, rather, silicate layered silicate nanocomposites.
domains, whose characteristic size independent of lay- The ø dependence of the low-frequency plateau value
ered silicate loading, as mentioned by Solomon and cow- of G", presented in Figure 26, shows that
orkers [98] to interpret results from flow reversal
experiments. However, Mederic et al. proposed that the G' ∝ø 2.4 (8)
existence of a unique characteristic fc ∼ 0.2 Hz, making
a change in the power-law frequency dependency of G' This observation tends to show that, contrary to elastic
for all nanocomposites samples, is another macroscopic contributions, viscous contribution depend on the layered
time signature of these domains. silicate loading in a more significant way. It is also
Not only that, it should be noted here that for a given confirmed by the ø dependence of the frequency marking
Rheology of Polymer/Layered Silicate Nanocomposites 833

Figure 28. The steady-shear rheological behavior for a series of


delaminated nanocomposites of polydimethylsiloxane with di-
Figure 26. Plateau storage modulus G" as a function of nano- methyl ditallow montmorillonite at 25 °C. The steady-shear
filler volume fraction. The solid line corresponds to Eq. (8). viscosity shows an increase with respect to that of the pure-pol-
The square of the correlation coefficient r2 is 0.95. Reproduced ymer at low shear rates but still obeys Newtonian type behav-
from ref. [88] by permission of The Society of Rheological, ior, even at the highest silicate loadings examined. Reproduced
USA. from ref. [130] with permission of Springer-Verlag Berlin
Heidelberg.

composites based on poly(dimethyldiphenyl siloxane)


(PDS) and dimethyl ditallow ammonium modified
montmorillonite (2M2T-MMT). Figure 27 represents the
steady-shear rheological behavior of a series of
intercalated poly(dimethyldiphenyl siloxane) [it is a
copolymer of poly(dimethyl siloxane) and poly (diphenyl
siloxane)]/2M2T‐MMT nanocomposites with varying
silicate contents at 28 °C. The viscosity of nano-
composites was increased considerably and for a fixed
shear rate, increases monotonically with silicate loadings.
In the low-shear rate region ( < 2 s‐1) neat polymer
showed Newtonian behavior, whereas intercalated
nanocomposites showed shear thinning behavior and this
Figure 27. The steady-shear rheological behavior for a series of
shear thinning behavior was more prominent with higher
intercalated nanocomposites of poly(dimethyl 0.95-diphenyl silicate loading. At high shear rate both neat polymer and
0.05 siloxane) with layered silicate (dimethyl ditallow mont- nanocomposites displayed strong shear thinning be-
morillonite) at 25 °C. The silicate loading is varied and are not- havior. This observation indicates highly anisotropic
ed in the legend. Reproduced from ref. [130] with permission intercalated layered silicate particles alter the relaxation
of Springer‐Verlag Berlin Heidelberg. dynamics of the polymer, leading to the non-Newtonian
viscosity behavior of nanocomposites at all shear rates.
the change in the power-law dependency of the loss More interesting behavior was observed in the case of
modulus. The solid content dependence of the dissipative poly(dimethyl siloxane)/2M2T-MMT nanocomposites,
properties could be assigned to the multiscale origin of where layered silicates are fully delaminated in the
the viscous dissipation. However, viscous dissipation polymer matrix, the steady-shear viscosity behavior of
comes from many contributions such as from viscous nanocomposites systematically increased with silicate
dissipation due to relative motions of domains on a loading but still obeys the Newtonian behavior in the
mesoscopic length scale but also from relative motions of lower shearrate region. Figure 28 represents the steady-
exfoliated dispersed layered silicate particles on nano- or shear rheological behavior of a series of intercalated
micro-scopic length scales. poly(dimethyl siloxane)/2M2T‐MMT nanocomposites
with varying silicate contents at 28 °C This indicates
Steady Shear Rheology delaminated silicate layers have no effect on the
In 1996, Krishnamoorti and coworkers [77] first relaxation dynamics of poly(dimethyl siloxane). There-
described the steady shear rheological behavior of nano- fore, the rheological behavior of these nanocomposite
834 Suprakas Sinha Ray

Figure 29. shear rates dependent viscosity of pure PLA and Figure 30. Shear viscosity vs. shear rate of BAP/OMMT nano-
various PLACNs measured at 175 oC. composites for various OMMT loadings at 140 ºC. Reproduced
from ref. [124] with permission of American Chemical Society,
systems is similar to the rheological behavior of USA.
conventional filler filled systems.
Figure 29 represents the shear rates dependent viscosity Like dynamic viscoelastic properties, steady shear vis-
of pure PLA and various PLACNs measured at 175 oC. cosity of nanocomposites also systematically increased
While the pure PLA exhibits almost Newtonian behavior with increasing favorable interaction between the PBSA
at all shear rates, PLACNs exhibited a non-Newtonian matrix and organic modifier. For comparison purposes,
behavior. All PLACNs showed a very strong shear-thin- the small amplitude oscillatory shear viscosity is also re-
ning behavior at all measured shear rates and this behav- ported on the Figure 31. As can be expected, the steady
ior is analogous to the results obtained in the case of os- shear viscosity and the dynamic viscosity overlap very
cillatory shear measurements (see Figure 13). Additio- well for the pure PBSA for ω =  (Cox-Merz rule
nally, at very high shear rates, the steady shear vis- holds). However, for the nanocomposites, such a rule
cosities of PLACNs were comparable to that of pure fails. The steady shear viscosity becomes lower than the
PLA. These observations suggest that the silicate layers dynamic shear viscosity. At high shear rate, the nano-
are strongly oriented towards the flow direction at high composites loose their solid‐like behavior and a rather
shear rates. classical. Newtonian behavior is observed for the three
Like PLA-based nanocomposite systems, all PBSCNs nanocomposites. This might be due to the destruction of
show very strong shear thinning behavior at all shear the percolating network and the alignment of the silicate
rates whereas neat PBS exhibits Newtonian behavior. At layers in the direction of flow. The polymeric chains en-
very high shear rates the viscosities of the nano- trapped between these thin channels of silicate layers are
composites are comparable to that of pure PBS. highly oriented and flow easily due to the loss of the den-
The shear viscosity as a function of shear rate for sity of their entanglements with the neighbouring chains.
BAP/OMMT and BAP is presented in Figure 30 This loss of local entropy decreases the local frictions
[124,125]. Like previous systems, the viscosity of the and thus facilitates the chain movements in the direction
BAP/OMMT nanocomposites were also decreased with of flow. Such behavior is similar to that of layered poly-
increasing shear rate, however, at very low shear rates, mer blends or reinforced polymers that show a dramatic
the shear viscosity data exhibited a Newtonian plateau decrease in their viscosity when submitted to a strong
even for high OMMT content. With increasing shear shear flow [38-40].
rate, the nanocomposites exhibit higher degrees of shear- The increase in shear viscosity of the PLS-nano-
thinning behavior compared to the pure polymer. A composites was recently analyzed using a mean- field
similar behavior was also observed by Krishnamoorti theory [111]. The viscosity at high shear rates showed
and coworkers [78] in the case of delaminated PCL- more decreased values from the zero-shear viscosities
based nanocomposites with several silicate loadings. with increasing clay loading, and the values were
These observations suggest that the silicate layers are identical to those of virgin polymer. Although the exact
strongly oriented towards the flow direction at high shear mechanism which causes the shear-thinning behavior
rates, and that the shear thinning behavior observed at still is not clear, however, it can be deduced that the
high shear rates is dominated by that of pure polymer. orientation of the silicate layers under shear is the main
Rheology of Polymer/Layered Silicate Nanocomposites 835

Figure 31. Steady shear viscosity of neat PBSA and various Figure 32. Steady shear viscosityηas a function of shear rate 
nanocomposites as a function of shear rate. Reproduced from at (●) 0 %, (O) 0.25 %, (■) 0.5 %, (□) 0.75 %, and (∆) 1
ref. [123] with permission of Elsevier Science Ltd. %. Reproduced from ref. [88] by permission of The Society of
Rheological, USA.
cause. With increasing shear rate, the intercalated poly-
mer chain conformations change as the coils align the unfilled polymer melt, is a result of the ability of the
parallel to the flow [133]. Nevertheless, because of this two-dimensional silicate layers to be preferentially
shear-thinning property, the nanocomposites can be oriented by shear flow [135].
processed in the melt state using the conventional Solomon and coworkers [89] have performed flow re-
equipment available in a manufacturing line. versal experiments to examine changes in structure due
PLS nanocomposites [1] always exhibit significant de- to flow in PP based nanocomposites. The transient flow-
viations from the empirical Cox-Merz relation [134], reversal stress response exhibits a stress overshoot,
while all neat polymers obey the empirical relation, whose magnitude increases with increased quiescent
which stipulated that, for  = ω, the viscoelastic data waiting time between flow reversals. While a scaling of
obeys the relationship the stress overshoot by 1/c (c being the concentration)
leads to the development of a mastercurve with respect to
η( ) = |η*(ω)| (9) time, the start-up transient stress exhibits a simple strain
scaling. These results are interpreted in the context of the
Typically, the quiescent state linear dynamic oscillatory silicate layers being non-Brownian and the stress re-
η*(ω) exceeds the steady shear η( ), with the sponse being dictated by hydrodynamics alone.
discrepancy being largest at low shear rates. Figure 32 represents the steady shear apparent viscosity
Furthermore, a comparison of the steady shear η( ) with η, for PA12-based nanocomposite samples at low load-
the aligned state ηal*(ω) clearly demonstrates that η( ) ings, from 0 wt% up to 1 wt%. On the basis these data,
≥ηal*(ω), with the values at high shear rates being Aubry and coworkers tried to determine an intrinsic vis-
comparable. These results suggest that even at low shear cosity [η] and an interaction constant k, by fitting the
rates, the application of steady shear results in at least curve to the second‐order Einstein‐type equation:
some alignment of the silicate layers. The first normal 2
stress difference (N12=σ11 - σ22), a measure of the η0 = 1+[η]ø + k([η]ø) (10)
elasticity, for the disordered PS-PI based nanocomposites
when compared at the same shear stress (σ12), is The least-squares fit leads to [η] ∼100 and k ∼0.5,
independent of silicate loading and identical to that of the which confirms the same results obtained by Utracki and
unfilled polymer [135]. Those measurements were coworkers [136] for the similar organoclay based nano-
typically restricted to high shear rates (to obtain reliable composite systems.
values of N12), where the silicate layers are believed to be In 2003 Jeon and coworkers [110,111] proposed an
oriented and only relatively small changes are observed alternative method to calculate the [η] by fitting the
in the viscosity data. It has been suggested that the near relative viscosity versus volume fraction plot using a
independence of the recoverable strain (=N12/σ12) on the modified Krieger’s empirical equation, replacing the
silicate loading and the near equivalence to that of effective maximum packing volume fraction by the
836 Suprakas Sinha Ray

obtained from equation 5, given the uncertainty in the de-


termination of the critical volume fraction øvp. By
knowing the value of [η] and the aspect ratio p (ratio of
diameter and thickness) of the disk-like layered silicate
particles can be inferred:
1.47
[η] = 2.5 + 0.025 (1 + p ) (12)

Again by considering the value of [η] in between 65 to


100, p is shown ti lie between 200 and 260. This result is
about one-half of the average aspect ratio of individual
MMT particles [135]. This indicates exfoliation of lay-
ered silicates particles is not perfectly achieved, even at
very low layered silicate content.
The volume fraction of freely rotating disks with radius
2
r and thickness e (aspect ratio, p = 2r/e) isπr e/4 = 3/2p
3
(3πr ) times the volume fraction of equivalent of hard
spheres. Assuming a hard sphere random close packing
∼64 %, they calculated the maximum packing volume
fraction of disk-like particles (ømcalc) with an aspect ratio
200 and 260 respectively and this value is about 0.48 %
and 0.38 % respectively. Further, assuming a critical vol-
ume fraction of spherical particles at percolation to be ∼
30 %, the freely rotating disk-like layered silicate par-
ticles are expected to be percolated at ømcalc ∼0.23 %
and 0.18 % respectively. All of these calculations are ap-
proximate, insofar as they do not take into account the
presence of large aggregates. Present author believe that
the real close packing volume fraction and percolation
threshold most likely occur at higher volume fractions,
possibly explaining the relatively “high” value of øvp.
Figure 33. (a) Relative Newtonian viscosity as a function of Elongation Flow Rheology
nanofiller volume fraction, ø. The solid line corresponds to
Okamoto and coworkers [137] first conducted elonga-
Eq. (10), with [η] = 100 and k = 0.5. The square of the corre-
lation coefficient r2 in 0.83. (b) Relative Newtonian viscosity as
tion tests of polypropylene based nanocomposite
a function of normalized nanofiller volume fraction. The line (PPCN4) in the molten state at constant Hencky strain

corresponds to Eq. (11), with øvp = 1 % and [η] = 65. The rate, ε 0 using elongation flow rheometer. Figure 34
2
square correlation coefficient r = 0.85. Reproduced from ref. shows double-logarithmic plots of the transient elonga-

[88] by permission of The Society of Rheological, USA. tion viscosity ηE(ε 0 ; t) versus time t, observed for
PPCN4 (4 wt% of MMT) at 150 oC with different ε ‧
0 val-
‐1
viscosity percolation threshold øvp as ues ranging from 0.001 to 1.0 s . The solid line on the
‧ 
curve ε 0 represents 3η0(  ; t), the 3-fold shear viscosity
−[η ]φvp
‐1
⎡ of PPCN4, measured with a constant  of 0.001 s at 150
φ ⎤ o ‧
C. At short times, ηE(ε0 ; t) gradually increases with
η 0r = ⎢1 − ⎥
⎣⎢ φ vp ⎦⎥ (11) ‧
time, but is independent of ε 0 ; this is generally called the

linear region of the ηE (ε0 ; t) curve. After a certain
On the basis of Figure 32, they considered øvp is equal time tηE, generally called the up-rising time (marked
to 1 wt% is the viscosity percolation threshold corre- with the upward arrows), there is a rapid upward devia-

sponds to the critical volume fraction at which the vis- tion of ηE (ε 0 ; t) from the curves of the linear region.
cosity tends to diverge, showing no Newtonian plateau at On the other hand, the shear viscosity curve shows two
low shear rates. Figure 33b shows that the relative vis- distinctive features: first, the extended Trouton rule is not
cosity versus volume fraction plot can be well fitted by valid in this case and second, the shear viscosity of
Eq. (11); [η] obtained from the beast least-squares fit is PPCN4 increases continuously with time within the ex-
∼ 65. This result is in good agreement with the result perimental time span. This time-dependent thickening
Rheology of Polymer/Layered Silicate Nanocomposites 837

Figure 34. Time dependence of elongation viscosity for


PP/clay (4) (PPCN4) melt at 150 ºC. The upward arrows in- Figure 36. Double‐logarithmic plots of the transient elongation

dicate up-rising time for different strain rate. The solid line viscosity ηE(ε 0 ; t) versus time t, observed for PLACN5 at 170
o
shows three times the shear viscosity, taken at a low shear rate C with different ε&0 values ranging from 0.01 to 1.0 s‐1.
‐1
of 0.001 s on a cone plate. Reproduced from from ref. [137]
with permission of American Chemical Society, USA. differences in the time-dependent responses reflect
differences in the shear flow-induced versus elongation-
behavior, as mentioned before, is called rheopexy. These induced internal structure formation in the case of

Figure 35. TEM micrographs showing PPCN4 elongated at 150 ºC with (a) ε0 = 1.0 s‐1 up to ε0 = 1.3 s‐1 (λ = 3.7) and (b) ε0 =
‐1 ‐1
0.001 s up to ε0 = 0.5 s (λ = 1.7); respectively. Upper pictures are in the x-y plane, and lower are in x-z plane along the stretch-
ing direction. Reproduced from from ref. [138] by permission of American Chemical Society, USA.
838 Suprakas Sinha Ray

hardening behavior for PLACN5. In the early stage, ηE



gradually increases with t but almost independent of ε 0.
This is generally calling the linear region of the viscosity
curve. After a certain time, tηE which is the up-rising
time (marked with the upward arrows in figure), was

strongly dependent on ε 0 , and a rapid upward deviation
of ηE from the curves of the linear region was observed.
On the other hand, Sinha Ray and coworkers [118] tried
to measure the elongational viscosity of pure PLA but
they failed to do that accurately. Low viscosity of pure
PLA may be the main reason. However, they confirmed
that neither strain‐induced hardening in elongation nor
rheopexy in shear flow took place in the case of pure
PLA having the same molecular weights and poly-
dispersity with that of PLACN3 [118].
Figure 37. Hencky strain rate dependence of the up-rising Like polypropylene/OMLS systems, the extended Trou-
‧ o
Hencky strain (εηE) = ε0 × tηE taken for PLACN3 at 170 C. ton rule, 3η0 ( ; t) ≅ ηE (ε ‧
0 ; t), does not hold for
PLACN5 melt, as opposed to the melt of pure polymers
PPCN4 in the molten state. The same experiments [138]. These results indicate that in the case of PLACN5,
conducted with a PP-MA matrix without clay did not the flow induced internal structural changes also
exhibit any strain-induced hardening or rheopexy occurred in elongation flow, but the changes were quite
behavior in the molten state. different from shear flow. The strong rheopexy observed
Figure 35 represents typical TEM photographs of the in shear measurements for PLACN5 at very slow shear
center portions of the recovered samples after elongation. rate reflected a fact that the shear-induced structural
The x- and y-axes of the elongated specimen correspond change involved a process with an extremely long
to directions parallel and perpendicular to the stretching relaxation time.
direction, respectively. A converging flow is applied to As to the elongation‐induced structure development,
the thickness direction (y- and z-axis) with stretching if Figure 37 represents the Hencky strain rate dependence
the assumption of an affine deformation without volume of the up‐rising Hencky strain (εηE) =ε ‧
0 × tηE taken for
change is valid. Interestingly, for the specimen elongated o
PLACN3 at 170 C. The εηE values increased systemati-
with a high strain rate, there is perpendicular alignment ‧ ‧
cally with the ε 0 . The lower the value of ε0 , the smaller
of the silicate layers (edges) along the stretching
is the value of εηE. This tendency probably corresponds
direction (x-axis) in the x-y plane. For the x-z plane, the
to the rheopexy of PLACN5 under slow shear flow.
silicate layers (edges) disperse into the PP-MA matrix
along the z-axis direction rather than randomly, but these
faces cannot be seen in this plane. On the basis of experi- Conclusions
mental results and two-directional TEM observations, the
authors concluded that the formation of a “house-of- Rheological properties of several polymer nano-
cards” like structure was observed under slow elongatio- composites with layered silicates have been discussed
nal flow. Details regarding the data collection and herein. On a global scale, the linear viscoelastic behavior
explanations are presented in reference [137]. of the polymer chains in the nanocomposites, as detected
Sinha Ray and coworkers [118] conducted elongation by conventional rheometry, is dramatically altered after
tests of PLACN5 (PLA nanocomposites prepared with 5 nanocomposites formation. Linear rheological material
wt% of C18MMT) in the molten state at constant functions showed that the nanocomposites are charac-

Hencky strain rate, ε 0 using elongation flow rheometry terized by a solid-like behavior as revealed by the pres-
[138]. On each run of the elongation test, samples of 60 ence of a low-frequency plateau on storage modulus. The
3
× 7 × 1 mm size were annealed at a predetermined tem- magnitude of such plateau or pseudo-plateau is found to
perature for 3 min before starting the run in the rhe- strongly depend on the level of interactions between the
ometer, and uniaxial elongation experiments were con- polymer matrix and the surfactant of the modified lay-
‧ ‐1
ducted at various ε 0 ranging from 0.01 to 1 s . Figure 36 ered silicate. Such specific interactions favour the state of
shows double-logarithmic plots of the transient elonga- dispersion-distribution of the layered silicate lamellae

tion viscosity ηE(ε 0 ; t) versus time t, observed for within the polymeric matrix that lead to the formation of
o ‧
PLACN5 at 170 C with different ε 0 values ranging a long range or short range percolating network.
‐1
from 0.01 to 1.0 s . Figure shows a strong strain-induced However, such a network does not seem very stable
Rheology of Polymer/Layered Silicate Nanocomposites 839

vis-a-vis of steady shear flow. In fact steady shear flow References


destroys such connectivity and imparts the material with
a rather Newtonian behavior in the high-shear rate 1. (a) S. Sinha Ray and M. Okamoto, Prog. Polym.
region. For the moment it is not clear whether or not such Sci., 28, 1539 (2003); (b) M. Okamoto, J. Ind. Eng.
percolating network can be maintained when the nano- Chem., 10, 1156 (2004); (c) J. W. Kim, S. G. Kim,
composites are reprocessed in machines such as ex- H. J. Choi, and M. S. Jhon, Macromol. Rapid
truders and injection moldings, where the flow has a Commun., 20, 450 (1999); (d) C. H. Hong, Y. B.
complex pattern combining shear, elongational, hyper- Lee, J. W. Bae, J. Y. Jho, B. U. Nam, and T. W.
bolic and eventually chaotic flows. However, some of Hwang, J. Ind. Eng. Chem., 11, 76 (2005); (e) J. H.
these systems show close analogies to other intrinsically Sung, M. S. Cho, H. J. Choi, and M. S. Jhon, J. Ind.
anisotropic materials such as block copolymers and Eng. Chem., 10, 1217 (2004).
smectic liquid crystalline polymers and provide model 2. J. A. Rojas-Chapana and M. Giersig, J. Nanosci.
systems to understand the dynamics of polymer brushes. Nanotechnol., 6, 316 (2006).
Under uniaxial elongational flow, the nanocomposites 3. R. J. Nussbaumer, W. R.Caseri, and P. Smith, J.
generally exhibit very high viscosity and a tendency to Nanosci. Nanotechnol., 6, 459 (2006).
strong strain‐induced hardening, which author believes 4. C. N. R. Rao, F. L. Deepak, G. Gundish, and A.
originate from the perpendicular alignment of the silicate Govindaraj, Prog. Solid State Chem., 31, 5 (2003).
layers towards the stretching direction. 5. M. S. Dresselhaust and G. Dresselhaust, Adv. Phys.,
Although a significant amount of work has already been 30, 139 (1981).
reported on rheological properties of PLA-nanocom- 6. H. Shioyama, Synth. Met., 114, 1 (2000).
posites, much research still remains in order to under- 7. F. M. Uhl and C. A. Wilkie, Polym. Degrad. Stab.,
stand the complex structure-property relationships in var- 76, 111 (2002).
ious PLA-nanocomposites. 8. R. Hiroi, S. Sinha Ray, M. Okamoto, and T. Shiroi,
Macromol. Rapid Commun., 25, 1359 (2004).
9. (a) A. K. Mohanty, A. Wibowo, M. Misra, and L. T.
Abbreviations Drzal, Polym. Eng. Sci., 43, 1151 (2003); (b) J. S.
Choi, S. T. Lim, H. J. Choi, S. M. Hong, A. K.
BPA, biodegradable aliphatic polyester; CEC, cation Mohanty, L. T. Drzal, M. Misra, and A. C. Wibowo,
exchange capacity; Deff, effective diffusion coefficient; Macromol. Symp., 224, 297 (2005).
EMA, ethylene-co-methyl acrylate; EVA, ethylene-co- 10. M. Joshi and B. S. Butola, J. Macromol. Sci. Part C-
vinyl acetate; HMW, high molecular weight; IANS, in- Polym. Rev., 44, 389 (2004).
termediate-angle neutron scattering; LCST, lower critical 11. N. Sinha, J. Ma, and J. T. W. Yeow, J. Nanosci.
solution temperature; LMW, low molecular weight; MD, Nanotechnol., 6, 289 (2006).
molecular dynamics; MMT, montmorillonite; 2M2T- 12. X. Cui, M. H. Engelhard, and Y. Lin, J. Nanosci.
MMT, dimethyl ditallow ammonium modified montmor- Nanotechnol., 5, 547 (2006).
illonite; MMW, medium molecular weight; N6, nylon 6; 13. W. Zhang, J. Suhr, and N. Koratkar, J. Nanosci.
NMR, nuclear magnetic resonance; OMLS, organically Nanotechnol., 6, 960 (2006).
modified layered silicate; PA12, polyamide 12; PBS, 14. K. Awasthi, A. Srivastava, and O. N. Srivastava, J.
poly(butylene succinate); PBSA, poly[(butylene succi- Nanosci. Nanotechnol., 5, 1616 (2005).
nate)-co-adipate]; PCL, poly(ε-caprolacton); PEO, poly 15. H. Miyagawa, M. Misra, and A. K. Mohanty, J.
(ethylene oxide); PDS, poly(dimethyl siloxane); PLA, Nanosci. Nanotechnol., 5, 1593 (2005).
polylactide; PLS, polymer layered silicate; PP, poly- 16. X. L. Xie, Y. W. Mai, and X. P. Zhou, Mater. Sci.
propylene; PPCH, polypropylene hybride; PS, poly- Eng., R49, 89 (2005).
styrene; QA+, quaternary ammonium cation; Rg, radius of 17. (a) R. Andrews and M. C. Weisenberger, Curr.
gyration; RIS, rotational-isomeric-state; SE, Stokes- Opinion Solid State Mater. Sci., 8, 31 (2004); (b) Y.
Einstein; SFM, synthetic fluorine mica; TEM, trans- S. Lee and T. H. Cho, J. Ind. Eng. Chem., 10, 631
mission electron microscopy, XRD, x-ray diffraction, (2004); (c) S. J. Park, M. S. Cho, S. T. Lim, H. J.
POSS, polyhedral oligomeric silsesquioxanes; PVA, poly Choi, and M. S. Jhon, Macromol. Rapid Commun.,
(vinyl alcohol); SCF; self-consistent field; SANS, small- 24, 1070 (2003).
angle neutron scattering. 18. (a) S. Sinha Ray and M. Bousmina, Prog. Mater.
Sci., 50, 962 (2005); (b) C. H. Hong, Y. B. Lee, J.
W. Bae, J. Y. Jho, B. U. Nam, and T. W. Hwang, J.
Ind. Eng. Chem., 11, 293 (2005); (c) S. J. Park, K.
Li, and S. K. Hong, J. Ind. Eng. Chem., 11, 561
840 Suprakas Sinha Ray

(2005). 42. D. V. Kuznetsov and A. C. Balazs, J. Chem. Phys.,


19. M. Biswas and S. Sinha Ray, Adv. Polym. Sci., 155, 112, 4365 (2000).
167 (2001). 43. Y. Lyatskaya and A. C. Balazs, Macromolecules,
20. S. Sinha Ray and M. Bousmina, in Nanostructured 31, 6676 (1998).
Polymers and Their Applications, H. S. Nalwa ed., 44. V. V. Ginzburg and A. C. Balazs, Macromolecules,
Ch. 23, American Scientific Publishers, Los Angeles 32, 5681 (1999).
(2006). 45. A. J. Lui and G. Fredrickson, Macromolecules, 26,
21. K. Yano, A. Usuki, A. Okada, T. Kurauchi, and O. 2817 (1993).
Kamigaito, J. Polym. Sci. Part A: Polym. Chem., 31, 46. H. W. Chiu and T. Kyu, J. Chem. Phys., 103, 7471
2493 (1993). (1995).
22. R. K. Bharadwaj, Macromolecules, 34, 1989 (2001). 47. H. W. Chiu and T. Kyu, Phys. Rev. E, 53, 3618
23. S. Sinha Ray, K. Yamada, M. Okamoto, and K. (1996).
Ueda, Nano Letts., 2, 1093 (2002). 48. H. W. Chiu and T. Kyu, J. Chem. Phys., 107, 6859
24. J. W. Gilman, Appl. Clay. Sci., 15, 31 (1999). (1997).
25. J. W. Gilman, C. L. Jackson, A. B. Morgan, Jr. R. 49. P. Tarazona, Phys. Rev., A 31, 2672 (1985).
Harris, E. Manisa, E. P. Giannelis, M. Wuthenow, D. 50. R. A. Vaia, K. D. Jandt, E. J. Kramer, and E. P.
Hilton, and S. H. Phillips, Chem. Mater., 12, 1866 Giannelis, Macromolecules, 28, 8080 (1995).
(2000). 51. B. Bhushan, J. N. Israelachvili, and U. Landman,
26. M. Bartholmai and B. Schertel, Polym. Adv. Tecnol., Nature (London), 374, 607 (1995).
15, 355 (2004). 52. R. G. Horn and J. N. Israelachvili, Macromolecules,
27. A. B. Morgan, Polym. Adv. Technol., 17, 206 (2006) 21, 2836 (1988).
28. S. Sinha Ray, K. Yamada, M. Okamoto, and K. 53. (a) H. K. Christenson, D. W. R. Gruen, R. G. Horn,
Ueda, Macromol. Mater. Eng., 288, 203 (2003). and J. N. Israelachvili, J. Chem. Phys., 87, 1834
29. S. Sinha Ray and M. Okamoto, Macromol. apid. (1987); (b) R. N. Kono, S. Izumisawa, M. S. Jhon,
Commun., 24, 814 (2003). C. A. Kim, and H. J. Choi, IEEE Trans. Magn., 37,
30. S. Sinha Ray, K. Yamada, M. Okamoto, and K. 1827 (2001).
Ueda, Polymer, 44, 857 (2003). 54. G. Reiter, A. L. Demirel, and S. Granick, Science,
31. A. Balazs, V. V. Ginzburg, Y. Lyatskaya, C. Singh, 263, 1741 (1994).
and E. Zhulina. In Polymer-Clay Nanocomposites, 55. A. L. Demirel and S. Granick, Phys. Rev. Lett., 77,
T. J. Pinnavaia and G. W. Bell ed., Ch. 14, Wiley, 2261 (1996).
New York (2001). 56. P. A. Thompson, M. O. Robbins, and G. S. Grest.
32. E. Manias and V. Kuppa. In Polymer Nanocompo- Isr. J. Chem., 35, 93 (1995).
sites, R. A. Vaia and R. Krishnamoorti eds., ACS 57. E. Manias, G. Hadziioannou, I. Bitsanis, and G. Ten
Symposium series, vol. 804, Oxford University Brinke, Europhys. Lett., 24, 99 (1993).
Press, pp. 193∼207 (2002). 58. E. Manias, I. Bitsanis, G. Hadziioannou, and G. Ten
33. N. Güven. In Hydrous Phyllosilicates. S. W. Bailey, Brinke, Europhys. Lett., 33, 371 (1996).
Ed., Mineralogical Society of America, Washington 59. E. Manias, A. Subbotin, G. Hadziioannou, and G.
DC, pp.497∼560 (1988). Ten Brinke, Mol. Phys., 85, 1017 (1995).
34. A. M. Somoza and P. Tarazona, J. Chem. Phys., 91, 60. A. R. C. Baljon and M. O. Robbins, Science, 271,
517 (1989). 482 (1996).
nd
35. H. van Olphen, Clay colloid chemistry, 2 ed., 61. A. R. C. Baljon and M. O. Robbins, MRS Bull., 22,
Johan Wiley, New York (1977). 22 (1997)
36. P. Aranda and E. Ruiz-Hitzky, Chem. Mater., 4, 62. S. A. Gupta, H. D. Cochran, and P. T. Cummings, J.
1395 (1992). Chem. Phys., 107, 10316 (1997).
37. D. J. Greenland, J. Colloid Sci., 18, 647 (1963). 63. M. J. Stevens, M. Mondollo, G. S. Grest, S. T. Cui,
38. R. A. Vaia and E. P. Giannelis, Macromolecules, 30, H. D. Crochan, and P. T. Cummings, J. Chem.
7990 (1997). Phys., 106, 7303 (1997).
39. R. A. Vaia and E. P. Giannelis, Macromolecules, 30, 64. I. A. Bitsanis and C. Pan, J. Chem. Phys., 99, 5520
8000 (1997). (1993).
40. A. C. Balazs, C. Singh, and E. Zhulina, Macromole- 65. R. K. Ballamudi and I. A. Bitsanis, J. Chem. Phys.,
cules, 31, 8370 (1998). 105, 7774 (1996)
41. G. Fleer, M. A. Cohen‐Stuart, J. M. H. M. 66. P. A. Thompson and S. M. Troian, Nature
Scheutjens, and T. V. Cosgrove. Polymers at (London), 389, 360 (1997).
Interfaces, Chaoman and Hall, London (1993). 67. P. A. Thompson and M. O. Robbins, Phys. Rev., A
Rheology of Polymer/Layered Silicate Nanocomposites 841

41, 6830 (1990). 92. M. Sepehr, L. A. Utracki, X. Zheng, and C. A.


68. R. F. Cracknell, D. Nicholson, and K. E. Gubbins, J. Wilkie, Polymer, 46, 11569 (2005).
Chem. Soc., Faraday Trans., 91, 1377 (1995). 93. L. Xu, S. Reeder, M. Thopasidharan, J. Ren, D. A.
69. R. F. Cracknell, D. Nicholson, and N. Quirke, Phys. Shipp, and R. Krishnamoorti, Nanotechnology, 16,
Rev. Lett., 74, 2463 (1995). S514 (2005).
70. D. Nicholson, R. F. Cracknell, and N. Quirke, 94. Y. Zhong, Z. Zhu, and S. Q. Wang, Polymer, 46,
Langmuir, 12, 4050 (1996). 3006 (2005).
71. E. J. Maginn, A. T. Bell, and D. N. Theodorou, J. 95. J. Zhao, A. B. Morgan, and J. D. Harris, Polymer,
Phys. Chem., 97, 4173 (1993). 46, 8641 (2005).
72. J. Y. Lee, A. R. C. Baljon, R. F. Loring, and A. Z. 96. O. Meincke, B. Hoffmann, C. Dietrich, and C.
Panagiotopoulos, J. Chem. Phys., 109, 10321 Friedrich, Macromol. Chem. Phys., 204, 823 (2003).
(1998). 97. G. Galgali, C. Ramesh, and A. Lele, Macromole-
73. K. Kremer and G. S. Grest, J. Chem. Phys., 92, 5057 cules, 34, 176 (2001).
(1990). 98. M. J. Solomon, A. S. Almusallam, K. F. Seefeldt, A.
74. V. Tries, W. Paul, J. Baschnagel, and K. Binder, J. Somwangthanarij, and P. Varadan, Macromolecules,
Chem. Phys., 106, 738 (1997). 34, 1864 (2001).
75. A. R. C. Baljon, J. Y. Lee, and A. F. Loring, J. 99. J. Li, C. Zhou, G. Wang, and D. Zhao, J. Appl.
Chem. Phys., 111, 9068 (1999). Polym. Sci., 89, 3609 (2003).
76. E. Manias, H. Chen, R. Krishnamoorti, J. Genzer, E. 100. L. Jian, C. Zhou, W. Gang, Y. Wei, T. Ying, and L.
J. Kramer, and E. P. Giannelis, Macromolecules, 33, Qing, Polym. Comp., 24, 323 (2003).
7955 (2000). 101. S. Y. Gu, J. Ren, and Q. F. Wang, J. Appl. Polym.
77. S. Sinha Ray and M. Bousmina, “Handbook of Sci., 91, 2427 (2004).
Biodegradable Polymeric Materials and Their 102. W. Lertwimolnun and B. Vergnes, Polymer, 46,
Applications”, Vol. 1, S. Mallapragada, N. Narasim- 3462 (2005).
han, Eds., American Scientific Publishers, California 103. P. Mederic, T. Razafinimaro, T. Aury, M. Moan,
2005, chapter 9. and M. H. Klopffer, Macromol. Symp., 221, 75
78. R. Krisnhamoorti and E. P. Giannelis, Macromole- (2005).
cules, 30, 4097 (1997). 104. M. Gelfer, H. H. Song, L. Liu, C. Avila-Orta, L.
79. J. Ren, A. S. Silva, and R. krishnamoorti, Yang, M. Si, B. Hsiao, B. Chu, M. Rafailovich, and
Macromolecules, 33, 3739 (2000). A. H. Tsou. Polym. Eng. Sci., 42, 1841 (2002).
80. K. Lakdawala and R. Salovey, Polym. Eng. Sci., 27, 105. V. Pasanovic-Zujo, R. K. Gupta, and S. N.
1035 (2004). Bhattacharyya, Rheol. Acta, 43, 99 (2004).
81. Ji-Zhao Liang, Polym. Int., 51, 1473 (2002). 106. M. Y. Gelfer, C. Buger, B. Chu, B. S. Hsiao, A. D.
82. D. R. Saini, A. V. Shenoy, V. M. Nadkarni, Polym. Drozdov, M. Si, M. Rafailovich, B. B. Sauer, and J.
Compos., 7, 193 (2004). W. Gilman, Macromolecules, 38, 3765 (2005).
83. M. L. Williams, R. F. Landel, and J. D. Ferry, J. Am. 107. K. M. Lee and C. D. Han, Macromolecules, 36,
Chem. Soc., 77, 3701 (1955). 7165 (2003).
84. T. D. Fornes, P. J. Yoon, H. Keskkula, and D. R. 108. A. J. Hsieh, P. Moy, F. L. Beyer, P. Madison, E.
Paul, Polymer, 42, 9929 (2001). Napadensky, J. Ren, and R. Krishnamoorti, Polym.
85. J. S. Choi, S. T. Lim, H. J. Choi, A. Pozsgay, L. Eng. Sci., 44, 825 (2004).
Szazdi, and B. Pukanszky, J. Mater. Sci., 41, 1843 109. K. M. Lee and C. D. Han, Polymer, 44, 4573
(2006). (2003).
86. R. Ahmad and S. J. Lee, Adv. Mater., 38, 19 (2006). 110. (a) H. J. Choi, S. G. Kim, Y. H. Hyun, and M. S.
87. W. Shao, Q. Wang, F. Wang, and Y. Chen, J. Polym. Jhon, Macromol. Rapid. Commun., 22, 320 (2001);
Sci. Part B: Polym. Phys., 44, 249 (2006). (b) T. H. Kim, L. W. Jang, D. C. Lee, H. J. Choi,
88. T. Aubry, T. Razafinimaro, and P. Mdric, J. Rheol., and M. S. Jhon, Macromol. Rapid Commun., 23,
49, 425 (2005). 191 (2002); (c) H. B. Kim, C. H. Lee, J. S. Choi, B.
89. B. Hoffman, J. Kressler, G. Stoppelmann, C. J. Park, S. T. Lim, and H. J. Choi, J. Ind. Eng.
Friedrich, G. M. Kim, Colloid Polym. Sci., 278, 629 Chem., 11, 769 (2005).
(2000). 111. Y. H. Hyun, S. T. Lim, H. J. Choi, and M. S. Jhon,
90. K. A. Ravishankar and I. L. Arkady, Rheol. Acta., Macromolecules, 34, 8084 (2001).
43, 283 (2004). 112. Z. Shen, G. P. Simon, and Y. B. Cheng, Polymer,
91. G. M. Russo, G. P. Simmon, and L. Incarnato, 43, 5011 (2002).
Macromolecules, 39, 3855 (2006). 113. B. Lepoittevin, M. Devalckenaere, N. Pantoustier,
842 Suprakas Sinha Ray

M. Alexandre, D. Kubies, C. Calberg, R. Jerome, 41, 679 (2005).


and P. Dubois, Polymer, 43, 4017 (2002). 126. W. Huang and C. D. Han, Macromolecules, 39, 257
114. S. Sinha Ray, Dispersion characteristics and vis- (2006).
coelastic properties of Poly(ε-caprolactone) nano- 127. J. Bandyopadhyay, S. Sinha Ray, and M. Bousmina,
composites. Unpublished results Polymer, submitted (2006).
115. A. Sanchez-Solis, I. Romero‐Ibarra, M. R. Estrada, 128. W. S. Chow, Z. A. M. Ishak, and J. Karger-Kocsis,
F. Calderas, and O. Manero, Polym. Eng. Sci., 44, Macromol. Mater. Eng., 290, 122 (2005).
1094 (2004). 129. S. Sinha Ray, M. Bousmina, and A. Moozouz,
116. S. Sinha Ray, K. Okamoto, K. Yamada, and M. Polym. Eng. Sci., 46, 1121 (2006).
Okamoto, Macromolecules, 35, 3104 (2002). 130. J. D. F. Ramsay, J. Colloid Interface Sci., 109, 441
117. S. Sinha Ray, K. Yamada, M. Okamoto, and K. (1986).
Ueda, Polymer, 44, 857 (2003). 131. R. G. Avery and J. D. F. Ramsay, J. Colloid
118. S. Sinha Ray and M. Okamoto, Macromol. Mater. Interface Sci., 109, 448 (1986).
Eng., 288, 936 (2003). 132. R. Sohm and Th. F. Tadros, J. Colloid Interface
119. K. Okamoto, S. Sinha Ray, and M. Okamoto, J. Sci., 132, 62 (1989).
Polym. Sci. Part B: Polym. Phys., 41, 3160 (2003). 133. R. Krishnamoorti, R. A. Vaia, and E. P. Giannelis,
120. S. Sinha Ray, K. okamoto, and M. Okamoto, Chem. Mater., 8, 1728 (1996).
Macromolecules, 36, 2355 (2003). 134. W. P. Cox and E. H. Merz, Correlation of dynamic
121. S. Sinha Ray, K. Okamoto, and M. Okamoto, J. and steady flow viscosities. J. Polym. Sci., 28, 619
Appl. Polyl. Sci., 102, 777 (2006). (1958).
122. S. Sinha Ray, M. Bousmina, and K. Okamoto, 135. E. P. Giannelis, R. Krishnamoorti, and E. Manias,
Macromol. Mater. Eng., 290, 759 (2005). Adv. Polym. Sci., 138, 107 (1999).
123. S. Sinha Ray and M. Bousmina, Polymer, 46, 136. L. A. Utracki and J. Lyngaae-Jorgensen, Rheol.
12430 (2005). Acta, 41, 394 (2002).
124. S. T. Lim, Y. H. Hyum, H. J. Choi, and M. S. Jhon, 137. M. Okamoto, P. H. Nam, P. Maiti, T. Kotaka, N.
Chem. Mater., 14, 1839 (2002). Hasegawa, and A. Usuki, Nano. Lett., 1, 295
125. (a) S. T. Lim, C. H. Lee, H. J. Choi, and M. S. Jhon, (2001).
J. Polym. Sci. Part B: Polym. Phys., 41, 2052 138. T. Kotaka, A. Kojima, and M. Okamoto, Rheol.
(2003); (b) H. B. Kim, J. S. Choi, C. H. Lee, S. T. Acta, 36, 646 (1997).
Lim, M. S. Jhon, and H. J. Choi, Euro. Polym. J.,

You might also like