You are on page 1of 385

Advanced Structured Materials

For further volumes:


http://www.springer.com/series/8611
Vikas Mittal  Jin Kuk Kim  Kaushik Pal
Editors

Recent Advances in Elastomeric


Nanocomposites

123
Editors Prof. Dr. Jin Kuk Kim
Dr. Vikas Mittal Department of Polymer Science and
Berner Weg 26 Engineering
67069 Ludwigshafen School of Nano and Advanced Materials
Germany Engineering
e-mail: vikas.mittal@chem.ethz.ch Gyeongsang National University
Gazwa-Dong 900, Jinju, Gyeongnam
Present Address 600-701
Dr. Vikas Mittal Korea, Republic of (South Korea)
Chemical Engineering Program e-mail: rubber@nongae.gsnu.ac.kr
The Petroleum Institute
Abu Dhabi Dr. Kaushik Pal
2533 Department of Polymer Science and
UAE Engineering
e-mail: vikas.mittal@chem.ethz.ch School of Nano and Advanced Materials
Engineering
Gyeongsang National University
Gazwa-Dong 900, Jinju, Gyeongnam
600-701
Korea, Republic of (South Korea)
e-mail: pl_kshk@yahoo.co.in

ISSN 1869-8433 e-ISSN 1869-8441

ISBN 978-3-642-15786-8 e-ISBN 978-3-642-15787-5

DOI 10.1007/978-3-642-15787-5

Springer Heidelberg Dordrecht London New York

Ó Springer-Verlag Berlin Heidelberg 2011

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcast-
ing, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this
publication or parts thereof is permitted only under the provisions of the German Copyright Law of
September 9, 1965, in its current version, and permission for use must always be obtained from
Springer. Violations are liable to prosecution under the German Copyright Law.

The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.

Cover design: WMXDesign GmbH, Heidelberg

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

Elastomers are a very important class of polymer materials and the generation of
their nanocomposites by the incorporation of nano-fillers has led to the enhance-
ment of their properties significantly and hence expansion of their application
potential. The book aims to specifically review the recent progresses in the syn-
thesis, processing as well as applications of the elastomeric nanocomposites. The
contents of the book are classified into three broad categories: first one dealing
with introduction and preparation of the elastomeric nanocomposites, the second
one focusing on the characterization and properties of the formed composites,
whereas the third one describing the application potential of these materials.
Chapter 1 describes the role of different nanoparticles in reinforcing elastomers.
Homogenous dispersion of the filler and subsequent microstructure development in
the composites have been focused upon. Chapter 2 explains the important syn-
thesis methodology of in situ preparation of elastomeric nanocomposites. Other
synthesis methodologies have also been described in brief. Chapter 3 focuses on
the relaxation phenomena in elastomeric nanocomposites. The results are pre-
sented for both non-polar and polar polymer matrices. Modeling and simulation of
nanocomposite processing have been described in Chap. 4 using molecular
dynamics and Monte Carlo methodologies. Chapter 5 shows the deformation
induced structural changes in elastomeric nanocomposites. Polymer matrices
reinforced with various fillers like clay, nanofibers, nanotubes, carbon black, etc.
have been considered. Thermally stable and flame retardant elastomeric nano-
composites are the focus of Chap. 6 whereas recycling of the elastomeric nano-
composites has been demonstrated in Chap. 7. In the applications section, Chap. 8
describes the considerations for the use of elastomeric nanocomposites in tyre
applications. Subsequently, synthesis of nanocomposites suitable for use in tyre
applications has been reported. Chapter 9 shows the use of elastomeric nano-
composites made from polyurethane and epoxy matrices for potential use in
packaging applications. Elastomeric nanocomposites suitable for biomedical
applications have been described in Chap. 11. Considerations of elastomeric
nanocomposite systems for use in aerospace applications have been focused in
Chap. 12. Chapter 13 describes the friction and wear of polymer nanocomposites

v
vi Preface

containing clay and nanotubes as fillers thus providing the potential of their use in
such applications.
The editors would also like to express heartfelt thanks to Springer, Germany for
their acceptance to publish the book.

Dr. V. Mittal
Prof. Dr. J.K. Kim
Dr. K. Pal
Contents

Part I Introduction & Preparation

Role of Different Nanoparticles in Elastomeric Nanocomposites . . . . . 3


Jin Kuk Kim, Kaushik Pal and V. Sridhar

In Situ Synthesis of Rubber Nanocomposites . . . . . . . . . . . . . . . . . . . 57


Massimo Messori

Part II Characterization & Properties

Relaxation Phenomena in Elastomeric Nanocomposites. . . . . . . . . . . . 89


G. C. Psarras and K. G. Gatos

Modeling and Simulation of Polymeric Nanocomposite Processing . . . 119


Teik-Cheng Lim

Deformation-Induced Structure Changes in Elastomeric


Nanocomposites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Shigeyuki Toki and Benjamin S. Hsiao

Thermally Stable and Flame Retardant Elastomeric


Nanocomposites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot

Recycling of Elastomeric Nanocomposites . . . . . . . . . . . . . . . . . . . . . 179


L. Reijnders

vii
viii Contents

Part III Application

Elastomeric Nanocomposites for Tyre Applications. . . . . . . . . . . . . . . 201


Kaushik Pal, Samir K. Pal, Chapal K. Das and Jin Kuk Kim

Elastomer Clay Nanocomposites for Packaging. . . . . . . . . . . . . . . . . . 233


V. Mittal

Elastomeric Nanocomposites for Biomedical Applications . . . . . . . . . . 257


Nicole Fong, Anne Simmons and Laura Poole-Warren

Actuators and Energy Harvesters based on Electrostrictive


Elastomeric Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Kaori Yuse, Pierre-Jean Cottinet and Daniel Guyomar

Elastomeric Nanocomposites for Aerospace Applications . . . . . . . . . . 307


James Njuguna, Krzysztof Pielichowski and Agnieszka Leszczyńska

Friction and Wear of Rubber Nanocomposites Containing Layered


Silicates and Carbon Nanotubes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
D. Felhös and J. Karger-Kocsis

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
Part I
Introduction & Preparation
Role of Different Nanoparticles
in Elastomeric Nanocomposites

Jin Kuk Kim, Kaushik Pal and V. Sridhar

Abstract The use of different types of fillers in preparation and characterization


of rubber based nanocomposites is depicted in this chapter followed by different
types of techniques used for dispersion of nanofillers in the rubber matrix. The
effects of the different fillers (CNT, CNF, nanographite) are generally discussed in
terms of morphology (investigated by electron microscopy, small angle X-ray
scattering, AFM), mechanical properties (modulus, stress and strain) and electrical
properties (impedance, percolation threshold etc.). The main aim of this research is
the dispersion of the carbonaceous materials in the elastomers and TPE gels which
is a significant obstacle to the scientists. In this chapter we have tried to overcome
this problem by using different types of fillers and it has been proved by the results
obtained from different types of characterization methods.

1 Introduction

Elastomeric composites are widely used because of their light weight, design
flexibility, and processability. However, these composites exhibit less attractive
mechanical properties such as low strength and low elastic modulus as compared
to metals and ceramics. Adding micron- or nano-sized inorganic filler particles to

J. K. Kim (&), K. Pal and V. Sridhar


Elastomer Lab, Polymer Science and Engineering,
School of Nano and Advanced Materials,
Gyeongsang National University, Jinju 660-701, Gyeongnam, Korea
e-mail: rubber@gnu.ac.kr
K. Pal
e-mail: pl_kshk@yahoo.co.in

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 3


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_1,
 Springer-Verlag Berlin Heidelberg 2011
4 J. K. Kim et al.

reinforce the polymeric materials has been standard practice in the composite
industry for decades. The term ‘‘filler’’ in rubber technology is often misleading,
implying a material that is primarily intended to reduce the cost of the more costly
rubber. But the modern-day fillers change one or more of these properties: optical
properties and color; improve surface characteristics and dimensional stability;
change thermal, magnetic, and electrical properties; improve mechanical proper-
ties, durability, and rheology; affect chemical reactivity, biodegradability, etc. The
mechanical and physical properties of the composite are mostly dominated by the
nature of the filler, whereas the polymer matrix determines the environmental
characteristics of the composite. Therefore, the overall composite properties can be
tailored to fit the desired application through the proper choice of filler and matrix
resin. Therefore, a judicious choice of the type of filler and its concentration in the
composite will augment the overall performance of the composite [1].

1.1 Classification of Fillers in Elastomers

Although many types of fillers (e.g., fibers, whiskers, particulates, etc.) are widely
used in the rubber industry, particulate fillers form a major share. Particulate fillers
are broadly classified as reinforcing and non-reinforcing, depending on whether or
not they enhance the performance characteristics of the final product.

1.1.1 Non-Reinforcing Fillers

Fillers that only lead to small increases in viscosity of the compound, cause
deterioration of the mechanical properties of the vulcanizate, and do not exhibit
any reinforcing action are called non-reinforcing or inactive fillers (e.g., calcium
silicate, chalk powder, etc.). These are often called extenders, and are used to
reduce the production cost of rubber goods [2]. Generally, powder minerals are
used as fillers and a few are listed below.
The most widely used non-reinforcing fillers are china clay and calcium car-
bonate. Clay is probably the most commonly used non-reinforcing fillers and is
classified as a hard clay or a soft clay, depending on their particle size (soft clay,
[2 lm; and hard clay, \2 lm). These are low-cost fillers that can be used at high
volumes to provide cheap compounds. However, clays impart improvement in
properties such as hardness and abrasion resistance, and also reduce mold
shrinkage.
Calcium carbonate, also known as whiting, is extensively used in rubber for-
mulations, primarily to impart color and reduce the cost the product [3, 4]. The two
main types are ground limestone and precipitated calcium carbonate. The former is
made by grinding mineral limestone and the latter is obtained by chemical pre-
cipitation from salt solution. Whiting provides smooth compounds that are easily
processed and often used in microcellular foams. Ground whiting gives low tear
Role of Different Nanoparticles 5

resistance but the finer particle size precipitated materials exhibit better charac-
teristics in this regard and have proved to have good hot tear strength.
Other non-reinforcing fillers are barytes (barium sulfate), mica, titanium
dioxide (TiO2), and silicates of calcium and zinc [5–9]. Barytes is generally used
in medical applications because of their unique chemical resistance and inertness.
Mica is used in composites where high thermal expansion and high resistivity are
needed. Titanium dioxide is incorporated in rubber to mask the inherent color of
the matrix.

1.1.2 Reinforcing Fillers

The basic feature that distinguishes between non-reinforcing and reinforcing fillers
is enhancement in performance characteristics such as the tensile strength, mod-
ulus, etc. of the composite. Reinforcement in vulcanized elastomers can be defined
as the simultaneous increase in stiffness and resistance to fracture by the addition
of filler [10–12]. A broader definition of reinforcement vis-à-vis the rubber
industry is ‘‘the improvement in abrasion, tear, cutting and rupture resistance,
stiffness and hardness of vulcanized compounds’’ [13]. Particulates such as carbon
blacks and silica are the most widely used reinforcing fillers in the rubber industry.
Carbon black, whose reinforcing character has been extensively exploited in
rubber engineering, was discovered in the early twentieth century (at Silverton,
United Kingdom, in 1910) and is the most widely used particulate filler in the
rubber industry. Carbon black reinforcement became a subject of scientific interest
only in the 1940s due to the development of suitable investigating tools and the
growing use of synthetic rubbers in demanding applications, namely automotive
and truck tires [14]. Carbon black is manufactured by a variety of processes,
namely furnace, thermal, acetylene, and channel. The characteristics of the carbon
blacks and their reinforcing nature in a polymer matrix are widely dependent on
the type of manufacturing process.

1.1.3 Nano-Sized Fillers

New-generation nano scaled fillers are challenging the domination of traditional


fillers such as carbon blacks and silica in the rubber industry. Nanoscaled fillers
such as layered silicates, carbon nanotubes, carbon nanofibers (CNFs), exfoliated
graphite, etc. dispersed as a reinforcing phase in an elastomer matrix are emerging
as a relatively new form of useful materials. These composites exhibit a change in
composition and structure over a nanometer length scale and possess remarkable
property enhancements relative to the pure polymer. Owing to the nanometer-size
particles obtained by dispersion, these nanocomposites exhibit superior mechani-
cal, thermal, optical, and dynamic mechanical properties at lower concentrations
compared with either the pure polymer or conventional micron-sized composites.
Their unique properties stem from a combination of factors such as their high
6 J. K. Kim et al.

aspect ratio (length to diameter), surface area, and the molecular bonds formed
between the nano-sized filler and the polymer during compounding.
The utility of nano-scaled fillers as possible reinforcing agents in elastomeric
composites can be traced to the earliest work of Schmidt, who reported that
colloidal stannic oxide, silica, Prussian Blue, polystyrene, and casein have very
strong reinforcing effects in styrene butadiene rubber (SBR) and concluded that
the ‘‘small particle size of the pigment (filler) is of prime importance in elas-
tomer reinforcement, whereas the chemical nature of the filler appears to be of
secondary importance’’ [15]. Although this hypothesis cannot be conclusively
proven experimentally, there is, nevertheless, good evidence to show that small
particle size is a necessary requirement, and very likely the predominant
requirement for reinforcement effect in rubber is the small size of the fillers.
Consequently, one can observe an increase in reinforcement with decreasing
primary particle size in carbon blacks as reported by Studebaker [16], who
showed a gradation of properties from the thermal blacks, with particle diam-
eters above 300 nm, that show little reinforcement, through the ‘‘semi-rein-
forcing’’ furnace grades, with particle diameters in the region of 100–200 nm,
thence to the ‘‘high abrasion’’ furnace grades (about 40 nm) and finally to the
‘‘intermediate super-abrasion’’ or ‘‘super-abrasion’’ grades with primary particle
diameters below 35 nm. Wagner [17] reported that nano-sized fillers such as
soft clays having a particle in the range of 1,000–8,000 nm as ‘‘diluent’’ fillers
and hard clays, zinc and titanium oxides precipitated calcium carbonates in the
range of 100–1,000 nm as ‘‘semi-reinforcing’’ and precipitated calcium car-
bonates, silicas, calcium silicates or silicoaluminates, or anhydrous silicas with
particle diameters in the range of 10–100 nm as ‘‘reinforcing’’ fillers in natural
rubber.
Theoretical investigations on the utility of nano-scaled fillers in elastomers have
been carried out by Fowkes and Gent [18–20]. Fowkes was the first researcher to
report that when functional filler particles such as carbon blacks are dispersed in a
rubber matrix, the polymer wets and adheres to the surface and is held by moderate
intermolecular attractive forces and by surface tension. Huber and Heinrich [21,
22] presented detailed theoretical investigations concerning the hydrodynamic
reinforcement contribution in elastomeric composites with rigid filler particles of
fractal nature (carbon black and silica aggregates), spherical core–shell particles
with a soft core and hard shell, spherical core–shell particles with a hard core and
soft shell (carbon black particles with bound rubber). In the context of carbon-
black-filled elastomers, the contribution to reinforcement on small scales can be
attributed to the complex structure of the branched filler aggregates, as well as to a
strong surface polymer interaction, leading to the so-called bound rubber. Thus,
the filler particles are coated with polymer chains, and the binding (physically or
chemically) of elastomer chains to the surface of the filler particles significantly
changes the elastic properties of the macroscopic material. On larger scales, the
hydrodynamic aspect of the reinforcement dominates the physical picture.
Hydrodynamic reinforcement of elastic systems plays a major role not only in
carbon-black-filled elastomers, but also in composite systems with hard and soft
Role of Different Nanoparticles 7

inclusions. Finally, at macroscopic length scales, the existence of filler networking


at medium and high filler volume fractions plays the dominate role.

1.1.4 Carbonaceous Nano-Fillers

Carbon black is the most commonly used filler in elastomers. The past few years
have seen the extensive use of nanoparticles because of the small size of the filler
and the corresponding increase in the surface area, allowing achieving tremendous
increase in mechanical properties even at very low filler loadings. Recently,
considerable research efforts have focused on nano-scale variants of carbon black
(viz. carbon nanotubes, carbon nanofibers, and exfoliated nanographite) as pos-
sible reinforcing fillers in elastomers. Among these, nanotubes are attracting the
most attention.

Carbon Nanotubes

The discovery of carbon nanotubes can be traced back to the origin of fullerene
chemistry (Buckyball, C60) in 1985 [23]. In 1991, Ijima [24] discovered carbon
nanotubes (CNTs) that are elongated fullerenes, where the walls of the tubes are
hexagonal carbon and often capped at each end. Carbon nanotubes are needle-
shaped, single crystals whose properties depend on the atomic arrangement, chi-
rality, diameter, and length of the tube and the overall morphology. Carbon
nanotubes can be synthesized by different techniques, including arc-discharge
laser ablation and various catalytic chemical vapor depositions (CCVDs) [25–29].
There are two types of CNTs: (1) multiwalled carbon nanotubes (MWCNTs) and
singlewalled carbon nanotubes (SWCNTs). An SWCNT (or SWNT) is best
described as a twodimensional graphene sheet (a hexagonal array of carbon atoms)
rolled into a tube with pentagonal rings as end caps. SWNTs have aspect ratios of
1,000 or more, and an approximate diameter of 1 nm, or 10 Å. Similarly, MWCNTs
(or MWNTs) can be described as multiple layers of concentric graphene cylinders,
also with pentagonal ring end caps. Both SWCNTs and MWCNTs have physical
characteristics of solids and are microcrystals with high aspect ratios of 1,000 or
more, although their diameter is close to molecular dimensions. Table 1 shows the
theoretical and experimental properties of carbon nanotubes.
From Table 1 it is clear that CNTs have a unique combination of mechanical,
electrical, thermal, and electrical properties that make them excellent candidates to
substitute or complement the conventional fillers in the fabrication of elastomeric
composites.

Carbon Nanofibers

The history of carbon nanofibers goes back more than a century. In a patent
published by Hughes and Chambers [30] in 1889, it was reported that carbon
8 J. K. Kim et al.

Table 1 Theoretical and experimental properties of carbon nanotubes


Property CNT Graphite Carbon black
3 3
Specific gravity 0.8 g/cm for SWNT 2.26 g/cm 1.2–2.4 g/cm3
1.8 g/cm3 for MWNT
Modulus *1 TPa for SWNT 1 TPa (in plane) –
*0.3–1 TPa for MWNT
Strength 50–500 GPa for SWNT – –
10–60 GPa for MWNT
Resistivity 5–50 lX-cm 50 lX-cm (in plane) 25–40 lX-cm
Thermal conductivity 3,000 W/m-K 3,000 W/m-K (in plane) –
Specific surface area 100–250 m2/g 10–20 m2/g 8–130 m2/g

filaments were grown from carbon-containing gases using a metallic crucible.


However, 80 years later it was Robertson [31] who was the first to recognize that
it was the interaction of methane and metal surfaces that led to the growth of
graphitic carbon at relatively low temperatures. For the first 80 years of this
century, the occurrence of carbon nanofibers—then often referred to as carbon
filaments or filamentous carbon—was considered a nuisance. In the 1980s,
several workers explored the use of carbon nanofibers for such applications as
reinforcing fillers in polymers and as catalyst support material. Carbon nanofi-
bers based on different precursors (PAN, rayon, and pitch) show various kinds of
surface structures [32]. The basic structure of a typical carbon fiber (PAN)
consists of long primary units (lateral aromatic molecules) lying parallel to the
fiber axis and bonding together to form a stretched network of branched fibrils
that apparently run the full length of the fiber. They form ribbon-shaped mon-
atomic layers of sp2-type carbon, with an average width of 6 nm and a length of
several hundred nanometers [33, 34]. The ribbons display an irregular contour,
which appears ‘‘turbostratic’’ in the crystalline structure, and a certain number of
these ribbons run parallel to form the microfibrils that have a preferred orien-
tation parallel to the fiber axis.

Layered Nano Dimensional Fillers

Layered fillers such as mica and talc have been used as fillers for elastomers due
to their low cost, easy availability and outstanding electrical, heat, and chemical
resistance. The extent of reinforcement by mica depends on the aspect ratio (i.e.,
the, ratio of length to thickness). The breakdown of mica particles during mixing
and processing causes a reduction in the aspect ratio, which can be countered
partially by using special equipment such as hot runners, longer sprues, and
streamlining to reduce particle breakdown [35]. But mica is an inert filler devoid
of any chemical groups, thereby having very low reinforcing capability when
compared to traditional rubber fillers such as carbon black and silica. A variety
of surface treatments, such as reacting mica platelets with titania, chlorinated
Role of Different Nanoparticles 9

paraffins, and silane coupling agents, has been studied in order to increase the
mica-polymer interactions [36–38]. Other physical means, such as ultrasonic
delamination, have been investigated by Tausz and Chaffey [39]. Mica-enhanced
anisometric swelling of nitrile rubber vulcanizate and reduction of permeability
of gases have been reported [40]. Mica, due to its layered structure, has good
shock-absorbing and vibration-damping properties and has been investigated in
polyurethane [41]. The anisotropic nature of mica will be useful in barrier
applications [42]. Yoshikawa et al. [43] patented a rubber hose based on nitrile
rubber–mica composites for fuel line systems in which swelling, permeability,
and low-temperature properties were optimized by selection of the mica filler
without any surface treatment. Debnath et al. [44] reported the effect of a silane
coupling agent on vulcanization conditions, network structure, polymer-filler
interaction, physical properties, and failure mode of mica-filled styrene butadiene
rubber.
The other type of layered filler is nanoclay. During the past 10 years, nanoclays
have raised considerable interest in the scientific community due to their size and
wide range of outstanding material properties. Of specific interest is the use of
nanoclay-reinforced composites for structural applications, where experimental
results demonstrate that substantial improvements in the mechanical behavior of
the polymer can be attained through the addition of very small amounts of
nanoclay. Understanding the structure–property relations in polymer–clay nano-
composites is of great importance in designing materials with the desired prop-
erties. In general, the improvement in properties of a polymer–clay nanocomposite
originates from the nature of the layered inorganic fillers, their extent of dispersion
in the polymer matrix, and by the distinctive interactions between a specific
polymer and the nano-clay.
Significant improvements in the physical and mechanical properties of layered
alumina silicate reinforced polymeric composites have been reported. Some of the
properties that undergo substantial improvement due to addition of nano-phases
include mechanical properties such as strength, modulus, and dimensional stability;
decreased permeability to gases, water, and hydrocarbon vapors; thermal stability
and heat distortion temperature; flame retardance; surface appearance; and elec-
trical conductivity. The most common types of nano-reinforcements are layered
natural silicates such as montmorillonite, hectorite, and saponite [45–47]. Synthetic
clays such as synthetic mica and synthetic hectorite have also been used [48, 49].
However, the most extensively used and studied is montmorillonite (MMT).
Dispersion within the polymer matrix is required to form the inorganic–organic
composite. To make MMT compatible with the organic polymer matrix, a cation
exchange reaction is used to replace the sodium, potassium, and calcium ions with
alkyl ammonium long-chain ions. This result in expansion between the clay gal-
leries due to the larger molecules inserted between the layers. The reaction also
changes the clay from hydrophilic to hydrophobic, making it more compatible
with the organic matrix. Clay minerals are composed of nanoplatelets with a length
of about 0.5–1 lm and a thickness of 1 nm, leading to large aspect ratios (500–
2,000).
10 J. K. Kim et al.

Processing of nanoclay composites is a real challenge because the elemen-


tary nanolayers or tactoids of a few platelets must be uniformly distributed in
order to develop the highest amount of interfacial area, thus increasing the
contact surface with the polymer chains. The main advantage of polymer-lay-
ered silicate nanocomposites is that great improvements in mechanical prop-
erties can be obtained even at very low clay content, which is economically
interesting. However, to achieve this, there must be good dispersion of nanoclay
platelets in the polymer matrix. But the intercalation or exfoliation of polymer
chains in the nanoclay inter-gallery layers is a complex reaction that depends
on many variables, such as polymer–clay interactions, chain conformation of
the polymer, etc. However, the general consensus is that the major difficulty for
the polymer chain to intercalate in the inter-gallery layers arises from the
inherent structure of the clay particles, which have closely stacked silicate
sheets. The diameter of these sheets typically lies between 20 and 200 nm,
while the thickness is on the order of 1 nm. The spacing between the sheets is
roughly on the order of 1 nm, which is less than the radius of gyration of
typical polymers. Therefore, there is a large entropic barrier that inhibits the
polymer from penetrating this gap and intermixing with the clay. Even when
the sheets are successfully separated and interspersed into the polymer matrix,
these high aspect-ratio platelets can form ordered or crystalline structures within
the polymers or can phase-separate from the matrix material. To reduce the
entropy of the nanoclay platelets, it is modified by treating with alkyl ammo-
nium modifiers. These alkyl ammonium modifiers react electrokinetically with
the cations inside the galleries of the clay by ion exchange, thereby causing the
clay platelets to move apart, resulting in a reduction in the level of platelet-
platelet attraction and in increased access for the polymer chains to intercalate
[50]. Of the many techniques employed to achieve a high degree of dispersion
of nanoclay platelets, in situ polymerization of monomers initially intercalated
between silicate layers, and melt intercalation dispersing nanoclay platelets in a
polymer solution are the most common [51–56].
There are few studies available on rubber-clay nanocomposites [57]. The
important rubberclay nanocomposites thus far studied are elastomeric polyure-
thane, natural rubber, epoxidized natural rubber, styrene butadiene rubber, buta-
diene rubber, nitrile butadiene rubber, EPDM, silicone rubber, fluoroelastomers,
butyl rubber, chlorobutyl rubber, etc. [58–68].

Nanographite

Nanographite shows a similar layered structure and has a high aspect ratio in the
range of 1,000–1,500 [69]. Graphite can provide additional advantages such as
excellent electrical and thermal conductivity. Exfoliated graphite is prepared by
rapidly heating a graphite intercalation compound (GIC).
An exfoliated graphite nanoparticle is composed of stacks of nanosheets that
can vary from 4 to 40 nm. They also show good affinity for both organic
Role of Different Nanoparticles 11

Table 2 Comparisons of properties of nanoclays and exfoliated nanographites


Exfoliated clay Graphite nanoplatelets
Physical structure Platelets, *1 nm 9 100–200 nm Platelets, *1 nm 9 100–200 nm
Chemical structure SiO2, Al2O3, MgO, K2O, Fe2O3 Graphene
Type of interactions Hydrogen bonds, dipoles p–p
between platelets
Tensile modulus 0.17 TPa *1 TPa
Tensile strength *1 GPa *10–20 GPa
Electrical resistivity 1010–1016 X-cm 50 9 10-6 X-cmk
*1X-cm\
Thermal 6.7 9 10-1 W/m-K 3,000 W/m-Kk
conductivity 6 W/m-K\
Specific gravity 2.8–3 g/cm3 2 g/cm3

compounds and polymers, therefore some monomers and polymers can be


absorbed into the pores and galleries of exfoliated graphite. A comparison between
layered silicates and exfoliated nanographite is shown in Table 2. From this table
it can be observed that nanographite has additional advantages, including
increased electrical and thermal conductivity. So, if one can effectively disperse
nanographite in a polymer composite, many additional properties such as con-
ductivity and electromagnetic shielding resistance can be obtained.

2 Problems Associated with Nanofillers in Elastomers

2.1 Dispersion

It is recognized that there is still a long way to go to create CNT-reinforced


composites fully realizing the potential of high stiffness and strength of CNTs.
Many of the reported CNT polymer composites have not presented anticipated
macroscopic properties such as high stiffness and strength, although the individual
CNTs have excellent mechanical properties. There are many reasons that influence
the macroscopic mechanical properties of nanocomposites, such as the dispersion,
alignment, and waviness of CNTs. The dispersion of CNTs in polymers is rela-
tively poor due to the nanotubes having a strong tendency to agglomerate as a
result of their strong van der Waals’ forces. Their very stable chemical charac-
teristics and lack of functional sites on the surface also complicate the dispersion
issue. Moreover, the length of CNTs prepared from thermally activated CVD
ranges from meters to several millimeters, which is undesirable for practical
applications. Both physical and chemical approaches have been adopted to
reduce the length of CNTs to a certain extent that is suitable for blending.
The physical dispersion route generally includes ultrasonication, ball milling,
grinding, and high-speed shearing. The as-prepared CNT product exists as loose
12 J. K. Kim et al.

multi-agglomerates, which can be separated by physical methods. Here we present


a brief discussion regarding some of the problems associated with the fabrication
of nano-scale filler-reinforced elastomers.
There are four main system requirements for effective reinforcement [70]: (1) a
large aspect ratio, (2) good dispersion, (3) alignment, and (4) interfacial stress
transfer. The aspect ratio must be large to maximize the load transfer to the
nanotubes. This is crucial in order to optimize composite strength and stiffness.
Dispersion is probably a more fundamental issue. Dispersion of fillers in a polymer
matrix results from the application of electrochemical and mechanical forces to the
interface of the inorganic filler/polymer so as to cause complete de-agglomeration
to the attired or original particle size in an organic phase, complete elimination of
air voids and water, and the creation of a true continuous inorganic–organic
composition [71].
The optimal performance of such nano filler reinforced polymer composites is
achieved when the nano-sized fillers are uniformly dispersed in the polymer
matrix. Fabricating such homogeneous mixtures poses considerable synthetic
challenges. The major difficulty arises from the inherent anisotropic nature of
the clay particles, which are composed of broad, closely stacked silicate sheets.
The diameter of the sheets lies typically between 20 and 200 nm, depending on the
specific type of clay, and each sheet is roughly 1 nm thick.
The spacing (or ‘‘gallery’’) between the sheets is also on the order of 1 nm,
which is smaller than the radius of gyration of typical polymers. Consequently,
there is a large entropic barrier that inhibits the polymer from penetrating this gap
and intermixing with the clay. Theoretical calculations made by Balazs’ group in a
series of articles show that to understand the exfoliation process in polymer–clay
nanocomposites, one must consider the kinetic and thermodynamic aspects of the
penetration of polymer into the gap between the clay sheets [72–74]. So, based on
this, the effective dispersion of nano-sized fillers in a polymer matrix remains
more of an art than a precise science.

2.2 Low Mechanical Properties

Nanotubes must be uniformly dispersed as isolated nanotubes individually coated


with polymer. This is imperative in order to achieve efficient load transfer to the
nanotube network. This also results in a more uniform stress distribution and min-
imizes the presence of stress-concentration centres. Alignment is, in some ways, a
less crucial issue. From geometric considerations, the difference between random
orientation and perfect alignment is a factor of five in composite modulus. While
alignment is necessary to maximize strength and stiffness, it is not always beneficial.
According to the Mori–Tanaka theory, aligned plate-like objects intrinsically rein-
force less than aligned fibers [75–78]. Consequently, in the absence of disorder,
modulus enhancement in clay nanocomposites would theoretically be less than that
of carbon nanotubes. Angular averaging has less impact on platelets than on fibers,
Role of Different Nanoparticles 13

so randomly aligned platelets are ultimately more effective than randomly aligned
fibers in terms of modulus enhancement [79].
The interface behavior can significantly affect the mechanical properties of
nanocomposites. Nanotubes and nanofibers are normally produced at very high
temperatures, thereby leading to the formation of highly graphitized carbon structures
devoid of any surface groups. So, to increase the interactions, some sort of surface
functionalization or activation of nanotubes and nanofibers is necessary. Rubber
technologists were the first to recognize this phenomenon, and consequently a large
amount of research and efforts has focused on this. They postulated that for a filler to
be considered an effective reinforcing agent, primary valence bonding between
polymer and filler is a prerequisite. The general conclusion of many studies is that
higher the ‘‘surface activity’’ (due to presence of active surface groups such as
carboxy, carbonyl, lactones, quinones, etc.) of the carbon black, the greater will be
the reinforcement. While it is true, as indicated above, that reinforcement does occur
in systems where strong interactions are highly unlikely, there is much evidence that a
degree of reinforcement is less and is greater in fillers with strong interactions and is
important in optimizing the properties of practical rubber compositions [80]. The
most familiar example for understanding is the removal of reaction sites from carbon
black (by graphitization) causes a drastic reduction in the modulus and abrasion
resistance properties imparted by the parent (ungraphitized) carbon. Strong bonding
may be of importance in at least two distinct ways. First, it may be used as a means of
pulling apart filler agglomerates during the mixing process, thereby providing
improved dispersion of the ultimate particles in the rubber, and opposing any ten-
dency toward subsequent particle re-agglomeration [81]. Second, it may contribute to
the adjustment of important physical characteristics, such as modulus, extensibility,
and resilience in the vulcanizate.
Carbon nanotubes are highly anisotropic in nature. Bradshaw et al. [82] pro-
posed that curvature of embedded CNTs or ‘‘waviness’’ significantly reduced their
reinforcement capabilities (by factors from 50 to 200) compared to straight CNTs.
It was also noted that other indistinguishable factors contribute to the low values
measured in experimental data, including weak interfacial bonding, insufficient
dispersion, and degradation of the CNTs due to processing. At very low strains, it
was suggested that the effect of poor interfacial shear strength should not affect the
composite’s modulus, implying a measurable elastic response below the strain
detrimental to the CNT interface [83]. Studies by Schadler et al. [84–86] have
shown that carbon nanotubes in general do not bond well in polymers and their
interactions result mainly from the weak Van der Waals forces. Consequently,
CNTs may slide inside the matrix and may not provide much reinforcing effect.
It is, however, important to assess whether the poor interface behavior is indeed
responsible for the shortfall of CNT-reinforced composites in order to reach their
expected properties. This is attributed to the absence of stress transfer to internal
layers of MWNTs; only the outermost layer contributes to tensile reinforcement.
It is noted that the additions of pristine and functionalized carbon nanotubes and
carbon black particles lead to relatively large enhancements in Young’s modulus
and fracture toughness [87, 88].
14 J. K. Kim et al.

3 Some Possible Solutions

The high aspect ratio of the nanotubes coupled with a strong intrinsic van der
Waals attraction between nanotubes combine to produce ropes and bundles of
CNTs, particularly in SWNTs where the attractive force is approximately 0.5 eV
per nanometer of nanotube-to-nanotube contact [89]. In the case of multi-walled
carbon nanotubes (MWCNTs), the molecular forces between the individual
MWCNTs might not be significant, so that sp2/sp3 orbit-hybridization in the walls
of the MWCNTs rarely occurs. Therefore, MWCNTs are formed in the individual
state with diameters at the high end of the nanometer scale. However, the dis-
persion of SWNTs is more difficult due to their highly crystalline nature and very
strong sp2/sp3 hybridization.

3.1 Physical Techniques for Good Dispersion

3.1.1 Grinding and Rubbing

There are few reports on the subject of rubbing or grinding carbon nanotubes to
decrease the size. Rubbing is more destructive than any other method. The process
introduces cuts and bends in the SWNT [90]. A less damaging method is chem-
ically cutting SWNTs by grinding them in a fluid (alpha- or beta-cyclodextrin)
using mortar and pestle. Both tube lengths and bundle diameters were noticeably
reduced. Other grinding agents were used as well but did not give as good results;
samples contained mostly long tubes [91].

3.1.2 High Shear Mixing-Shear Stress through a Nozzle

Another mechanical method is to apply shear force to pull agglomerates apart—


that is, high shear mixing. Usually narrow passages, and/or relatively high rates of
flow, are required to generate high shear; in a lot of cases, rotor and stator con-
struction is used [92]. Hilding et al. [93] applied a diesel fuel injector to create the
high shear and correlated the reduction in viscosity with nanotube breakage.

3.1.3 High-Energy Ball Milling

Pierard et al. [94] reported a comprehensive study on the effect of high-energy ball
milling on the structure of SWNTs. Based on physical properties such as diameter,
length, and surface area, they concluded that the optimum ball milling time is 2 h.
Konya et al. [95] reported a simple ball milling technique in specific atmospheres
to introduce functional groups such as thiol, amines, amides, carbonyl, etc.
Role of Different Nanoparticles 15

Lee et al. [96] reported a cryogenic crushing in the presence of liquid nitrogen to
produce CNTs of 500-nm length. Other interesting technique to cut CNTs is the
use of gamma-radiation in the presence of dilute sulfuric acid [97].

3.2 Chemical Techniques for Good Dispersion

Carbon nanotubes and nanofibers are devoid of any chemical functional groups on
their surfaces and hence have low interfacial interactions between them and the
polymer matrix. Some researchers have made considerable efforts in the chemical
modification of NTs, which might pave the way to many useful applications. The
main approaches for the modification of these quasi-one-dimensional structures
can be grouped into three categories [96]:
(1) Covalent attachment of chemical groups through reactions with the ð-conju-
gated skeleton of CNT;
(2) Non-covalent adsorption or wrapping of various functional molecules
(3) Endohedral filling of their inner empty cavity
A comprehensive review of the grafting and coating of CNTs by polymers is
beyond the scope of this chapter. Some representative methods include attachment
of long alkyl chains and polymers, fluorination, and radical reactions; these have
provided access to tip and sidewall functionalization, eventually leading to a
relative increase in solubility of CNTs in the polymer matrix [97–100]. The utility
of silane coupling agents such as aminopropyl triethoxy silane and TESPD
{poly(trimethyl dihydro quinoline)} to improve the rubber-filler interactions has
been well investigated [101]. Shanmugharaj et al. [102] have functionalized CNTs
with aminosilanes and explored the utility of the same as reinforcing filler in
natural rubber. In addition to the above techniques, grafting or wrapping of con-
ductive nano-sized metals on the surface of CNTs, especially for electromagnetic
(EMI) shielding applications, have also been investigated. Techniques such as sol–
gel, vapor deposition, and electrochemical methods have been used [103, 104].
The basic step for all chemical functionalizations of nanotubes or nanofibers is
acid treatment. Although a concentrated H2SO4/HNO3 (vol:vol = 3:1) treatment
is efficient in severing entangled nanotubes to enable their dispersion as individ-
uals, damage to the tube-wall layers is serious and unavoidable; it has been
reported that highly dispersed rod-like carbon can be achieved by sonication in
negatively charged polyelectrolyte solutions [105, 106]. It has also been reported
that acid treatment of CNT could improve the processability and performance of
composites by introducing carboxylic acid groups on the surface of CNT, which
leads to stabilization in polar solvents and helps to covalently link polymers.
It must be mentioned that both acid treatment and ultrasonication have been
reported to cause fragmentation (comminution) and damage of the nanotubes,
which can be modeled using the moments of the length distribution. As with other
solids, the breakage rate of carbon nanotubes depends on their lengths, with the
16 J. K. Kim et al.

longest particles experiencing the highest breakage rates. Perhaps the earliest
reports of the effect of ultrasonic treatment on the properties of CNT was reported
by Lu et al. [107], who reported extensive bending, buckling, damage, and for-
mation of defects on the surface of nanotubes.
The utility of surfactants to increase the dispersion of CNTs in the polymer
matrix has been investigated. A surfactant’s property of accumulation at surfaces
or interfaces has been widely utilized to promote stable dispersions of solids in
polymeric media [108]. These amphiphilic molecules (i.e., compounds having
both polar and apolar groups) adsorb at the interface between immiscible bulk
phases, such as oil and water, air and water, or particles and solution, act to reduce
the surface tension. The distinct structural feature of a surfactant originates from
its ‘‘duality’’: the hydrophilic region of the molecule or the polar head group, and
the hydrophobic region or the tail group that usually consists of one or few
hydrocarbon chains. Although surfactants increase the dispersion of CNTs in the
polymer matrix, it has many detrimental effects on the performance characteris-
tics. Hernandez et al. have reported that by addition of the 1 wt% of CNT, which
was initially dispersed by Triton X-100 surfactant (polyethylene oxide (9) nonyl
phenyl ether), the improvement in storage modulus was only 50% when compared
to over 230% in untreated CNT [109]. A similar tendency was found when the
glass transition temperatures, Tg, of the neat polymer and surfactant-contained
polymer were compared. This surfactant-induced plasticization effect was also
reported for epoxy and polyethylene glycol matrices [110, 111].

3.3 Analysis of Dispersion

In addition to the obvious difficulty in obtaining stable and homogeneous dis-


persions of nanotubes, another complication is finding a valid method to evaluate
their state of dispersion. The most widely used visualization technique is
microscopy: optical, scanning electron, and transmission electron microscopies.
Visualization of CNT-based samples by optical microscopy enables to access
mainly micrometer-sized agglomerates, while atomic force microscopy (AFM) is
used to monitor suspended CNTs at the nano-scale level. But by using AFM, one
can probe only a few nanotubes at a time, which might not be representative of the
entire sample. Imaging of CNT-based polymeric composites by scanning or
transmittance electron microscopy (SEM or TEM, respectively) often requires
pretreatment by means of gold or carbon sputtering or microtome slicing of the
sample, which might cause a defect in the original pattern of the composite.
Solutions of carbon nanotubes are best viewed with cryo-TEM, which is ideally
suited for imaging of wet samples.
Scattering techniques such as x-ray diffraction, small angle neutron scattering,
and dynamic light scattering have been used to study MWNT suspensions by
particle size and agglomerate size, both in solution and polymer composites. The
data analysis techniques have many complications, including assumptions of
Role of Different Nanoparticles 17

perfectly spherical particles, aligned rods, etc. As a result, the reported dimensions
of nanotubes are in over-fitting range.
The discovery of nanotube fluorescence has led to a precise method of detecting
individual nanotube dispersion [112]. A nanotube in an aligned bundle does not
emit because of energy transfer to neighboring tubes, particularly to the metallic
ones. Thus, the dispersion process can be monitored by examining transient
fluorescent emission as a function of various parameters, such as the type of
surfactant used, sonication time, and surfactant concentration and functionalization
[113]. In case of MWNTs, fluorescence spectra data can be misleading because
MWNTs are a mixture of conducting (metallic) and semi-conducting nanotubes
[114].

4 Utility of Carbonaceous Nanofillers in Elastomers and


TPE Gels

The next part of this chapter briefly discusses the preparation and properties of
carbonaceous nano filler reinforced polymeric composites studied the following
items by our group.
• Exfoliated nano graphite in fluoroelastomers [115, 116]
• Multiwalled carbon nanotubes in fluoroelastomers [117]
• Vapor-grown carbon nanofibers in chlorobutyl elastomers [118]
• Multiwalled carbon nanotubes reinforced thermoplastic gels [119]
Elastomeric nanocomposites were prepared by traditional two-roll mill mixing
technique, whereas swelling of SEBS thermoplastic in mineral oil was used to
fabricate the TPE (thermoplastic elastomer) gels. Special attention focused on the
dielectric relaxation of fluoroelastomer–nanographite composites, whereas strain-
dependent dynamic mechanical properties were extensively studied for chlo-
robutyl-CNF composites.

4.1 Morphology

The dispersion and morphology of the nanofillers in the composites have been
studied by AFM, SEM, and TEM (Figs. 1, 2, 3).

4.1.1 Nanographite

The dispersion of graphite platelets in fluoroelastomers has been studied using


TEM andrepresentative microphotographs at 0.5 phr and 3.5 phr nanographite
loadings are shown in Fig. 4a and b, respectively. From this figure it can be
18 J. K. Kim et al.

Fig. 1 General structure of


sodium montmorillonite (Na-
MMT)

Fig. 2 General procedure used to make exfoliated nanographite

observed that the graphite nanosheets consist of thin graphite nano-lamellae with
thicknesses of 1–5 nm or thinner, and with inter-gallery spacings of 3.37 Å. These
inter-gallery distances reflect the XRD plots (Fig. 5), which show that both pure
graphite platelets and graphite-FKM exhibit an intense peak at a diffraction angle
of 26.4, corresponding to a basal spacing of 3.37 Å. A similar observation was
reported by Yasmin and Daniel [120], who concluded that graphite platelets
remain multilayered and maintain their original d-spacing. The dispersion of the
nanographite platelets in the polymer matrix has been further studied by atomic
force microscopy. Figures 6 and 7 show a representative topology of 0.5 and 3.5
phr nanographite reinforced. From these figures, multilayered nanographite
platelets having a parallel orientation can be observed.
Role of Different Nanoparticles 19

Fig. 3 Dispersion and fabrication of nanoclay-reinforced polymer composites

Fig. 4 TEM of nanographite


dispersed fluoroelastomer
composites at a 0.5 phr and b
3.5 phr
20 J. K. Kim et al.

Fig. 5 WAXD diffracto-


graphs of the matrix and the
nanographite/fluoroelastomer
composites

Fig. 6 a Topology and 3-D AFM image of 0.5 phr nanographite-reinforced fluoroelastomer
composites, respectively; b cropped topographical image of the nanographite platelets; and c line
diagram
Role of Different Nanoparticles 21

Fig. 7 AFM topology images of 3.5 phr nanographite-reinforced fluoroelastomer composites

4.1.2 Carbon Nanofibers

Sections of samples of chlorobutyl rubber compounds containing increasing


amounts of VGCNFs (vapor-grown carbon nanofibers) were examined using SEM.
Figure 8a–c show the SEM microphotographs at increasing VGCNF loadings. All
the microphotographs show excellent distribution of the filler particles. From the
figures it can be observed that at all filler loadings there is uniform distribution of
the VGCNFs, the majority of which are individual nanofibers rather than aggre-
gates, thus implying uniform distribution of nanofibers in the polymer matrix.
SEM microphotographs (Fig. 8d, e) taken at lower length scales (2 lm) clearly
show this phenomenon.
The dispersion of VGCNFs in the polymer matrix was further studied by
electron probe microanalysis (EPMA). The main advantage of EPMA is
selective elemental scans. In the present study, carbon and oxygen scans have
22 J. K. Kim et al.

Fig. 8 SEM microphotographs of VGCNF-reinforced chlorobutyl elastomers at increasing filler


loadings: a 5 phr, b 10 phr, and c 15 phr. SEM microphotographs of VGCNF-reinforced chlo-
robutyl elastomers at: d 5 phr, e 10 phr, and f 15 phr taken at smaller length scale (2 lm)

been carried out and a representative carbon and oxygen scan of a VGCNF-
loaded CIIR compound is shown in Fig. 9a and b, respectively. Figure 9a
shows the distribution of nanofibers in the polymer matrix as the carbon scan,
wherein it can be observed that VGCNFs are well dispersed with islands of
carbon nanofibers, the majority of which are individual nanofibers rather than
aggregates and implying uniform and excellent distribution of the nanofibers in
the polymer matrix. Figure 9b shows the oxygen scan. Islands of dark color can
be observed in the figures that indicate the presence of oxygen on the surface of
the nanofibers.
Role of Different Nanoparticles 23

Fig. 9 Two-dimensional
distribution of carbon nanofi-
bers in chlorobutyl elastomer:
a carbon scan and b oxygen
scan

4.1.3 Carbon Nanotubes

Figure 10a–c, and d show the SEM microphotographs of CNT-reinforced fluoro-


elastomers at increasing MWNT loadings. All the microphotographs show
excellent distribution of the filler particles. From the figures it can be observed that
at all filler loadings there is uniform distribution of the MWNTs, the majority of
which are individual nanotubes rather than aggregates and implying complete
exfoliation of the nanotubes in the polymer matrix. The MWNTs (bright dots) are
found to disperse homogeneously in the rubbery matrix. The excellent distribution
of nanotubes in the polymer matrix can be explained in terms of colloids, wherein
the adsorbed polymer stabilizes the nanotube dispersion and protects it against
bridging flocculation and depletion aggregation caused by the free polymer [121].
Representative TEMs of a 6 phr MWNT-reinforced fluoroelastomer composite are
shown in Fig. 11a and b. Good dispersion of the nanotubes in the polymer matrix
can be observed. From Fig. 11b, taken at a smaller length scale (20 nm), an
absorbed polymer layer on the walls of nanotube can be observed, indicating
wetting of the nanotubes in the polymer matrix.
A novel technique, magnetic force microscopy (MFM), has been utilized to
further study the dispersion of nanotubes in the polymer matrix. MFM, a variant of
24 J. K. Kim et al.

Fig. 10 Scanning electron microscopy of razor-cut section of MWNT-reinforced fluoroelas-


tomer vulcanizates: a 1.5 phr, b 3 phr, c 4.5 phr, and d 6 phr (Reproduced from Ref. [117] with
permisssion from Wiley Interscience)

Fig. 11 TEM microphotographs of MWNT-fluoroelastomer composites (Reproduced from Ref.


[117] with permisssion from Wiley Interscience)

scanning force microscopy, is a tool capable of revealing a magnetic sample’s


domain structure in real space. A representative topological picture of 6 phr
nanotube reinforced fluoroelastomer is shown in Fig. 12. From the figure it is
observed that the topographical image of MWNT/fluoropolymer composite is not
as clear as has been observed by traditional techniques such as AFM and SPM; this
can be attributed to the reduced ability of image contrast in MFM. It is generally
known that it is not easy to separate the magnetic contrast from other background
forces in MFM topography images. However, polymer/nanotube composites are
two-phase materials with two distinct magnetic properties. The nanotubes are
paramagnetic or diamagnetic, depending on their orientation, whereas the polymer
Role of Different Nanoparticles 25

Fig. 12 Topology of MWNT-fluoroelastomer nanocomposites measured by magnetic force


microscopy (Reproduced from Ref. [117] with permisssion from Wiley Interscience)

matrix is paramagnetic. So when an MFM tip moves across the nanocomposite


sample, many interactions, which include magnetic interactions, electrostatic
interactions, Van der Waals interactions, short-range and capillary forces, take
place. Another reason for the poor quality of MFM images of nanotube-polymer
composites has been attributed to the basic working principle of MFM.
Our group reported the preparation and properties of MWNT reinforced TPE
gels [119]. Figure 13a and b display a pair of TEM images collected from NCTPE
(nanocomposite thermoplastic elastomer) gels modified with 5 wt% MWCNTs
taken from different sites on the same ultra-thin sample. The images of NCTPE
gels exhibit morphology composed of a micellar of SEBS in hydrocarbon oil and
MWCNTs. The irregularly shaped, long, and opaque features in both figures
identify the incorporation of MWCNTs that measure on the order of 10–16 nm
across and appear to flocculate into large-scale aggregates within the swollen
midblock of the SEBS triblock copolymer network. In some areas as indicated by
arrow, it is clear that these nanoparticles are flocculated into large-scale aggregates
due to high solvent content and hinder to a lesser extent the bridging efficacy of
individual copolymer molecules. The relatively high magnification image shows
the more detailed examination of the NCTPE gel morphology and confirmed that
the matrix consists of SEBS micelles measuring ca. 20 nm in core diameter uni-
formly dispersed throughout the hydrocarbon oil. It is clear from these related
26 J. K. Kim et al.

Fig. 13 TEM images of 5 wt% MWNT-based NCTPE gels at a low magnification and b high
magnifications

images that the micelles do not exhibit any discernible indication of long-range
order, a face- or, more likely, body-centered cubic lattice. The irregularly shaped
dark features in both figures identify the incorporation of MWCNTs, which appear
to exist as aggregates that measure on the order of 10–16 nm across. It is clear that
MWCNTs are flocculated into large-scale aggregates due to high solvent content
and hinder to a lesser extent the bridging efficacy of individual copolymer mol-
ecules, which possibly is dictated by poor dispersion.

4.2 Mechanical Properties

The variation in tensile strength and modulus with increasing nanofiller loadings is
shown in Fig. 14 (nanographite-reinforced fluoroelastomers), Fig. 15 (VGCNF-
reinforced chlorobutyl vulcanizates), Fig. 16 (MWNT-reinforced fluoroelastom-
ers), and Fig. 17 (MWNT-reinforced TPE gels). From the figures it is observed
that irrespective of the polymer matrix, there is a steady increase in tensile strength
and modulus with increasing concentrations of the nanofillers. Tensile strength can
be regarded as catastrophic tearing of cracks initiated by accidental flaws,
microvoids, dewetting, or cavitation from the filler surface. If the elastomeric
network is capable of dissipating the input energy into heat, then less elastic
energy will be available to break this polymer network. Incorporation of fillers is
the major source of energy dissipation. Increasing amounts of filler lead to a large
Role of Different Nanoparticles 27

Fig. 14 Variation in tensile


strength and modulus with
increasing nanographite con-
centrations in fluoroelastomer
composites

Fig. 15 Variation in tensile


strength and modulus with
increasing carbon nanofiber
concentrations in chlorobutyl
vulcanizates

number of polymer chains adhering to the polymer surface, thereby leading to a


greater probability of molecular slippage and thus increasing the fracture energy.
From the figures it is also observed that the intensity of increase in tensile strength
28 J. K. Kim et al.

Fig. 16 Variation in modu-


lus and tensile strength with
increasing MWNT concen-
trations in fluoroelastomer
composites

Fig. 17 Mechanical properties of NCTPE gels with MWNT: a tensile strength and b elongation
at break
Role of Different Nanoparticles 29

Fig. 18 a High-resolution SEM microphotograph of exfoliated nanographite sheets; and b high-


resolution SEM microphotograph of nanographite sheets extracted from the polymer compound
after solvent (MEK) extraction) (Reproduced from Ref. [116] with permisssion from Wiley
Interscience)

is more pronounced when compared to modulus. The nanofillers used in the


present study have been modified by acid treatment. This acid treatment of the
nanofillers imparts functional groups like –COOH and –C=O on the surface,
thereby increasing their oxygen functionality. The presence of these oxygen
groups increases the polymer-filler interactions due the formation of complex
physico-chemical bonds between the filler surface and the polymer matrix.
Increased filler loading leads to an increase in polymer-filler interactions, thereby
making a portion of the polymer matrix attach to the filler surface (the so-called
bound rubber phenomenon). SEM microphotographs of raw nanographite powder
and solvent extracted graphite from the polymer compound using methyl ethyl
ketone (MEK) taken at the same length scales are shown in Fig. 18a and b,
respectively. In Fig. 18b, adsorption of polymer chains onto the nanographite
surface can be observed. The mechanical properties of TPE gels with increasing
MWNT concentration are presented in Fig. 17a and b. For comparison, the tensile
strength of triblock copolymer gels is 0.067 MPa and % elongation at break is 82,
which corresponds to zero content in the figure. With a small addition of
MWCNTs in TPE gels, there is a modest increase in tensile strength as the con-
centration increases. This can be explained by the fact that the incorporation of
fillers is a major source of energy dissipation and due to the aspect ratio, thereby
increasing the tensile strength of the TPE gels. As the particle size decreases, large
surface areas become available.

4.3 Fractography

The observation of fracture surface can give new insights regarding the failure
mechanism of the composites. Although there have been many fracture surface
studies in traditional filler (carbon black, carbon silica dual-phase filler, and silica)
reinforced vulcanizates, seldom has it been investigated in nanofiller-reinforced
30 J. K. Kim et al.

elastomers. Figure 19a–d show representative scanning electron micrographs of


the fracture surface of nanographite-reinforced fluoroelastomer composites at
decreasing length scales. The crack growth phenomenon can be considered as a
succession of elementary sequences from initiation to final failure. As the crack tip
is stretched, cavities are induced by de-cohesion (or debonding) between fillers
and the rubber matrix, forming microvoids. When these microvoids reach a critical
volume, they coalesce into microcracks. With further application of stress, these
microcracks propagate perpendicularly to the crack growth direction and eventu-
ally lead to tearing of the samples. Figure 19a reveals the presence of weld lines.
The presence of weld lines can be attributed to parallel orientation of the graphite
platelets in the plane of the weld line, thereby inhibiting the interdiffusion of
polymer chains [122]. SEM micrographs at lower length scales (Fig. 19c, d) show
the presence of aggregated and de-bonded nano-graphite particulates, which
confirms the theory that the failure occurring during deformation of the composites
is primarily via the coalescence of the voids forming critical cracks. From Fig. 19b
it is also observed that the crack surface is rough, which indicates that crack
propagation is difficult.

Fig. 19 Fracture morphology at different length scales in 3.5 phr nanographite-reinforced flu-
oroelastomer composites
Role of Different Nanoparticles 31

Fig. 20 Tensile fracture morphology of MWNT-reinforced fluoroelastomer nanocomposites


(Reproduced from Ref. [117] with permisssion from Wiley Interscience)

Figure 20a and b show representative SEM microphotographs of fractured


surfaces at two different length scales. From Fig. 20a the characteristic fracture
morphology of nanofiller-reinforced elastomers, the so-called ‘‘cross-hatch pat-
tern,’’ can be observed. This morphology is composed of numerous webs and steps
of different sizes in particulate-reinforced vulcanizates. But in our case of MWNT-
reinforced fluoroelastomers, the numerous webs and steps appear to be nearly the
same size. This can be attributed to the mechanism of crack propagation in
MWNT-reinforced elastomers. Recent studies using H-NMR imaging demon-
strated the presence of several microvoids in the elastomer matrix. These small
voids tend to nucleate at the critical stage for crack growth, leading to lower
mechanical properties. Stress–strain forces are amplified near a void, thereby
creating nucleation sites for craze and crack growth. In a typical tensile test the
rubber composite undergoes an immediate elastic deformation, after which the
position of the crack changes slowly and follows a path of least resistance to its
propagation. Increasing filler concentration in the polymer matrix reduces this
propagation.

4.4 Dielectric Relaxation

One of the main advantages of using exfoliated graphite is that it imparts good
electrical conduction and dielectric properties to the composites. Special attention
has focused on these properties. One of the most valuable tools for characterizing
the relaxation behaviour of polymer systems is dielectric relaxation spectroscopy
(DRS). DRS is a useful complement to the more customary mechanical methods of
probing the viscoelastic properties of polymers [123]. Dielectric spectra reflect the
same chain motions as the mechanical modulus; however, there is reduced
interference due to symmetry from shorter time processes, thus making it more
accurate than traditional dynamic mechanical analysis [124].
32 J. K. Kim et al.

4.4.1 Impedance Analysis

Figure 21 shows the variation of the imaginary component of the complex


impedance (Z00 ) with frequency as a function of increasing filler loading in
exfoliated nanographite-reinforced fluoroelastomer vulcanizates. From the figure it
is observed that with an increase in frequency, there is a gradual increase in
complex impedance for all the vulcanizates, reaching a maximum in the region of
10 Hz at low filler loadings (0.5 and 1.5 phr) and around 100 Hz at higher loadings
(2.5 and 3.5 phr). This can be attributed to the secondary relaxation of the polymer
chains of the fluoroelastomer matrix. The observation of this additional peak in the
imaginary component of impedance can also be attributed to the relaxation of the
interfacial region between the polymer and the filler. Numerous examples of
additional damping peaks have been cited in the field of particulate reinforced
polymeric systems [125, 126]. This can be explained on the basis of the
mechanical and viscoelastic properties of cross-linked and reinforced multiphase
polymeric materials. Generally, polymer composites are cross-linked multiphase
materials, the relaxation of which depends on molecular relaxation processes and
morphology. Although these relaxations can usually be associated with each
component, their appearance depends on the chemical and physical interactions of
the two phases (i.e., filler and polymer matrix). The filler used in this study
(exfoliated nanographite, prepared by acid intercalation technique), due to its high
surface activity and the presence of oxygen, shows good interaction with the
polymer matrix, thereby leading to the formation of a strong interphase.

Fig. 21 Variation in com-


plex impedance with increas-
ing nanographite loadings
(Reproduced from Ref. [118]
with permisssion from Wiley
Interscience)
Role of Different Nanoparticles 33

Fig. 22 Variation in real part


of impedance with increasing
nanographite loadings
(Reproduced from Ref. [118]
with permisssion from Wiley
Interscience)

The thickness of this interphase is inversely proportional to the interfacial tension


between the polymeric phases. A substantial portion of the polymer chains is
expected to be immobilized on the filler surface, thus leading to the formation of
regions of spatial heterogeneity. A polymer layer having a higher stiffness than the
bulk polymer in the vicinity of the dispersed phase surface is created from
restricted molecular mobility due to interactions between the phases [127]. So, as
the filler loading increases, more and more polymer is adsorbed onto the filler
surface, thereby making its relaxation difficult. A similar explanation can be
extended to explain the decrease in the real part of impedance with frequency
(Fig. 22). This trend of a continuous drop in impedance with applied frequency is
characteristic of a pure capacitor.

4.4.2 Nyquist Plots

Figure 23 shows the Nyquist plot [the relationship between the imaginary part of
impedance (Z00 ) and the real part of impedance (Z0 )] of fluoroelastomer vulca-
nizates as a function of filler loading. It can be observed that increasing filler in the
composite has a sizable effect on the dielectric properties of this system at all
frequencies. From the figure it can also be observed that, irrespective of the filler
loadings, the plots yield good semicircles, indicating the occurrence of polarization
with a single relaxation time taking place, that is, a local mode process dominated.
However, at higher filler loadings (2.5 and 3.5 phr), the semicircles did not reach
the origin and had a small positive intercept on the Z0 axis, indicating a build-up of
ions at the interphase between the filler and polymer matrix [128].
Several attempts have been made to interpret the impedance spectroscopy of
polymerfiller systems using the resistance–capacitance parallel (R–C) circuit
34 J. K. Kim et al.

Fig. 23 Nyquist plot of fluoroelastomer-nanographite nanocomposites (Reproduced from Ref.


[116] with permisssion from Wiley Interscience)

model. With increasing filler loading, the distance between the aggregates
decreases. This gap can be approximated by a parallel plate capacitor with an area
A, separation distance (d), and capacitance (C) eA/x, where e is the dielectric
constant of the polymer. Each filler aggregate has a resistance Ra, the resistance
within the aggregate. The impedance in a composite can be written as:

Rcs xR2cs Cs
Z ¼ Ras þ 2 2 2
j ð1Þ
1 þ x Rcs Cs 1 þ x2 R2cs þ Cs2
The respective imaginary and real parts of impedance can be expresses as:
Rcs
Z 0 ¼ Ras þ
1 þ x2 R2cs Cs2

and

xR2cs Cs
Z 00 ¼  ð2Þ
1 þ x2 R2cs Cs2

and the dielectric loss tangent can be expressed as:

Z 00 xR2cs Cs
tan d ¼  0
¼ ð3Þ
Z Ras þ Rcs þ x2 R2cs Ras Cs2
Role of Different Nanoparticles 35

Table 3 Radius and center


Filler loading (phr) Center (x, 0) Radius
for Nyquist plots in
nanographite-reinforced 0.5 3.424 9 105, 0 1.98 9 105
fluoroelastomer vulcanizates 1.5 2.273 9 105, 0 0.87 9 105
2.5 4.358 9 104, 0 4.42 9 104
3.5 4.462 9 104, 0 4.34 9 104

From above equations, the relationship between Z0 and Z00 is:


   2
0 2Ras þ Rcs 2 2 Rcs
Z  þZ2 ¼ ð4Þ
2 2
Therefore, a plot of Z00 and Z0 will give a half circle that has its center at
{(2Ras ? Rcs)/2, 0} and a radius of Rcs/2. Wang et al. [129] proposed that because
the circular curve of Z00 versus Z0 occurs only for the parallel resistor circuit, then
the above analysis can be used to confirm the existence of the capacitor effect. The
capacitor effect also confirms that the gaps between the nanographite control
electron conduction via non-Ohmic contacts between the filler aggregates. The
variation in the values of radius and center of the half-circle can also be used as a
measure of the gaps between the filler aggregates. Using the above equations, the
radius and center have been calculated and are tabulated in Table 3. It can be
observed that with increasing filler loadings, the radius reduces and the center
shifts to lower values, which implies increasing capacitance.

4.4.3 Percolation

According to classical percolation theory, the percolation limit of any property of


the material MP in a polymer reinforced with a filler and the volume fraction f of
particles in the mixture above percolation threshold f c can be expressed as:

MP ¼ MP0 ðf  fc Þt ð5Þ
where t is the critical exponent that is related to the dimensionality of the system
and MP0 is the material property in pure polymer.
Changing Eq. 5 to log scale, one can calculate the value of critical exponent t.
A system obeying the percolation model should give a straight line:
log MP ¼ t logðf  fc Þ þ log MP0 ð6Þ
Two material properties—conductivity as obtained from dielectric relaxation
spectroscopy (at various frequencies) and modulus values obtained from stress–
strain measurements—have been used to study the percolation phenomenon and
are plotted in Figs. 24 and 25, respectively.
From Figs. 24 and 25 it can be observed that percolation is occurring at 2.5 phr
in electrical conductivity (at all tested frequencies), above 1.5 phr in storage
36 J. K. Kim et al.

Fig. 24 Variation in AC
conductivity with increasing
nanographite loadings mea-
sured at increasing frequen-
cies (Reproduced from Ref.
[116] with permisssion from
Wiley Interscience)

Fig. 25 Variation in modu-


lus at increasing nanographite
loadings (Reproduced from
Ref. [118] with permisssion
from Wiley Interscience)

modulus (at all temperatures tested), and above 2.5 phr in modulus as obtained
from stress–strain curves. The values of the critical exponent t are shown in
Table 4.
From Table 4 it is observed that the value of the critical exponent t decreases
with both increasing frequency and increasing temperature (as in storage modu-
lus). A similar observation of a decrease in the value of t with increasing tem-
perature was reported by Levya et al. [130], who attributed this phenomenon to the
Role of Different Nanoparticles 37

Table 4 Value of critical


Property Test conditions Exponent t
exponent t
Conductivity from DRS 100 Hz 2.67
1,000 Hz 2.54
10,000 Hz 2.36
100,000 Hz 2.17
Modulus from stress–strain +22C 3.03

occurrence of multiple percolation thresholds. A similar observation in carbon-


black-reinforced polymer blends was also reported by Levon et al. [131] and
theoretically proven by Roberts et al. [132].

4.5 Dynamic Mechanical Properties

One of the main usages of butyl and halobutyl elastomers is as inner tubes in tires
and vibration-isolating applications in the automotive industry. Therefore, in the
present study, special attention focused on the dynamic mechanical properties of
VGCNF-reinforced chlorobutyl vulcanizates.

4.5.1 Effect of Temperature on Loss Tangent

Figure 26 shows the loss tangent spectra of chlorobutyl vulcanizates reinforced


with VGCNF as a function of temperature. Increasing the amount of filler in the
vulcanizates has no significant effect on the temperature or location of the max-
imum value of loss tangent (tandmax), but the magnitude of the peak decreases
increasing filler loading. All samples showed the glass transition in the range of
-29 to -34C.
When a polymer is cooled through the glass transition region, the physical
properties of the polymer in the non-equilibrium state (at temperature less than Tg),
such as volume and enthalpy, gradually recover to new equilibrium values through
configurational rearrangement of the polymer segments. The rate of the rear-
rangement or relaxation process depends on the local environment surrounding the
relaxation entities and hence reflects the extent of environmental restriction on
those entities. However, the magnitude of the Tg shift is marginal in our case. This
can be explained on the basis of the ‘‘sluggish’’ nature of polyisobutylene relax-
ation dynamics. The glass-to-liquid transition primarily involves correlated local
motions of a few backbone chains [124]. However, the segmental relaxation
in polyisobutylene- (PIB-) based elastomers is weak, due to the steric hindrance
by the methyl side groups. It has been suggested that low barriers to internal
rotation about the skeletal bonds are a contributing factor in maintaining the
segmental mobility necessary for elastomeric behavior. In polyisobutylene (PIB)
(–C–(CH3)2–CH2–), the methyl groups bonded to alternate chain carbon atoms
38 J. K. Kim et al.

Fig. 26 Variation in loss


tangent as a function of tem-
perature in VGCNF-rein-
forced chlorobutyl
vulcanizates

produce steric crowding. Although partially relieved by distortion of the (–C–


CH2–C–) and (–CH2–C–CH2–) bond angles (127 and 109, respectively), it is still
significant and hence PIB is not very flexible. Because butyl and halobutyl elas-
tomers are copolymers of 97 wt% isobutylene and 3% isoprene, the relaxation in
CIIR (Chlorobutyl Rubber) is predominantly dominated by the polyisobutylene
chains. Further, the distribution of the end-to-end distance of chain segments
between topologically adjacent cross-links plays a crucial role in the relaxation
process [133]. Because the isobutylene content of CIIR has no chemically reactive
site (unsaturated bond or double bond), and the only available reaction sites are Cl
(1.35%) and the unsaturation of isoprene, the average chain length between the
two adjacent reactive sites is very large. Consequently, the vulcanization network
in CIIR is expected to be more homogeneous than those in other rubbers; thereby,
the glass transition solely depends on polyisobutylene chain dynamics. Beyond the
glass transition temperature (i.e., in the rubbery region) at higher VGCNF loadings
(15 and 10 phr), it is observed that there is a small hump in the loss tangent curves.
This can be attributed to the relaxation of the interfacial region of the polymer-
filler in the composites. A similar behaviour is also observed in carbon-black-filled
bromobutyl, chlorobutyl, and styrene butadiene rubbers where a similar phe-
nomenon has been observed [134, 135].

4.5.2 Effect of Temperature on Storage and Loss Modulus

The variation of storage modulus with temperature for elastomeric composites is a


very important design criterion. Lewis and Nielsen [125] have pointed out that an
elastomer composition with dynamic mechanical properties less sensitive to
temperature changes is required for vibration isolation applications. Figure 27
shows the variation in storage modulus (E0 ) for different vulcanizates over a
Role of Different Nanoparticles 39

Fig. 27 Variation in storage


modulus as a function of
temperature in VGCNF-rein-
forced chlorobutyl
vulcanizates

temperature range of -80 to +100C. In the glassy region (-80 to –30C), the
augmentation in storage modulus E0 in the composite with the addition of fillers
can be attributed to the hydrodynamic effect of the filler particle embedded in the
polymer continuum. In the rubbery region (-30 to 100C), the polymer-filler and
filler-filler networking (i.e., structure) and aggregate interactions have a pro-
nounced effect on the value of E0 . Carbonaceous fillers such as carbon blacks and
carbon silica dual-phase fillers affect the elastic properties of the vulcanizates
more than the hydrodynamic reinforcement. The filler used in the present study,
VGCNF, is also carbon-based filler that has been oxidized by treating with acid.
This acid treatment increases the oxygen functionality of the fillers and has a
profound effect on polymer-filler interactions, as widely reported by Serizawa
et al. [136] in isobutylenebased elastomers.
Figure 28 represents the temperature dependence of the loss modulus (E00 ) for
various compositions. For all filled compositions, a distinct transition peak
(a transition) is observed around -50C that can be attributed to conformational
transitions occurring in the polyisobutylene backbone caused by micro-Brownian
motions. The amount of filler has no significant effect on a peak location and
a
intensity. The amount of filler does not affect the TE00 max temperature (temper-
ature at a peak); however, the intensity of a peak increases slightly with filler
loading. Higher filler loadings result in a percolated network of filler particles
that can influence relaxation on different scales. The percolation effect is usually
considered effective for relaxation of longer time scales, such as the terminal
relaxation observed in some rheological measurements. Recent dynamic
mechanical experiments for composite solids seem to indicate that restriction
effects do, in fact, result from the formation of a percolation network [137].
Nevertheless, reports regarding the restriction effects of percolated networks on
the segmental relaxation are not fully conclusive, and the issue awaits more
detailed and systematic study.
40 J. K. Kim et al.

Fig. 28 Variation in loss


modulus as a function of
temperature in VGCNF-rein-
forced chlorobutyl
vulcanizates

4.5.3 Strain-Dependent Dynamic Mechanical Properties

Although there have been many reports about strain-dependent dynamic properties
(Payne effect) in vulcanizates reinforced with carbon-based fillers such as carbon
blacks and carbon silica dual-phase fillers, it has never been reported in VGCNF-
reinforced elastomers. Because a typical elastomeric composite must withstand a
wide range of temperatures in many practical applications, the Payne effect was
studied at three temperatures in this study: -30C (near the glass transition
temperature), room temperature (+30C), and high temperature (+70C).

Effect of Strain on tand and Storage Modulus at -30C

The variation in tand and storage modulus with strain at -30C is shown in
Fig. 29a and b, respectively. The storage modulus decreases and increases with
increasing strain for all the compounds. All the vulcanizates show highest storage
modulus (E0 ) at the lower strains. At the lowest strain of 0.07%, the three-
dimensional filler-filler and filler-polymer structure acts as a rigid unit against the
imposed strain and hence gives the highest modulus value.
The strain input associated at low strains is not sufficient to cause any signif-
icant change in network structure. The reinforcement at a moderate strain (\2%) is
greatly affected by a disruption of the continuous network of filler that interpen-
etrates the rubber matrix. Under application of strain, the molecules of smaller
chain lengths between the densely packed network points get oriented and form
crystallites, whereas the molecules of much longer chain lengths would remain in
random coil states. CIIR seldom exhibits strain-induced crystallization at normal
operating temperatures, but recent studies have shown that butyl elastomers are
capable of forming oriented lamellar structures at low temperatures [138]. They
form a relatively high oriented amorphous fraction (*50%). The strain-induced
Role of Different Nanoparticles 41

Fig. 29 Effect of dynamic


strain amplitude on a loss
tangent and b storage modu-
lus in VGCNF-reinforced
chlorobutyl vulcanizates at -
30C at increasing filler
loadings

crystallites are in the form of a microfibrillar structure containing extended chain


conformation. These crystallites are connected by oriented amorphous tie chains,
which form a new network structure that is immobilized in a large volume of
random amorphous chains. Luch et al. showed the disappearance of lamellar
crystallites with increasing cross-link density, whereas recently Shimizu et al.
showed that in uncrosslinked natural rubber, there is a growth of lamellar crys-
tallites [139, 140]. It must be mentioned that the direction of orientation of these
crystallites is controversial. Keller theoretically showed that the strain induced
crystallization in polymers grow in the direction parallel to the chain axis, but
Conradt showed the development of lamellar crystallites in the direction perpen-
dicular to the molecular chain axis [141, 142]. However, the general consensus is
that the strain-induced crystallization increases the storage modulus. Moreover, the
42 J. K. Kim et al.

filler particles are anisotropic in nature and get oriented in the direction of applied
strain, thus leading to higher storage modulus values.

Effect of Strain on tan d and Storage Modulus at Room Temperature (+30C)

The variation in tand and storage modulus with strain at room temperature (+30C)
is shown in Fig. 30a and b, respectively. The familiar phenomenon of a decrease in
storage modulus and an increase in tand with increasing strains for all the com-
pounds is observed. With increasing strain, it is observed that there is a decrease in
the storage modulus. This can be explained on the basis of polymer-filler inter-
actions, the desorption and reabsorption of the hard rubber shell surrounding the
filler aggregate, or breaking and reforming of the effective cross-link in the vul-
canizates-forming transition zone between the bound rubber and the bulk rubber.

Fig. 30 Effect of dynamic


strain amplitude on a loss
tangent and b storage modu-
lus in VGCNF-reinforced
chlorobutyl vulcanizates at
30C at increasing filler
loadings
Role of Different Nanoparticles 43

At lower strains, the three-dimensional filler-filler and filler-polymer structure


acts as a rigid unit against the imposed strain and hence will give a higher mod-
ulus. The strain input associated with low strains is not sufficient to cause any
significant change in network structure. However, at higher strains, this secondary
structure tends to break down. Some researchers tend to equate ‘‘microstructure’’
with ‘‘agglomeration’’ or filler networking, and therefore filler agglomeration is
also widely cited for the nonlinear behavior of filled polymer melts. Many
researchers showed that the chemical structure of the polymer chains and fillers
have a profound influence on relaxation behavior.
Many studies have shown that the surface oxidation of carbon-based fillers leads
to increased interactions with the polymer matrix [143]. At higher filler loadings,
there may be two competitive phenomena occurring. Higher filler loading will give
rise to the availability of a larger surface area of the fillers, thereby increasing the
polymer-filler interactions. On the other hand, the secondary aggregate structures also
increase at higher loadings, especially at 15 phr. Because the inter-particle potential is
a strong function of the separation distance, this induces a steep dependence of the
energy stored in the network of the particles on the macroscopic strain.

Effect of Strain on tan d and Storage Modulus at +70C

Figure 31a and b show the variation of tand and storage modulus with strain at
+70C. There is a decrease in storage modulus with increasing strain. But at all filler
loadings, the decrease is much more pronounced when compared to -30C and
room temperature (+30C). Increased filler loading in the composites form sec-
ondary aggregate structures. This may lead to new damping mechanisms such as
particle–particle friction, particle-polymer friction where there is essentially no
adhesion at the interface, and excess damping at the interface due to induced
thermal stresses or changes in polymer conformation or morphology. Some recent
studies show that the linear viscoelastic behavior of filled elastomers is governed by
the glass transition gradient near the particle surface, combined with the arrange-
ment of filler particles within the elastomermatrix [144]. If an increasing strain is
applied to a reinforced system, its dynamic modulus decreases while the modulus of
the nonreinforced matrix remains constant. Thus, the strain-softening of the glassy
shell surrounding the solid particles is responsible for the Payne effect.

4.6 Rheological Properties

One of the most important properties of TPE gels is the MOT (maximum operating
temperature), which is characterized by the transition from a solid-like state to a
viscoelastic liquid state. A rheological technique has been utilized in this study to
find the MOT of MWNT-reinforced TPE gels. Hybrid gels were subjected to
dynamic mechanical studies to discern the effect of nanoparticles on the
44 J. K. Kim et al.

Fig. 31 Effect of dynamic


strain amplitude on a loss
tangent and b storage modu-
lus in VGCNF-reinforced
chlorobutyl vulcanizates at
70C at increasing filler
loadings

rheological behavior of hybrid triblock copolymer gels using a strain-controlled


Rheometrics Mechanical Spectrometer (RMS 800, Rheometric Scientific, USA.),
operated with 25-mm parallel plate geometry and a 1.5- or 2.5-mm gap height in
the temperature range between 30 and 140C. The oscillatory shear measurements
focused on the variations of elastic (in-phase) G0 the out-of-phase, G00 , as a
function of temperature and frequency.

4.6.1 Effect of Frequency

Measurements of the real and imaginary parts, G0 and G00 , of the shear modulus
were also made as a function of frequency, w, of a small deformation strain of 1%
Role of Different Nanoparticles 45

for nanoparticles. The effect of frequency of NCTPE gels containing different


amounts of MWCNTs is shown in Fig. 32. The results show that G0 exceeds G00
over the entire experimental frequency range, and G0 depends on the frequency at
this 1% strain, which is typical of a physical gel. Here, we mean physical gels as a
liquid-rich system exhibiting solid-like behavior, which the characteristics of
showing flat mechanical spectrum in an oscillatory shear experiment [145]. The
slight increase in G0 as the frequency increases may still be considered negligible.
It can also be seen from Fig. 32 that the modulus increases as the concentration
increases. The effect of frequency at higher temperature was also studied. The
onset of maximum operating temperature, such as 70C for MWCNTs and 90C
for parent TPE gels, was chosen to investigate the effect of temperature as a
function of frequency (Fig. 33a, b). A further increase in temperature causes the
moduli to become frequency dependent with increasing w. This only implies that
the gel is no longer physical gel at these temperatures and the system loses its
elasticity. In the case of NCTPE gels with MWCNTs, G00 is lower than G0 at lower
frequency and then crossover in G0 and G00 is also observed as frequency increases,
thereby indicating that the gels become viscous liquids. Comparable behavior was
observed in TPE gels with 20% SEBS concentration, suggesting that just below
this temperature, the gel network reinforcing TPE gels undergo some supramo-
lecular rearrangement, perhaps similar to an orderdisorder transition (ODT). Note
that the loss of gel behavior (i.e., the onset of frequency dependence in G0 ) is most
pronounced at low frequencies, as is typically observed for physical gels.

Fig. 32 Storage modulus G0


(closed symbols) and loss
modulus G00 (open symbols)
of NCTPE gels with
MWCNTs presented as a
function of oscillatory fre-
quency (w) at a strain ampli-
tude of 1% at 30C
46 J. K. Kim et al.

Fig. 33 Variation in storage modulus G0 (closed symbols) and G00 (open symbols) as a function
of oscillating frequency at 70C for (a) 5 wt% MWNT and (b) 20 wt% MWNT loaded TPE gels

4.6.2 Effect of Temperature

Figure 34 shows the change in G0 with temperature heated throughout the melting
range for NCTPE gels with 5 wt% MWCNTs and parent TPE gels (Fig. 35). The
parent TPE gel consists of four characteristics region: (1) an initial plateau over
which G0 remains relatively constant or slightly increases with increasing tem-
perature, which means that the rubbery PS domains become glassy; (2) an abrupt
reduction in G0 attributed to a lattice disordering transition in the vicinity of 75C;
(3) a small second plateau in G0 at 100C; and (4) a steep decrease in G0 , signifying

Fig. 34 Dynamic storage


modulus as a function of
temperature at strain ampli-
tude of 1% for parent TPE
gels (squares) and NCTPE
gels with 5 wt% MWCNT
(circles)
Role of Different Nanoparticles 47

Fig. 35 Dynamic storage


modulus G0 (closed symbols)
of NCTPE gels with
MWCNTs of 0.5 wt% (hexa-
gon), 1 wt% (triangle), 3 wt%
(down triangle), and 5 wt%
(star) and dynamic loss
modulus (G00 ) (open symbols)

the collapse of the copolymer network at ca. 108C. The existence of a second
plateau reflects the persistence of a residual, load-bearing nanostructure that either
remains intact or develops upon increasing the temperature. Montensen et al. [146]
have demonstrated that similar TPE gels exhibit a high-temperature, body-
centered cubic phase that possesses an unusually high degree of nanostructured
order. The nanosystem TPE gels have only three major characteristic regions
without the second plateau observed in the parent TPE gels, that is, without any
morphological change during the heating procedure. The precipitous decrease in
G0 at very low temperature was observed for all NCTPE gels, which are much
lower than their parent TPE gels. An abrupt reduction in G0 was reduced in the
vicinity of 75C. This reduction in G0 indicates the collapse of the triblock
copolymer network in which the lattice disordering transition. If the MWCNT
loading increases, G0 also increases.

4.6.3 Effect of MWNT Concentration on the Storage Modulus

At a temperature between 30 and 40C below the gel point, the NCTPE gels have
the property of elastic moduli, where G0 [ G0 . This phenomenon was found at
every nanoparticle composition. This indicates that at ambient temperature, a
physical network is still present despite the addition of nanoparticles. In this
regard, the values of the plateau G0 extracted from the data between 30 and 40C is
displayed in Fig. 35 as a function of nanoparticle. The parent TPE gel with 20 wt%
SEBS concentration was included here as reference material for comparison.
48 J. K. Kim et al.

Low concentrations yield low G0 values with their parent TPE gels. MWCNTs are
probably flocculated into large-scale aggregates due to high solvent content and
hinder to a lesser extent the bridging efficacy of individual copolymer molecules,
which are possibly dictated by dispersion problems encountered, and are further
explained in morphological observation. The further addition of 3–5 wt% con-
centrations to SEBS/oil gel likewise promotes a modest increase in G0 , which may
be due to the high aspect ratio nature of this material.

4.7 Thermal Property

Thermo-gravimetry (TG) is a suitable method to evaluate the thermal degradation


properties of elastomers. The thermal stability of TPE has been extensively studied
because of the great importance of this group of materials. Reinforcing the inor-
ganic filler improves the thermal stability of thermoplastic elastomers and plastics.
Because the reinforcing of nanoparticles improves several properties, which
promises wide applicability, it would be interesting to study the thermal stability of
NCTPE gels. The derivative thermogravimetric (DTG) and thermo-gravimetric
(TG) curves for NCTPE gels with MWCNTs and parent TPE gels are shown in
Fig. 36a and b. The TG curve has two distinct regions of weight loss that are
reflected in two peaks (maxima) in the DTG curve, implying that at least two
stages of degradation occur in this sample. The initial degradation in stage I results
primarily from the decomposition of the swollen EB (ethylene butylene) block in
hydrocarbon oil, while stage II is due to the SEBS matrix. For each sample, the

Fig. 36 a DTG and b TGA of NCTPE gels with increasing amounts of MWNTs in the parent
TPEG
Role of Different Nanoparticles 49

thermogram revealed that the DTG plot (Fig. 36a) shows a maximum rate of
weight loss (Fig. 36b). It can be clearly observed from these figures the thermo-
gram shift toward higher temperature as the heating rate increases. The shift of
temperature is more pronounced at lower temperatures.
It appears that the particles reside in the region of the EB block swollen by a
high content of oil, as seen in TEM image, thereby increasing the distinct region of
oil degradation temperature. It can also explained that may be due to the filler
effect that is prominent and retards the degradation of TPE gels and the retardation
effect is attributed to the interaction between nanoparticles and macroradicals
generated during the degradation process. However, the second peak that appears
at higher temperature remains constant with addition of MWCNTs. The weight
loss for swollen EB containing oil and SEBS matrix remains constant with the
addition of MWCNTs. The residual yields of the NCTPE gels increased with
increasing MWCNT content, indicating that thermal decomposition of the polymer
matrix was retarded in the NCTPE gels with higher residual yield. This result may
be attributed to a physical barrier effect due to the fact that MWCNTs would
prevent the transport of decomposition products in the polymer nanocomposites.
Similar observations have been reported that the thermal stability of polypropyl-
ene/layered silicate nanocomposites was improved by a physical barrier effect;
hence it is enhanced by ablative reassembling of the silicate layer [147]. Com-
paring the residue of TPE gels without MWCNTs, there is little residue because
the component of the gels consists only of carbon and hydrogen.

References

1. Kraus, G.: Reinforcement in Elastomers. Wiley-Interscience, New York (1965)


2. Morton-Jones, D.H.: Polymer Processing. Chapman and Hall, London (1989)
3. Ismail, H., Poh, B.T., Tan, K.S., Moorthy, M.: Effect of filler loading on cure time and
swelling behaviour of SMR L/ENR 25 and SMR L/SBR blends. Polym. Int. 52, 685 (2003)
4. Ismail, H., Poh, B.T., Tan, K.S.: Effect of filler loading on tensile and tear properties of
SMR L/ENR 25 and SMR L/SBR blends cured via a semi-efficient vulcanization system.
Polym. Test 21, 801 (2002)
5. Spurr, R.T.: Fillers in friction materials. Wear 22, 367 (1972)
6. Debnath, S.: Mica as reinforcing filler in elastomeric composites. Ph.D. thesis, Rubber
Technology Center, IIT Kharagpur, India (2000)
7. Bajaj, P., Babu, G.N., Khanna, D.N., Varshney, S.K.: Room temperature-vulcanized
silicone elastomer: effect of curing conditions and the nature of filler on mechanical and
thermal properties. J. Appl. Polym. Sci. 23, 3505 (1979)
8. Kwei, T.K., Kunins, C.A.: Polymer-filler interaction: vapor sorption studies. J. Appl. Polym.
Sci. 8, 1483 (1964)
9. Krysztafkiewicz, A.: Modified calcium silicates as active rubber fillers. J. Mater. Sci. 22,
478 (1987)
10. Varma, A.J., Deshpande, M.D., Nadkarni, V.M.: Morphology and mechanical properties of
silicate filled polyurethane elastomers based on castor oil and polymeric MDI. Angew.
Makromol Chem. 132, 203 (1985)
50 J. K. Kim et al.

11. Shepard, N.A., Street, J.R., Park, C.R.: Fillers and reinforcing agents. In: Davis, C.C.,
Blake, J.T. (eds.) ACS Monograph Series. New York, Chap 11 (1937)
12. Hamed, G.R., Park, B.H.: The mechanism of carbon black reinforcement of SBR and NR
vulcanizates. Rubber Chem. Technol. 72, 946 (1999)
13. Medalia, A.I., Kraus, G.: Reinforcement of elastomers by particulate fillers. In: Mark, J.E.,
Erman, B., Eirich, F.R. (eds.) Science and Technology of Rubber, p. 378. Academic Press,
San Diego (1994)
14. Leblanc, J.L.: Rubber-filler interactions and rheological properties in filled compounds.
Prog. Polym. Sci. 27, 627 (2002)
15. Schmidt, E.: Effect of colloidal noncarbon pigments on elastomer properties. Ind. Eng.
Chem. 43, 679 (1951)
16. Studebaker, M.L.: A useful new system for categorizing rubber grade carbon blacks.
Technical Bulletin-Phillips Petroleum Co (1971)
17. Wagner, M.P.: Non black fillers in elastomers. Rubber World 164, 46 (1941)
18. Fowkes, F.M.: Attractive forces at interfaces. Ind. Eng. Chem. 56, 40 (1964)
19. Gent, A.N., Tompkins, D.A.: Nucleation and growth of gas bubbles in elastomers. J. Appl.
Phys. 40, 2520 (1969)
20. Gent, A.N.: Detachment of an elastic matrix from a rigid spherical inclusion. J. Mat. Sci. 15,
2884 (1980)
21. Huber, G., Vilgis, T.A.: On the mechanism of hydrodynamic reinforcement in elastic
composites. Macromolecules 35, 9204 (2002)
22. Heinrich, G., Kluppel, M., Vilgis, T.A.: Reinforcement of elastomers. Curr. Opin. Solid
State Mater. Sci. 6, 195 (2002)
23. Kroto, H.W., Heath, J.R., O’Brien, S.C., Curl, R.F., Smalley, R.E.: C60:
buckminsterfullerene. Nature 318, 162 (1985)
24. Ijima, S.: Helical microtubules of graphitic carbon. Nature 354, 56 (1991)
25. Hutchison, J.L., Kiselev, N.A., Krinichnaya, E.P., Krestinin, A.V., Loutfy, R.O., Morawsky,
A.P., Muradyan, V.E., Obraztsova, E.D., Sloan, J., Terekhov, S.V., Zakharov, D.N.:
Double-walled carbon nanotubes fabricated by a hydrogen arc discharge method. Carbon
39, 761 (2001)
26. Saito, Y., Nakahira, T., Uemura, S.: Growth conditions of double-walled carbon nanotubes
in arc discharge. J. Phys. Chem. B 107, 931 (2003)
27. Arepalli, S.: Laser ablation process for single-walled carbon nanotube production.
J. Nanosci. Nanotech. 4, 317 (2004)
28. Jiang, W., Molian, P., Ferkel, H.: Rapid production of carbon nanotubes by high-power laser
ablation. J. Manuf. Sci. Eng. 127, 703 (2005)
29. Endo, M., Hayashi, T., Kim, Y.A., Muramatsu, H.: Development and application of carbon
nanotubes. Jpn. J. Appl. Phys. 45, 4883 (2006)
30. Hughes, T.V., Chambers, C.R.: Manufacture of carbon filaments. US Patent 405480 (1889)
31. Robertson, S.D.: Graphite formation from low temperature pyrolysis of methane over some
transition metal surfaces. Nature 221, 1044 (1969)
32. Audier, M., Oberlin, A., Coulon, A.: Crystallographic orientations of catalytic particles in
filamentous carbon; case of simple conical particles. J. Cryst. Growth 55, 549 (1981)
33. Schouten, F.C., Kaleveld, E.W., Bootsma, G.A.: AES-LEED-ellipsometry study of the
kinetics of the interaction of methane with Ni(110). Surf. Sci. 63, 460 (1977)
34. Zaikovskii, V.I., Chesnokov, V.V., Buyanov, R.A.: High-resolution electron microscopic
study of the structure of filamentary carbon on iron and nickel catalysts. Appl. Catal. 38, 41
(1988)
35. Fisa, B., Sanschagrin, B., Favis, B.: Mechanical degradation of mica during processing with
polypropylene. Polym. Compos. 5, 264 (1984)
36. Kamisnki, S.J.: Resinous polymer sheet materials having selective, decorative effects. US
Patent No. 4126727 (1976)
37. Garton, A.: Characterization of the interface: some initial results for mica-polypropylene
composites. Polym. Compos. 3, 189 (1982)
Role of Different Nanoparticles 51

38. Favis, B.D., Blanchard, L.P., Leonard, J., Prudapos-Homme, R.E.: The formation of
coupling agent monolayers on the surface of mica. Polym. Compos. 5, 11 (1984)
39. Tausz, S.E., Chaffey, C.E.: Ultrasonically delaminated and coarse mica particles as
reinforcements for polypropylene. J. Appl. Polym. Sci. 27, 4493 (1982)
40. Eldred, R.J.: Dimensional control of swelling in nitrile elastomers by anisometric fillers.
Rubber World 188, 26 (1983)
41. Chen, C.H., Ma, C.C.M.: Pultruded fiber reinforced blocked polyurethane (PU) composites.
II. Processing variables and dynamic mechanical properties. J. Appl. Polym. Sci. 46, 949
(1992)
42. Jilken, L., Malhammar, G., Selden, R.: The effect of mineral fillers on impact and tensile
properties of polypropylene. Polym. Test 10, 329 (1991)
43. Yoshikawa, M., Niwa, H., Fukuura, Y., Naito, K.: Low permeable nitrile rubber hose. US
Patent No. 5476121 (1995)
44. Debnath, S., De, S.K., Khastgir, D.: Effect of silane coupling agent on vulcanization,
network structure, polymer-filler interaction, physical properties and failure mode of mica-
filled styrene-butadiene rubber. J. Mater. Sci. 22, 4453 (1987)
45. Brown, J.M., Curliss, D., Vaia, R.A.: Thermoset-layered silicate nanocomposites.
Quaternary ammonium montmorillonite with primary diamine cured epoxies. Chem.
Mater. 12, 3376 (2000)
46. Zanetti, M., Camino, G., Canavese, D., Morgan, A.B., Lamelas, F.J., Wilkie, C.A.: Fire
retardant halogen-antimony-clay synergism in polypropylene layered silicate nanocomposites.
Chem. Mater. 14, 189 (2002)
47. Chang, J.H., Jang, T.G., Ihn, K.J., Lee, W.K., Sur, G.S.: Poly(vinyl alcohol) nanocomposites
with different clays: pristine clays and organoclays. J. Appl. Polym. Sci. 90, 3208 (2003)
48. Kanzaki, Y., Hayashi, M., Minami, C., Inoue, Y., Kogure, M., Watanabe, Y., Tanaka, T.:
Intercalation and bilayer formation of phospholipids in layered synthetic mica-2. Solvent
effect of the intercalation reaction of natural and reduced-type phosphatidylcholines.
Langmuir 13, 3674 (1997)
49. Carrado, K.A., Xu, L., Csencsits, R., Muntean, J.V.: Use of organo- and alkoxysilanes in the
synthesis of grafted and pristine clays. Chem. Mater. 13, 3766 (2001)
50. Osman, M.A., Mittal, V., Morbidelli, M., Suter, U.W.: Polyurethane adhesive
nanocomposites as gas permeation barrier. Macromolecules 36, 9851 (2003)
51. Liu, L., Qi, Z., Zhu, X.: Studies on nylon 6/clay nanocomposites by melt-intercalation
process. J. Appl. Polym. Sci. 71, 1133 (1999)
52. Alexandre, M., Dubois, P.: Polymer-layered silicate nanocomposites: preparation,
properties and uses of a new class of materials. Mater. Sci. Eng. R 28, 1 (2000)
53. Lee, D.C., Jang, L.K.: Preparation and characterization of PMMA-clay hybrid composite by
emulsion polymerization. J. Appl. Polym. Sci. 61, 1117 (1996)
54. Vaia, R.A., Vasudevan, S., Krawiec, W., Scanlon, L.G., Giannelis, E.P.: New polymer
electrolyte nanocomposites: melt intercalation of poly(ethylene oxide) in mica-type
silicates. Adv. Mater. 7, 154 (1995)
55. Hitzky, E.R., Pilar, A.: Polymer-salt intercalation complexes in layer silicates. Adv. Mater.
2, 545 (1990)
56. Liu, L., Qi, Z., Zhu, X.: Studies on nylon 6/clay nanocomposites by melt-intercalation
process. J. Appl. Polym. Sci. 1133 (1999)
57. Messersmith, P.B., Stupp, S.I.: Synthesis of nanocomposites: organoceramics. J. Mater. Res.
7, 2599 (1992)
58. Okada, A., Fukumori, A., Usuki, A., Kojima, Y., Kuruachi, T., Kamigaito, O.: Rubber-clay
hybrid synthesis and properties. Polym. Prep. 32, 540 (1991)
59. Wang, Z., Pinnavaia, T.J.: Nanolayer reinforcement of elastomeric polyurethane. Chem.
Mater. 10, 3769 (1998)
60. Joly, S., Garnaud, G., Ollitrault, R., Bokobza, L., Mark, J.E.: Organically modified layered
silicates as reinforcing fillers for natural rubber. Chem. Mater. 14, 4202 (2002)
52 J. K. Kim et al.

61. Varghese, S., Karger-Kocsis, J., Gatos, K.G.: Melt compounded epoxidized natural rubber/
layered silicate nanocomposites: structure-properties relationships. Polymer 44, 3977 (2003)
62. Arroyo, M., Lopez-Manchado, M.A., Herrero, B.: Organo-montmorillonite as substitute of
carbon black in natural rubber compounds. Polymer 44, 2447 (2003)
63. Mousa, A., Karger-Kocsis, J.: Rheological and thermodynamical behavior of styrene/
butadiene rubber-organoclay nanocomposites. Macromol. Mater. Eng. 286, 260 (2001)
64. Kojima, Y., Fukumori, K., Usuki, A., Okada, A., Kurauchi, T.: Gas permeabilities in
rubberclay hybrid. Mater. Sci. Lett. 12, 12 (1993)
65. Chang, Y.W., Yang, Y., Ryu, S., Nah, C.: Preparation and properties of EPDM/
organomontmorillonite hybrid nanocomposites. Polym. Int. 51, 319 (2002)
66. Wang, S., Long, C., Wang, X., Li, Q., Qi, Z.: Synthesis and properties of silicone rubber/
organomontmorillonite hybrid nanocomposites. J. Appl. Polym. Sci. 69, 1557 (1998)
67. Kader, M.A., Nah, C.: Influence of clay on the vulcanization kinetics of fluoroelastomer
nanocomposites. Polymer 45, 2237 (2004)
68. Kokabi, M., Razzaghi-Kashani, M., Hasankhani, H.: Improvement in physical and
mechanical properties of butyl rubber with montmorillonite organo-clay. Ir. J. Polym.
Sci. 16, 10 (2007)
69. Sridhar, V., Chaudhary, R.N.P., Tripathy, D.K.: Effect of fillers on the relaxation behavior
of chlorobutyl vulcanizates. J. Appl. Polym. Sci. 100, 3161 (2006)
70. Novak, I., Krupa, I.: Electro-conductive resins filled with graphite for casting applications.
Eur. Polym. J. 40, 1417 (2004)
71. Coleman, J.N., Khan, U., Gunko, Y.K.: Mechanical reinforcement of polymers using carbon
nanotubes. Adv. Mater. 18, 689 (2006)
72. Monte, S.J.: Titanate coupling agents. In: Xanthos, M. (ed.) Functional Fillers for Plastics.
Wiley-VCH Verlag GmbH (2005)
73. Lyatskaya, Y., Balazs, A.C.: Modeling the phase behavior of polymer-clay composites.
Macromolecules 31, 6676 (1998)
74. Ginzburg, V.V., Balazs, A.C.: Calculating phase diagrams of polymer-platelet mixtures
using density functional theory: implications for polymer/clay composites. Macromolecules
32, 5681 (1999)
75. Ginzburg, V.V., Singh, C., Balazs, A.C.: Theoretical phase diagrams of polymer/clay
composites: the role of grafted organic modifiers. Macromolecules 33, 1089 (2000)
76. Mori, T., Tanaka, K.: Average stress in matrix and average elastic energy of materials with
misfitting inclusions. Acta Metall. 21, 571 (1973)
77. Tandon, G.P., Weng, G.J.: The effect of aspect ratio of inclusions on the elastic properties of
unidirectionally aligned composites. Polym. Compos. 5, 327 (1984)
78. Fornes, T.D., Paul, D.R.: Modeling properties of nylon 6/clay nanocomposites using
composite theories. Polymer 44, 4993 (2003)
79. Fischer, H.: Polymer nanocomposites: from fundamental research to specific applications.
Mater. Sci. Eng. C 23, 763 (2003)
80. Schaefer, D.W., Justice, R.S.: How nano are nanocomposites. Macromolecules 40, 501
(2007)
81. Kraus, G.: Reinforcement of elastomers by carbon black. Adv. Polym. Sci. 8, 115 (1971)
82. Payne, A.R.: Effect of dispersion on the dynamic properties of filler-loaded rubbers. J. Appl.
Polym. Sci. 9, 2273 (1965)
83. Bradshaw, R.D., Fisher, F.T., Brisnon, L.C.: Fiber waviness in nanotube-reinforced polymer
composites. II. Modeling via numerical approximation of the dilute strain concentration
tensor. Compos. Sci. Technol. 63, 1705 (2003)
84. Thostenson, E.T., Chou, T.W.: On the elastic properties of carbon nanotube-based
composites: modelling and characterization. J. Phys. D Appl. Phys. 36, 573 (2003)
85. Schadler, L.S., Giannaris, S.C., Ajayan, P.M.: Load transfer in carbon nanotube epoxy
composites. Appl. Phys. Lett. 73, 3842 (1998)
86. Ajayan, P.M., Schadler, L.S., Giannaris, S.C., Rubio, A.: Single-walled carbon
nanotubepolymer composites: strength and weakness. Adv. Mater. 12, 750 (2000)
Role of Different Nanoparticles 53

87. Wong, M., Paramsothy, M., Xu, X.J., Ren, Y., Li, S., Liao, K.: Physical interactions at
carbon nanotube-polymer interface. Polymer 44, 7757 (2003)
88. Gojny, F.H., Wichmann, M.H.G., Fiedler, B., Schulte, K.: Influence of different carbon
nanotubes on the mechanical properties of epoxy matrix composites—a comparative study.
Compos. Sci. Technol. 65, 2300 (2005)
89. Gojny, F.H., Nastalczyk, J., Roslaneic, Z., Schulte, K.: Surface modified multi-walled
carbon nanotubes in CNT/epoxy-composites. Chem. Phys. Lett. 370, 820 (2003)
90. Dyke, C.A., Tour, J.M.: Covalent functionalization of single-walled carbon nanotubes for
materials applications. J. Phys. Chem. A 108, 11151 (2004)
91. Haluska, M., Hulman, M., Hirscher, M., Becher, M., Roth, S., Stepanek, I., Bernier, P.:
Proceedings of AIP conference on electronic properties of molecular nanostructures, 591,
603 (2001)
92. Chen, J., Dyer, M.J., Yu, M.F.: Cyclodextrin-mediated soft cutting of single-walled carbon
nanotubes. J. Am. Chem. Soc. 123, 6201 (2001)
93. Hilding, J., Grulke, E.A., Zhang, Z.G., Lockwood, F.: Dispersion of carbon nanotubes in
liquids. J. Disp. Sci. Technol. 24, 1 (2003)
94. Pierard, N., Fonseca, A., Colomer, J.F., Bossout, C., Benoit, J.M., Van Tendeloo, G.,
Pirard, J.P., Nagy, J.B.: Ball milling effect on the structure of single-wall carbon nanotubes.
Carbon 42, 1691 (2004)
95. Konya, Z., Vesselneyi, K., Niesz, A., Demortier, A., Fonseca, J., Delhalle, Z.: Large scale
production of short functionalized carbon nanotubes. Chem. Phys. Lett. 360, 429 (2002)
96. Lee, J.H., Jeong, T., Heo, J., Park, S.H., Lee, D.H., Park, J.B., et al.: Short carbon nanotubes
produced by cryogenic crushing. Carbon 44, 2984 (2006)
97. Peng, J., Qu, X.X., Wei, G.S., Li, J.Q., Qiao, J.L.: The cutting of MWNTs using gamma
radiation in the presence of dilute sulfuric acid. Carbon 42, 2741 (2004)
98. Tasis, D., Tagmatarchis, N., Bianco, A., Prato, M.: Chemistry of carbon nanotubes. Chem.
Rev. 106, 1105 (2006)
99. Chen, J., Hamon, M.A., Hu, H., Chen, Y., Rao, A.M., Eklund, P.C., Haddon, R.C.: Solution
properties of single-walled carbon nanotubes. Science 282, 95 (1998)
100. Riggs, J.E., Guo, Z., Carroll, D.L., Sun, Y.P.: Strong luminescence of solubilized carbon
nanotubes. J. Am. Chem. Soc. 122, 5879 (2000)
101. Sridhar, V., Gupta, B.R., Tripathy, D.K.: Bound rubber in chlorobutyl compounds: influence
of filler type and storage time. J. Appl. Polym. Sci. 102, 715 (2006)
102. Shanmugharaj, A.M., Bae, J.H., Lee, K.Y., Noh, W.H., Lee, S.H., Ryu, S.H.: Physical and
chemical characteristics of multiwalled carbon nanotubes functionalized with aminosilane
and its influence on the properties of natural rubber composites. Comp. Sci. Technol. 67,
1813 (2007)
103. Jitianu, A., Cacciguara, T., Berger, M.H., Benoit, R., Beguin, F., Bonnamy, S.: New carbon
multiwall nanotubes-TiO2 nanocomposites obtained by the sol-gel method. J. NonCryst.
Sol. 345, 596 (2004)
104. Fan, W., Gao, L., Sun, J.: Anatase TiO2-coated multi-wall carbon nanotubes with the vapour
phase method. J. Am. Ceram. Soc. 89, 731 (2006)
105. Wang, Y., Wu, J., Wei, F.: A treatment method to give separated multi-walled carbon
nanotubes with high purity, high crystallization and a large aspect ratio. Carbon 41, 2939
(2003)
106. Schaefer, D.W., Zhao, J., Brown, J.M., Anderson, D.P., Tomlin, D.W.: Morphology of
dispersed carbon single-walled nanotubes. Chem. Phys. Lett. 375, 369 (2003)
107. Lu, K.L., Lago, R.M., Chen, Y.K., Green, M.L.H., Harris, P.J.F., Tsang, S.C.: Mechanical
damage of carbon nanotubes by ultrasound. Carbon 34, 814 (1996)
108. Singh, B.P., Menchavez, R., Fuji, M., Takahashi, M.: Characterization of concentrated
colloidal ceramics suspension: a new approach. J. Colloid Interface Sci. 300, 163 (2006)
109. Velsaco-Santos, C., Martinez-Hernandez, A.L., Fisher, F., Ruoff, R., Castano, V.M.:
Dynamical mechanical and thermal analysis of carbon nanotube-methyl-ethyl methacrylate
nanocomposites. J. Phys. D Appl. Phys. 36, 1423 (2003)
54 J. K. Kim et al.

110. Cui, S., Canet, A., Derre, M., Couzi, M., Delhaes, P.: Characterization of multiwall carbon
nanotubes and influence of surfactant in the nanocomposite processing. Carbon 41, 797
(2003)
111. Vaisman, L., Marom, G., Wagner, H.D.: Dispersions of surface-modified carbon nanotubes
in water-soluble and water-insoluble polymers. Adv. Funct. Mater. 16, 357 (2006)
112. Graff, R.A., Swanson, J.P., Barone, P.W., Baik, S., Heller, D.A., Strano, M.S.: Achieving
individual-nanotube dispersion at high loading in single-walled carbon nanotube
composites. Adv. Mater. 17, 980 (2005)
113. Vaisman, L., Wagner, H.D., Marom, G.: The role of surfactants in dispersion of carbon
nanotubes. Adv. Colloid Interface Sci. 128, 37 (2006)
114. Fukushima, H., Drzal, L.T., Rook, B.P., Rich, M.J.: Thermal conductivity of exfoliated
graphite nanocomposites. J. Therm. Anal. Calor. 85, 235 (2006)
115. Xu, D., Sridhar, V., Pham, T.T., Kim, J.K.: Dispersion, mechanical and thermal properties
of nano graphite platelets reinforced flouroelastomer composites. E-Polym 23, 1 (2008)
116. Xu, D., Sridhar, V., Mahapatra, S.P., Kim, J.K.: Dielectric properties of exfoliated nano
graphite reinforced flouroelastomer composites. J. Appl. Polym. Sci. 111, 1358 (2009)
117. Pham, T.T., Sridhar, V., Kim, J.K.: Floroelastomer-MWCNT nanocomposites-1: dispersion,
morphology, physico-mechanical and thermal properties. Polym. Comp. 30, 121 (2009)
118. Sridhar, V., Xu, D., Pham, T.T., Kim, J.K.: Dielectric and dynamic mechanical and
relaxation behavior of nano graphite reinforced flouroelastomer composites. Polym. Comp.
30, 334 (2009)
119. Paglicawan, M.A., Balasubramanian, M., Kim, J.K.: Study on nanocomposite thermoplastic
elastomer gels. Macromol. Symp. 249, 601 (2007)
120. Yasmin, A., Daniel, I.M.: Mechanical and thermal properties of graphite platelet/epoxy
composites. Polymer 45, 8211 (2004)
121. Shaffer, M.S.P., Windle, A.H.: Fabrication and characterization of carbon nanotube/
poly(vinyl alcohol) composites. Adv. Mater. 11, 937 (1999)
122. Vu-Kanh, T., Fisa, B.: Fracture behavior of mica-reinforced polypropylene: effects of
coupling agent, flake orientation, and degradation. Polym. Compos. 7, 216 (1986)
123. Zheng, W., Wong, S.C.: Electrical conductivity and dielectric properties of PMMA/
expanded graphite composites. Compos. Sci. Technol. 63, 225 (2003)
124. Roland, C.M., Bero, C.A.: Normal mode relaxation in linear and branched polyisoprene.
Macromolecules 29, 7521 (1996)
125. Lewis, T.B., Nielsen, L.E.: Dynamic mechanical properties of particulate-filled composites.
J. Appl. Polym. Sci. 14, 1449 (1970)
126. Eklind, H., Schantz, S., Maurer, F.H.J., Jannasch, P., Wesslen, B.: Characterization of the
interphase in PPO/PMMA blends compatibilized by P(S-g-EO). Macromolecules 29, 984
(1996)
127. Shalaby, S.W.: In: Turi, E. (ed.) Thermal Characterization of Polymeric Materials.
Academic Press, London (1981)
128. Jawad, S.A., Alnajjar, A.: Frequency and temperature dependence of ac electrical properties
of graphitized carbon-black filled rubbers. Polym. Int. 44, 208 (1997)
129. Wang, Y.J., Pan, Y., Zhang, X.W., Tan, K.: Impedance spectra of carbon black filled
highdensity polyethylene composites. J. Appl. Polym. Sci. 98, 1344 (2005)
130. Levya, M.E., Barra, G.M.O., Moreira, A.C.F., Soares, B.G., Khastgir, D.: Electric,
dielectric, and dynamic mechanical behavior of carbon black/styrene-butadiene-styrene
composites. J. Polym. Sci. B Polym. Phys. 41, 2983 (2003)
131. Levon, K., Margolina, A., Patashinsky, A.Z.: Multiple percolation in conducting polymer
blends. Macromolecules 26, 4061 (1993)
132. Roberts, A.P., Knackstedt, M.A.: Structure-property correlations in model composite
materials. Phys. Rev. E 54, 2313 (1996)
133. Curro, J.G., Mark, J.E.: A non-Gaussian theory of rubberlike elasticity based on rotational
isomeric state simulations of network chain configurations. II. Bimodal
poly(dimethylsiloxane) networks. J. Chem. Phys. 80, 4521 (1984)
Role of Different Nanoparticles 55

134. Dutta, N.K., Tripathy, D.K.: Effects of types of fillers on the molecular relaxation
characteristics, dynamic mechanical, and physical properties of rubber vulcanizates. J. Appl.
Polym. Sci. 44, 1635 (1992)
135. Smit, P.P.A.: The glass transition in carbon black reinforced rubber. Rheol. Acta 5, 277
(1966)
136. Serizawa, H., Nakamura, T., Ito, M., Tanaka, K., Nomura, A.: Effects of oxidation of carbon
black surface on the properties of carbon black- natural rubber systems. Polym. J. 14, 201
(1983)
137. Angell, C.A., Ngai, K.L., McKenna, G.B., McMillan, P.F., Martin, S.W.: Relaxation in
glass forming liquids and amorphous solids. J. Appl. Phys. 88, 3113 (2000)
138. Toki, S., Hsiao, B.S.: Nature of strain-induced structures in natural and synthetic rubbers
under stretching. Macromolecules 36, 5915 (2003)
139. Luch, D., Yeh, G.S.Y.: Morphology of strain-induced crystallization of natural rubber.
I. Electron microscopy on uncrosslinked thin film. J. Appl. Phys. 43, 4326 (1972)
140. Shimizu, T., Tanaka, Y., Kutsumizu, S., Yano, S.: Ordered structure of
poly(fluoroalkylfluoroacrylates). Macromolecules 26, 6694 (1993)
141. Keller, A.: Growth and perfection of crystals. In: Doremus, R.I., Roberts, B.W., Turnbull,
D. (eds.) Proceedings of conference on crystal growth. New York, p. 499 (1958)
142. Conradt, R.N.J., Heise, B., Kilian, H.G.: Crystallization in stretched and unstretched rubber-
thermodynamics and kinetics. Prog. Colloid Polym. Sci. 87, 84 (1992)
143. Bandyopadhyay, S., De, P.P., Tripathy, D.K., De, S.K.: Influence of surface oxidation of
carbon black on its interaction with nitrile rubbers. Polymer 37, 353 (1996)
144. Sternstein, S.S., Zhu, A.I.: Reinforcement mechanism of nanofilled polymer melts as
elucidated by nonlinear viscoelastic behaviour. Macromolecules 35, 7262 (2002)
145. Kavanagh, G.M., Ross-Murphy, S.B.: Rheological characterisation of polymer gels. Prog.
Polym. Sci. 23, 533 (1998)
146. Mortensen, K., Theunissen, E., Kleppinger, R., Almdal, K., Reynaers, H.: Shear-induced
morphologies of cubic ordered block copolymer micellar networks studied by in situ
smallangle neutron scattering and rheology. Macromolecules 35, 7773 (2002)
147. Xu, Z., Zhao, C., Gu, A., Fang, Z.: Electric conductivity of PS/PA6/carbon black
composites. J. Appl. Polym. Sci. 103 (2004)
In Situ Synthesis of Rubber
Nanocomposites

Massimo Messori

Abstract The preparation and characterization of rubber based nanocomposites


prepared by in situ generation of inorganic oxides by means of the hydrolytic sol–
gel process are reviewed in the present chapter. The sol–gel approach has been
applied to several rubber matrices to prepare reinforced vulcanized and unvul-
canized rubbers. Several synthetic procedures are presented while the most
investigated filler is silica obtained by hydrolysis and condensation of tetraeth-
oxysilane. The effects of the different preparation conditions and of the filler
content are generally discussed in terms of morphology (investigated by electron
microscopy and small angle X-ray scattering) and mechanical properties (modulus,
strength and extensibility). The mechanical properties of the in situ filled nano-
composites are generally better than those of the corresponding materials prepared
with the conventional mechanical mixing of preformed particulates and elasto-
mers. This enhancement is generally attributed to a lower tendency to filler–filler
aggregation due to a lower particle surface interaction resulting from the ‘bottom-
up approach’ of the sol–gel process applied to the preparation of organic–inorganic
hybrid materials.

1 Introduction

The improvement of the mechanical properties (reinforcement) of elastomeric


materials by addition of rigid fillers represent one of the most important aspects
in the field of rubber science and technology [1]. The concurrent enhancement of

M. Messori (&)
Dipartimento di Ingegneria dei Materiali e dell’Ambiente,
Università di Modena e Reggio Emilia,
Via Vignolese 905/A, 41125 Modena, Italy
e-mail: massimo.messori@unimore.it

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 57


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_2,
Ó Springer-Verlag Berlin Heidelberg 2011
58 M. Messori

stiffness (elastic modulus) and, possibly, of elongation at break due to the


presence of rigid particles derives from hydrodynamic effects mainly depending
on the filler volume fraction but also affected by the filler shape factor (aspect
ratio). The interactions between rubber and filler, which can be increased if good
dispersion and distribution of the particulate are achieved, play a fundamental
role for the increase of elastic modulus. The surface characteristics of the par-
ticles (presence of reactive/functional groups, wettability and surface energy) and
the chemical nature of the rubber represent the key parameters for these
interactions.
Concerning the field of fillers having dimensions in the nanometric scale,
polymer matrix nanocomposites have attracted extraordinary attention in the last
decade on the basis of their excellent mechanical and barrier properties compared
to the conventional microcomposites, usually at very low filler content. Layered
silicates, ceramic nanoparticles (such as silica, titania, zirconia, etc.), carbon
nanofibers and nanotubes are typical examples of materials used as nanosize
reinforcing additive [2]. The usual method for the preparation of nanocomposites
is based on the top-down approach according to which preformed nano-objects are
dispersed within the polymeric matrix by physical–mechanical dispersion and
distribution (melt or solution mixing). Even if very attractive from an industrial
point of view, this method presents some severe limitations related to the diffi-
culties to obtain an effective dispersion due to the strong tendency to particles
aggregation phenomena and the significant increase of melt viscosity because of
the complex rheology of nanocomposite systems.
In particular, in the specific case of rubber composites, it is well known that
carbon black represents the most effective reinforcing additive for elastomeric
compounds notwithstanding the strong limitation due to the black coloration
imparted to the final part. Alternatively to carbon black, also silica is another filler
widely employed in the rubber industry, thanks to some advantages deriving from
its use such as high tear strength, good abrasion resistance and reduction in heat
build-up [3]. The main drawbacks associated with the use of conventional silica as
reinforcing additive instead of carbon black are the higher compound viscosity, the
incompatibility of silica and rubber, the more difficult mixing and processing, the
longer vulcanization time, the lower cross-linking density and the absorption of
curative additives by silica. Most of these disadvantages are due to the very strong
interaction occurring among silica particles and caused by the hydrogen bonding
of the silanol groups present onto the silica surface. This interaction inhibits an
uniform dispersion of the filler within the rubber matrix giving rise to the for-
mation of particles’ aggregates. Silica agglomeration and silica–rubber incom-
patibility are generally reduced by using different types of silanes as coupling
agents.
In order to minimize these problems and alternatively to the usually employed
processing technique (mechanical mixing) for the preparation of filled vulcanized
rubber, the in situ generation of silica (or other inorganic oxides) by using the sol–
gel process has been recognized as a novel and interesting technique for the
preparation of rubber composites.
In Situ Synthesis of Rubber Nanocomposites 59

The classical aqueous (or non-hydrolytic) sol–gel process [4] consists of a


two-step hydrolysis-condensation reaction starting with metal alkoxides M(OR)x,
typically tetraethoxysilane (TEOS), according to the following scheme:
Step 1: Hydrolysis

MðORÞx þ n H2 O ! MðORÞxn ðOHÞn þ n ROH

Step 2: Condensation

MOH þ HOM ! MOM þ H2 O ðwater condensationÞ

and/or
MOR þ HOM ! MOM þ ROH ðalcohol condensationÞ
The presence in the reactive system of an organic oligomer or polymer (bearing
or not suitable groups reactive towards to the sol–gel process) leads to the for-
mation of organic–inorganic hybrid structures composed of metal oxide (silica or
other) and organic phases intimately mixed each other. This synthetic procedure
belongs to the so-called ‘bottom-up’ approach for the preparation of hybrid
materials and, depending on the experimental conditions, permits the synthesis of
composite structures in which the dimensions of the dispersed phase are under
100 nm (nanocomposites). The optical, physical and mechanical properties of
these nanocomposites are strongly dependent not only on the individual properties
of each component, but also on important aspects of the chemistry involved such
as uniformity, phase continuity, domain size and the molecular mixing at the phase
boundaries.
The morphologies of the hybrid materials are strictly dependent on the char-
acteristics of the organic polymer such as the molecular weight, the presence and
the number of reactive functionalities as well as the solubility of the polymer in the
sol–gel system.
The nature of the interface between organic and inorganic phases is generally
used to grossly divide these materials into two distinct classes. In Class I, organic
and inorganic components are simply embedded and only weak bonds (hydrogen,
dipolar and van der Waals bonds) give the cohesion to the whole structure. On the
contrary, in Class II materials, the two phases are linked together through strong
chemical bonds (covalent or ionic bonds).
Several applications of organic–inorganic hybrid nanocomposites prepared by
sol–gel and other procedure have been already extensively reviewed [5–8]. It is
also important to underline that the use of the sol–gel procedure for the preparation
of rubber matrix nanocomposites could present some peculiarity mainly due to the
possibility to vulcanize a preformed organic polymer (‘green’ rubber).
60 M. Messori

The majority of the reported papers deals with the preparation and character-
ization of elastomers filled with in situ generated silica obtained from TEOS as
precursors while a much more limited literature on other inorganic oxides such as
titania, zirconia and alumina is available. Several synthetic procedures are pro-
posed and the characterization is mainly devoted to mechanical and/or thermo-
mechanical properties, such as quasi-static tensile measurements or dynamic
mechanical analysis, and to morphology investigation by means of scanning and
transmission electron microscopy and/or small-angle X-ray scattering. One of the
main targets is the development of nanocomposites with improved properties with
respect to similar materials obtained with the conventional mechanical mixing of
rubber and preformed nanoparticles.
In the present chapter a general overview on the in situ generation of silica or
other reinforcing fillers within elastomeric phases reported in literature is pre-
sented and discussed for different elastomeric matrices.

2 General Synthetic Strategies

From a generic point of view, the in situ generation of an inorganic oxide phase
(usually having a spherical or quasi-spherical shape) by means of the sol–gel
process within an elastomeric polymer can be carried out according to different
strategies, as step-by-step schematically described in the following.
Solution procedure (procedure A)
(i) dissolution of both the metal alkoxide (metal oxide precursor) and unvul-
canized rubber in a common solvent;
(ii) addition of water (directly added or absorbed from the external humid
atmosphere), sol–gel catalysts (basic or acidic) and vulcanization ingredients
(optional) and activation of the sol–gel process at a given temperature and for
a given reaction time;
(iii) removal of solvent and by-products (usually water and alcohols) by evapo-
ration or precipitation in a non-solvent;
(iv) vulcanization of the filled ‘green’ rubber (optional).

Swelling of unvulcanized rubber (procedure B)


(i) immersion and swelling of films or sheets of unvulcanized rubber in metal
alkoxide;
(ii) immersion of the swollen rubber in a basic or acidic aqueous solution and
activation of the sol–gel process at a given temperature and for a given
reaction time;
(iii) mechanical mixing (usually by using a conventional two-roll mill) of the
filled ‘green’ rubber with the vulcanization ingredients and vulcanization
(optional).
In Situ Synthesis of Rubber Nanocomposites 61

Swelling of vulcanized rubber (procedure C)


(i) immersion and swelling of films or sheets of vulcanized rubber in metal
alkoxide;
(ii) immersion of the swollen vulcanized rubber in a basic or acidic aqueous
solution and activation of the sol–gel process at a given temperature and for a
given reaction time.
It can be anticipated that the procedure A (in solution) ensures a highly
homogeneous dispersion of the in situ generated filler within the rubber matrix also
for high reinforcing metal oxide contents. The swelling of unvulcanized rubber
(procedure B) is highly attractive especially for industrial applications thanks to its
solvent-free approach while the swelling of vulcanized rubber (procedure C) is
very simple and can be in principle applied to already moulded three-dimensional
articles (even if limitations in the maximum amount of filler generated can be
present due to the cross-linked structure of the swollen rubber).

3 Rubber Based Nanocomposites

3.1 Polydimethylsiloxane Based Nanocomposites

Polydimethylsiloxane (PDMS) is the most important member of the class of


polysiloxanes and presents some interesting properties thanks to the presence of
Si–O–Si linkages in the polymeric backbone such as good thermal stability, water
repellency, excellent resistance to oxygen, ozone and UV-light, anti-stickiness and
low chemical reactivity. As drawbacks, PDMS is characterized by very poor
mechanical properties, in particular tensile strength, and it requires the addition of
reinforcing additives (usually mineral fillers) in order to achieve the mechanical
properties required depending on the final application.
PDMS is traditionally reinforced with silica and the filler–matrix interactions
are ensured through hydrogen bonding between silanol groups of silica surface and
oxygen atoms of PDMS macromolecules. Surface modification of silica particles
permits the optimization of the interactions between mineral filler and polymer.
Due to the concurrent presence of Si–O groups in both silica and PDMS, it is not
surprising that these represent one of the first systems investigated in the field of in
situ generation of silica within polymeric matrix, as indicated by the pioneeristic
works of Mark [9, 10].

3.1.1 PDMS-Silica Nanocomposites

According to a slight modification of the above described procedure C (swelling


of vulcanized rubber), Rajan et al. [11] reported a method (named controlled
hydrolysis, CH) for the in situ generation of silica particles in which, after swelling
62 M. Messori

Table 1 PDMS/SiO2
Catalyst Silica Ultimate Elongation
composites: sol–gel catalysts,
(wt%) strength (MPa) (relative length)
filler content and mechanical
at rupture
properties (adapted from
Ref. [11]) None 0.0 0.082 1.77
Dibutyltin diacetate 5.4 0.538 1.68
8.9 0.818 1.63
14.2 0.712 1.09
Dibutyltin dilaurate 6.0 0.543 2.01
8.0 0.831 1.93

of cross-linked PDMS in TEOS, the required water of hydrolysis was simply


absorbed from the air and the catalyst was generated from tin salts (dibutyltin
diacetate or dilaurate). These PDMS nanocomposites had a silica content up to 14
wt% and were characterized by an unusually high transparency (if compared to
similar materials obtained with aqueous ammonia as catalysts) as quantitatively
judged by UV–VIS spectroscopy. Electron microscopy showed that silica domains
were very small (30–50 nm in diameter) and well dispersed, as expected from the
transparency of the composites. Tensile stress–strain measurements indicated that
the particles provide very good reinforcement. Data reported in Table 1 indicate
that the ultimate tensile strength and Young’s moduli increased with higher silica
content and elongation at break remained almost the same of the unfilled PDMS
(up to a silica content of about 9 wt%).
The dependence of silica particle dimensions on PDMS cross-linking density,
silica content and catalyst concentration was investigated by small-angle X-ray
scattering [12].
Of particular interest were the relationships between particle size and molecular
weight of the network chains (Mc, mesh sizes), amount of filler introduced and
catalyst concentration.
Silica particle sizes were smallest for the smallest values of Mc, presumably due
to constraining effects deriving from the very short network chains. At fixed Mc
and filler concentrations, higher catalyst concentrations gave larger particles.
Increase in silica concentration generally had little effect on particle size at low
and high loadings, but markedly increased sizes at intermediate levels
(10–20 wt%), presumably caused by coalescence of isolated small particles into
considerably larger aggregates.
Films comprised of PDMS and in situ generated SiO2 were applied on silicon,
aluminum and polystyrene substrates. The surfaces in contact with air and with a
substrate were investigated by using several surface technique [13]. The hybrid
sample surfaces generated in contact with air were characterized by a silica-free
PDMS top layer of about 2 nm; in the surfaces in contact with the substrates SiO2
was located at or just beneath the outermost atomic layer. In contact with polar
liquids such as water, polar hydroxy groups present at the surface of SiO2 can
easily stretch out to the outer-most atomic layer. Quite surprisingly, no significant
correlation was found between the roughness of the surfaces and the amount of in
situ generated SiO2 present in the materials.
In Situ Synthesis of Rubber Nanocomposites 63

Fig. 1 TEM image of


PDMS/SiO2 composites
(silica content 10 phr)
(reproduced with permission
from Ref. [14])

Bokobza et al. [14, 15] applied the same synthetic procedure to obtain similar
composites by starting from a cross-linked vinyl-terminated PDMS. Authors found
that the in situ generated silica structures were uniformly dispersed within the
elastomeric matrix (see Fig. 1) but different morphologies were obtained as a
function of the used tin-based catalyst and thus of the type of growth processes.
The hydrophilic character of the silica surface was responsible of extensive
interaction with PDMS chains leading to a significant improvement in the
mechanical properties of the composites.
An interesting comparison among PDMS reinforced with different types of
fillers has been published by Bokobza [16]. Three different techniques of incor-
poration of silica were compared: the conventional mechanical mixing process, in
situ filling process and introduction of spherical colloidal silicas (Stöber silicas). In
addition, results obtained with other types of fillers (layered silicates and fibrous
clays) were reported and discussed. Among the different types of reinforcement of
PDMS imparted by silica particles, the in situ filling process was by far the most
efficient. As reported in Table 2, the modulus, ultimate stress and extensibility
increased by increasing the filler loading. Moreover, larger increases were
observed over the silica particles percolation threshold.

Table 2 PDMS/SiO2 composites: mechanical properties (adapted from ref. [16])


Compound Stress at 100% Stress at 200% Tensile strength Elongation
strain (MPa) strain (MPa) (MPa) at break (%)
Unfilled sample n.r. n.r. 0.43 91
10 phr SiO2 1.44 4.98 7.64 242
18 phr SiO2 2.49 7.86 10.70 240
30 phr SiO2 5.19 10.35 21.54 313
40 phr SiO2 7.99 16.38 27.34 275
64 M. Messori

3.1.2 PDMS-Titania and PDMS-Zirconia Nanocomposites

The same CH method was used by Murugesan and Mark [17, 18] for the
preparation of PDMS containing in situ generated titania and zirconia starting
from different alkoxides and the obtained materials were compared to those
prepared according the so-called conventional ‘water excess process’ (WE, based
on the direct addition of stoichiometric water to the cross-linked PDMS swollen
in the corresponding metal alkoxide). The composites prepared by the CH
method were characterized by a greater amount of filler (higher conversion of
alkoxides to zirconia or titania) for the same reaction time. The optical trans-
parency of the prepared materials is reported in Fig. 2 and UV–VIS analysis
showed that the transmittance values of PDMS/ZrO2 were significantly higher
than those of PDMS/TiO2 (47 and 28% for composites containing 6 wt% of filler
and 16 and 0.01% for composites containing 28 wt% of filler, respectively)
according to the fact that titania particles formed small aggregates compared to
zirconia, as evidenced by small-angle X-ray scattering analysis and leading to a
higher opacity.
Thermogravimetric analysis showed that PDMS/ZrO2 composites had a lower
thermal stability with respect to PDMS/TiO2 composites, which were inherently
more stable due to the formation of different phases at high temperatures. Con-
cerning the mechanical properties, all the PDMS/ZrO2 and PDMS/TiO2 compos-
ites (independently on the preparation method) had mechanical properties much
improved with respect to the unfilled elastomer. The reinforcement of zirconia
filled composites was greater compared to titania filled composites even for similar
amount of filler presumably due to the smaller dimension of ZrO2 particles,
consistently with the above discussed transparency.

Fig. 2 Transparency com-


parison of PDMS/ZrO2 and of
PDMS/TiO2 composites
(reproduced with permission
from Ref. [18])
In Situ Synthesis of Rubber Nanocomposites 65

3.1.3 PDMS-(mixed oxides) Nanocomposites

Wen and Mark [19] reported the in situ generation of silica–titania mixed oxide
within a PDMS network (previously cross-linked with TEOS). TEOS and tetra-
butyl titanate (TBT) were the precursors of the correspondent oxides. Since the
TEOS hydrolysis is extremely slow compared to titanium alkoxides, hydrolysis of
a mixture of them generally results in phase segregation leading to a physical
mixture of two oxides instead of a true mixed oxide. This problem has been
avoided carrying out a partial hydrolysis of pure TEOS with subsequent addition
of the more reactive TBT. The sol–gel process has been activated by placing the
swollen PDMS in 2 wt% diethyleneamine aqueous solution at room temperature
leading to composites with a silica–titania content ranging from 10 to 22 wt%.
TEM analysis carried out on PDMS/SiO2–TiO2 composites showed a filler average
diameter of 20–25 nm, with a relatively narrow diameter distribution and with
very little aggregation of particles. A further interesting observation was the
increase of the particle size by decreasing the cross-linking density of the rubber,
as already evidenced for PDMS/SiO2 nanocomposites [12].
The analysis of equilibrium swelling data on the basis of the Kraus’ theory [20]
suggested that the composites were of the adhering type in which the rubber matrix
is restricted by the filler through attachments onto the filler surface.
Concerning mechanical properties, stress–strain isotherms represented as plots
of modulus against reciprocal elongation according to the Mooney-Rivlin equation
[21, 22] indicated that, compared to the silica-filled PDMS networks, the mixed-
oxide based composites had better extendibility and had upturns which occurred at
higher elongations (but were smaller because of the lower reinforcing effect
provided by TiO2).
The presence of in situ generated silica–titania mixed oxides also increased the
onset temperature for thermal degradation of PDMS, on the contrary of that
showed by physical mixing of preformed silica and/or titania nanoparticles.
The same authors reported a further study [23] on the preparation and char-
acterization of PDMS/SiO2 and PDMS/TiO2–SiO2 composites having inorganic
contents ranging from 30 to 100 wt% and in which the sol–gel reaction (con-
version of TEOS in SiO2 and of TBT in TiO2 and corresponding mixed oxides)
occurred simultaneously with the cross-linking of PDMS due to the co-conden-
sation of hydroxyl-terminated PDMS and TEOS. Experimental data indicated that
this last process was dominant and that the majority of the PDMS was incorporated
in the silica network. The addition of PDMS was found to shorten the gelation
times and to increase the rates of increase in modulus of the network structure. The
impact strength and the fracture surfaces of the materials were studied with the
finding that the presence of PDMS significantly increases their impact strength and
ductility, as shown in Fig. 3.
The growth processes and resulting structures of the reinforcing fillers were
investigated by small-angle X-ray scattering by Breiner and Mark [24]. The sys-
tems were found to yield dense particles with fractally rough surfaces and
the results were used to interpret mechanical properties of these composites.
66 M. Messori

Fig. 3 Dependence of
impact strengths on PDMS
content for PDMS/SiO2
composites. The impact
strength were obtained from a
the Charpy pendulum impact
test and b the falling-weight
impact test (reproduced with
permission from Ref. [23])

The reinforcement with mixed oxides seems to synergistically combine the best
properties of titania and silica: the titania provides additional reinforcement at low
strains and allows for higher elongations to be obtained, whereas the silica con-
tributes to the reinforcement over the entire range of allowable strain and gives rise
to the very desirable upturn in modulus at high elongations.
Wen and Mark [25] also reported the precipitation of silica–zirconia and silica–
alumina mixed oxides (at a concentration ranging from 10 to 22 wt%) into PDMS
networks using a sol–gel approach according the above described method. The
resulting filled networks were found to have very good mechanical properties. In
comparison with networks filled only with silica, these materials had good
extensibilities as well as high strengths. Filler particle diameters were generally
several hundred angstroms, but also in this case were found to decrease with
increase in cross-linking density of the networks. The distributions of particle size
were relatively narrow, and there was very little particles aggregation. The pres-
ence of in situ generated silica–titania and silica–zirconia mixed oxides also
improved the thermal stability of the PDMS.

3.2 Natural Rubber Based Nanocomposites

Bokobza and co-workers [26, 27] reported the preparation and the characterization
of natural rubber (NR) composites containing in situ generated silica particles by
hydrolysis and condensation of TEOS before the vulcanization of the rubber
matrix in solution (according to the procedure A) and after the vulcanization of the
rubber (according to the procedure C) with or without the addition of a silane
coupling agent (bis(3-triethoxysilylpropyl) tetrasulfide, TESPT). The silica content
of the prepared composites varied from 8 to 21 phr.
Tensile stress–strain experiments showed that, at a given strain, the stress
increased by increasing the silica content showing a significant reinforcing effect
due to the presence of in situ generated silica.
In Situ Synthesis of Rubber Nanocomposites 67

It is generally accepted that the increase of the stiffness (modulus) deriving


from the generation or incorporation of active filler within a polymeric matrix is
related to two factors: (i) a hydrodynamic effect arising from the presence of rigid
particles and (ii) an increase in the cross-linking density due to polymer–filler
interactions. The first term can be quantitatively described by the following Guth
and Gold equation [28, 29]:

G ¼ G0 ð1 þ 2:5  / þ 14:1  /2 Þ
where G and G0 represent the moduli of the composite and the unfilled polymeric
matrix, respectively, and / represents the volume fraction of the filler.
As shown in Fig. 4, the stress–strain curve calculated by using the Guth and
Gold equation was lower with respect to the experimental curve indicating that,
besides a hydrodynamic reinforcement, the presence of filler–rubber interactions
must be considered. Similar conclusions were also reported analysing the stress–
strain data according to the Mooney-Rivlin plots and on the basis of equilibrium
swelling tests.
Authors also demonstrated that silica particles generated before rubber vulca-
nization (according to the procedure A) inhibit the cross-linking reaction of the
rubber compound by sulphur. The presence of a filler networking structure, due to
the aggregation of silica particles via the silanol groups present on the surface, was
suggested by the Mooney-Rivlin plot analysis. On the other hand, the tendency to
form a particle–particle network was avoided when silica was generated in the
already vulcanized rubber (according to the procedure C) and small and well
dispersed particles were observed.
Pissis and co-workers [30, 31] investigated the molecular dynamic of the above
discussed materials. Broadband dielectric relaxation spectroscopy investigations
reported in these papers demonstrated that, in addition to the a relaxation related to
the glass transition of the rubber matrix, a slower a relaxation was observed and
assigned to polymer chains close to the polymer/silica interphase whose mobility

Fig. 4 Experimental and


calculated (Guth and Gold
equation) stress–strain curves
for NR/SiO2 composites
(reproduced with permission
from Ref. [26])
68 M. Messori

is restricted due to interaction with the silica surface. Authors reported an esti-
mation of the thickness of this interphase of about 2.1–2.4 nm.
Poompradub and co-workers [32, 33] prepared in situ silica reinforced NR by
means of the procedure in solution (procedure A) with n-hexylamine as basic
catalysts without any subsequent vulcanization and investigated the preparation
conditions and the rheological and mechanical properties of the prepared rubber
composites (having a silica content ranging from 10 to 70 phr).
Different solvents were used for the synthesis (CCl4, CHCl3 and tetrahydro-
furan, THF) and it was found that the lower amount of silica in rubber matrix was
generated when CCl4 and CHCl3 were used compared with THF. The best
behaviour of THF was attributed to its higher water solubility and polarity.
As expected, they observed an increase in Mooney viscosity by increasing the
in situ generated silica content. Interestingly, at the same filler content the vis-
cosity of composites containing in situ generated silica was lower compared to that
of rubber reinforced with preformed silica dispersed by mechanical mixing. This
behaviour has been attributed to the presence of a lower amount of silanol groups
onto the surface of in situ generated silica with respect to commercial preformed
silica, as further supported by dynamic mechanical analysis which suggested that
the silica–silica interactions of the in situ generated silica were weaker, resulting in
a better dispersion in the rubber matrix.
Also the moduli and compression set values of NR filled with in situ generated
silica were improved with respect to the commercial silica ones. The reinforce-
ment has been modelled by using the Guth and Gold equation with a shape factor
f = 2.53 (according to the morphology evidenced by TEM analysis). Authors
concluded that the volume fraction of filler, the particle–particle interaction and
the anisotropy of silica particles represent the three main factors affecting the
reinforcing effect due to the in situ generation of silica within the NR matrix.
Ikeda and Poompradub [34–37] reported the preparation and the characteriza-
tion of NR composites containing in situ generated silica particles from TEOS in
the presence of amines as catalyst before the vulcanization of the rubber (thus
according to the procedure B). As a general procedure, sheets of NR having a
thickness of about 1 mm were prepared by two-roll milling and immersed in
TEOS at room temperature or slightly higher and for different times. The swollen
NR sheets were further immersed in aqueous solution of amine-based catalyst at
40°C for 75 h in order to carry out the conversion of TEOS to silica. NR sheets
containing the in situ generated silica were mixed with sulphur-based vulcanizing
agents in a two-roll mill at room temperature. Then, the rubber compound was
moulded by hot-pressing to produce vulcanized sheets of about 1 mm of thickness.
As expected, the amount of in situ generated silica was found strongly
dependent on the amount of TEOS in the swollen NR and on the type of catalyst.
Authors found that the polarity and the basicity of amine catalyst were crucial for
controlling the conversion of TEOS to silica. In this respect, primary alkylamines
with suitable hydrocarbon segments (n-hexylamine, n-heptylamine and n-octyl-
amine) produced the highest amount of in situ generated silica (up to about
80 phr). Due to its higher solubility in water, n-hexylamine was found to be the
In Situ Synthesis of Rubber Nanocomposites 69

Fig. 5 Stress-elongation
curves for NR/SiO2 compos-
ites (filled circle in situ gen-
erated SiO2; filled square
SiO2 mechanically mixed)
(reproduced with permission
from Ref. [34])

most effective catalyst. Almost independently on the type of catalyst, all the silica
particles were generated in the level of nanometric size (average particle size
ranging from 25 to 45 nm) and with spherical shape.
The tensile properties of the materials have been investigated and compared
to those of rubber unfilled and filled with preformed nanosized silica by using
the conventional mechanical mixing method. Stress–strain curves reported in
Fig. 5 indicate that for the same filler content (71 phr) and for comparable
network chain densities, the in situ silica reinforced rubber showed the lower
stress at the elongation up to about 200% and higher stress at the elongation
beyond about 200%.
The high modulus at low elongation for conventional silica filled rubber has
been assumed to be due to the formation of larger aggregates of particles sug-
gesting a better dispersion in the case of in situ silica reinforced rubber. On the
other hand, the high modulus at high elongation for in situ silica filled rubber has
been attributed to a stronger interaction between NR matrix and in situ generated
silica particles.
The reinforcement effect due to in situ generated silica was compared to that of
carbon black (both low structure and high structure carbon black stocks) [38].
Physical and mechanical properties of NR reinforced with in situ generated silica
were in between those of carbon black filled composites and conventional silica
filled composites (both prepared by mechanical mixing).
Ikeda and co-workers [39, 40] also reported a detailed investigation on the
effect of a coupling agent (c-mercaptopropyltrimethoxysilane, c-MPS) on NR
filled with in situ generated silica. The addition of c-MPS was found to improve
the reinforcement effect. The diameter of in situ generated silica particles was in
70 M. Messori

the range 20–45 nm and their dispersion within the rubber matrix became
homogeneous by adding the silane coupling agent thanks to a reduction of silica–
silica interactions. The concurrent use of in situ generated silica and c-MPS also
prevented the delay of sulphur curing and increased the wettability of NR onto the
particles’ surface, which resulted in the increase of reinforcement of the vulca-
nized composite.
Ikeda et al. [41, 42] recently reported an interesting investigation on the
applicability of 3D-TEM, a technique that combines transmission electron
microscopy with computerized tomography, to visualize and analyze the three-
dimensional state of NR containing in situ generated silica.
Tangpasunthal and co-workers [43, 44] described the in situ generation of silica
within NR latex. Mixtures of TEOS and other alkyltriethoxysilanes (TESPT,
vinyltriethoxysilane, ethyltriethoxysilane and i-butyltriethoxysilane) were directly
added to a commercial grade NR latex, containing ammonia and thus permitting
the direct formation of silica through the sol–gel reaction thanks to the presence of
a basic pH. Compounds with sulfur-based vulcanizing agents were subsequently
prepared by two-roll milling and vulcanized by hot pressing.
The conversion of alkoxysilanes to silica within the rubber was almost
complete for TEOS but decreased when the alkyl group of the alkyltriethox-
ysilane increased in size. Silica particles with sizes between 100 and 500 nm
and evenly dispersed without extensive aggregation were obtained. The pres-
ence of TESPT, a coupling agent widely used in rubber industry, resulted in an
increase of the mechanical properties and the rate of sulphur cure. Hardness,
tensile and tear properties of vulcanized NR reinforced with in situ generated
silica were higher than those of similar composites prepared by conventional
mechanical mixing. Among the different alkyltriethoxysilane investigated, vi-
nyltriethoxysilane seemed to be the most promising taking into account the
high enhancement in tensile modulus and resistance to tear of the vulcanized
rubber.

3.3 Epoxidized Natural Rubber Based Nanocomposites

Epoxidized natural rubber (ENR)/silica nanocomposites were prepared by


Bhowmick and co-workers [45, 46] according to the solution procedure A.
TEOS (as precursor for the in situ generation of silica) and ENR were dissolved
in THF and the sol–gel process was activated at room temperature under
hydrochloric acid catalysis. Alternatively to NR, ENR with adequately high
epoxy content was chosen as rubber matrix taking into account that epoxy
groups, randomly distributed along the chain backbones, give higher glass
transition temperature and, most importantly, increased polarity and thus stronger
interaction with the in situ generated silica. In fact, it is well known that under
acidic conditions, the epoxy groups are likely to open up as a diol moiety which
can undergo intermolecular hydrogen bonding with the silanol groups. After
In Situ Synthesis of Rubber Nanocomposites 71

solvent elimination, some samples were vulcanized with benzoyl peroxide or


dicumyl peroxide. The composite films appeared transparent up to 50 wt% of
TEOS loading. Dispersion of the discrete silica particles having dimensions
ranging from 15 to 100 nm (by increasing the TEOS content) was observed.
Infrared spectroscopic studies indicated the occurrence of chemical interaction
within the rubber–silica interfaces which was further supported by the insolu-
bility of the respective samples in THF under the ambient conditions. Dynamic
mechanical analysis further corroborated the interactive nature of the in situ
generated silica with the ENR matrix.
All the ENR/SiO2 composites showed a significant improvement in
mechanical properties with increased TEOS loading within the rubber. A max-
imum increment in tensile strength (200%) and tensile modulus (170%) with
respect to the unfilled ENR was observed for the uncured composites with the
highest TEOS content (50 wt%). Further reinforcement was noticed when the
rubber in the nanocomposites was cured with either benzoyl peroxide or dicumyl
peroxide. The dicumyl peroxide cured hybrid composites displayed 112%
improvement in tensile strength over the control cross-linked rubber sample,
probably due to synergisms of nanosilica reinforcement and cross-linking of the
rubber phase in the hybrids.
The effect of polymer-silica interaction on THF swelling and dynamic
mechanical properties have been investigated and compared to those of acrylic
rubber/silica and poly(vinyl alcohol)/silica composites [47]. A further comparative
study on structure–property relationship of hybrids prepared under different pH
levels has been published by the same authors [48]. The silica particles were
formed in the nanometer scale (average diameter \100 nm) at low pH (equal or
lower than 2) beyond which aggregation occurred, although the conversion of
TEOS to silica was not strictly influenced by the various pH conditions. These
nanocomposites were optically clear and showed superior mechanical reinforce-
ment over the microcomposites containing aggregated silica structures with lower
optical clarity. Furthermore, the nanocomposites exhibited higher storage modulus
both at the glassy and the rubbery regions as compared to those of microcom-
posites. As evidenced by dynamic mechanical analysis, the tand peak heights were
also minimum and the glass transition temperature shifted to higher temperature
for those nanocomposites.
An interesting and alternative synthetic procedure has been proposed by
Hashim et al. [49, 50] according to which ENR sheets were first cross-linked with
3-aminopropyltriethoxysilane (APS) by hot pressing. After vulcanization, ENR
sheets were swollen in TEOS and subsequently subjected to a sol–gel reaction in
butylamine aqueous solution. Silica content was up to about 30 wt% according to
conversions of TEOS to silica higher than 60%. The obtained sol–gel vulcanized
rubber were more rigid and stronger than a typical sulphur-cured vulcanized ENR
containing comparable amount of silica. Comparative stress–strain and dynamic
mechanical property analysis suggested that chemicals bonds were formed
between the silica particles and the rubber network thanks to the dual reactivity of
APS with respect rubber vulcanization and sol–gel reaction.
72 M. Messori

3.4 Isoprene Rubber Based Nanocomposites

Messori and Bignotti [51] recently published a work on the in situ generation of
silica from TEOS within synthetic cis-1,4-polyisoprene (isoprene rubber, IR), the
synthetic counterpart of NR. According to the solution procedure A, the sol–gel
process was activated in solution of toluene, in the presence of all components and
by absorbing the water required for the hydrolysis of TEOS from the external humid
atmosphere. After solvent elimination, vulcanization of the rubber was attained for
some sample by thermal activation of dicumyl peroxide as vulcanizing agent. The
conversion of TEOS to silica was quantitative for samples with a low initial TEOS
concentration while for initial concentrations higher than 20 wt% the yield in silica
was in the range 60–80% almost independently on the formulation and the curing
conditions (presence or absence of coupling agent, vulcanization or not of IR phase).
The in situ generated silica particles were homogeneously dispersed in the
vulcanized rubber with a spherical shape and an average dimension which
increased from a few nanometers to the submicron scale (300–400 nm) by
increasing the concentration of inorganic phase. Swelling experiments evidenced
that good polymer-filler adhesion was observed in the presence of coupling agents,
while no definite conclusions could be drawn for vulcanized materials produced in
the absence of the coupling agent. The dynamic mechanical behaviour of the
various elastomers became increasingly nonlinear for silica contents higher than
20 wt%. In addition, only in this range of compositions the filler exerted on the
low amplitude storage modulus a remarkable reinforcement, which was related to
the silica content through a power law with exponent a = 4, in agreement with the
prediction of a model proposed by Huber and Vilgis [52] as reported in Fig. 6.

Fig. 6 Huber and Vilgis model: excess storage modulus at the lowest strain amplitude inves-
tigated (cmin) versus silica volume fraction (U0) for IR/SiO2 nanocomposites (open diamond
unvulcanized rubbed; filled circle vulcanized rubber; open circle vulcanized rubber with coupling
agent). The straight line represents the best-fitting power law with exponent a = 4 (see text)
(reproduced with permission from Ref. [51])
In Situ Synthesis of Rubber Nanocomposites 73

A synthetic procedure based on direct addition of water to the reactive medium


and of octyltriethoxysilane as coupling/surfactant agent at different reaction times
was recently proposed by Messori et al. [53]. For long enough reaction time, the
sol–gel process led to a quantitative conversion of TEOS precursor to silica with
the obtainment of a homogeneous distribution of spherically shaped particles. The
delayed addition of octyltriethoxysilane was effective for controlling the average
size and the aggregation phenomena of the in situ generated silica. Dynamic
mechanical analysis carried out on filled IR showed a significant reinforcement (in
terms of storage modulus increment) with respect to the pristine elastomer. On the
other hand, both swelling and extraction tests suggested that the sol–gel process
perturbed the vulcanization process of IR leading to a slight decrease of the cross-
linking degree of the rubber matrix.

3.5 Styrene-Butadiene Rubber Based Nanocomposites

Ikeda et al. [54, 55] described the preparation of styrene-butadiene rubber (SBR)
reinforced with in situ generated silica from TEOS under different catalytic con-
ditions. Composites were prepared by swelling of vulcanized SBR in TEOS or in
TEOS-THF mixture in the presence of basic (n-butylamine) or acid (hydrochloric
acid) catalysts for the sol–gel reaction. Hydrochloric acid was found to be inade-
quate for the sol–gel reaction in the system comprised of vulcanized SBR swollen
in TEOS because the aqueous solution of hydrochloric acid was not well dissolved
in TEOS and, as a consequence, silica was formed only in the surface layer of the
vulcanized rubber [55]. In order to avoid this problem, pre-swelling of vulcanized
SBR was carried out in THF before the sol–gel reaction and, under these conditions,
in both the acidic and basic aqueous solutions mixed with THF, the in situ for-
mation of silica occurred homogeneously in the SBR matrix. The size of in situ
generated silica was observed to be influenced by the cross-linking density, i.e. the
larger the cross-linking density, the smaller the size of in situ formed silica parti-
cles, according to what already reported for other systems [12, 19, 25]. Comparing
with the silica-filled vulcanized SBR prepared by conventional mechanical mixing,
the homogeneity of dispersion of the silica particles was found to be important for
the reinforcement of vulcanized rubber. Stress–strain curves for different materials
are reported in Fig. 7, from which a marked improvement of mechanical properties
(modulus and tensile strength) is evident for composites prepared under basic
condition with respect to SBR unfilled and filled with conventional silica. In
addition, the size of silica particles obviously affects the reinforcement of the
rubber. The larger the in situ silica particles, the better the mechanical properties
found in this study.
Interestingly, and contrary to what generally expected, the tand peaks (Tg)
detected by dynamic mechanical analysis were found to decrease by about 2–4°C
with respect to unfilled rubber. Authors proposed that this to be due to the swelling
in TEOS, which may contribute to the disentanglement of the SBR chains in the
74 M. Messori

Fig. 7 Stress–strain curves


of vulcanized a SBR; b SBR/
SiO2 (mechanically mixed,
SiO2 22 wt%); c SBR/SiO2
(pre-swelling in THF, acid
catalysis, SiO2 13 wt%);
d SBR/SiO2 (pre-swelling in
THF, basic catalysis, SiO2
24 wt%) and e SBR/SiO2
(basic catalysis, SiO2 23
wt%) (reproduced with per-
mission from Ref. [54])

vulcanizate followed by lowering of Tg. The plasticization of the rubber by residual


oligomers from the sol–gel reaction might contribute to this lowering of Tg.
Ikeda et al. [56] also reported an investigation on the effect of TESPT as
coupling agent in the preparation of SBR filled with in situ generated silica. The
presence of TESPT resulted in a much higher reinforcing efficiency with respect to
conventional mechanical mixing and the in situ method without TESPT. The
higher reinforcing efficiency was attributed to the formation of a silica–rubber
network, which also changed the dynamic mechanical behaviour of the vulcanized
rubber. Transmission electron microscopy analysis showed in situ silica incorpo-
ration of very fine particles in comparison to the sol–gel process without TESPT.
Similarly to the above described approach used by Hashim for ENR [49, 50],
de Luca and co-workers [57, 58] prepared a Class II hybrids comprising of
epoxidized SBR, TEOS (as silica precursor) and APTS (as coupling agent) in order
to enhance the interaction between organic and inorganic phases. The epoxy
groups of the rubber reacted with the amino groups of APTS forming an inter-
mediate amino-silane/rubber compound which in turn was reactive towards the
sol–gel reaction of TEOS to silica. The pre-reaction between epoxidized SBR and
APTS was carried out in THF and TEOS and water were subsequently added (the
alkaline pH for the sol–gel synthesis was provided by the APTMS itself). Large
amounts of silica were incorporated using combinations of the inorganic precur-
sors TEOS and APTS. Non-solubility of the materials in THF indicated the for-
mation of a network, the microstructure of which varied according to the
concentrations of the inorganic precursors employed. The mechanical properties
In Situ Synthesis of Rubber Nanocomposites 75

increased considerably by increasing the amount of silica incorporated and SEM


analysis revealed the presence of phase separation for high TEOS content.
In situ silica generation from TEOS has been also applied in the case of blends
of SBR and reclaim rubber (RR) by De and co-workers [59]. They presented an
interesting comparison of mechanical properties of nanocomposites comprised of
SBR/RR blends containing silica both in situ generated and mechanically mixed.
They also investigated the effect of the presence of TESPT as coupling agent.
Authors concluded that the mechanical properties of conventional nanocomposites
were better than those of in situ generated ones in the absence of TESPT but a
reverse trend was observed in the presence of the coupling agent.

3.6 Acrylonitrile-Butadiene Rubber Based Nanocomposites

In situ silica reinforcement was applied to acrylonitrile-butadiene rubber (NBR)


vulcanizates which were swelled in TEOS and subsequently soaked in an aqueous
solution of ethylenediamine [60]. The amount of in situ generated silica within the
NBR vulcanizates (conversion of TEOS to silica of 63%) was limited due to the
high polarity of NBR and the resulting low degree of swelling of NBR in TEOS.
The presence of c-MPS in the rubber vulcanizate increased the conversion of
TEOS to silica during the sol–gel reaction (conversion of TEOS to silica higher
than 90%), compared to the system without c-MPS. The obtained silica particles
were very fine and very homogeneously dispersed.
The same authors also proposed the in situ generation of silica from TEOS
within NBR vulcanizates pre-mixed with conventional silica both in the presence
or not of c-MPS [61]. They observed that the reinforcement efficiency tended to
increase with the increase of mechanically pre-mixed silica. Both transmission
electron and scanning electron microscopies showed that the simultaneous use of
pre-mixed silica and c-MPS promoted the formation of large silica particles and
clusters with a relatively good dispersion by the sol–gel reaction of TEOS in the
rubber vulcanizate as further evidenced by the results of hysteresis measurements.
This behaviour has been attributed to the surface modification of conventional
silica by the sol–gel reaction of TEOS and the presence of c-MPS which worked as
a dispersion agent for silica particles.

3.7 Butadiene Rubber Based Nanocomposites

In situ silica filling of butadiene rubber (BR) was carried out by the sol–gel process
using TEOS [62]. BR was sulphur-cured and the resultant cross-linked BR was
firstly swollen in TEOS and subsequently immersed in an aqueous solution of
n-butylamine at 30°C for 24 h and at 50°C for 72 h to activate the sol–gel reaction
of TEOS. The in situ generated silica was homogeneously dispersed in the rubbery
76 M. Messori

Fig. 8 Stress–strain curves


of vulcanized a BR; b BR/
SiO2 (mechanically mixed,
SiO2 22 wt%); c BR/SiO2 (in
situ generated, SiO2 16 wt%)
(reproduced with permission
from Ref. [62])

matrix with spherical shape and size in the range 15–35 nm. As already reported
for other systems, the size of in situ generated silica was influenced by the cross-
linking density, i.e. the larger the cross-linking density, the smaller the size of in
situ silica formed in the vulcanized rubber. Concurrently, the interaction between
filler and BR matrix seemed to become larger. Compared to the conventional
silica-filled BR vulcanizate, which was prepared by mechanical mixing of the
silica particles, the vulcanized rubber with the in situ generated silica showed
better mechanical properties (in terms of both modulus and strength, as shown in
Fig. 8).
Mark and Zhou reported the preparation and characterization of trans-
1,4-polybutadiene (tPBD) reinforced with in situ generated silica after swelling of
the cross-linked tPBD with TEOS in the presence of a tin-based catalyst and by
absorbing the required water from the external humid atmosphere [63]. The silica
dimensions varied from below 100 nm at low silica content (\2 wt%) to 2–5 lm
at higher content (up to 20 wt%). Concerning the mechanical properties, small
amounts of silica (up to 2 wt%) increased the extensibility and stress at rupture
with a subsequent increment of toughness.

3.8 Acrylic Rubber Based Nanocomposites

Following an approach very similar to those of ENR, Bhowmick and Bandyo-


padhyay [46, 64] reported the preparation and the characterization of acrylic
rubber (ACR) reinforced with in situ generated silica by means the acid-catalyzed
In Situ Synthesis of Rubber Nanocomposites 77

hydrolysis-condensation of TEOS in solution of THF (procedure A). The pro-


portion of TEOS was from 0 to 50 wt% (macrophase separation occurred beyond
this last maximum content of TEOS). For comparison, composites were also
prepared with preformed silica up to 30 wt% of its loading. After solvent elimi-
nation, some sample was vulcanized with benzoyl peroxide or with a mixed curing
system comprised of hexamethylenediamine and ammonium benzoate. All the
hybrid composites prepared by sol–gel technique were transparent while, on the
contrary, the composites of ACR with mechanically mixed preformed silica were
opaque at all compositions. Infrared spectroscopic analysis revealed the absence of
significant shifts in peak position for carbonyl absorption in all the hybrid com-
posites, suggesting that a homogeneous dispersion of silica particles within the
organic matrix occurred without any significant interaction between the two
phases. This conclusion was further supported by the complete dissolution of the
uncured hybrid in THF at room temperature. Morphological analysis revealed the
presence of discrete spherical silica particles with an average diameter that
increase from 20 to 90 nm by increasing the TEOS content (from 10 to 50 wt%).
The storage modulus detected by dynamic mechanical analysis above glass
transition temperature increased by increasing the TEOS content according to the
reinforcing effect of nanosilica network structure present in the rubber matrix. The
tand peak height was gradually reduced, the peak broadened and the Tg values of
the composites shifted towards higher temperatures by increasing the inorganic
filler content in the composites.
Concerning mechanical properties of the hybrid composites, the tensile strength
increased by increasing the silica content. For the same silica content, the tensile
strength of composites filled with in situ generated silica was significantly higher
than that of conventional composites. The nanolevel mixing of in situ generated
silica particles and the formation of strong Si–O–Si network within the ACR
matrix were responsible for greater reinforcement in the rubber matrix. In the case
of conventional silica filled rubber, the bigger sized silica particles did not provide
higher surface area for interactions such as that of in situ generated silica and
hence they were not as effective as the former.
On curing the rubber phase, with either benzoyl peroxide or hexamethylen-
ediamine/ammonium benzoate, the tensile strength was further increased, although
benzoyl peroxide cured samples showed lesser improvements than the corre-
sponding mixed cross-linking system.
In order to verify the effect of the polarity of the rubber matrix on the properties
of this class of composites, Bhowmick and co-workers [65, 66] synthesized acrylic
copolymers and terpolymers by bulk polymerization of ethyl acrylate (EA), butyl
acrylate (BA) and acrylic acid (AA) and used them as rubber matrix for the
preparation of silica filled composites by using the usual procedure based on the
acid-catalyzed hydrolysis-condensation of TEOS in THF solution. Authors
reported a morphological investigation showing the presence of silica particles
with an average diameter lower than 100 nm. The average diameter of silica
particles was lower in the case of high extent of polarity and hydrophilicity of the
rubber matrix, suggesting that these molecular parameters control the in situ
78 M. Messori

generated silica particles size. In all the cases, no significant chemical interaction
occurred at the silica–rubber interface, as depicted from spectroscopic analysis
(FT-IR and NMR). In general, the mechanical properties increased by increasing
the polarity of the matrix due to better rubber-silica interaction and uniform
nanosilica dispersion. Also in this case, cross-linking further improved the
mechanical properties.

3.9 Ethylene–Propylene–Diene Monomer Rubber Based


Nanocomposites

Ethylene–propylene–diene monomer (EPDM) rubber was modified with TESPT


in an internal mixer in order to graft triethoxysilyl groups onto the polymeric
backbone [67]. Triethoxysilyl-grafted EPDM sheets were swollen with TEOS
and subsequently immersed in n-butyl amine aqueous solution to in situ generate
silica particles. The silica filled EPDM rubber was finally mixed with vulcanizing
agents and cured by hot pressing. Authors concluded that TESPT fragments
present in the macromolecular chain of EPDM acted as nucleation sites for the
growing of silica which in turn led to the formation of a strong silica-EPDM
network chemically bonded to the rubber matrix. The pendant triethoxysilyl
groups on the EPDM backbone played a fundamental role in uniformly dis-
persing silica particles within the rubber. For similar silica loadings, EPDM
modified with in situ generated silica showed superior reinforcing efficiency with
respect to materials prepared by ex situ process (mechanical mixing of precipi-
tated silica within the rubber).

3.10 Other Rubber Based Nanocomposites

Matêjka and co-workers [68, 69] reported some studies on rubbery epoxy resins
reinforced with in situ generated silica from TEOS and with a silica content
ranging from 6 to 22 wt%. The epoxy resin and the hardener were diglycidyl ether
of bisphenol A (DGEBA) and polyoxypropylene-diamine (JeffamineÒ D2000),
respectively. Long flexible polyether chain lead to a rubbery cured material and
moreover it solubilizes siloxane structures formed in the sol–gel process resulting
in a transparent hybrid. Authors reported several preparation methods:
(i) one-stage process, in which all reactants were mixed in iso-propyl alcohol
solution and reacted simultaneously;
(ii) two-stage ‘simultaneous’ process, in which TEOS was pre-hydrolyzed under
acidic conditions in iso-propyl alcohol solution and subsequently added to
DGEBA and JeffamineÒ D2000 to activate the simultaneous formation of
both organic and inorganic networks;
In Situ Synthesis of Rubber Nanocomposites 79

(iii) two-stage ‘sequential’ process with preformed epoxide network, in which the
cured epoxy resin was firstly prepared by stoichiometric reaction between
DGEBA and JeffamineÒ D2000 and subsequently swelled in the sol–gel
solution (TEOS, water and acid in iso-propyl alcohol) to activate the in situ
generation of silica.
Compact and large silica aggregates (100–300 nm in diameter) were observed in
the case of hybrids obtained with one-stage process, whereas open polymeric
structures of smaller aggregates (50–100 nm in diameter) composed of fractal
particles were formed during the two-stage procedure with the pre-hydrolyzed
TEOS. Moreover, this two-stage process resulted in a much faster gelation of the
silica. In these ‘simultaneous’ processes, the silica network was formed faster than
the epoxide one and its structure was not influenced by the presence of the epoxide
and amine. However, in the case of the two-stage ‘sequential’ process, the pre-
formed organic network suppressed growth of the silica aggregates by inter-particle
condensation and the relatively smallest silica domains (10–20 nm) were formed.
Authors concluded that the synthetic procedure controls the reaction mecha-
nism, final structure and morphology. In particular, contrary to what shown by
basic and neutral catalysis, acid catalysis promotes rapid hydrolysis of siloxane
groups which results in the high content of silanol groups and an extensive grafting
to the epoxide network. Grafting between organic and inorganic phases resulted in
the formation of an interphase epoxide layer with reduced mobility and increased
glass transition temperature.
Concerning the mechanical properties, an increase in modulus by two orders of
magnitude was achieved at a low silica content (10 vol%). Interestingly, dynamic
mechanical analysis revealed the presence of a co-continuous morphology of the
epoxy matrix and of the silica phase (and the silica-glassy epoxide phase) con-
tinuously extending through the macroscopic sample.
Sunada and co-workers [70] reported the synthesis of alkoxysilane-modified
polychloroprene latex by the emulsion copolymerization of 2-(3-triethoxysilyl-
propyl)-1,3-butadiene and chloroprene. This latex was mixed with unmodified
polychloroprene (CR) latex and TEOS to obtain composites by sol–gel reaction in
the latex under basic catalysis. After vacuum drying, the composites were com-
pounded with curing agents by two-roll milling and subjected to vulcanization by
hot pressing. Electron microscopy showed that the silica particles in unvulcanized
composites had various diameters ranging from 0.1 to 0.6 lm, and their size
became larger with the decrease of the silica content. Compared to similar com-
posites prepared by conventional mechanical mixing with preformed silica, vul-
canized CR/SiO2 composites obtained with in situ generation of filler showed that
the tensile modulus and tear strength improved with an increase of the amount of
modified CR.
The metallocene-based poly(ethylene-octene) (POE) elastomer, which was
developed using a metallocene catalyst by Dow and Exxon, has received much
attention because of its peculiar properties such as uniform distribution of
comonomer content and narrow molecular weight distribution.
80 M. Messori

Wu et al. [71, 72] reported a systematic investigation on composites comprised


of POE or POE grafted with maleic anhydride or acrylic acid (POE-g-MAH or
POE-g-AA, respectively) as elastomeric matrix and in situ generated silica or
silica–titania mixed oxides. The composites were prepared by addition of solutions
of (i) TEOS, water and hydrochloric acid or (ii) silicic acid, tetraisopropyl orth-
otitanate (TTIP), water and hydrochloric acid to POE or POE-g-MAH or POE-g-
AA melted in an internal mixer at a temperature of 160–170°C.
FT-IR, 29Si-NMR and XRD analysis showed that Si–O–C, Ti–O–C, Ti–O–Ti,
Si–O–Si and Si–O–Ti linkages were formed in silica and silica–titania reinforced
elastomers supporting the expected formation of covalent bonds between organic
and inorganic phases and of SiO2–TiO2 mixed oxides. It was found that there are
maximum values of tensile strength and glass transition temperature at about
10 wt% of inorganic filler (see Table 3). Authors explained this behaviour
assuming that an excess of SiO2 or SiO2–TiO2 particles might cause separation
between the organic and inorganic phases with a reduction of their compatibility.
From the reported data it is interesting to observe that superior properties were
obtained in the case of hybrids based on POE grafted with maleic anhydride or
acrylic acid (POE-g-MAH or POE-g-AA) with respect to pure POE. The higher
polarity and the presence of reactive groups (anhydride or carboxylic) in the
grafted elastomer allowed the formation of stronger chemical bonds between
organic and inorganic phases.
The literature concerning the in situ generation of inorganic oxides through the
sol–gel process within thermoplastic elastomers (TPE) is relatively limited.
Lai et al. [73, 74] reported the preparation of thermoplastic polyurethane (TPU)
reinforced with in situ generated silica starting from TEOS as precursor or from
polysilicic acid extracted from an aqueous sodium metasilicate solution with THF.
A solution procedure (procedure A) with acidic catalysis (hydrochloric acid or
acetic acid) was adopted in both approaches to prepare the hybrid materials.
Spectroscopic and dynamic mechanical analysis showed the presence of
interfacial interactions between the organic and inorganic phases. Storage modulus
increased at all concentrations of silica due to its reinforcement effect while the
tensile strength exhibited a maximum value for filler content of 10–15 wt%. In
contrast, the cutting strength decreased, probably due to a reduction of the energy
dissipation from silica as physical cross-links. The acetic acid catalyzed system

Table 3 Glass transition temperature and tensile strength (TS) values POE-based composites
Filler POE/SiO2 POE-g-MAH/SiO2 POE/SiO2–TiO2 POE-g-AA/SiO2–TiO2
content
Tg TS Tg TS Tg TS Tg TS
(wt%)
(°C) (MPa) (°C) (MPa) (°C) (MPa) (°C) (MPa)
0 -60 27 -58 15 -60 27 -59 17
3 -58 28 -56 38 -58 28 -54 37
10 -55 31 -52 52 -56 32 -48 56
20 -58 26 -56 27 -58 28 -57 30
In Situ Synthesis of Rubber Nanocomposites 81

showed better optical properties than the hydrochloric acid catalyzed system. The
modified process with extraction showed improved mechanical properties and
better optical properties than the conventional sol–gel process without extraction.

4 Concluding Remarks

The sol–gel process has been established as a versatile, effective and innovative
method for the in situ preparation of rubber nanocomposites. The great majority of
the published studies concerns the generation of silica from TEOS within the most
common elastomers even if some example of in situ generation of other oxides
such titania, zirconia, alumina or mixed oxides is present in the literature. Different
synthetic strategies can be applied to obtain materials with tailored properties, in
particular the in situ generation of inorganic fillers can be obtained in solution or
directly in the rubber, either swollen in metal alkoxides or in the melt state.
Depending on the preparation conditions, the vulcanization of the rubber matrix
can be carried out before, simultaneously or after the sol–gel process and the
resulting morphologies and ultimate properties are generally affected by the spe-
cific synthetic procedure. As a general observation, finer morphologies (lower
filler dimensions) are realized by increasing the cross-linking density of the vul-
canized rubber presumably due to constraining effects deriving from the organic
network. It seems also very promising that the mechanical properties (modulus,
strength and extensibility) of the in situ filled nanocomposites are generally better
than those of the corresponding materials prepared with the conventional
mechanical mixing of preformed particulates and elastomers. This enhancement is
generally attributed to a lower tendency to filler–filler aggregation due to a lower
particle surface interaction resulting from the ‘bottom-up approach’ of the sol–gel
process applied to the preparation of organic–inorganic hybrid materials.
Future developments on in situ synthesis of rubber nanocomposites should
involve further systematic studies on the possibility to use silica precursors
alternative to TEOS as well as the incorporation of metal oxides different than
silica, exploring the different reinforcing effect of titania, zirconia or other. Also
the use of non-hydrolytic sol–gel process, which is characterized by the absence of
water as reactant for the generation of inorganic oxides, should be an interesting
field of research and development.

References

1. Donnet, J.-B., Custodero, E.: Reinforcement of elastomers by particulate fillers. In: James,
E.M., Burak, E., Frederick, R.E. (eds.) Science and Technology of Rubber, 3rd edn.
Academic Press, Burlington (2005)
2. Klein, L.C., Wojcik, A.B.: Polymer-ceramic nanocomposites: polymer overview. In:
Buschow Jr, K.H., Robert, W.C., Merton, C.F., et al. (eds.) Encyclopedia of Materials:
Science and Technology. Elsevier, Oxford (2001)
82 M. Messori

3. Hewitt, N.: Silica as a reinforcing filler. In: Compounding Precipitated Silica in Elastomers.
William Andrew Publishing, Norwich, NY (2007)
4. Brinker, C., Scherer, G.: Sol–Gel Science: the Physics and Chemistry of Sol–Gel Processing.
Academic Press, Boston (1990)
5. Sanchez, C., Julian, B., Belleville, P., et al.: Applications of hybrid organic–inorganic
nanocomposites. J. Mater. Chem. 15, 3559–3592 (2005)
6. Novak, B.M.: Hybrid nanocomposite materials—between inorganic glasses and organic
polymers. Adv. Mater. 5, 422–433 (1993)
7. Mark, J.E.: The sol–gel route to inorganic–organic composites. Heterogeneous Chem. Rev. 3,
307–326 (1996)
8. Mark, J.E.: Ceramic-reinforced polymers and polymer-modified ceramics. Polym. Eng. Sci.
36, 2905–2920 (1996)
9. Mark, J.E.: Some novel polymeric nanocomposites. Acc. Chem. Res. 39, 881–888 (2006)
10. Mark, J.E.: Ceramic-modified elastomers. Curr. Opin. Solid State Mater. Sci. 4, 565–570
(1999)
11. Rajan, G.S., Sur, G.S., Mark, J.E., et al.: Preparation and characterization of some unusually
transparent poly(dimethylsiloxane) nanocomposites. J. Polym. Sci. Polym. Phys. 41, 1897–
1901 (2003)
12. Breiner, J.M., Mark, J.E., Beaucage, G.: Dependence of silica particle sizes on network chain
lengths, silica contents, and catalyst concentrations in in situ-reinforced polysiloxane
elastomers. J. Polym. Sci. Polym. Phys. 37, 1421–1427 (1999)
13. Li, Z.L., Han, W., Kozodaev, D., et al.: Surface properties of poly(dimethylsiloxane)-based
inorganic/organic hybrid materials. Polymer 47, 1150–1158 (2006)
14. Dewimille, L., Bresson, B., Bokobza, L.: Synthesis, structure and morphology of poly
(dimethylsiloxane) networks filled with in situ generated silica particles. Polymer 46, 4135–
4143 (2005)
15. Bokobza, L.: New developments in rubber reinforcement. Kgk-Kaut Gummi Kunst 62, 23–27
(2009)
16. Bokobza, L.: Elastomeric composites. I. Silicone composites. J. Appl. Polym. Sci. 93, 2095–
2104 (2004)
17. Murugesan, S., Mark, J.E., Beaucage, G.: Structure-property relationships for
poly(dimethylsiloxane) networks in situ filled using titanium 2-ethylhexoxide and zirconium
n-butoxide. ACS Symp. Ser. 838, 163–169 (2003)
18. Murugesan, S., Sur, G.S., Mark, J.E., et al.: In situ catalyst generation and controlled
hydrolysis in the sol–gel precipitation of zirconia and titania particles in poly
(dimethylsiloxane). J. Inorg. Organomet. P 14, 239–252 (2004)
19. Wen, J., Mark, J.E.: Precipitation of silica–titania mixed-oxide fillers into poly
(dimethylsiloxane) networks. Rubber Chem. Technol. 67, 806–819 (1994)
20. Kraus, G.: J. Appl. Polym. Sci. 7, 861 (1963)
21. Mark, J.E.: Rubber Chem. Technol. 48, 495 (1975)
22. Mark, J.E., Flory, P.J.: Stress–strain isotherms for poly-(dimethylsiloxane) networks. J. Appl.
Phys. 37, 4635–4639 (1966)
23. Wen, J.Y., Mark, J.E.: Sol–gel preparation of composites of poly(dimethylsiloxane) with
SiO2 and SiO2/TiO2 and their mechanical properties. Polym. J. 27, 492–502 (1995)
24. Breiner, J.M., Mark, J.E.: Preparation, structure, growth mechanisms and properties of
siloxane composites containing silica, titania or mixed silica–titania phases. Polymer 39,
5483–5493 (1998)
25. Wen, J.A., Mark, J.E.: Synthesis, structure, and properties of poly(dimethylsiloxane)
networks reinforced by in situ-precipitated silica–titania, silica–zirconia, and silica–alumina
mixed oxides. J. Appl. Polym. Sci. 58, 1135–1145 (1995)
26. Bokobza, L., Chauvin, J.P.: Reinforcement of natural rubber: use of in situ generated silicas
and nanofibres of sepiolite. Polymer 46, 4144–4151 (2005)
27. Bokobza, L.: Some new developments in rubber reinforcement. Compos. Interface 13, 345–
354 (2006)
In Situ Synthesis of Rubber Nanocomposites 83

28. Guth, O., Gold, E.: Phys. Rev. 53, 322 (1938)
29. Guth, O.: J. Appl. Phys. 16, 20 (1945)
30. Fragiadakis, D., Pissis, P., Bokobza, L.: Modified chain dynamics in poly(dimethylsiloxane)/
silica nanocomposites. J. Non-Cryst. Solids 352, 4969–4972 (2006)
31. Fragiadakis, D., Pissis, P., Bokobza, L.: Glass transition and molecular dynamics in
poly(dimethylsiloxane)/silica nanocomposites. Polymer 46, 6001–6008 (2005)
32. Poompradub, S., Chaichua, B., Kanchanaamporn, C., et al.: Synthesis of silica in natural
rubber solution via sol–gel reaction. Kgk-Kaut Gummi Kunst 61, 152–155 (2008)
33. Chaichua, B., Prasassarakich, P., Poompradub, S.: In situ silica reinforcement of natural
rubber by sol–gel process via rubber solution. J. Sol-Gel Sci. Technol. 52, 219–227
(2009)
34. Ikeda, Y., Poompradub, S., Morita, Y., et al.: Preparation of high performance nanocomposite
elastomer: effect of reaction conditions on in situ silica generation of high content in natural
rubber. J. Sol-Gel Sci. Technol. 45, 299–306 (2008)
35. Ikeda, Y., Kameda, Y.: Preparation of ‘‘green’’ composites by the sol–gel process: in situ
silica filled natural rubber. J. Sol-Gel Sci. Technol. 31, 137–142 (2004)
36. Poompradub, S., Kohjiya, S., Ikeda, Y.: Natural rubber/in situ silica nanocomposite of a high
silica content. Chem. Lett. 34, 672–673 (2005)
37. Kohjiya, S., Murakami, K., Iio, S., et al.: In situ filling of silica onto ‘‘green’’ natural rubber
by the sol–gel process. Rubber Chem. Technol. 74, 16–27 (2001)
38. Murakami, K., Iio, S., Tanahashi, T., et al.: Reinforcement of NR by silica generated in situ:
comparison with carbon black stock. Kaut Gummi Kunstst 54, 668–672 (2001)
39. Murakami, K., Iio, S., Ikeda, Y., et al.: Effect of silane-coupling agent on natural rubber filled
with silica generated in situ. J. Mater. Sci. 38, 1447–1455 (2003)
40. Kohjiya, S., Ikeda, Y.: In situ formation of particulate silica in natural rubber matrix by the
sol–gel reaction. J. Sol-Gel Sci. Technol. 26, 495–498 (2003)
41. Kohjiya, S., Kato, A., Ikeda, Y.: Visualization of nanostructure of soft matter by 3D-TEM:
nanoparticles in a natural rubber matrix. Prog. Polym. Sci. 33, 979–997 (2008)
42. Kohjiya, S., Katoh, A., Shimanuki, J., et al.: Three-dimensional nano-structure of in situ silica
in natural rubber as revealed by 3D-TEM/electron tomography. Polymer 46, 4440–4446
(2005)
43. Tangpasuthadol, V., Intasiri, A., Nuntivanich, D., et al.: Silica-reinforced natural rubber
prepared by the sol–gel process of ethoxysilanes in rubber latex. J. Appl. Polym. Sci. 109,
424–433 (2008)
44. Siramanont, J., Tangpasuthadol, V., Intasiri, A., et al.: Sol–gel process of alkyltriethoxysilane
in latex for alkylated silica formation in natural rubber. Polym. Eng. Sci. 49, 1099–1106
(2009)
45. Bandyopadhyay, A., De Sarkar, M., Bhowmick, A.K.: Epoxidized natural rubber/silica
nanoscale organic–inorganic hybrid composites prepared by sol–gel technique. Rubber
Chem. Technol. 77, 830–846 (2004)
46. Bandyopadhyay, A., Maiti, M., Bhowmick, A.K.: Synthesis, characterisation and properties
of clay and silica based rubber nanocomposites. Mater. Sci. Tech.-Lond. 22, 818–828
(2006)
47. Bandyopadhyay, A., De Sarkar, M., Bhowmick, A.K.: Polymer-filler interactions in sol–gel
derived polymer/silica hybrid nanocomposites. J. Polym. Sci. Polym. Phys. 43, 2399–2412
(2005)
48. Bandyopadhyay, A., De Sarkar, M., Bhowmick, A.K.: Structure–property relationship in sol–
gel derived polymer/silica hybrid nanocomposites prepared at various pH. J. Mater. Sci. 41,
5981–5993 (2006)
49. Hashim, A.S., Kohjiya, S., Ikeda, Y.: Moisture cure and in situ silica reinforcement of
epoxidized natural-rubber. Polym. Int. 38, 111–117 (1995)
50. Hashim, A.S., Kawabata, N., Kohjiya, S.: Silica reinforcement of epoxidized natural rubber
by the sol–gel method. J. Sol-Gel Sci. Technol. 5, 211–218 (1995)
84 M. Messori

51. Messori, M., Bignotti, F., De Santis, R., et al.: Modification of isoprene rubber by in situ
silica generation. Polym. Int. 58, 880–887 (2009)
52. Huber, G., Vilgis, T.A.: Universal properties of filled rubbers: mechanisms for reinforcement
on different length scales. Kaut Gummi Kunstst 52, 102–107 (1999)
53. Messori, M., Fiorini, M.: In situ formation of silica particles in isoprene rubber by the sol–gel
method. J. Sol-Gel Sci. Technol. (2010, submitted)
54. Ikeda, Y., Tanaka, A., Kohjiya, S.: Reinforcement of styrene-butadiene rubber vulcanizate by
in situ silica prepared by the sol–gel reaction of tetraethoxysilane. J. Mater. Chem. 7, 1497–
1503 (1997)
55. Ikeda, Y., Tanaka, A., Kohjiya, S.: Effect of catalyst on in situ silica reinforcement of
styrene-butadiene rubber vulcanizate by the sol–gel reaction of tetraethoxysilane. J. Mater.
Chem. 7, 455–458 (1997)
56. Hashim, A.S., Azahari, B., Ikeda, Y., et al.: The effect of bis(3-triethoxysilylpropyl)tetrasulfide
on silica reinforcement of styrene-butadiene rubber. Rubber Chem. Technol. 71, 289–299
(1998)
57. de Luca, M.A., Jacobi, M.M., Orlandini, L.F.: Synthesis and characterisation of elastomeric
composites prepared from epoxidised styrene butadiene rubber, 3-aminopropyltriethoxysilane
and tetraethoxysilane. J. Sol-Gel Sci. Technol. 49, 150–158 (2009)
58. de Luca, M.A., Machado, T.E., Notti, R.B., et al.: Synthesis and characterization of
epoxidized styrene-butadiene rubber/silicon dioxide hybrid materials. J. Appl. Polym. Sci.
92, 798–803 (2004)
59. De, D., Das, A., Panda, P.K., et al.: Reinforcing effect of silica on the properties of styrene
butadiene rubber-reclaim rubber blend system. J. Appl. Polym. Sci. 99, 957–968 (2006)
60. Tanahashi, H., Osanai, S., Shigekuni, M., et al.: Reinforcement of acrylonitrile-butadiene
rubber by silica generated in situ. Rubber Chem. Technol. 71, 38–52 (1998)
61. Murakami, K., Osanai, S., Shigekuni, M., et al.: Silica and silane coupling agent for irt situ
reinforcement of acrylonitrile-butadiene rubber. Rubber Chem. Technol. 72, 119–129 (1999)
62. Ikeda, Y., Kohjiya, S.: In situ formed silica particles in rubber vulcanizate by the sol–gel
method. Polymer 38, 4417–4423 (1997)
63. Zhou, D.H., Mark, J.E.: Preparation and characterization of trans-1,4-polybutadiene
nanocomposites containing in situ generated silica. J. Macromol. Sci. Pure A41, 1221–1232
(2004)
64. Bandyopadhyay, A., Bhowmick, A.K., De Sarkar, M.: Synthesis and characterization of
acrylic rubber/silica hybrid composites prepared by sol–gel technique. J. Appl. Polym. Sci.
93, 2579–2589 (2004)
65. Patel, S., Bandyopadhyay, A., Vijayabaskar, V., et al.: Effect of acrylic copolymer and
terpolymer composition on the properties of in situ polymer/silica hybrid nanocomposites. J.
Mater. Sci. 41, 927–936 (2006)
66. Patel, S., Bandyopadhyay, A., Vijayabaskar, V., et al.: Effect of microstructure of acrylic
copolymer/terpolymer on the properties of silica based nanocomposites prepared by sol–gel
technique. Polymer 46, 8079–8090 (2005)
67. Das, A., Jurk, R., Stockelhuber, K.W., et al.: Silica-ethylene propylene diene monomer
rubber networking by in situ sol–gel method. J. Macromol. Sci. Part A-Pure Appl. Chem. 45,
101–106 (2008)
68. Matêjka, L., Dukh, O., Kolarik, J.: Reinforcement of crosslinked rubbery epoxies by in situ
formed silica. Polymer 41, 1449–1459 (2000)
69. Matêjka, L., Dusek, K., Plestil, J., et al.: Formation and structure of the epoxy-silica hybrids.
Polymer 40, 171–181 (1999)
70. Sunada, K., Takenaka, K., Shiomi, T.: Synthesis of polychloroprene-silica composites by
sol–gel method in the presence of modified polychloroprene containing triethoxysilyl group.
J. Appl. Polym. Sci. 97, 1545–1552 (2005)
71. Wu, C.S., Liao, H.T.: Modification of polyethylene-octene elastomer by silica through a sol–
gel process. J. Appl. Polym. Sci. 88, 966–972 (2003)
In Situ Synthesis of Rubber Nanocomposites 85

72. Wu, C.S.: Synthesis of polyethylene-octene elastomer/SiO2–TiO2 nanocomposites via in situ


polymerization: properties and characterization of the hybrid. J. Polym. Sci. Polym. Chem.
43, 1690–1701 (2005)
73. Lai, S.M., Wang, C.K., Shen, H.F.: Properties and preparation of thermoplastic polyurethane/
silica hybrid using sol–gel process. J. Appl. Polym. Sci. 97, 1316–1325 (2005)
74. Lai, S.M., Liu, S.D.: Properties and preparation of thermoplastic polyurethane/silica hybrids
using a modified sol–gel process. Polym. Eng. Sci. 47, 77–86 (2007)
Part II
Characterization & Properties
Relaxation Phenomena in Elastomeric
Nanocomposites

G. C. Psarras and K. G. Gatos

Abstract Elastomeric nanocomposites are technologically important engineering


materials, mostly because of their thermomechanical and electrical behaviour.
Relaxation phenomena arising in rubber nanocomposites include contributions
from both the elastomeric matrix and the presence of nanofiller, and in many cases
reflect the interactions between matrix and filler. Dynamic mechanical analysis
and broadband dielectric spectroscopy are two mutual complementary experi-
mental techniques, which record the relative response of the tested material under
the influence of a harmonically time varying mechanical or electrical field.
Occurring relaxations originate from molecular dynamic effects, interfacial phe-
nomena and phase changes. Studying relaxation phenomena provide valuable
information upon the structure–property relationships of the nanosystems.
Elastomeric nanocomposites can be classified according the employed matrix to
non-polar and polar. Additionally, can be classified according the aspect ratio of
the used filler. Suitably selecting the type and the amount of nanoinclusions the
service performance of nanocomposites can be tailored.

Abbreviations
ABS Antilock braking system
AMIC 1-Allyl-3methyl imidazolium chloride
BDS Broadband dielectric spectroscopy

G. C. Psarras (&)
Department of Materials Science, School of Natural Sciences, University of Patras,
26504 Patras, Greece
e-mail: G.C.Psarras@upatras.gr
K. G. Gatos
Megaplast S.A. Research & Development Center, 4 Makedonias Str., 16672, Athens,
Greece
e-mail: kgatos@gmail.com

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 89


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_3,
Ó Springer-Verlag Berlin Heidelberg 2011
90 G. C. Psarras and K. G. Gatos

BR Butadiene rubber
CB Carbon black
CNT Carbon nanotube
CP Cross polarization
DEA Dynamic electrical analysis
DMA Dynamic mechanical analysis
DSC Differential scanning calorimetry
ENR Epoxidized natural rubber
EPDM Ethylene-propylene-diene rubber
FGPNC Functionally graded polymeric nanocomposite
FHT Sodium fluorohectorite
HNBR Hydrogenated nitrile rubber
HXNBR Hydrogenated carboxylated acrylonitrile butadiene rubber
IP Interfacial polarization
LS Layered silicates
MMT Montmorillonite
MQ(NMR) Multiple quantum (NMR)
MWCNT Multi-wall carbon nanotubes
MWS Maxwell–Wagner–Sillars
NMR Nuclear magnetic resonance
NR Natural rubber
OMMT Organically modified montmorillonite
phr Parts per hundred rubber
POSS Polyhedral oligomeric silsesquioxanes
PUR Polyurethane
SBR Styrene butadiene rubber
SDS Sodium dodecyl sulfate
TEM Transmission electron microscopy
UDPNC Uniformly dispersed polymeric nanocomposite
VFT Vogel–Fulcher–Tamann
XNBR Carboxylated nitrile rubber
XRD X-ray diffraction

1 Introduction

The term relaxation phenomenon or effect refers to the process, which undergo a
system or substance under the influence of an external field by changing its ori-
ginal equilibrium state to a new dynamic one. After removing the exciting force
the system relaxes back to its initial equilibrium state [1]. The time scale of the
system’s response function compared to the exerting time of the external field
governs the relaxation process. If the time scale of the disturbing field is much
Relaxation Phenomena in Elastomeric Nanocomposites 91

longer than the system’s response time, then response function is allowed to reach
its maximum value. If, on the other hand, the time scale of the field is short relative
to the system’s response time, then limited variation of the equilibrium position is
expected to occur, holding the response function to low or even minimum value.
The application of an external electric field on a polar material constitutes a typical
example of such a process. If the time of applying the field is long enough for the
dipoles to be orientated to the direction of the field then the material becomes fully
polarized and its response function (polarization) takes maximum value. On the
contrary, if the time needed by the dipoles to become parallel to the applied field is
much longer than the time of field application a weak disturbance of the initial
state will take place and the achieved polarization of the material will remain low
[1, 2].
Relaxation phenomena occur in many different materials (such as glasses,
liquids, suspensions, disordered solids, ceramics, magnetic materials, polymers,
composites, etc.). Among them, elastomers and their micro- and nano-composites
have attracted the interest of the scientific community. Studying their relaxation
phenomena provide valuable information concerning the relaxation time, the
distribution or not of relaxation times, the activation energy of the process, the
temperature and/or pressure dependency of the effect and the dynamics of
molecular entities with respect to the applied stimulus and the exerted thermo-
dynamic variables.
Rubber technologists belong to the pioneers who utilized fillers of nano-
dimensions (e.g. carbon black) however, recent practice reveals their stiffness
in accommodating the late advances in nanotechnology. The reason behind
is likely the multi-component rubber vulcanization recipes, along with the
complex chemistry involved during curing. Therefore, restrained utilization of
novel fillers within the conservative and rather empirically built formulations took
place.
Since the initiation of the industrial applications of elastomers in the nineteenth
century, fillers played a dominant role. Reinforcement of rubber stocks with chalk,
talc, kaolin, zinc oxide and CB were among the first, which were exploited [3].
Usually, the fillers are classified as active and inactive. In the first case the fillers
aim mainly to reinforce the rubber stock, whereas in the latter case the fillers act as
extenders assisting the compounding process. Several factors have been initially
recognized to determine the action of the fillers within an elastomeric matrix.
These involve (i) the extend of the interface between elastomer and filler usually
expressed in square meters per gram of filler, (ii) the nature of this surface evident
in adsorption properties and chemical reactivity, (iii) the shape of filler particles
involving aggregates and agglomerates, defining the so-called ‘structure’ and
(iv) the porosity of the filler particles [4]. The latter characteristic is structurally
classified into micro-, meso- and macro-porous with pore sizes of less than 2 nm,
between 2 and 50 nm and larger than 50 nm, respectively [5].
The selection of the appropriate filler to reinforce a specific rubber requires a
number of considerations. A high specific surface area along with the respec-
tive loading determines the effective contact area between filler and rubber [6].
92 G. C. Psarras and K. G. Gatos

This constitutes a requisite for enhanced reinforcement emanated by the


increased occluded rubber. Additionally, the type and number of the active sites
on a filler surface plays a crucial role on the degree of bonding with the elas-
tomeric matrix. In this respect treatment with coupling agents is able to modify
accordingly the surface energy of the related filler [7, 8]. The induced rubber–
filler interactions restrain the mobility of the macromolecular chains in the
vicinity of the latter. This conformational entropy loss of the macromolecular
chain is compensated by its enthalpy gain. For uncured compounds the created
‘bound-rubber’ remains un-dissolved during an extraction process in a good
solvent [9, 10]. On the other hand, in vulcanizates, that immobilized rubber is
characterized as rubber shell. Its presence results in an interphase creation with
discernible properties [11]. Two or more fillers might come close enough so that
a joint rubber shell is created. Usually, a filler network formed via joint shell
construction is less rigid than that formed by direct contact of fillers. The forces
induced by surface charges of fillers are able to give rise to agglomeration.
Rubber trapped within a filler network would behave more as filler than as a
matrix, at least in terms of mechanical properties. Under circumstances, in
reinforced vulcanizates, overlaps among the three types of immobilized rubber
can be in action i.e. part of occluded rubber may also be trapped rubber or rubber
shell and vice versa [11].
The agglomerates found in a rubber matrix comprising of several small stable
units called aggregates. These latter are made of a number of homogeneous
primary entities tight together. For example, CB has a primary particle with
cross-sectional dimension of 5–100 nm, whereas its aggregates consist of mul-
tiple primary particles and present dimensions of 100–500 nm. Condensing a
plethora of aggregates into agglomerates a cross-sectional dimension of 1–40 lm
is obtained [12]. The term ‘structure’ used in the literature for CB describes the
arrangement of aggregates while the term ‘secondary structure’ is connected with
the conformation of its agglomerates [3]. Nowadays, the fillers mixed with
elastomeric matrices have variety of sizes and dimensions. These might include
spherical-like (i.e. 0D) [13] and rod-like forms (i.e. 1D) [14–16] to platelet-like
(i.e. 2D) [17] and other complex configurations (i.e. 3D) [18, 19]. Therefore,
their ‘structure’ and eventually, their ‘secondary structure’ are diversified
accordingly. Crucial role on the dispersion of the fillers within the elastomeric
matrix plays the character of the rubber itself. Polar or non-polar elastomeric
matrices must be chosen with caution so that to fit to the surface activity of the
respective filler, and vice versa.
The quality of rubber–filler interphase determines in great extent the properties
of the vulcanizate. Its response to external stimuli is differentiated as the vulca-
nizate operates from low temperatures (i.e. glassy behaviour) to temperatures
above the glass transition (i.e. rubber-like behaviour) [20]. At the same time, the
relaxation response of the overall reinforced elastomer is altered, accordingly. The
methods developed to monitor the relaxation performance of the vulcanizate
supply valuable information for the determination of its structure-property
relationships.
Relaxation Phenomena in Elastomeric Nanocomposites 93

2 Relaxation Analysis Methods

2.1 Dynamic Mechanical

The type and extend of the segmental motion of the macromolecular chain in elas-
tomeric nanocomposites along with its relaxation phenomena is linked to the
physical properties of the vulcanizates. The restrained mobility of the rubber in the
vicinity of filler alters its response to an applied force. This alteration is able to
be closely examined when the force is periodic, induces limited deformation and
does not cause any failure or appreciable fatigue [21]. Additionally, applying a
sinusoidal stress and performing a temperature or frequency sweep across the
desirable range, the induced strain will also alternate sinusoidal, however, presenting
a lag (i.e. out of phase). This viscoelastic performance of the vulcanizate is sus-
ceptible to the various motions activated within the temperature or frequency range
tested. Appropriate technique to monitor such phenomena has proved to be dynamic
mechanical analysis (DMA). By this method parameters like the ability of the vul-
canizate to store energy (i.e. storage modulus), to lose energy (i.e. loss modulus) and
the ratio of these effects (i.e. loss factor) are plausible determined [22]. The latter
dimensionless parameter is a measure of dynamic hysteresis of the vulcanizate and it
is generally characterized as damping. DMA testing might be associated with dif-
ferent types of fixtures suitable for tensile, flexure or compression modes. The tensile
mode is usually in action. It should be mentioned at this point that the dynamic
properties of rubber vulcanizates investigated by a sinusoidal strain dynamic tester,
which involve low to moderate deformations able to destroy the created filler net-
work (i.e. Payne effect), are not considered in the current analysis [23, 24].
Examining the thermal transitions in elastomers, changes in free volume or
relaxation times supply information for the various segmental motions. Making
use of the crankshaft model, as the free volume increases, the ability of the chain
segments to move in various directions also increase. Usually at very low
temperatures, the localized bond movement of bending and stretching is ascribed
as gamma (c) transition, which may also involve associations with water. As the
temperature and the free volume increases, whole side chains and localized
groups of four to eight backbone atoms are in action. Such transitions are called
beta (b) transitions. The large scale motions of the amorphous regions as heating
increases further are related to the glass transition [22]. Above that threshold
the rubbery state appears. The modulus in that region is proportional to the number
of cross-links. In elastomers, glass transition temperature often defines the lower
limit of the operating temperature. For DMA testing, a frequency value close to
the real world application should be chosen. Due to the fact that this appears in
most of the cases to be inconvenient for the instrumentation, a set of the same
parameters for all samples in the data set is considered. Generally speaking, a shift
of the temperature of a transition to lower values appears by lowering the
frequency. Moreover, as the testing frequency increases, the modulus usually
exhibits also an increase.
94 G. C. Psarras and K. G. Gatos

2.2 Dielectric Spectroscopy

Electrical relaxation phenomena occurring in polymer and polymer composites


can be studied experimentally by means of broadband dielectric spectroscopy
(BDS), which is also called dynamic electrical analysis, underlying that the
method is the electrical equivalent of DMA. Both methods exert a time dependent
field (electrical or mechanical) and record the material’s response, with varying
parameters the frequency of the applied stimulus, the sample’s temperature (in
some cases the applied pressure) and the amplitude of the exerted field. In DMA
the responding entities of the material are masses, while in BDS the responding
ones are permanent or induced dipoles. As it can be easily understood BDS can
record faster relaxation effects (characterized by sorter relaxation times) than
DMA.
Broadband Dielectric Spectroscopy is a powerful technique for the investiga-
tion of physical effects occurring in polymers and polymer micro- and/or nano-
composites, such as molecular mobility, polarization, conductivity, interfacial
phenomena, phase changes, polymerization, crystallization, etc. [25–28]. In our
days modern and convenient devices are available in reasonable prices, facilitating
thus the usage of dielectric spectroscopy in basic and applied research. The
employed experimental set up (device and dielectric cell) varies according to the
examined frequency range. However, an analytical description of the required
experimental set ups is out of the scope of the present chapter, useful experimental
information can be retrieved elsewhere [25–28]. Dielectric data are recorded under
isothermal or isochronal conditions scanning the frequency of the applied field in
the first case and specimen’s temperature in the second one. Data can be analysed
via different formalisms that is in terms of: (a) dielectric permittivity, (b) electric
modulus, (c) ac conductivity, and (d) complex impedance. Electrical phenomena
present in materials can be described and studied via all four formalisms. How-
ever, under specific circumstances one of them could be proved more helpful in
extracting information concerning the observed physical effects. Typically,
dielectric data are expressed via the real and imaginary part of dielectric permit-
tivity and electric modulus according to Eqs. 1 and 2, respectively. Equation 3
defines tan d which is a measure of the dissipated energy upon the stored energy at
each charging cycle.
e ¼ e0  je00 ð1Þ

1 1 e0 e00
M ¼ ¼ ¼ þ j ¼ M 0 þ jM 00 ð2Þ
e e0  je00 e02 þ e002 e02 þ e002

e00 M 00
tan d ¼ ¼ 0; ð3Þ
e0 M
where e0 , M0 and e00 , M00 are the real and the imaginary part of dielectric permit-
tivity and electric modulus, respectively. The electric modulus presentation offers
Relaxation Phenomena in Elastomeric Nanocomposites 95

advantages in the interpretation of dielectric data, since the large values of (e0 ) and
(e00 ) which obscure relaxation effects in the low frequency range and at high
temperatures are minimized. Further, the influence of electrode polarization
becomes negligible [29–32].
Relaxation processes in polymer nanocomposites arise from both the polymer
matrix and the presence of the filler. Dynamics of relaxation processes can be
investigated by recording the frequency of dielectric loss peaks varying the tem-
perature of the sample. Relaxation dynamics express the influence of temperature
upon relaxation time or dielectric loss peak frequency and allow investigating the
type of the distribution of relaxation times (i.e. symmetrical, non symmetrical, and
superposition of symmetrical and non symmetrical) [29]. From these results
valuable information for the matrix/filler interactions might be obtained [33–37].

2.3 Others

Among other techniques, transient mechanical testing has been used in the liter-
ature to examine relaxation phenomena in elastomeric vulcanizates. These may
involve creep, set and stress relaxation, all methods investigating the result of an
applied stress or strain as a function of time. Parallel to that, these methods are
used as measures for ageing characteristics, low temperature resistance and
resistance to chemicals [38].
A technique which currently receives increased attention for dynamic molec-
ular interaction processes in elastomer vulcanizates is solid state nuclear magnetic
resonance (NMR) spectroscopy. Utilizing methods like magic angle spinning and
cross polarization, the line broadening has been reduced increasing at the same
time the NMR signal sensitivity [39]. Thus, solid state NMR might be considered
of comparable quality to the traditional liquid NMR. The central NMR observable
is the residual dipolar couplings characterizing local chain order. In this respect,
multiple quantum NMR has been utilized for the measuring of weak residual
dipolar couplings in elastomeric nanocomposites [40].

3 Relaxation Phenomena in Nanocomposites with


Non-Polar Elastomeric Matrices

3.1 Dynamic Mechanical Analysis Characterization

Natural rubber (NR) has traditionally received most of researchers’ attention


among the non-polar elastomers. Regarding the spherical-like nanofillers, which
have been mixed with NR, CB constitutes a special subcategory wherein filler
agglomeration is rather desirable for real-life applications. This emanates from the
96 G. C. Psarras and K. G. Gatos

fact that an elastomer reinforced with dispersed CB agglomerates behaves, in


terms of its viscoelastic response, as its filler concentration is higher than its real
volume [41, 42]. This discrepancy between theory and practice has been settled
considering the occluded rubber as part of the filler thus, introducing the term of
‘effective’ filler [21, 43]. Due to broad existence of huge agglomerates in high
filler loadings, CB filled elastomers usually decline from their classification as
nanocomposites.
Nevertheless, CB has been employed in several cases as standard filler for
identifying mechanisms involved in elastomeric (nano) composites. Reduction of
the CB particle size in butadiene rubber (BR) resulted in a respective decrease of
the damping performance as this was revealed in a tan d versus temperature graph.
The compound with the lowest damping presented also the highest bound rubber
value [44]. Styrene butadiene rubber (SBR) was mixed with 40 phr CB through
melt blending in various percentages of sulfur curatives. A number of calendered
stocks were prepared, wherein each possessed its particular percentage of cura-
tives. By hot-pressing the individual layers together a so-called functionally gra-
ded polymeric nanocomposite (FGPNC) was produced. The increase in crosslink
density of the cured layers caused a shift of the glass transition temperature from -
19 to +20 °C as detected during dynamic mechanical tests. Thus, damping proved
to be a sensitive indicator of crosslinking. Additionally, comparing a uniformly
dispersed polymeric nanocomposite (UDPNC) with the respective FGPNC, both
retaining the same equivalent of curatives, interesting results were obtained. More
specifically, due to the variation of crosslink density along the thickness, FGPNC
broaden the glass transition region as observed in a tan d versus temperature
graph. Moreover, its tan d peak value at the Tg was lower compared to that of
UDPNC, whereas glass transition temperature remained rather unaffected [45].
Considering the loss factor as a measure of material internal losses per cycle
deformation during periodical excitation, its monitoring might reveal notable
information for the filler performance under wear conditions. Tire applications
currently involve particulate fillers like CB and silica. In this respect, master
curves of tan d at the reference temperature of 20 °C and at strain amplitude of
0.5% were produced for SBR vulcanizates filled with silica and CB in 80 and
81 phr filler loading, respectively. At the low frequency region silica produced
lower tan d values compared to the CB stock, indicating better rolling perfor-
mance. On the other hand, as frequency increased (1–1,000 kHz) silica yielded
higher tan d values compared to CB, thus higher energy dissipation. The latter
frequency range is related to the excitation frequencies involved during the anti-
lock braking system braking phase [46].
The last decade, rod-like fillers for rubber reinforcement have been the subject of
vivid research. Pre-vulcanized NR latex has been mixed with water solution
of multi-walled carbon nanotubes (MWCNT) and sodium dodecyl sulfate (SDS).
After stirring and sonication the mixture was cast and left to dry. The DMA
investigation of the vulcanizates showed an almost linear decrease of the mechanical
loss factor peak value at the glass transition temperature (Tg) versus the MWCNT
weight percentage. Efficient stress transfer to the filler was provoked by the
Relaxation Phenomena in Elastomeric Nanocomposites 97

treatment of the MWCNTs with nitric acid, which created carboxyl functionalities
on the side walls of the purified nanotubes. The significantly increased storage
modulus in the rubbery state of the elastomeric nanocomposites with higher filler
weight percentages was accompanied by higher tan d values compared to the neat
matrix [47]. Additionally, at higher MWCNT percentages (i.e. 5.4 and 8.3 wt%)
the tan d peak at Tg was shifted to lower temperatures, likely due to the increased
amount of SDS along with the filler aggregation observed at these loadings.
Generally speaking, the enhanced mechanical performance of elastomers
reinforced with carbon nanotubes (CNT) is not always accompanied by a peak-
shift to higher temperatures of the Tg in a loss tangent (tan d) versus temperature
graph. Considering that the existence of such peak-shift hints build-up of rubber–
filler interactions, its absence raises various argumentations [48]. SBR containing
23.5 wt% bound styrene was dissolved in toluene and stirred with 0.66 wt%
MWCNT along with sulfur type curatives. The purified nanotubes were sonicated
in ethanol and dried prior their solution mixing with SBR. The ability of the
vulcanizate to absorb energy was enhanced as shown by the increase of loss
tangent peak in the DMA spectrum. Additionally, the glass transition temperature
was shifted slightly to higher temperatures (i.e. 2.5 °C) compared to the neat
matrix. The prominent damping characteristics of the SBR/MWCNT nanocom-
posites were accompanied by improved storage modulus in the glassy and rubbery
state [49].
The creation of functional groups on the surface of CNTs is often requisite for
mixing with non-polar elastomers in order to improve the load transfer to the filler
[50]. In this respect, CNTs were first pretreated with nitric and sulfuric acid
followed by treatment with hydrated silica, resorcinol and hexamethylene tetra-
mine. Ball-milling technique was exploited for the CNT prior their solution mixing
with NR. It was found that the CNT preparation process created carbonylic and
hydroxylic functional groups on the filler surface. Addition of 25 phr CNT in the
NR matrix restricted mobility of the NR backbone lowering the damping perfor-
mance of the vulcanizate along with a shift of the Tg to higher temperatures [51].
Going a step further, introducing MWCNTs in an ionic liquid ethanol solution
followed by its melt mixing with an SBR/BR blend (in 1/1 ratio) significant
interfacial interactions were created. The ionic liquid 1-allyl-3methyl imidazolium
chloride (AMIC) was thought on the one hand to participate in rubber vulcani-
zation via its reactive double bond and on the other hand to be tethered on the CNT
surface by cation-p interactions. This strong rubber–filler coupling was responsible
for the enhanced storage modulus in the rubbery region and the decreased damping
at the glass transition temperature by adding 3 phr of MWCNTs. The interaction
of the AMIC with the rubber and the CNT formed a thin polymeric layer around
each CNT of significantly reduced mobility. As revealed in the mechanical loss
factor versus temperature spectrum, apart from the main a-relaxation peak of the
matrix another smaller peak at about 80 °C was detected [52]. Supportive data to
the idea of a tight shell around the CNTs were given by the clear peak-shift in the
Raman spectrum towards higher wavenumbers of the G-band, which is associated
with the tangential C–C stretching of the graphite-like structures.
98 G. C. Psarras and K. G. Gatos

The interactions of the macromolecular chains with the respective substrate


have been found responsible for the promotion of relaxation phenomena and the
shift of the glass transition temperature [53, 54]. Thus, platelet-like nanofillers is
expected to vigorously define various relaxation behaviors for elastomer nano-
composites according to their affinity with the matrix. Waxy maize starch nano-
crystals with a thickness of 6–8 nm, a length of 40–60 nm and a width of 15–
30 nm has been mixed with NR latex. This type of nanofiller enhanced signifi-
cantly the storage modulus of the respective nanocomposites especially in the
rubbery region as recorded by means of DMA. Addition of 30 wt% of starch
nanocrystals resulted strongly in a reduction of the damping performance of the
vulcanizates, shifting slightly the Tg to lower temperatures [55].
Lately, a lot of research has been conducted on rubber/layered silicate nano-
composites. For this 2D filler type (referred also as clay if it is of natural origin), an
elastomeric nanocomposite is obtained favorably when the individual silicate sheets
are surrounded by the macromolecular chains of the rubber. This scenario includes
rubber confined within the silicate galleries (cf. intercalated nanocomposite) and/or
individual silicate sheets dispersed in the matrix (cf. exfoliated nanocomposite) [56].
Significant role on such preferred silicate dispersion in the rubber matrix plays
the organic modification of the sheets with the appropriate intercalant [57].
The organically modified clay is generally called ‘organoclay’. The relaxation
phenomena involved in intercalated structures emanate from the favorable rubber-
organoclay interactions and the shielded rubber by the silicate layers. Thus, less
amount of elastomer is able to be deformed under dynamic conditions reducing the
amount of dissipated energy [58]. In this respect, NR melt-mixed with organically
modified montmorillonite (OMMT) presented reduced damping performance and a
shift of the Tg to higher temperatures in a loss tangent versus temperature graph [59,
60]. In certain cases, the confined macromolecular chains within the silicate layers
were charged with the appearance of a second relaxation peak (or of shoulder-like
form) positioned at higher temperatures in a tan d versus temperature spectrum. Such
behavior has been reported for NR/fluorohectorite nanocomposites produced via the
latex route [61] and NR melt-mixed with 10 phr montmorillonite (MMT) modified
with octadecylamine in tensile mode at a frequency of 10 Hz [62]. The presence of
the intercalants on the surface of the clay layers should additionally affect the
dynamics of the rubber chains in their vicinity. In order to amplify this behavior, SBR
and ethylene-propylene-diene rubber, were filled with OMMT up to 150 phr. In both
highly filled rubber/OMMT nanocomposites, a transition around 60–70 °C was
detected by means of DMA. The relaxation peak raised at that temperature region in a
tan d versus temperature graph was connected with the thermally induced transition
of the surfactant (i.e. alkyl chains) from solid-like to liquid-like state [63].
Enhancement of the interactions of non-polar rubbers with organoclay layers
through coupling agents was found to be a convenient tool for increased mechanical
performance of said nanocomposites [64, 65]. SBR latex was mixed with a 2 wt%
aqueous suspension of MMT and 3-aminopropyltrimethoxysilane followed by co-
coagulation in 2% dilute sulfuric acid solution. The dried compound was melt-
blended with the sulfuric curatives and 2 phr bis(triethoxysilyl-propyl)tetrasulfide
Relaxation Phenomena in Elastomeric Nanocomposites 99

(i.e. Si69) prior vulcanization. The coupling of Si69 through its ethoxysilyl-propyl
and sulfide groups with SBR and the Si–OH of the nanoclay decreased the mobility
of the rubber chains near the platelets. Thus, the Tg of SBR/clay compounds was
shifted from -49.8 to -34.8 °C when the clay layers were micro- and nano-scale
dispersed, respectively [66].
In exfoliated structures of rubber/organoclay nanocomposites, the single
platelets dispersed within the matrix are accompanied by their tethered interca-
lants. While the first restrain the rubber chain mobility through the network created
[67], the latter act as plasticizer in the nanocomposite [58]. Thus, a moderate peak-
shift of the glass transition temperature due to the above-mentioned interplay is
expected to be revealed in a loss factor versus temperature spectrum [68].
Nowadays, demanding applications of elastomers might require a combined
action of fillers [69, 70]. In such cases, the relaxation phenomena involved in rel-
evant vulcanizates are related to each filler component, as well as, to the reinforcing
filler-blend efficiency. In this respect, SBR was melt-mixed with CB and organo-
clay. The DMA spectra of the vulcanizates reinforced with CB and OMMT both in
20 phr loading resulted in significantly increased storage modulus and decreased
damping compared to the vulcanizates filled only with CB or OMMT [71].

3.2 Dielectric Probing

Dielectric relaxations are related to dipolar effects, thus non-polar rubbers appear to
be inactive dielectrically. Despite the technological importance of NR, which was
stated in the previous paragraph, scarce studies on its dielectric response have been
reported in the literature [32, 36, 59]. NR is weakly active dielectrically because of
its non-polar nature. The absence of polar side groups in the polymer chain of NR is
responsible for lacking of any secondary relaxation process in its dielectric spectra.
Thus, the only present relaxation process in its spectra refers to the glass to rubber
transition of NR (a-mode) [36]. This is located around -64 °C as it has been
detected by means of differential scanning calorimetry [36, 72]. The corresponding
loss peak recorded at low temperatures appears to be weak and is detected also in
the NR based nanocomposites [36]. Figure 1 depicts the dielectric response of NR
via the variation of the imaginary part of electric modulus (M00 ) as a function of both
temperature and frequency. The dielectric response of NR has been examined in the
temperature range from -80 to 50 °C and in the frequency range of the field from
0.1 Hz to 10 MHz. The presence of a single relaxation process is evident in Fig. 1.
Electric modulus loss maximum (attributed to glass to rubber transition) shifts, with
increasing frequency, from the vicinity of -60 to almost 0 °C. Figure 2 presents the
dielectric response of a NR nanocomposite. NR has been reinforced by 10 phr
(parts per hundred of NR) of layered silicates (LS). The employed LS exhibit very
high aspect ratio (viz.[1,000). Intergallery distance of LS, as determined by X-ray
diffraction (XRD) scattering, was found to be 0.95 nm. In the case of the nano-
composite system the basal spacing expands to 1.19 nm, indicating intercalation of
100 G. C. Psarras and K. G. Gatos

10-1
10-2
10-3
Modulus''
10-4
10-5
10-6
α-mode
8
6 10
50 10
4
0 10
-50 10
2
y[ H z]
[°C]
10
0 enc
equ
rature -100 -2
Tem pe 10 Fr

Fig. 1 Imaginary part of electric modulus versus frequency and temperature for pure NR

IP

10-1
10-2
10-3
Modulus''
10-4
10-5
10-6

α-mode 10
8
6
50 10
4
0 10
2
H z]
] -50
0
10
cy [
ture [°C ra -100 -2 10 q uen
Tempe 10 Fre

Fig. 2 Imaginary part of electric modulus versus frequency and temperature for NR/LS nano-
composite in 10 phr filler loading

LS by latex components [36, 72, 73]. As it can be seen in Fig. 2 the NR/LS
nanocomposite spectra resemble to the ones of the pure matrix, including the
primary relaxation process of a-mode. However, it should be noted that in a
composite system it is expected to detect interfacial polarization (IP). IP also known
as Maxwell–Wagner–Sillars (MWS) effect occurs in heterogeneous and complex
Relaxation Phenomena in Elastomeric Nanocomposites 101

systems, because of the accumulation of mobile charges at the interface of the


constituents. At the interfaces large dipoles are formed. The relaxation of these
dipoles under the influence of the applied field give rise to a slow relaxation process
(characterized by enhanced relaxation time), which in most of the cases is recorded
at relatively high temperatures and low frequencies [29, 74–77]. In the case of the
NR/LS nanocomposite, this peak can be found in the low frequency region of its
dielectric spectra in Fig. 2. At temperatures higher than 20 °C the values of (M00 ) for
both systems increase rapidly likely due to the softening of the matrix.
It is well known that the frequency-temperature superposition shifts dielectric
loss peaks to higher frequencies with raising temperature. Relaxation mechanisms
differ in their peak shift rate, reflecting the type of the involved dipoles, the influence
of the molecular environment and the effect (if any) from the filler. The relaxation
dynamics can be expressed in terms of the temperature dependence of dielectric loss
peak frequency position, or in terms of the temperature dependence of the relaxation
time. In the second case, the distribution of relaxation times can also be investigated
via suitable distribution functions, which are available in the literature [26, 29]. Fast
relaxation processes, such as local motions of polar side groups and small segments
of the polymer chain (b- and c-mode respectively) typically follow Arrhenius type
temperature dependence. IP relaxation exhibits the same type temperature depen-
dence, although it is recorded at the opposite edge of the time scale compared to local
motions. Arrhenius behaviour is expressed by Eq. 4:
 
EA
fpeak ¼ f0 exp  ; ð4Þ
kB T

where fpeak is the frequency of the dielectric loss peak, f0 pre-exponential factor,
EA the activation energy of the process, kB the Boltzmann constant and T tem-
perature (in Kelvin). The relaxation process related to glass to rubber transition, is
not characterized by a constant loss peak shift rate, and thus its temperature
dependence cannot be described via Eq. 4. The temperature dependence of a-mode
follows the Vogel–Fulcher–Tamann (VFT) equation, which considers that relax-
ation rate increases rapidly at lower temperatures because of the reduction of the
free volume. VFT relation is expressed by Eq. 5:
 
AT0
fpeak ¼ f0 exp  ; ð5Þ
T  T0

where fpeak is the frequency of the dielectric loss peak, f0 pre-exponential factor, A
a constant being the measure of activation energy, T temperature (in Kelvin) and
T0 the Vogel temperature or ideal glass transition temperature. Figure 3 presents
the temperature dependence of a-mode for pure NR and NR reinforced with 10 phr
LS, while the values of parameters A and T0 evaluated via fitting data with Eq. 5
are listed in Table 1. The increase of parameter A in the case of the nanocomposite
indicates an enhancement of the required activation energy for the onset of the
process, suggesting indirectly that the presence of nanofiller exerts some restric-
tions on macromolecules rearrangement.
102 G. C. Psarras and K. G. Gatos

Fig. 3 Loss peak position as


a function of the reciprocal
temperature of a-relaxation
process, for the NR and NR
reinforced with 10 phr LS
systems. Lines are produced
by fitting data via Eq. 5

Table 1 Fitting parameters of Eq. 5 for the glass to rubber relaxation process of NR and NR
reinforced with 10 phr LS
Sample A T0 (K)
NR 25.7 128
NR ? 10 phr LS 35.0 120

Recently, the effect of type and amount of LS loading upon the dielectric response
of NR nanocomposite was studied in detail [32]. Pure clay and an organically
modified type were used for the production of NR/LS nanocomposites (vulcanized
and non-vulcanized versions) [32]. IP or MWS effect is present in all nanocom-
posites. Glass to rubber relaxation process, detected in all systems, slows down its
dynamics when the NR matrix is vulcanized. Further, in vulcanized systems the
temperature dependence of the dielectric loss peak position appears to be unaffected
by the type of the employed LS. Relaxation spectra of non-vulcanized systems
include normal-mode relaxation, in which the end-to-end polarization vector
characterizes the translational motion of the whole chain. This process vanishes after
vulcanization, because of the formation of cross-links [32] and becomes not
detectable. An additional slow relaxation mode was observed in all NR/organoclay
systems. Its origin was attributed to partially immobilized polymer chains or
restricted segmental mobility at the interface layer around clay particles [32].

4 Relaxation Phenomena in Nanocomposites with


Polar Elastomeric Matrices

4.1 Dynamic Mechanical Analysis Characterization

In several cases elastomeric nanocomposite manufacturing requires increased


rubber polarity for optimized filler dispersion in the respective matrix. However,
Relaxation Phenomena in Elastomeric Nanocomposites 103

tailoring the interface quality and performance in polar elastomeric nanocom-


posites is not an easy task. While precipitated silica dispersion in nitrile rubber is
favored by increasing acrylonitrile content, a negligent effect might appear for CB
reinforcement [44]. Nano-sized silica has been melt-mixed with carboxylated
nitrile rubber (XNBR) at a filler loading up to 50 phr. DMA investigations in the
torsion mode at a frequency of 1 Hz revealed a slight reduction of the damping
performance of the nanocomposites along with a marginal peak-shift of the Tg to
lower temperatures [78]. This seems to be related to the silica agglomeration and
de-wetting phenomena at the respective interface. In order to overcome floccu-
lation of silica within the rubber matrix, the first is synthesized in-situ during
compounding. Silica particles were synthesized from N-(2amino-ethyl)-3-amino-
propyltrimethoxysilane in XNBR. The rubber type used was containing 6.4% of
carboxyl groups and 26.3% of acrylonitrile. As the amount of aminosilane varied
in the rubber recipe from 5 to 20 phr, in order to produce the set of reinforced
vulcanizates, crosslink density was accordingly changed. The ionic nature of the
crosslinks between the carboxyl groups of the XNBR and the amino groups of the
silane restrained the mobility of the macromolecular chains at the interface. This
was reflected in the tan d versus temperature graph wherein apart from the a-
relaxation at about 0 °C, which was almost constant for all vulcanizates, another
one at higher temperature (&40 °C) appeared, characterized as a0 -relaxation. As
the aminosilane amount in the vulcanizates increased, the loss tangent value at the
glass transition temperature decreased along with an intensity increase of the a0 -
relaxation [79]. The emergence of slower segmental dynamics than in the bulk
rubber have been also referred for hybrid polyurethane (PUR) reinforced with
polyhedral oligomeric silsesquioxanes [80]. For the same polar rubber, when it
was mixed with 10 phr alumina nanoparticles of various sizes via the latex route
and left to dry, decrease of the particle size of the filler resulted in a reduction of
the intensity of the loss factor peak at the glass transition temperature. At the same
time, a marginal tan d peak-shift to higher temperatures was detected for the
compounds with small particle diameter [37]. It seems that by enhancing rubber–
filler interactions less macromolecular chains contribute to a-relaxation due to
their immobilization in the vicinity of the filler. As shown in Fig. 4, the alumina
nanoparticles (being 25 nm in diameter) resulted in a significant enhancement of
the storage modulus at the rubbery state along with a shift of the Tg to higher
temperatures compared to the neat matrix. In addition, the damping of the nano-
composite was strongly reduced.
Rubber–filler interface enhancement for polar elastomers has also been inves-
tigated in case of rod-like nanofillers [81]. For MWCNT compounded with
hydrogenated carboxylated acrylonitrile butadiene rubber by melt blending, the
presence of polar groups had small effect on the glass transition temperature [82].
It seems that significantly slower segmental dynamics require apart from the
adequate rubber polarity a properly modified filler surface, and vice versa. Nev-
ertheless, enhancement of the mechanical performance is generally reported.
For platelet-like fillers mixed with polar rubbers, extended research has been
conducted during the last decade. Epoxidized natural rubber (ENR) has been used
104 G. C. Psarras and K. G. Gatos

Fig. 4 Storage modulus and


mechanical loss factor of
PUR and PUR/alumina
nanocomposite (in 10 phr
filler loading) versus
temperature (in tension
mode at 10 Hz frequency).
Note that the line represents
the neat PUR whereas line
plus symbol the PUR
reinforced with nanoparticles
of 25 nm in diameter

as compatibilizer for NR mixed with MMT modified by octadecyltrimethylamine.


Increasing the degree of epoxidization from 25 to 50 mol% (designated as ENR 25
and ENR 50, respectively) resulted in a pronounced immobilization of rubber
chains in the vicinity of the platelets as this was corroborated by attenuated
damping behavior [83]. Decreased damping and shift of the Tg to higher tem-
peratures has been also reported for the same system mixed with 50 phr CB or
30 phr silica [84]. However, melt blending of only 2 phr organoclay with NR
compatibilized by 10 phr of ENR 50 resulted in pronounced intercalated structures
and improved strength. Moreover, in a loss factor versus temperature graph, apart
from the a-relaxation at about -45 °C another peak at about -5 °C was raised.
The latter was connected with the intercalated rubber populations due to the
increased polarity of ENR 50 [83]. Similar segmental dynamics were marked also
when the organoclay loading increased up to 10 phr [85]. The ability of ENR 50 to
penetrate within silicate layers has been elucidated when it was melt mixed
directly with organoclay [86]. As it is shown in Fig. 5, addition of 10 phr
organoclay (MMT modified by octadecylamine) enhanced the storage modulus of
the vulcanizate compared to a silica filled compound of same loading. Addition-
ally, the a-relaxation at about -5 °C was reduced in intensity followed by a
‘shoulder’ positioned at about 10 °C, due to the segmental motions of a part of
ENR 50 matrix activated at a higher temperature.
Dynamic mechanical analysis do not always detect the discernible relaxation
phenomena due to the intercalated rubber populations in polar rubbers reinforced
with platelet-like nanofillers [87]. Nevertheless, a tan d maximum value decrease
is usually observed in the relevant DMA spectra [88–90]. As the interaction of the
clay intercalant with the rubber matrix increases the damping reduction might
become more pronounced. In such an example, hydrogenated nitrile rubber
(HNBR), which was reinforced with MMT modified with various intercalants
resulted in a decreased intensity of the tan d peak at the glass transition temper-
ature for all nanocompounds. However, when a methyltallow-bis(2-hydroxyethyl)
quaternary ammonium salt was in action, the induced hydrogen bonds restrained
Relaxation Phenomena in Elastomeric Nanocomposites 105

Fig. 5 Storage modulus and


mechanical loss factor of
ENR 50 vulcanizates mixed
with silica (line curve) or
organoclay (line plus symbol
curve) in 10 phr loading
versus temperature (in
tension mode at 10 Hz
frequency)

Fig. 6 Storage modulus and


mechanical loss factor of neat
PUR/NR blend (line curve)
and reinforced with 10 phr
sodium fluorohectorite (line
plus symbol curve) versus
temperature (in tension mode
at 10 Hz frequency)

further the mobility of the nitrile rubber, thus, an even lower tan d peak intensity
was measured [91].
Rubber blends are usually in action in order to overcome cost and property
related issues. In such an example, PUR rubber was mixed with a quite non-
expensive pre-vulcanized NR via the latex route followed by the addition of
sodium fluorohectorite (FHT) slurry in 10 phr filler loading [73]. While the first
rubber bears polar groups able to create enhanced interface with FHT, the latter
apolar one interacts weak with the FHT layers. The characteristics of the dried
compounds were resolved by means of DMA. As shown in Fig. 6, two distinct
a-relaxation peaks rise at about -55 and 5 °C for NR and PUR rubber, respec-
tively, revealing the immiscibility among these two rubbers. Addition of FHT in
the compound shifted the glass transition temperature of PUR at about 10 °C,
while the Tg for NR remained rather unaffected. On the other hand, the increased
tan d values at the rubbery region for the nanocomposite suggest a more efficient
energy dissipation mechanism than that of the plain blend. This is likely due to the
‘house of cards’ structure built by the dispersed FHT platelets, which are actually
106 G. C. Psarras and K. G. Gatos

expelled by the pre-vulcanized NR latex particles [73]. Nevertheless, the nanod-


ispersion of FHT platelets in the blend enhanced significantly the storage modulus
of the compound in the whole temperature range.

4.2 Dielectric Probing

Polar rubbers are, as expected, dielectrically active materials. Their response


include contributions arising from glass to rubber transition (a-mode), local
motions of polar side groups (b-mode) and from the rearrangement of small
segments of the polymer main chain (c-mode). In cases where additives and/or
plasticizer are present, as well as in semi-crystalline rubbers IP could be detected.
The dielectric response of nanocomposites based on polar matrices compromises
dipolar events arising from the matrix as well as from the presence of the filler.
Further, the presence of the nanofiller might affect relaxation phenomena ema-
nating from the polymer matrix. Synthetic PUR is a typical polar rubber used for
the development of many nanocomposites. In most of the studies the employed
nanofillers are alumina particles, LS and CNT. Slow and fast relaxation processes
can be observed in the dielectric spectra of PUR. The latter compromise local
motions of small parts of the main chain, like the ‘crankshaft’ type motion of the
[CH2]x sequence (c-mode), and re-orientation of polar side groups of the main
chain (b-mode), while the former are a-mode (glass to rubber transition) and IP
[36, 72]. The PUR macromolecular backbone consists of hard and soft segments
[36, 75, 76]. At the interface between soft and hard regions unbounded charges,
arising from the stage of samples preparation, are concentrated forming large
induced dipoles, which participate to the polarization effect of the system at the
low frequency and high temperature regions. PUR/NR blends are also used as base
for the development of rubber nanocomposites. Figures 7 and 8 show the variation
of the imaginary part of electric modulus as a function of temperature and fre-
quency for PUR and PUR/NR systems. All four aforementioned relaxations can be
easily detected in the PUR spectra, while blend’s response contains also the
contribution from the a-mode of NR. Reasons for choosing the electric modulus
formalism have been, briefly, referred in Sect. 2.2.
Relaxations of a PUR mixed with 10 phr alumina nanoparticles are depicted in
the spectra of Fig. 9. Qualitatively the picture refers to the response of PUR. The
presence of nanofiller within a polymer matrix, could affect the matrix glass
transition temperature. For favorable polymer–filler interactions at the nanoscale
Tg shifts to higher value, compared to that of the pure bulk polymer [92, 93]. On
the other hand, repulsive interactions at the interface between matrix and filler, as
well as the existence of pore or voids because of poor wetting level of the
inclusions, or insufficient adhesion between polymer and nanofiller result in
enhancement of the cooperative mobility of the macromolecules, and thus in
decreasing glass transition temperature [94, 95]. Another reason, which could
influence Tg in rubber/LS nanocomposites is their structural configuration [33, 36].
Relaxation Phenomena in Elastomeric Nanocomposites 107

IP

10-1
10-2
Modulus''
10-3
γ-mode

10-4
α-mode
8
0
6 1
β-mode 10
4
10 ]
Hz
50
2
10 y[
0
c
en
-50 0
0
-2 1 equ
-100
re [°C] -150
Temperatu 10 F r

Fig. 7 Imaginary part of electric modulus versus frequency and temperature for pure PUR

100
IP
10-1
Modulus''
10-2
10-3
10-4

β,γ-modes
α-mode
of PUR
of PUR 0
8
α-mode 6 1
10
4
of NR 10 ]
Hz
100
2
y[
50
0 10 c
0 en
-50
0
-2 1 equ
-100
re [°C] -150
Temperatu 10 F r

Fig. 8 Imaginary part of electric modulus versus frequency and temperature for PUR/NR blend

In the case of intercalated structures rubber molecules lying within the LS gal-
leries, are able to move faster than bulk polymer chains. Confining of polymer
macromolecules in a narrow space at the nanoscale level, results in isolated
macromolecules, which can move or relax faster or easier, and by thus shift glass
transition temperature to lower values [33]. Glass transition temperature variation
108 G. C. Psarras and K. G. Gatos

10-1
IP

10-2
Modulus''
10-3
10-4
γ-mode

10-5
α-mode
8
0
β-mode 6 1
10
4
10 ]
Hz
100
2
y[
50
0 10 c
0 en
-50
0
-2 1 equ
-100
re [°C] -150
Temperatu 10 F r

Fig. 9 Imaginary part of electric modulus versus frequency and temperature for PUR reinforced
with 10 phr alumina particles. The mean diameter of alumina particles is 25 nm

can be detected and studied by means of BDS. In the vicinity of Tg macromole-


cules acquire sufficient energy to increase their mobility and to relax under the
influence of the applied alternating electric field. In nanocomposites shifting the
loss peak of a-mode, under isothermal conditions, to higher frequency compared to
that of pure bulk polymer, signifies lowering of Tg. This is quite reasonable, since
shifting the loss maximum to higher frequency denotes decreasing of relaxation
time, due to the facilitation of the process. On the contrary, when loss peak
position of a-mode in a nanocomposite is shifted to lower frequency, with respect
to the loss position of bulk polymer, Tg value enhances. It should be noted that
BDS can be used not only for qualitative studying the modification of glass
transition temperature in nanocomposites, but also in studying the molecular
relaxation dynamics, and in identifying the type of the occurring interactions at the
system’s interface.
Secondary relaxations (i.e. b- and c-mode) are weak, sometimes suffer by
experimental noise, and characterized by short relaxation time, and thus recorded
at low temperatures and high frequencies. In some cases they mutually superim-
posed and/or are not easily detectable. Usually their loss peak position remains
unaffected by the presence of nanofiller. The dielectric permittivity loss maximum
could increase in rubber nanocomposites indicating higher consumption of energy
at each charging cycle. Enhanced energy consumption could be attributed to
restrictions, exerted by nanoinclusions, to local motions and reorientations [36,
72]. Figure 10 depicts the variation of real (e0 ) and imaginary part (e00 ) of dielectric
permittivity with frequency, in PUR and PUR/alumina nanocomposite systems at
-100 °C. Nanocomposites exhibit increased values of both parts of dielectric
Relaxation Phenomena in Elastomeric Nanocomposites 109

10 PUR
0.20
PUR+10phr alumina (220nm)
PUR+10phr alumina ( 90nm)
9 PUR+10phr alumina ( 25nm)
PUR+40phr alumina (220nm)
0.16
8

7
0.12

e ′′
e′

6
0.08
5

4
0.04
3

2 0.00
-1 0 1 2 3 4 5 6
logf

Fig. 10 Real and imaginary part of dielectric permittivity as a function of frequency, at


-100 °C, for PUR and PUR/alumina systems, varying the mean particle size and/or the alumina
content

permittivity, which in the case of the system with alumina particles of 25 nm mean
size attain maximum values. As responsible for this behaviour could be considered
the enhancement of interfacial area with diminishing the mean alumina’s particle
diameter. Additionally, the selection of the amount and the type of the employed
nanoinclusions offer the possibility to tailor the dielectric response of the
nanosystems.
Hydrogenated nitrile rubber, commercially introduced in the 1980s, addresses
advanced requirements of the automobile industry because of its excellent heat and
oil resistance properties combined with superior mechanical performance [96].
HNBR is a polar elastomer wherein its acrylonitrile content and its crystallinity
may vary accordingly [96, 97]. HNBR/CNT nanocomposites are reported to
improve, further, the mechanical properties of HNBR vulcanizates, as well as,
their wear performance [98]. The electrically conductive character of the CNTs
offers an additional advantage in applications, where moving or rotating members,
fabricated with HNBR, are present [99]. On the surface of moving or rotating
insulating members electrostatic charges can be easily developed leading to
undesirable sparks, which can cause an early failure of the component. A possible
solution to the problem is to incorporate within the insulating matrix conductive
inclusions, which will not lower the mechanical performance of the system and at
the same time they will be able to create conductive paths, inside the composite
systems, through which leakage current could flow. HNBR/CNT nanocomposite
appear to satisfy these requirements, since CNTs act as reinforcing phase to both
mechanical and electrical behaviour. Recently, the electrical response of HNBR
and HNBR/CNT nanocomposites was examined thoroughly [99]. In HNBR and
HNBR/CNT systems with low concentration of CNTs four distinct relaxation
processes were detected. These are IP at the interface of crystalline and amorphous
110 G. C. Psarras and K. G. Gatos

regions of HNBR (the existence of crystalline phase was confirmed via XRD
spectra) and/or at the interface between of HNBR and CNTs, glass to rubber
transition of the amorphous part of HNBR, rearrangement of polar side groups,
such as –CN, and local motions of small segments of the main elastomer chain
[99]. At higher CNTs concentrations, above a critical value also known as per-
colation threshold, the electrical response of the system varies significantly,
because it switches from insulating to conductive performance. Relaxation pro-
cesses vanish or notably suppressed, while the electrical behaviour of the system is
dominated by the increased values of electrical conductivity [99].
The dielectric response of PUR/LS and PUR/NR/LS nanocomposites (in 10 phr
filler loading) is presented in Figs. 11 and 12 respectively. Although, the observed
relaxations are those mentioned previously, some modifications are present in the
spectra of nanocomposites. The IP loss peak of the PUR/LS system appears to be
broader and its maximum is suppressed to lower values compared to the corre-
sponding of PUR (Fig. 7), PUR/NR (Fig. 8) and PUR/NR/LS (Fig. 12). The
occurring modification has been connected to variations of the morphology of the
examined systems [36, 72]. Previous morphological studies, via transmission
electron microscopy and XRD [72, 73] on the PUR reinforced with 10 phr LS
system, revealed the existence of two intercalated populations with different
interlayer distances, namely 1.23 and 1.73 nm. Note that the basal spacing of LS
powder was 0.95 nm. Further, the presence of isolated silicate layers due to partial
exfoliation cannot be excluded. Under this point of view, interfaces with varying
geometrical characteristics co-exist within the nanosystem, contributing to inter-
facial relaxation phenomena with different dynamics or relaxation times. The
10-1

IP
10-2
Modulus''
10-3

α-mode
10-4

β-mode γ-mode 8
0
6 1
10
4
10 ]
Hz
50
2
10 y[
0
c
en
-50 0
0
-2 1 equ
C] -100
re [° -150
Temperatu 10 Fr

Fig. 11 Imaginary part of electric modulus versus frequency and temperature for PUR/LS
nanocomposite (in 10 phr filler loading)
Relaxation Phenomena in Elastomeric Nanocomposites 111

10-1
IP

10-2
Modulus''
10-3
α-mode
of PUR

10-4
α-mode β, γ-modes
of PUR 8
of NR 0
6 1
10
4
10 ]
Hz
50
2
10 y[
0
c
en
-50 0
0
-2 1 equ
-100
re [°C] -150
Temperatu 10 F r

Fig. 12 Imaginary part of electric modulus versus frequency and temperature for PUR/NR/LS
nanocomposite (in 10 phr filler loading)

recorded broad peak can be considered as the superposition of all interfacial


effects. Interfacial relaxation phenomena prominent in the low frequency range
and relatively high temperatures are characterized by high values of both real and
imaginary part of dielectric permittivity [31, 100, 101]. Further increase of the
intensity of interfacial effects results in even higher values of (e0 ) and (e00 ).
However, in the electric modulus formalism the increase of intensity of the IP
process is demonstrated by reduced values of (M0 ) and (M00 ), because of its defi-
nition via Eq. 2. Thus, the recorded lower values of the modulus loss index (M00 )
constitute a strong indication for the existence of pronounced interfacial phe-
nomena in the PUR/LS system. In all other systems, the formed IP loss peaks
resemble to each other. The PUR contribution seems to dominate the spectra of all
systems. Morphological inspection saw that the PUR intercalation of LS is con-
siderably better than the one of NR [73]. This effect can be assigned to the higher
polarity of PUR compared to NR, which favors the compatibility with LS [73].
Since, NR and PUR are not compatible (amplified by the prevulcanization of NR),
and the intercalation of LS by NR is poor, silicate stacks are preferentially located
in the PUR regions [73]. The amplitude of the IP loss peak in the case of the PUR/
NR/LS nanocomposite attains lower value compared to that of PUR/NR and PUR
systems, implying the relative enhancement of IP due to the increase of the sys-
tem’s heterogeneity.
The temperature dependence of the dielectric loss peak position for a-mode and
IP of the systems: (i) pure PUR, (ii) PUR/LS, (iii) PUR/NR, and (iv) PUR/NR/LS
is shown in Fig. 13. Note that filler loading was set in 10 phr. As it can be seen IP
follows an Arrhenius type behaviour, described by Eq. 4, while a-mode’s peak
112 G. C. Psarras and K. G. Gatos

Fig. 13 Loss peak position


as a function of the reciprocal
temperature for the IP and
a-relaxation processes, for
PUR, PUR/LS, PUR/NR, and
PUR/NR/LS systems (LS is
incorporated in 10 phr in the
respective compounds). Lines
are produced by fitting data
via Eqs. 4 and 5

Table 2 Activation energy for the IP process and fitting parameters of Eq. 5 for the glass to
rubber relaxation process of the PUR based systems
Sample IP-process a-Relaxation process
EA (eV) A T0 (K)
PUR 1.29 8.3 215
PUR/NR 1.37 5.8 225
PUR ? 10 phr LS 0.60 33.0 165
PUR/NR ? 10 phr LS 1.29 2.4 246

shift is in agreement with VFT relation, as expressed by Eq. 5. Activation energy


values and values of parameters A and T0 calculated via Eqs. 4 and 5 are listed in
Table 2. The increase of parameter A in the case of the PUR/LS nanocomposite
indicates an enhancement of the required activation energy for the onset of the
process, suggesting indirectly that the presence of nanofiller exerts some restric-
tions on macromolecules rearrangement. Moreover the same system exhibits the
lowest value of activation energy for IP denoting that its morphological charac-
teristics, mentioned previously, facilitate the process.

5 Summary

Dynamic mechanical analysis and BDS are two effective experimental techniques
for studying rubber nanocomposites. Their spectra not only reveal the mechanical
and dielectric behaviour of the under test systems, but also provide valuable
information regarding molecular dynamics, interfacial effects, filler/matrix inter-
actions, and phase changes increasing the impact of our knowledge upon struc-
ture–property relationships. DMA and BDS are basically complementary methods,
although they are able to describe relaxations of the same type (primary and
Relaxation Phenomena in Elastomeric Nanocomposites 113

secondary modes). Relaxation spectra compromise contributions originating from


glass to rubber transition, local motions of side groups or small side branches,
rearrangement of small segments of the rubber main chain, IP at heterogeneous
systems and phase changes. DMA is a well established method for the determi-
nation of glass transition temperature, secondary relaxations of side groups, and
the investigation of mechanical energy storage capacitance and damping charac-
teristics of rubber nanocomposites. BDS besides glass to rubber transition is able
to follow faster molecular relaxation processes arising from reorientation of polar
side groups and small chain segments, IP effects due to the accumulation of
unbounded charges at the interface of the constituents and conductivity phenom-
ena. Further, qualitative information for the type of the occurring interactions
between nanofiller and elastomeric matrix can be revealed.

References

1. Haase, W., Wróbel, S.: Relaxation phenomena in physics and chemistry. In: Haase, W.,
Wróbel, S. (eds.) Relaxation Phenomena. Springer, Heidelberg (2003)
2. Havriliak Jr., S., Havriliak, S.J.: Dielectric and Mechanical Relaxation in Materials. Hanser,
Munich (1997)
3. Sommer, F.: Füllstoffe, Vernetzungsmittel, Additive. In: Röthemeyer, F., Sommer, F. (eds.)
Kautschuktechnologie. Hanser, Munich (2001)
4. Brennan, J.J., Jermyn, T.E., Boonstra, B.B.: Carbon black–polymer interaction: a measure
of reinforcement. J. Appl. Polym. Sci. 8, 2687–2706 (1964)
5. Su, F., Zhou, Z., Guo, W., Liu, J., Tian, X.N., Zhao, X.S.: Template approaches to preparing
porous carbon. In: Radovic, L.R. (ed.) Chemistry and Physics of Carbon. CRC Press, Boca
Raton (2008)
6. Conzatti, L., Costa, G., Falqui, L., Turturro, A.: Microscopic imaging of rubber compounds.
In: White, J., De, S.K., Naskar, K. (eds.) Rubber Technologist’s Handbook, vol. 2, 2nd edn.
Smithers Rapra Technology Limited, Shawbury (2009)
7. Sae-Oui, P., Sirisinha, C., Hatthapanit, K., Thepsuwan, U.: Comparison on reinforcing
efficiency between Si-69 and Si-264 in an efficient vulcanization system. Polym. Test. 24,
439–446 (2005)
8. Reuvekamp, L.A.E.M., ten Brinke, J.W., van Swaaij, P.J., Noordermeer, J.W.M.: Effects of
mixing conditions: reaction of TESPT silane coupling agent during mixing with silica filler
and tire rubber. Kautsch Gummi Kunstst 55, 41–47 (2002)
9. Choi, S.S.: Effect of bound rubber on characteristics of highly filled styrene-butadiene
rubber compounds with different types of carbon black. J. Appl. Polym. Sci. 93, 1001–1006
(2004)
10. Leblanc, J.L.: Elastomer–filler interactions and the rheology of filled rubber compounds. J.
Appl. Polym. Sci. 78, 1541–1550 (2000)
11. Wang, M.J.: Effect of polymer–filler and filler–filler interactions on dynamic properties of
filled vulcanizates. Rubber Chem. Technol. 71, 520–589 (1998)
12. Vilgis, T.A., Heinrich, G., Klüppel, M.: Reinforcement of Polymer Nano-composites:
Theory, Experiments and Applications. Cambridge University Press, New York (2009)
13. Wang, Q., Yang, F., Yang, Q., Chen, J., Guan, H.: Study on mechanical properties of nano-
Fe3O4 reinforced nitrile butadiene rubber. Mater. Des. 31, 1023–1028 (2010)
14. López-Manchado, M.A., Arroyo, M.: Short fibers as reinforcement of rubber compounds.
Polym. Compos. 23, 666–673 (2002)
114 G. C. Psarras and K. G. Gatos

15. Moraleda, J., Segurado, J., Llorca, J.: Effect of interface fracture on tensile deformation of
fiber-reinforced elastomers. Int. J. Solids Struct. 46, 4287–4297 (2009)
16. Frogley, M.D., Ravich, D., Wagner, H.D.: Mechanical properties of carbon nanoparticle-
reinforced elastomers. Comp. Sci. Technol. 63, 1647–1654 (2003)
17. Karger-Kocsis, J., Wu, C.M.: Thermoset rubber/layered silicate nanocomposites. Status and
future trends. Polym. Eng. Sci. 44, 1083–1093 (2004)
18. Wang, Z.L.: Nanostructures of zinc oxide. Mater. Today 7, 26–33 (2004)
19. Zhou, Z., Liu, S., Gu, L.: Studies on the strength and wear resistance of tetrapod-shaped
ZnO whisker-reinforced rubber composites. J. Appl. Polym. Sci. 80, 1520–1525 (2001)
20. Bukhina, M.F., Kurlyand, S.K.: Low Temperature Behaviour of Elastomers. VSP, Leiden
(2007)
21. Medalia, A.I.: Effect of carbon black on dynamic properties of rubber vulcanizates. Rubber
Chem. Technol. 51, 437–523 (1978)
22. Menard, K.P.: Dynamic Mechanical Analysis: A Practical Introduction. CRC Press, Boca
Raton (1999)
23. Payne, A.R.: The dynamic properties of carbon black loaded natural rubber vulcanizates.
Part II. J. Appl. Polym. Sci. 21, 368–372 (1962)
24. Payne, A.R., Whittaker, R.E.: Effect of vulcanization on the low-strain dynamic properties
of filled rubbers. J. Appl. Polym. Sci. 16, 1191–1212 (1972)
25. Kremer, F., Arndt, M.: Broadband dielectric measurement techniques. In: Runt, J.P.,
Fitzgerald, J.J. (eds.) Dielectric Spectroscopy of Polymeric Materials. American Chemical
Society, Washington (1997)
26. Kremer, F., Schönhals, A.: Analysis of dielectric spectra. In: Kremer, F., Schönhals, A.
(eds.) Broadband Dielectric Spectroscopy. Springer, Berlin (2003)
27. Riande, E., Díaz-Calleja, R.: Electrical Properties of Polymers. Marcel Dekker, New York
(2004)
28. Vassilikou-Dova, A., Kalogeras, M.: Dielectric analysis (DEA). In: Menczel, J.D., Prime,
R.B. (eds.) Thermal Analysis of Polymers, Fundamentals and Applications. Wiley,
Hoboken (2009)
29. Tsangaris, G.M., Psarras, G.C., Kouloumbi, N.: Electric modulus and interfacial
polarization in composite polymeric systems. J. Mater. Sci. 33, 2027–2037 (1998)
30. Yu, S., Hing, P., Hu, X.: Dielectric properties of polystyrene-aluminum-nitride composites.
J. Appl. Phys. 88, 398–404 (2000)
31. Psarras, G.C., Manolakaki, E., Tsangaris, G.M.: Electrical relaxations in polymeric
particulate composites of epoxy resin and metal particles. Compos. A Appl. Sci. Manuf. 33,
375–384 (2002)
32. Hernández, M., Carretero-González, J., Verdejo, R., Ezquerra, T.A., López-Manchado,
M.A.: Molecular dynamics of natural rubber/layered silicate nanocomposites as studied by
dielectric relaxation spectroscopy. Macromolecules 43, 643–651 (2010)
33. Anastasiadis, S.H., Karatasos, K., Vlachos, G., Manias, E., Giannelis, E.P.: Nanoscopic-
confinement effects on local dynamics. Phys. Rev. Lett. 84, 915–918 (2000)
34. Mijovic, J., Lee, H., Kenny, J., Mays, J.: Dynamics in polymer-silicate nanocomposites as
studied by dielectric relaxation spectroscopy and dynamic mechanical spectroscopy.
Macromolecules 39, 2172–2182 (2006)
35. Clayton, L.M., Knudsen, B., Cinke, M., Meyyappan, M., Harmon, J.P.: DC conductivity and
interfacial polarization in PMMA/nanotube and PMMA/soot composites. J. Nanosci.
Nanotechnol. 7, 3572–3579 (2007)
36. Psarras, G.C., Gatos, K.G., Karahaliou, P.K., Georga, S.N., Krontiras, C.A., Karger-Kocsis,
J.: Relaxation phenomena in rubber/layered silicate nanocomposites. Express Polym. Lett.
1, 837–845 (2007)
37. Gatos, K.G., Martínez-Alcázar, J.G., Psarras, G.C., Thomann, R., Karger-Kocsis, J.:
Polyurethane latex/water dispersible boehmite alumina nanocomposites: thermal,
mechanical and dielectrical properties. Comp. Sci. Technol. 67, 157–167 (2007)
38. Brown, R.: Physical Testing of Rubber, 4th edn. Springer, New York (2006)
Relaxation Phenomena in Elastomeric Nanocomposites 115

39. Arantes, T.M., Leão, K.V., Tavares, M.I.B., Ferreira, A.G., Longo, E., Camargo, E.R.:
NMR study of styrene-butadiene rubber (SBR) and TiO2 nanocomposites. Polym. Test. 28,
490–494 (2009)
40. Valentín, J.L., Mora-Barrantes, I., Carretero-González, J., López-Manchado, M.A., Sotta,
P., Long, D.R., Saalwächter, K.: Novel experimental approach to evaluate filler–elastomer
interactions. Macromolecules 43, 334–346 (2010)
41. Smallwood, H.M.: Limiting law of the reinforcement of rubber. J. Appl. Phys. 15, 758–766
(1944)
42. Guth, E.: Theory of filler reinforcement. J. Appl. Phys. 16, 20–25 (1945)
43. Heinrich, G., Klüppel, M., Vilgis, T.A.: Reinforcement of elastomers. Curr. Opin. Solid
State Mater. Sci. 6, 195–203 (2002)
44. Sirisinha, C., Prayoonchatphan, N.: Study of carbon black distribution in BR/NBR blends
based on damping properties: influences of carbon black particle size, filler, and rubber
polarity. J. Appl. Polym. Sci. 81, 3198–3203 (2001)
45. Ahankari, S.S., Kar, K.K.: Processing of styrene butadiene rubber–carbon black
nanocomposites with gradation of crosslinking density: static and dynamic mechanical
characterization. Mater. Sci. Eng. A 491, 454–460 (2008)
46. Le Gal, A., Guy, L., Orange, G., Bomal, Y., Klüppel, M.: Modelling of sliding friction for
carbon black and silica filled elastomers on road tracks. Wear 264, 606–615 (2008)
47. Bhattacharyya, S., Sinturel, C., Bahloul, O., Saboungi, M.L., Thomas, S., Salvetat, J.P.:
Improving reinforcement of natural rubber by networking of activated carbon nanotubes.
Carbon 46, 1037–1045 (2008)
48. Bokobza, L.: Multiwall carbon nanotube elastomeric composites: a review. Polymer 48,
4907–4920 (2007)
49. De Falco, A., Goyanes, S., Rubiolo, G.H., Mondragon, I., Marzocca, A.: Carbon nanotubes
as reinforcement of styrene-butadiene rubber. Appl. Surf. Sci. 254, 262–265 (2007)
50. Das, A., Stöckelhuber, K.W., Jurk, R., Saphiannikova, M., Fritzsche, J., Lorenz, H.,
Klüppel, M., Heinrich, G.: Modified and unmodified multiwalled carbon nanotubes in high
performance solution-styrene-butadiene and butadiene rubber blends. Polymer 49, 5276–
5283 (2008)
51. Sui, G., Zhong, W.H., Yang, X.P., Yu, Y.H.: Curing kinetics and mechanical behavior of
natural rubber reinforced with pretreated carbon nanotubes. Mater. Sci. Eng. A 485, 524–
531 (2008)
52. Das, A., Stöckelhuber, K.W., Jurk, R., Fritzsche, J., Klüppel, M., Heinrich, G.: Coupling
activity of ionic liquids between diene elastomers and multi-walled carbon nanotubes.
Carbon 47, 3313–3321 (2009)
53. Forrest, J.A., Dalnoki-Veress, K., Dutcher, J.R.: Interface and chain confinement effects on
the glass transition temperature of thin polymer films. Phys. Rev. E 56, 5705–5716 (1997)
54. van Zanten, J.H., Wallace, W.E., Wu, W.L.: Effect of strongly favorable substrate
interactions on the thermal properties of ultrathin polymer films. Phys. Rev. E 53, 2053–
2056 (1996)
55. Angellier, H., Molina-Boisseau, S., Dufresne, A.: Mechanical properties of waxy maize
starch nanocrystal reinforced natural rubber. Macromolecules 38, 9161–9170 (2005)
56. Gatos, K.G., Karger-Kocsis, J.: Rubber/clay nanocomposites: preparation, properties and
applications. In: Thomas, S., Stephen, R. (eds.) Rubber Nanocomposites. Wiley, Singapore
(2010)
57. Avalos, F., Ortiz, J.C., Zitzumbo, R., López-Manchado, M.A., Verdejo, R., Arroyo, M.:
Effect of montmorillonite intercalant structure on the cure parameters of natural rubber. Eur.
Polym. J. 44, 3108–3115 (2008)
58. Schön, F., Thomann, R., Gronski, W.: Shear controlled morphology of rubber/organoclay
nanocomposites and dynamic mechanical analysis. Macromol. Symp. 189, 105–110 (2002)
59. Carretero-González, J., Retsos, H., Verdejo, R., Toki, S., Hsiao, B.S., Giannelis, E.P.,
López-Manchado, M.A.: Effect of nanoclay on natural rubber microstructure.
Macromolecules 41, 6763–6772 (2008)
116 G. C. Psarras and K. G. Gatos

60. Sun, Y., Luo, Y., Jia, D.: Preparation and properties of natural rubber nanocomposites with
solid-state organomodified montmorillonite. J. Appl. Polym. Sci. 107, 2786–2792 (2008)
61. Varghese, S., Karger-Kocsis, J.: Natural rubber-based nanocomposites by latex
compounding with layered clays. Polymer 44, 4921–4927 (2003)
62. Varghese, S., Karger-Kocsis, J.: Melt-compounded natural rubber nanocomposites with
pristine and organophilic layered silicates of natural and synthetic origin. J. Appl. Polym.
Sci. 91, 813–819 (2004)
63. Lu, Y.L., Li, Z., Yu, Z.Z., Tian, M., Zhang, L.Q., Mai, Y.W.: Microstructure and properties
of highly filled rubber/clay nanocomposites prepared by melt blending. Comp. Sci. Technol.
67, 2903–2913 (2007)
64. Jia, Q.X., Wu, Y.P., Lu, M., He, S.J., Wang, Y.Q., Zhang, L.Q.: Interface tailoring of
layered silicate/styrene butadiene rubber nanocomposites. Compos. Interfaces 15, 193–205
(2008)
65. Shojaei, A., Faghihi, M.: Physico-mechanical properties and thermal stability of thermoset
nanocomposites based on styrene-butadiene rubber/phenolic resin blend. Mater. Sci. Eng. A
527, 917–926 (2010)
66. Jia, Q.X., Wu, Y.P., Wang, Y.Q., Lu, M., Zhang, L.Q.: Enhanced interfacial interaction of
rubber-clay nanocomposites by a novel two-step method. Comp. Sci. Technol. 68, 1050–
1056 (2008)
67. Ramorino, G., Bignotti, F., Pandini, S., Riccó, T.: Mechanical reinforcement in natural
rubber/organoclay nanocomposites. Comp. Sci. Technol. 69, 1206–1211 (2009)
68. Zhang, Z., Zhang, L., Li, Y.: Effect of the addition of toluene on the structure and properties
of styrene-isoprene-butadiene rubber/montmorillonite nanocomposites. Macromol. Mater.
Eng. 290, 430–437 (2005)
69. Ganter, M., Gronski, W., Semke, H., Zilg, T., Thomann, C., Mühlhaupt, R.: Surface-
compatibilized layered silicates: a novel class of nanofillers for rubbers with improved
mechanical properties. Kaut Gummi Kunstst 54, 166–171 (2001)
70. Lorenz, H., Fritzsche, J., Das, A., Stöckelhuber, K.W., Jurk, R., Heinrich, G., Klüppel, M.:
Advanced elastomer nano-composites based on CNT-hybrid filler systems. Comp. Sci.
Technol. 69, 2135–2143 (2009)
71. Praveen, S., Chattopadhyay, P.K., Albert, P., Dalvi, V.G., Chakraborty, B.C.,
Chattopadhyay, S.: Synergistic effect of carbon black and nanoclay fillers in styrene
butadiene rubber matrix: development of dual structure. Compos. A 40, 309–316 (2009)
72. Psarras, G.C., Gatos, K.G., Karger-Kocsis, J.: Dielectric properties of layered silicate-
reinforced natural and polyurethane rubber nanocomposites. J. Appl. Polym. Sci. 106,
1405–1411 (2007)
73. Varghese, S., Gatos, K.G., Apostolov, A.A., Karger-Kocsis, J.: Morphology and mechanical
properties of layered silicate reinforced natural and polyurethane rubber blends produced by
latex compounding. J. Appl. Polym. Sci. 92, 543–551 (2004)
74. Hedvig, P.: Dielectric Spectroscopy of Polymers. Adam Hilger Ltd, Bristol (1977)
75. Pissis, P., Apekis, L., Christodoulides, C., Niaounakis, M., Kyritsis, A., Nedbal, J.: Water
effects in polyurethane block copolymers. J. Polym. Sci. B Polym. Phys. 34, 1529–1539
(1996)
76. Korzhenko, A., Tabellout, M., Emery, J.R.: Influence of a metal–polymer interfacial
interaction on dielectric relaxation properties of polyurethane. Polymer 40, 7187–7195
(1999)
77. Steeman, P.A.M., van Turnhout, J.: Dielectric properties of inhomogeneous media. In:
Kremer, F., Schönhals, A. (eds.) Broadband Dielectric Spectroscopy. Springer, Berlin
(2003)
78. Fritzsche, J., Klüppel, M.: Molecular dynamics of silica and organoclay filled XNBR-
nanocomposites. Kaut Gummi Kunstst 1, 16–22 (2009)
79. Pietrasik, J., Gaca, M., Zaborski, M., Okrasa, L., Boiteux, G., Gain, O.: Studies of molecular
dynamics of carboxylated acrylonitrile-butadiene rubber composites containing in situ
synthesized silica particles. Eur. Polym. J. 45, 3317–3325 (2009)
Relaxation Phenomena in Elastomeric Nanocomposites 117

80. Raftopoulos, K.N., Pandis, Ch., Apekis, L., Pissis, P., Janowski, B., Pielichowski, K.,
Jaczewska, J.: Polyurethane-POSS hybrids: molecular dynamic studies. Polymer 51, 709–
718 (2010)
81. Du, M., Guo, B., Lei, Y., Liu, M., Jia, D.: Carboxylated butadiene-styrene rubber/halloysite
nanotube nanocomposites: interfacial interaction and performance. Polymer 49, 4871–4876
(2008)
82. Lu, L., Zhai, H., Zhang, Z., Ong, C., Guo, S.: Reinforcement of hydrogenated carboxylated
nitrile-butadiene rubber by multi-walled carbon nanotubes. Appl. Surf. Sci. 255, 2162–2166
(2008)
83. Teh, P.L., Mohd Ishak, Z.A., Hashim, A.S., Karger-Kocsis, J., Ishiaku, U.S.: Effects of
epoxidized natural rubber as a compatibilizer in melt compounded natural rubber-
organoclay nanocomposites. Eur. Polym. J. 40, 2513–2521 (2004)
84. Teh, P.L., Mohd Ishak, Z.A., Hashim, A.S., Karger-Kocsis, J., Ishiaku, U.S.: On the
potential of the organoclay with respect to conventional fillers (carbon black, silica) for
epoxidized natural rubber compatibilized natural rubber vulcanizates. J. Appl. Polym. Sci.
94, 2438–2445 (2004)
85. Teh, P.L., Mohd Ishak, Z.A., Hashim, A.S., Karger-Kocsis, J., Ishiaku, U.S.: Physical
properties of natural rubber/organoclay nanocomposites compatibilized with epoxidized
natural rubber. J. Appl. Polym. Sci. 100, 1083–1092 (2006)
86. Varghese, S., Karger-Kocsis, J., Gatos, K.G.: Melt compounded epoxidized natural rubber/
layered silicate nanocomposites: structure–properties relationships. Polymer 44, 3977–3983
(2003)
87. Fritzsche, J., Das, A., Jurk, R., Stöckelhuber, K.W., Heinrich, G., Klüppel, M.: Relaxation
dynamics of carboxylated nitrile rubber filled with organomodified nanoclay. Express
Polym. Lett. 2, 373–381 (2008)
88. Pradhan, S., Costa, F.R., Wagenknecht, U., Jehnichen, D., Bhowmick, A.K., Heinrich, G.:
Elastomer/LDH nanocomposites: synthesis and studies on nanoparticle dispersion,
mechanical properties and interfacial adhesion. Eur. Polym. J. 44, 3122–3132 (2008)
89. Gatos, K.G., Karger-Kocsis, J.: Effects of primary and quaternary ammine intercalants
on the organoclay dispersion in a sulfur cured EPDM rubber. Polymer 46, 3069–3076
(2005)
90. Ahmadi, S.J., G’Sell, C., Huang, Y., Ren, N., Mohaddespour, A., Hiver, J.M.: Mechanical
properties of NBR/clay nanocomposites by using a novel testing system. Comp. Sci.
Technol. 69, 2566–2572 (2009)
91. Gatos, K.G., Sawanis, N.S., Apostolov, A.A., Thomann, R., Karger-Kocsis, J.:
Nanocomposite formation in hydrogenated nitrile rubber (HNBR)/organo-montmorillonite
as a function of the intercalant type. Macromol. Mater. Eng. 289, 1079–1086 (2004)
92. Rittigstein, P., Torkelson, J.M.: Polymer–nanoparticle interfacial interactions in polymer
nanocomposites: confinement effects on glass transition temperature and suppression of
physical aging. J. Polym. Sci. B Polym. Phys. 44, 2935–2943 (2006)
93. Patsidis, A., Psarras, G.C.: Dielectric behaviour and functionality of polymer matrix—
ceramic BaTiO3 composites. Express Polym. Lett. 2, 718–726 (2008)
94. Ash, B.J., Schadler, L.S., Siegel, R.W.: Glass transition behavior of alumina/
polymethylmethacrylate nanocomposites. Mater. Lett. 55, 83–87 (2002)
95. Ash, B.J., Siegel, R.W., Schadler, L.S.: Glass-transition temperature behaviour of alumina/
PMMa nanocomposites. J. Polym. Sci. B Polym. Phys. 42, 4371–4383 (2004)
96. Wrana, C., Reinartz, K., Winkelbach, H.R.: TherbanÒ—the high performance elastomer for
the new millenium. Macromol. Mater. Eng. 286, 657–662 (2001)
97. Severe, G., White, J.L.: Physical properties and blend miscibility of hydrogenated
acrylonitrile-butadiene rubber. J. Polym. Sci. B Polym. Phys. 78, 1521–1529
(2000)
98. Felhös, D., Karger-Kocsis, J., Xu, D.: Tribological testing of peroxide cured HNBR with
different MWCNT and silica contents under dry sliding and rolling conditions against steel.
J. Appl. Polym. Sci. 108, 2840–2851 (2008)
118 G. C. Psarras and K. G. Gatos

99. Sofos, G.: Electrical response of HNBR and composite rubber blends HNBR/FKM which
incorporate MWCNTs. MSc Thesis, University of Patras, Patras, Greece (2009)
100. Psarras, G.C., Manolakaki, E., Tsangaris, G.M.: Dielectric dispersion and ac conductivity
in—iron particles loaded—polymer composites. Compos. A 34, 1187–1198 (2003)
101. Tsangaris, G.M., Psarras, G.C.: The dielectric response of a polymeric three-component
composite. J. Mater. Sci. 34, 2151–2157 (1999)
Modeling and Simulation of Polymeric
Nanocomposite Processing

Teik-Cheng Lim

Abstract As a consequence of its large surface area to volume ratio, the


embedment of nano-particles in polymers leads to large interface area and hence
considerable interphase volume region between the nanoparticles and the polymer.
The resulting bulk properties in the context of nano-particle filled polymers
therefore differ from the use of conventional particle reinforcement. While the
mechanical, electrical and other solid state physical properties of polymer nano-
composites can be easily obtained due to the static nature of the molecular and
continuum modeling, the same cannot be said so for the case of polymer nano-
composite processing due to the dynamic flow nature inherent in the latter. This
chapter lays down the common rules adopted in modeling of polymer nanocom-
posite processing. Beginning from the various interatomic and intermolecular
potential energy functions that are indispensable for molecular modeling, the
chapter presents two major approaches for molecular modeling of polymer flow.
Recent results arising from the use of molecular modeling is then summarized with
emphasis on the glass transition temperature and the rheological properties of the
polymer melt with the presence of nano-scale fillers. The chapter concludes with
the advantages of molecular modeling techniques for understanding the nano-
particle-filled polymer in the context of flow processing.

Keywords Modeling  Nanocomposites  Nanoparticles  Processing

T.-C. Lim (&)


School of Science and Technology, SIM University, 535A Clementi Road,
599490 Singapore, Republic of Singapore
e-mail: alan_tc_lim@yahoo.com

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 119


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_4,
Ó Springer-Verlag Berlin Heidelberg 2011
120 T.-C. Lim

1 Introduction

As the phrase suggests, the processing of polymer nanocomposites involves


flow of melt polymers and polymeric solutions with nano-scale fillers, i.e.
fillers whereby at least one of the dimensions is equal to or less than 100 nm.
As in conventional polymer composite processing—such as compression
molding, extrusion, injection molding, pultrusion, resin casting, rotational
molding, spinning, electrospinning, thermoforming and transfer molding—
processing of polymer nanocomposites are governed by the same processing
parameters and boundary conditions associated with the processing machinery
and flow pathway. Unlike the conventional polymer composite processing, the
unique properties arising from the large surface area to volume ratio of the
nano-scale fillers, hence the large interfacial interaction between the surface
area of the nano-scale fillers with the adjacent polymeric chain. The nano-scale
fillers can possess aspect ratio greater than 1 such as carbon nanotubes or other
nanorods [1–5], aspect ratio of 1 such as buckminsterfullerene, carbon black or
other nanoparticles [6–10], and aspect ratio less than 1 such as montmorillonite
clay or other nano-platelets [11–15]. Polymers that were processed with nano-
scale fillers include polypropylene [16], polyimide [12, 17], elastomers [18,
19], etc.
Modeling of polymer nanocomposite processing differs significantly from
computer simulation of bulk polymeric melt flow with and without con-
ventional fillers, which uses rheological properties of the polymer, the pro-
cessing condition, the filler volume fraction, the filler shape (aspect ratio)
and the mold geometry. The accuracy of any simulation of the structural
properties of a nanomaterial, and its processing thereof, are ultimately
governed by the appropriate use of interatomic and intermolecular potential
energy functions used in the simulation, and the length of time-step incre-
ment imposed on the simulation. The interatomic potentials covered include
(1) van der Waals systems, (2) covalent interactions and (3) ionic interac-
tions. Computer modeling of polymeric nanocomposites can be categorized
into two broad methods, namely the molecular dynamics (MD) approach and
the Monte Carlo (MC) approach. The MD approach has been adopted by
Douglas et al. [20, 21], Smith et al. [22–26], Keblinski et al. [27–29] and
others. The MC approach was employed by Mark et al. [30–32], Mattice
et al. [33–35], Vacatello [36–38] and Termonia [39], among others. This
chapter begins with a review of the various interatomic potential energy
functions that are used in both the MD and MC approaches. The funda-
mentals of MD and MC methodologies are then discussed with reference to
polymer processing. Finally, results from the MD and MC techniques
from various research groups are discussed with emphasis on the effect of
nano-scale fillers on the rheological and other flow properties of the melt
polymers with nano-scale fillers.
Modeling and Simulation of Polymeric Nanocomposite Processing 121

2 Potential Energy Functions

In molecular modeling, one aims to obtain the flow prediction and the corre-
sponding polymer chain conformation either on the basis of minimum energy or on
the basis of steepest descent of energy. The Hamiltonian

X
N X
N 1   XN 2  
H¼ Vi þ U2b ri;iþ1 þ U3b hi;iþ1;iþ2
i¼1 i¼1 i¼1

X
N 3   XN 1 X
N  
þ U4b si;iþ1;iþ2;iþ3 þ UNb rij ð1Þ
i¼1 i¼1 j  iþ3

gives the overall energy of the molecular system, whereby V is the kinetic energy
of each atom, U is the potential energy, N being the number of atoms in the
system, while the bond length, bond angle and torsional angle are denoted as r, h,
and s respectively. The kinetic energy
pi
Vi ¼ ; ð2Þ
2m
where pi and m refer to the momentum of ith atom and its mass, applies during
polymer flow. The interatomic potentials, U, can be broadly classified into bonded
and non-bonded interactions. The bonded interactions can be further categorized
into 2-body (stretching) potential, 3-body (bending) potential and 4-body (tor-
sional) potential. The non-bonded interaction takes place between two atoms not
directly bonded within the same molecule as well as between atoms of different
molecules, as a result of van der Waals interaction (electrodynamics) and also due
to Coulombic interaction (electrostatics).
The 2-body potential considers the energy stored when two adjacent atoms,
bonded to one another, undergoes relative displacement. The simplest potential is
based on a simple harmonic oscillator consisting of a mass m attached to a rigid
wall by a spring of stiffness kr. Hence the load F applied in stretching the spring by
an amount r can be written as
F ¼ kr r ð3Þ
according to Hooke’s Law, and

d2 r
F ¼ ma ¼ m ð4Þ
dt2
according to Newton’s Second Law. Since the summation of energy, H, is constant
1 2
mv þ U ¼ H; ð5Þ
2
122 T.-C. Lim

then differentiating with respect to time, with due consideration of (4), leads to
Hinchliffe [40]
 
dr dU
Fþ ¼ 0: ð6Þ
dt dr
Since (dr/dt)=0, we have
Z
U ¼  Fdr: ð7Þ

Substitution of (3) into (7) yields


1  2
U2b ¼ kr rij  rij;e ð8Þ
2
where rij is the bond length between ith and jth atom with subscript e referring
to equilibrium bond length, and rij - rij,e = r [41–49]. An obvious disadvan-
tage of this potential, also known as the harmonic potential, is that the potential
energy curve forms a quadratic curve, which is symmetrical about the equi-
librium bond length, thereby predicting an erroneously large energy at long
distance. This drawback was overcame by the introduction of the Morse
potential [50],
    2
U2b ¼ D 1  exp a rij  rij;e ð9Þ

where D and a are the Morse parameters. These, as well as the equilibrium bond
length, can be obtained by performing curve-fitting of data from ab initio results or
more traditionally from spectroscopic data. This potential function has been
widely used when large bond stretching is anticipated [51–55]. In most cases, the
harmonic potential would be sufficient except where long polymeric chains are
simulated due the significant bond stretching during molecular chain
entanglement.
The 3-body potential considers the molecular energy stored when the bond
angle, h, formed by three atoms is changed as a consequence of relative dis-
placement. The harmonic version of bond bending can be written in a way that
resembles (8), i.e.
1  2
U3b ¼ kh hijk  hijk;e ð10Þ
2

where kh refers to the bending stiffness by the angle formed by atom i, j and k [45–
49]. A more frequently adopted alternative to (10) is the Quadratic Potential in
Cosine [41–44, 51–56]
1  2
kh cos hijk  cos hijk;e : ð11Þ
2
Modeling and Simulation of Polymeric Nanocomposite Processing 123

Equation (11) gives rise to lower energy in comparison to (10) when h [ h0 but
higher when h \ h0. Furthermore, the asymmetric curve from (11) is more
realistic.
The 4-body potential quantifies the amount of energy stored when a series of 4
atoms undergoes torsion about its own axis. A direct way to describe the change in
energy resulting from the change in torsional angle of s is the periodic function
[45, 46, 48]
1
U4b ¼ ks ½1 þ s cosðnsÞ ð12Þ
2
in which ks is the torsional stiffness, n is the order of symmetry (or periodicity of
the potential) and s is the barrier to rotation (or phase factor). A more frequently
adopted torsional potential is written in the form [41–44]
X 5
U4b ¼ ks an cosn s ð13Þ
n¼0

whereby the coefficients an are obtained empirically [57] as a0 = 1, a1 = 1.31,


a2 = -1.414, a3 = -0.3297, a4 = 2.828 and a5 = -3.3943. In other forms to
(13), the torsional stiffness and coefficients are re-grouped such that
X
3
U4b ¼ Cn cosn s ð14Þ
n¼0

where Cn (in kJ/mol) are empirically obtained coefficients [47, 49, 52–54, 56,
58–60].
The non-bonded potential energy functions are the van der Waals interaction
imposed upon atoms that are not bonded together in the same molecule, between
atoms of different molecules, and also between molecules or clusters or atoms. The
force between such atoms is highly repulsive when brought closer together but
mildly attractive, with a minimum and followed by decay, when separated. There
are two major sets of potential functions used: exponential-6 function and the
Lennard-Jones (12-6) potential. The exponential-6 function is written [47, 54, 61]
in the original version as
 
UX6 ¼ Arij6 þ B exp Crij ð15Þ

or in the loose form as [46]


(  "  6 #)
6 h
rij i f r
UX6 ¼ 4e exp f 1   ð16aÞ
f6 r f  6 rij

or
(    "  6 #)
6 rij f rij;e
UX6 ¼ e exp f 1   ð16bÞ
f6 rij;e f6 rij
124 T.-C. Lim

whereby A, B and C are coefficients which depends on the system under consid-
eration and f is a scaling factor which reduces to the conventional Lennard-Jones
form at long range when f = 12 and coincides with the minimum well-depth of
the conventional Lennard-Jones potential when f = 13.772, with e being the
dissociation energy while r is the internuclear range at which U = 0. Both e and r
are also known as the LJ parameters, which are comparable to the more frequently
applied Lennard-Jones function [47–49, 52–55]
"   6 #
r 12 r
ULJ ¼ 4e  : ð17aÞ
rij rij

or
"   6 #
rij;e 12 rij;e
ULJ ¼e 2 : ð17bÞ
rij rij

Since the potential energy decays at long distance, the Lennard-Jones potential
is truncated (e.g. [41–44]) such that
( h 12  6 i
4e r=rij  r=rij ; rij  2:5r
ULJ ¼ ð18Þ
0; rij [ 2:5r

because the potential energy magnitude would be less than e/60 when rij = 2.5r.
Other forms of the Lennard-Jones potential energy have been attempted by Noid
and Pfeffer [51]
"    6 #
r 9 r
ULJðNPÞ ¼ e 2 3 ð19Þ
rij rij

and for the case of interaction between molecules,


 12   6 
r r
ULJðMolÞ ¼ 24e 2 13  7 : ð20Þ
r r
The Coulombic interactions (also known as electrostatic or ionic interactions) is
written as
X q2 Aq2
Uc ¼ ¼ ð21Þ
i6¼j
4pe0 rij 4pe0 R

whereby e0 = permittivity of free space, R = nearest-neighbor separation,


rij = interionic distance = aijR, and A is the Madelung constant defined as
X 1
A¼ : ð22Þ
i6¼j
aij
Modeling and Simulation of Polymeric Nanocomposite Processing 125

Due to the dynamical nature of the polymeric chain as well as the enormous
number of atoms involved, simplifying assumptions have been introduced [41–45,
48, 51–54, 56]. Firstly the groups of CH2 and CH3 are lumped together to form
united atom models, and secondly the cross terms are neglected. In some instances,
bond stretching is neglected (e.g. [56]) while others neglected bond torsional
energy [51].
The individual atoms in CH2 and CH3 groups has been taken into account by
Noid et al. [47] and Schoen et al. [62] to check the extent of improved accuracy. It
was found that modeling of CH4 and CF4 each as united atom models provide
better static results whereas consideration of the 4-site bonds for these molecules
results in better dynamic properties when matched against experimental results
[62, 63]. As such, the united atom model is applicable to structural and materials
property simulation, while consideration of all atoms is highly recommended for
process modeling.

3 The Molecular Dynamics Methodology

Arising from Newton’s second law, (4) and the force-potential relation, (7), we
have

d 2 ri dU
mi 2
¼ ð23Þ
dt dri
The finite difference approach has been adopted to solve the equations of
motion by numerical method so that the positions and velocities of each atom at
time t leads to those properties at the next time step, t ? dt. The differential
equation shown in (23), can be easily solved using the finite difference method.
One of the most commonly used approximation is the Verlet [64] algorithm, which
requires information pertaining to the position ri(t) of each atom and the corre-
sponding atom’s acceleration ai(t), as well as the position of the atom from the
previous time step, ri(t - dt). To pave a way for finite difference solution, Taylor’s
expansions for ri(t ± dt) about ri(t) is performed to give
1
ri ðt þ dtÞ ¼ ri ðtÞ þ vi ðtÞdt þ ai ðtÞðdtÞ2 þ    ð24Þ
2
and
1
ri ðt  dtÞ ¼ ri ðtÞ  vi ðtÞdt þ ai ðtÞðdtÞ2     ð25Þ
2
respectively. Substitution of (25) into (24) leads to the Verlet algorithm:

ri ðt þ dtÞ ¼ 2ri ðtÞ  ri ðt  dtÞ þ ai ðtÞðdtÞ2 þ    ð26Þ


126 T.-C. Lim

of which the accurate is valid up to the fourth order (dt)4. Subtraction of (25) from
(24) gives the atomic velocity
ri ðt þ dtÞ  ri ðt  dtÞ
vi ðtÞ ¼ ð27Þ
2dt
which is accurate to the order of (dt)2.
With these equations in place, the Verlet algorithm is performed as follows.
Starting with the atomic positions from two time steps, i.e. r(t) and r(t - dt), the
atom’s acceleration a(t) is calculated from (4). The position of each atom at
the next time step r(t ? dt) can then be calculated from (26). Due to the
dynamical nature in flow processes, there is a need to obtain the kinematic
energy. Hence the atomic velocity v(t) is calculated from (27). As information of
r(t - dt) and r(t) leads to r(t ? dt), the information of r(t) and r(t ? dt) leads
to r(t ? 2dt) in the next round of calculation. The advantages of the
Verlet algorithm are that the advancement of positions is all performed in one
step and conserves energy well even with relatively long time step. In addition,
this technique is time-reversible, highly compact and easy to program. However,
the velocities at t can be calculated only after r(t ? dt) are known. Another
disadvantage is that a trajectory can only be started if information of two time
steps are available, i.e. r(t) and r(t - dt) instead of just one time step, i.e.
r(t) and v(t).
As an alternative to (27), the Verlet algorithm was extended to the Velocity
Verlet algorithm [65], which can be generally summarized as
1
vðt þ dtÞ ¼ vðtÞ þ ½aðtÞ þ aðt þ dtÞ: ð28Þ
2
The algorithm begins with r(t) and v(t) to calculate a(t). The next position
r(t ? dt) is calculated using (24). Then the mid-step velocity is calculated as
 
dt 1
v tþ ¼ vðtÞ þ aðtÞdt: ð29Þ
2 2

followed by acceleration a(t ? dt) at the next time step. The velocity move is then
completed using
 
dt 1
vðt þ dtÞ ¼ v t þ þ aðt þ dtÞdt: ð30Þ
2 2
The main advantage of the Velocity Verlet algorithm over the original
Verlet algorithm is that the former can start with positions and velocities at time t,
and that the kinetic energy at time (t ? dt) is readily available. As with the original
Verlet algorithm, the Velocity Verlet algorithm is also numerically stable, simple
to program, time-reversible, and conserves energy well even with relatively long
time steps. The only drawback of the Velocity Verlet algorithm is that it requires
the calculation of two velocity steps.
Modeling and Simulation of Polymeric Nanocomposite Processing 127

The MD approach has been applied for simulating various physical behavior of
polymeric systems by numerous researchers including Rigby et al. [41–44], Noid
et al. [47, 51–55] and Jin et al. [48].

4 The Monte Carlo Methodology

The Monte Carlo (MC) method uses random numbers to find particle displace-
ments during molecular simulation. A frequent way of implementing an MC
calculation is to use the Metropolis method. Beginning from a starting configu-
ration of s a random distribution of particles (atoms) and molecules, one atom i is
picked randomly and displaced in a random direction by a random amount, from
rm n
i to ri , subject to the maximum displacement being the adjustable parameter
drmax. The alteration of potential energy of the system, dUmn, as a consequence of
this prescribed atomic movement is then obtained for an assumed form for the
interatomic potential. Suppose the atomic displacement gives rise to decreased
energy (dUmn B 0), then the new position is unconditionally accepted. However if
the move leads to increase in energy (dUmn C 0), then this prescribed atomic
movement is accepted only conditionally, subject to Boltzmann probability factor,
exp(-dUmn/kBT).
The MC method was the first simulation technique to obtain the equilibrium
behavior of liquids and solids [66]. With sufficiently large number of configura-
tions sampled, the average value of the energy fluctuations give the constant
volume-specific heat in liquid and, in the case of solid, fluctuations in the stress
tensor determine to the elastic constants [67]. The MC method has also been
applied to bulk polymers (e.g. [68–72]).
The MC technique by Vao-soongnern et al. [73] simulates the molecular
structure of polyethylene as a nanofiber. As in previous works (e.g. [68]), these
molecular chains are placed on a diamond lattice whereby every second site has
been removed, thereby giving rise to the second nearest neighbor diamond (2nnd)
lattice. Computation of the polymeric nanofiber chain consisted of two parts, firstly
the generation of the initial polymer nanofiber structure, which was then followed
by relaxation of the initial structure to thermodynamic equilibrium. The initial step
employs only self-avoiding walks with the excluded volume condition. In the
following step, both the intermolecular as well as the intramolecular potential
functions are then imposed [68, 69]. The initial structure is then relaxed so as to
minimize its potential energy according to the Dynamic Monte Carlo technique
[69–71]. The equilibrated free-standing nano-film structure was calculated from
previous work by Mattice et al. [72]. The fiber is then formed via collapsing this
thin film after extending, about 3 to 4 times, on one side of the periodic box in the
direction that is perpendicular to the normal axis of the thin film surface. This new
box would be sufficiently large in order that no parent chain interacts with its
images. As such, the periodic boundary appears to apply only in one direction for
this step. Two methods for fiber formation were then introduced, namely the
128 T.-C. Lim

elongation method and the cut-and-press method. In the elongation method, the
nanofiber is formed by generating in a similar way to that of thin film formation
from bulk. In the cut-and-press method, a new nanofiber is obtained from an
existing larger fiber structure. Specific details of these two methods can be found
elsewhere [73]. It has been shown that when the number of chains and chain length
are the same, both approaches lead to equivalent nanofiber structure. The
parameters simulated include the shape of the cross-section (which is almost
circular), the radial density profile, the bead distribution, the segregation of the
chain ends, the local orientation, the global equilibrium properties of the chains, as
well as the surface energy of the nanofiber.

5 Simulation Results

A number of simulation results have been reported in the literature. In this section,
the findings are reviewed, and commented with particular reference to the effect of
nano-scale filler presence during polymeric processing.
An MD study on the influence of control parameters on nanoparticle clustering
in polymer was performed by Douglas et al. [20]. Simulated results give quanti-
tative description in which (1) the polymer-matrix strength increases, at a
decreasing rate, with the nanoparticle loading; and that (2) the polymer-matrix
strength decreases with increasing temperature [20]. Investigations on the glass
transition temperature of pure polymers and those with nano-scale fillers were
performed by Lamm and Yann [74], Papakonstantopoulos et al. [75] and others.
Lamm and Yann [74] obtained the glass transition temperature of pure polyimide
(PI) as well as PI with 20.23% weight of octaaminophenyl silsesquioxanone
(OAPS) and PI with 10.45% weight of octahydro silsesquioxanone (OHS), in
which the transition temperature is marked by a change in slope of specific tem-
perature plotted against temperature. The MD result reveals an increase and
decrease of the glass transition temperature with the addition of OAPS and OHS
respectively. The modeling approach by Papakonstantopoulos et al. [75], on the
other hand, obtained the glass transition temperature by the change of slope of the
polymer density against temperature. Again, a shift on the glass transition tem-
perature is detected via MD with the inclusion of nano-fillers in the pure polymer.
In modeling the viscosity g of liquid polymers, one considers the shear stress s
arising from the shear strain rate (dc/dt) as
 S
dc
s¼g ð31Þ
dt

whereby a Newtonian flow is denoted by S = 1 while the range 0 \ S \ 1 cor-


responds to a Power-Law fluid. For the case of Newtonian fluid, the inclusion on
an additional term that describes the yield shear stress sy denotes a Bingham
plastic, i.e.
Modeling and Simulation of Polymeric Nanocomposite Processing 129

dc
s ¼ sy þ g : ð32Þ
dt
On the other hand, the Casson fluid appears to be a blend of the Power-law fluid
and Bingham plastic, and is written as
rffiffiffiffiffi
pffiffiffi pffiffiffiffi dc
s ¼ sy þ g ð33Þ
dt
or
rffiffiffiffiffiffiffiffiffi
dc dc
s ¼ sy þ 2g sy þ g2 : ð34Þ
dt dt
As can be seen in all the above fluids, the viscosity g of the fluid affects the
shear stress of the flow, which in turn influences flow speed and required applied
pressure for processing. The viscosity g and shear stress relaxation modulus G(t)
of a liquid polymer nanocomposite can be calculated by the Einstein relations
[76, 77]
* 2 t 32 +
V XX Z
g ¼ lim 4 Pab ðt Þdt 5
0 0
ð35Þ
t!1 12kB Tt
a b6¼a
0

and
V  
GðtÞ ¼ Pab ðtÞPab ð0Þ ð36Þ
kB T
where V = system volume, kB = Boltzmann’s constant, T = temperature,
Pab(0) = initial value of the off-diagonal element of the stress tensor, Pab(t) instan-
taneous value of the off-diagonal element of the stress tensor at time t, and h i averaged
value over the entire trajectory and all 6 off-diagonal elements of the stress tensor.
The normalized polymer nanocomposite viscosity, as a function of nano-filler
volume fraction /f is given by [26]
5
Fð/f Þ ¼ 1 þ /f þ 4:94/2f þ C/3f ð37Þ
2
where C = constant. Plots of normalized viscosity of polymer matrix nanocom-
posite against the specific interfacial area between polymer and nanoparticles show
that the normalized viscosity increases rapidly and weakly with the specific
interfacial area when the nanoparticle and polymer is attractive and neutral
respectively [26]. The same simulation also shows that when the nanoparticle and
polymer is repulsive, the normalized viscosity of polymer matrix nanocomposite
decreases with the specific interfacial area between polymer and nanoparticles.
Using a chain length of N = 20, Douglas et al. [21] showed that (1) the nano-
filler reinforced polymer, just like the pure polymer, experiences viscosity decay
130 T.-C. Lim

with increasing shear rate, and that (2) the viscosity of nano-filler reinforced
polymer has higher viscosity than that of pure polymer by one order. While these
two results are not surprising, it is striking to note that, with other parameters fixed,
the nano-rod fillers give the polymer the highest viscosity, followed by icosahedra
(or spherical-like) nano-fillers. The nano-platelet fillers give the lowest viscosity to
the polymers [21]. Although the difference in the polymer viscosity for the use of
these three nano-fillers are small compared to that of pure polymer, the simulated
result nevertheless provides evidence on the influence of nano-filler shape on the
overall performance of nanocomposite processing.

6 Conclusions

Arising from its length scale, nano-scale structures are best modeled at the
molecular level, which justifies the need to incorporate knowledge on the inter-
action between atoms, cluster of atoms and molecular chains. This aspect sets
nano-scale modeling apart from simulation of bulk properties of conventional
polymer composites, which can be predicted using composite mechanics
principles.
Modeling and simulation of polymer nanocomposites allow a way for visual-
ization and analysis of the processing performance in regard to the speed of
production as well as the cost or energy consumption in the processing method for
various combinations of polymer melt/solution, filler geometries, processing
temperature & pressure, and other conditions before a prototype processing is
tested.
Visualization of molecular flow enables the detection of any hazardous colli-
sion of nanoparticles on the inner surface of the tool geometry, so that the tool
designer can design or select tool geometries that can ensure smoother flow so as
to reduce or eliminate the generation of debris, which can influence the flow
pattern of the polymer nanocomposite processing as well as the properties of the
final product. The visualization also enables the detection of multi-phase flow, and
the extent of this flow if any, so as to redesign the tool geometry and/or recalibrate
the processing condition in order to reduce energy loss arising from the second
phase flow. Developmental cost is hence effectively reduced.

References

1. Ajayan, P.M., Stephan, O., Colliex, C., Trauth, D.: Aligned carbon nanotube arrays formed
by cutting a polymer resin-nanotube composite. Science 265, 1212–1214 (1994)
2. Xie, X.L., Mai, Y.W., Zhou, X.P.: Dispersion and alignment of carbon nanotubes in polymer
matrix: a review. Mater. Sci. Eng. R 49, 89–112 (2005)
3. Tjong, S.C.: Structural and mechanical properties of polymer nanocomposites. Mater. Sci.
Eng. R 53, 73–197 (2006)
Modeling and Simulation of Polymeric Nanocomposite Processing 131

4. Hussain, F., Hojjati, M., Okamoto, M., Gorga, R.E.: Poly-mer matrix nanocomposites,
processing, manufacturing, and application: an overview. J. Compos. Mater. 40, 1511–1575
(2006)
5. Moniruzzaman, M., Winey, K.I.: Polymer nanocomposites containing carbon nanotubes.
Macromolecules 39, 5194–5205 (2006)
6. Ou, Y., Yang, F., Yu, Z.: A new conception on the toughness of nylon 6/silica
nanocomposites prepared via in situ polymerization. J. Polym. Sci. Part B: Polym. Phys.
36, 789–795 (1998)
7. Jiang, L., Lam, Y.C., Tam, K.C., Chua, T.H., Sim, G.W., Ang, L.S.: Strengthening
acrylonitrile-butadiene-styrene (ABS) with nano-sized and micron-sized calcium carbonate.
Polymer 46, 243–252 (2005)
8. Chen, Y., Zhou, S., Yang, H., Gu, G., Wu, L.: Preparation and characterization of
nanocomposites polyurethane. J. Colloid Interface Sci. 276, 370–378 (2004)
9. Ma, D., Hugener, T.A., Siegel, R.W., Christerson, A., Martensson, E., Onneby, C., Schadler,
L.: Influence of nanoparticle surface modification on the electrical behaviour of polyethylene
nanocomposites. Nanotechnology 16, 724–731 (2005)
10. Zhang, M., Singh, R.: Mechanical reinforcement of un-saturated polyester by Al2O3
nanoparticles. Mater. Lett. 58, 408–412 (2004)
11. Chen, G., Wu, C., Weng, W.: Preparation of polystyrene/graphite nanosheet composite.
Polymer 44, 1781–1784 (2003)
12. Yano, K., Usuki, A.: Synthesis and properties of polyimide-clay hybrid. J. Polym. Sci. Part A:
Polym. Chem. 31, 2493–2498 (1993)
13. Ray, S.S., Okamoto, M.: Polymer/layered silicate nano-composite: a review from preparation
to processing. Prog. Polym. Sci. 28, 1539–1641 (2003)
14. Shen, J.W., Chen, X.M., Huang, W.Y.: Structure and electrical properties of grafted
polypropylene/graphite nanocomposites prepared by solution intercalation. J. Appl. Polym.
Sci. 88, 1864–1869 (2003)
15. Park, J.H., Jana, S.: Mechanism of exfoliation of nano-clay particles in epoxy-clay
nanocomposites. Macromolecules 36, 2758–2768 (2003)
16. Kawasumi, M., Hasegawa, N., Kato, M., Usuki, A., Okada, A.: Preparation and mechanical
properties of polypropylene–clay hybrids. Macromolecules 30, 6333–6338 (1997)
17. Yano, K., Usuki, A., Okada, A.: Synthesis and properties of polyimide–clay hybrid films. J.
Polym. Sci. Part A: Polym. Chem. 35, 2289–2294 (1997)
18. Maiti, M., Bhattacharya, M., Bhowmick, A.K.: Elastomer nanocomposites. Rubber Chem.
Technol. 81, 384–469 (2008)
19. Zhang, L.Q., Wang, Y.Z., Wang, Y.Q., Sui, Y., Yu, D.S.: Morphology and mechanical
properties of clay/styrene-butadiene rubber nanocomposites. J. Appl. Polym. Sci. 78, 1873–
1878 (2000)
20. Starr, F.W., Douglas, J.F., Glotzer, S.C.: Origin of particle clustering in a simulated polymer
nanocomposite and its impact on rheology. J. Chem. Phys. 119, 1777–1788 (2003)
21. Knauert, S.T., Douglas, J.F., Starr, F.W.: The effect of nanoparticle shape on polymer-
nanocomposite rheology and tensile strength. J. Polym. Sci. Part B: Polym. Phys. 45, 1882–
1897 (2007)
22. Smith, J.S., Bedrov, D., Smith, G.D.: A molecular dynamics simulation study of nanoparticle
interactions in a model polymer-nanoparticle composite. Compos. Sci. Technol. 63, 1599–
1605 (2003)
23. Bedrov, S.D., Smith, G.D., Smith, J.S.: Matrix-induced nanoparticle interactions in polymer
melts. A molecular dynamics simulation study. J. Chem. Phys. 119, 10438–10447 (2003)
24. Borodin, O., Smith, G.D., Bandyopadhyaya, R., Byutner, E.: Molecular dynamics simulation
of solid interfaces on poly(ethylene oxide) structure and dynamics. Macromolecules 36,
7873–7883 (2003)
25. Smith, J.S., Borodin, O., Smith, G.D., Kober, E.M.: A molecular dynamics simulation and
quantum chemistry study of poly(dimethylsiloxane)-silica nanoparticle interactions. J.
Polym. Sci. Part B: Polym. Phys. 45, 1599–1615 (2007)
132 T.-C. Lim

26. Smith, G.D., Bedrov, D., Li, L.W., Byutner, O.: A molecular dynamics simulation study of
the viscoelastic properties of polymer nanocomposites. J. Chem. Phys. 117, 9478–9489
(2002)
27. Desai, T., Keblinski, P., Kumar, S.K.: Molecular dynamics simulations of polymer transport
in nanocomposites. J. Chem. Phys. 122, 134910 (2005)
28. Thomin, J.D., Keblinski, P., Kumar, S.K.: Network effects on the nonlinear rheology of
polymer nanocomposites. Macromolecules 41, 5988–5991 (2008)
29. Sen, S., Thomin, J.D., Kumar, S.K., Keblinski, P.: Molecular underpinnings of the
mechanical reinforcement in polymer nanocomposites. Macromolecules 40, 4059–4067
(2007)
30. Yuan, Q.W., Kloczkowski, A., Mark, J.E., Sharaf, M.A.: Simulations on the reinforcement of
poly(dimethylsiloxane) elastomers by randomly distributed filler particles. J. Polym. Sci. Part
B: Polym. Phys. 34, 1647–1657 (1996)
31. Mark, J.E.: Some simulations on filler reinforcement in elastomers. Mol. Cryst. Liq. Cryst.
374, 29–38 (2002)
32. Mark, J.E., Abou-Hussein, R., Sen, T.Z., Kloczkowski, A.: Some simulations on filler
reinforcement in polymers. Polymer 46, 8894–8904 (2005)
33. Erguney, F.M., Mattice, W.L.: Response of matrix chains to nanoscale filler particles.
Polymer 49, 2621–2623 (2008)
34. Erguney, F.M., Lin, H., Mattice, W.L.: Dimensions of matrix chains in polymers filled with
energetically neutral nanoparticles. Polymer 47, 3689–3695 (2006)
35. Lin, H., Erguney, F., Mattice, W.L.: Collapsed chains as models for filler particles in a
polymer melt. Polymer 46, 6154–6162 (2005)
36. Vacatello, M.: Molecular arrangements in polymer-based nanocomposites. Macromol Theory
Simul 11, 757–765 (2002)
37. Vacatello, M.: Chain dimensions in filled polymers: an intriguing problem. Macromolecules
35, 8191–8193 (2002)
38. Vacatello, M.: Monte Carlo simulations of polymer melts filled with solid nanoparticles.
Macromolecules 34, 1946–1952 (2001)
39. Termonia, Y.: Monte Carlo modeling of dense polymer melts near nanoparticles. Polymer 50,
1062–1066 (2009)
40. Hinchliffe, A.: Modelling molecular structures. Wiley, England (1996)
41. Rigby, D., Roe, R.J.: Molecular dynamics simulation of polymer liquid and glass. I. Glass
transition. J. Chem. Phys. 87, 285–7292 (1987)
42. Rigby, D., Roe, R.J.: Molecular dynamics simulation of polymer liquid and glass. II. Short
range order and orientation correlation. J. Chem. Phys. 89, 5280–5290 (1988)
43. Rigby, D., Roe, R.J.: Molecular dynamics simulation of polymer liquid and glass. 3. Chain
conformation. Macromolecules 22, 2259–2264 (1989)
44. Rigby, D., Roe, R.J.: Molecular dynamics simulation of polymer liquid and glass. 4. Free-
volume distribution. Macromolecules 23, 5312–5319 (1990)
45. Sun, Z., Morgan, R.J., Lewis, D.N.: Calculation of crystalline modulus of syndiotactic
polystyrene using molecular modeling. Polymer 33, 725–727 (1992)
46. Fan, C.F., Cagin, T., Chen, Z.M., Smith, K.A.: Molecular modeling of polycarbonate. 1.
Force field, static structure, and mechanical properties. Macromolecules 27, 2383–2391
(1994)
47. Noid, D.W., Tuzun, R.E., Sumpter, B.G.: On the importance of quantum mechanics for
nanotechnology. Nanotechnology 8, 119–127 (1997)
48. Jin, Y., Boyd, R.H.: Subglass Chain dynamics and relaxation in polyethylene: a molecular
dynamics simulation study. J. Chem. Phys. 108, 9912–9923 (1998)
49. Fukui, K., Sumpter, B.G., Barnes, M.D., Noid, D.W.: Molecular dynamics studies of the
structure and properties of polymer nano-particles. Comp. Theor. Polym. Sci. 9, 245–254
(1996)
50. Morse, P.M.: Diatomic molecules according to the wave mechanics. II. Vibrational levels.
Phys. Rev. 34, 57–64 (1929)
Modeling and Simulation of Polymeric Nanocomposite Processing 133

51. Noid, D.W., Pfeffer, G.A.: Short time molecular dynamics simulations: stressed polyethylene
results. J. Polym. Sci. Part B: Polym. Phys. 27, 2321–2335 (1989)
52. Sumpter, B.G., Noid, D.W., Wunderlich, B.: Computer experiments on the internal dynamics
of crystalline polyethylene: mechanistic details of conformational disorder. J. Chem. Phys.
93, 6875–6889 (1990)
53. Noid, D.W., Sumpter, B.G., Wunderlich, B.: Molecular dynamics simulation of twist motion
in polyethylene. Macromolecules 24, 4148–4151 (1991)
54. Sumpter, B.G., Noid, D.W., Wunderlich, B.: Computational experiments on the motion and
generation of defects in polymer crystals. Macromolecules 25, 7247–7255 (1992)
55. Tuzun, R.E., Noid, D.W., Sumpter, B.G.: The dynamics of molecular bearings.
Nanotechnology 6, 64–74 (1995)
56. Brown, D., Clarke, J.H.R.: Molecular dynamics of an amorphous polymer under tension. 1.
Phenomenology. Macromolecules 24, 2075–2082 (1991)
57. Ryckaert, J.P., Bellemans, A.: Molecular dynamics of liquid n-butane near its boiling point.
Chem. Phys. Lett. 30, 123–125 (1975)
58. Steele, D.J.: An ab initio investigation of the torsional potential function of n-butane. J.
Chem. Soc. Faraday Trans. II 81, 1077–1083 (1985)
59. Boyd, R.H.: Method for calculation of the conformation of minimum potential energy and
thermodynamic functions of molecules from empirical valence force potentials—application
to the cyclophanes. J. Chem. Phys. 49, 2574–2583 (1968)
60. Sorensen, R.A., Liam, W.B., Boyd, R.H.: Prediction of polymer crystal structures and
properties: a method utilizing simultaneous inter- and intramolecular energy minimization.
Macromolecules 21, 194–199 (1988)
61. Williams, D.E.: Nonbonded potential parameters derived from crystalline hydrocarbons. J.
Chem. Phys. 47, 4680–4684 (1967)
62. Schoen, M., Hoheisel, C., Beyer, O.: Liquid CH4, liquid CF4 and the partially miscible liquid
mixture of CH4/CF4: a molecular dynamics study based on both a spherically symmetric and
a four-centre Lennard-Jones model. Mol Phys 58, 699–709 (1986)
63. Tashiro, K.: Molecular dynamics calculation to clarify the relationship between structure and
mechanical properties of polymer crystals: the case of orthorhombic polyethylene. Comp
Theo Polym Sci 11, 357–374 (2001)
64. Verlet, L.: Computer ‘‘experiments’’ on classical fluids. I. Thermodynamical properties of
Lennard-Jones molecules. Phys. Rev. 159, 98–103 (1967)
65. Swope, W.C., Anderson, H.C., Berens, P.H., Wilson, K.R.: A computer simulation method
for the calculation of equilibrium constants for the formation of physical clusters of
molecules: application to small water clusters. J. Chem. Phys. 76, 637–649 (1982)
66. Metropolis, N.A., Rosenbluth, A.W., Rosenbluth, M.N., Teller, A.H., Teller, E.: Equation of
state calculations by fast computing machines. J. Chem. Phys. 21, 1087–1092 (1953)
67. Klein, M.L., Murphy, R.D.: Elastic constants of solid Ar, Kr, and Xe: a Monte Carlo study.
Phys. Rev. B 6, 2433–2442 (1972)
68. Rapold, R.F., Mattice, W.L.: Introduction of short and long range energies to simulate real
chains on the 2nd lattice. Macromolecules 29, 2457–2466 (1996)
69. Cho, J.H., Mattice, W.L.: Estimation of long-range interaction in coarse-grained rotational
isomeric state polyethylene chains on a high coordination lattice. Macromolecules 30, 637–
644 (1997)
70. Doruker, P., Mattice, W.L.: Reverse mapping of coarse-grained polyethylene chains from the
second nearest neighbor diamond lattice to an atomistic model in continuous space.
Macromolecules 30, 5520–5526 (1997)
71. Doruker, P., Mattice, W.L.: Dynamics of bulk polyethylene on a high coordination lattice.
Macromol. Symp. 133, 47–70 (1999)
72. Doruker, P., Mattice, W.L.: Simulation of polyethylene thin films on a high coordination
lattice. Macromolecules 31, 1418–1426 (1998)
73. Vao-soongnern, V., Doruker, P., Mattice, W.L.: Simulation of an amorphous polyethylene
nanofiber on a high coordination lattice. Macromol. Theor. Simul. 9, 1–13 (2000)
134 T.-C. Lim

74. Yani, Y., Lamm, M.H.: Molecular dynamics simulation of polyimide-octahydrido


silsesquioxane and polyimide-octaaminophenyl silsesquioxane systems. Polymer 50, 1324–
1332 (2009)
75. Papakonstantopoulos, G.J., Yoshimoto, K., Doxastakis, M., Nealey, P.F., de Pablo, J.J.: Local
mechanical properties of polymeric nanocomposites. Phys. Rev. E Stat. Nonlinear Soft
Matter Phys. 72, 031801 (2005)
76. Einstein, A.: Eine neue Bestimmung der Molekuldi-mensionen. Annalen der Physik 19, 289–
306 (1906)
77. Einstein, A.: Errata: Eine neue Bestimmung der Mole-kuldimensionen. Annalen der Physik
34, 591–592 (1911)
Deformation-Induced Structure Changes
in Elastomeric Nanocomposites

Shigeyuki Toki and Benjamin S. Hsiao

Abstract During tensile deformation, nanofillers can orient and align with
polymer chains to reinforce the strength and modulus of elastomeric nanocom-
posites. The presence of nanofillers can also enhance the orientation of sur-
rounding polymer chains and accelerate the strain-induced crystallization
behavior. Conventionally, natural rubber, styrene-butadiene rubber and thermo-
plastic elastomer (ethylene-propylene copolymer and poly-urethane) are routinely
reinforced with fillers. In this chapter, the behavior of nanofiller-enhanced strain-
induced crystallization of elastomeric nanocomposites, based on natural/synthetic
rubbers and fillers of varying sizes (from microns to nanometers) such as carbon
black, multi-walled carbon nanotube (MWCNT), carbon nanofiber (CNF), nano-
clay (NC) during deformation, is described.

1 Introduction

Elastomer is defined as the kind of soft material, which can be stretched to a large
deformation and exhibits a spring-like characteristic, capable of recovering almost
to the original length when released. This behavior is caused by the mobility of the
long flexible polymer chains (with a low Tg) and the corresponding network
structure that can receive and transfer the stress due to the chain movement
(i.e., the origin of entropy modulus). The most widely studied elastomer is natural
rubber (NR), which also has a very broad range of practical applications. In fact,
NR is the second most consumed biopolymer on earth (cellulose is the most

S. Toki and B. S. Hsiao (&)


Department of Chemistry, Stony Brook University, Stony Brook, NY 11794-3400, USA
e-mail: bhsiao@notes.cc.sunysb.edu

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 135


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_5,
Ó Springer-Verlag Berlin Heidelberg 2011
136 S. Toki and B. S. Hsiao

consumed one). NR contains naturally polymerized poly-isoprene chains as the


major component, as well as proteins, carbohydrates and phospholipids. These
non-rubber components also reinforce the rubber component (polyisoprene chains)
as nanofillers.
Carbon black is the most widely used filler to reinforce the mechanical
properties of NR. During deformation, the carbon black particulates often behave
as additional (physical) network points and accelerate strain-induced crystalliza-
tion of the NR matrix. Recently, with the advance of nanomaterial synthesis, many
newly available nanofillers have also been used to reinforce the mechanical
properties of both natural and synthetic elastomers. For example, multi-walled
carbon nanotubes (MWCNT) have been routinely added to styrene-butadiene
rubber (SBR), NR and thermoplastic elastomers to enhance their mechanical
properties. As the size of MWCNT is much smaller, the amount of MWCNT
needed to achieve the same reinforcing properties is only about one-third of the
amount of carbon black typically used. As MWCNT has a thread-like shape (its
diameter is usually tens of nanometers but its length is in microns), its homoge-
neous dispersion capability in the viscous/rubbery matrix is usually a concern. For
this purpose, proper surface modification of MWCNT is necessary to induce good
dispersion capability. Because of the unique shape of MWCNT, its filled elastomer
also showed peculiar behavior after deformation due to the complex relaxation and
interactions of polymer chains and MWCNT.
Carbon nanofiber (CNF) is an analog of MWCNT but with a larger diameter. It
has also been used to reinforce the properties of various elastomers. The surface
modification of CNF is relatively easier than that of MWCNT, thus CFN may be
more convenient to use to induce strain-induced crystallization of elastomers (such
as ethylene-propylene copolymer).
Another class of nanofillers that has also been demonstrated to enhance the
mechanical and physical properties (such as gas permeability) of elastomers is
nanoclay. For example, surfactant-modified montmorillonite can also accelerate
strain-induced crystallization in NR and thus increase the total crystallinity.
Similar to nanofibers, nanoclays can also orient and align upon stretching, where
the presence of nanoclays can also induce orientation of surrounding polymer
chains and overall strain-induced crystallization, resulting in reinforcement of the
mechanical properties of this class of nanocomposite.

2 Non-Rubber Components in Natural Rubber

Natural rubber (NR) is composed of 94 wt% of rubber components and 6 wt% of


non-rubber components such as proteins, carbohydrates, phospholipids and metal
ions. Rubber components mainly consist of polyisoprene (with 100% of cis-1,4
structure) having functional groups at chain ends. The synthetic analogue of nat-
ural rubber is polyisoprene (e.g. IR2200), which is 100% polyisoprene with 98.5%
of cis-1,4 structure. IR2200 showed significantly inferior mechanical properties to
Deformation-Induced Structure Changes 137

NR. Although the inferior properties have been partially attributed to the lower
content of cis-1,4 structure (98.5%), the main reason is due to the absence of non-
rubber components in the compound, which have been confirmed recently [1, 2].
It is now clear that in NR, non-rubber components create naturally occurring
networks and multi-scaled structures simultaneously. Recently, small-angle X-ray
scattering (SAXS) and small-angle neutron scattering (SANS) techniques have
been used to reveal the existence of multi-scaled structures (from nanometers to
micrometers) in NR [1, 2]. Figure 1 illustrates the integrated SAXS profiles of NR,
IR2200 and de-proteinized natural rubber (DPNR). The comparison of these
profiles indicates that the excess scattered intensity in NR is due to the scattering
from proteins and other non-rubber components.
The corresponding wide-angle X-ray diffraction (WAXD) patterns of NR,
DPNR, Gel and Sol fractions of the NR sample (Gel and Sol are classified as the
heavy part and light part of toluene dissolved NR) are shown in Fig. 2. It was seen
that a small amount of large crystals is present in NR, as well as in Gel samples,
but not in DPNR and Sol samples. Therefore, protein agglomerates and crystals
(identified as quebrachitol), as well as phospholipid micelles, are distributed
throughout the NR matrix, which can be confirmed by optical microscopy shown
in Fig. 3.
The decrease in the sizes of these structures by mechanical processes can be
seen from the SAXS profiles of NR and NR0 samples shown in Fig. 4, where NR
was the original sample and NR0 was subjected to a two-roll milling for 5 min at
room temperature. It is seen that the SAXS profile of NR0 (dotted line) at the
scattering vector q = 0.003 shows much lower scattering intensity than that of
NR. This confirms that the large scattering structures (in micron size) are
decreased by mechanical milling. On the other hand, SAXS profiles of NR0 at the
scattering vector q = 0.04 show a higher scattered intensity than those of NR. This
indicates that the amount of smaller scattering structures (in nanometer size) is

Fig. 1 SAXS intensity (I(s)) 1e+5


plotted as a function of the
scattering vector for several
NR
rubbers (NR, DPNR and IR) 1e+4
(Reproduced from Reference
DPNR
Inntensity (arb.)

[2] with permission from


1e+3
John Wiley & sons Inc.) IR

1e+2

1e+1

1e+0
0.00 0.05 0.10 0.15 0.20
1
s / nm -1
138 S. Toki and B. S. Hsiao

Fig. 2 2D-WAXD patterns


of Gel, NR, DPNR, and Sol
samples under static condi-
tions (i.e., no deformation)
(Reproduced from Reference
[2] with permission from
John Wiley & sons Inc.)

Fig. 3 Optical micrograph (A) and cross polarized micrograph (B) of NR. Three kinds of
microstructures in NR were identified: (a) protein aggregates, (b) quebrachitol crystals and (c)
micelles of phospholipids and fatty acids (Reproduced from Reference [2] with permission from
John Wiley & sons Inc.)

increased by mechanical milling (the above observations are indicated by two


arrows in Fig. 4).
In NR, the functional groups at both ends of polyisoprene chains can react with
non-rubber components to form a naturally occurring network. This naturally
occurring network is responsible for the high green strength and high viscous
behavior of unvulcanized NR, where the effect can be clearly identified in Fig. 5
(the stress–strain curves of unvulcanized NR, DPNR and IR).
It is seen that unvulcanized IR shows lower modulus and no spring-back
behavior. In fact unvulcanized IR behaves as a polymer melt because it has a low
Tg (-70°C) and no network structure. Unvulcanized NR shows strain-induced
crystallization, which was clearly supported by the in situ WAXD measurement
during extension and retraction (as illustrated in Fig. 6).
Deformation-Induced Structure Changes 139

Fig. 4 SANS (symbols) and


SAXS profiles (thin lines) of
nature rubber (a) before (NR)
and (b) after milling (NR0)
without cross-linking
(Reproduced from Reference
[1] with permission from
American Chemical Society.)

Fig. 5 Stress–stress (elonga-


tion ratio) curves of NR,
DPNR, and IR samples at
room temperature. Repro-
duced from Reference [1]
with permission from Ameri-
can Chemical Society.)

Based on the above studies, the non-rubber components in NR, such as protein
agglomerates, quebrachitol crystals and phospholipid micelles, can be considered
as nanofillers that offer additional physical crosslinking points to enhance the
properties.

3 Carbon Black Filled Natural Rubber

Carbon black has been used extensively to reinforce general rubber compounds
such as NR and styrene-butadiene rubber (SBR). The amount of carbon black in
general rubber compounds is often from 30 to 50 phr (i.e., parts per hundred parts
of rubber). Carbon black is composed of primary particles (the diameter ranges
from 20 to 30 nm) and aggregates (primary particles make aggregates, the radius
140 S. Toki and B. S. Hsiao

0.4
Stress / MPa

0.3

0.2

0.1
S

0.0

0 1 2 3 4 5 6
Strain

Fig. 6 Stress–strain curve and selected WAXD patterns during extension and retraction of
un-vulcanized NR at 25°C. Each image was taken at the average strain indicated by the arrow
(Reproduced from Reference [3] with permission from Elsevier Science)

of gyration of the aggregates is around 1 micron and the aspect ratio is around
5–6). These numbers vary with the kind of carbon blacks used.
The presence of carbon black generally increases modulus and tensile strength
however, it decreases elongation at the break of rubber compounds. The most
effective means to enhance the tensile strength is to increase elongation at break
through strain-induced crystallization [3]. The stress–strain relation and selected
3D WAXD patterns of carbon black-filled NR during extension and retraction are
shown in Fig. 7. Each WAXD image was taken at the average strain value indi-
cated by the arrow. It is seen that the presence of carbon black in fact accelerates
the strain-induced crystallization process [4].
It is well known that the flexibility of polymer chains in NR depends mainly on
the content of the cis-1,4 structure (the trans-1,4 structure is too rigid). During
vulcanization, the transformation of cis- to trans- would occur. The higher curing
temperature and the longer curing time can cause more transformation. Therefore,
the curing condition is one of the important factors to keep the polymer chains
flexible. The sulfur vulcanization process would create mono-sulfur bridges,
di-sulfur bridges and multi-sulfur bridges in vulcanized compounds. To assure the
chain flexibility between the network points, the choice of accelerator is also
important. The choice of carbon black and the amount of carbon black also turns
Deformation-Induced Structure Changes 141

25

20
Stress / MPa

15

10

0 1 2 3 4 5 6
Strain

Fig. 7 The stress–strain curve and selected 3D WAXD profile of the sample B-2 during
extension and retraction. Each image was taken at the average strain indicated by the arrow.
Reproduced from Reference [4] with permission from KHK Kautschuk Gummi Kunststoffe)

out to be an effective tool to keep the polymer chain flexible. For example, the
crystal fraction of B2 (i.e., NR prepared at lower cure temperature, shorter cure
time and less carbon black amounts) is higher than that of B1 (i.e., NR prepared at
higher cure temperature, longer cure time and more carbon black amounts) in
Fig. 8.
The stress–strain curves showed that the elongation at break and tensile strength
of B2 are larger than those of B1 (as illustrated in Fig. 9). It is interesting to note
that the tensile strength of the B2 compound officially reached 42.5 MPa (which is
the highest value ever reported [3] in the literature we searched.) under the con-
dition of ISO-37. Therefore, the presence of carbon black accelerates strain-
induced crystallization and simultaneously increases the elongation at break
leading to very high tensile strength.

4 Multi-Walled Carbon Nanotube Filled Elastomers

Multi-walled carbon nanotube (MWCNT) has become one of the most widely
investigated nanofillers in elastomeric composites. [5,6] The diameter of MWCNT
142 S. Toki and B. S. Hsiao

Fig. 8 The changes of the 30


anisotropic fraction in the Crystal
samples B-1 and B-2 during 25
extension. Reproduced from B-2

Anisotropic fraction / %
Reference [4] with permis-
20
sion from KHK Kautschuk
B-1
Gummi Kunststoffe)
15

10

0
Oriented amorphous

-5
0 1 2 3 4 5 6
Strain

Fig. 9 Stress–strain curves 50


for samples B-1 and B-2
during extension. Reproduced 40
from Reference [4] with per- B-1
Stress / MPa

mission from KHK Kauts-


chuk Gummi Kunststoffe) 30

20

10 B-2

0 1 2 3 4 5 6 7
Strain

ranges from one nanometer to tens of nanometers and the length ranges from
several micrometers to millimeters or even centimeters (as illustrated in
Fig. 10a–d). Figure 10a shows that MWCNT has a broad distribution of length
(0.1–5 lm) and diameter (10–50 nm). Figure 10b shows the curling structure of an
isolated tube. Figure 10c shows the entanglements of the tubes and the formation
of an interconnected network structure. Figure 10d shows the nanostructure of a
multi-walled carbon nanotube with several layers of carbon graphite and a hollow
core.
The dispersion of MWCNT in rubber is usually poor, as shown in Fig. 10e–h
(TEM images of SBR with 4 phr of MWNTs). In these figures, a bundle (Fig. 10e)
and black spots (Figs. 10f, g) are observed. The magnified view of one bundle
(Fig. 10h) shows that MWCNTs are orientated. AFM images of SBR filled with
10 phr of MWCNTs (Fig. 11) showed a larger scale aggregation than that
Deformation-Induced Structure Changes 143

Fig. 10 TEM images of MWNT and corresponding composites. a–d: pure MWNT; e–h:
SBR/4 phr MWNT composite (Reproduced from Reference [6] with permission from Elsevier
Science)
144 S. Toki and B. S. Hsiao

observed in Fig. 10, as large bright domains were seen. In AFM imaging, these
bright domains can be ascribed to filler agglomerates.
Stress-strain curves for SBR and MWCNT/SBR composite samples are shown
in Fig. 12. The reinforcement effect by MWCNT was quite significant on
mechanical properties of composites, since the presence of 1 phr of MWCNTs has
increased the modulus by 45% and tensile strength by 70%.
This behavior can be explained by theoretical predictions [5], which indicate that
the modulus can be increased by both volume fraction and aspect ratio of the fillers.
Figure 13 shows the comparison of experimental results and theoretical predictions
using the Guth model and Halpin/Tsai model, respectively. To fit the experimental
data, the aspect ratios of the fillers for both models were chosen to be 40 and 45,

0.0 2.0 4.0 µm 0.0 2.0 4.0 µm 0.0 2.0 4.0 µm


Topography Signal error Phase

Fig. 11 AFM images of SBR filled with 10 phr of MWNT (Reproduced from Reference [6] with
permission from Elsevier Science)

Fig. 12 Stress–strain curves


for pure SBR and MWNT/ 10 phr
7 phr
SBR composites. The filler
content is expressed in phr 4
(i.e., parts per hundred parts
5 phr
Nominal stress / MPa

of rubber) (Reproduced from


Reference [6] with permis- 3
sion from Elsevier Science) 4 phr

2 2 phr

1 phr

1
0 phr

0
0 100 200 300 400
Strain (%)
Deformation-Induced Structure Changes 145

Fig. 13 Experimental 8
modulii plotted against the
Experimental points
volume fraction of MWNT 7
and comparison with theoret- Guth model (f=40)
ical predictions (Reproduced
from Reference [6] with 6 Halpin / Tsai model (f=45)
permission from Elsevier
Science) 5

E / E0
4

0
0 0.01 0.02 0.03 0.04 0.05
Volume fraction of filler

Stretching direction

0.0 2.0 4.0 µm 0.0 2.0 4.0 µm


Extension ratio : α = 1.8 Extension ratio : α = 2.8

Fig. 14 AFM phase-contrast images for the 10 phr MWNT-filled SBR at two different strains
(Reproduced from Reference [6] with permission from Elsevier Science)

respectively. The results suggest that MWCNTs in SBR were aggregated since the
real aspect ratio of each MWCNT particle is more than 100 or even 1000.
The behaviors of filler aggregates during deformation were also observed by
AFM, where typical results are shown in Fig. 14. In this figure, tensile deformation
of the sample resulted in orientation of the filler aggregates, which appeared as the
146 S. Toki and B. S. Hsiao

Fig. 15 Mullins hysteresis 5


for the 10 phr MWNT-filled

Nominal stress / MPa


SBR composite (Reproduced 4
from Reference [6] with per-
mission from Elsevier First and second stretch
Science) 3

0
0 50 100 150 200 250
Strain (%)

white regions. At the highest stretch ratio, the bundle aggregate was found to break
up into long straight structures.
After stretching and retraction, the SBR compound filled with 10 phr MWNT
exhibited a large degree of permanent orientation, where the stress–strain rela-
tion is shown in Fig. 15. The hysteresis or stress-softening (Mullins effect)
behavior could be explained by the loss of elastic chains taking place at the
polymer-filler interface. It was found that the reinforcement effect of MWCNT
in NR is almost the same as discussed above, which has also been reported in
the literature [6].
MWCNT filled polyurethane elastomer showed peculiar behavior because of
the complicated interaction between MWCNT and polymer crystals [7]. The
concept of the peculiar response of MWCNT filled polyurethane was that both
MWCNTs and strain-induced crystals of soft segments in polyurethane can act as
physical cross-linkers, controlling the mechanical and stimuli responsive behavior
of this class of composites.

5 Carbon Nanofiber Filled Elastomer

Carbon nanofiber (CNF) is a tubular structure with the sidewalls composed of


angled graphite sheets, whose configuration is quite different from carbon nano-
tube (CNT). Chemically modified CNFs (MCNFs) have been used to reinforce
ethylene-propylene rubber (EPR) [8]. In the modification scheme, the surface of
MCNFs was grafted with short polypropylene chains, resulting in good dispersion
in the EPR matrix. The chosen CNFs had a diameter ranging from 60 to 150 nm
and a length ranging from 30 to 100 lm. The EPR has 84.3 mol% of propylene
content co-polymerized with ethylene monomer using a metallocene catalyst. The
polypropylene segments would crystallize and act as network points, whereby
MCNF can increase the crystal content of EPR as nuclear agents for crystallization
of the PP segments.
Deformation-Induced Structure Changes 147

Fig. 16 Stress–strain curves


and selected WAXD patterns
acquired during stretching of
pure copolymer (unfilled cir-
cles) and 10 wt% nanocom-
posite (filled circles) at room
temperature (Reproduced
from Reference [9] with per-
mission from Elsevier
Science)

The stress–strain curves and selected WAXD patterns acquired during


stretching of the unfilled EPR and 10 wt% MCNF/EPR nanocomposite samples at
room temperature are compared in Fig. 16. It was found that the WAXD patterns
exhibited differences between filled and unfilled samples, where the presence of an
outer ring (at a scattering vector q = 18.4 nm-1) was seen in the filled sample but
not in the unfilled sample. This scattering peak corresponds to the 3.4 Å spacing,
which is inter-shell spacing within the CNF structure (d002).
The circularly averaged WAXD intensity profiles for both filled (10 wt%
MCFN) and unfilled EPR samples taken at different strains are shown in Fig. 17.
It was seen that during deformation, crystal structure transformation from c to a
phase occurred. (We note that the crystal morphology of isotactic polypropylene
usually contains a phase lamellae with epitaxial growth of c phase). A small
amount of c phase crystals was found to remain in the filled samples, but the
corresponding amount of the c phase in the unfilled samples was lower. Appar-
ently, the presence of MCNF affected the process of destruction of c phase crystal
during deformation.
The stress–strain relation of unfilled and 10 wt% MCFN filled EPR elasto-
mers taken at 55°C are shown in Fig. 18. It is clear that the tensile strength,
elongation at break and toughness of the 10 wt% MCNF filled samples were
significantly higher than the unfilled sample. Selected integrated WAXD profiles
of filled and un-filled samples taken at different strains are shown in Fig. 19.
148 S. Toki and B. S. Hsiao

Fig. 17 Deconvolution of circularly averaged WAXD intensity profiles into amorphous and
crystalline components. The WAXD patterns were acquired during stretching at room temper-
ature of pure copolymer (a–c) and 10 wt% nanocomposite (d–f). The corresponding strains are
indicated (Reproduced from Reference [9] with permission from Elsevier Science)

It was found that the higher initial crystallinity and higher orientation of polymer
chains in the filled sample could lead to thicker crystals that melt at higher
temperatures. The load experienced by the polymer matrix is effectively reduced
by the network formation of MCNF fillers. The reduced load effect became more
efficient at higher temperatures. Therefore, the strong bonding in the MCNF-
polymer interface is essential for the mechanical reinforcement of nanofiller
filled elastomeric composites. The efficient stress transfer in the MCNF/EPR
composite can be attributed to two factors (1) MCFN is an effective crystal
nucleating agent for crystallization of the matrix and (2) the presence of MCNF
Deformation-Induced Structure Changes 149

Fig. 18 Stress–strain curves


and selected WAXD patterns
acquired during stretching of
the pure copolymer (unfilled
circles) and 10 wt% nano-
composite (filled circles) at
55°C (Reproduced from Ref-
erence [9] with permission
from Elsevier Science)

forms an additional network structure that enhances the stress transfer during
deformation.

6 Nano-Clay Filled Natural Rubber

Nano-clay filled NR usually does not contain exfoliated clay layers, rather it
contains finely dispersed clay tactoids (i.e., nano-clay stacks containing around 10
layers) in the NR matrix. These tactoids are largely isolated and separated from
each other without forming a filler framework. The average thickness of the clay
tactoids is 100–500 nm, but the average lateral size is considerably larger. The
presence of clay tactoids can result in a faster crystallization rate and different
morphology of NR upon deformation. This is because the large interfacial surface
areas introduced by nano-clays facilitate the overall chain orientation during
deformation, resulting in an increase in total crystalline fraction. The experimental
observations are described as follows.
Figure 20 shows TEM images and SAXS patterns of a chosen nano-clay/NR
nanocomposite. The TEM image indicated the tactoids were finely dispersed and
separated at 10–40 nm apart. Directional SAXS patterns (A: face-on view; B and
C: edge-on views) showed that nano-clay stacks were aligned parallel to the film
plane with preferred orientation due to processing.
Figure 21 shows stress–strain relations during stretching and retraction for
unfilled and three filled (NR/Na+-MMT, NR/O-MMT, and NR/O-LAP)
150 S. Toki and B. S. Hsiao

Fig. 19 Deconvolution of circularly averaged WAXD intensity profiles into amorphous and
crystalline components. The patterns were acquired during stretching at 55°C of the pure
copolymer (a, b) and the 10 wt% nanocomposites (c, d). The strain was zero for (a) and (c),
2.8 for (b) and 22.7 for (d) (Reproduced from Reference [9] with permission from Elsevier
Science)

Fig. 20 Representative TEM


images (I: scale bar 1000 nm;
II: scale bar 50 nm) and
SAXS patterns (A: face-on
view; B, C: edge-on views) of
the NR-NC1 nanocomposite
(Reproduced from Reference
[10] with permission from
American Chemical Society)
Deformation-Induced Structure Changes 151

Fig. 21 Stress–strain curves and selected synchrotron WAXD patterns during extension and
retraction cycles for the chosen samples. The corresponding stress–strain curve for the NR/
O-LAP nanocomposites is included for comparison. 2D WAXD patterns for (A) the unfilled NR,
(B) NR/Na+-MMT, and (C) NR/O-MMT are illustrated at selected strains (Reproduced from
Reference [10] with permission from American Chemical Society)

nanocomposites where Na+-MMT represents sodium montmorillonite, O-MMT is


organically modified montmorillonite and O-LAP is modified synthetic hectrite
laponite clay). Selected WAXD patterns of these samples showed that nano-clays
increased strain-induced crystallization and decreased the onset strain of strain-
induced crystallization.
Figure 22 shows the evolution of crystallinity index (CI) with strain, which
exhibited significant hysteresis behavior as in the stress–strain relations in Fig. 21.
Contrary to the stress values during the extension-retraction cycle, the CI values
are larger during retraction than those during extension. The onset strain of
crystallization was determined by the interception of the regression line in the plot
of CI against the strain in Fig. 22. The onset strains of all these nano-clay
composites are smaller than those of unfilled NR. The difference in the aspect ratio
of these nano-clays might play an important role in the onset strain and the
maximum CI of these composites. At the stretching ratio a \ 3, the orientation and
alignment of highly anisotropic nano-clay particles occurred; while at 3 \ a \ 4,
152 S. Toki and B. S. Hsiao

Fig. 22 Crystallinity index (CI) as a function of strain during the stretch-recovery cycle for (a)
unfilled NR, (b) NR/Na+-MMT, (c) NR/OMMT, and (d) NR/O-LAP samples. The solid line is
only a guide to the eye (Reproduced from Reference [11] with permission from American
Chemical Society)

Un-stretched state Stretched state

nano-clay

Sulfur bridges NR crystal

Fig. 23 Schematic model of deformation state of nano-clay filled natural rubber (Reproduced
from Reference [11] with permission from American Chemical Society)

these highly anisotropic nano-clay particles are completely aligned along the
deformation direction forming a physical network. This physical network would
favor the alignment of the NR chains resulting in the increase of crystallization
rate.
Deformation-Induced Structure Changes 153

The presence of interfacial adhesion between nano-particles and rubber matrix


could induce an early promotion and enhancement of overall crystallization of NR
chains under uniaxial stretching.
Based on the above results, a schematic model of the deformation process in
nano-clay filled NR is shown in Fig. 23.

7 Conclusions

Based on recent studies, the inclusion of nano-fillers in the elastomeric matrix can
effectively generate another level of physical crosslinking network. Under defor-
mation, the presence of nanofiller network can accelerate the orientation of sur-
rounding polymer chains and increase the strain-induced crystallization behavior;
it can also enhance the stress transfer efficiency of the composite materials.
Without question, the role of nanofillers to enhance the mechanical and physical
properties of elastomers will become increasingly important as more versatile and
functional nanofillers are generated in the future.

Acknowledgements The financial support of this work was provided by the National Science
Foundation (DMR-0405432) and Bridgestone.

References

1. Karino, T., Ikeda, Y., Yasuda, Y., Kohjiya, S., Shibayama, M.: Non-uniformity in natural
rubber as revealed by small-angle neutron scattering, small-angle X-ray scattering, and
atomic force microscopy. Biomacromolecules 8, 693 (2007)
2. Toki, S., Hsiao, S.H., Amnuaypornsri, S., Sakdapipanich, J., Tanaka, Y.: Multi-scaled
microstructures in natural rubber characterized by synchrotron X-ray scattering and optical
microscopy. J. Polym. Sci. Polym. Phys. 46, 2456 (2008)
3. Toki, S., Hsiao, B.S., Amnuaypornsri, S., Sakdapipanich, J.: New insights into the
relationship between network structure and strain-induced crystallization in un-vulcanized
and vulcanized natural rubber by synchrotron X-ray diffraction. Polymer 50, 2142–2148
(2009)
4. Toki, S., Minouchi, N., Sics, I., Hsiao, B.S., Kohjiya, S.: Synchrotron X-ray scattering:
tensile strength and strain-induced crystallization in carbon black filled natural rubber.
Kautschuk Gummi Kunststoffe 61, 85 (2008)
5. Tosaka, M., Kawakami, D., Senoo, K., Kohjiya, S., Ikeda, Y., Toki, S., Hsiao, B.S.:
Crystallization and stress relaxation in highly stretched samples of natural rubber and its
synthetic analogue. Macromolecules 39, 5100–5105 (2006)
6. Bokobza, L.: Multiwall carbon nanotube elastomeric composites: a review. Polymer 48,
4907–4920 (2007)
7. Atieh, M.A., Girun, N., Mahdi, E., Tahir, H., Guan, C.T., Alkhatib, M.F., Ahmadun, F.R.,
Baik, D.R.: Fullerenes. Nanotubes Carbon Nanostruct. 14, 641–649 (2006)
8. Koerner, H., Price, G., Pearce, N.A., Alexander, M., Vaia, R.A.: Remotely actuated polymer
nanocomposites—stress-recovery of carbon-nanotube-filled thermoplastic elastomers. Nat.
Mater. 3, 115–120 (2004)
154 S. Toki and B. S. Hsiao

9. Kelarakis, A., Yoon, K., Sics, I., Somani, R.H., Hsiao, B.S., Chu, B.: Uniaxial deformation of
an elastomer nanocomposite containing modified carbon nanofibers by in situ synchrotron X-
ray diffraction. Polymer 46, 5103–5117 (2005)
10. Carretero-Gonzalez, J., Verdejo, R., Toki, S., Hsiao, B.S., Giannelis, E.P.,
López-Manchado, M.A.: Real time crystallization of organoclay nanoparticle filled
natural rubber under stretching. Macromolecules 41, 2295–2298 (2008)
11. Carretero-Gonzalez, J., Retsos, H., Verdejo, R., Toki, S., Hsiao, B.S., Giannelis, E.P.,
Lopez-Manchado, M.A.: Effect of nanoclay on natural rubber microstructure.
Macromolecules 41, 6763–6772 (2008)
Thermally Stable and Flame Retardant
Elastomeric Nanocomposites

O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot

Abstract This chapter is dedicated to thermally stable and flame retardant elas-
tomeric composites. Two approaches are considered: the synthesis of elastomeric
nanocomposites, where the nanoparticles are dispersed at the nanoscale, and the
incorporation of nanofillers at high loadings where agglomerate of nanoparticles
are observed in the elastomeric matrix. The chapter is mainly focused on the key
parameter influencing the flame retardancy, that is to say the dispersion state, and
the ways to improve it by the modification of the matrix and/or of the nanofiller. A
particular attention is also paid to the creation of synergies between different types
of nanoparticles or between nanoparticles and conventional flame retardant
additives.

1 Introduction

Elastomeric materials are used in a plethora of application fields: packaging,


aerospace and transports, electrical devices, etc. Some of these fields are directly
concerned by fire regulations. Indeed, the inherent flammability of elastomers can

O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot


Univ Lille Nord de France, 59000 Lille, France
O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot
ENSCL, ISP-UMET, 59652 Villeneuve d’Ascq, France
O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot
USTL, ISP-UMET, 59655 Villeneuve d’Ascq, France
O. Cerin, G. Fontaine, S. Duquesne and S. Bourbigot (&)
CNRS, UMR 8207, 59652 Villeneuve d’Ascq, France
e-mail: serge.bourbigot@ensc-lille.fr

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 155


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_6,
 Springer-Verlag Berlin Heidelberg 2011
156 O. Cerin et al.

be the source of fire hazards. Important safety aspects in case of fire should permit
an early detection of fire and should find efficient means to fight it, but the best
way to guarantee the safety of people is to provide a non-flammable or at least a
material exhibiting low flammability.
Various methods can be used to protect elastomers or in general polymeric
materials from fire. It can be cited the graft-polymerization consisting in inserting
into the polymeric backbone flame-retardant monomers [1] or more simply the
blending of the matrix with a polymer exhibiting higher thermal resistance
properties [2]. The most commonly used approach to make thermally stable or
low-flammable materials is the incorporation in the polymer of flame retardant
particles, micro- or nano-dispersed in the matrix [3]. This method is generally
preferred to the others because it is easily compatible with industrial processes and
because it offers a good compromise between economical considerations,
mechanical, thermal and fire properties. Moreover, there is a growing interest upon
nanoparticles because they can be used in small quantities (a few percent) pro-
viding the same performance and thus limiting the degradation of the mechanical
properties of the resulting material.
This chapter is organized in five parts. First the basics of flame retardancy will
be commented in order to provide the reader the flame retardancy strategies, and in
particular the mode of action of nanoparticles. In the second part the fire behaviour
of the different elastomer classes and the ways to improve it will be investigated.
The distinction between the nanocomposites, were a real nanodispersion is
achieved, and the elastomers containing nanoparticles will be highlighted. The
flame retardancy induced by nanoparticles in elastomeric nanocomposites will
then be presented, and in another part the role of the incorporation of nanofillers in
elastomeric matrices in terms of reaction to fire will also be commented. The
potential future advances for designing flame retardant elastomers will be dis-
cussed in the last part.

2 Basics of Flame Retardancy

To understand the various strategies to make flame retardant materials, a synthetic


overview of the basics of flame retardancy will be presented in the following.
At first, why is a polymer flammable? When it is exposed to heat or flame,
during a fire scenario for example, the temperature of a polymer increases, leading
to its thermal degradation. Thus chemical bonds of the polymeric chains are
broken to generate highly flammable volatiles. These volatile compounds spon-
taneously form combustible mixtures with air which ignite easily and burn with a
high velocity. So, a low thermal stability associated to the release of highly
flammable volatile molecules is responsible for the flammability of the material.
Second, how is it possible to decrease flammability? Two major ways can be
followed: influencing physically the combustion process, or leading to a chemical
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 157

action in the condensed and/or in the gas phase. Among the processes dedicated to
the condensed phase can be cited:
• the artificial cooling of the matrix, by the addition of fillers which decomposes
endothermically during combustion (metal hydroxides, carbonates, etc.), thus
delaying the polymer degradation [4];
• the creation of a protective layer. The additives can form during combustion a
shield with a low thermal conductivity which can reduce heat transfer. It then
reduces the degradation rate of the polymer and decreases the release rate of the
pyrolysis gases. This is the basic principle of the intumescence process.
Intumescence is defined by the creation during combustion of an expanded
foamed cellular charred layer on the surface which protects the underlying
material slowing down heat and mass transfer [5].
The gas phase modifications concern:
• the dilution of the flame-feeding gases by non-flammable ones. For example the
use of metal hydroxides releasing water, or ammonium compounds generating
ammonia [6];
• the flame inhibition. The radical mechanism of the combustion process occur-
ring in the gas phase is interrupted by the radicals generated by the flame
retardants. Halogenated and some phosphorus-based compounds are involved in
such reactions.
The various ways presented above give an overview of the flame retardant
principles. Nevertheless the described processes generally not occur singly but are
part of a complex process in which many modifications occur simultaneously with
one dominating.
Third, what are the common flame retardant nanoparticles and what is their
influence on the flame retardancy of nanocomposites?
Some categories of nano-additives are widely used to enhance the polymers
properties, among them the thermal stability and/or the flame retardancy. The most
common ones are layered silicates (clays) and silica, metal hydrates and oxides
(Fe2O3, TiO2), calcium carbonates and graphene-based nanofillers. More exoti-
cally can be used polyhedral silsesquioxane (POSS) or layered double hydroxides
LDH. The mechanisms of action of the nanoparticles in elastomeric composites
are commented in the following.
Clay minerals are part of the larger class of silicate minerals. Included in this
layered silicates family are the natural montmorillonites (MMT). Undoubtedly
montmorillonite clay is the most commonly used silicate for producing thermal
resistant and flame retardant nanocomposites. According to Gilman et al. [7], the
flame retardant mechanism of clay based nanocomposites is a build-up of a high
performance carbonaceous silicate char on the surface during burning. Two
hypotheses can explain how MMT-rich char surface is produced [8]. There is an
assumption that the MMT is precipitated during pyrolysis as a result of progressive
gasification of the polymer, but another approach suggested a migration of MMT
during and after decomposition of the nanocomposite structure due to the lower
158 O. Cerin et al.

surface energy of MMT [9]. This layer insulates the underlying material and slows
the mass loss rate of decomposition products.
Recently Leszczynska et al. [10] studied the mechanism that controls thermal
stability of polymer/MMT nanocomposites. According to these authors, there are
several factors that influence thermal stability, among which:
• a labyrinth effect induced by intercalated or exfoliated structure of MMT, which
limits oxygen diffusion inside the nanocomposite sample,
• a steric effect, due to the interactions between the polar MMT layers and the
polymer matrix thus limiting the polymer chain motion,
• a barrier effect, which protects the bulk of sample from heat, decreases the rate
of mass loss during thermal degradation of polymer nanocomposite and gen-
erates more intensive char formation on the surface,
• a catalytic effect of the clay effectively promoting char-forming reactions.
Apart from the layered silicates, another class of nanofillers is able to impart
thermal stability or flame resistance through the creation of a physical protective
barrier: the graphene-based nanoparticles such as graphite, carbon nanotubes and
carbon nanofibres. Carbon nanotubes can be seen as elongated fullerenes: they are
graphene sheets rolled up into a hollow cylinder, with each end capped with half of
a fullerene molecule. There are two major types of carbon nanotubes: single-
walled carbon nanotubes (SWNTs) and multi-walled carbon nanotubes (MWNTs).
Carbon nanofibres CNFs are also composed of graphene particles, stacked together
and forming fibre-like structures. These particles act principally in forming a
charred residue, playing a protective role during combustion, as shown in
poly(methylmethacrylate) by Kashiwagi et al. [11]. Concerning graphite nano-
platelets, two types can be mentioned: natural graphite (NG), and expanded
graphite (EG). The effective method of preparing the expanded graphite is by
rapidly heating the pre-treated natural graphite to a high temperature, which
separates the platelets and also to some extent destroys the crystalline structure.
This treatment generally allows a better dispersion of the platelets. Their mode of
action is quite similar to that of carbon nanotubes [12].
The most popular fillers among the nano-metal hydrates are nano-aluminium
trihydrate (ATH) Al(OH)3 and nano-magnesium dihydroxide (MDH) Mg(OH)2
because of their high level of flame retardant performance associated to a low cost.
Nevertheless a relative high amount of filler (typically 60 wt% in thermoplastics)
is needed to achieve sufficient flame retardant properties [13]. Magnesium dihy-
droxide Mg(OH)2 acts in decomposing endothermically at relative low tempera-
ture (353C) into magnesium oxide and water, thus releasing non combustible gas
(water steam). This process dilutes the concentration of any other gaseous products
and also decreases the concentration of fuel available for combustion. The
decomposition also generates the oxide residue MgO, creating a protective barrier
that has relatively high heat capacity, reducing the amount of thermal energy
available to further degrade the substrate. Acting in a similar way as MDH, ATH
releases energy from the flames by decomposing into aluminium oxide and water,
and creates an alumina barrier. Another particle exhibits comparable behaviour:
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 159

the nano-layered double hydroxides (LDH). Its structure is based on brucite


(Mg(OH)2)-like layers. The flame retardation mechanism of LDH has been studied
in some matrices. According to Jiao et al. [14] they act in EVA as a combination of
aluminium trioxide and magnesium dihydroxide. LDH decomposes and absorbs
the heat in different temperature range. However this interpretation of the mode of
action of LDH is probably not complete: according to Zhang et al. [15]
nano-LDHs have a capacity to catalytically oxidize a char-rich compound in the
presence of O2; they can accelerate the carbonization of the char-rich compound to
obtain a more sustainable char, improve the graphitization degree of the formed
char, and further encourage the formation of an intumescent and heat-insulating
carbonaceous layer to some extent. This is confirmed by the work of Zammarano
et al. [16, 17] who highlighted that some modifiers (i.e. sulfonate anions) catalyze
charring reactions in LDH based nanocomposites during thermal degradation,
enhancing the formation of a carbonaceous char and decreasing the release rate of
combustible volatiles.
We have seen that nanocomposites exhibit similar mechanisms of flame re-
tardancy, that is to say the creation of a protective layer hindering the combustion
of polymer. Three parameters are considered crucial to achieve good flame re-
tardancy: the nanodispersion state, the resistance and the formation rate of the
protective barrier and the viscosity of the melt [18]. Particular attention has to be
paid to the morphology of nanocomposites. Indeed, the filler must be nanodi-
spersed to talk about nanocomposites, and this dispersion is a key parameter of
flame retardancy.
Finally, which parameters describe fire behaviour or thermal stability? The
thermal stability of materials is generally characterized by thermogravimetric
analyses (TGA). The mass loss of the material is recorded as a function of a
temperature ramp. Two types of thermal degradation can be implemented: non-
oxidative decomposition, under nitrogen or another inert gas, and the oxidative
decomposition under air.
Fire behaviour can be described through three major parameters: the ignit-
ability, the contribution to flame spread and the heat release. Moreover side effects
can be observed, the most important one is the emission of smoke during
combustion.
Depending on the material application fields some specific tests can be
implemented to simulate the desired conditions, for example the Federal Aviation
Administration (FAA) seat cushion flammability test, which characterizes the
involvement of aircraft seat cushions in cabin fires. More generally, three tests are
usually used since each of them quantifies the ignitability, the contribution to flame
spread and/or the heat release. These tests are the Limiting Oxygen Index (LOI),
the UL-94 test and the cone calorimeter test. LOI is determined according to ISO
4589 [19] and ASTM D2863 [20]. It consists in determining the concentration of
oxygen which will just support combustion. Another small-scaled test described in
UL-94 [21] determines the tendency of a sample either to extinguish or to spread
the flame once the specimen has been ignited. There are twelve classifications
specified in UL-94 that are assigned to materials based on their behaviour
160 O. Cerin et al.

regarding burning, flame propagation and dripping. The most common ones are
V-0 to V-2 classifications (V-0 considered the best) for the vertical burning test
and HB for the horizontal one. Cone calorimeter tests [22, 23] allow the simulation
of the conditions of a fire in a small bench scale test. The measured parameters are
among others the Rate of Heat Release (RHR), the Total Heat Release (THR) and
the Time To Ignition (TTI). These parameters permit to evaluate the contribution
to a fire of the material (reaction to fire).

3 Fire Behaviour of Elastomers

Depending on their own thermal stability and reaction to fire, the different elas-
tomers do not need the same level of protection against fire.
There exist various types of elastomers. Can be cited the hydrocarbon rubbers,
and among them, some thermoplastic and polyolefin elastomers. The thermoplastic
rubbers include polyisoprene/natural rubber (NR), thermoplastic polyurethane
(TPU) and derived flexible foams, ethylene-vinylacetate copolymer (EVA),
polybutadiene (BR), block copolymers such as styrene-butadiene-styrene (SBS).
The thermoplastic polyolefins assign poly(octene-ethylene) elastomer (POE) and
the ethylene-propylene-diene monomer (EPDM). All these polymers present a
high flammability (LOI of 17 vol% for NR [24] and 18 vol% for SBS as an
example) as they thermally decompose into smaller hydrocarbon molecules, these
molecules forming high fuel-value gases. As an example, PU foams are known to
be great contributors to fire hazard [25].
Other types of elastomers, the inorganic or semi-inorganic rubbers, have the
advantage to present intrinsically high thermal resistance and low flammability.
Indeed, replacing some hydrocarbon part by an inorganic one results in a polymer
with higher heat resistance. Compared to hydrocarbons, the high inorganic content
of these elastomers have lower fuel value, thus reducing the flammability. These
polymers are fluoroelastomers, silicones and chloroprene. Silicones, in particular
polydimethylsiloxanes (PDMS) exhibit an intrinsically flame retardant behaviour
with a peak of RHR (PRHR) comprised between 60 and 150 kW/m2 [26]. Lyon
et al. [27] reported the development of some brand-new materials: polysilphen-
ylene-siloxane and polyphosphazene elastomers, specially designed to present
intrinsically high fire resistance (Fig. 1).

Fig. 1 Chemical structure


of polysilphenylene-siloxane
and polyphosphazene
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 161

Polysilphenylene-siloxanes are ultra thermally stable elastomers. Compared to


PDMS their thermal decomposition occurs 200C later, that is to say beyond
500C. Concerning polyphosphazenes, their high nitrogen and phosphorus content
is responsible for their flame retardant properties. Indeed, the cone calorimeter
tests done at an external heat flux of 50 kW/m2 revealed the heat release rate of a
polyphosphazene is four times lower that current aircraft polyurethane elastomer.
The better flame retardancy is attributed to a gas phase mechanism, where phos-
phorus compounds in a low oxidation state inhibits the radicals present in the
flame. However, higher smoke production (five times more) is induced by the
presence of phosphorus.
The flammability of an elastomer depends thus on its chemical composition. So,
a way to modify its combustion behaviour would be to chemically modify it.
Different strategies are reported here: the design of new thermally stable and flame
retardant elastomers (such as polyphosphazene) and the modifications of the
polymer backbone by grafting.
In their review, Levchik et al. [28] report the modifications of the structure of
thermoplastic polyurethane (TPU) leading to better flame retardant properties. Can
be cited the incorporation of a phosphorus-based molecules in the pendant groups
[29] or in skeleton [30]. Polyurethanes with the phosphorus in the pendant groups
were prepared by N-alkylation of a linear PU. LOI data showed that the fire
resistance of TPU is slightly enhanced. To incorporate phosphorus in the skeleton
some phosphorus-based monomers were copolymerized with 2-hydroxyethyl-
methacrylate to form a hydroxy-containing copolymer, used as a polyol in the
synthesis of PU. The so-modified PUs were characterized through TGA, which
revealed their earlier degradation temperature but concurrently their higher char
yield.
Najafi-Mohajeri et al. [31] synthesized a series of ferrocene-modified poly-
urethane block copolymers and evaluated their thermal stability and reaction to fire
through TGA, LOI and cone calorimetry. It was shown that a small amount of
ferrocene reduced the PHRR by 40–80% depending on the ferrocene type and
content, but the LOI was not significantly modified. The interest of the incorpo-
ration in the backbone of the flame retardant is also demonstrated, since only
0.51% ferrocene in the backbone structure reduce the PRHR by 60% whereas PU
containing ferrocene at 15% as an additive showed a reduction of only 54%
compared to unmodified PU.
Levchik et al. [28] also report the copolymerization of TPU and silicones. An
amine-functionalized siloxane incorporated at 5 wt% in the TPU chain allows a
PRHR reduction of 79% compared to the virgin polymer. In comparison a non-
functionalized siloxane used as an additive at 5 wt% leads to a PRHR decrease of
70%.
The chemical modification of the polymer backbone allows the synthesis of
thermally stable and/or flame resistant elastomers. Up to now, the examples pre-
sented in this chapter concerned raw polymers, with no mention to nanocom-
posites. However, it is noteworthy that the ways described in this part could be
perfectly applicable to elastomeric nanocomposites, since the decreased
162 O. Cerin et al.

flammability of the matrix imparts better flame resistance to the whole material.
Moreover, nanoparticles themselves can provide some specific fire properties to
elastomer nanocomposites.

4 Flame Retardant Nanocomposites

Very few scientific literature deals with elastomeric nanocomposites and flame
retardancy. Indeed, up to now, only 17 articles and 9 patents were published on
this subject (Database SciFinder, keywords ‘‘elastomeric nanocomposites’’ and
‘‘flame retardancy’’, March 2010). Moreover, the aspect of nanodispersion when
investigating flame retardancy is not always commented and is often assumed
when nanoparticles are incorporated in polymeric matrices without further char-
acterization. Since we know it is a primordial aspect, we will make the distinction
between the nanocomposites, were a real nanodispersion is achieved, and the
elastomers containing nanoparticles.
The following part focuses on nanodispersed flame retardant particles in elas-
tomers, and on the parameters controlling this dispersion.

4.1 Nanofillers in Nanocomposites

Some nanofillers can induce the formation of a residue during degradation, this
residue acting as flame retardant by a barrier effect. The two major additive
families playing this role in elastomers are the layered silicates and the graphene-
based additives.

4.1.1 Layered Silicates

The efficiency of clays is linked to their structure modifications. By introducing


trivalent aluminium and iron ions into the clay silica sheets, or bivalent magnesium
and iron ions into the central sheet, the sheet package is negatively charged, charge
which is compensated by the presence of cations such as Na+ (for raw clay)
between the individual sheet packages. These cations can be easily replaced by
others having a stronger binding affinity with polymer matrices. Organic cations
are used in order to improve the compatibility between the organic polymer matrix
and the montmorillonite layers. Montmorillonite containing these organic cations
are called organo-modified montmorillonite.
Some of these organo-modified clays were dispersed at 2.5 wt% at different
degrees in elastomeric PU by in-situ polymerization [32]. The organic cation used
as a modifier in the clay structure is methylbis-2-hydroxyethyltallow alkyl qua-
ternary ammonium, presented in Fig. 2.
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 163

Fig. 2 Quaternary ammo-


nium salt used as a modifier
in montmorillonite

The investigation of the fire behaviour of the nanocomposite was carried out
through UL-94 test (3.2 mm thickness). The clay nanocomposite allows the sup-
pression of the dripping phenomenon during the test. Char develops quickly upon
ignition; both the char layer and reduced mobility of polymer chains prevent the
flow and dripping caused by polymer melting. Nevertheless, the nanocomposite
remains non-classified.
Montmorillonite clays (organo-modified or not) through their action on the
rheology of the nanocomposites towards temperature and their char-forming cat-
alytic effect, allow the increase of the thermal stability of the materials and
sometimes of their flame retardant properties. Some other nanoparticles, among
the layered silicates, can be proposed as potential flame retardants as they are
supposed to provide similar properties. Can be cited talc, mica (known to increase
thermal stability of PDMS [33]) or kaolin.

4.1.2 Graphene-based Nanofillers

The incorporation of multi-walled carbon nanotubes was evaluated in ethylene-


octene copolymer POE [34]. Thermogravimetric analyses revealed that thermal
decomposition of the polymer matrix was retarded in the MWNTs nanocompos-
ites. This result may be attributed to a physical barrier effect, resulting from the
fact that MWNTs would prevent the transport of decomposition products of the
polymer nanocomposites.
The use of such nanoparticles was also reported in elastomeric EVA (40% vinyl
acetate) [35]. Multi-walled carbon nanotubes MWNT and carbon nanofibres CNF
at 4 wt% were incorporated by solution blending. The thermogravimetric analyses
of the obtained materials revealed a heat stability increase with MWNT (decreased
degradation rate and higher maximum degradation temperature) as shown in
Table 1.
When compared to montmorillonite or LDH, MWNT prove to be more efficient
in terms of reduction of peak of RHR, as shown in Table 2. The reduction reaches
66% while the sample with clay shows a reduction of only 49%. In the residual
char structure produced by the combustion, an integrated structure with surface

Table 1 Thermal
Sample Tmax (C) Maximum rate of
degradation data of various
degradation (%/C)
EVA40 nanocomposites at
4 wt% filler loading (20C/ EVA40 453 2.24
min under nitrogen) [35] EVA40–4MWNT 460 1.37
EVA40–4CNF 456 2.39
164 O. Cerin et al.

Table 2 Cone calorimeter


Sample PRHR (kW/m2) Reduction (%)
data of various EVA19
nanocomposites at 3 wt% EVA19 1772 ± 170 –
filler loading (35 kW/m2) [36] EVA19–3MWNT 597 ± 30 66
EVA19–3LDH 1090 ± 58 39
EVA19–3MMT 903 ± 24 49

cracks was formed from the clay/EVA nanocomposites while only limited char
fragments were obtained in the composite containing the nanotubes. These
structures do not appear to support the results on the peak of heat release rate if the
fire retardant action of the nanotubes occurs in the condensed phase.

4.1.3 Combinations

The use of the cited nanoadditives in elastomer nanocomposites proved to have a


positive effect in terms of thermal resistance and flame retardancy. Nevertheless, it
seems to be possible to combine the properties of the nanoparticles to obtain more
resistant composites. That is why many researches are dedicated to potential
synergistic effects between nanoparticles.
A promising combination in terms of flame retardancy is the clay-carbon
nanofiller mixture. Thermoplastic polyurethane elastomer was modified with dif-
ferent loadings of montmorillonite nanoclays and carbon nanofibers (CNFs) [37].
It was found that thermoplastic polyurethane elastomer with 10 wt% CNFs and
with 5 wt% nanoclays gave the best thermal performance with respect to pro-
tecting a substrate. Indeed, the combination of the two nanoparticles resulted in the
formation of a char layer, thus increasing the thermal insulative properties of the
material.
The same combination was reported in the literature in EVA [38, 39].
It appeared that the PRHR reduction is slightly increased when the two particles
are nanodispersed in the matrix, but the aspect of the remaining char is radically
different. The char produced by the clay-composite showed cracks on its surface,
while the combination provides a smooth-surfaced carbonaceous shield.
The roles of composites have been studied by Gao et al. It has been found that
nanotubes may act as nucleation agents for graphitisation, leading to the formation
of turbostratic and graphitic carbons. The formation of graphitic carbon in char
may contribute directly to the reduction of the peak of heat release rate of the
composites. This effect has been enhanced when both nanotubes and clay are used,
since clay enhanced nanocomposites appear to have better resistance to char
oxidation. The nanotubes also have the function to reduce surface cracks of chars,
leading to the increase of barrier resistance to the evolution of flammable volatiles
and to limit the oxygen transfer to the condensed phase.
The use of carbon nanotubes with MMT in nanocomposites seems of interest as
MWNT can provide mechanical properties to the char produced by the composite
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 165

during combustion. That leads to an increase of the shield-playing role of the char
and also of the fire resisting properties of the material.

4.2 Influence of the Dispersion

The efficiency of the nanoadditives in polymeric materials is directly linked to


their dispersion state. The presence of tactoïds or the total exfoliation of the
particles influences the thermal or fire behaviour of nanocomposites, as reported
by George et al. [35, 40] about nano-graphite dispersed in EVA60 (60% vinyl
acetate). Transmission electron microscopy (TEM) pictures revealed that the
natural graphite (NG) particles formed agglomerates inside the rubber matrix at
4phr as well as 8phr, on the contrary to expanded graphite (EG) platelets which
were homogeneously distributed in the EVA matrix (Fig. 3). The given explana-
tion was that the greater surface area of the expanded graphite can lead to higher
interaction between the platelets and the macromolecules.
These results were correlated to the measured thermal properties: the presence
of NG adversely affected most of the properties of EVA while the incorporation of
expanded graphite provided improvement in mechanical properties, thermal con-
ductivity and thermal stability of EVA. A shift of 14C in maximum rate of
degradation was observed for 4phr EG addition.
The main parameter responsible for the dispersion quality in the composites is
the compatibility of the nanofillers with the polymer matrix. So, to improve the
dispersion of the nanoadditives it is important to increase this compatibility by
modifying the filler or the matrix.

4.2.1 Compatibility with the Polymeric Matrix: Influence of the Surfactant

The influence of the compatibility of EVA and different types of montmorillonite


was investigated by Duquesne et al. [41]. They showed that the fire performance of

Fig. 3 TEM pictures of EVA60 containing a 4phr EG, b 8phr EG, c 4phr NG [Reproduced from
Ref. 40]
166 O. Cerin et al.

Fig. 4 SEM pictures of the EVA-Cloisite composites [Reproduced from Ref. 41]

EVA/C30B (Cloisite 30B is a MMT, organomodified by the incorporation of an


alkyl ammonium ion) is very good at low loading (50% PRHR decrease at 5 wt%
loading) whereas in the case of EVA/Na+ (pristine MMT), the improvement is
lower (20% decrease for the same loading). In a second part, it has been dem-
onstrated that the Cloisite Na+ is less compatible with the polymer because the ion
is hydrophilic.
On the contrary, the compatibility of Cloisite 30B in EVA is improved by the
presence of the ammonium ions between the clay-layers. So, the structure of the
composite obtained with Cloisite 30B is a nanocomposite-type structure whereas
in the case of Cloisite Na+, a micro-structure is obtained. This assumption has been
checked by small-angle X-ray diffraction analysis and by SEM analysis (Fig. 4).
As a consequence, it was possible to conclude that a nano-structure enables to
achieve better fire performance than a micro-structure. In fact, the presumed
‘‘diffusion effect’’ which leads to such an improvement, occurs in a nano-structure
but not in a micro-composite.
The work of Tang et al. [42] confirms the previous results. EVA/clay nano-
composites have been synthesized by melt intercalation from pristine MMT by
adding compatibilizer C16 (hexadecyltrimethylammonium bromide). Figures 5
and 6 indicate that the clay compatibilizers have influence on the HRR. The initial

Fig. 5 Heat Release Rate for


pure EVA and different
MMT composites (MMTa:
nanocomposite, MMTb: mi-
crocomposite) [Reproduced
from Ref. 42]
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 167

Fig. 6 Heat Release Rate for


pure EVA and EVA/MMT
nanocomposites [Reproduced
from Ref. 42]

HRR for the nanocomposite is higher at the beginning of combustion probably


because of decomposition of the organic alky ammonium cation (C16) and of
accelerated evolution of acetic acid owing to catalytic effect of acidic sites of the
layered silicates.
The modification of the surfactant of the clay layers is an efficient way to
increase MMT dispersion and thus thermal and flame retardant properties. Another
possibility to achieve better dispersion of additive having low compatibility with
the elastomer is to modify the polymer matrix by grafting.

4.2.2 Use of a Compatibilizer: Grafting

Since it does not include any polar group in its backbone, it is difficult to make
intercalation of POE chains into clay layers. According to the literature, incor-
poration of polymer functionalized with maleic anhydride or hydroxyl groups as
compatibilizer has been proven as a successful way to facilitate interactions
between these two dissimilar components [9, 10]. This compatibilization was
tested between maleic anhydride grafted POE (POE-g-MAH) and organo-modified
montmorillonite to investigate the influence of MMT on the structure and flam-
mability properties of POE-based nanocomposites [43].
The grafting allows a better dispersion of the MMT layers: the tactoïds present
in the raw composite disappear in the grafted composite, where a complete ex-
foliation is achieved (Fig. 7).
The enhancement of the dispersion quality is accompanied by an increase of the
flame retardant properties. As shown in Table 3 the PRHR decrease of the grafted
(and thus exfoliated) nanocomposite is higher than that of the composite pre-
senting MMT tactoïds.
The use of maleic anhydride as a compatibilizer for MMT was also reported in
ethylene-propylene diene monomer EPDM [44]. Dispersion was characterized
through X-ray diffraction (XRD) analyses, which showed an enhancement of the
168 O. Cerin et al.

Fig. 7 TEM pictures of a POE/MMT and b POE-g-MAH/MMT nanocomposites [Reproduced


from Ref. 43]

Table 3 Cone calorimeter


Sample PRHR (kW/m2)
data of POE, grafted POE and
their nanocomposites POE 1,727
(35 kW/m2) [43] POE-g-MAH 1,650
POE/MMT 1,280
POE-g-MAH/MMT 793

platelets dispersion. Cone calorimeter was used to investigate the reaction to fire of
the sample. It indicated that the PRHR decreases from 1374 kW/m2 for pure
EPDM to 906 kW/m2 for the formulation with the compatibilizer whereas no
PRHR decrease occurs with the organomodified clay.
The compatibility of the nanoadditive in the polymeric matrix has a real impact
on the thermal and fire properties of the resulting composite, since it directly
influences the filler-polymer interactions and the dispersion quality. So, this
compatibility is key parameter to achieve the best flame retardant performances.
Thus, the synthesis of thermally stable or flame retardant elastomeric nanocom-
posites has to be designed to provide the best dispersion state, so as the
nanoparticles can fully express their flame retardant qualities.

5 Nanoparticles in Elastomers

Some nanoadditives are incorporated in polymeric matrices at relatively high


loadings (typically 60 wt% for nano-MDH). However there is an interrogation
concerning the relevance of incorporating nanoparticles at such loadings: making a
calculation of an ideal dispersion (nanoparticles evenly dispersed individually)
will reveal that the distance between 2 molecules is too low to avoid the
agglomeration of particles. So the term ‘‘nano’’ is not suitable in that case, except
for the elementary size of the particle.
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 169

However, we propose to enlarge the notion of ‘‘nanocomposite’’ to composites


filled with nanoparticles but where nanodispersion is not always achieved. That is
why this part of the chapter is dedicated to the elastomeric matrices, flame retarded
with nanoadditives but not nanodispersed.

5.1 Nanofillers in Elastomers

Three major types of nanoadditives are used at high loadings in elastomers: the
metal hydrates, the layered double hydroxides and calcium carbonate or sulphate.

5.1.1 Metal Hydrates and LDH

The flame retardancy and mechanical properties of EVA containing 18% vinyl
acetate (EVA18) containing Mg(OH)2 have been investigated by Huang et al.
with reference to the particle size of MDH [45]. At a filling level of 55 wt%,
there is little difference in mechanical properties and LOI values among four
composites varying the particle size of MDH. The fire retarded blends, what-
ever the nano or micro-size MDH exhibit equivalent fire retardant properties.
According to these results the expected effect of particle size does not exist.
But it could be explained by the poor dispersion of nano-MDH observed in
EVA compared with micro sized MDH since nano-size particles aggregate
easier than micro-size one.
At the opposite Lu et al. [46] with lamellar-like nano MDH noticed interesting
properties. If UL-94 is not modified with nanoparticles compared to the same
microparticles, the LOI is sharply increased (46 vol% for nano-MDH versus
39 vol% for micro-MDH at the same loading of 150phr). This might be explained
by the better dispersion of the nanoparticles compared with microparticles, even if
the presence of tactoids is detected by TEM analyses.
It appears that dispersion is an essential parameter determining the efficiency of
the flame retardant effect of MDH. Considering the need of mechanical properties
and flame retardancy of composites, it is suggested that the smaller sized MDH
should be selected as flame retardant, but the appropriate methods must be con-
ducted to improve the dispersion of smaller particles.
Concerning LDH, they have been tested in EVA with 14% vinylacetate
(EVA14) [14] and proved to be efficient, as a LOI value of 42 vol% and a V-0 UL-
94 rating are reached with a filler content of 150phr (Table 4).

5.1.2 Calcium Carbonate and Sulphate

Calcium carbonate mode of action implies the release of CO2 gases during
combustion, which cool down the polymer substrate and in the same time dilute
170 O. Cerin et al.

Table 4 Effect of nano-LDH


EVA–LDH (phr) UL-94 rating Dripping LOI (vol%)
concentration LOI and UL-94
rating of the EVA14–LDH 100–50 Fail Yes 27
blends (specimens 100–80 Fail Yes 32
117 9 12.7 9 3 mm) [14] 100–100 Fail No 37
100–120 V1 No 40
100–150 V-0 No 42

flammable gases. Therefore the heat feedback to the polymer substrate is reduced
and the emission of flammable gases decreases. Moreover, calcium carbonate
contributes to the constitution of a mineral or ceramic-like calcium silicate residue
after degradation which acts as flame retardant by a barrier effect [47]. Never-
theless this additive has to be incorporated at relatively high loadings to be effi-
cient in polymeric matrices.
Hamdani et al. [48] reported the use of calcium carbonate in a silicone matrix
as a flame retardant. Thermo gravimetric analyses (TGA) typically show
enhancement of thermal stability of silicone by addition of CaCO3 (30 wt%).
Indeed, PDMS degrades around 300C, whereas when mixed with calcium
carbonate, some residue remains in samples treated at 500C. The silicone
elastomer stabilized by the additive thus resists at higher temperatures in addition
to taking an active part in the formation of the protective structure. After
treatment at 500C, the intumescent structure only consists of calcium carbonate
and silicon oxides.
This additive is also known to be used in styrene-butadiene rubber SBR.
Nanosize-CaCO3-filled rubber showed a significant reduction in flammability in
comparison with that without filler [49] according to the ASTM D4804 (the
sample is clamped at a 45, a free end is exposed to a specified gas flame for 30 s,
then the time required for the burning and the relative rate of burning are mea-
sured). The rate of flame retarding of nano-CaCO3 was more than that of com-
mercial micrometric CaCO3. In this case, the rate of flame retarding for 9-nm
CaCO3 was 2.55 s/mm, whereas for commercial CaCO3 it was 1.78 s/mm. The
increase in flame retardancy with a decrease in the size was attributed to the
homogeneous dispersion of the nanofillers.
In the same matrix calcium sulphate CaSO4 at various loadings (between 2 and
12 wt%) and various particle size was compounded [50]. Thermal stability
increased slightly when the particle size decreased: the temperature at the maxi-
mum degradation rate increases from 440C for neat SBR to 446C for filler
loading of 4% (23 nm) or 460C at the same loading with a lower particle size
(10 nm). The enhancement of the thermal stability is attributed to the homoge-
neous dispersion of nano-CaSO4 throughout the matrix (confirmed by scanning
electron microscopy SEM). The nanoscale inorganic filler in the rubber nano-
composites also promotes the formation of the char layer, which acts as an
excellent insulator and mass-transfer barrier, because of which this effect drasti-
cally improves the burning resistance.
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 171

The particle size of the nanoadditives is also a crucial factor influencing


the thermal stability or the reaction to fire of a material. Indeed, the specific
surface increases with decrease particle diameter, leading to higher surface
contact with the polymeric chains and thus to a modification of the composite
behaviour.
The effect of the particle size of CaSO4 was investigated in styrene-butadiene
rubber [50]. The particle diameter was 10 nm, 15 nm, 23 nm or 20 lm. The
thermal stability was evaluated through TGA measurements, which show an
increase of the thermal stability with the decreasing particle size. The enhance-
ment of the thermal stability is due to the uniform dispersion of nano-CaSO4
throughout the matrix, and the increased specific surface of the particle allow
larger interactions with the polymeric chains, keeping them intact at higher tem-
perature. The influence of the particle diameter also concerns the flammability of
the material. The test ASTM D4804 reveals a lower flammability of the nano-
composites compared to the microcomposite, and this effect is more pronounced
with low particle diameters. This is explained by the char-promoting action of the
nanoscale inorganic additive in the rubber, which catalytic effect is higher with
high specific area.
The use of nano-MDH, LDH or nano-sulphate or carbonate in elastomers has
been reported in the literature. These additives proved to be efficient in terms of
LOI, mass loss calorimetry and UL-94 test, even if very high loadings (about 60%
or more) are still needed to meet an adequate level of flame-retardant property.
However, the nanoparticles are still detrimental to the mechanical properties. So,
to develop their use in elastomers, it seems to be necessary to find synergistic
systems allowing the decrease of the fillers content without decreasing the flame
retardant properties.

5.2 Synergies and Use with Conventional Fame Retardant


Additives

Combining the nanofillers is an attractive way to reduce the overall fillers content
in the formulation while achieving an acceptable level of performance, by the
creation of synergies. As an example combinations of metal hydroxides with
organomodified clay have been studied as a potential fire retardant system for
polypropylene [51]. The combination of polypropylene with 20 wt% metal
hydroxide and 5 wt% clay gives a reduction in the PRHR similar to that obtained
with 40 wt% metal hydroxide.
The use of nanofillers as synergists with other nanoadditives enables the cre-
ation of flame retardant elastomeric nanocomposites. Another way to enhance the
thermal stability or flame resistance of such materials is the use of conventional
flame retardant additives, whose action can be reinforced by appropriate
nanosynergists.
172 O. Cerin et al.

5.2.1 Combinations with Conventional Flame Retardants

The incorporation of nanofillers to improve the flame retardancy of elastomers,


flame retarded with conventional additives, is reported in the scientific literature
[52]. In the case of elastomers the main combinations involve metal hydroxide and
nanoadditives, such as clays, inorganic whiskers, carbon nanotubes and fumed
silica.
The decrease of the overall filler content is illustrated by the work of Beyer [53]
about ATH, clays and carbon nanotube combinations. The RHR is maintained at
200 kW/m2 by the incorporation of only 5 wt% nanofiller while the ATH content
is decreased from 65 to 45 wt%.
Szep et al. [52] added modified (IMM = polybuthene/polysiloxane intercalated
MMT) and unmodified montmorillonite (MMT) to EVA/MDH formulations. It
appeared that the addition of MMT and IMM was detrimental to the UL-94 rating
and slightly beneficial for LOI. The simultaneous use of modified and non-mod-
ified montmorillonite in EVA with magnesium hydroxide increases the effec-
tiveness of flame retardancy (Table 5). The synergistic effect between MDH and
MMT ? IMM was explained by the increased and sustained viscosity, by the early
pre-carbonisation before the beginning of intense degradation and by the formation
of a tough, stable ceramic-like residue from the montmorillonites and metal
hydroxide interaction in the combustion process.
Magnesium hydroxide sulphate hydrate whiskers/silicone rubber (SR) nano-
composites, were flame retarded with microencapsulated red phosphorus (MRP)
used as a synergist [54]. It is found that whiskers can effectively improve the flame
retardancy of SR composites due to endothermic degradation of whiskers with the
release of water vapour diluting fuel supplied in the flame [55], accompanied by
the formation of fibrous magnesia acting as a barrier layer. The incorporation of
MRP in the system had a synergistic flame retardant effect in the condensed phase:
the water released from whiskers promotes the formation of phosphoric acid from
MRP, forming a protective phosphorus glassy layer [56]. A radical gas phase
action is also suspected [57].
Silica particles can also be used as synergists in combination with metal
hydroxides. The influence of fumed silica of different particle size on the fire
retardant properties of EVA-ATH-MMT-silica blends was investigated by Laoutid
et al. [58]. The partial substitution of ATH by 5 wt% of silica in an EVA/ATH/
MMT system did not improve significantly the overall fire behaviour. The pres-
ence of silica reduces the PRHR, especially for silica particles of small size and

Table 5 LOI and UL-94 data of EVA28 compounds containing MMT and MMT–MDH [52]
Material Weight ratio LOI UL-94
EVA ? MDH 1:1 33 V-1
EVAMDH ? MMT 8:8:1 34 V-2
EVA ? MDH ? IMM 8:8:1 36 V-2
EVA ? MDH ? IMM ? MMT 16:16:1: 2 43 V-0
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 173

large specific surface area, but reduces the appearance time of the second PRHR,
due to the poor cohesion of the char formed with silica. The addition of silica in
EVA-ATH-oMMT blends seems not beneficial regarding to fire resistance prop-
erties as it decreases char cohesion during combustion
Flame-retardant methyl vinyl silicone rubber (MVMQ)/montmorillonite nano-
composites were prepared, using magnesium hydroxide (MH) and red phosphorus
(RP) as synergistic flame-retardant additives, and aero silica (SiO2) as synergistic
reinforcement filler [59]. In addition to LOI test, thermal properties were evaluated
by thermogravimetric analysis (TGA). The decomposition temperature of the
nanocomposite can be higher than that of MVMQ as basal polymer matrix. This
kind of silicone rubber nanocomposite is considered by the authors a promising
flame-retardant polymeric material.

5.2.2 Synergistic Combinations of Nanofillers

Combinations with flame retardant additives are conventionally used to improve


the flame retardancy of elastomers while decreasing flame retardant loading. In
this part we will present the use of some exotic nanofiller which can be combined
with other nanoadditives: the hydroxyl aluminium oxalate (HAO).
HAO was tested as a synergist in clay/LDPE/EPDM composites [60]. Nano-
HAO (presented in Fig. 8) acts as heat remover and flammable gas diluter with the
release of water and carbon dioxide after decomposing, while MMT mainly acts as
heat and gas barrier, so the ternary composition has the aim to combine the two
flame retarding properties, thus exhibiting superior flame resistance.
It was found that the addition of MMT can produce synergistic effect on flame-
retarding nano-HAO/LDPE/EPDM system. The substitution of HAO by various
amounts of clay leads to an increase of the LOI values. When the ratio of MMT
and nano-HAO was 1:3, the LOI of the composites was up to 34 vol%, whereas
HAO alone in the polymer matrix exhibits only a LOI of 31 vol%. Through the
analysis of TGA, infra-red spectroscopy (FTIR) and SEM, it was proven that the
barrier effect of the char layer was promoted when MMT was added and thus the
decomposition of nano-HAO and polymer matrix under the char layer was
retarded.
The same approach was implemented with the combination of nano-kaolin and
nano-hydroxyl aluminium oxalate in LDPE/EPDM composites [61]. From the LOI
tests and UL94 tests, it was found that when 12 wt% nano-kaolin substituted nano-
HAO in the composites, the LOI was enhanced from 31.0 to 35.5 vol% (Fig. 9),
and the composites passed the UL94 V-0 standard, which proved that nano-kaolin

Fig. 8 Composition of nano-


HAO
174 O. Cerin et al.

Fig. 9 LOI values of the


nano kaolin/HAO/LDPE/
EPDM composites (60 wt%
filler content) [Reproduced
from Ref. 61]

had a synergistic effect with nano-HAO on flame retardancy in the LDPE/EPDM


system. Through the analysis of FTIR and SEM, it was found that the synergistic
effect of nano-kaolin and nano-HAO on flame retarding LDPE/EPDM composites
was attributed to the improved ability of rebuilding compact char barrier.
Montmorillonite and nano-kaolin are efficient synergists in combination with
nano-HAO. However, the very high nanoadditive loading (60 wt%) raises once
again the question of the interest of using nanoparticles instead of microadditives,
since the interest of nanofillers is to reduce the overall filler content without
decreasing the flame retardant performances.

6 Conclusion

In this chapter the design of thermally stable and/or flame retardant elastomeric
materials was discussed. Two approaches were considered: the synthesis of elas-
tomeric nanocomposites, where a real nano-dispersion is achieved, and the
incorporation of nanofillers at high loadings, where agglomerates of nanoparticles
are observed in the elastomeric matrix.
Elastomer nanocomposites achieve a satisfactory dispersion state are flame
retardant materials in specific conditions (fire testing involving mainly radiative
phenomena) but generally fail flammability tests (LOI, UL-94). The overview of
the nanoparticles mechanisms of action reveals that the formation of a protective
layer during combustion is always involved whatever the used nano-additive.
Moreover, the quality of the dispersion of the nanoparticles was found to be a key
parameter in terms of flame retardancy. Since the dispersion is directly linked to
the compatibility of the filler with the polymeric matrix, this aspect has to be
considered, with the modification of the matrix (by grafting for example) or with
the compatibilization of the filler (by coating or use of adapted surfactants).
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 175

Moreover a particular attention should be paid to the mode of preparation of the


nanocomposites, since it greatly influences the final dispersion state.
The second approach suggested in this chapter concerns the incorporation of
nanofillers in elastomers but at high loadings, thus preventing the nanodispersion.
The use of nano-additives compared to micrometric ones was found more efficient
in terms of flame retardancy or thermal stability, but some mechanical properties
of the elastomers might be negatively influenced by such additives. The best flame
retardant properties were obtained by the use of synergies between nanoparticles
and combinations with conventional flame retardants.
Promising developments are also expected in the creation of synergies, espe-
cially in nanocomposites. So, it is crucial to have a good understanding of the
nanofillers mode of action, and of the type of synergisms created by the combi-
nation of nano-additives. Research work has to be focused on this particular
technology as it provided nanocomposites exhibiting superior flame retardant
properties.

References

1. Wilkie, C.A., Suzuki, M., Dong, X., et al.: Grafting to achieve flame retardancy. Polym.
Degrad. Stab 54, 117–124 (1996)
2. Jayakody, C., Myers, D., Crocker, M., et al.: A new neneration of fire-resistant seat
cushioning materials for airline industry. In: 12th Annual BCC Conference on Flame
Retardancy (2001)
3. Gilman, J.W., Kashiwagi, T., Lichtenhan, J.D.: Nanocomposites: a revolutionary new flame
retardant approach. SAMPE J 33, 40–46 (1997)
4. Rothon, R.N., Hornsby, P.R.: Flame retardant effects of magnesium hydroxide. Polym.
Degrad. Stab 54, 383–385 (1996)
5. Bourbigot, S., Le Bras, M., Duquesne, S., et al.: Recent advances for intumescent polymers.
Macromol. Mater. Eng 289, 499–511 (2004)
6. Cullis, C.F., Hirschler, M.M., Tao, Q.M.: Studies of the effects of phosphorus-nitrogen-
bromine systems on the combustion of some thermoplastic polymers. Eur. Polym. J 27, 281–
289 (1991)
7. Gilman, J.W., Jackson, C.L., Morgan, A.B., et al.: Flammability properties of polymer—
layered-silicate nanocomposites. Polypropylene and polystyrene nanocomposites. Chem.
Mater 12, 1866–1873 (2000)
8. Lewin, M.: Some comments on the modes of action of nanocomposites in the flame
retardancy of polymers. Fire Mater 27, 1–7 (2003)
9. Gilman, J.W., Kashiwagi, T., Morgan, A.B., et al.: Recent advances in flame retardant
polymer nanocomposites. In: International SAMPE Symposium and Exhibition
(Proceedings) (2001)
10. Leszczynska, A., Njuguna, J., Pielichowski, K., et al.: Polymer/montmorillonite
nanocomposites with improved thermal properties. Part I. Factors influencing thermal
stability and mechanisms of thermal stability improvement. Thermochim. Acta 453, 75–96
(2007)
11. Kashiwagi, T., Du, F., Winey, K.I., et al.: Flammability properties of polymer
nanocomposites with single-walled carbon nanotubes: effects of nanotube dispersion and
concentration. Polymer 46, 471–481 (2005)
176 O. Cerin et al.

12. Bian, X.C., Tang, J.H., Li, Z.M., et al.: Dependence of flame-retardant properties on density
of expandable graphite filled rigid polyurethane foam. J. Appl. Polym. Sci. 104, 3347–3355
(2007)
13. Bourbigot, S., Le Bras, M., Leeuwendal, R., et al.: Recent advances in the use of zinc borates
in flame retardancy of EVA. Polym. Degrad. Stab. 64, 419–425 (1999)
14. Jiao, C.M., Wang, Z.Z., Ye, Z., et al.: Flame retardation of ethylene-vinyl acetate copolymer
using nano magnesium hydroxide and nano hydrotalcite. J. Fire Sci. 24, 47–64 (2006)
15. Zhang, Z., Lan, B., Mei, X., et al.: Study on fire retardant mechanism of nano-LDHs in
intumescent system. Sci. China Ser. B Chem. 50, 392–396 (2007)
16. Zammarano, M., Bellayer, S., Gilman, J.W., et al.: Delamination of organo-modified layered
double hydroxides in polyamide 6 by melt processing. Polymer 47, 652–662 (2006)
17. Zammarano, M., Franceschi, M., Bellayer, S., et al.: Preparation and flame resistance
properties of revolutionary self-extinguishing epoxy nanocomposites based on layered double
hydroxides. Polymer 46, 9314–9328 (2005)
18. Bourbigot, S., Duquesne, S.: Fire retardant polymers: Recent developments and
opportunities. J. Mater. Chem 17, 2283–2300 (2007)
19. ISO4589-1 Determination of burning behaviour by oxygen index (1996)
20. ASTMD2863-09 Standard test method for measuring the minimum oxygen concentration to
support candle-like combustion of plastics (oxygen index)
21. UL-94 (1996) Tests for Flammability of Plastic Materials for Parts in Devices and
Appliances, 5th edn
22. ISO5660 Reaction to fire tests—heat release, smoke production and mass loss rate (2002)
23. ASTME1354 Standard test method for heat and visible smoke release rates for materials and
products using an oxygen consumption calorimeter (2009)
24. Ramesan, M.T.: Thermogravimetric analysis, flammability and oil resistance properties in
natural rubber and dichlorocarbene modified styrene butadiene rubber blends. React. Funct.
Polym 59, 267–274 (2004)
25. Sarkos, C.P.: Application of full-scale fire tests to characterize and improve the aircraft
postcrash fire environment. Toxicology 115, 79–87 (1996)
26. Buch, R.R.: Rates of heat release and related fire parameters for silicones. Fire Saf. J 17, 1–12
(1991)
27. Lyon, R.E., Speitel, L., Walters, R.N., et al.: Fire-resistant elastomers. Fire Mater 27, 195–
208 (2003)
28. Levchik, S.V., Weil, E.D.: Thermal decomposition, combustion and fire-retardancy of
polyurethanes—a review of the recent literature. Polym. Int 53, 1585–1610 (2004)
29. Liaw, D.J., Lin, S.P.: Phosphorus-containing polyurethanes based on bisphenol-S, prepared
by N-alkylation. Eur. Polym. J 32, 1377–1380 (1996)
30. Wang, T.L., Cho, Y.L., Kuo, P.L.: Flame-retarding materials. II. Synthesis and flame-
retarding properties of phosphorus-on-pendent and phosphorus-on-skeleton polyols and the
corresponding polyurethanes. J. Appl. Polym. Sci 82, 343–357 (2001)
31. Najafi-Mohajeri, N., Nelson, G.L., Benrashid, R.: Synthesis and properties of new ferrocene-
modified urethane block copolymers. J. Appl. Polym. Sci 76, 1847–1856 (2000)
32. Berta, M., Saiani, A., Lindsay, C., et al.: Effect of clay dispersion on the rheological
properties and flammability of polyurethane-clay nanocomposite elastomers. J. Appl. Polym.
Sci 112, 2847–2853 (2009)
33. Burnside, S.D., Giannelis, E.P.: Synthesis and properties of new poly(dimethylsiloxane)
nanocomposites. Chem. Mater 7, 1597–1600 (1995)
34. Bom, D., Andrews, R., Jacques, D., et al.: Thermogravimetric analysis of the oxidation of
multiwalled carbon nanotubes: evidence for the role of defect sites in carbon nanotube
chemistry. Nano. Lett 2, 615–619 (2002)
35. George, J.J., Bhowmick, A.K.: Influence of matrix polarity on the properties of ethylene vinyl
acetate-carbon nanofiller nanocomposites. Nanoscale Res. Lett 4, 655–664 (2009)
36. Costache, M.C., Heidecker, M.J., Manias, E., et al.: The influence of carbon nanotubes,
organically modified montmorillonites and layered double hydroxides on the thermal
Thermally Stable and Flame Retardant Elastomeric Nanocomposites 177

degradation and fire retardancy of polyethylene, ethylene–vinyl acetate copolymer and


polystyrene. Polymer 48, 6532–6545 (2007)
37. Koo, J.H., Khiet, S.D., Ngujen, C., et al: Flammability studies of a novel class of
thermoplastic elastomer nanocomposites J. Fire Sci. 28 (2010)
38. Gao, F., Beyer, G., Yuan, Q.: A mechanistic study of fire retardancy of carbon nanotube/
ethylene vinyl acetate copolymers and their clay composites. Polym. Degrad. Stab 89, 559–
564 (2005)
39. Beyer, G.: Short communication: carbon nanotubes as flame retardants for polymers. Fire
Mater 26, 291–293 (2002)
40. George, J.J., Bhowmick, A.K.: Ethylene vinyl acetate/expanded graphite nanocomposites by
solution intercalation: preparation, characterization and properties. J. Mater. Sci 43, 702–708
(2008)
41. Duquesne, S., Jama, C., Le Bras, M., et al.: Elaboration of EVA-nanoclay systems—
characterization, thermal behaviour and fire performance. Compos. Sci. Technol 63, 1141–
1148 (2003)
42. Tang, Y., Hu, Y., Wang, S., et al.: Preparation and flammability of ethylene-vinyl acetate
copolymer/montmorillonite nanocomposites. Polym. Degrad. Stab 78, 555–559 (2002)
43. Lu, H., Hu, Y., Li, M., et al.: Clay intercalation and influence on flammability and
crystallization behaviors of POE-based nanocomposites. Polym. Compos 29, 1358–1363
(2008)
44. Kang, D., Kim, D., Yoon, S.H., et al.: Properties and dispersion of EPDM/modified-
organoclay nanocomposites. Macromol. Mater. Eng 292, 329–338 (2007)
45. Huang, H., Tian, M., Liu, L., et al.: Effect of particle size on flame retardancy of Mg(OH)2-
filled ethylene vinyl acetate copolymer composites. J. Appl. Polym. Sci 100, 4461–4469
(2006)
46. Lv, J.P., Liu, W.H.: Flame retardancy and mechanical properties of EVA nanocomposites
based on magnesium hydroxide nanoparticles/microcapsulated red phosphorus. J. Appl.
Polym. Sci 105, 333–340 (2007)
47. Hermansson, A., Hjertberg, T., Sultan, B.A.: The flame retardant mechanism of polyolefins
modified with chalk and silicone elastomer. Fire Mater 27, 51–70 (2003)
48. Hamdani, S., Longuet, C., Perrin, D., et al.: Flame retardancy of silicone-based materials.
Polym. Degrad. Stab 94, 465–495 (2009)
49. Mishra, S., Shimpi, N.G.: Mechanical and flame-retarding properties of styrene-butadiene
rubber filled with nano-CaCO3 as a filler and linseed oil as an extender. J. Appl. Polym. Sci
98, 2563–2571 (2005)
50. Mishra, S., Shimpi, N.G.: Effect of the variation in the weight percentage of the loading and
the reduction in the nanosizes of CaSO4 on the mechanical and thermal properties of styrene-
butadiene rubber. J. Appl. Polym. Sci 104, 2018–2026 (2007)
51. Zhang, J., Wilkie, C.A.: Fire retardancy of polypropylene- metal hydroxide nanocomposites.
A.C.S Symposium Series 922, 61–74 (2006)
52. Szep, A., Szabe, A., Teth, N., et al.: Role of montmorillonite in flame retardancy of ethylene-
vinyl acetate copolymer. Polym. Degrad. Stab 91, 593–599 (2006)
53. Beyer, G.: Flame retardancy of nanocomposites based on organoclays and carbon nanotubes
with aluminum trihydrate. Polym. Adv. Technol 17, 218–225 (2006)
54. Fang, S., Hu, Y., Song, L., et al.: Mechanical properties, fire performance and thermal
stability of magnesium hydroxide sulfate hydrate whiskers flame retardant silicone rubber. J.
Mater. Sci 43, 1057–1062 (2008)
55. Ma, P., Wei, Z., Xu, G., et al.: Dehydration and desulfuration of magnesium oxysulfate
whisker. J. Mater. Sci. Lett 19, 257–258 (2000)
56. Braun, U., Schartel, B.: Flame retardant mechanisms of red phosphorus and magnesium
hydroxide in high impact polystyrene. Macromol. Chem. Phys 205, 2185–2196 (2004)
57. Braun, U., Balabanovich, A.I., Schartel, B., et al.: Influence of the oxidation state of
phosphorus on the decomposition and fire behaviour of flame-retarded epoxy resin
composites. Polymer 47, 8495–8508 (2006)
178 O. Cerin et al.

58. Laoutid, F., Gaudon, P., Taulemesse, J.M., et al.: Study of hydromagnesite and magnesium
hydroxide based fire retardant systems for ethylene-vinyl acetate containing organo-modified
montmorillonite. Polym. Degrad. Stab 91, 3074–3082 (2006)
59. Yang, L., Hu, Y., Lu, H., et al.: Morphology, thermal, and mechanical properties of flame-
retardant silicone rubber/montmorillonite nanocomposites. J. Appl. Polym. Sci 99, 3275–
3280 (2006)
60. Chang, Z.H., Guo, F., Chen, J.F., et al.: Synergic flame retardancy mechanism of
montmorillonite in the nano-sized hydroxyl aluminum oxalate/LDPE/EPDM system.
Polymer 48, 2892–2900 (2007)
61. Chang, Z.H., Guo, F., Chen, J.F., et al.: Synergistic flame retardant effects of nano-kaolin and
nano-HAO on LDPE/EPDM composites. Polym. Degrad. Stab 92, 1204–1212 (2007)
Recycling of Elastomeric Nanocomposites

L. Reijnders

Abstract Recycling elastomeric nanocomposites is discussed in view of resource


cascading, which aims at maximum exploitation of the value and service time of
resources. Special attention is given to particulate releases linked to the elasto-
meric nanocomposite lifecycle. Recycling of tyres containing, at least partly
nanoparticulate, carbon black is discussed as a practical example of the potential
for resource cascading, recycling and particulate releases. In the case of elasto-
meric nanocomposites in general, prevention of degradation and exploring options
for self-healing are worth considering in view of resource cascading. Recycling
options for elastomeric nanocomposites include: reuse of the product, remanu-
facturing of the product and use of nanocomposite granulate (where appropriate,
devulcanized). Further options are ‘chemical recycling’ (recycling of constituents
or conversion products of nanocomposites) and incineration with energy recovery
(‘thermal recycling’). Possibilities for the reduction of nanoparticulate releases
linked to the elastomeric nanocomposite life cycle and socioeconomic arrange-
ments favoring recycling are briefly outlined.

1 Introduction

Firstly, I will clarify what ‘elastomeric nanocomposites’, also called nanocom-


posite elastomers, and ‘recycling’ actually mean in this chapter. Thereafter, in
Sect. 1.3, the rest of this chapter will be briefly outlined.

L. Reijnders (&)
IBED, University of Amsterdam, Nieuwe Achtergracht 166,
1018 WV, Amsterdam, Netherlands
e-mail: l.reijnders@uva.nl

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 179


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_7,
Ó Springer-Verlag Berlin Heidelberg 2011
180 L. Reijnders

1.1 Elastomeric Nanocomposites

Nanocomposite elastomers consist of elastomers and engineered nanoparticles, the


latter being \100 nm in at least one dimension. In this chapter only inorganic
nanoparticles will be considered. For instance, cellulose nanoparticles which are
also considered for application in elastomeric nanocomposites [1] will not be dealt
with, here. Use of engineered inorganic nanoparticles in elastomers has an already
long tradition. Carbon black, traditionally used, and still the most popular filler, in
rubber tyres, is at least partly nanoparticulate [2–5].
There is increasing interest in the use of inorganic nanoparticles in elastomers
to confer improved performance regarding e.g. tensile strength, tear strength,
abrasion resistance, crack growth resistance, barrier function, stiffness, hardness,
permeability control, flame retardancy, electrical conduction, weatherability and
stability [4, 6–31]. The nanoparticles studied in this context include ‘nanoclay’
(e.g. montmorillonite, fluorohectorite, bentonite, kaolinite), silica (SiO2), CaCO3,
titania (TiO2) and carbon nanofibers and nanotubes, Silica particles are now
commonly used in nanoparticulate elastomers [4].
In producing nanocomposites, often compatibilizing or functionalizing sub-
stances are used to promote dispersion of nanoparticles into polymeric matrices.
Such compatibilizers include: (quaternary) alkylammonium compounds, alkyl-
phosphonium compounds, polyolefin elastomers, maleated compounds and long
chain amines.

1.2 Recycling

Recycling is a loosely used concept. As such, it covers a range of activities varying


from product reuse to incineration combined with energy recovery (which has
been called ‘thermal recycling’).
Recycling is important to the environmental performance of products (e.g. [32–
34]). To the extent that elastomer nanocomposites can be recycled a number of
times as product, material or constituent substances with limited inputs into the
recycling process, the environmental performance per cycle of usage might
improve.
The environmental benefit of many use cycles is at variance with a strategy
aiming at biodegradable nanocomposite elastomers, which involves biodegrada-
tion after one cycle of usage [35–39]. Nanocomposites have drawn special
attention in this context because nanoparticles may enhance biodegradability [37].
From the environmental point of view a strategy aiming at biodegradation would
only seem worth considering when inputs in the recycling process are very high
and the hazard of substances released by biodegradation, which in the case of
nanocomposite elastomers would probably include hazardous nanoparticles (see
Sect. 3), is very low. It would seem hard to come up with a case in which such
conditions would be met by elastomeric nanocomposites.
Recycling of Elastomeric Nanocomposites 181

The applicability of recycling to elastomeric nanocomposites will be discussed


against the background of a normative strategy regarding the resources used for
producing nanocomposite elastomers. This normative strategy is based on the
argument that recycling should aim at ‘resource cascading’, which in one version
has been specified as the maximum exploitation of the value and service time of
resources [40]. Alternatively, it has been argued that recycling should aim at
maximization of the number of times that a material is utilized [41]. The difference
between these two specifications is that the former values a more demanding
application higher than a less demanding application and relatively long usage
higher than short usage [42], whereas the latter gives equal value to all uses. Here
the first specification will be followed.

1.3 Outline of the Chapter

As pointed out in Sect. 1.1, tyres are a longstanding example of elastomeric


nanocomposites. They belong to the category of thermoset elastomers, which one
would expect to be more difficult to recycle at a high level of resource value than
the category of thermoplastic elastomers [43]. Notwithstanding the latter, there is a
long standing tradition of recycling tyres. So, to get a practical perspective on the
recycling of nanocomposites, it would seem useful to look at current tyre recycling
practices and the extent to which these are in conformity with resource cascading,
as defined in Sect. 1.2. This will be done in Sect. 2. Also in Sect. 2 environmental
effects of current usage and recycling of tyres will be outlined, including effects on
the workplace environment. In this section special attention will be given to
releases of particles.
Section 3 will consider other nanoparticulate materials than carbon black which
are currently used, or studied for use, in elastomeric nanocomposites. The nano-
particles which will be discussed in this section are: silica nanoparticles, carbon
nanofibers and nanotubes, nanoclay and nano TiO2. This section will also consider
what is known about the hazards of these nanoparticles to human health.
Section 4 will consider the resource cascade and associated recycling options
for elastomeric nanocomposites other than rubber tyres, and the potential to limit
particulate hazards linked to these elastomers. Section 5 will present concluding
remarks with a focus on the socio-economic arrangements which might be con-
ducive to recycling.

2 Recycling of Tyres and Resource Cascading

Tyres commonly contain a minimum of 20–30% carbon black and/or silica, which
at least partially has a nanoparticulate character [4, 5, 8].
182 L. Reijnders

Worldwide about 65% of all end-of-life tyres are landfilled or dumped, whereas
*35% are the object of some form of recycling [44].
The main types of recycling are: retreading, recycling as rubber powder, rubber
granulate or crumb rubber, and combustion with energy recovery [45]. Rubber
powder or granulate, whether or not subjected to further treatment to improve
performance, may find its way to a wide variety of applications, ranging from new
tyres, brake pedals, rubber wheels, carpet tiles and livestock mats to athletic tracks
and asphalt paving (e.g. [46–52]).
The secondary product is often made of rubber. The properties of the
secondary product are usually such that its performance tends to be at least
somewhat below the standards required for the original tyre, even if there is a
devulcanization step for secondary rubber and admixture of virgin rubber
[53, 54]. To optimize product performance, the modulus of the recycled gran-
ulate should be matched to that of the virgin materials in the matrix [55].
Applications of rubber granulate may also be in composites of rubber and other
materials as in the case of some varieties of asphalt and of rubber/polyolefin
composites [56–59]. Renewed recycling of the latter type of composites should
evade high temperatures because such temperature may degrade rubber and this
in turn may negatively affect important properties such as tensile strength, tear
strength and hardness [58].
‘Thermal recycling’ is often in cement kilns, pulp and paper mills and power
stations [49] (Beukering et al. 2001).
Retreading and use of rubber powder or granulate tends to be favored by
relatively high prices for mineral oil and new tyres [45]. Use of the major recycling
options varies however widely among countries [45, 46, 60].
Of the main recycling options, retreading, which is in the more general category
of remanufacturing [61], would come out relatively high in the resource cascade,
directly after reuse of the product-as-it-is. Use of rubber powder or granulate
would be lower in this resource cascade, and incineration with energy recovery
still lower (cf. [62]).
There are other options for recycling than retreading, use of rubber powder/
granulate and ‘thermal recycling’. The first thereof is pyrolysis. Pyrolyis of tyres
has been studied extensively but has been applied commercially in a very limited
way.
Pyrolysis (also called thermolysis) of tyres may produce products varying
from carbon black and oil to methane, hydrogen and monomeric feedstocks
[44, 49, 63–68]. A problem that may arise in the pyrolysis of tyres is the presence
of contaminants which interfere with further use and/or are relatively toxic.
Examples of the former are N-compounds present in pyrolysis oil which can
poison catalysts [69] and deposits on carbon black which necessitate further
treatment before useful application [70]. Pyrolysis of tyres may also generate
chars, which may be converted into activated carbon [71]. Dependent on the output
of non-fuel products, pyrolysis may score higher in the resource cascade than
incineration with energy recovery (cf. [62]).
Recycling of Elastomeric Nanocomposites 183

There are also other, in practice, minor recycling options. These include the
construction of artificial reefs, the use in playground equipment, erosion control,
bumpers for boats and highway barriers [46].
Finally there is research into self-healing of rubber [72] which is interesting in
view of resource cascading. Self healing of tyres is as yet without substantial
commercial applications. When such applications would come about, retreading
would in the resource cascade perhaps come after self-healing tyres, depending on
the actual composition of the tyre and the implications thereof for recycling (see
Sect. 4.2).
Important to resource cascading is that resources are conserved during a
product life cycle. The tyre life cycle which includes recycling is associated with
substantial resource losses, reflected in releases into the workplace and the wider
environment [45, 73]. Such releases may have negative impacts on health. Also
some of the recycling options, such as for instance the usage of tyre granulate in
asphalt, may be associated with substantial resource losses.
As to the health impact of resource loss, the following may be noted.
Occupational exposure to carbon black has been found associated with
increased pulmonary and cardiovascular illness, and exposure to carbon black
has also been linked to oxidative stress and mutagenic/genotoxic effects [1, 5,
74–77]. Carbon black has been classified as possibly carcinogenic to human
beings [76].
A tyre tends to lose *10% of its weight before it is disposed of [45]. Most of
this material comes from the rubber tread and a substantial part of the particles
which are released is in the respirable category (particulate matter with a diameter
\10 lm or PM 10) [73]. Indeed, at busy city roads *5–7% of PM 10 may be tyre
particles [73]. PM 10 is likely to contribute to increased risk of pulmonary and
cardiovascular disease for people chronically exposed to particulate matter levels
common along busy roads [73]. In scrap-tyre shredding facilities in Taiwan
respirable particulate concentrations have been found which, on chronic exposure,
would entail considerable risk of pulmonary and cardiovascular disease [73, 78].
These particles have moreover been found to be mutagenic, which may entail
increased cancer risk on exposure [73, 78].
Also linked to the tyre life cycle, is the loss into the environment of Zn (zinc),
which tends to be present in tyres at a level of about 1.5% [79, 80].

3 Nanoparticles for Use in Nanocomposite Elastomers


and Their Hazards to Human Health

The nanoparticles which will be discussed in this section are: silica nanoparticles,
carbon nanofibers and nanotubes, nanoclay and nano-TiO2. As to the latter, also
TiO2 nanoparticles coated with SiO2 will be considered. In this type of TiO2
nanoparticles the photocatalytic activity is suppressed. Regarding the hazards of
these nanoparticles to human health, the following has been found. There is
184 L. Reijnders

evidence that the nanoparticles considered here, when inhaled, are hazardous to
lung tissue and, dependent on size, to the cardiovascular system [81–85]. Hazard
to the cardiovascular system is probably linked to translocation from the lungs.
Such translocation may also negatively affect other organs.
The actual hazard following inhalation is dependent on number, diameter,
surface area, structure, surface charge, surface energy, agglomeration and shape
of the nanoparticles [81, 83, 86–96]. Ceteris paribus, carbon nanofibers and
nanotubes which are longer than 10 lm would seem to be more hazardous than
e.g. spherical nanoparticulate carbon black because macrophages which clean the
lungs would seem to have much more difficulty in clearing such carbon fibers
and tubes than carbon nanospheres [87, 97] and because carbon nanotubes
inhaled by mice can reach the outer lining of the lungs and cause scarring [98].
Also, ceteris paribus, anatase TiO2 nanoparticles seem to be more hazardous
than rutile TiO2 nanoparticles [99, 100]. There is furthermore suggestive
evidence that nanoparticles can be translocated from the nasal area to the central
nervous system via the olfactory nerve, thus posing a hazard of inhaled
nanoparticles to the central nervous system including enhanced inflammation
[83, 101–105].
There is some evidence that the nanoparticles considered here may be
hazardous after ingestion [43, 106]. This hazard regards the intestines and, after
translocation from the digestive tract, other organs.
Uncoated TiO2 nanoparticles and carbon nanotubes, especially when residues
of catalyst are present, may be a health hazard on dermal exposure [84, 107–109].
TiO2 nanoparticles have also been found to exhibit immunotoxicity [110].
A main molecular mechanism underlying the hazard of nanoparticles to human
health may be oxidative damage caused by nanoparticles [81, 83, 85, 102, 107,
111–117]. However, this is probably not the only mechanism underlying hazard, as
SiO2-coated TiO2 particles are linked to a relatively high hazard of lung inflam-
mation, which is apparently not caused by oxidative damage [118]. Similarly, the
necrotic effect of anatase TiO2 nanoparticles is unlikely to be explained on the
basis of oxidation reactions only [99].
In evaluating the hazard of nanoparticles, one should be aware of the possibility
that nanoparticles may absorb hazardous substances, and thus may cause exposure
to combinations of substances [86, 93, 119].

4 Resource Cascading and Recycling of Nanocomposites


Other Than Rubber Tyres

Proper recycling is dependent on the availability of suitable technology and


socio-economic arrangements conducive to recycling. With respect to the latter,
tyres are not necessarily representative for all elastomeric nanocomposites. A first
cause of differences is the matter of identification. Tyres are easily recognized and
have a relatively well defined composition, and this may not hold for other
Recycling of Elastomeric Nanocomposites 185

nanocomposite elastomers. To remedy the problem of identification it has been


proposed to tag products with information about composition which is relevant to
recycling [120, 121]. Also in other respects there may be differences. For instance,
car components may come under extended producer responsibility [45, 60], which
is conducive to recycling, but this is not necessarily so for all nanocomposite
elastomers.
Designing optimum socio-economic arrangements for recycling is a complex
matter (e.g. [122]). In this chapter the focus will be on the technical aspects of
resource cascading and associated recycling. However in Sect. 5 the social and
economic arrangements which might be conducive to recycling of nanocomposite
elastomers will be briefly discussed.

4.1 Preventing Degradation of Elastomeric Nanocomposites

Whether degradation during usage and recycling can be limited, is important to


retaining high value for resource cascading in the case of nanocomposite
elastomers.
When oxygen is available, mineral nanoparticles such as SiO2, TiO2, C, and
clay nanoparticles, generate reactive oxygen species, which may cause scission of
polymer chains. An increased size of surface area, specific crystal structures and
the chemical nature of the surface are important determinants of the capacity to
generate reactive oxygen species [3, 5, 83, 112, 123, 124]. Polymer scission by
reactive oxygen species may be suppressed by antioxidants.
In the presence of sunlight, scission of polymers due to the formation of
reactive oxygen species may also occur when nanoparticles turn out to be phot-
ocatalytically active. The latter applies to uncoated TiO2 nanoparticles [108] and
to montmorillonite nanoclay [125–128]. Such nanoparticles may, when the
nanocomposite is indeed exposed to sunlight (or another UV source), lead to a
decrease in durability of the polymer [125–127], which in turn negatively affects
the potential for resource cascading as defined in Sect. 1.2.
The photocatalytic properties of the TiO2 nanoparticle can be suppressed by
coating by e.g. silica, which for this purpose should preferably be complete [108].
Alternatively, increased additions of anti-oxidants, HALS light stabilizers and UV
absorbents may be employed to reduce the deterioration of polymers by photo-
catalytically active nanoparticles.
Degradation may also be linked to thermo-oxidative stress (e.g. during pro-
cessing) and to delamination in the case that layered silicates (e.g. nanoclays) are
used [124, 129–132]. The impact of mineral nanoparticles on the actual thermo-
oxidative degradation of elastomers is somewhat uncertain as empirical studies
have given rise to contradictory results [124, 132]. However, it seems that mineral
nanoparticles are likely to be conducive to thermo-oxidative degradation [124,
129–134]. This may be linked to the generation of reactive oxygen species by
mineral nanoparticles. It also appears that compatibilizers used in nanocomposites,
186 L. Reijnders

such as long chain amines, quaternary ammonium salts and maleated compounds in
the case of nanoclay, may enhance degradation, especially at elevated temperatures
which might for instance occur during reprocessing [130, 132, 133, 135]. This has
triggered a search for compatibilizers that have better thermal stability [135].
As degradation of nanocomposites may be linked to thermo-oxidative stress
[132], there is a preference for low temperature recycling processes, which may
also be preferable because such processes are associated with low energy inputs.
Short residence time and limiting exposure to oxygen in processing elastomers
during recycling may also be helpful in limiting thermo-oxidative stress [43].
Additives used to suppress oxidative damage due to the causes discussed here,
may migrate out of the product, which may lead to poorer stability of the nano-
composite over time. Thus, minimization of migration, also at elevated tempera-
tures that may for instance be necessary for reprocessing, should be important in
the choice of additives [43]. But even so, it should be borne in mind that increasing
chemical complexity of the product linked to the addition of additives may be
unfavorable to further resource cascading [43, 126].

4.2 Self Healing of Elastomers

To the extent that product degradation does occur, self-healing of nanopartic-


ulate elastomers may be conducive to maintaining a high value in the resource
cascade. Though there seems to be as yet no commercial application of self-
healing elastomeric nanocomposites, there is increasing research into its
potential. Experimentally, self-healing of neat polymers has been demonstrated
for rubber (based on hydrogen bonding), for polydimethylsiloxane (based on
microcapsules with organotin catalysts) and for Diels–Alder elastomers (using
encapsulated furan/maleimide) [72]. Also, there seems to be potential for self-
healing nanocomposites with carbon fibers and ferrite particles [72]. From the
point of view of further resource cascading it is important that the inclusion of
substances conferring self-healing properties does not interfere with later stages
in the resource cascade.

4.3 Recycling Options for Elastomeric Nanocomposites

From the point of view of resource cascading, the best recycling option would be
reuse of the same product and recycling of production scrap into products.
In analogy to rubber tyres and where appropriate, the second best option for
recycling would seem remanufacturing of the product. In the case of nanocom-
posites exposed to wear and tear, this might for instance imply providing the
product with a new surface layer.
Thirdly, one can consider the recycling of nanocomposite elastomer granulate
or powder, in view of minimizing the environmental burden preferably with low
Recycling of Elastomeric Nanocomposites 187

additional inputs such as energy, restabilizers and compatibilizers (cf. [62, 136,
137]). Among these options, reprocessing, whether or not combined with virgin
material and—where appropriate—devulcanization, generating nanocomposite
products meeting criteria that are not much inferior to the original product, is an
interesting option [55, 137–144]. There is at least some scope for this approach.
Thompson and Yeung [132] have shown that reprocessing of a nanoclay-ther-
moplastic olefin elastomer nanocomposite did induce some degradation in the
nanocomposite, and that damage was reinforced by use of maleated compatibi-
lizer, but that the rheological and mechanical properties of the recycled nano-
composite remained significantly better than those of the neat elastomer. Shim
et al. [145] studied the option of devulcanization of silica-filled poly(dimethylsi-
loxane) and found this to be suitable to recycling, though revulcanization did
restore original properties largely, but not completely.
An alternative possibility is to include elastomer granulate in a matrix of
elastomer with a different composition, using compatibilizer to achieve acceptable
properties [146, 147]. It should be noted though that meeting performance
requirements in the case of renewed recycling of elastomers composed of more
than one type of polymer may be more difficult than in the case that one type of
polymer is present [43].
Fourthly, one might think of recycling constituents or conversion products of
the nanocomposite ‘chemical recycling’.
For instance, dependent on the type of elastomer, one might consider
chemical decomposition, to reclaim the original constituents of the elastomer.
This has been demonstrated for flexible polyurethane, based on processes such
as aminolysis, hydrolysis, alcoholysis and glycolysis, which regenerate con-
stituent polyols [148–152]. Such processes require severe reaction conditions
(which may include much elevated temperatures), which entail considerable
inputs, and tend to generate substantial amounts of degraded by-products [153].
Also, it has been suggested to degrade polyurethanes by esters of phosphonic or
phosphoric acid, and subsequently treat the reaction products with propylene-
oxide [154]. This would generate fire retardants and other building blocks for
the production of rigid polyurethane foams [154]. Again this is a type of
‘chemical recycling’ which requires considerable inputs. Inputs in polyurethane
recycling processes and the generation of degraded by-products can be reduced
by changes in polyurethane composition, which allow for decomposition at
room temperature [153].
There are also options which focus on ‘chemical recycling’ of a small part of
the nanocomposite. For instance, in the case that the nanoparticles applied are very
valuable, as would currently apply to carbon nanotubes, one might consider
reclaiming the nanoparticles from the nanocomposite e.g. by solvolysis of the
elastomer, as also suggested for reclaiming longer carbon fibers from composites
[155–157]. When the focus is on recycling a small part of the nanocomposite, one
would expect a relatively poor environmental performance.
Pyrolysis of nanocomposite elastomers is still another option in the context of
‘chemical recycling’, as it is in the case of rubber tyres (see Sect. 2) and in the case
188 L. Reijnders

of neat elastomers (e.g. [62, 158–160]). Pyrolysis requires relatively high inputs of
energy (e.g. [155]. One might aim at different outcomes, such as activated carbon,
synthesis gas, oil and monomers. There is also a catalytic variant of pyrolysis,
which might narrow the variety of substances generated [161]. The presence of
mineral nanoparticles, such as clay, SiO2 and TiO2 nanoparticles, may consider-
ably change the degradation patterns of polymers on pyrolysis [159, 162–168].
The presence of mineral nanoparticles in the nanocomposite might for instance
change the composition of the gases produced, including the yield of monomers, or
increase the generation of chars. Apart from pyrolysis, one might consider gasi-
fication and nanocomposite hydrocracking [169].
In resource cascading, incineration or ‘thermal recycling’ is usually the lowest
stage of the cascade [62, 160]. It might however be that, in case of elastomeric
nanocomposites, some types of ‘chemical recycling’ (e.g. focusing on recycling a
small part of the nanocomposite or ‘chemical recycling’ involving large inputs) are
more of an environmental burden than well managed ‘thermal recycling’.

4.4 Reduction of the Release of Hazardous Particles


Associated with the Nanocomposite Life Cycle
and Reduction of Nanoparticle Hazard

In the production stage of nanocomposites, the release of hazardous particles can


be reduced by in situ generation of nanoparticles in elastomers, such as described
for nanosilica by Mark [170], Banyopadhyay et al. [7] and Ikeda et al. [12]. In the
usage stage, robustness of the nanocomposite against degradation (cf. Sect. 4.1)
and wear and tear [4] might reduce the release of nanoparticles. In recycling
contained processing might reduce the release of nanoparticles [171, 172].
To the extent that nanoparticles are released during the life cycle of nanopar-
ticulate elastomers, there may be scope for reduction of hazard by appropriate
coatings for, and functionalization and purification of, nanoparticles. For instance
it has been argued that the hazard of TiO2 nanoparticles can be reduced by
coatings [173]. The hazard of amorphous SiO2 nanoparticles may be reduced by
functionalization with amidosilanes [174]. The hazard of carbon nanotubes can
probably be reduced by purification, which leads to the elimination of metal-
catalyst residues, and the hazard of carbon nanotubes can be modulated by
introducing functional groups [90, 175, 176].

5 Concluding Remarks

It has appeared there is scope for resource cascade-type recycling of nanocom-


posites, and low nanoparticle releases along the product life cycle. Whether the
Recycling of Elastomeric Nanocomposites 189

technical options available will be used, strongly depends on social and eco-
nomical arrangements. Both the examples of recycling of tyres and of the recy-
cling of plastics (which does probably not exceed 10% of virgin plastics
worldwide) illustrate this [43, 45].
Box 1 summarizes social and economic arrangements conducive to the recy-
cling of nanocomposite elastomers [42, 43, 122].

Box 1. Social and Economic Arrangements Conducive


to Cascade-type Recycling of Nanocomposite Elastomers
– High prices of virgin nanocomposites;
– A tradition of (product) design for recycling;
– Labeling to identify the chemical nature of the nanocomposite;
– Extended producer responsibility of nanocomposites;
– Bans on incineration and landfilling;
– Well-organized recycling centers;
– Separate collection of end-of-life products;
– Well developed ‘reverse-logistics’ of end-of-life products.

References

1. Cao, X., Dong, H., Li, C.M.: New nanocomposite materials reinforced with flax cellulose
nanocrystals in waterborne polyurethane. Biomacromolecules 8, 899–904 (2007)
2. Kohjiya, S., Katoh, A., Shimanuki, J., Hasegawa, T., Ikeda, Y.: Nanostructural observation
of carbon black dispersion in natural rubber matrix by three-dimensional transmission
electron microscopy. J. Mater. Sci. 40, 2553–2555 (2005)
3. Kolke, E., Kobayashi, T.: Chemical and biological oxidative effects of carbon black
nanoparticles. Chemosphere 65, 946–951 (2006)
4. Pal, K., Rajasekar, R., Kang, D.J., Zhang, Z.X., Pal, S.K., Kim, J.K., Das, C.K.: Effect of
fillers and nitrile blended PVC on natural rubber/ high styrene rubber with nanosilica
blends: morphology and wear. Mater. Des. 21, 25–34 (2010)
5. Stone, V., Shaw, J., Brown, D.M., Macnee, W., Faux, S.P., Donaldson, K.: The role of
oxidative stress in the prolonged inhibitory effect of ultrafine carbon black on epithelial cell
function. Toxicol. In Vitro 12, 649–659 (1998)
6. Ansarifar, A., Azhar, A., Ibrahim, N., Shiah, S.F., Lawton, J.M.D.: The use of silanized
silica filler to reinforce and crosslink natural rubber. Int. J. Adhesion Adhesives 25, 77–86
(2005)
7. Bandyopadhyay, A., de Sarkar, M., Bhowmick, A.K.: Epoxidized natural rubber/silica
hybrid nanocomposites by sol–gel technique: effect of reactants on the structure and the
properties. J. Mater. Sci. 40, 53–62 (2005)
8. Carratero-Gonzalez, J., Retsos, H., Verdejo, R., Toki, S., Hsiao, B.S., Giannelis, E.F.,
Lopez-Manchado, M.A.: Effect of nanoclay on natural rubber microstructure.
Macromolecules 41, 6763–6772 (2008)
9. Hamdani, S., Longuet, C., Perrin, D., Lopez-Cuesta, J., Ganachaud, F.: Flame retardancy of
silicone-based materials. Polym. Degrad. Stab. 94, 465–495 (2009)
190 L. Reijnders

10. Hanu, L.G., Simon, G.P., Cheng, Y.: Thermal stability and flammability of silicone polymer
composites. Polym. Degrad. Stab. 91, 1373–1379 (2006)
11. Hrachova, J., Chodak, I., Komadel, P.: Modification and characterization of
montmorillonite fillers used in composites with vulcanized natural rubber. Chem. Pap.
63, 55–61 (2006)
12. Ikeda, Y., Poompradub, S., Morita, Y., Kohjiya, S.: Preparation of high performance
nanocomposite elastomer: effect of reaction conditions on in situ silica generation of high
content in natural rubber. J. Sol Gel Sci. Technol. 45, 299–306 (2008)
13. Ismail, H., Osman, H., Jaafar, M.: Hybrid-filler filled polypropylene/(natural rubber)
composites: effects of natural weathering on mechanical and thermal properties and
morphology. J. Vinyl Addit. Technol. 142–151 (2008). doi:10.1002/vnl.20156
14. Kelarakis, A., Yoon, K., Sics, I., Somani, R.H., Hsiao, B.S., Chu, B.: Uniaxial deformation
of an elastomer nanocomposite containing modified carbon nanofibers by in situ
synchrotron X ray diffraction. Polymer 46, 5103–5117 (2005)
15. Khosrokhavar, R., Bakhshandseh, G.R., Ghoreishy, M.H.R., Naderi, G.H.: PP/EPDM
blends and their developments up to nanocomposites. J. Reinf. Plast. Compos. 28, 613–639
(2009)
16. Koo, J.H., Nguyen, K.C., Lee, J.C., Ho, W.K., Bruns, M.C., Ezekoye, O.A.: Flammability
studies of a novel class of thermoplastic elastomer nanocomposites. J. Fire Sci. 28, 49–85
(2010)
17. Lewicki, J.P., Liggat, J.J., Pethrick, R.A., Patel, M., Rhoney, I.: Investigating the aging
behavior of polysiloxane nanocomposites by degradative thermal analysis. Polym. Degrad.
Stab. 93, 158–168 (2008)
18. Lewicki, J.P., Liggat, J.J., Hayward, D., Pethrick, R.A., Patel, M.: Degradative thermal
analysis and dielectric spectroscopy studies of aging in polysiloxane nanocomposites. In:
Celina, M. (ed.) Polymer Degradation and Performance. American Chemical Society,
Washington (2009)
19. Li, Y., Shimizu, H.: Towards a stretchable, elastic, and electrically conductive nanocomposite:
morphology and properties of poly[styrene-b-(ethylene-co-butylene)-b-styrene]/multiwalled
carbon nanotube composites fabricated by high-shear processing. Macromolecules 42, 2587–
2593 (2009)
20. Lopez-Manchado, M.A., Arroyo, M., Herrero, B., Biagotti, J.: Vulcanization kinetics of
natural rubber-organoclay nanocomposites. J. Appl. Polym. Sci. 89, 1–15 (2003)
21. Nambiar, A.N., Mahalik, N.P.: Trends in food packaging and manufacturing systems and
technology. Trends Food Sci. Technol. (2010). doi:10.1016/j.tifs.2009.12.006
22. Peng, Z., Kong, L.X., Li, S., Chen, Y., Huang, M.F.: Self-assembled natural rubber/silica
nanocomposites: its preparation and characterization. Compos. Sci. Technol. 67, 3130–3139
(2007)
23. Ratanasom, N., Saowapark, T., Deepraertkul, C.: Reinforcement of natural rubber with
silica/carbon black hybrid filler. Polym. Test. 26, 360–377 (2007)
24. Sae-Oui, P., Sirisinha, C., Thepsuwan, U., Hatthapanit, K.: Dependence of mechanical and
aging properties of chloroprene rubber on silica and ethylene thiourea loadings. Eur. Polym.
J. 43, 185–193 (2007)
25. Shimpi, N.G., Mishra, S.: Synthesis of nanoparticles and its effect on properties of
elastomeric nanocomposites. J. Nanopart. Res. (2010). doi:10.1007/s11051-009-9768-x
26. Song, L., Li, M., Hu, Y., Lu, H.: Study on preparation and properties of silane-crosslinked
polyethylene/magnesium hydroxide/montmorillonite nanocomposites. J. Fire Sci. 26, 493–
507 (2008)
27. Utracki, L.A., Sepehr, M., Boccaleri, E.: Synthetic, layered nanoparticles for polymeric
nanocomposites. Polym. Adv. Technol. 18, 1–37 (2007)
28. Wahit, M.U., Hassan, A., Rahmat, A.R., Lim, J.W., Mohd Ishak, Z.A.: Effect of organoclay
and ethylene-octene copolymer inclusion on the morphology and mechanical properties of
polyamide/polypropylene blends. J. Reinf. Plast. Compos. 26, 9333–9955 (2006)
Recycling of Elastomeric Nanocomposites 191

29. Wang, T., Dalton, A.B., Keddie, J.L.: Importance of molecular friction in a soft polymer-
nanotube nanocomposite. Macromolecules 41, 7656–7661 (2008)
30. Yang, L., Seryowati, K., Li, A., Gong, S., Chen, J.: Reversible infrared actuation of carbon
nanotube-liquid crystalline elastomer nanocomposites. Adv. Mater. 20, 2271–2275 (2008)
31. Zhu, L., Wool, R.P.: Nanoclay reinforced bio-based elastomers: synthesis and characterization.
Polymer 47, 8106–8115 (2006)
32. Khanna, V., Bakshi, B.R.: Carbon nanofiber polymer composites: evaluation of life cycle
energy use. Environ. Sci. Technol. 43, 2078–2084 (2009)
33. Lloyd, S.M., Lave, L.B.: Life cycle economic and environmental implications of using
nanocomposites in automobiles. Environ. Sci. Technol. 37, 3458–3466 (2003)
34. Vogtländer, J.G., Brezet, H.C., Hendriks, C.F.: Allocation in recycling systems. Int. J. Life
Cycle Assess. 5, 344–355 (2001)
35. El Fray, M.: Polymer matrix nanocomposites from biodegradable thermoplastic elastomers.
Adv. Eng. Mat. 11, B35–B39 (2009)
36. Liao, H., Wu, C.: Synthesis and characterization of polyethylene-octene elastomer/clay/
biodegradable starch nanocomposites. J. Appl. Polym. Sci. 97, 397–404 (2005)
37. Singh, N.K., Purkayastha, B.D., Roy, J.K., Banik, J.M., Yashpal, M., Singh, G., Malik, S.,
Maiti, P.: Nanoparticles-induced controlled biodegradation and its mechanism in
poly(caprolactam). Appl. Mater. Interfaces 2, 69–81 (2010)
38. Stevenson, K., Stallwood, B., Hart, A.G.: Tire rubber recycling and bioremediation.
Bioremediat. J. 12, 1–11 (2008)
39. Zhao, R., Torley, P., Halley, P.J.: Emerging biodegradable materials: starch and protein
based bionanocomposites. J. Mater. Sci. 43, 3058–3071 (2008)
40. Sirkin, T., ten Houten, M.: The cascade chain. Resour. Conserv. Recycl. 11, 215–277 (1994)
41. Yamashita, H., Kishino, H., Hanyu, K., Hayashi, C., Abe, K.: Circulation indexes: new tools
for analyzing the structure of material cascades. Resour. Recycl. Conserv. 28, 85–104
(2000)
42. Reijnders, L.: A normative strategy for resource choice and recycling. Resour. Conserv.
Recycl. 28, 121–133 (2000)
43. Reijnders, L.: Improving resource cascading. In: Loeffe, C.V. (ed.) Trends in Conservation
and Recycling of Resources. Nova Science Publishers, New York (2007)
44. Al-Salem, S.M., Lettieri, P., Baeyens, J.: Kinetics and product distribution of end of life
tyres (ELTs) pyrolysis: a novel approach in polyisoprene and SBS thermal cracking.
J. Hazard. Mater. 172, 1690–1694 (2009)
45. van Beukering, B.J.H., Janssen, M.A.: Trade and recycling of used tyres in Western and
Eastern Europe. Resour. Conserv. Recycl. 33, 253–265 (2001)
46. Amari, T., Themelis, N.J., Wernick, I.K.: Resource recovery from used rubber tires. Resour.
Policy 25, 179–188 (1999)
47. Du, M., Guo, B., Jia, D.: Effect of thermal and UV induced grafting of bismaleide on
mechanical performance of reclaimed rubber/natural rubber blend. J. Polym. Res. 12, 473–
482 (2005)
48. Nelson, F.A., Kutty, S.K.N.: Effect of silane coupling agent on cure characteristics and
mechanical properties of chloroprene rubber/ reclaimed rubber blend. Polym. Plast. Eng. 43,
1141–1156 (2004)
49. Smith, D.G., Daniels, E.J., Teotia, A.P.S.: Testing and evaluating commercial applications
of new surface-treated rubber technology utilizing waste tires. Resour. Conserv. Recycl. 15,
133–144 (1995)
50. Sonnier, R., Leroy, E., Clerc, L., Bergeret, A., Lopez-Cuesta, J., Bretelle, A., Ienny, P.:
Compatibilizing thermoplastic/ground tyre rubber powder blends: efficiency and limits.
Polym. Test. 27, 901–907 (2008)
51. Taha, M.M.R., El-Dieb, A.S., ElWahab, M.A.A., Abdel-Waheed, M.E.: Mechanical,
fracture, and microstructural investigations of rubber concrete. J. Mater. Civil Eng. 20, 640–
649 (2008)
192 L. Reijnders

52. Zhang, S.L., Xin, Z.X., Zhang, Z.X., Kim, J.K.: Characterization of the properties of
thermoplastic elastomers containing waste rubber tire powder. Waste Manage. 29, 1480–
1485 (2009)
53. Kojima, M., Tosaka, M., Ikeda, Y., Kohjiya, S.: Devulcanization of carbon black filled
natural rubber using supercritical carbon dioxide. J. Appl. Polym. Sci. 95, 137–143 (2005)
54. Li, S., Lamminmäki, J., Hanhi, K.: Improvement of mechanical properties of rubber
compounds using waste rubber/virgin rubber. Polym. Eng. Sci. 45, 1239–1246 (2005). doi:
10.1002/pen.20402
55. Kumar, P., Fukahori, Y., Thomas, A.G., Busfield, J.J.C.: Volume changes under strain
resulting from the incorporation of rubber granulates into a rubber matrix. J. Polym. Sci. B
45, 3169–3180 (2007)
56. Acierno, D., Ciccarelli, I., Romano, V., Russo, P.: Mechanical and thermal analysis of
composites based on rubbers from used tires. Macromol. Symp. 247, 244–251 (2007)
57. Lee, S.H., Balsubramanian, M., Kim, J.K.: Dynamic reaction inside co-rotating twin screw
extruder. I truck tire model material/polypropylene blends. J. Appl. Polym. Sci. 106, 3193–
3208 (2007)
58. Sae-Oui, P., Sirisinha, C., Sa-nguanthammarong, P., Thaptong, P.: Properties and recyclability
of thermoplastic elastometer prepared from natural rubber powder (NRP) and high density
polyethylene (HDPE). Polym. Test. (2010). doi:10.1016/j.polymertesting.2009.12.101
59. Xin, Z.X., Zhang, Z.X., Pal, K., Byeon, J.U., Lee, S.H., Kim, J.K.: Study of microcellular
injection-molded polypropylene/waste ground rubber tire powder blend. Mater. Des. 31,
589–593 (2010)
60. Milanez, B., Bührs, T.: Extended producer responsibility in Brazil: the case of tyre waste. J.
Clean. Prod. 17, 608–615 (2009)
61. Parkinson, H.J., Thompson, G.: Analysis and taxonomy of remanufacturing. Proc. Inst.
Mech. Eng. 218, 1–13 (2003)
62. Perugini, F., Mastellone, M.L., Arena, U.: A life cycle assessment of mechanical and
feedstock recycling options for management of plastic packaging wastes. Environ. Prog. 24,
137–154 (2005)
63. Al-Salem, S.M., Lettieri, P., Baeyens, J.: The valorization of plastic solid waste (PSW) by
primary and quaternary routes: from reuse to energy and chemicals. Progress Energy
Combust. Sci. 36, 103–129 (2010)
64. Kalitko, V.A.: Steam-thermal recycling of tire shreds: calculation of the rate of explosion-
proof feed of steam. J. Eng. Phys. Thermophys. 81, 781–786 (2008)
65. Kaminsky, W., Mennerich, C.: Pyrolysis of synthetic tire rubber in a fluidised-bed reactor to
yield 1, 3-butadiene, styrene and carbon black. J. Anal. Appl. Pyrol. 58–59, 803–811 (2001)
66. Kaminsky, W., Mennerich, C., Zhang, Z.: Feedstock recycling of synthetic and natural
rubber by pyrolysis in a fluidized bed. J. Anal. Appl. Pyrol. 85, 334–337 (2009)
67. Shah, J., Jan, M.R., Mabood, F.: Catalytic conversion of waste tyres into valuable
hydrocarbons. J. Polym. Environ. 15, 207–211 (2007)
68. Williams, B.T., Brindle, A.J.: Fluidized bed catalytic pyrolysis of scrap tyres: influence of
catalyst; tyre ratio and catalyst temperature. Waste Manage. Res. 20, 546–555 (2002)
69. Mirmiran, S., Pakdel, H., Roy, C.: Characterization of used tire vacuum pyrolysis oil:
nitrogenous compounds from the naphtha fraction. J. Anal. Appl. Pyrol. 22, 205–215 (1992)
70. Zhou, J., Yu, T., Wu, S., Xie, Z., Yang, Y.: Inverse gas chromatography investigation of
rubber reinforcement by modified pyrolytic carbon black from scrap tires. Ind. Eng. Chem.
Res. 49, 1691–1896 (2010)
71. Mui, E.L.K., Cheung, W.H., Valix, M., McKay, G.: Mesoporous activated carbon from
waste tyre rubber for dye removal from effluents. Microporous Mesoporous Mater. 130,
287–294 (2010)
72. Murphy, E.B., Wudl, F.: The world of smart healable materials. Prog. Polym. Sci. 35, 223–
251 (2010)
Recycling of Elastomeric Nanocomposites 193

73. Mantecca, P., Sancini, G., Moschini, E., Farina, F., Gualthieri, M., Rohr, A., Miserocchi, G.,
Palestini, P., Camatini, M.: Lung toxicity induced by intratracheal installation of size
fractionated tire particles. Toxicology 189, 206–214 (2009)
74. Brockmann, M., Fischer, M., Müller, K.M.: Exposure to carbon black: a cancer risk? Int.
Arch. Occup. Environ. Health 71, 85–95 (1998)
75. Nikula, K.J., Snipes, M.B., Barr, E.B., Griffith, W.C., Henderson, R.F., Mauderly, J.L.:
Comparative pulmonary toxicology and carcinogenicity of chronically inhaled diesel
exhaust and carbon black in F 344 rats. Fundam. Appl. Toxicol. 25, 80–94 (1995)
76. International Agency for Research on Cancer (IARC): Monographs on the evaluation of
carcinogenic risk to humans. Vol. 93. Carbon black, titanium dioxide and non-asbestiform
talc. IARC, Lyon (2006)
77. Inoue, H., Shimada, A., Kaewamatawong, T., Naota, M., Morita, T., Ohta, Y., Inoue, K.,
Takano, H.: Ultrastructural changes of the air-blood barrier in mice after intratracheal
instillation of lipopolysaccharide and ultrafine carbon black particles. Exp. Toxicol. Pathol.
61, 51–58 (2009)
78. Chen, Y., Ton, S., Lee, M., Chia, T., Shu, H., Wu, Y.: Assessment of occupational health
hazards in scrap-tire shredding facilities. Sci. Total Environ. 309, 33–46 (2003)
79. van Beers, D., Bertram, M., Fuse, K., Spatari, S., Graedel, T.E.: The contemporary
Oceania zinc cycle: one years stocks and flows. J. Mater. Cycles Waste Manage. 6, 125–
144 (2004)
80. Sörme, L., Bergabäck, B., Lohm, U.: Goods in the anthroposphere as a metal emission
source. Water Air Soil Pollut. Focus 1, 213–227 (2001)
81. Borm, P.J.A., Beruba, D.: A tale of opportunities, uncertainties, and risks. NanoToday 3,
56–59 (2008)
82. Nurkiewicz, T.R., Porter, D.W., Hubbs, A.F., Stone, S., Chen, B.T., Frazer, D.G.,
Boegehold, M.A., Castranova, V.: Pulmonary nanoparticle exposure disrupts systemic
microvascular nitric oxide signaling. Toxicol. Sci. 110, 191–203 (2009)
83. Oberdörster, G., Stone, V., Donaldson, K.: Toxicology of nanoparticles: a historical
perspective. Nanotoxicology 1, 2–25 (2007)
84. Shvedova, A.A., Kagan, V.E.: The role of nanotoxicology in realizing the ‘helping without
harm’ paradigm of nanomedicine: lessons from studies of pulmonary effects of single-
walled carbon nanotubes. J. Intern. Med. 267, 106–118 (2010)
85. Ye, Y., Chen, M., Sun, L., Lan, M.: In vitro toxicity of silica nanoparticles in myocardial
cells. Environ. Toxicol. Pharmacol. (2010). doi:10.1016/j.etap.2009.12.002
86. Auffan, M., Rose, J., Bottero, J., Lowry, G.V., Jolivet, J., Wiesner, M.A.: Towards a
definition of inorganic nanoparticles from an environmental, health and safety perspective.
Nat. Nanotechnol. 4, 634–641 (2009)
87. Borm, P., Castranova, V.: Toxicology of nanomaterials: permanent interactive learning.
Part. Fibre Toxicol. 6, 28 (2009)
88. Brown, S.C., Kamal, M., Nasreen, N., Baumuratov, A., Sharma, P., Antony, V.B., Moudgil,
B.M.: Influence of shape, adhesion and simulated lung mechanics on amorphous silica
nanoparticle toxicity. Adv. Powder Technol. 18, 69–79 (2007)
89. Casals, E., Vazquez-Campos, S., Bastus, N.G., Puntes, V.: Distribution and potential
toxicity of engineered inorganic nanoparticles and carbon nanostructures in biological
Systems. Trends Anal. Chem. 27, 672–683 (2008)
90. Cheng, C., Müller, K.H., Koziol, K.K.K., Skepper, J.N., Midgley, P.A., Welland, M.E.,
Porter, A.E.: Toxicity and imaging of multi-walled carbon nanotubes in human macrophage
cells. Biomaterials 30, 4152–4160 (2009)
91. Choi, H.S., Liu, W., Liu, F., Nasr, K., Misra, P., Bawendi, M.G., Frangioni, J.V.: Design
conditions for tumor targeted nanoparticles. Nat. Nanotechnol. 5, 42–47 (2010)
92. Jiang, J., Oberdörster, G., Biswas, P.: Characterization of size, surface charge, and
agglomeration state of nanoparticle dispersions for toxicological studies. J. Nanopart. Res.
11, 77–80 (2009)
194 L. Reijnders

93. Li, Y., Li, Y., Li, G., Li, J., Li, J., Huang, G., Li, W.: The acute pulmonary toxicity in mice
induced by multiwall carbon nanotube, benzene and their combination. Environ. Toxicol.
(2010). doi:10.1002/tox.20512
94. Mueller, N.C., Nowack, B.: Exposure modeling of engineered nanoparticles in the
environment. Environ. Sci. Technol. 42, 4447–4453 (2008)
95. Poland, C.A., Duffin, R., Kinloch, I., Maynard, A., Wallace, W.A.H., Seaton, A., Stone, V.,
Brown, S., MacNee, W., Donaldson, K.: Carbon nanotubes introduced into the abdominal
cavity of mice show asbestos-like pathogenicity in a pilot study. Nat. Nanotechnol. 3, 423–
428 (2008)
96. Yu, Y., Zhang, Q., Ma, Q., Zhang, B., Yan, B.: Exploring the immunotoxicity of carbon
nanotubes. Nanoscale Res. Lett. 3, 271–277 (2008)
97. Pauluhn, J.: Multiwalled carbon nanotubes (Baytubes): approach for derivation of occupational
exposure limit. Regul. Toxicol. Pharmacol. (2010). doi:10.1016/j.yrtph.2009.12.012
98. Donaldson, K., Poland, C.A.: New insights into nanotubes. Nat. Nanotechnol. 4, 708–710
(2009)
99. Braydich-Stolle, L.K., Schaeublin, N.M., Murdock, R.C., Jiang, J., Biswas, P., Schlager,
J.J., Hussain, S.M.: Crystal structure mediates mode of cell-death in TiO2 nanotoxicity.
J. Nanopart. Res. 11, 1361–1374 (2009)
100. Sayes, C.M., Wahi, R., Kurian, P.A., Liu, V., West, J.L., Ausman, K.D., Warheit, D.B.,
Colvin, V.L.: Correlating nanoscale titania structure with toxicity and inflammatory
response study with human dermal fibroblasts and human lung epithelial cells. Toxicol. Sci.
92, 174–185 (2006)
101. Long, T.C., Tajuba, J., Sama, P., Salkeh, N., Schwartz, C., Parker, J., Hester, S., Lowry,
G.V., Veronesi, B.: Nanosize titanium dioxide stimulates reactive oxygen species in brain
microglia and damages neurons in vitro. Environ. Health Perspect. 115, 1631–1637 (2007)
102. Ma, L., Liu, J., Li, N., Wang, J., Duan, Y., Yan, J., Liu, H., Wang, H., Hong, D.: Oxidative
stress in the brain of mice caused by translocated nanoparticulate TiO2 delivered to the
abdominal cavity. Biomaterials 31, 99–105 (2010)
103. Shin, J.A., Lee, E.J., Seo, S.M., Kim, H.S., Kang, J.L., Park, E.M.: Nanosized titanium
dioxide enhanced inflammatory responses in the septic brain of mouse. Neuroscience 165,
445–454 (2010)
104. Wang, J., Chen, C., Liu, Y., Jiao, F., Li, W., Lao, F., Li, Y., Li, B., Ge, C., Zhou, G., Gao,
Y., Zhao, Y., Chai, Z.: Potential neurological lesion after nasal installation of TiO2
nanoparticles in the anatase and rutile crystal phases. Toxicol. Lett. 183, 72–80 (2008)
105. Wang, J., Liu, Y., Jiao, F., Lao, F., Li, W., Gu, Y., Li, Y., Ge, C., Zhou, G., Li, B., Zhao,
Y., Chai, Z., Chen, C.: Time-dependent translocation and potential impairment on central
nervous system by intranasally installed TiO2 nanoparticles. Toxicology 254, 82–90
(2008)
106. Gerloff, K., Albrecht, C., Boots, A.W., Forster, I., Schins, R.P.F.: Cytotoxicity and
oxidative DNA damage by nanoparticles in human intestinal Caco-2 cells. Nanotoxicology
3, 355–364 (2009)
107. Murray, A.K., Kisin, E., Leonard, S.S., Young, S.H., Kommineni, C., Kagan, V.E.,
Castranova, V., Shevedova, A.A.: Oxidative stress and inflammatory response in dermal
toxicity of single-walled nanotubes. Toxicology 257, 161–171 (2009)
108. Reijnders, L.: Safety of nanoparticles in sunscreens. Househ. Pers. Care Today 3, 16–17
(2009)
109. Wu, J., Liu, W., Xue, C., Zhou, S., Lan, F., Bi, L., Xu, H., Yang, X., Zeng, F.: Toxicity and
penetration of TiO2 nanoparticles in hairless mice and porcine skin after subchronic dermal
exposure. Toxicol. Lett. 191, 1–8 (2009)
110. Palomäki, J., Karisola, P., Pylkkänen, L., Savolainen, K., Alenius, H.: Engineered
nanomaterials cause cytotoxicity and activation on mouse antigen presenting cells.
Toxicology 267, 125–131 (2010)
111. Barillet, S., Simon-Deckers, A., Herlin-Boime, N., Layne-L’Hermite, M., Reynaud, C.,
Cassio, D., Gouget, B., Carriere, M.: Toxicological consequences of TiO2, SiC nanoparicles
Recycling of Elastomeric Nanocomposites 195

and multi-walled carbon nanotubes exposure in several mammalian cell types: an in vitro
study. J. Nanopart. Res. (2010). doi:10.1007/s11051-009-96944
112. Donaldson, K., Berwick, P.H., Gilmour, P.S.: Free radical activity associated with the
surface of particles; a unifying factor in determining biological activity. Toxicol. Lett. 88,
293–298 (1996)
113. Falck, G.C.M., Lindberg, H.K., Suhonen, M., Vippola, M., Vanhala, E., Catalan, J.,
Savolainen, K., Norppa, H.: Genotoxic effects of nanosized and fine TiO2. Hum. Exp.
Toxicol. 28, 339–351 (2009)
114. Jacobsen, N.R., Pojana, G., White, P., Moller, P., Cohn, C.A., Smith-Korsholm, K., Vogel,
U., Marcomini, A., Loft, S., Wollin, H.: Genotoxicty, cytotoxicity, and reactive oxygen
species induced by single-walled carbon nanotubes and C60 fullerenes in the FEI-Muta
mouse lung epithelial cells. Environ. Mol. Mutagen. 49, 476–487 (2009)
115. Ravichandran, P., Periuakaruppan, A., Sadanandan, B., Ramesh, V., Hall, J.C., Jejelowo, O.,
Ramesh, G.T.: Induction of apoptosis in rat lung epithelial cells by multiwalled carbon
nanotubes. J. Biochem. Mol. Toxicol. 23, 333–344 (2009)
116. Wang, F., Gao, F., Lan, M., Yuan, H., Huang, Y., Liu, J.: Oxidative stress contributes to
silica nanoparticle-induced cytotoxicity in human embryonic kidney cells. Toxicol. In Vitro
23, 808–815 (2009)
117. Yu, K.O., Grabinski, C.M., Schrand, A.M., Mursock, R.C., Wang, W., Gu, B., Schlager, J.J.,
Hussain, S.M.: Toxicity of amorphous silica nanoparticles in mouse keratinocytes.
J. Nanopart. Res. 11, 15–24 (2009)
118. Rossi, E.M., Pylkänen, L., Koivisto, A.J., Vippola, M., Jensen, K.A., Miettinen, M., Sirola,
K., Nykäsenoja, H., Karisola, F., Sternvall, T., Kiilunen, M., Pasanen, P., Mäkinen, M.,
Hämeri, K., Joutsenaari, F., Jokiniemi, J., Wolff, H., Savolainen, K., Matikainen, S.,
Alenius, H.: Airway exposure to silica coated TiO2 nanoparticles induces pulmonary
neutrophilia in mice. Toxicol. Sci. 113, 422–433 (2010)
119. Liu, S., Xu, L., Zhang, T., Ren, G., Yang, Z.: Oxidative stress and apoptosis induced by
nanosized titanium dioxide in PC 12 cells. Toxicology 267, 172–177 (2010)
120. Binder, C.R., Quirici, R., Domnitcheva, S., Stäubli, B.: Smart labels for waste and resource
management. J. Ind. Ecol. 12(2), 207–228 (2008)
121. Luttropp, C., Johansson, J.: Improved recycling with life cycle information tagged to the
product. J. Clean. Prod. 18, 346–354 (2010)
122. Krook, J., Eklund, M.: The strategic role of recycling centres for environmental performance
of waste management options. Appl. Ergon. (2010). doi:10.1016/j.aspergo.2009.06.012
123. Hussain, S., Boland, S., Baeza-Squiban, A., Hamel, R., Thomassen, L.C.J., Martens,
J.A., Billon-Galland, M.A., Fleury-Feith, J., Morsean, F., Pairon, J., Marano, F.:
Oxidative stress and proinflammatory effects of carbon black and titaniumdioxide
nanoparticles: role of particle surface area and internalized amount. Toxicology 200,
142–149 (2009)
124. Zhou, Q., Xanthos, M.: Nanosize and microsize clay effects on the kinetics of the thermal
degradation of polylactides. Polym. Degrad. Stab. 94, 327–338 (2009)
125. Bottino, D.A., Di Pasquale, G., Fabbri, E., Orestano, A., Pollicino, A.: Influence of
montmorillonite nano-dispersion on polystyrene photo-oxidation. Polym. Degrad. Stab. 94,
369–374 (2009)
126. Morlat-Therias, S., Mailhot, B., Gonzalez, D., Gardette, J.: Photooxidation of
polypropylene/montmorillonite nanocomposites. Interaction with anti-oxidants. Chem.
Mater. 17, 1072–1078 (2005)
127. Morlat-Therias, S., Fanton, E., Tomer, N.S., Rana, S., Singh, R.P., Gardette, J.:
Photooxidation of vulcanized EPDM/montmorillonite nanocomposites. Polym. Degrad.
Stab. 91, 3033–3039 (2006)
128. Qin, H., Zhao, C., Zhang, S., Chen, G., Yang, M.: Photo-oxidative degradation of
polyethylene/montmorillonite nanocomposite. Polym. Degrad. Stab. 81, 497–500 (2003)
129. Hong, S., Liao, C.: The surface oxidation of a thermoplastic olefin elastomer under ozone
exposure: ATR analysis. Polym. Degrad. Stab. 49, 437–447 (1995)
196 L. Reijnders

130. Pfaendner, R.: Nanocomposites: industrial opportunity or challenge? Polym. Degrad. Stab.
95, 369–373 (2010)
131. Ramirez-Vargas, F., Navarro-Rodriguez, D., Blagueto-Menchaca, A.I., Huerta-Martinez,
B.M., Palacios-Mezta, M.: Degradation effects on the rheological and mechanical properties
of multi-extruded blends of impact modified polypropylene and poly(ethylene-a-
vinylacetate). Polym. Degrad. Stab. 86, 301–307 (2004)
132. Thompson, M.R., Yeung, K.K.: Recyclability of layered silicate-thermoplastic olefin
elastomer nanocomposite. Polym. Degrad. Stab. 91, 2396–2407 (2006)
133. Kumanayaka, T.O., Parthasarathy, R., Jollands, M.: Accelerating effect on montmorillonite
oxidative degradation of polyethylene nanocomposites. Polym. Degrad. Stab. (2010). doi:
10.1016/jpoldegradstab.2009.11.030
134. Wang, Y., Chen, F., Li, Y., Wu, K.: Melt processing of polypropylene/clay nanocomposites
modified with maleated polypropylene compatibilizers. Compos. B 35, 111–124 (2004)
135. He, A., Wang, L., Yao, W., Huang, B., Wang, D., Han, C.C.: Structural design of
imidazolium and its application in PP/montmorillonite nanocomposites. Polym. Degrad.
Stab. (2010). doi:10.1016/jpolymerdegradstab.2009.12.003
136. Gupta, S., Pallavi, M.B., Som, A., Krishnamurthy, R., Bhowmick, A.K.: Anomalous
mechanical behavior upon recycling of poly(phenylene-ether)-based thermoplastic
elastomer. Polym. Eng. Sci. 48, 496–504 (2008)
137. Vilaplana, F., Karlsson, S.: Quality concepts for the improved use of recycled polymeric
materials. Marcomol. Mater. Eng. 293, 274–297 (2008)
138. Dementienko, O.V., Kuznetsova, O.P., Tikhonov, A.P., Prut, E.V.: The effect of dynamic
vulcanization on the properties of polymer–elastomer blends containing crumb rubber.
Polym. Sci. Ser. A 49, 1218–1225 (2007)
139. Ghosh, A., Rajeev, R.S., Bhattacharya, A.K., Bhowmick, A.K., De, S.K.: Recycling of
silicone rubber waste: effect of ground silicone rubber vulcanizate powder in the properties
of silicone rubber. Polym. Eng. Sci. 43, 279–296 (2003)
140. Grigoryeva, O.P., Fainleib, A.M., Tolstov, A.L., Starostenko, O.M., Lievana, E., Karger-
Kocsis, J.: Thermoplastic elastomers base don recycled high-density polyethylene, ethylene-
propylene-diene monomer rubber, and ground tire rubber. J. Appl. Polym. Sci. 95, 659–671
(2005)
141. Guo, A., Xiang, D., Duan, G., Mou, P.: A review of mechanochemistry applications in
waste management. Waste Manage. 30, 4–10 (2010)
142. Jacob, C., De, P.P., Bhowmick, A.K., De, S.K.: Recycling of EPDM waste. II. Replacement
of virgin rubber by ground EPDM vulcanizate in EPDM/PP thermoplastic elastomeric
composition. J. Appl. Polym. Sci. 82, 3304–3312 (2001)
143. Neto, J.R.A., Visconte, L.L.Y., Tavares, M.I.B., Pacheco, E.B.A.V., Futrado, C.R.G.:
Regeneration of vulcanized compounds based on butadiene-styrene copolymer. Int.
J. Polym. Mater. 56, 565–578 (2007)
144. Susanto, P., Picchioni, F., Janssen, L.P.B.M., Dijkhuis, K.A.J., Dierkes, W.K.,
Noordermeer, J.W.M.: EPDM rubber reclaim from devulcanized EPDM. J. Appl. Polym.
Sci. 102, 5948–5957 (2006)
145. Shim, S.E., Isayev, I., von Meerwall, E.: Molecular mobility of ultrasonically devulcanized
silica-filled poly (dimethylsiloxane). J. Polym. Sci. B 41, 454–465 (2003)
146. Meszaros, L., Tabi, T., Kovacs, J.G., Barany, T.: The effect of EVA content on the
processing parameters and the mechanical properties of LDPE/ground tire rubber blends.
Polym. Eng. Sci. 868–874 (2008). doi:10.1002/pen.21022
147. Mnif, N., Massardier, V., Kallel, T., Elleuch, B.: Study of the modification of
properties of PP/EPR blends with a view to preserving natural resources when
elaborating new formulation and recycling polymers. Polym. Compos. 805–811 (2009).
doi:10.1002/pc
148. Molero, C., de Lucas, A., Rordiguez, J.F.: Recovery of polyols from flexible polyurethane
by ‘split-phase’ glycolysis with new catalysts. Polym. Degrad. Stab. 91, 894–901 (2006)
Recycling of Elastomeric Nanocomposites 197

149. Molero, C., de Lucas, A., Rordiguez, J.F.: Recovery of polyols for flexible polyurethane
foam by splits phase glycolysis; study on the influence of reaction parameters. Polym.
Degrad. Stab. 93, 353–361 (2008)
150. Molero, C., de Lucas, A., Rodriguez, J.P.: Activities of octoate salts as novel catalysts for
the transesterification of flexible polyurethane foams with di-ethylene glycol. Polym.
Degrad. Stab. 94, 533–539 (2009)
151. Watando, H., Say, S., Fakaya, T., Fujieda, S., Yamamoto, M.: Improving chemical
recycling rate by reclaiming polyurethane elastomer form polyurethane foam. Polym.
Degrad. Stab. 91, 3354–3359 (2006)
152. Wu, C., Chang, C., Cheng, C., Huang, H.: Glycolysis of waste flexible polyurethane foam.
Polym. Degrad. Stab. 80, 103–111 (2003)
153. Hashimoto, T., Mori, H., Urushisaki, M.: Poly(tetramethylene ether) glycol containing
acetal linkages: mew PTMG-based polyol for chemically recyclable polyurethane
thermoplastic elastomer. J. Polym. Sci. A 46, 1893–1901 (2008)
154. Grancharov, G., Mitoba, V., Shenkov, S., Topliyska, A., Gitsov, I., Toev, K.: Smart polymer
recycling: synthesis of novel polyurethanes using phosphorus-containing ologomers formed
by controlled degradation of microporous polyurethane elastomer. J. Appl. Polym. Sci. 105,
302–308 (2007)
155. Cheul-Kyu, L., Yong-Ki, K., Phirada, P., Jung-Suk, K., Kun-Mo, L., Chang-Sik, J.:
Assessing environmentally friendly recycling methods for composite bodies of railway
rolling stock using life-cycle analysis. Transp. Res. D (2010). doi:10.1016/jtrd.2010.02.001
156. Pinero-Hernanz, R., Garcia-Serna, J., Dodds, C., Hyde, J., Poliakoff, M., Cocero, M.J.,
Kingman, S., Pickering, S., Lester, E.: Chemical recycling of carbon fibres composites using
alcohols under subcritical and supercritical conditions. J. Supercrit. Fluids 46, 83–92 (2008)
157. Yuyan, L., Gohua, S., Linhui, M.: Recycling of carbon fibre reinforced composites using
water in subcritical condition. Mater. Sci. Eng. A 520, 179–183 (2009)
158. Im, E.J., Kim, S.H., Lee, K.: Optimization of pyrolysis conditions of polyurethane recycling
of solid products. J. Anal. Appl. Pyrol. 82, 184–190 (2008)
159. Kaminsky, W., Predel, M., Sadiki, A.: Feedstock recycling of polymers by recycling in a
fluidized bed. Polym. Degrad. Stab. 85, 1045–1050 (2004)
160. Wollny, V., Dehoust, G., Fritsche, U.R., Weinem, P.: Comparison of plastic packaging
waste management options: feedstock recycling versus energy recovery in Germany. J. Ind.
Ecol. 5(3), 49–63 (2008)
161. Keane, M.A.: Catalytic transformation of waste polymers to fuel oil. ChemSusChem 2,
207–214 (2009)
162. Gilman, J.W.: Flammability and thermal stability studies of polymer layered-silicate (clay)
nanocomposites. Appl. Clay Sci. 15, 31–49 (1999)
163. Gilman, J.W., Harris Jr, R.H., Shields, J.R., Kashiwagi, T., Morgan, A.B.: A study of
flammability reduction mechanism of polystyrene-layered silicate nanocomposite: layered
silicate reinforced carbonaceous char. Polym. Adv. Technol. 17, 263–271 (2006)
164. Holmes, R.L., Campbell, J.A., Burford, R.P., Katchevseva, I.: Pyroysis behaviour of
titanium dioxide-poly(vinyl pyrrolidone) composite materials. Polym. Degrad. Stab. 94,
1882–1889 (2009)
165. Lewicki, J.P., Liggat, J.J., Patel, M.: The thermal degradation behaviour of
polydimethylsiloxane/monymorillonite nanocomposites. Polym. Degrad. Stab. 94,
1548–1557 (2009)
166. Peng, Z., Kong, L.X.: A thermal degradation mechanism of polyvinyl alcohol/silica
nanocomposites. Polym. Degrad. Stab. 92, 1061–1071 (2007)
167. Semenzato, S., Lorenzetti, A., Modesti, M., Ugel, E., Hrelja, D., Besco, S., Michelin, R.A.,
Sassi, A., Facchin, G., Zorzi, F., Bertani, R.: A novel phosphorous polyurethane foam/
montmorillonite nanocomposite: preparation, characterization and thermal behaviour. Appl.
Clay Sci. 44, 35–42 (2009)
168. Zhang, J., Cui, P., Tian, X., Zheng, K.: Pyrolysis studies of polyethylene terephthalate/silica
nanocomposites. J. Appl. Polym. Sci. 104, 9–14 (2007)
198 L. Reijnders

169. Sasse, F., Emig, G.: Chemical recycling of polymers. Chem. Eng. Technol. 21, 777–789
(1998)
170. Mark, J.E.: Some interesting things about polysiloxanes. Acc. Chem. Res. 27, 346–353
(2004)
171. Hansen, S.F., Maynard, A., Baun, A., Tickner, J.A.: Late lessons from early warnings for
nanotechnology. Nat. Nanotechnol. 3, 444–447 (2008)
172. Reijnders, L.: Hazard reduction in nanotechnology. J. Ind. Ecol. 12, 297–306 (2008)
173. Fadeel, B., Garcia-Bennett, A.E.: Better safe than sorry: understanding toxicological
properties of inorganic nanoparticles manufactured for biomedical applications. Adv. Drug
Deliv. Rev. (2010). doi:10.1016/addr.2009.11.006
174. He, X., Liu, F., Wang, K., Ge, J., Qin, D., Gong, P., Yan, W.: Bioeffects of different
functionalized silica nanoparticles on HeCaT cell line. Chin. Sci. Bull. 51, 1939–1946
(2008)
175. Kang, S., Mauter, M.S., Elimelech, M.: Physicochemical determinants of carbon nanotube
bacterial cytotoxicity. Environ. Sci. Technol. 42, 7528–7534 (2008)
176. Tong, H., McGee, J.K., Saxens, R.K., Kodavanti, U., Devlin, R.B., Gilmour, M.I.: Influence
of acid functionalization on cardiopulmonary toxicity of carbon nanotubes and carbon black
particles in mice. Toxicol. Appl. Pharmacol. 239, 224–232 (2009)
Part III
Applications
Elastomeric Nanocomposites for Tyre
Applications

Kaushik Pal, Samir K. Pal, Chapal K. Das and Jin Kuk Kim

Abstract In this study the epoxidized natural rubber (ENR) and organoclay (Cloisite
20A) composites were prepared by solution mixing process. The obtained nanocom-
posites were incorporated in natural rubber (NR) and styrene butadiene rubber (SBR)
blends in presence of varying types of carbon black as reinforcing fillers. Morphology,
curing characteristics, mechanical and thermal properties were characterized and
analyzed. Also, the wear characteristics of the nanocomposites against Du-Pont and
DIN abrader were determined and discussed. The morphology of the organoclay
incorporated in ENR shows a highly intercalated structure. ISAF type of carbon black
shows a significant effect on curing and mechanical properties by reacting at the
interface between SBR and NR matrix. Blends containing ISAF N234 type of carbon
black shows high abrasion resistant properties against Du-Pont and DIN abrader.

1 Introduction

1.1 General

In automobile industry, the design of engine and other mechanical components


receives prime importance, while the tyres are often overlooked. Most means of

K. Pal (&) and J. K. Kim


Polymer Engineering and Science, School of Nano and Advanced Materials,
Gyeongsang National University, Jinju, Gyeongnam, 660-701, South Korea
e-mail: pl_kshk@yahoo.co.inpl.kshk@gmail.com
S. K. Pal
Mining Engineering Department, IIT, Kharagpur 721302, India
C. K. Das
Materials Science Centre, IIT, Kharagpur 721302, India

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 201


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_8,
 Springer-Verlag Berlin Heidelberg 2011
202 K. Pal et al.

transport run by the use of tyres, thus, tyre plays an important and effective role in
transportation. In automobiles, tyre is an inseparable assemblage of materials with
very wide range of properties whose manufacture demands great precision.
Tyre industries consume more than 60% of the rubber product; and the prime
factors of consideration are safety and tread life. Thus, 30% of the mechanical
work available from the fuel is dissipated in the tyres. A reduction of 15% in tyre
rolling resistance may improve fuel consumption by 4.5% [1].
Throughout every industry, whatever type of machine is used, the concerns are
the same: lowering operating costs and maintaining or improving the level of
safety on site. In both areas, tyres are an integral factor. Tyres are one of the
greatest consumable costs in surface mining and in underground trackless opera-
tions, representing as much as 20% of the operating costs of some machines [2].
Increasing the abrasion resistance of rubbers and rubber products is one of the
problems of the rubber industry and is particularly important for the tyre industry. An
analysis of the reasons, why tyres wear out?, from the data of tests on hundreds of
thousands of series of production tyres shows that from 60 to 90% of tyres go out of
use because of wear in the tread. On modern scales of tyre production, every 10%
increase in useful life means a saving of significant costs. It is therefore important to
the economy of the countries that the quality of tyres should be improved [3].

1.2 What is Tyre?

A tyre is a composite, in other words an inseparable assembly of materials with


very different properties, whose manufacture demands great precision.
Tyres, or tires (in American and British English, respectively), are either
pneumatic enclosures, or solid items (including rubber, metals and plastic com-
posites). They are used to protect and enhance the effect of road wheels.
Pneumatic tyres are used on many types of vehicles, from bicycles, motorcy-
cles, cars, trucks, to earthmovers and aircrafts. Tyres enable vehicle performance
by providing for traction, braking, steering, and load support. Tyres provide a
flexible cushion between the vehicle and the road, which smoothes out shock, and
provides comfort (Fig. 1).
Fig. 1 Structure of tyre
Elastomeric Nanocomposites for Tyre Applications 203

It is made up of the following semi-finished products:


(1) The inner liner,
(2) The casing ply,
(3) The lower bead area,
(4) Bead wires,
(5) Sidewalls,
(6) Bracing plies,
(7) Tread
The fundamental tyre functions are:
(1) Proving load carrying capacity,
(2) Provide cushioning and enveloping,
(3) Transmit driving and breaking torque,
(4) Producing tractive force,
(5) Provide dimensional stability,
(6) Resist abrasion,
(7) Provide steering response,
(8) Have low rolling resistance,
(9) Provide minimum noise.
(10) Permit minimum road vibration,
(11) Be durable and safe.

1.2.1 Materials for Tyre

The basic materials for the production of tyres are:


(1) Base Rubber: NR, SBR, NBR, PBR, PUR, XNBR etc.
(2) Fillers: carbon black, China clay
(3) Additives: sulfur, peroxide, process oil (aromatic or aliphatic), wax, acceler-
ator (CBS, MBT, MBTS, TMTD, DCBS, TBBS, DPG etc.), accelerator
activator (ZnO, Stearic acid etc.), antioxidant and antiozonant (IPPD, HQ, TQ
etc.), silica, nano fillers (clay, nanotuber, fibre etc.)
(4) Wire: steel, brass, nylon, polyvinyl, polyamide, polyester cord etc.

1.3 Tyre Life and the Causes of Tyre Wear

There are many factors that influence tyre life. Some of these are:
1. Cuts
2. Contamination
3. Dual tyre matching
4. Tyre rotation
5 Vehicle maintenance
204 K. Pal et al.

a. Misaligned wheels
b. Wheel balancing
c. Mechanical irregularities
6. Rolling resistance
7. Inflation pressure
a. Over inflation/under loading
b. Under inflation/over loading
8. Heat
9. Incorrect alignment
10. Grip
11. Load
12. Abnormal tyre wear

1.3.1 Damage and Wear of Tyres

Following are some reasons leading to damage and wear of tyres:


1. Tread detachment
2. Air pockets
3. Sidewall cuts and rupture of the sidewall
4. Rupture resulting from a cut in the tread
5. Impact ruptures
6. Pinch rupture caused by road shock

Fig. 2 Different types of wounds in tyres


Elastomeric Nanocomposites for Tyre Applications 205

7. Bead damage
a. Incorrect fitting or removal
b. Due to over inflation
The images of several types of damages are depicted in the Fig. 2.

1.4 Phenomenon of Wear in Rubber

Wear is ‘the process in which the tyre becomes unfit for use during a certain
minimum service life’, which to considerable extent results from, or is related to,
continuous damage of the tread [4].
Investigation by Rymuza [5, 6] shows that the wear dynamics of polymer–
polymer and polymer-metal systems is determined by properties of the polymer
such as surface energy, modulus of elasticity, specific heat, thermal conductivity
and various operating conditions. Ratner [5, 7], Lewes [8], Rhee [9], Lancaster
[10], Atkinson [11] and others have developed various forms of equations and
relationships for the wear of polymers using variables such as load/pressure, speed,
sliding length, sliding duration, shear strength of polymer etc. In 1974, Kar and
Bahadur [12] developed a wear equation in terms of the sliding variables, pressure,
speed, time and the material properties, modulus of elasticity, surface energy,
thermal conductivity and specific heat.
A proper understanding is necessary for control and prediction of polymer
performance [13]. A vast amount of literature has appeared over the years in which
relations between tribological performance and polymer properties are described
in terms of mechanical parameters, such as yield and shear stress, toughness (as
defined by the product of stress- and strain-to-break), plasticity index, Young’s
modulus, and hardness [14–17]. These studies and reviews have helped in pre-
dicting polymer behaviour in sliding wear and friction to a certain level.

1.4.1 Relation between Abrasion Resistance and Mechanical Properties


of Rubber

Regarding physical ideas on the nature of abrasion, Schallamach [18, 19] was the
first to examine the simple case of the failure of rubber by the action of a hard
projection. The coefficient of friction l can also be expressed as a function of the
elastic and hysteresis properties of the rubber and of the configuration of the
abrading surface [18, 19].

1.4.2 Effect of Temperature on the Resistance to Wear

In certain operating conditions of tyres e.g. sudden braking and acceleration, sharp
bends and high speeds etc., high temperature develops in the contact area [20].
206 K. Pal et al.

It has been observed [19] that during intense abrasion in sliding contact, high
temperature is developed, and consequently, the abrasion resistance of the rubber
depends to large extent on the resistance to high temperature and heat. One sug-
gestion is that chemical degradation of the sliding interface of the rubber can be
attributed to a thermal effect [21]. In the wear of tyres, the temperature of the
rubber at the interface needs to be considered in relation to abrability, which is a
function of temperature [22].

1.4.3 Mechanism of Wear of Tread Rubbers

Tread wear in pneumatic tyres may be measured as weight loss or decrease in


crown thickness, or more commonly as a loss of tread depth over prolonged
periods. Pavement texture, as might be supposed, plays a major role in determining
the extent and severity of the wear mechanism, in addition to driving habits,
climatic conditions and operational factors. There are several mechanisms
involved for tyre abrasion, such as,
(1) Fatigue, or hysteresis wear
(2) Abrasive, or catastrophic wear
(3) Cohesive tearing, or wear by roll formation
(4) Tread reversion and blistering
(5) Smearing of rubber
(6) Threshold strength of rubber

1.4.4 Relation between Abrasion and Tensile Strength of Rubber

The investigation of abrasion by scratching rubber with needle led Schallamach to


the conclusion that wear of rubber on sharp abrasives was due to tensile failure
[23, 24]. Buist and Davies [25] proposed an empirical relation between volume (V)
of wear and physical properties of rubber as
V ¼ C0 þ C1  Shore hardness þ C2  Tensile strength ð1Þ

where, C0, C1, and C2 are constants. Thornly [26] has similarly been able to
correlate tyre wear with hardness and tensile properties.
The effects of the particle size and structure of various carbon blacks on friction
and abrasion behavior of filled natural rubber (NR), styrene-butadiene rubber
(SBR) and polybutadiene rubber (BR) compounds were investigated [27] using a
modified blade abrader. Characteristic parameters like particle size and the
structure of carbon blacks were observed to have a linear relationship with the
Young’s modulus. The frictional coefficient depended not only on the particle size,
but also on the structure of carbon black. The rates of abrasion were decreased
with increasing surface area and developing structure of carbon blacks.
Parkins [28] showed that abrasion resistance of the carbon black filled rubber
would increase if the tensile strength is higher.
Elastomeric Nanocomposites for Tyre Applications 207

Rattanasoma et al. [29] showed that the vulcanizates containing 20 and 30 phr
of silica in hybrid filler exhibit better overall mechanical properties as well as
abrasion resistance properties.
Persson [30] reported that friction force decreased as the velocity increased in
the high velocity range depending on the nature of the substrate, surface roughness
and the mechanical properties of rubber. Persson’s attention was focused on
examining the behavior of bulk rubber undergoing continuous motion.
Many attempts have been made to find a relationship between the rate of
abrasion loss and the physical and mechanical properties of rubber [22, 31–33].
Uchiyama [33] found the following relationship for determining the wear volume
of rubber abrasion:
lP
V ¼ k1 L ð2Þ
rB
where V is the wear volume, l represents the friction coefficient, P denotes the
normal load, L the length of rubbing distance and k1 is a constant. The parameter
rB is expressed by the following equation:
rB ¼ rN ð3Þ
where r is maximum amplitude of tensile stress and N is the number of cycles.
It has been shown that the wear resistance could be correlated with the
mechanical properties of the vulcanizates [34].

1.5 Development of Wear Resistant Rubber Blends

Polymer blends are being used extensively in numerous applications, especially in


tyre manufacture.
In the time of the twelfth century, the source of the whole of the rubber
supplied to the industry was natural rubber. In the nineteenth century, Charles
Goodyear invented the process of ‘sulfur vulcanization’ and suggested the use
of ground natural rubber to overcome its limitations [35]. As a consequence of
the first world war, Germany introduced Buna rubbers which are purely syn-
thetic rubber and it increased the curiosity of polymer chemists all over the
world. Various attempts were made in laboratory to enhance the properties of
natural rubber thus transforming it into a material of desirable properties. As a
result, synthetic rubbers with tailor-made properties were born. Consequently,
different chemicals and methods for vulcanization and processing were
developed.
Apart from blends of common rubbers, specialty rubber is also utilized for
such purposes, depending on service demands and components of the tyre [36,
37]. Many reports covering a wide range of rubber blends have been published.
The use of carbon black is synonymous with the history of tyres. Although it
208 K. Pal et al.

has lost some ground to other reinforcing fillers such as silica, but, by virtue of
its unrivalled performance, it is still the most popular and widely used rein-
forcing filler. However, the primary properties of carbon blacks are normally
controlled by particle size, surface area, structure, surface activity and they are
in most cases interrelated [38]. In tyre treads, silica can yield a lower rolling
resistance at equal wear resistance and wet grip than carbon black [39]. Natural
rubber (NR) is a virgin rubber, having properties resembling those of synthetic
rubbers. Natural rubber (NR) is known to exhibit numerous outstanding prop-
erties; reinforcing fillers are necessarily added into NR in most cases in order to
gain the appropriate properties for specific applications. A wide variety of
particulate fillers are used in the rubber industry for various purposes, of which
the most important are reinforcement, reduction in material costs and
improvements in processing [40]. Reinforcement is primarily the enhancement
of strength and strength-related properties, abrasion resistance, hardness and
modulus. It can offer unique properties such as good oil resistance, low gas
permeability, improved wet grip and rolling resistance, coupled with high
strength. A lot of research has been carried out on NR and SBR blends by
varying the quantity and composition of additives and fillers. The abrasion
resistance of styrene-butadiene tread rubbers is observed to depend to large
extent on the molecular weight distribution, the average molecular weight and
certain other factors [41–43].
Solution styrene butadiene rubber (S-SBR) is used in a wide variety of appli-
cations, including the production of tyres, footwear, conveyor belts, hoses, flooring
and adhesives [44]. Solid solution polymerized styrene butadiene rubber, produced
by anionic batch polymerization is available in a wide variety of styrene and vinyl
contents. S-SBR rubbers provide excellent balance between wet grip, rolling
resistance and dry handling in silica and carbon black compounds for high-per-
formance tyres [45]. They are also used for the manufacture of high quality
technical rubber goods. As it is well known, the performance of motor car tyres
must be improved, whilst at the same time reducing the amount of natural rubber
used. This problem applies even more to large truck tyres, and can be solved
provided that new types of stereo-regular synthetic rubbers like isoprene [46–48]
and butadiene [46, 49–51] are used.
The effects of mixing method, blend ratio, content, type of carbon black, and
vulcanization system have also been compared. Tyres used in mining vehicles are
very costly and need regular maintenance, and it is expensive to replace them
within a very short term. The rugged working conditions in mining industries
reduce the life span of tyres on account of cuts, contamination, abrasion, wear,
speed fluctuations etc. There are several types of damage which occur in the dump-
truck tyre such as tread detachment, sidewall cuts, impact ruptures, bead damage
etc. [39, 52]. The idea of blending synthetic rubbers with natural rubber is certainly
not a new one, but it is only now that this can be shown to be possible with
consistently positive results, by the use of new techniques developed over the last
years.
Elastomeric Nanocomposites for Tyre Applications 209

1.5.1 Compatibility of Polymer Blends

Polymer blend is the intimate mixture of two or more polymers and/or copolymers
resulting from common processing steps [53]. Throughout the last decades, sci-
entific and technical literatures in this area have expanded remarkably as evi-
denced by appearance of several books and proceedings of various conferences
[54]. The word compatibility has been used by many investigators to describe
single phase behavior. However, the terms compatible and incompatible refer to
the degree of intimacy of the blends, which depends on the measurement proce-
dure during examination. A blend could be considered as a compatible blend, if it
does not exhibit the gross symptoms of component polymer segregation, whereas a
heterogeneous blend at a macroscopic level is incompatible. Compatible system of
blends containing high molecular weight polymers have been identified usually
when a favorable specific interaction such as hydrogen bonding, dipole interaction
or ionic interaction exists between two components. Although the majority of
thermoplastics/elastomer blends are heterogeneous, the components may be
referred to as compatible if some technically advantageous combination of prop-
erties can be realized from the blends. Partial compatibility implies that above a
particular level either the minor or the major components remain as a dispersed
phase.
There are some technical problems, which are frequently the result of some
type of mutual incompatibility, which provide an inferior set of properties when
dissimilar polymers are blended together [55]. The blending process and the
quality of the blends can be improved by adjusting the individual raw polymer
viscosity. As a result, the effective viscosities of the phase will no longer mis-
match. The thermodynamic incompatibility can be overcome if the surface energy
differences between polymers are small enough to permit the formation of very
small micro domains of the individual polymer phases and there is sufficient
adhesion between the phases by formation of crosslinks across the interface during
blending [56, 57]. The compatibility of various components and the generation of
single phase from multiphase system play a major role in influencing the physical
properties of the polymer blends. The most significant need of the designer of
polymer blends is to ensure good stress transfer between the components of the
multi-component system which can only vouch for the efficient utilization of
component physical properties of the blends. Numerous techniques have been
utilized to determine the compatibility of the blends but only a few predict good
results [58, 59].

1.6 Nanofillers

Nanofillers have for many years high significance in the plastics industry.
Nanofillers are basically understood to be additives in solid form, which differ
from the polymer matrix in terms of their composition and structure. They
210 K. Pal et al.

Fig. 3 Nanosilica as
observed from TEM

generally comprise inorganic materials, more rarely organic materials. Inactive


fillers or extenders raise the quantity and lower the prices, while active fillers bring
about targeted improvements in certain mechanical or physical properties. The
activity of active fillers may have a variety of causes, such as the formation of a
chemical bond (e.g., cross linking by carbon black in elastomers) or filling of a
certain volume and disruption of the conformational position of a polymer matrix,
and also the immobilization of adjacent molecule groups and possible orientation
of the polymer material [60].
There are many grades of nano fillers, e.g. carbon black, carbon nanotubes,
carbon fiber, activated clay, natural clay (mined, refined, and treated), clay (syn-
thetic), natural fiber, silica, zinc oxide etc. (Fig. 3).

1.6.1 Effects of Nanofillers

Nano filler are expected to improve the properties of materials significantly, more
even at lower loading than conventional/micro-fillers. There are some reasons for
that [61]:
(1) The size effect: large number of particles with smaller inter-particle distance
and high specific surface area results in larger interfacial area with the matrix.
(2) The interactivity and potential reactivity of the nanofillers with the medium.

1.6.2 Nanofiller Reinforcement

The difference between the behaviors of micro and nano-reinforced polymers can
be analyzed by observing the specific changes in properties in nanoscale, in which
polymer chain lengths approach the filler dimensions so that they might display
particular interaction influencing the macroscopic behaviour of the materials.
Elastomeric Nanocomposites for Tyre Applications 211

Fig. 4 Different phases of


nanocomposites

Indeed, many parameters can be taken into account for the reinforcement effi-
ciency of filler into given medium (Fig. 4) [62, 63].
(1) chemical nature of fillers
(2) shape and orientation of the fillers
(3) average size, size distribution, specific area of the particle
(4) volume fraction
(5) dispersion state
(6) interfacial area
(7) respective mechanical properties of each phase

1.6.3 Several Uses of Nanofillers

• Solar energy—tougher, more efficient solar cells are already under develop-
ment, with the promise of drastic cost reductions on the horizon. Some will even
produce hydrogen.
• Fuel cells
• Display technologies and e-paper—e-paper and carbon-nanotube-based field-
emission displays expected to be slugging it out with liquid–crystal displays
(with carbon-nanotube-based backlights, of course) in the next 2 years.
• Nanotubes—both as raw materials and as products. Multi-walled nanotubes are
already used in composites, to increase conductivity at much lower filler loads.
Single-walled nanotubes will have a much bigger effect in the longer term.
• Catalysis has a huge potential geopolitical impact, especially after recent
developments in the energy business.
• Nanocomposites—mainly clay-based for structural applications (increased
strength) or with novel properties. These are already penetrating the
automotive and aerospace industries.
212 K. Pal et al.

• Storage technologies—magnetic random access memory (RAM), nanotube


RAM and terabyte hard drives in the next few years.
• Nanocrystalline bulk materials or steels containing nanoparticulates—some
companies are already using steel with nanoparticulate carbon added during the
rolling process.
• Coatings—extra hard or with special properties, such as being electrochromic or
self-cleaning, are under investigation by everyone from car manufacturers to
architects.
• Sensors—bio and chemical sensors made from nanowires and nanotubes are
currently probed.
• Bio analysis—devices using atomic force microscopes and quantum dots are
already being readied for market.
• Textiles—nanofibres in stain-resistant trousers are already available, with
electrospun nanofibres and nanotube-enhanced fibers coming soon.

1.7 Tyre Retreading

Tyres that are fully worn can be re-manufactured to replace the worn tread.
Retreading is the process of buffing away the worn tread and applying a new tread.
Retreading is economical for truck tyres because the cost of the new tread is small
compared to the cost of the tyre carcass. Retreading is less economical for pas-
senger tyres because the cost is high compared to the cost of a new tyre. However,
commercial truck drivers run the risk of ‘‘blow-outs’’, separation, and tread peeling
from the casing, due to constant re-use of the casing.
Commercial trucking companies have taken their own initiative as well. Many
only run retreads on their trailers, and keep ‘‘Virgin Casings’’ (new tyres) on their
Steer and Drive wheels. This ensures that in the event that a retread blows out, the
driver maintains control over the truck.

1.8 Objectives and Scope of Work

As mentioned earlier, apart from the blends of common elastomers, specialty


elastomers are also utilized for tyre applications, depending on service demands
and components of the tyre [36, 37] and carbon black is one the most commonly
used fillers [44, 64–66].
The primary aspect in preparing organoclay nanocomposites is to attain a very
high degree of dispersion of organoclay aggregates that afford to very large surface
areas. Hence, it exhibits significant improvements in physical, mechanical and
thermal properties in relation to the polymer host [67–69]. Though many organo-
clay thermoplastics have been prepared and studied [70–73], less attention has been
Elastomeric Nanocomposites for Tyre Applications 213

paid to use organically modified layered silicates in reinforcing elastomers [74–76].


In this study, rubber-organoclay nanocomposites have been prepared and analyzed.
The accomplishment of highly dispersed organoclay nanocomposites has two
requirements. The first requirement involves the compatibility between the polymer
and nanoclay to obtain better dispersion of organoclay in the polymer matrix.
Organoclay can be easily dispersed in polar polymers than in non-polar polymers.
The formulation of organoclay/polar polymeric systems usually contains a poly-
meric compatibilizer [77, 78] e.g. ENR. ENR is obtained by epoxidation of 1,
4-polyisoprene, depicts higher glass transition temperature and increased polarity.
Accordingly, as a fine dispersion of organoclay is needed, ENR was also chosen as
a compatibilizer in this study. Arroyo et al. [52], Teh et al. [79] and Varghese et al.
[80] have carried out few resplendent works using ENR as compatibilizer for
organoclay/natural rubber nanocomposites. In our previous literature, we have
already analyzed the effect of ENR as a compatibilizer in natural rubber-nanoclay
gum compounds [81] and in presence of carbon black [82].
The second requirement is the preparation methods of nanocomposites. Various
methods have been adopted for the preparation of rubber/organoclay nanocom-
posites that include in situ polymerization intercalation [83], solution intercalation
[84] and melt intercalation [85], and finally co-coagulation of rubber latex and clay
aqueous suspension [86]. In this study, the solution intercalation process has been
used to make the organoclay nanocomposites.
Tyres used in mining vehicles are very costly and need higher abrasion resis-
tance. An increase in the abrasion resistance of rubber products can be achieved by
studying the mechanism of wear of rubber under different operating conditions.
The wear of rubber is a complex phenomenon and dependent on a combination of
processes such as mechanical, mechano-chemical and thermo-chemical etc.
Schallamach [87] and later Grosch [88] reviewed abrasion of rubber and tyre wear.
Champ et al. [89] and Thomas [90] suggested that abrasion takes place through a
cyclic process of cumulative growth of cracks and tearing. Kragelskii et al. [91]
and Schallamach [92] examined the simple case of the failure of rubber by the
action of a hard projection moving over its surface. It has been observed that
during intense abrasion in sliding contact, a high temperature is developed, and
consequently the abrasion resistance of the rubber depends, to a large extent, on its
resistance to high temperature and heat [92]. The earlier literature review has
demonstrated several types of damages and its causes in the tyre [93–112].
Blending of elastomers has been often used to obtain an optimum number of
desirable combinations, physical properties, processability and cost. The elasto-
mers selected in this study were natural rubber (NR) and styrene butadiene rubber
(SBR). Since NR and SBR are non-polar, epoxidized natural rubber (ENR) is used
as a compatibilizer to improve the dispersion of organoclay for nanocomposite
preparation. In this study, incorporation of organoclay in ENR was done by
solution mixing, in order to obtain uniform dispersion of the nanoclay in ENR. The
obtained ENR-organoclay nanocomposites were incorporated in the NR/SBR
blends with varying types of carbon black. The cure characteristics, morphologi-
cal, mechanical and thermal properties of nanocomposites were investigated.
214 K. Pal et al.

These rubber compounds were examined in a specially fabricated experimental


set-up for evaluating their wear resistance properties, when abraded against vari-
ous rock types.

1.9 Experimental

1.9.1 Materials Used in Rubber Preparation

Natural rubber (RMA-1X) was supplied by the Rubber Board, Kottayam, Kerala.
The styrene butadiene rubber (SBR) used was Krylene HS 260, No.-5 of 1948
grade of Bayer AG. Its styrene content is 23.5 ± 1.0, specific gravity is 0.94 and
Mooney viscosity at 100C is 50 ± 5. Epoxidized natural rubber containing 47%
epoxidation unit was supplied by Agricultural Product Processing Research
Institute, Zhangiang, PR China. Cloisite 20A, a natural montmorillonite (MMT)
modified with a quaternary ammonium salt with cation exchange capacity of
95 meq/100 g clay (Southern Clay, Inc, USA), was used as a nanofiller in the
preparation of the nanocomposites. Zinc oxide, stearic acid, N-cyclohexyl-2-
benzothiazyl sulfonamide (CBS) and N-isopropyl-N-phenyl-p-phenylenediamine
(IPPD) were supplied by Bayer (India) Ltd. Standard rubber grade process oil
(Elasto 710) and paraffinic wax were purchased locally. Carbon black was sup-
plied by Birla Carbon.

1.9.2 Solution Mixing Method

Epoxidized natural rubber was dissolved in methyl ethyl ketone (MEK). The ratio
of the rubber to solvent was 1:3 (weight/volume). Continuous stirring was per-
formed at room temperature, until the rubber was completely dissolved in the
solvent. Subsequently 100 wt% of nanoclay (Cloisite 20A) was added to the rubber
solution and stirring was continued. The resultant solution was then cast over in a
thoroughly cleaned plane glass plate. The sample was kept in the same condition
until the solvent was completely evaporated. Appearance of a transparent film was
observed. The obtained nanocomposites contained 1:1 ratio of ENR and nanoclay.

1.9.3 Preparation of Nanocomposites

The compounds were prepared in two-roll mixing mill operated at room temper-
ature. The speed ratio of the rotors was 1:1.4. Initially the natural rubber and
styrene butadiene rubber was masticated followed by incorporating ENR/nanoclay
composites. The reinforcing filler (carbon black) was added along with the process
oil followed by curatives and shown in Table 1.
Also, three types of carbon blacks and one semi reinforcing filler were used,
such as, SAF N110, ISAF N231, ISAF N234 and SRF N774. For vulcanization, the
Elastomeric Nanocomposites for Tyre Applications 215

Table 1 Compound formulation


Compounds Sample codes
A B C D E F
Weight in wt%
Natural rubber (NR) 75 75 75 65 65 65
Styrene butadiene rubber (SBR) 25 25 25 35 35 35
Carbon blacks
1. SAF (N110) 20 – – 20 – –
2. SRF (N774) 20 – – 20 – –
3. ISAF (N234) – 40 – – 40 –
4. ISAF (N231) – – 40 – – 40
ENR/nanoclay (Cloisite 20A) 10 10 10 10 10 10
Stearic acid 2 2 2 2 2 2
Antioxidant (HQ) 1 1 1 1 1 1
Accelerator (CBS) 1 1 1 1 1 1
Zinc oxide 5 5 5 5 5 5
Process oil 2 2 2 2 2 2
Sulfur 2 2 2 2 2 2

amounts of additives such as sulfur, process oil, CBS were based on 100 wt% of
rubber and the samples had the codes ‘A’, ‘B’, ‘C’, ‘D’, ‘E’, and ‘F’, respectively.
The physical properties of the different types of carbon black used in this study
have already been discussed in our earlier literature [113].

1.9.4 Experimental Techniques

Experimental techniques followed in the present research are as follows.

Cure Characteristics of Rubber Compound

The cure characteristics of the rubber compound were studied with the help of a
Monsanto Oscillating Disc Rheometer (ODR—100 s) at 150C as per ASTM
D-2084-07. From the graphs, the optimum cure time, scorch time and Cure Rate
Index (CRI) could be determined.

Determination of Crosslink Density

The cross-link density was determined by immersing a small amount of sample in


100 ml benzene for 72 h to attain equilibrium swelling. After swelling, the sample
was taken out from benzene and the solvent was blotted from the surface of the
sample and weighed immediately. This sample was then dried out at 80C to
remove all the solvent, and reweighed. The volume fracture of rubber in swollen
216 K. Pal et al.

gel Vr, which represented the relative cross-link density of the vulcanizate, was
determined by the following equation.
m0 Uð1  aÞ=qr
Vr ¼ ð4Þ
m0 Uð1  aÞ=qr þ ðm1  m2 Þ=qs

where m0 is the sample mass before swelling, m1 and m2 are sample masses before
and after drying, U is the mass fraction of the rubber in the vulcanizate, a is the
loss of gum EVM vulcanizate during swelling, qr and qs are the rubber and solvent
densities respectively.
The physical cross-link density was calculated using modified [114] Flory-
Rehner equation from swelling measurement in benzene reported earlier [115].

1 ½Vr þ vVr2 þ lnð1  Vr Þ


l
¼ 1=3
ð5Þ
Mc qr VS ðVr  Vr =2Þ
where, 1/M’c, Vr, Vs, v and qr are the physical cross-link density, volume fraction
of rubber, molar volume of swelling medium, the Flory–Huggins solvent-rubber
interaction parameter and density of rubber respectively.
Lastly, the chemical cross-link density was determined from the following
equation.
1 1 1
¼ l
þ  1:55  105 ð6Þ
2Mc 2Mc M

where 1/2Mc, 1/M’c and M are the chemical cross-link density, physical cross-link
density and molecular weight of rubber vulcanizate respectively.

X-Ray Diffraction Measurements (XRD)

X-ray diffraction was performed with a PW 1840 X-ray diffractometer with a


copper target (Cu-Ka) at a scanning rate of 0.050 2h/s, chart speed 10 mm/2h,
range 5,000 c/s, and a slit of 0.2 mm, applying 40 kV, 20 mA to assess the change
of crystallinity of the blends as a function of blend ratio [116]. The range of 2h
scanning of X-ray intensity employed was 1.5–10.
The degree of crystallinity (vc) was measured using the following relationship:
vc ¼ Ic =ðIa þ Ic Þ ð7Þ

where, Ia and Ic are the integrated intensity of the crystalline and amorphous
region, respectively.
The crystallite sizes (P) and the interplaner distance (d) are calculated as
follows:
P ¼ Kk=b cos h ð8Þ
Elastomeric Nanocomposites for Tyre Applications 217

D ¼ k=2 sin h ð9Þ

where b is the half height width (in radian) of the crystalline peak, k is the
wavelength of the X-ray radiation (1.548 for Cu), and k is the Scehrrer constant
taken as 0.9 [117].

Mechanical Characterization (Tensile and Tear)

Vulcanized slabs were prepared by compression molding, and the dumb-bell


shaped specimens for tensile tests and crescent shaped specimens for tear tests
were punched out from a molded sheet by using ASTM Die C. The tests were
performed on a universal tensile testing machine (Hounsfield H10KS) under
ambient condition (25 ± 2C), following the ASTM D 412-06 and ASTM D 624-
00 (2007). The modulus at 100, 300 and 500% elongation, tensile strength, tear
strength and elongation at break (%) were measured at room temperature. The
initial length of the specimens was 25 mm and the speed of the jaw separation was
500 mm/min. Samples were tested five times for each set of conditions, at the
same elongation rate. The values of the tensile strength, modulus at 100% elon-
gation, 300% elongation, 500% elongation and elongation at break were averaged.
The relative error was below 5%.
The hardness was measured by Shore A hardness tester following ASTM
D2240-05 standards.

Thermal Characterization

Thermal characterization (TGA) studies were carried out Shimadzu-DT-40


instrument in presence of air at a rate of 10C/min, using temperature range of 25 to
650C. Degradation temperature of the composites was studied through this anal-
ysis. It is well known that because of the high flexing of tyre off the road (dump-
truck), the temperature rises. In tropical countries, the temperature can go as high as
150C. Hence, the study of thermal resistance of the compounds is desirable.

Transmission Electron Microscopy (TEM)

The dispersion morphology was observed in the high resolution transmission


electron microscope (HRTEM, JEOL 2100). The samples were ultramicrotomed at
-20C for ENR/nanoclay films and -70C for ENR/nanoclay composites in NR.

Scanning Electron Microscopy

The tensile fracture surface of the samples was studied in a scanning electron
microscope (JSM-5800 of JEOL Co.; acceleration voltage of 20 kV; gold coating
218 K. Pal et al.

and 500 and 1000 times magnification. Scanning electron microscopy (SEM) was
used to study the morphology like filler dispersion and indentation of the abrader
on the abraded rubber surface of the samples prepared.

Du-Pont Abrasion Test

Du-Pont abrasion test was done by the Du-Pont Craydon type of abrasion tester for
determining the abrasion resistance of compounds of vulcanized rubber recom-
mended by the Indian Standards Institution vide IS:3400 (Part-III)—1965.

DIN Abrasion Test

DIN abrasion test was done by the DIN abrasion tester for determining the
abrasion resistance of compounds of vulcanized rubber recommended by the
Indian Standards Institution vide IS:3400 (Part 3)—1987.

Heat Buildup Study

Heat build up study was carried out using Goodrich Flexometer for the selective
compounds having higher tensile strength.
The test pieces were prepared in cylindrical shape having diameter
17.8 ± 0.15 mm and height of 25 ± 0.25 mm by compression molding at 150C.
The test pieces were kept at initial temperature of 50C and stoke of
4.45 ± 0.03 mm. The temperature and the load were kept constant throughout the
study. The temperature attained by the samples after the time periods of 10 and
20 min was recorded.

1.10 Results and Discussions

1.10.1 Cure Characteristics of the Rubber Compounds

The optimum cure time (t90) for ‘C’ and ‘E’ rubber sample is higher than other
rubber vulcanizates as shown Table 2. It is possibly due to the mixing of 25 wt%
of SBR irrespective of carbon black grade. The t90 of the compounds sharply
decreased may be due to the amine functionalities in the filler after the modifi-
cation process or ion exchange process. The rate of cure {tmax - tmin} always
increased with increasing concentration of NR. This increase in cure rate can be
due to the fact that an increasing concentration of NR caused the increase in
vulcanization reaction and created more active cross link sites in the rubber
compound.
Elastomeric Nanocomposites for Tyre Applications 219

Table 2 Cure characteristics


Properties A B C D E F
Min. Torque (dN m) 11.39 12.19 11.32 10.40 11.35 10.83
Max. Torque (dN m) 37.58 32.75 59.39 31.29 46.87 43.39
TS2 induct time (min) 1.71 1.82 1.60 2.15 1.81 1.91
TS5 scorch time (min) 2.32 2.40 2.37 2.75 2.51 2.33
TC90 opt. cure time (min) 5.64 5.97 10.07 6.85 9.15 7.85
Opt. cure (min) 35.23 29.62 53.75 38.58 40.13 37.81
Cure rate index (min-1) 23.86 22.54 13.04 19.32 14.88 16.94
Crosslink density (moles/g) 9 10-5 2.27 1.89 4.02 2.66 3.12 3.68

Maximum torque can be considered as a measure of stock modulus [118]. The


torque difference (MH–ML) which shows the extent of cross linking, [119] is found
to have higher variation from one compound to other. The lesser torque difference
is found for the compound ‘B’ and ‘D’ which contain less gel fraction [120]
(Table 2) compared to other rubber vulcanizates, which reduces the maximum
rheometric torque. The obtained cross-link density values from Flory-Rehner
equation [121] correspond with the variation in torque differences.
The lower cross-link density in 10 wt% of ENR/nanoclay, which contains 3.09
wt% of nanoclay for ‘B’ hinders the formation of chemical cross-links and
physical cross-links are formed by the clay bundles [122]. As a result of this, the
decrease in MH–ML value is observed. By the use of semi reinforcing type of
carbon black the scorch time reduces. This decrease in scorch time was due to
presence of active cross-linking sites in the vulcanized rubber [123]. Faster cure
rate index is observed in Table 2 for the compounds containing 25 wt% of SBR
and 75 wt% of NR. The decrease in cure rate may be due to the greater thermal
history formed during mixing, as a result of their higher compound viscosities.
Also, the possible formation of a Zn complex in which sulfur and ammonium
modifier participate may facilitate for the increase in rate of cure [124].

1.10.2 XRD Analysis

X-ray diffraction patterns of epoxidized natural rubber with 100 wt% of nanoclay
loading are shown in Fig. 5. The d-spacing (spacing between the planes in the
atomic lattice) values were calculated using the Bragg’s Law.
The organically modified nanoclay patterns showed an intense peak around
2h = 3.144, corresponding to the basal spacing of 2.58 nm (d001). The EC pattern
showed that the d001 main diffraction peak shifted towards the lower angle
2h = 2.12, corresponding to the basal spacing of 3.79 nm (d001). The peak shift
to a lower angle corresponds to the increased distance between interlayers. The
higher d-spacing value is observed with ENR which signified an intercalated
structure, suggesting that rubbery polymer was incorporated into the interlayer
spacing. XRD data showed that the extent of intercalation was a function of the
polarity of the rubber.
220 K. Pal et al.

Fig. 5 XRD study of ENR


and nano clay

1.10.3 Mechanical Properties of the Rubber Samples

Tensile strength, modulus, elongation at break, tear strength for all the compounds
are shown in Table 3. The tensile strength of ‘A’, ‘E’ and ‘F’ is higher in the
system, because in the case of rubber vulcanizates, the rubber chains orient
themselves in the direction of stretching creating crystallites. These crystallites tie
together a large number of network chains and contribute to high tensile strength
and elongation. But for ‘E’ sample, the tensile strength increases with carbon black
ISAF N234, possibly due to the outstanding reactivity of the carbon black acting as
filler, thus enhancing the properties of the samples. While for other blend types,
the tensile and tear strength starts to decrease, the filler is uniformly dispersed in
the natural rubber matrix which can be attributed to the aggregation of clay
nanolayers [125], also confirmed by the SEM images shown in Fig. 7. The
aggregation leads to the formation of weak points in the NR matrix, accordingly
reducing the elastomeric strength [126, 127]. The filler has high aspect ratio which
leads to improved interfacial bonding and form filler-rubber interactions because
of the high specific surface area of the filler. The mechanical properties of rubber
vulcanizates markedly depended on the number of conjugate double bonds

Table 3 Mechanical properties of the blends


Sample Tensile Elongation 100% 300% Tear Hardness
code strength at break (%) modulus modulus strength (Shore A)
(MPa) (MPa) (MPa) (N/mm)
A 12.19 577 2.25 6.61 50.6 61
B 9.44 457 2.01 5.01 25.2 57
C 8.55 505 1.85 3.45 19.7 64
D 10.19 526 2.44 5.66 32.4 69
E 13.17 529 3.01 7.77 33.8 78
F 12.65 524 2.89 6.32 36.6 70
Elastomeric Nanocomposites for Tyre Applications 221

(sample ‘E’ and ‘F’). These observations suggest that more SBR reacts with the
carbon–carbon double bonds, slower is the reversion reaction rate and hence
increases the mechanical properties of the vulcanizates. The tear resistance of
elastomers is mainly dependent on the processes by which stress dissipation near
the tip of the growing crack takes place. Several processes such as slippage or
breakage of crosslinks or chain entanglements or arresting of the growing crack by
filler particles take place during the tear failure of elastomers [128]. The tear
strength for all the samples is moderate except sample ‘A’, but it varies with
varying the matrix ratio. So the system with carbon black SAF N110 and SRF
N774 gives better reinforcing effect as well as tear strength. It is believed that at
lower filler content, the filler can be dispersed well in the rubber matrix and the
filler can extend further propagation. However, at higher filler content, the filler
tends to form agglomerates, thus decreases the tear properties of composite. The
modulus of all the NR vulcanizates increased with increasing concentration of
SBR. This was because of the following possible reasons: the restriction of
molecular chain mobility, and an increase in the cross-link density. The maximum
torque (MH) is generally correlated with the durometer hardness and modulus. This
indicates that the incorporation of organoclay filler increases the stiffness of the
rubber. The hardness of ‘D,’ ‘E’ and ‘F’ rubber sample was higher than other
blends as shown in Table 3. The increase in hardness of those rubber samples
probably increases the cross-link density.

1.10.4 Thermal Analysis

High temperature Thermal Analysis (TGA) (50–650C) curves for the sample are
shown in Table 4. The temperature for the onset of degradation (T1), the tem-
perature at which 10% degradation occurred (T10), the temperature at which 50%
degradation occurred (T50) and the temperature at which 90% degradation
occurred (T90) were calculated from the TGA plots.
It was observed that the onset degradation temperature was more or less same
for all the samples except sample ‘B’. The onset degradation temperature thereby
probably decreased in the case of rubber sample due to a decrease in cross link
density (Table 4). Cross linking increased the rigidity of the system, which in turn
increased the thermal stability [129, 130]. This proves that increasing percentage
of SBR content is responsible for the increase in thermal stability.

Table 4 Thermal properties


Sample code T10 T50 T90
of the rubber compounds (C)
A 315 339 558
B 307 339 558
C 315 339 552
D 317 410 549
E 315 414 540
F 318 413 558
222 K. Pal et al.

Fig. 6 TEM pictograph of


ENR and nanoclay

1.10.5 TEM Study

TEM is the most lineal method to observe the dispersion of nanoclay. TEM image
of ENR with nanoclay composites is shown in Fig. 6. TEM photomicrographs
showed uniform distribution of the organo-clay in the NR-SBR matrix. Nanoclay
clusters were observed from the image. The dark lines are the silicate layers, in
which bulks of the nanoclay dispersion are in the intercalated state. The clay
structures did not break down during mixing in the NR-SBR matrix and interca-
lation was observed from the TEM photomicrographs. It reveals that nanoclay
dispersion is in the intercalated state, which affirms the better dispersion of
nanoclay in ENR [124].
It should also be noted that TEM showed that the clay layers were dispersed in
the NR matrix at the nano level but XRD indicated that there were some non-
exfoliated MMT layers in the NR matrix.

1.10.6 SEM Study

The tensile fracture samples were scanned after gold coating, and are represented
in Fig. 7. The smooth fracture surfaces and smooth filler dispersion and unidi-
rectional tear path oriented along the direction of flow, which is smooth rubbery in
nature [131], were observed for all rubber samples. Extended nanoclay platelets
which are partially wrapped by the matrix due to the adsorption of the polymer on
nanoclay with some tear line in branching were observed. The micrographs of the
entire rubber sample are characterized by a smooth, rubbery failure (which is a
smooth failure in the case of rubber samples without the formation of necking)
Elastomeric Nanocomposites for Tyre Applications 223

Fig. 7 SEM images of different types of blends

where the additives are clearly seen and the appearance is associated with a low
tensile strength. But for ‘A’, ‘E’ and ‘F’ fatigue, intermolecular and ductile type of
failure was clearly observed. In ‘F’, many holes compared with the fracture surface
of other samples were noticed. This hole formation may be assigned to low rubber-
filler interaction as a result of detachment of the filler from the natural rubber
[132]. Such holes could act as initial flaws leading to localised stress concentration
during deformation. Finally, premature failure of the rubber compound occurred.
This perhaps explains the reduction of both tensile and tear strength with higher
filler content. Some samples like ‘E’ and ‘F’ have also shown rupture type of
failure. More serious effects of hysteresis arise from chemical changes to the
rubber structure at higher sustained temperatures; these effects include the rubber
cross-linking, and thermal degradation leading to explosive rupture (blowout).
224 K. Pal et al.

This phenomenon was studied by Gent and Hindi [133]. They heated rubber
specimens in a microwave oven and showed that blowout was due to the gener-
ation of gases in the interior of rubber components.

1.10.7 Du-Pont Abrader Study

The mass loss of rubber against Du-Pont abrader is given in Fig. 8. The mass loss
of rubber for E and F samples are lower compared to other four type blends under
the same conditions. It is clear that higher abrasion resistance is mainly due to the
presence of ISAF N234 type of carbon black in 65 wt% of NR, 35 wt% of SBR
and 10 wt% of ENR/nanoclay in the blend system. It is also seen that samples
containing 65 wt% of NR, 35 wt% of SBR and 10 wt% of ENR/nanoclay showed
good abrasion property, where as SAF N110 and SRF N774 type of carbon black
with 75 wt% of NR, 25 wt% of SBR and 10 wt% of ENR/nanoclay showed the
high abrasion against Du-Pont abrader.

1.10.8 DIN Abrader Study

Figure 9 refers to the DIN abrasion test result in terms of mass loss of rubber
compounds. Compounds ‘B’ and ‘F’ showed higher abrasion resistance mainly
due to the presence of ISAF N234 type of carbon black for first one and ISAF
N231 type of carbon black for later. The compound C, D, and E exhibited mod-
erate abrasion resistance property, where as SAF N110 and SRF N774 type of
carbon black with 75 wt% of NR, 25 wt% of SBR and 10 wt% of ENR/nanoclay
showed the high abrasion against DIN abrader.

1.10.9 Heat Buildup Study

The values of heat build-up for the compounds, which showed good abrasion
resistant properties against DIN and Du-Pont abrader, are shown in Table 5. For

Fig. 8 Du-Pont abrasion


results
Elastomeric Nanocomposites for Tyre Applications 225

Fig. 9 DIN abrasion results

Table 5 Heat build-up of the


Sample code Temperature (C)
rubber samples
Initial 10 min 20 min
B 50 57 62
D 50 59 66
F 50 58 64

both the compounds, the temperature development is higher, due to the presence of
40 wt% of carbon black and 10 wt% of nano clay. This may be due to the
disproportionate breaking of the carbon black structure and reformation of the
inter-aggregate bonds of carbon black. The compound A shows lesser heat buildup
compared to B and F. The compound ‘A’ contains SAF N110 and SRF N774,
whereas the compound ‘B’ and ‘F’ contain ISAF N234 high structured carbon
black and ISAF N231 low structured carbon black, respectively. The use of semi
reinforcing filler and 80 wt% of NR may be responsible for low heat buildup.
These high temperature containing samples accelerate the fatigue of rubber
components [134]. Higher tyre temperature usually means higher energy dissi-
pation and thus higher fuel consumption [135]. Hence, it is proved that lower
percentage of HSR leads to less heat build-up. It may be concluded that there is an
important connection between heat build-up and the crosslinking system. Thus, the
higher degree of network stability given by sulfur system generally causes high
heat generation. Heat generation tests before and after aging indicate a low degree
of heat build-up can be expected, even when the degree of crosslink densities is
kept similar.

1.11 Summary

Preparation of abrasion resistant tyre tread rubber with the help of an open two-
roll-mixing mill represents a novel method for making high value rubber tyre
226 K. Pal et al.

tread. The NR and SBR with addition of organoclay nanocomposites obtained


from this process has very good mechanical properties which can withstand the
rugged working condition of automobile tyre.
From this study, faster scorch time and cure time has been observed for the
compounds having ISAF N231 type of carbon black. In addition, they show
increase in maximum torque, which correlates with the cross-link density results.
The morphology of the organoclay dispersion in ENR by solution mixing dem-
onstrates the higher intercalation of organoclay based on the XRD results and
TEM images. The FTIR study proves the interaction between ENR and organo-
clay. The overall mechanical properties increases for the compounds containing
ISAF type carbon blacks. In DSC study, the all the samples show the same Tg
except ‘E’, may be due to the ISAF type of carbon black reinforcement with 35
wt% SBR. Higher thermal stability is found for the nanocomposites containing 35
wt% SBR content. It was found that onset degradation temperature was higher for
samples containing 35 wt% of SBR. From the SEM micrographs, fatigue, ductile
and intermolecular fracture type of failure is clearly observed for ‘A’, ‘E’ and ‘F’,
respectively. Samples containing ISAF N234 type of carbon black show higher
abrasion resistance property against Du-Pont and DIN abrader. Also, sample
containing 25 wt% SBR with ISAF N234 type of carbon black shows the lowest
heat generation among all the samples.

References

1. Pal, K., Pal, S.K.: National Seminar on ‘‘Tires in Mining & Allied Sectors: Status And
Outlook’’. ISM, Dhanbad, India (2003)
2. Potts, A.: Mining Magazine, May Ed (2004)
3. James, D.I. (ed.): Wear of Rubber. Maclaren and Sons Ltd., London (1967)
4. Pal, S.K.: Experimental investigations on wear of rubber by rocks. Ph.D. dissertation, IIT
Kharagpur, India (2001)
5. Viswanath, N., Bellow, D.G.: Wear 181–183:42–49 (1995)
6. Rymuza, Z.: Wear in polymer micro-pairs, Proceedings of 3rd international conference on
wear of materials, pp. 125–132 (1981)
7. Ratner, S.B., et al.: Sovt. Plast. 7, 37 (1964)
8. Lewis, R.B.: Mech. Eng. 86, 32–35 (1964)
9. Rhee, S.K.: Wear mechanisms for asbestos-reinforced automotive friction materials. Wear
29, 391–393 (1974)
10. Lancaster, J.K.: Friction and wear. In: Jenkins, A.D. (ed.) Polymer Sciences, Chapter 4
(1972)
11. Atkinson, J.R., Brown, K.J., Dawson, D.: The wear of high molecular weight
polyethylene—Part I: the wear of isotropic polyethylene against dry stainless steel in
unidirectional motion. J. Lubr. Technol. 100, 208–218 (1978)
12. Kar, M.K., Bahadur, S.: The wear equation for unfilled and filled polyoxymethylene. Wear
30, 337–348 (1974)
13. Mergler, Y.J., Schaake, R.P.: Relation between strain hardening and wear resistance of
polymers. J. Appl. Polym. Sci. 92, 2689–2692 (2004)
14. Stejin, R.P.: In: Brostow, W., Corneliussen, R.D. (ed.): Failure of Plastics. Hanser
Publishers, Munich, Chapter 19 (1986)
Elastomeric Nanocomposites for Tyre Applications 227

15. Briscoe, B.J.: In: Friedrich, K. (ed.) Composites Materials Series, vol. 1, Friction and Wear
Of Polymer Composites. Elsevier, Amsterdam, Chapter 2 (1986)
16. Ratner, S.B., Farberova, I.I.: In: James, D.I. (ed.) Abrasion of Rubber. Maclaren and Sons
Ltd., London (1967)
17. Lancaster, J.K.: Abrasive wear of polymers. Wear 14, 223–239 (1969)
18. Kragelskii, I.V., Nepomnyashchil, E.F.: Fatigue Mechanism of the Wear of the Tyre Treads,
Abrasion of Rubber. Maclaren and Sons Ltd., London, pp. 3–13 (1967)
19. Schallamach, A.: Khin I Technol Polimersu 4 (1959)
20. Viehmann, W.: Surface heating by friction and abrasion by thermal decomposition. Rubber
Chem. Technol. 31, 925 (1958)
21. Kragelskii, I.V., Nepomnyashcii, E.F.: In: James, D.I. (ed.) Wear of Rubber. Maclaren and
Sons Ltd., London (1967)
22. Muhr, A.H., Roberts, A.D.: Rubber abrasion and wear. Wear 158, 213–228 (1992)
23. Schallamach, A.: Abrasion of rubber by a needle. J. Polym. Sci. 9, 385 (1952)
24. Schallamach, A.: The velocity and temperature dependence of rubber friction. Proc. Phys.
Soc. B 66, 386 (1953)
25. Bekkedahl, N., Stiehler, R.D.: Natural and synthetic rubbers. Anal. Chem. 21(2), 266 (1949)
26. Box, G.E.P.: Problems in the analysis of growth and wear curves. Biometrics 6(4), 362
(1950)
27. Hong, C.K., Kim, H., Ryu, C., Nah, C., Huh, Y.I., Kaang, S.: Effects of particle size and
structure of carbon blacks on the abrasion of filled elastomer compounds. J. Mater. Sci. 42,
8391–8399 (2007)
28. Parkins, D.: The reinforcement of rubber by carbon black. Br. J. Appl. Phys. 2, 273–280
(1951)
29. Rattanasoma, N., Saowapark, T., Deeprasertkul, C.: Reinforcement of natural rubber with
silica/carbon black hybrid filler. Polym. Testing 26, 369–377 (2007)
30. Persson, B.: Theory of rubber friction and contact mechanics. J. Chem. Phys. 115, 3840–
3860 (2001)
31. Fukahori, Y., Yamazaki, H.: Mechanism of rubber abrasion part 3: how is friction linked to
fracture in rubber abrasion? Wear 188, 19–26 (1995)
32. Schallamach, S.: Abrasion of rubber. Prog. Rubber Technol. 46, 107–142 (1984)
33. Uchiyama, Y.: Studies on the friction and wear of rubbers. Part I: influence of mechanical
properties on the abrasive wear of rubbers. Int. Polym. Sci. Technol. 11, 74–80 (1984)
34. Kurian, J., Nando, G.B.: Scanning electron microscopy studies on wear of HDPE-filled
natural rubber vulcanizates. Wear 127, 139–147 (1988)
35. Goodyear, C.: British Patent 2933, 16 Dec 1853; Miller, G.W.: In: Davis, C.C. (ed.)
Chemistry and technology of rubber, p. 720. Reinhold Publishing Corp., New York (1937)
(New York, state, United States: New York, Middle Atlantic state of the United States. It is
bordered by Vermont, Massachusetts, Connecticut, and the Atlantic Ocean (E), New Jersey
and Pennsylvania (S), Lakes Erie and Ontario and the Canadian province of
http://encyclopedia2.thefreedictionary.com/New+York
36. Gunasekaran, S., Natarajan, R.K., Kala, A.: FTIR spectra and mechanical strength analysis
of some selected rubber derivatives. Spectrochim. Acta Part A 68, 323–330 (2007)
37. Fern’andez-Berridi, M., Gonz’alez, J.N., Mugica, A., Bernicot, C.: Pyrolysis-FTIR and
TGA techniques as tools in the characterization of blends of natural rubber and SBR.
Thermochim. Acta 444, 65–70 (2006)
38. Nasir, M., Choo, C.H.: Cure characteristics and mechanical properties of carbon black filled
styrene-butadiene rubber and epoxidized natural rubber blends. Polym. J. 25, 355–359
(1989)
39. Brinke, J.W.T., Debnath, S.C., Reuvekamp, L.A.E.M., Noordermeer, J.W.M.: Mechanistic
aspects of the role of coupling agents in silica–rubber composites. Compos. Sci. Technol.
63, 1165–1174 (2003)
40. Tinker, A.J., Jones, K.P.: Blends of Natural Rubber, Novel Techniques for Blending with
Speciality Polymer. Chapman & Hall, London (1998)
228 K. Pal et al.

41. Meyer, A.W., Hampton, R.R., Davison, J.A.: Structure of alkali metal-catalyzed butadiene
polymers. J. Am. Chem. Soc. 74, 2294 (1952)
42. Engels, H.-W., Weidenhaupt, H.-J., Pieroth, M., Hofmann, W., Menting, K.-H.,
Mergenhagen, T., Schmoll, R., Uhrlandt, S.: Rubber, 4. Chemicals and Additives.
Ullmann’s Encyclopedia of Industrial Chemistry, Wiley (2004)
43. Dalton, W.O.: Rubber latex agglomeration by acid anhydride addition. J. Colloid Interface
Sci. 43, 339 (1973)
44. Heinrich, G., Kluppel, M., Vilgis, T.A.: Reinforcement of elastomers. Curr. Opin. Solid
State Mater. Sci. 6, 195–203 (2002)
45. Hashim, A.S., Azahari, B., Ikeda, Y., Kohjiya, S.: The effect of bis(3-triethoxysilylpropyl)
tetrasulfide on silica reinforcement of styrene-butadiene rubber. Rubber Chem. Technol. 71,
289–299 (1998)
46. Pankratov, V.A., Zakharkin, O.A., Zakharov, N.D., Kostrykina, G.I., Zhukov, A.A.: Effect
of pulverization on the molecular structure of synthetic cis-polyisoprene. Polym. Sci.
U.S.S.R. 16, 1568 (1974)
47. Horne, S.E., Kiehl, J.P., Shipman, J.J., Folt, V.L., Gibbs, C.F., Willson, E.A., Newton, E.B.,
Reinhart, M.A.: Ameripol SN—A Cis-,4-Polyisoprene. Ind. Eng. Chem. 48, 784 (1956)
48. Huang, D.C., Lin, Y.C., Tsiang, R.C.C.: Synthesis of SBS thermoplastic block copolymers
in cyclohexane in the presence of diethylether used as a structure modifier. J. Polym. Res.
2(2), 91–98 (1995)
49. Portal, J., Carrot, C., Majeste, J.C., Cocard, S., Pelissier,V., Anselme-Bertrand, I.:
Quantification of the distribution of carbon black in natural rubber/polybutadiene blends by
differential scanning calorimetry. Polym. Eng. Sci. 49, 1544 (2009)
50. Garvey, B.S.: Elastomers. Materials of construction review. Ind. Eng. Chem. 53(10), 856–
859 (1961)
51. Sturrock, A.T., Sarbach, D.V.: Antioxidants for cis-polybutadiene compounds. Rubber Age
92, 723 (1963)
52. Arroyo, M., Lopez-Manchadoa, M.A., Valentina, J.L., Carretero, J.: Morphology/behaviour
relationship of nanocomposites based on natural rubber/epoxidized natural rubber blends.
Compos. Sci. Technol. 67, 1330–1339 (2007)
53. Das, T.: Speciality polymer blends of high performance thermoplastics and elastomers with
liquid crystalline polymers. Ph.D. Dissertation, IIT Kharagpur, India (2007)
54. Nielsen, L.E.: Mechanical Properties of Polymer and Composites. Marcel Dekker, New
York (1975)
55. Ahmed, S., Jones, F.R.: A review of particulate reinforcement theories for polymer
composites. J. Mater. Sci. 25, 4933 (1990)
56. Nakamae, K., Nishino, T., Xu, A.R., Matsumoto, T., Miyamoto, T.: Studies on mechanical
properties of polymer composites by X-ray diffraction. I. Residual stress in epoxy resin by
X-ray diffraction. J. Appl. Polym. Sci. 40, 2231 (1990)
57. Black, W.B.: High-modulus wholly aromatic fibers: introduction to the symposium and
historical perspective. J. Macromol. Sci. Chem. A7, 3 (1973)
58. Morgan, P.W.: Synthesis and properties of aromatic and extended chain polyamides.
Macromolecules 10, 1381 (1977). doi:abs/10.1021/ma60060a040
59. Wei, K.H., Ho, J.C.: A study on blends of liquid crystalline copolyesters with
polycarbonate. III. Mechanical properties of compatibilized blends. J. Appl. Polym. Sci.
63, 1527 (1997)
60. Park, D.S., Kim, S.H.: Miscibility study on blend of thermotropic liquid crystalline
polymers and polyester. J. Appl. Polym. Sci. 87, 1842 (2003)
61. Auer, C., Kalinka, G., Krause, T., Hinrichsen, G.: Crystallization kinetics of pure and fiber-
reinforced poly(phenylene sulfide). J. Appl. Polym. Sci. 51, 407 (1994)
62. Liu, J., Tang, G., Qg, U., Zhou, H., Guo, Q.: Crystallization of rare earth oxide-filled
polypropylene. J. Appl. Polym. Sci. 47, 2111 (1993)
63. Phillips, R., Manson, J.E.: Prediction and analysis of nonisothermal crystallization of
polymers. J. Polym. Sci. B Polym. Phys. 35, 875 (1997)
Elastomeric Nanocomposites for Tyre Applications 229

64. Nasir, M., Choo, C.H.: Chemical modification of natural rubber latex with peracetic acid.
Polymer 25, 355 (1989)
65. Tinker, A.J., Jones, K.P.: Blends of Natural Rubber, Novel Techniques for Blending with
Speciality Polymers. Chapman & Hall, London (1998)
66. Choi, S.S., Nah, C., Lee, S.G., Joo, C.W.: Effect of filler–filler interaction on rheological
behaviour of natural rubber compounds filled with both carbon black and silica. Polym. Int.
52, 23 (2002)
67. Alexandre, M., Dubois, P.: Polymer-layered silicate nanocomposites: preparation,
properties and uses of a new class of materials. Mater. Sci. Eng. R Rep. 28, 1–63 (2000)
68. Giannelis, E.P., Krishnamoorti, R., Manias, E.: Polymer-silicate nanocomposites: model
systems for confined polymer and polymer brushes. Adv. Polym. Sci. 138, 107 (1999)
69. Arroyo, M., Lopez-Manchado, M.A., Herrero, B.: Organo-montmorrilonite as substitute of
carbon black for natural rubber compounds. Polymer 44, 2447 (2003)
70. Vaia, R.A., Giannelis, E.P.: Synthesis and properties of two-dimensional nanostructures by
direct intercalation of polymer melts in layered silicates. Chem. Mater. 5, 1694 (1993)
71. Vaia, R.A., Jandt, K.D., Kramer, E.J., Giannelis, E.P.: Kinetics of polymer melt
intercalation. Macromolecules 28, 8080 (1995)
72. Hackett, E., Manias, E., Giannelis, E.P.: Computer simulation studies of PEO/layer silicate
nanocomposites. Chem. Mater. 12, 2161 (2000)
73. Bujdak, J., Hackett, E., Giannelis, E.P.: Effect of layer charge on the intercalation of
poly(ethylene oxide) in layered silicates: implications on nanocomposite polymer
electrolytes. Chem. Mater. 12, 2168 (2000)
74. Burnside, S.D., Giannelis, E.P.: Synthesis and properties of new poly(dimethylsiloxane)
nanocomposites. Chem. Mater. 7, 1597 (1995)
75. Joly, S., Garnaud, G., Ollitrault, R., Bokobza, L., Mark, J.E.: Organically modified layered
silicates as reinforcing fillers for natural rubber. Chem. Mater. 14, 4202 (2002)
76. Zhang, Z., Zhang, L., Li, Y., Xu, H.: New fabricate of styrene–butadiene rubber/
montmorillonite nanocomposites by anionic polymerization. Polymer 46, 129 (2005)
77. Karger-Kocsis, J., Zhang, Z.: Mechanical Properties of Polymers Based on Nanostructure
and Morphology, p. 547. CRC Press, New York (2005)
78. Reichert, P., Nitz, H., Klinke, S., Brandsch, R., Thomann, R., Mulhaupt, R.:
Poly(propylene)/organoclay nanocomposite formation: influence of compatibilizer
functionality and organoclay modification. Macromol. Mater. Eng. 275, 8 (2000)
79. Teh, P.L., Mohd Ishak, Z.A., Hashim, A.S., Karger-Kocsis, J., Ishiaku, U.S.: Effects of
epoxidized natural rubber as a compatibilizer in melt compounded natural rubber–
organoclay nanocomposites. Eur. Polym. J. 40, 2513 (2004)
80. Varghese, S., Karger-Kocsis, J., Gatos, K.G.: Melt compounded epoxidized natural rubber/
layered silicate nanocomposites: structure-properties relationships. Polymer 44, 3977 (2003)
81. Pal, K., Rajasekar, R., Kang, D.J., Zhang, Z.X., Kim, J.K., Das, C.K.: Effect of epoxidized
natural rubber–organoclay nanocomposites on NR/high styrene rubber blends with fillers.
Mater. Des. 30(10), 4035–4042 (2009)
82. Rajasekar, R., Pal, K., Heinrich, G., Das, A., Das, C.K.: Development of nitrile butadiene
rubber–nanoclay composites with epoxidized natural rubber as compatibilizer. Mater. Des.
30(9), 3839–3845 (2009)
83. Zilg, C., Thomann, R., Mülhaupt, R., Finter, J.: Polyurethane nanocomposites containing
laminated anisotropic nanoparticles derived from organophilic layered silicates. Adv. Mater.
11, 49 (1999)
84. Ganter, M., Reichert, P.: Rubber nanocomposites: morphology and mechanical properties of
BR and SBR vulcanizates reinforced by organophilic layered silicates. Rubber Chem.
Technol. 74, 221 (2001)
85. Zhang, L., Wang, Y., Wang, Y., Sui, Y., Yu, D.: Morphology and mechanical properties of
clay/styrene-butadiene rubber nanocomposites. J. Appl. Polym. Sci. 78, 1873 (2000)
86. Schallamach, A.: In: Bateman, L. (ed.) Chemistry and Physics of Rubber-Like Substances.
Maclaren and Sons Ltd., London, Chapter 3, p. 382 (1963)
230 K. Pal et al.

87. Grosch, K.A., Schallamach, A.: Tire friction on wet roads. Rubber Chem. Technol. 49, 862
(1976)
88. Pal, K., Pal, S.K., Das, C.K., Kim, J.K.: Influence of fillers on NR/SBR/XNBR blends.
Morphology and wear. Tribol. Int. 43(8), 1542–1550 (2010)
89. Thomas, A.G.: Factors influencing the strength of rubbers. J. Polym. Sci. Polym. Symp. 48,
145 (1974)
90. Kragelskii, I.V., Nepomnyashchil, E.F. In: James, D.I. (ed.) Abrasion of Rubber. Maclern
and Sons Ltd., London, Chapter 3 (1967)
91. Schallamach, A.: A theory of dynamic rubber friction. Wear 6(5), 375 (1963)
92. Nayek, S., Bhowmick, A.K., Pal, S.K., Chandra, A.K.: Wear behavior of silica filled tire
tread compounds by various rock surfaces. Rubber Chem. Technol. 78, 705 (2005)
93. Kragelskii, I.V., Nepomnyashchil, E.F.: In: James, D.I. (ed.) Abrasion of Rubber. Maclern
and Sons Ltd., London, Chapter 3 (1967)
94. Schnumann, R., Warlow-Davies, E.: The elastomeric component of the force of sliding
friction. Proc. Phys. Soc. 54(1), 14–27 (1942)
95. Pal, K.: Speciality elastomer blends for abrasion resistant tyre tread of dump-trucks. Ph.D
Thesis, IIT Kharagpur, India (2009)
96. Bhowmick, A.K.: Ridge formation during the abrasion of elastomers. Rubber Chem.
Technol. 55, 1055 (1979)
97. Southern, E., Thomas, A.: Studies of rubber abrasion. Rubber Chem. Technol. 52, 1008
(1979)
98. Medalia, A.I., Alesi, A.I., Mead, J.L., Simonean, R.: Paper no. 34, Rubber Division, ACS,
Cincinnati, Ohio, October, 18–21, 1988; abstract in Rubber Chem Technol 62, 165 (1989)
99. Viswanath, N., Bellow, D.G.: Development of an equation for the wear of polymers. Wear
181–183, 42 (1995)
100. Rymuza, Z.: Wear in polymer micro-pairs. Proceedings of 3rd international conference on
wear of materials 125 (1981)
101. Ratner, S.B.: Connection between the wear resistance of plastics and other mechanical
properties. Sov. Plast. 7, 37 (1964)
102. Lewis, R.B.: Predicting the wear of sliding plastic surfaces. Mech. Eng. 86, 32 (1964)
103. Rhee, S.K.: Wear equation for polymers sliding against metal surfaces. Wear 16, 431 (1970)
104. Lancaster, J.K.: Friction and wear. In: Jenkins, A.D. (ed.) Polymer Sciences. North-Holland
Publishing Co., Amsterdam, Chapter 14 (1972)
105. Atkinson, J.R., Brown, K.J., Dowson, D.: The wear of high molecular weight polyethylene:
part I: the wear of isotropic polyethylene against dry stainless steel in unidirectional motion.
Trans. ASME J. Lubr. Technol. 100, 208 (1978)
106. Schallmach, A.: Abrasion pattern on rubber. Rubber Chem. Technol. 26, 230 (1953)
107. Kragelsky, I.V., Nepomnyashchil, E.F.: Friction wear of polymers. Khimiya 5 (1964)
108. Klitenik, G.S., Ratner, S.B.: Friction wear of polymers. Khimiya 77 (1964)
109. Brodsky, G.I.: Comprehensive evaluation of cord-to-rubber adhesion. Rubber World 190(5),
29–39 (1984)
110. Brodsky, G.I., Reznikovsky, M.M., Sizikov, N.N.: Rezina konstruktsionnyi material
sovremennogo machinostroeniya. Khimiya 118 (1967)
111. Reznikovskii, M.M.: In: James, D.I. (ed.) Abrasion of Rubber. Maclaren, London, p. 41
(1967)
112. Rudakov, A., Kuvshinskii, E.: The mechanism of abrasion of vulcanized rubber. Rubber
Chem. Technol. 37, 291 (1964)
113. Pal, K., Das, T., Rajasekar, R., Pal, S.K., Das, C.K.: Wear characteristics of styrene
butadiene rubber/natural rubber blends with varying carbon blacks by DIN abrader and
mining rock surfaces. J. Appl. Polym. Sci. 111, 348 (2009)
114. Dayantis, J.: The effect of pressure on the determination of the Flory-Huggins v parameter
by vapour pressure measurements. Polymer 33(1), 219–221 (1992)
115. Bhatnagar, S.K., Banerjee, S.: Viscosity and molecular weight of masticated styrene-
butadiene rubber. Rubber Chem. Technol. 38, 961 (1965)
Elastomeric Nanocomposites for Tyre Applications 231

116. Ishida, H., Miller, J.D.: Substrate effects on the chemisorbed and physisorbed layers of
methacryl silane-modified particulate minerals. Macromolecules 17, 1659 (1984)
117. Maiti, S.N., Mohapatro, P.K.: Mechanical properties of i-PP/CaCO3 composites. J. Appl.
Polym. Sci. 42, 3101 (1991)
118. Teh PL, T.: Effects of epoxidized natural rubber as a compatibilizer in melt compounded
natural rubber–organoclay nanocomposites. Eur. Polym. J. 40, 2513 (2004)
119. Sezna, J.A., Pawlowski, H.A., DeConinck, D.: New test results from rotorless curemeters.
Proceeding of 136th meeting of the ACS-rubber division (1989)
120. Yehia, A.A., Ismail, M.N., Hefny, Y.A., Abdel-Bary, E.M., Mull, M.A.: Mechano-chemical
reclamation of waste rubber powder and its effect on the performance of NR and SBR
vulcanizates. J. Elasto Plast. 36, 109 (2004)
121. Manik, S.P., Banerjee, S.: Determination of chemical cross-links in rubbers. Die
Angewandte Makromolekuiare Chemie 6, 171 (1979)
122. Zhu, L., Wool, R.P.: Nanoclay reinforced bio-based elastomers: synthesis and
characterization. Polymer 47, 8106 (2006)
123. Rajasekar, R., Pal, K., Heinrich, G., Das, A., Das, C.K.: Development of NBR-nanoclay
composites with epoxidized natural rubber as compatibilizer. Mater. Des. 30, 3839 (2009)
124. De, D., Maiti, S., Adhikary, B.: Reclaiming of rubber by a renewable resource material
(RRM). III. Evaluation of properties of NR reclaim. J. Appl. Polym. Sci. 75, 1493 (2000)
125. Agag, T., Koga, T., Takeichi, T.: Studies on thermal and mechanical properties of
polyimide–clay nanocomposites. Polymer 42, 3399 (2001)
126. Nielsen, L.E. (ed.): Mechanical Properties of Polymers and Composites. Marcel Dekker,
New York, Chapter 2 (1974)
127. Sharifa, J., Yunus, W.M.Z.W., Dahlan, K.Z.H.M., Ahmad, M.H.: Preparation and properties
of radiation crosslinked natural rubber/clay nanocomposites. Polym Test 24(2), 211–217
(2005)
128. Thomas, S., Kuriakose, B., Gupta, B.R., De, S.K.: Scanning electron microscopy studies on
tensile, tear and abrasion failure of plasticized poly (vinyl chloride) and copolyester
thermoplastic elastomers. J. Mater. Sci. 21, 711 (1986)
129. Maity, M., Khatua, B.B., Das, C.K.: Effect of processing on the thermal stability of the
blends based on polyurethane: part IV. Polym. Degrad. Stab. 72, 499 (2000)
130. Gann, R.G., Dipert, R.A., Drews, M.J.: Flammability. In: Kroschwitz, J.I. (ed) Encyclopedia
of Polymer Science and Engineering, 2nd edn. John Wiley & Sons, Inc., New York, 7:154–
210 (1985)
131. Pal, K., Das, T., Pal, S.K., Das, C.K.: Use of carboxylated nitrile rubber and natural rubber
blends as retreading compound for OTR tires. Polym. Eng. Sci. 48, 2410 (2008)
132. Siriwardena, S., Ismail, H., Ishiaku, U.S.: A comparison of white rice husk ash and silica as
fillers in ethylene–propylene–diene terpolymer vulcanizates. Polym. Int. 50, 707 (2001)
133. Gent, A.N., Hindi, M.: Heat build-up and blowout of rubber blocks. Rubber Chem. Technol.
63, 892 (1988)
134. Medalia, A.I.: Heat generation in elastomer compounds: causes and effects. Rubber Chem.
Technol. 64, 481–492 (1991)
135. Park, D.M., Hong, W.H., Kim, S.G., Kim, H.J.: Heat generation of filled rubber vulcanizates
and its relationship with vulcanizate network structures. Eur. Polym. J. 36, 2429–2436
(2000)
Elastomer Clay Nanocomposites
for Packaging

V. Mittal

Abstract Packaging application of the materials require strong barrier as well as


mechanical properties. Majority of the reports in the literature have studied the
mechanical properties while neglecting the barrier properties and it is generally
assumed that the improvement in mechanical properties leads to automatic
enhancement of barrier performance. However, it may not be true in all the cases as
the permeation properties are more sensitive to interface between the polymer and
filler phases. The incompatibility between the filler surface modification and poly-
mer or the presence of excess surface modification on the filler surface can nega-
tively impact the permeation properties, while the mechanical performance may still
enhance. The current chapter focuses on the two very important elastomeric mate-
rials used in large amounts in packaging industry. Various factors affecting the
barrier performance have been quantified. Use of not only montmorillonite, but also
vermiculite has been presented. The importance of clean filler surface on the
composite properties has also been underlined. The determination of average aspect
ratio as a tool to control or tune the microstructure and properties of the nano-
composites has been shown with the help of finite element computer models.

V. Mittal (&)
Institute of Chemical and Bioengineering, Department of Chemistry
and Bio-Engineering, ETHZ Zurich, 8093, Zurich, Switzerland
e-mail: vikas.mittal@chem.ethz.ch
Present Address:
V. Mittal
Polymer Research, BASF SE, 67056, Ludwigshafen, Germany
V. Mittal
Chemical Engineering Program, The Petroleum Institute, 2533,
Abu Dhabi, United Arab Emirates

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 233


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_9,
Ó Springer-Verlag Berlin Heidelberg 2011
234 V. Mittal

 
Keywords Gas barrier Mechanical properties Delamination Clay Aspect  
ratio 
Intercalation 
Degradation 
Interface 
Polyurethane Epoxy  
 
Montmorillonite Vermiculite Finite element models

1 Introduction

Elastomeric nanocomposites form a significant proportion of all the nanocom-


posites synthesized in general. As is common with polymer nanocomposites based
on polypropylene, polyethylene, etc., majority of the studies on elastomeric
nanocomposites also report enhancements in the mechanical performance of the
polymers. The other properties of immense importance like gas barrier properties
have been generally neglected. It is commonly believed that the improvements in
the mechanical properties lead to automatic reinforcement in other properties too,
but it may not be true in all the cases. The reason for such difference is the
significant dependence of the gas barrier properties on the interface development
between the organic and inorganic phases at nanoscale. Any minor incompatibility
at the interface may not affect the mechanical properties, but it significantly affects
the gas barrier properties as the incompatibility at interface may lead to a source of
increased diffusion though the materials thus negating the whole purpose of
polymer reinforcement with the high aspect ratio inorganic clay platelets.
Polymers find increasing use in packaging applications where the polymer
packaging materials are required to have high strength and high barrier to gas
(oxygen, water vapor, carbon dioxide, aroma, etc.) among other properties. Thus, it
is important to study the gas barrier properties of the elastomeric composites to
establish the factors which affect the barrier performance of these nanocomposites
along with the mechanical performance enhancement. These factors vary with the
nature of the polymer, nature of the surface modification of the inorganic filler as
well as processing conditions used for the synthesis of the nanocomposites.
Understanding these factors can lead to optimal design of the high barrier (and high
strength) organic–inorganic hybrid materials for packaging applications. The
chapter would focus on elastomer polyurethane and epoxy polymers for the syn-
thesis of nanocomposites to achieve high gas barrier materials owing to their
extensive use in the packaging industry mostly as adhesives. Figure 1 shows an
simplified representation of a packaging laminate where the poly(ethylene ter-
epthalate) and polypropylene foils are adhered together by the use of an adhesive
comprising of either epoxy or polyurethane [1]. The foils are chosen by the virtue of

Fig. 1 Representation of a
commercial packaging lami-
nate. Reproduced from [1]
with permission from Nova
Science Publishers
Elastomer Clay Nanocomposites for Packaging 235

the properties expected from them, but the choice of the epoxy or polyurethane
based adhesive is solely to bond together the other foils and it does not contribute
towards the barrier performance of the laminate. By the inclusion of the layered
silicate fillers with high aspect ratio crystalline platelets, the adhesive can also be
functionalized to contribute towards the barrier performance of the laminate and by
virtue of thus functionality, the thickness of the laminate can be reduced or one or
two foils of the laminate can be avoided thus leading to significant material savings.
The elastomeric nanocomposites thus can also be developed as free standing barrier
foils for packaging.

2 Polyurethane Clay Nanocomposites

Polyurethane rubbers or elastomers have many applications in foams, coatings,


adhesives of packaging laminates, fibers, etc. Use of polyurethane for the pack-
aging applications requires the enhancement of both barrier properties as well as
mechanical performance for which the plate like clay particles are added to the
matrix as the thin platelets lead to efficient stress transfer between the organic and
inorganic phases, increase interfacial contacts between the organic and inorganic
phases as well as increase the mean path length of the permeant molecules. Thus
the permeant molecules have to wiggle around the platelets in a random walk
caused by the tortuouity in the path of the permeant molecules owing to the
incorporation of clay platelets.
In one such representative study, elastomeric polyurethane nanocomposites
were studied by incorporating montmorillonite modified with different surface
modifications [2]. The chemical modifications exchanged on the surface of the
montmorillonite were bis(2-hydroxyethyl) hydrogenated tallow ammonium
(Nanofil 804), alkylbenzyldimethylammonium (benzalkonium, Nanofil 32), and
dimethyl dihydrogenated tallow ammonium (Nanofil 15). The different surface
modifications were chosen in order to study the effect of the chemical structure of
the modification and subsequent interactions with the polymer on the composite
microstructure and properties. Especially the modification with hydroxyethyl
moieties was expected to have much better polarity match with the polymer than
the other modifications. The nanocomposites were generated by solvent interca-
lation approach and nanocomposite coatings were drawn on suitable substrates.
Figure 2a shows the X-ray diffractograms of the Nanofil 804 filled polyurethane
nanocomposites with different filler volume fractions. The diffraction pattern of
the pure clay has also been shown. As is clear from the diffractograms, the dif-
fraction peak of the pure filler was shifted to lower angles in the composites thus
indicating intercalation of the polymer in the filler interlayers. However, the
presence of the diffraction peak in all the composites indicates that the filler was
not completely delaminated. Also, the same peak position was observed in the
composites irrespective of the filler amount indicating the interlayers were inter-
calated by similar amount of polymer. As the intensity of the X-ray diffraction
236 V. Mittal

Fig. 2 a X-ray diffracto-


grams of the polyurethane
nanocomposites containing
different volume fractions of
the Nanofil 804 modified
montmorillonite and b TEM
micrograph of the polyure-
thane nanocomposites com-
prising of 2.86 vol% of the
Nanofil 804 modified mont-
morillonite. Reproduced from
[2] with permission from
American Chemical Society

peaks is qualitative in nature owing to the dependence on sample preparation, filler


impurities, orientation, etc., the composite microstructure was also examined with
transmission electron microscopy as shown in Fig. 2b. The clay platelets in the
composites with 2.86 vol% of the Nanofil 804 filler were observed to be exten-
sively exfoliated. Other volume fractions of the Nanofil 804 filler in the com-
posites led to the generation of similar morphology. The composites generated
with other fillers like Nanofil 32 and Nanofil 15 on the other hand had more
intercalated morphology than exfoliated. The increase in the basal plane spacing
observed in the X-ray diffractograms was also the maximum in the case of Nanofil
804 composites, even though the other fillers had initial higher basal spacing. The
interaction of the hydroxyl groups on the surface of the filler platelets were
expected to have caused better interactions with the polyurethane polymer thus
leading to higher extents of filler exfoliation. Even a chemical reaction between
the two can be expected thus chemically tethering the polymer chains with the
filler platelets, but it could not be confirmed spectroscopically.
Elastomer Clay Nanocomposites for Packaging 237

Fig. 3 Relative a oxygen


and b water vapor permeation
through the polyurethane
nanocomposites as a function
of filler volume fraction for
different fillers. Reproduced
from [2] with permission
from American Chemical
Society

Usability of the generated nanocomposites for the packaging applications was


tested by measuring their barrier performance. Figure 3 shows the oxygen and
water vapor permeation through the nanocomposites as a function of filler volume
fraction. The oxygen permeation in the composites with fillers of Nanofil 804 and
Nanofil 32 decreased as a function of filler volume fraction. 30% reduction could
be achieved at a filler fraction of 3 vol%. Though the Nanofil 804 filler was
expected to chemically bind with the polymer thus improving the barrier perfor-
mance, however, the oxygen permeation behavior of Nanofil 804 filled composites
was similar to Nanofil 32 filled composites. It indicated that either the Nanofil 32
238 V. Mittal

filler owing to the presence of benzyl rings in the surface modification also had a
better interaction with the polymer or the chemical reaction in the case of Nanofil
804 filler with the polymer did not take place. More surprising was the result
obtained for Nanofil 15 filled composites, where the permeation was observed to
increase with increasing the filler volume fraction. As the only difference in the
three different systems lies in the nature of the surface modification ionically
bound to the filler platelets, therefore, it is clear that the chemical architecture of
the surface modification did affect the composite microstructure and properties.
The modification in the case of Nanofil 15 is completely non-polar which may not
mix well with the polar polymer chains thus causing incompatibilities at the
interface. These incompatibilities are not detected in X-ray or microscopy but were
observed to affect the interface sensitive barrier properties. The incompatibility
can lead to the generation of voids or areas of increased free volume which are
responsible for the increase in the permeation. It also explains the increase of
permeation as a function of filler volume fraction. The water vapor permeation
behavior through the composites is also depicted in Fig. 3b. Here the permeation
decreased in all the three cases as a function of filler volume fraction and there was
no increase in the case of Nanofil 15. It is owing to the different mode of transport
of water molecules through the polymer matrix. Water molecules interact with
each other as well as with the polymer matrix thus forming big clusters which
cannot pass through the voids generated by the mismatch between the polymer and
filler surface modification. Thus, the water vapor permeation is more of a function
of hydrophobicity of the filler. More hydrophobic the filler is, higher the decrease
in the water vapor permeation through the composites.
Similarly, other studies have highlighted the use of polyurethane composites for
packaging applications. Relative water vapor permeation of a poly(urethane urea)-
dimethyl dehydrogenated tallow ammonium modified montmorillonite composite
was reported to show a decrease of five times at a 6 vol% filler concentration [3].
Oxygen permeation through polyurethane nanocomposites was observed to
decrease by 50 or 15% at 4 wt% of montmorillonite surface treated with hexa-
decylammonium or dodecyltrimethylammonium salts, respectively, and by 35% at
4 wt% Cloisite 25A (dimethyl dehydrogenated tallow, 2-ethylhexylammonium)
[4]. Similarly, water vapor permeation of polyurethane was reported to have
significant decrease up to 20 wt% organically modified montmorillonite concen-
tration by Tortora et al. [5].
Mechanical performance of the polyurethane nanocomposites has been reported
in many studies [6–13]. Figure 4 shows the improvement in the tensile modulus of
the polyurethane nanocomposites as a function of filler fraction [6]. The modulus
increased more than twice the value of pure polymer at a filler loading of 10 wt%.
Apart from tensile modulus, both tensile strength as well as strain at break were
also reported to significantly improve as a function of filler fraction. The authors
suggested that the improvements in tensile modulus and strength can be attributed
to the reinforcing effect of dispersed clay platelets whereas improvement in the
strain at break can be a result of the plasticizing effect of the ammonium ions
ionically bound in the filler interlayers.
Elastomer Clay Nanocomposites for Packaging 239

Fig. 4 Tensile modulus of


the polyurethane nanocom-
posites as a function of filler
weight fraction. Reproduced
from [6] with permission
from American Chemical
Society

Fig. 5 Force-extension
curves for the pure polymer
as well as polymer nano-
composites containing differ-
ent vol% of the filler.
Reproduced from [3] with
permission from American
Chemical Society

Xu et al. [3] also reported force-extension curves for the poly(urethane urea) based
nanocomposites as shown in Fig. 5. The authors reported a significant increase in
modulus and strength of the composites owing to the nano dispersed filler platelets. It
was observed that for 20 wt% composite, the modulus and tensile strength increased
by more than 300 and 30%, respectively. Apart from that, there was no negative
impact on the ductility of the composites as the elongation to break of the 20 wt%
nanocomposite was observed to increase by 50% as compared to pure polymer.

3 Epoxy-Montmorillonite Nanocomposites

Similar to the polyurethane nanocomposites, elastomeric epoxy nanocomposites


were also synthesized in order to present their use for packaging applications.
240 V. Mittal

Fig. 6 Chemical structures


of the filler modifications as
well as epoxy–amine system
used fro the synthesis of
epoxy nanocomposites.
Reproduced from [14] with
permission from American
Chemical Society

The interaction of filler surface modification with the epoxy polymer was also
similarly carried out by using various surface modifications as shown in Fig. 6
[14]. The modifications used were benzyldibutyl(2-hydroxyethyl)ammonium
chloride (Bz1OH), benzylbis(2-hydroxyethyl)butylammonium chloride (Bz2OH),
benzyltriethanolammonium chloride (Bz3OH), and benzyl(2-hydroxyethyl)-
methyloctadecylammonium chloride (BzC18OH), benzyldimethylhexadecylam-
monium chloride (BzC16) and dioctadecyldimethylammonium chloride (2C18).
The modifications differed in extent of polarity by the virtue of different chemical
structures. The modifications with one, two and three hydroxyl groups per mole-
cule were exchanged on the surface in order to observe the effect of the hydroxyl
groups on the resulting composite microstructure as well as properties. Similarly,
modifications also included benzyl groups, combination of benzyl groups with
hydroxyl groups and octadecyl chains. Apart from that, modification with only
octadecyl chains was also used. Similar to the case of polyurethane nanocom-
posites, the hydroxyl group containing modifications were expected to tether the
polymer chains with the filler surface. Apart from that, benzyl groups in the
modification molecules were expected to bring stronger van der Waals forces of
interaction between the filler and the polymer. The long alkyl chains were used to
facilitate better interlayer spacing, which may be of help in intercalating more and
more polymer thus leading to the filler delamination.
X-ray diffraction studies on the filler dispersions in solvent and composites
revealed interesting insights into the systems. The prepolymer was observed to
Elastomer Clay Nanocomposites for Packaging 241

have negative interaction with I, II and III surface modifications owing to mis-
match of polarity. The dispersions of the filler in solvent were observed to have
lowering of the basal plane spacing after the addition of the prepolymer to the
dispersion. The diffractogram in the case of benzyldimethylhexadecylammonium
modified clay suspension in solvent had no diffraction peak indicating complete
exfoliation, but a diffraction peak at 3.55 nm appeared after the addition of pre-
polymer. Similarly for dioctadecyldimethylammonium modified clay, a diffraction
peak for the filler suspension in solvent was observed at 3.69 nm which became
3.54 after the addition of prepolymer. For benzyl(2-hydroxyethyl)methyloctade-
cylammonium modified clay, a decrease from 4.12 to 3.89 nm was similarly
observed. On the other hand, the modifications with hydroxyl groups (especially
one hydroxyl group per molecule) had positive interaction with the prepolymer.
The basal plane spacing of especially Bz1OH and Bz2OH modified fillers was not
affected or increased by the addition of the epoxy prepolymer. These fillers
retained the basal plane spacing values around 1.9 nm before and after the addition
of prepolymer. Composite basal plane spacing values were observed to decrease in
all the cases as compared to the filler suspensions in solvent owing to the evap-
oration of solvent during crosslinking.
As the X-ray diffraction provided only qualitative insight into the system, it was
further synergized with the microscopy investigations. The TEM investigation of the
3.5 vol% composites with IV and III modified montmorillonites is shown in Fig. 7.
The composite with benzyldibutyl(2-hydroxyethyl)ammonium modified clay was
observed to have extensive filler exfoliation, whereas, the composite with benzyl-
dimethylhexadecylammonium modified clay was more intercalated in nature.
Composites with V as filler modification were also exfoliated in nature, whereas
with VI modified clay as filler, the composites had intercalated morphology prob-
ably owing to high extent of polarity of the interlayer which may also not match well

Fig. 7 TEM micrographs of the 3.5 vol% filler epoxy nanocomposites using modifications
a benzyldibutyl(2-hydroxyethyl)ammonium and b benzyldimethylhexadecylammonium. Repro-
duced from [14] with permission from American Chemical Society
242 V. Mittal

Table 1 Oxygen and water


Composite Permeability coefficient Water vapor
vapor permeation through the
(oxygen) [cm3 lm/ transmission rate
3.5 vol% epoxy
(m2 d mmHg)]a [g lm/(m2 d mmHg)]a
nanocomposites
Neat epoxy 2.0 10
Cloisite Na+ 1.6 23
BzlOH 0.77 6.7
Bz20H 0.78 5.8
Bz30H 1.0 7.1
BzC180H 2.2 5.7
BzC16 1.6 5.3
2C18 3.7 6.8
Reproduced from [14] with permission from American Chemical
Society
a
Relative probable error 5%

with the epoxy polymer. The other modifications also led to the generation of the
intercalated nanocomposites. The basal plane spacing values of the fillers modified
with Bz1OH as well as Bz2OH were much smaller than the modifications containing
long octadecyl chains. Even then, the higher extent of filler exfoliation was observed
in composites containing Bz1OH as well as Bz3OH modified fillers. It again con-
firms the importance of the interactions between the organic and inorganic phases in
developing the composite microstructure. The initial higher basal plane spacing is
not as important as matching the compatibility between the phases.
The observations from X-ray diffraction as well as microscopy were also
reflected in the composite properties. Table 1 shows the oxygen and water vapor
barrier properties of the epoxy nanocomposites at 3.5 vol% filler fraction. A plot
between the relative oxygen permeation and the basal spacing of the filler in the
composite led to an indication that the permeation actually increased with the
increase in basal plane spacing, which cannot be true. Thus, there is an absence of
correlation between the two factors indicating that the increase in the basal plane
spacing by the intercalation of polymer in the interlayers is not responsible for the
improvement in the oxygen barrier. It is rather the exfoliated platelets which are
more responsible towards this barrier as the aspect ratio of the platelets is fully
recovered, whereas the intercalated platelets have no increase in the aspect ratio.
The oxygen permeation through the composites containing BzC16 modified filler
was observed to be similar as the composites with untreated montmorillonite. The
composites with 2C18 as well as BzC18OH modified montmorillonites were
observed to have increased oxygen permeation as compared to pure epoxy matrix.
This behavior of incompatibility of the alkyl chains with polar polymer matrix
confirms the X-ray and TEM findings. Similar response was also observed for the
polyurethane nanocomposites. Even the presence of alkyl chain and hydroxyl
groups together in the surface modification had negative influence on the micro-
structure and properties of the composites. Maximum decrease in the oxygen
permeation was observed for the Bz1OH and Bz2OH modified montmorillonites
also indicated earlier by the virtue of positive interaction of the filler modification
Elastomer Clay Nanocomposites for Packaging 243

with the epoxy prepolymer. Increasing the number of OH groups in the surface
modification to three was not very effective as it seemed to have become too polar
for the polymer matrix to mix well. However, the performance of the composites
with Bz3OH modified montmorillonite was much better than the composites with
montmorillonites modified with long alkyl chains as 50% reduction in the oxygen
permeation as compared to pure polymer could still be achieved. The water vapor
permeation through the composites is also represented in Table 1. The various
surface modifications though had significant effect on the oxygen permeation
properties of the nanocomposites, the water vapor permeation was not significantly
affected by the different surface modifications and the decrease in the water vapor
permeation was a function of the hydrophobicity of the filler also observed in the
polyurethane composites. The composite with untreated montmorillonite had a
massive increase in the water vapor permeation owing to its high hydrophilicity.
The composites with surface modified montmorillonites had reduced hydrophi-
licity, thus, were also observed to have lower water vapor permeation. The
composites where the oxygen permeation was observed to deteriorate were also
observed to have improved water vapor resistance owing to the different modes of
diffusion of water vapor through the polymer matrices than oxygen as mentioned
in the case of polyurethane nanocomposites. These results thus direct the optimal
design of epoxy nanocomposites for their use in packaging applications.
Figure 8 also shows the oxygen and water vapor permeation through the epoxy
nanocomposites as a function of filler volume fraction. Composites with benzyl-
dibutyl(2-hydroxyethyl)ammonium or Bz1OH and benzyldimethylhexadecylam-
monium or BzC16 modified montmorillonites have been compared. Both the
composite series were observed to have reduced oxygen permeation as a function
of filler volume fraction, but the composites containing Bz1OH had much more
impressive decrease than the BzC16 composites. The oxygen permeation was
reduced to one fourth at 5 vol% loading of the Bz1OH modified filler, whereas in
the case of BzC16 composites, the oxygen permeation decreased by roughly 20%
at 5 vol% filler loading. The permeation results were also fitted with theoretical
models to generate an idea about the average aspect ratio of the platelets in the
composite. It was observed that the platelets in the case of BzC16 composites had
an average aspect ratio of 50, whereas the platelets in Bz1OH nanocomposites had
an average aspect ratio of 250–300 indicating much higher exfoliation of the filler
which subsequently affects the permeation properties and hence composite
applications. As the interaction of permeant with the polymer matrix defines the
diffusion behavior, the water vapor permeation was different from the oxygen
permeation trends as shown in Fig. 8b. The permeation through the composites
containing BzC16 modified montmorillonite decrease by 50% at 5 vol% loading
and the performance was also superior to the Bz1OH modified fillers.
Epoxy nanocomposites with enhanced mechanical properties have been exten-
sively reported [15–23]. In one such study, Lan and Pinnavaia [15] reported the
epoxy nanocomposites with montmorillonites modified with different chain length
surface modifiers. Tensile properties of the epoxy nanocomposites filled with
varying amounts of octadecyl ammonium modified montmorillonite have been
244 V. Mittal

Fig. 8 a Oxygen and b water


vapor barrier properties of
epoxy nanocomposites as a
function of filler volume
fraction and filler modifica-
tions of benzyldibutyl(2-
hydroxyethyl)ammonium
(Bz1OH) and benzyldimeth-
ylhexadecylammonium
(BzC16). The dotted lines act
as guide. Reproduced from
[14] with permission from
American Chemical Society

detailed in Fig. 9. It was observed that both tensile strength as well as tensile
modulus increased nearly linearly with increasing filler content in the composite.
More than ten times increase in both strength and modulus was observed at 15 wt%
of the filler.
Majority of the reported studies on epoxy nanocomposites have used non-
reactive surface modifications or sometimes modifications with van der Waals
interactions with the polymer matrix have also been employed. The modifications
containing hydroxyl groups have also been exchanged on the filler surface in the
hope of chemical reaction with the epoxy polymer [24], but the reaction could not
be confirmed. In another interesting study, Ma et al. ionically exchanged the
hardener M-xylylenediamine containing two amine groups. One reactive group of
the hardener was converted into a cation and was subsequently ion exchanged on
the surface of the montmorillonite platelets to generate hardener grafted clay as
shown in Fig. 10. The hardener modified clay was then mixed with epoxy pre-
polymer and blended well followed by heating. Following this procedure, the
Elastomer Clay Nanocomposites for Packaging 245

Fig. 9 Tensile strength and


tensile modulus of the epoxy
nanocomposites filled with
varying amounts of octadecyl
ammonium modified mont-
morillonite. Reproduced from
[15] with permission from
American Chemical society

polymer was cured by using 4-aminophenyl sulfone (DDS) to generate completely


disordered nanocomposites. The authors also compared the nanocomposites gen-
erated by using conventional modification of the filler (containing OH groups). X-
ray diffraction peaks were observed in the case of these composite, whereas the
diffractogram had no diffraction signal in the composites generated by hardener
grafting method.

4 Epoxy Vermiculite Nanocomposites

Majority of the studies on nanocomposites have focused on the use of montmo-


rillonite as the fillers. Owing to lower charge density of the montmorillonites, a
larger area per cation is available on the surface of the platelets. As the cross-
section area of the exchanged surface modification molecules is smaller than the
246 V. Mittal

Fig. 10 Representation of the modification of the clay with hardener and subsequent epoxy
crosslinking. Reproduced from [24] with permission from American Chemical Society
Elastomer Clay Nanocomposites for Packaging 247

area available per cation on the surface of the montmorillonite platelets, the sur-
face modifications do not lie straight on the surface. This leads to lower basal plane
spacing in the interlayers. However, if the medium charge density minerals like
vermiculite (0.5–0.8 equiv mol-1) are chosen, they can provide the potential of
obtaining higher basal plane spacing for the same ammonium modification as
compared to montmorillonites, the higher charge density leads to lower area per
cation on the surface, thus, more amount of organic modification can be exchanged
on the same area leading to more upright positioning of the molecules. However,
the higher charge density minerals do not swell easily in water and complete cation
exchange is generally not possible. But, an optimal exchange can still be generated
if the cation exchange process is properly controlled and on occasions also
repeated to enhance the extent of surface coverage with the organic modification.
It was mentioned in the earlier section that it is more the interaction between the
polymer and surface modification than the initial higher basal plane spacing, which
is responsible for the filler exfoliation. In the case of epoxy montmorillonite
nanocomposites, Bz1OH modification though generated little basal plane spacing
increment in the filler, but the composites were extensively exfoliated than the case
of modifications which generated initial higher basal plane spacing because of
incompatibility of these modifications with the prepolymer. However, the inter-
actions are also dependant on the nature of inorganic substrate, thus, it is of
importance to study the epoxy nanocomposites with vermiculite system too along
with comparison with the montmorillonite system [25]. Also, the aspect ratio of
the platelets is higher in case of vermiculite than the montmorillonite which is
further helpful for the application of composite materials for packaging.
The vermiculite platelets in the reported study [25] were modified with surface
modifications: benzyldibutyl(2-hydroxyethyl)ammonium (Bz1OH), benzyldim-
ethylhexadecylammonium (BzC16) and benzyl(2-hydroxyethyl)methyloctadecy-
lammonium (BzC18OH). The exchange with Bz1OH cations was however not
successful and the cations did not exchange the sodium cations present on the
surface of the vermiculite platelets. The other two cations BzC16 and BzC18OH
could be successfully exchanged and the extent of cation exchange was also near
complete. Figure 11 shows the thermogravimetric analysis of the modified ver-
miculite samples. The degradation of the organic matter in the temperature range
of 200–350°C as one peak indicates the absence of any excess modification
molecules which would degrade at lower temperatures. The presence of excess
surface modification molecules in the filler interlayers can have negative impact on
the composite properties, as is explained in Sect. 5.
Table 2 details the X-ray diffraction analysis of the pure filler, filler suspension
in solvent as well as epoxy nanocomposites with 3.5 vol% of filler fraction. Basal
plane spacing of 3.40 and 3.25 nm were observed for BzC18OH and BzC16
modified vermiculite as compared to 1.22 nm for pristine vermiculite. These
values were much higher than the montmorillonites similarly modified (2.06 nm
for BzC18OH and 1.87 nm for BzC16 modification). The suspension the filler in
the solvent led to further enhancement of the basal plane spacing, but complete
exfoliation of the filler was not observed. The addition of the prepolymer to the
248 V. Mittal

Fig. 11 Thermogravimetric
analysis of benzyldimethyl-
hexadecylammonium
(BzC16) and benzyl(2-
hydroxyethyl)methyloctade-
cylammonium (BzC18OH)
modified vermiculite. Repro-
duced from [25] with per-
mission from Sage Publishers

Table 2 Basal plane spacing of the filler, solvent suspensions as well as nanocomposites with
3.5 vol% filler fraction
Filler d-Spacing d-Spacing of filler d-Spacing of filler d-Spacing in
filler powder suspended in DMF suspended in epoxy composite
(nm) (nm) DMF ? epoxy (nm) (nm)
Na-vermiculite 1.22 1.42 1.42 1.29
BzC180H 3.40 3.80 3.80 3.96
BzC16 3.25 3.34 3.53 3.68
Reproduced from [25] with permission from Sage Publishers

suspensions increased the basal plane spacing further though even by the addition
of prepolymer, complete exfoliation of the filler was not achieved. However, it still
indicates the intercalation of prepolymer in the filler interlayers and this behavior
is opposite from that observed in the case of montmorillonite system, indicating
the influence of the inorganic substrate on the interactions between the compo-
nents. The synthesis of the composite after curing of the polymer and removal of
solvent led to further increase in the basal plane spacing indicating more polymer
intercalated in the interlayers during composite synthesis. The basal plane spacing
of the filler in composites filled with montmorillonites was observed to decrease
after the solvent evaporation. However, the presence of the diffraction peak in the
composites indicated again the absence of complete filer exfoliation. The
microscopy analysis on the composites as reported in Fig. 12 for BzC16-ver-
miculite epoxy nanocomposites revealed the presence of mixed morphology where
a part of filler was exfoliated and a part was intercalated to varying extents.
Barrier performance of the nanocomposites revealed interesting differences in
comparison with the montmorillonite nanocomposites (Table 3). The oxygen per-
meation of the epoxy matrix is quite low indicating further decrease in this value is
not very straightforward. The composites with BzC18OH and BzC16 filler modifi-
cations were observed to have oxygen permeation values of 1.5 and 1.4 cm3/
Elastomer Clay Nanocomposites for Packaging 249

Fig. 12 TEM micrograph of


the 3.5 vol% BzC16-vermi-
cukte epoxy nanocomposite.
Reproduced from [25] with
permission from Sage
Publishers

Table 3 Oxygen and water vapor permeation through the 3.5 vol% vermiculite epoxy
nanocomposites
Composite Oxygen permeability Water vapor
coefficienta [cm3 lm/ transmission ratea
(m2 d mmHg)] [g lm/(m2 d mmHg)]
Neat epoxy 2.0 10.0
Na-vermiculite 1.7 37.0
BzC18OH 1.5 7.5
BzC16 1.4 9.7
Reproduced from [25] with permission from Sage Publishers
a
Relative probable error 5%

m2 d mmHg respectively indicating further decrease in the oxygen permeation of


epoxy could be obtained by the nanoscale reinforcement of polymer with vermiculite
platelets. The composites with same modifications for the montmorillonite system
had oxygen permeation values of 2.2 and 1.6 cm3/m2 d mmHg respectively indi-
cating that the vermiculite system had better polymer filler interactions than the
montmorillonite system. The water vapor permeation of the composites was less
affected owing to the residual polarity of the vermiculite fillers after cation exchange.

5 Effect of Excess Surface Modification

The surface modification of the filler platelets is a very important factor which
significantly affects the generation of composite microstructure and hence the
250 V. Mittal

Fig. 13 Representation of
presence of excess surface
modification molecules on
the filler surface. Reproduced
from [26] with permission
from Springer

resulting composite properties. During the surface modification, generally the


modification in excess of the cation exchange capacity of the mineral is added to
ensure complete cation exchange. However, the removal or washing of this excess
of surface modification molecules is not very simple and a number of washing
cycles are necessary to obtain completely clean filler [26]. Commercially treated
organo-montmorillonites have been generally observed to contain excess of
modifier molecules. Figure 13 shows the representation of the presence of excess
surface modification molecules on the filler surface [26]. The excess molecules
form a pseudo bilayer and get trapped between the ionically bound molecules. A
number of studies have reported that the presence of excess and unattached
molecules of surface modification on the filler surface can have a detrimental
effect on the microstructure and properties of the composites [27–29]. Two dif-
ferent phenomena by which these excess molecules may affect the composite
microstructure and properties have been observed: firstly, the unattached mole-
cules present as pseudo bilayer degrade thermally at lower temperature and the
resulting products can further degrade the polymer matrix thus resulting in the
deterioration of the interface between the polymer and filler. This point is more of
concern when high temperature compounding of the filler with the polymer is
involved. Secondly, the excess modification molecules can also specifically
interact with the polymer even at lower or room temperature. In case of epoxy
system, the free ammonium head groups can bond with the epoxy prepolymer thus
inducing system instability and disturbing filler polymer interfacial interactions.
Such interactions specifically impact the interface sensitive permeation properties,
thus, it is important to ensure the clean filler surface in order to eliminate any
deteriorating impact on these properties.
To study the effect of excess surface modification molecules on especially the
barrier properties, elastomer epoxy nanocomposites with dioctadecyldimethy-
lammonium (2C18), bis(2-hydroxyethyl) methylhydrogenatedtallowammonium
(C182OH) and benzylhexadecyldimethylammonium (BzC16) modified montmo-
rillonites were synthesized. Montmorillonites with two different cation exchange
capacities of 680 and 880 leq g-1 were used. Figure 14 shows the thermogravi-
metric analysis of commercially procured and self modified BzC16M680 and
2C18M880 montmorillonites. The presence of two thermal degradation peaks
Elastomer Clay Nanocomposites for Packaging 251

Fig. 14 Thermogravimetric analysis of a commercially modified and b self modified


BzC16M680; c commercially modified and d self modified 2C18M880. Reproduced from [26]
with permission from Springer

between 200 and 300°C in Fig. 14a, c indicates the presence of excess surface
modification molecules whose degradation occurs at lower temperature than the
molecules which are ionically bound to the filler surface. It was further observed
that roughly 15–20% of the total weight loss can be attributed to this excess
material indicating a significant amount of excess in the commercially modified
montmorillonites. The self modified montmorillonites were washed rigorously
following multiple cleaning cycles and the thermogravimetric analysis of two of
these montmorillonites is depicted in Fig. 14b, d. The absence of any low tem-
perature thermal degradation peak or the presence of only one degradation peak in
the region of 200–300°C confirmed the absence of any excess surface modification
molecules indicating the excess of surface modification molecules used during the
exchange reaction could be successfully washed off.
Table 4 shows the oxygen permeation through the 3.5 filler vol% epoxy
nanocomposites synthesized by using commercially modified clays. A value of
2 cm3 lm/(m2 d mmHg) was observed for pure epoxy which was observed to
increase to high values in all the composites indicating the negative impact of
252 V. Mittal

Table 4 Oxygen permeation


Composites Oxygen permeability coefficient
through the 3.5 vol% epoxy
[cm3 lm/(m2 d mmHg)]
nanocomposites synthesized
with commercially modified Neat epoxy 2.0
montmorillonites 2C18M680 4.8
C182OHM680 5.9
BzC16M680 2.9
BzC16M880 2.8
2C18M880 4.6
C182OHM880 5.5
Reproduced from [26] with permission from Springer

Table 5 Oxygen permeation


Composites Oxygen permeability coefficient
through the 3.5 vol% epoxy
[cm3 lm/(m2 d mmHg)]
nanocomposites synthesized
with self modified or washed Neat epoxy 2.0
montmorillonites BzC16M680 1.8
2C18M680 3.9
BzC16M880 1.6
2C18M880 3.7
C182OHM880 (washed) 1.5
Reproduced from [26] with permission from Springer

excess surface modification on the composite properties. Also, no effect of the


chemical architecture of the surface modification molecules as well as cation
exchange capacity of the montmorillonite on the oxygen permeation properties
was observed. Especially high was the oxygen permeation through the
C182OHM680 and C182OHM880 filled composites where the oxygen perme-
ation values were enhanced to 5.9 and 5.5 cm3 lm/(m2 d mmHg) respectively.
Table 5 describes the oxygen permeation of the composites when self treated or
washed montmorillonites were used as filler. Completely different oxygen per-
meation behavior through the composites was observed as compared to com-
mercially modified systems. Oxygen permeation of 1.8 cm3 lm/(m2 d mmHg)
instead of 2.9 cm3 lm/(m2 d mmHg) was observed for BzC16M680 system.
Similarly, oxygen permeation of 1.6 cm3 lm/(m2 d mmHg) was observed for
BzC16M880 instead of 2.8 cm3 lm/(m2 d mmHg) for commercially modified
system. Composites with 2C18 modified fillers though also saw a decrease in the
oxygen permeation when self modified montmorillonites were used, but it was
expected as already mentioned in the earlier section owing to mismatch between
the polymer and filler modification. Remarkable improvement in the oxygen
permeation behavior of C182OHM880 system was observed after washing.
Before washing, very high oxygen permeation of 5.5 cm3 lm/(m2 d mmHg) was
observed, which was reduced to 1.5 cm3 lm/(m2 d mmHg) after washing the
montmorillonites. These results clearly establish the importance of filler cleanli-
ness on the resulting composite properties. The higher cation exchange capacity of
Elastomer Clay Nanocomposites for Packaging 253

the montmorillonite was also observed to perform better than the low capacity
mineral. The more polar surface modification was also observed to have better
oxygen resistance properties owing to better interaction with polymer and hence
higher extent of filler exfoliation. Another important point to mention here is that
though the presence of excess of surface modification molecules in the filler
interlayers impacted the barrier properties significantly, but the composites with
commercially modified montmorillonites or self treated montmorillonites were
observed to have no difference in the microstructure as evaluated with X-ray and
microscopy thus further underlining the sensitive nature of permeation properties.

6 Aspect Ratio and Gas Permeation

As mentioned above, the increase in the aspect ratio of the filler in composite
represents its exfoliation and it is only the exfoliated (and not intercalated)
platelets which have maximum impact on the composite properties. Thus, the
aspect ratio of the platelets in the composite is an important tool which can be used

Fig. 15 a Representation of
a computer model consisting
of 50 randomly placed plate-
lets. The platelets had an
aspect ratio of 50 and filler
volume fraction of 3% and
b cross-section through the
centre of the model. Repro-
duced from [30] with per-
mission from Wiley
254 V. Mittal

Fig. 16 a Comparison
between the experimental gas
permeation values through
epoxy and polyurethane
nanocomposites as a function
of filler volume fraction and
the model predictions and b
influence of misalignment on
the permeation through the
composites. Reproduced from
[30] with permission from
Wiley

to compare the different systems, different surface modifications, or different


synthesis methodologies, etc. it can thus help to tune the microstructure devel-
opment and to design the nanocomposites according to requirement. Determina-
tion of the aspect ratio is not possible from the transmission electron micrographs
owing to significant extent of platelet bending, folding and misalignment. Other
means to obtain average aspect ratio of the filler in the composite are thus required.
Numerical finite element approach has been used to generate the computer
models which can be used to generate information on the average aspect ratio of
the platelets [30]. The computer models comprised of 50 non-overlapping identical
platelets with aspect ratios of 50 or 100, which were randomly distributed and
oriented in a periodic cubic box. The volume fraction of the platelets was varied
between 1 and 5%. Figure 15a shows the representation a computer model con-
sisting of 50 randomly oriented platelets with an aspect ratio of 50 and a filler
volume fraction of 3%. Figure 15b also shows a cross-section through the center of
the model. Figure 16a points towards the comparison of the experimental gas
permeation values through epoxy and polyurethane nanocomposites as a function
of filler volume fraction and the numerical predictions from the models as a
function of the increasing misaligned filler platelets with different aspect ratios.
Elastomer Clay Nanocomposites for Packaging 255

The experimental and predicted values match very well till 3 vol% of the filler in
the composite. At higher filler volume fractions, the comparison was observed to
deteriorate owing to the lower extent of exfoliation experimentally. It was also
observed that in this case, the filler in the polyurethane nanocomposites had higher
aspect ratio than the epoxy nanocomposites resulting from the better interactions
between the polymer and filler phases. Figure 16b also demonstrates the impact of
misalignment on the permeation properties of the composites, the misaligned
platelets were observed to be less resistant to permeation than the aligned platelets.

References

1. Mittal, V.: Barrier properties of composite materials. In: Mittal, V. (ed.) Barrier Properties of
Polymer Clay Nanocomposites. Nova Science Publishers, New York (2009)
2. Osman, M.A., Mittal, V., Morbidelli, M., Suter, U.W.: Polyurethane adhesive
nanocomposites as gas permeation barrier. Macromolecules 36, 9851–9858 (2003)
3. Xu, R., Manias, E., Snyder, A.J., Runt, J.: New biomedical poly(urethane urea)-layered
silicate nanocomposites. Macromolecules 34, 337–339 (2001)
4. Chang, J.H., An, Y.U.: Nanocomposites of polyurethane with various organoclays:
thermomechanical properties, morphology, and gas permeability. J. Polym. Sci. Polym.
Chem. 40, 670–677 (2002)
5. Tortora, M., Gorrasi, G., Vittoria, V., Galli, G., Ritrovati, S., Chiellini, E.: Structural
characterization and transport properties of organically modified montmorillonite/
polyurethane nanocomposites. Polymer 43, 6147–6157 (2002)
6. Wang, Z., Pinnavia, T.J.: Nanolayer reinforcement of elastomeric polyurethane. Chem.
Mater. 10, 3769–3771 (1998)
7. Zilg, C., Thomann, R., Mulhaupt, R., Finter, J.: Polyurethane nanocomposites containing
laminated anisotropic nanoparticles derived from organophilic layered silicates. Adv. Mater.
11, 49–52 (1999)
8. Chen, T.K., Tien, Y.I., Wie, K.H.: Synthesis and characterization of novel segmented
polyurethane/clay nanocomposites. Polymer 41, 1345–1353 (2000)
9. Ma, J., Zhang, S., Qi, Z.: Synthesis and characterization of elastomeric polyurethane/clay
nanocomposites. J. Appl. Polym. Sci. 82, 1444–1448 (2001)
10. Yao, K.J., Song, M., Hourston, D.J., Luo, D.Z.: Polymer/layered clay nanocomposites: 2
polyurethane nanocomposites. Polymer 43, 1017–1020 (2002)
11. Tien, Y.I., Wei, K.H.: Hydrogen bonding and mechanical properties in segmented
montmorillonite/polyurethane nanocomposites of different hard segment ratios. Polymer
42, 3213–3221 (2001)
12. Tien, Y.I., Wei, K.H.: High-tensile-property layered silicates/polyurethane nanocomposites
by using reactive silicates as pseudo chain extenders. Macromolecules 34, 9045–9052 (2001)
13. Tien, Y.I., Wei, K.H.: The effect of nano-sized silicate layers from montmorillonite on glass
transition, dynamic mechanical, and thermal degradation properties of segmented
polyurethane. J. Appl. Polym. Sci. 86, 1741–1748 (2002)
14. Osman, M.A., Mittal, V., Morbidelli, M., Suter, U.W.: Epoxy-layered silicate
nanocomposites and their gas permeation properties. Macromolecules 37, 7250–7257 (2004)
15. Lan, T., Pinnavaia, T.J.: Clay-reinforced epoxy nanocomposites. Chem. Mater. 6, 2216–2219
(1994)
16. Messersmith, P.B., Giannelis, E.P.: Synthesis and characterization of layered silicate-epoxy
nanocomposites. Chem. Mater. 6, 1719–1725 (1994)
17. Lan, T., Kaviratna, P.D., Pinnavaia, T.J.: Mechanism of clay tactoid exfoliation in epoxy-clay
nanocomposites. Chem. Mater. 7, 2144–2150 (1995)
256 V. Mittal

18. Zilg, C., Mulhaupt, R., Finter, J.: Morphology and toughness/stiffness balance of
nanocomposites based upon anhydride-cured epoxy resins and layered silicates. Macromol.
Chem. Phys. 200, 661–670 (1999)
19. Brown, J.M., Curliss, D., Vaia, R.A.: Thermoset-layered silicate nanocomposites. Quaternary
ammonium montmorillonite with primary diamine cured epoxies. Chem. Mater. 12, 3376–
3384 (2000)
20. Zerda, A.S., Lesser, A.J.: Intercalated clay nanocomposites: morphology, mechanics, and
fracture behavior. J. Polym. Sci. Polym. Phys. 39, 1137–1146 (2001)
21. Kornmann, X., Lindberg, H., Berglund, L.A.: Synthesis of epoxy-clay nanocomposites:
influence of the nature of the clay on structure. Polymer 42, 1303–1310 (2001)
22. Kornmann, X., Thomann, R., Mulhaupt, R., Finter, J., Berglund, L.: Synthesis of amine-
cured, epoxy-layered silicate nanocomposites: the influence of the silicate surface
modification on the properties. J. Appl. Polym. Sci. 86, 2643–2652 (2002)
23. Kong, D., Park, C.E.: Real time exfoliation behavior of clay layers in epoxy-clay
nanocomposites. Chem. Mater. 15, 419–424 (2003)
24. Ma, J., Yu, Z.Z., Zhang, Q.X., Xie, X.L., Mai, Y.W., Luck, I.: A novel method for
preparation of disorderly exfoliated epoxy/clay nanocomposite. Chem. Mater. 16, 757–759
(2004)
25. Mittal, V.: Epoxy-vermiculite nanocomposites as gas permeation barrier. J. Compos. Mater.
42, 2829–2839 (2008)
26. Mittal, V.: Effect of the presence of excess ammonium ions on the clay surface on permeation
properties of epoxy nanocomposites. J. Mater. Sci. 43, 4972–4978 (2008)
27. Osman, M.A., Atallah, A., Suter, U.W.: Influence of excessive filler coating on the tensile
properties of LDPE–calcium carbonate composites. Polymer 45, 1177–1183 (2004)
28. Morgan, A.B., Harris, J.D.: Effects of organoclay Soxhlet extraction on mechanical
properties, flammability properties and organoclay dispersion of polypropylene
nanocomposites. Polymer 44, 2313–2320 (2003)
29. Kadar, F., Szazdi, L., Fekete, E., Pukanszky, B.: Surface characteristics of layered silicates:
influence on the properties of clay/polymer nanocomposites. Langmuir 22, 7848–7854 (2006)
30. Osman, M.A., Mittal, V., Lusti, H.R.: The aspect ratio and gas permeation in polymer-layered
silicate nanocomposites. Macromol. Rapid Commun. 25, 1145–1149 (2004)
Elastomeric Nanocomposites for
Biomedical Applications

Nicole Fong, Anne Simmons and Laura Poole-Warren

Abstract Elastomeric nanocomposites are gaining considerable attention as new


materials for biomedical use. Elastomers such as polyesters, polyurethanes, and
silicone rubber are excellent candidates as biomaterials in applications including
tissue engineering due to properties such as their ease of synthesis and chemical
manipulation, biodegradability and biocompatibility. However, when used alone
these elastomers often fail to meet the mechanical and physical demands of the
specific application. Elastomeric nanocomposites are composite materials com-
prising nano-sized reinforcements dispersed throughout the polymer matrix. The
presence of such nanoparticles has been shown to improve the mechanical prop-
erties of the base elastomer as well as decrease their permeability properties
making them more suitable for tissue engineering scaffolds and controlled drug
release among other uses. Thus nanocomposite technology is creating greater
applicability of elastomers for biomedical use, however continued research is
required to better understand their behavior when material components are varied
and the effect of nanoparticles on biological systems.

Abbreviations
NC Nanocomposite
MMT Montmorillonite
CNT Carbon nanotube
QAC Quaternary ammonium compound
XRD X-ray diffraction
TEM Transmission electron microscopy
PHA Polyhydroxyalkanoate

N. Fong, A. Simmons and L. Poole-Warren (&)


Graduate School of Biomedical Engineering, University of New South Wales, Sydney
NSW, 2052, Australia
e-mail: l.poolewarren@unsw.edu.au

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 257


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_10,
Ó Springer-Verlag Berlin Heidelberg 2011
258 N. Fong et al.

PTMC Poly(tetramethylene carbonate)


UHMWPE Ultra high molecular weight polyethylene
PCL Poly(e-caprolactone)
PGS Poly(glycerol sebacate)
FTIR Fourier transform infrared
TGA Thermogravimetric analysis
PHB Poly(3-hydroxybutyrate)
HA Hydroxyapatite
MCNT Multiwalled carbon nanotube
PU Polyurethane
CHX Chlorhexidine diacetate
PDMS Polydimethylsiloxane

1 Introduction

Elastomers have been used in a range of biomedical applications since the 1940s
as they have comparable properties to many tissues in the body, making them
valuable materials in the tissue engineering, orthopaedics, and medical device
fields. Elastomers are defined as polymeric materials with rubber-like properties,
exhibiting an ability to elongate under applied force and return to its original state.
Therefore, they offer the advantage of being easily formed into complex shapes
and chemically manipulated to modulate properties such as compliance, tensile
strength and stiffness [1]. In biomedical applications such as orthopaedics these
capabilities make elastomers invaluable over the use of traditional metallic or
ceramic materials. For example, elastomer–calcium phosphate composites have
demonstrated more favorable elastic moduli to facilitate bone healing than their
traditional metal counterparts. For metallic bone implants, stress-shielding is a
common cause of bone refracture due to difference in elastic modulus between the
bone and metal. In addition to the better mechanical compatibility, the elastomeric
materials allow biological components such as bone marrow stromal cells, or
mesenchymal stem cells to be incorporated into the implant to aid the healing
process [2].
Despite these advantages, elastomers are often deemed inappropriate materials
for certain biomedical applications due to limitations associated with poor diffu-
sion or barrier properties of the pristine polymers and/or their long-term bio-
compatibility or mechanical resilience. For example, biomaterials currently used
for urinary catheters are silicone, latex and silicone-coated latex. Latex catheters
are easy and cheap to manufacture and they produce soft devices which are
comfortable for the patient. The associated compromise is that a thicker wall size
is required for sufficient rigidity. Latex is also associated with allergic reactions
Elastomeric Nanocomposites for Biomedical Applications 259

and is susceptible to infection and encrustation as well as having relatively poor


tissue compatibility [3]. Notwithstanding these shortcomings, latex remains a
common urinary device material due to patient comfort. The alternative, silicone,
does not tend to induce allergic reactions and offers greater rigidity compared to
latex, which allows for thinner walls and hence more efficient bladder drainage.
On the downside, the higher rigidity can cause patient discomfort [4].
Nanocomposite (NC) technology has the potential to address such shortcomings
of elastomers. The addition of nanoparticles within elastomers as well as other
polymer matrices allows for a range of material properties to be modulated,
including barrier, mechanical and thermal properties. This capability makes
elastomeric NCs excellent candidates as biomaterials for medical applications
since the desirable properties of the elastomer can be retained while other material
characteristics may be tailored for the required application. Through careful
selection of an appropriate polymer matrix, and selection and loading of nanof-
illers, mechanical properties that meet requirements of a particular application can
theoretically be obtained.
In this chapter, the most significant elastomeric biomaterials and their NCs
studied to date are reviewed, including their impact on the material properties of
the pristine polymer, and how nanotechnology is creating greater applicability of
elastomers in biomedical applications.

2 Polymer Nanocomposites

The origins of polymer NCs can be traced back to the early 1960s when Blumstein
successfully adsorbed layers of methyl methacrylate onto the surface of mont-
morillonite (MMT) clay nanoparticles [5–7]. However, it was not until 1993 that
the real potential of NCs was discovered and reported by the Toyota Central R&D
Labs in Japan [8–10]. This group pioneered the synthesis of a polymer NC where
homogeneous dispersion of MMT within a Nylon 6 (polycaprolactam) polymer
matrix was achieved. The NC was reported to have significant improvements in
mechanical and thermal properties at low loadings of 2–5% MMT, compared to
pristine Nylon 6 and traditional Nylon 6 composites. At a loading of 4.7% MMT,
tensile strength was seen to increase from 68.6 to 97.2 MPa, Young’s modulus
from 1.11 to 1.87 GPa, flexural strength from 89.3 to 143 MPa, and heat distortion
temperature from 65 to 152°C [11]. These findings attracted attention from
numerous industries ranging from the automotive and aerospace to the packaging
industries, and since then research into the development of polymer NCs,
employing a variety of nanoparticles and matrix polymers for a large scope of
applications, has gained momentum [12–17].
Polymer NCs are essentially comprised of two main components; (1) the
polymer matrix material and (2) a filler component that has at least one dimension
on the nanoscale [18], but in many cases a third organic modifier component is
260 N. Fong et al.

also required to distribute the nanofiller within the polymer matrix. A range of
elastomeric matrices for NC systems have been studied including polyurethane,
poly(e-caprolactone) and polysiloxane. Silicates are the usual nanofillers of choice
in NC systems, although other fillers such as single and multi-walled carbon
nanotubes (CNTs), alumina nanofibres, and carbon black have also been used. The
silicates can be naturally occurring (e.g. montmorillonite) or synthetic (e.g. layered
double hydroxides, fluorohectorite), provided at least one dimension, measures
between 1 and 100 nm [19].
Clay minerals are classified according to their crystal structure and surface
charge. MMT is one of the most frequently used clays for NC synthesis, and is
considered to be a smectite belonging to the family of phyllosilicates, which are
hydrous layered silicates of Al, Mg, Fe, and other elements, and large aspect ratios
[18]. Smectites have a 2:1 layered structure made up of two tetrahedral sheets
filled with silicon (Si4+) atoms edge-shared with a central octahedral sheet
occupied by either aluminium or magnesium hydroxide. A single silicate layer is
formed when the tetrahedral and octahedral sheets combine at the tips of the
tetrahedrons of the silica sheets with hydroxyl layers of the octahedral sheets,
making them in the order of 1 nm thick, with lateral dimensions ranging from
30 nm to several micrometres.
Silicate nanolayers occur in nature in stacked configurations called tactoids.
This, along with their hydrophilic nature makes them particularly incompatible
and difficult to disperse in hydrophobic polymers. Therefore, organic modifiers are
often introduced, whose role in polymer NC systems is to compatibilise the silicate
surface with the polymer. As a result, dispersion of the silicate nanoparticles can
be achieved. For this process, cationic surfactant molecules such as quaternary
ammonium compounds (QACs, e.g. dodecyltrimethyl ammonium, hexadecyl-
trimethyl ammonium, dimethyl ditallow ammonium) are primarily chosen as their
charged head groups can interact with the silicate particles, while their hydro-
phobic hydrocarbon tails facilitate silicate swelling and polymer intercalation
(Fig. 1). Along with QACs, primary amines (e.g. dodecylamine, octadecylamine)
have been the most widely used organic modifiers in NC fabrication.

Fig. 1 QACs such as octadecyltrimethyl ammonium chloride have a monocationic head group
with a hydrophobic tail
Elastomeric Nanocomposites for Biomedical Applications 261

2.1 Polymer Nanocomposite Synthesis

Several methods for polymer-NC synthesis have been described. These include an
in situ polymerisation method, polymer melt intercalation method and solution
cast method [20]. The in situ polymerisation method is a one-step method that was
used in the pioneering studies by the Toyota group [8–10] in the synthesis of
Nylon 6-MMT NCs. The monomer, in this case e-caprolactam, and nanoparticle,
MMT, were combined prior to polymerisation to allow the monomer to intercalate
between the silicate layers. The polymerisation process was then forced to occur
by heating to 250°C, thus resulting in a polymer NC.
Polymer melt intercalation involves heating of the polymer–clay mixture
beyond the melting point of the polymer to enable the polymer chains to move to
intercalate the clay layers. Vaia et al. [21] successfully demonstrated the use of this
method to produce the first polystyrene-MMT NC.
The solution cast method utilises an organic solvent to aid with the penetration
of polymer chains between clay layers. Through the addition of the silicate and
polymer to the solvent, the silicate layers expand and the dissolved polymer can
intercalate between the layers. The subsequent removal of the solvent results in the
polymer NC. Solution casting is commonly used for the manufacture of bio-
medical devices such as catheters, insulation for pacemaker leads and aortic bal-
loons [22] and is a suitable method for these applications as it can produce high
quality products at relatively low processing temperatures and cost. The low
processing temperatures allow for the incorporation of bioactive agents, making
them particularly attractive in medical applications. For NC synthesis, an addi-
tional advantage includes a relatively low cost increase per unit volume for sig-
nificant improvements in material performance, particularly when abundant and
naturally occurring nanoparticles such as silicates are used [18]. For these reasons,
this method has been widely used by researchers in the NC field [23–26].

2.2 Dispersion Morphology

Varying degrees of nanoparticle dispersion are usually achieved in polymer NCs.


These are generally categorized as (1) conventional composite, (2) intercalated
NC, or (3) exfoliated NC, as illustrated in Fig. 2. In conventional composites, the
clay particles, for example, exist in agglomerated stacks or tactoids and there is no
intercalation of polymer between clay layers, whereas intercalated NCs exhibit
some insertion of polymer layers between the clay. Exfoliated NCs are achieved
when the tactoids completely separate and the individual clay layers are dispersed
throughout the polymer matrix in either a uniform or disordered manner with an
average distance determined by clay loading [27].
To optimize the benefits that NCs have to offer with regards to material
properties, the filler particles should be exfoliated throughout the matrix material.
262 N. Fong et al.

Fig. 2 Schematic diagrams of an (a) conventional composite, where nanoparticles remain


agglomerated within the polymer matrix, (b) intercalated nanocomposite, where polymer chains
intercalate between the nanoparticles, and (c) exfoliated nanocomposite, where tactoids com-
pletely separate and individual nanoparticles are dispersed throughout the polymer matrix in
either a uniform or disordered manner with an average distance determined by filler loading

In an exfoliated form, maximal interfacial interactions with the polymer matrix


can be attained due to the large aspect ratio of the clay, and therefore stress transfer
to individual nanoparticles is increased, thereby contributing to the reinforcement
and mechanical improvement of the materials. Exfoliation, however, is difficult to
achieve in most polymer–clay systems since the clays’ preferred state is in stacked
configurations, and being hydrophilic in nature they are intrinsically incompatible
with hydrophobic polymers [28]. A major challenge in NC research then, is to
determine effective ways in the manipulation and synthesis of these materials to
optimize polymer intercalation between the clay layers.

2.3 Nanocomposite Characterization

The characterization of polymer NCs for biomedical applications generally


involves firstly the structural characterization of the materials followed by bio-
logical assays that are appropriate for the specific application. Structural analysis
typically involves X-ray diffraction (XRD), transmission electron microscopy
(TEM) and mechanical testing.

2.3.1 X-Ray Diffraction

Through the analysis of XRD, or scatter, valuable information on the structure of


polycrystalline materials can be obtained. In the case of NC characterization, XRD
results give an indication of the dispersion morphology of the nanoparticles, and
infer information on the orientation of the organic molecules in the intergallery
space.
Elastomeric Nanocomposites for Biomedical Applications 263

XRD operates by directing an incident X-ray beam at the sample, which


diverges from the X-ray source and passes through a divergence slit, to then strike
and be diffracted by the sample. The diffracted beam passes through an antiscatter
and receiving slit to be received by an X-ray detector. The divergence and anti-
scatter slits limit and direct the incident and diffracted beams and cause them to
collimate, thus controlling the area of the sample that is exposed to the X-ray
beam. The electronic detector converts the X-rays into pulses of electrical current
which are processed to provide information about the intensity of the X-rays
entering the detector [29]. Calculations used to determine the basal spacing of
OMS or the average distance between silicate layers in NCs are based on Bragg’s
Law (Eq. 1), where l is the known wavelength of the incident beam, 2h is the
diffraction angle, or the angle between the diffracted beam and the transmitted
beam, and d is the distance between silicate layers.
l ¼ 2d sin h ð1Þ

2.3.2 Transmission Electron Microscopy

TEM provides a means to visually image the internal structure of the NCs. In
polymer NC analysis, TEM is often used to complement XRD findings, as XRD
gives a quantitative indication of average silicate dispersion and TEM can provide
qualitative information on the dispersion morphology of the materials.
TEMs enable imaging of specimens at very high magnification through irra-
diation by electron beams under vacuum. Depending on the electron densities
within the sample material, the incident electrons either pass through the specimen
(transmission electrons), or are scattered (scattered electrons). It is from these
transmission and scattered electrons that high-magnification TEM images are
formed to aid in the identification of NC morphologies [30]. Representative
images of NCs obtained by TEM imaging are shown in Fig. 3.

Fig. 3 Representative TEM images of (left to right) elastomeric microcomposite, intercalated


nanocomposite, partially exfoliated nanocomposite
264 N. Fong et al.

2.3.3 Mechanical Testing

Testing of NC materials to assess their mechanical properties can be performed


using a range of techniques including tensile and compression testing, hardness/
nanoindentation tests and tensile hysteresis and stress relaxation tests [31–34]. Of
these, tensile tests are the most common and widely used method to gain an
understanding of how a material will behave in a finished product. The results
from the measurement of force as a specimen is subjected to deformation at a
constant rate provide information on the ultimate tensile strength, elasticity and
Young’s modulus of a material.
The mechanical properties of NCs can give an indirect indication of silicate
dispersion. Since the presence of the silicate layers within the material alters the
mobility of the polymer chains, changes in their mechanical properties are
observed [35]. In general the inclusion of nanofillers in polymer matrices causes
UTS to increase, particularly at lower inclusion loadings, or to be maintained.
Ultimate strain is seen to decrease in most polymer NC systems and Young’s
modulus is enhanced. It should be noted that these findings are highly dependent
on several factors including the matrix polymer properties and preparatory
methods, which explains variations in results [36].

2.4 Elastomeric Nanocomposites as Biomaterials

Among the aforementioned properties of elastomers that make them excellent for
biomedical use, elastomeric NCs provide additional benefits particularly in terms
of being able to tailor material properties for specific applications. As previously
discussed, the mechanical properties of the NC materials can be modulated
through careful selection of NC components and chemistry, and factors such as
nanoparticle loading and method of synthesis [36]. Barrier properties are influ-
enced in a similar manner [14, 36].
It is well understood that, with the addition of glass or mineral inclusions, an
improvement in barrier properties is experienced in conventional polymer com-
posite materials. However, other important material properties such as mechanical
strength and thermal stability are usually sacrificed due to the weak interfacial
interaction and stress concentrations that occur at the inclusion-polymer interface
[37]. A key goal in the development of polymer NCs is to maintain or enhance the
favorable diffusion properties of composite materials, as well as preserve other
important properties that are sacrificed in traditional composites. In the attempt to
achieve this ideal, the aspect ratio or dispersion of the inclusions is increased to
encourage a higher degree of interfacial interactions and lower stress concentra-
tions at the inclusion sites. In polymer NC systems, it is expected that a longer
diffusive path is afforded by the presence of particles with increased aspect ratio,
since the nanoparticle is assumed to be impermeable, thereby forcing the penetrant
to traverse around them, illustrated in Fig. 4. Such improved barrier properties of
Elastomeric Nanocomposites for Biomedical Applications 265

Fig. 4 Enhanced barrier properties are achieved with better dispersion of the nanoparticle

NC materials make them particularly attractive for biomedical applications such as


controlled drug delivery.
Due to their favorable physical and chemical material properties, the most
common elastomers studied for biomedical applications include polyesters,
polyurethanes, and silicone rubber. The following section reviews current research
into these elastomers and their NCs as biomaterials.

2.5 Polyester Elastomer Nanocomposites

Polyesters commonly used for biomedical applications such as polyglycolide and


polylactide, are limited in their use due to their stiffness and elastic deformation
when exposed to long-term strain, as well as their acidic degradation products [38].
The development of thermoplastic polyester elastomers (TPEE), however, has
helped to overcome these shortcomings whilst maintaining the favourable prop-
erties of the polyester.
TPEEs are multiblock copolyether esters with alternating long or short-chain
oxyalkylene glycols connected by ester linkages. Structurally, they contain
repeating hard and soft segments much like polyurethanes and polyamide elas-
tomers. The hard segments are typically composed of short-chain cyclic ester units
such as tetramethylyene terephthalate and the soft segments are derived from
aliphatic polyether glycols [39].
Due to the favourable low toxicity, appearance, and ease of processing and
sterilizing of traditional polyesters, TPEEs with their additional improved bio-
degradation properties, elasticity and tensile strength, have gained considerable
interest for use in biomedical applications such as controlled drug delivery, sur-
gical sutures and tissue engineering scaffolds [38].
266 N. Fong et al.

An excellent example of the use of polyester elastomers in biomedical appli-


cations is in tissue engineering, where it is desirable for the tissue scaffold to have
mechanical properties similar to that of the tissues they are proposed to replace,
and be suitably compliant to accommodate physical and chemical deformations
while being biocompatible [40]. Tissues have elastomeric properties and thus the
development of elastomeric polymers for tissue engineering are of continuing
interest, with the main biodegradable polyesters currently studied for these
applications being polyhydroxyalkanoates (PHAs) [41, 42] and the more recently
developed poly(trimethylene carbonate) (PTMC) [43, 44] copolymers [45–47],
elastomeric poly(e-caprolactone) (PCL), poly(glycerol sebacate) (PGS) [48, 49]
and poly(diol citrates) [50, 51].
Although elastomeric NCs are attracting much research interest for biomedical
use, studies into NCs of these polyester elastomers are still relatively scarce. A
selection of the few investigations into these elastomers as NC materials for
biomedical use is discussed in this section. It is of note, however, that no current
NC research using PGS has been reported in the literature to date.

2.5.1 Thermoplastic Polyester Elastomer

PHAs have generated much research attention for applications such as controlled
release, wound dressings and cardiovascular tissue engineering, as they are non-
toxic biodegradable polyesters that may be easily chemically manipulated to
modulate mechanical and degradation properties for a range of applications.
Excellent reviews on PHAs and their applications are provided by Philip et al. [52]
and Martin et al. [42].
PHAs are naturally produced from the cells of microorganisms and therefore
have the advantage of being free from any toxic residues that may occur in the
chemical synthesis of other polyesters, yet PHAs are still structurally comparable
to chemically derived polyesters [42].
The development of PHA NCs for biomedical applications is relatively new,
however it has been shown that the presence of dispersed nanoparticles within the
PHA matrix affords improved material properties for medical use. Lim et al. [41]
investigated an elastomeric polyester NC using the most commonly used PHA,
poly(3-hydroxybutyrate) (PHB), and MMT nanoparticles to determine the effects
of the nanoinclusions on polymer properties, in particular their thermal stability.
NCs were solvent cast using PHB/MMT in the ratios 100/0, 97/3, 94/6 and 91/9 by
weight and subsequently characterised using XRD, Fourier transform infrared
(FTIR) spectroscopy and thermogravimetric analysis (TGA). Their results showed
that intercalated PHB/MMT NCs were fabricated using low to moderate nano-
particle loadings, with increased thermal stability relative to pristine PHB. These
results implicated the improved biostability of the materials, thus increasing their
applicability as biodegradable biomaterials.
PHA NCs with nano-hydroxyapatite (HA) reinforcements have also been
studied for more specific orthopedic applications. Chen et al. [53] investigated the
Elastomeric Nanocomposites for Biomedical Applications 267

mechanical properties and bioactivity of a solution cast NC based on bioresorbable


polymer-poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBHV) and nano-sized
hydroxyapatite (HA) for bone replacement. Structural characterization of the
materials were conducted as well as thermal and dynamic mechanical tests.
Results from these studies showed that the HA nanoparticles were well dispersed
within the PHBHC matrix, which was hypothesized to be responsible for the
increased degradation temperature, and thus improved biostability, of the materials
as well as the increased storage modulus. Bioactivity of the PHBHV-HA NCs were
assessed by immersing materials into a simulated body fluid with solution ion
concentrations similar to those of blood plasma as observing the formation of
surface calcium-phosphate. A thin layer of calcium-phosphate was observed on the
NC surface after 30 days, indicating good bioactivity.

2.5.2 Poly(trimethylene carbonate)

Poly(trimethylene carbonate) (PTMC) has been studied as biodegradable surgical


sutures [45], controlled drug delivery vehicles [46, 47], bone cement [54], as
scleral buckles [55] and as tissue engineering scaffolds [56]. Among studies into
PTMC copolymers as tissue engineering scaffolds, Pego et al. [57, 58] demon-
strated that these elastomers show potential in their use in long-term degradable
devices for soft tissue engineering as they are both biocompatible [57] and
mechanically resilient [58].
Ribeiro et al. [59] developed a NC material based on PTMC reinforced with HA
and multiwalled carbon nanotubes (MCNT) for joint repair, where HA was used to
stimulate osteoinduction and the MCNT was used as a reinforcing agent and to
provide mechanical damping. The nanoparticles were incorporated into the
polymer matrix by in situ polymerization and materials were assessed for their
coefficient of friction and surface properties relative to cartilage and the conven-
tionally used biomaterial for joint repair, ultra high molecular weight polyethylene
(UHMWPE). Results suggested that the friction coefficient of PTMC composite
materials were closer to that of cartilage than UHMWPE and surface behavior of
the NCs was also comparable to articular cartilage. However, surface attraction
forces of UHMWPE and the PTMC NCs were found to be lower than that of
natural cartilage, implicating poor lubrication properties.

2.5.3 Elastomeric Poly(e-Caprolactone)

Amsden and colleagues [43, 44] synthesised and characterised biodegradable


thermoset elastomers using a star copolymer of D,L-lactide and e-caprolactone as
potential biomaterials for drug delivery and tissue engineering applications. They
found that the elastomers’ mechanical and degradation properties could be mod-
ulated through different crosslinking densities, and materials with properties
268 N. Fong et al.

comparable to elastin could be produced making them excellent candidates for soft
tissue engineering applications.
Di et al. [60] also demonstrated the possibility of modulating the mechanical
and barrier properties of PCL matrix through incorporation of organically modified
silicate nanoparticle. In their study Young’s modulus and ultimate strength
increased with increased nanoparticle loading and elongation at break decreased
and the materials’ permeability to air was observed to significantly decrease.
In further assessment of the material properties of PCL NCs in more specific
biomedical applications, Kim and Fuchs et al. investigated their biodegradation
and biocompatibility properties for bone regeneration. Kim [61] synthesized and
characterized exfoliated HA/PCL NCs where oleic acid was used as the organic
modifier. Morphological features along with the mechanical properties and cellular
responses compared to traditional HA/PCL composites were examined with results
showing that the NCs had better mechanical properties compared to the traditional
composites and pristine PCL as well as improved cellular interactions, where the
oleic acid was thought to improve osteoblastic response.
Fuchs et al. [62] assessed the suitability and biocompatibility of PCL NCs with
calcium-deficient hydroxyapatite (CDHA) nanoparticles in 11 and 24% PCL
matrices for bone tissue engineering and vascularisation through investigating
their ability to support the cell growth of endothelial cells and primary osteoblasts
under monoculture conditions. Biocompatibility tests were conducted using human
umbilical vein endothelial cells (HUVEC) and primary osteoblasts, with results
showing that the HUVECs on the 24% PCL NCs formed a more homogeneous
endothelial monolayer the materials with 11% PCL content. Osteoblasts remained
viable on all material surfaces, with no clear effect of material composition.
Further, assessment of the formation of pre-vascular structures using co-cultures of
the endothelial cells and osteoblasts suggested that this was time and PCL-loading
dependent as more pre-vascular structures were observed after 1 week on the 24%
PCL NC materials, whereas after 4 weeks of pre-vascular structure formation was
slightly higher on the 11% PCL materials.

2.5.4 Poly(Diol Citrates)

Poly(diol citrates) are a biodegradable elastomer recently developed by Yang and


colleagues [48] to specifically address mechanical requirements of scaffolds for
tissue engineering, particularly for small diameter blood vessels. The elastomer,
poly(1,8-octanediol-co-citric acid) (POC) is non-toxic, and easily synthesised with
mechanical and barrier properties that maybe be modulated. In the characterization
of the POC elastomer, Yang et al. [49] found that the mechanical properties of the
materials could be modified through the selection of diols and the postpolymer-
ization conditions. However it was also determined that the mechanical properties
of the POC elastomer may not be mechanically sufficient as scaffolds for tissues
that experience significant tensile or compressive forces such as cartilage. To
enhance the mechanical properties of the elastomeric poly(diol citrate) Webb et al.
Elastomeric Nanocomposites for Biomedical Applications 269

[63] used nanocomposite technology, incorporating a nanofibrous poly(L-lactic


acid) (PLLA) nanophase to the POC matrix. Results from tensile tests showed an
increase in tensile strength, modulus, and elongation at break mechanical prop-
erties comparable to those of human cartilage, ligaments, and blood vessels.
Additional assessment of the compressive modulus of the NC materials was found
to be similar to human and bovine articular cartilage. Thus these studies demon-
strated that poly(diol citrate)–PLLA NC scaffolds hold much promise for tissue
engineering.

2.6 Polyurethane Nanocomposites

Thermoplastic polyurethanes (PUs), are considered to be one of the most bio-


compatible and haemocompatible materials currently available, and their excellent
physical–mechanical properties such as toughness, fatigue resistance and dura-
bility, and chemical versatility that allows for the incorporation of bioactive
molecules, also make them popular choices in the manufacture of biomedical
devices such as the left ventricular assist device, endotracheal tubes, catheters and
vascular grafts.
PUs are defined by the presence of urethane linkages and their properties can
range from soft elastomeric to hard thermosetting materials. In general, biomedical
PUs have a segmented copolymer chemistry made from a combination of a diis-
ocyanate, macrodiol and a chain extender. Figure 5 shows their characteristic
alternating polydisperse ‘soft’ and ‘hard’ microdomains, which are attributed to
the macrodiol, and the diisocyanate linked with the chain extender, respectively.
Since both segments are chemically incompatible, a phase separated morphology
results, with the ‘soft’ macrodiol phase conferring elasticity, lubricity and softness,
and the ‘hard’ domains providing cohesive strength to the material [64].
The first biomedical PU was presented by Boretos and Pierce [65] who found
that a segmented PU experienced no failure or change in tensile properties after
9 days of implantation in calves. They also discovered that the material remained
essentially free from thrombus formation and emboli when used as atrial cannulas

Fig. 5 Characteristic alter-


nating polydisperse ‘soft’ and
‘hard’ microdomains of
polyurethane, which are
attributed by the macrodiol,
and the diisocyanate linked
with the chain extender,
respectively
270 N. Fong et al.

and vascular tubing in calves for 1 week and also when used in the vena cava of
dogs. Subsequent experiments also showed no tumor induction in response to the
material. A later study by Boretos et al. [66] further confirmed the use of PU as
potential biomaterials, performing particularly well in cardiovascular applications.
Since then, the study of PU for biomedical applications has confirmed that these
materials are one of the best available elastomers for biomedical applications
[67–69] even though they have been found to degrade in vivo via hydrolysis,
oxidation and stress-cracking [70, 71].
PU NCs have been studied for a range of biomedical applications including
infection control [72–74] and controlled drug delivery [75, 76]. Poole-Warren
et al. [72] demonstrated the use of bioactive agents as organic modifiers of MMT
nanoparticles within a poly(ether urethane) matrix to impart biological properties
to the polymer matrix as well as aid dispersion of the MMT throughout. More
specifically, the antibacterial agent chlorhexidine diacetate (CHX) and the
non-antibacterial compound dodecylamine (12CH3) were tested against the path-
ogen Staphylococcus epidermidis, and in platelet and cell adhesion studies
respectively. For both organic modifiers, NCs were solvent cast and were found to
have intercalated to exfoliated structures. PU-CHX-MMT NCs showed an excel-
lent 2-log reduction in adherent bacteria compared to control PU materials, and
PU-12CH3-MMT materials demonstrated a decrease in platelet and fibroblast
adhesion. It was therefore shown that organic modifiers with appropriate structure
may be used to alter biological interactions of PU NCs.
Fong et al. [73] expanded on this research and reported on the use of PU NCs as
antibacterial biomaterials for medical devices such as urinary catheters. It was
shown that molecules with appropriate structure may be dual-functional within a
NC system, behaving as both a silicate dispersant and bioactive agent. More
specifically, Fong et al. showed that the antibacterial dicationic molecule,
chlorhexidine diacetate, produced intercalated to partially exfoliated NCs when
used as an organic modifier and/or additive, and promoted a 2-log drog in
Staphylococcus epidermidis, an opportunistic microorganism significant in medi-
cal device-related infections [77]. Other similar studies have been reported sup-
porting such findings, including a study by Styan et al. [74] who used the
quaternary ammonium compound, EthoquadÒ O/12PG, with an 18 carbon alkyl
chain as the organic modifier.
Another approach to conferring antibacterial properties to PU NCs is through
the addition of nanoparticles with antibacterial properties. Hung and Hsu [68]
demonstrated this technique using silver nanoparticles as the filler component in a
poly(ether urethane) matrix. Their assessment of the biocompatibility and anti-
bacterial properties of the solvent cast NC films suggested that the presence of
silver nanoparticles within a PU matrix improved the elastomer’s biocompatibility
and cell proliferation properties, and reduced macrophage attachment, which may
improve material biostability [68].
Permeability of biomaterials to water and air can potentially lead to problems in
implanted devices. Xu et al. [78] demonstrated the advantages of nanocomposite
technology in significantly decreasing the materials’ permeability characteristics.
Elastomeric Nanocomposites for Biomedical Applications 271

NCs were prepared using poly(urethane urea)s (PUU) and two organically mod-
ified layered silicates, Cloisite 15A (Southern Clay Products) and Nanomer I.30
TC, which use alkyl ammonium compounds dimethyl ditallow ammonium and
octadecylammonium as organic modifiers, respectively. Structural assessment of
the solvent cast NCs suggested that intercalated and exfoliated NCs were formed
and Young’s modulus could be modulated with silicate loading while maintaining
the tensile strength and elasticity of pristine PUU. Permeability studies showed a
five-fold reduction in water vapour permeability at the highest silicate loading
(20 wt%) compared to the unfilled PUU control materials.
Williams et al. [79] studied the potential of PUNCs incorporating multiwalled
carbon nanotubes (CNTs) for electrical interfacing with neural tissue. Traditional
platinum electrodes used for this application are associated with fibrous encap-
sulation which increases the impedance at the neural interface. In the study by
Williams et al., NC technology was hypothesised to afford conductive properties
to the biocompatible PU matrix while maintaining the processibility and other
favourable properties of the PU elastomer. Although homogenous dispersion of the
CNTs throughout the PU matrix was not confirmed, minimal cell growth inhibition
of L929 fibroblasts was observed relative to pristine PU and the conductivity of
films was found to range from being non-conductive at 0% loading of CNTs to
6.55 9 10-2 S/cm at 20 wt% loading suggesting that these NC materials show
potential as electrically conductive biomaterials.

2.7 Silicone Nanocomposites

Silicones are synthetic polymers with a backbone of repeating silicon and oxygen
bonds, referred to as siloxane. Silicon is also typically bound to methyl groups, or
other organic groups. The presence of such organic groups to the inorganic
polymer backbone equips silicone with its unique properties such as thermal
stability and biocompatibility [80].
The most commonly used silicone is polydimethylsiloxane (PDMS), which are
linear polymers with a repeating unit of –(Si(CH3)2)O)–. These polymers can be
transformed into elastomeric silicone structures through cross-linking reactions
with radicals, by condensation or addition-cure reaction. Details of elastomeric
silicone synthesis have been described elsewhere [80].
Silicone rubber has been the most frequently used and widely studied polymeric
biomaterials, for applications including controlled drug release [81], bone bonding
and tissue engineering scaffolds [82] and conductive electrodes for neural stim-
ulation [83]. It is considered to be an excellent elastomer for biomedical purposes
as it is known to be biocompatible with excellent biostability [84] and possess
good elastomeric properties over a wide temperature range, thermal stability,
excellent resistance to biodegradation, oxidation and ageing [1]. Despite these
properties, silicone is associated with poor bioactivity and mechanical properties
that are often insufficient for certain applications. NC technology allows these
272 N. Fong et al.

properties to be modulated through the addition of nano-reinforcements and bio-


active agents.
Meng et al. [85, 86] demonstrated the biofunctional capabilities of PDMS NCs
by incorporating bioactive agents in the organic modification process. More spe-
cifically, PDMS/MMT NCs were synthesized using different antifungal agents as
organic modifiers in the development of NC materials for antibacterial urinary
catheter coatings. In their study using a topical and active allylamine antifungal
agent, terbinafine hydrochloride, as the organic modifier of MMT nanoparticles,
exfoliated NCs with enhanced thermal stability and excellent antifungal properties
against Candida albicans were formed using an in situ polymerization technique
[85]. In another study by Meng et al. [86], the antimicrobial agent chlorhexidine
acetate was used as the organic modifier. Exfoliated NCs were again successfully
formed using similar processing methods and an antibacterial effect was reported
for the NC materials against Staphylococcus aureus and Escherichia coli, whereas
control PDMS materials did not demonstrated any antibacterial activity.
Mechanical properties of the NC materials were improved in the presence of the
nanoparticles in both studies, increasing tensile strength and elongation at break by
up to 232, 370%, respectively.
Han et al. [87] aimed to enhance the bioactivity and bone-bonding properties of
octamethyltrisiloxane for bone regeneration by incorporating HA nanoparticles
into the polymer matrix. Structural characterization of the NCs showed that the
HA nanoparticles were uniformly distributed throughout the silicone rubber matrix
and mechanical tests indicated an improvement in tensile strength and Young’s
modulus as well as maintaining the high elongation of the silicone elastomer on
the addition of HA nanoparticles. In terms of bioactivity the interaction of pre-
osteoblasts with NCs was assessed. The presence of attached osteoblasts confirmed
the biocompatibility of the control silicone and NC materials, and cell density on
the NC materials significantly increased with time relative to pristine silicone, thus
indicating the superior biological properties of the silicone incorporating HA.
Depan et al. [88] combined the favorable biodegradable, biocompatible, non-
toxic, and biofunctional material properties of the naturally occurring polymer,
chitosan (CS), with the excellent properties of PDMS in a CS-PDMS-MMT NC to
overcome the limitations of using CS alone in wound dressing. These limitations
including its solubility only in aqueous acidic solutions and rigid structure could
theoretically be modulated using NC technology, leading to a material with greater
applicability as wound dressings. NC materials were made by the intercalation of
CS between MMT layers, followed by UV-irradiated grafting of PDMS onto CS.
Structural characterization of the NC materials showed that an intercalated NC
was formed with increased thermal stability and hydrophilicity as well as water
absorption properties that could be modulated by MMT loading.
Zhou et al. [89] fabricated a PDMS NC incorporating silver-chitosan
-clay nanoparticles to investigate the material’s antibacterial effect on bacteria
commonly associated with catheter-associated infections; Escherichia coli,
Pseudomonas aeruginosa, Staphylococcus aureus and Candida albicans. Exfoli-
ated NCs were made using a clay to silver-chitosan ratio of 1:1.3 through
Elastomeric Nanocomposites for Biomedical Applications 273

crosslinking at room temperature under vacuum. Excellent antibacterial activity of


the NC materials against E. Coli, P. aeruginosa and S. aureus was observed
compared to AgNO3 and studies of Ag release into artificial urine indicated
retarded rate of release in NC material relative to PDMS-silver-chitosan materials
in the absence of clay.
In another study by Zhou et al. [90], the PDMS film was reinforced with MMT
organically modified using phosphatidyl choline (PC), a phospholipid in eukary-
otes responsible for the fluidity and thickness of eukaryotic membranes. Consid-
ering the origins of PC, its haemocompatible properties make it an excellent
candidate for use in antithrombogenic applications. NCs were made by mixing
PDMS and the PC-MMT nanoparticles in solvent and crosslinking under vacuum
at room temperature. The PC-MMT nanoparticles were believed to be well dis-
persed within the PDMS matrix, and mechanical properties of PDMS were
maintained in the NCs. Haemolysis tests to assess blood compatibility of the NCs
demonstrated lower haemolysis in these materials compared to pristine PDMS,
thus suggesting potential for the PDMS-PC-MMT NCs as antithrombotic
biomaterials.
Another biomedical application in which NC technology may prove to be
beneficial is in transdermal drug delivery using pressure sensitive adhesive (PSA)
systems. Currently, this means of drug delivery is associated with problems with
drug permeation through the protective stratum corneum layer of the skin, and
poor adhesive properties. Shaikh et al. [76] studied the role of a PDMS based NC
as a PSA system for transdermal drug delivery. MMT nanoparticles were prepared
using octadecylamine as an organic modifier, and NCs were subsequently fabri-
cated using a range of nanoparticle loadings (0, 2, 5, and 10 wt%) by solvent
casting. It was hypothesized that the NC materials would offer increased control
over the drug release kinetics and adhesive properties relative to unfilled PDMS
based PSA materials. Results showed that partially exfoliated NCs were achieved,
promoting a more uniform release rate and significantly reduced the initial burst
release from the materials compared to pristine PDMS. Further, assessment of the
NC materials’ shear strength showed more than 200% improvement at the lower
clay loadings, implicating improved adhesive properties.

2.8 Styrenic Elastomer Nanocomposites

Styrenic elastomers are block copolymers containing polystyrene or polyethylene


hard segments, and butadiene, isoprene, butylene or ethylene butylene soft seg-
ments. In 1969, the Shell Oil Company developed the styrenic block copolymer,
Kraton, based on isoprene and butadiene [91]. It was found that these materials
were not feasible as biomaterials due to the poor oxidative stability of these soft
segments. The subsequent introduction of an ethylene–butylene segment to the
copolymers saw an improvement in oxidative stability, leading to the development
of poly (styrene-b-ethylene butylene-b-styrene) (SEBS) block copolymer. As well
274 N. Fong et al.

as SEBS, Poly(styrene-b-isobutylene-b-styrene) (SIBS) block copolymers are


among the most commonly studied styrenic elastomers. SIBS was first reported by
Kennedy in 1989 [92] and has attracted attention in the biomedical field due to its
biostability and biocompatibility [93–96].
The use of SIBS as a biomaterial was first studied by Pinchuk et al. [97] They
reported that the materials possessed superior biostability due to its quaternary
carbon backbone of the polyisobutylene polymer. Since then, they have be used in
a range of biomedical applications including drug eluting coatings for stents [98].
Further details of SIBs as biomaterials can be found in an excellent review by
Renade et al. [99].
Research into styrenic elastomer NCs is still relatively scarce, however studies
that have been conducted on SEBs based NCs have demonstrated similar
improvements to material properties as other polymer-NC systems. More specif-
ically, Hasegawa and Usuki [100] reported on SEBS-MMT NCs made using a
compression moulding technique. The resulting NCs were characterized using
TEM and XRD. Results showed that the polystyrene microsegments of the SEBS
polymer in the NC materials were arranged along the length of the nanoparticle
inclusions, suggesting a level of control of material structure and properties could
be achieved. In terms of structural and mechanical properties, Ganguly et al. [101]
studied SEBS-MMT NCs incorporating a series of organically modified and
commercially available clay nanoparticles, Cloisite 10A and Cloisite 20A. Results
from mechanical and structural characterization studies showed that exfoliated
NCs could be achieved in the SEBS at lower loadings of Cloisite 20A nanopar-
ticles and significantly improved dynamic mechanical properties as well as
superior Young’s modulus and tensile strength was achieved using Cloisite 20A
compared to the other nanoparticles investigated and unfilled SEBS.
To the best of our knowledge, SIBS based NCs have not yet been reported on in
the literature, however given the excellent biostability and biocompatibility
properties of these styrenic polymers, research into SIBS-NCs as biomaterials will
certainly flourish in the near future.

3 Limitations and Future Directions

Current research in elastomeric NCs for biomedical applications shows great


promise in their use in clinical settings due to the favorable material properties that
may be achieved through nanoparticle reinforcement of the pristine polymer.
However inconsistencies in their success for different applications need to be
better understood. This level of understanding is well under way although the
difficulty in this task results from the endless combinations of elastomer chemistry,
nanoparticle type and synthesis technique as well as the specific application for
which the material is destined to be used.
In terms of biocompatibility, many studies have been concerned with the tox-
icity of the NC materials intended for biomedical use and the degradation products
Elastomeric Nanocomposites for Biomedical Applications 275

of their polymers [102–104], however problems facing NCs for biomedical use is
now concerned with how nanoparticles interface with biological systems. Studies
into nanotoxicology are still in their infancy although research to date highlights
the pressing need to understand the fate of nanoparticles in different biological
systems and applications [105, 106]. In a review article by Lanone and
Boczkowski [105], the properties of nanomaterials suggesting potential nanotox-
icology issues in biomedical applications were discussed. In general, the factors
influencing the potential toxicology of nanoparticles include their chemical
composition and surface reactivity, nanoparticle size and shape, and particle
retention time [105, 107].
Of particular interest in current nanotoxicology research is CNTs. In a review
paper by Smart et al. [107] research into the toxicity and biocompatibility of CNTs
in biomedical applications was discussed. Overall, although no nanotoxicity of
CNTs has been confirmed to date, in vivo studies have shown adverse effects of
CNTs on the body. For example CNT lung inflammation and granuloma formation
was observed when CNTs were introduced into the trachea of mice and rats
[108–110]. Results from such studies warrant continued investigation into the fate
of nanoparticles in biological systems to be able to properly understand the
behavior and limitations of NC materials for biomedical applications. Thus, in
addition to the excellent properties of elastomeric NCs that have been evident in
the literature to date, results from a more detailed understanding of the behavior of
nanoparticles within the human body will provide a more complete understanding
of the materials, potentiating a more confident and widespread use of elastomeric
NCs as biomaterials for a range of applications.

References

1. Yoda, R.: Elastomers for biomedical applications. J. Biomater. Sci. Polym. Ed. 9(6),
561–626 (1998)
2. Serrano, M.C., Chung, E.J., Ameer, G.A.: Advances and applications of biodegradable
elastomers in regenerative medicine. Adv. Funct. Mater. 20, 192–208 (2010)
3. Denstedt, J.D., Wollin, T.A., Reid, G.: Biomaterials used in urology: current issues of
biocompatibility, infection, and encrustation. J. Endourol. 12(6), 493–500 (1998)
4. Lawrence, E.L., Turner, I.G.: Materials for urinary catheters: a review of their history and
development in the UK. Med. Eng. Phys. 27, 443–453 (2005)
5. Blumstein, A.: Polymerization of adsorbed monolayers i preparation of the clay–polymer
complex. J. Polym. Sci., Part A: Gen. Pap. 3(7), 2653–2664 (1965)
6. Blumstein, A.: Polymerization of adsorbed monolayers. ii. Thermal degradation of the
inserted polymer. J. Polym. Sci., Part A: Gen. Pap. 3(7), 2665–2672 (1965)
7. Blumstein, A., Billmeyer Jr., F.W.: Polymerization of adsorbed monolayers. iii. Preliminary
structure studies in dilute solution of the insertion polymers. J. Polym. Sci., Part B: Polym.
Phys. 4, 465–474 (1965)
8. Usuki, A., Kawasumi, M., Kojima, Y., Okada, A., Kurauchi, T., Kamigaito, O.: Swelling
behavior of montmorillonite cation exchanged for x-amino acids by e-caprolactam.
J. Mater. Res. 8(5), 1174–1178 (1993)
276 N. Fong et al.

9. Usuki, A., Kawasumi, M., Kojima, Y., Fukushima, Y., Okada, A., Kurauchi, T., Kamigaito,
O.: Synthesis of nylon 6-clay hybrid. J. Mater. Res. 8(5), 1179–1184 (1993)
10. Kojima, Y., Usuki, A., Kawasumi, M., Okada, A., Kurauchi, T., Kamigaito, O.: Synthesis of
nylon 6-clay hybrid by montmorillonite intercalated with e-caprolactam. J. Polym. Sci., Part
A: Polym. Chem. 31, 983–986 (1993)
11. Kojima, Y., Usuki, A., Kawasumi, M., Okada, A., Fukushima, Y., Karauchi, T., Kamigaito,
O.: Mechanical properties of nylon 6-clay hybrids. J. Mater. Res. 8(5), 1185–1189 (1993)
12. Wang, Z., Pinnavaia, T.J.: Hybrid organic–inorganic nanocomposites: exfoliation of
magadiite nanolayers in an elastomeric epoxy polymer. Chem. Mater. 10, 1820–1826
(1998)
13. Lan, T., Pinnavaia, T.J.: Clay-reinforced epoxy nanocomposites. Chem. Mater. 6,
2216–2219 (1994)
14. Yano, K., Usuki, A., Okada, A.: Synthesis and properties of polyimide-clay hybrid films.
J. Polym. Sci., Part A: Polym. Chem. 35, 2289–2294 (1997)
15. Jo, B.W., Park, S.K., Kim, D.K.: Mechanical properties of nano-MMT reinforced polymer
composite and polymer concrete. Constr. Build. Mater. 22, 14–20 (2008)
16. Chang, J.-H., An, Y.U., Cho, D., Giannelis, E.P.: Poly(lactic acid) nanocomposites:
comparison of their properties with montmorillonite and synthetic mica. Polymer 44,
3715–3720 (2003)
17. Hule, R.A., Pochan, D.J.: Polymer nanocomposites for biomedical applications. MRS Bull.
32, 354–358 (2007)
18. Utracki, L.A.: Clay-Containing Polymeric Nanocomposites. Rapra Technology Limited,
UK (2004)
19. Auffan, M., Rose, J., Bottero, J.-Y., Lowry, G.V., Jolivet, J.-P., Wiesner, M.R.: Towards a
definition of inorganic nanoparticles from an environmental, health and safety perspective.
Nat. Nanotechnol. 4(10), 634–641 (2009)
20. Kato, M., Usuki, A.: Polymer-clay nanocomposites. In: Polymer-clay Nanocomposites.
Wiley, New York (2000)
21. Vaia, R.A., Ishii, H., Giannelis, E.P.: Synthesis and properties of two dimensional
nanostructures by direct intercalation of polymer melts in layered silicates. Chem. Mater. 5,
1694–1696 (1993)
22. Lamba, N.M.K., Woodhouse, K.A., Cooper, S.L., Lelah, M.D.: Polyurethanes in
Biomedical Applications. CRC Press, London (1997)
23. Finnigan, B., Martin, B., Halley, P., Truss, R., Campbell, K.: Morphology and properties of
thermoplastic polyurethane nancomposites incorporating hydrophilic layered silicates.
Polymer 45, 2249–2260 (2004)
24. Han, B., Cheng, A.M., Ji, G.D., Wu, S.S., Shen, J.: Effect of organophilic montmorillonite
on polyurethane/montmorillonite nanocomposites. J. Appl. Polym. Sci. 91, 2536–2542
(2004)
25. Chang, J.-H., An, Y.U.: Nanocomposites of polyurethane with various organoclays:
thermomechanical properties, morphology, and gas permeability. J. Polym. Sci., Part B:
Polym. Phys. 40, 670–675 (2002)
26. Krikorian, V., Pochan, D.J.: Poly(L-lactic acid)/layered silicate nanocomposite: fabrication,
characterization, and properties. Chem. Mater. 15, 4317–4324 (2003)
27. Wang, Z., Massam, J., Pinnavaia, T.J. Epoxy-clay nanocomposites. In: Polymer–clay
Nanocomposites. Wiley, New York (2000)
28. LeBaron, P.C., Wang, Z., Pinnavaia, T.J.: Polymer-layered silicate nanocomposites: an
overview. Appl. Clay Sci. 15, 11–29 (1999)
29. Cullity, B.D., Stock, S.R.: Elements of X-Ray Diffraction, 3rd edn. Prentice-Hall, NJ (2001)
30. Horiuchi, S.: Fundamentals of High-Resolution Transmission Electron Microscopy.
Elsevier, Amsterdam (1994)
31. Ozkoc, G., Kemaloglu, S., Quaedflieg, M.: Production of poly(lactic acid)/organoclay
nanocomposite scaffolds by microcompounding and polymer/particle leaching. Polym.
Compos. 31, 674–683 (2010)
Elastomeric Nanocomposites for Biomedical Applications 277

32. Salahuddin, N.A.: Layered silicate/epoxy nanocomposites: synthesis, characterization and


properties. Polym. Adv. Technol. 15, 251–259 (2004)
33. Singh, M.K., Shokuhfar, T., de Almeida Gracio, J.J., Mendes de Sousa, A.C., Da Fonte
Fereira, J.M., Garmestani, H., Ahzi, S.: Hydroxyapatite modified with carbon nanotube-
reinforced poly(methyl methacrylate): a novel nanocomposite material for biomedical
applications. Adv. Funct. Mater. 18(5), 694–700 (2008)
34. Finnigan, B., Casey, P., Cookson, D., Halley, P., Jack, K., Truss, R., Martin, D.: Impact of
controlled particle size nanofillers on the mechanical properties of segmented polyurethane
nanocomposites. Int. J. Nanotechnol. 4, 496–515 (2007)
35. Lai, M., Kim, J.-K.: Effects of epoxy treatment of organoclay on structure, thermo-
mechanical and transport properties of poly(ethylene terephthalate-coethylene naphthalate)/
organoclay nanocomposites. Polymer 46, 4722–4734 (2005)
36. Alexandre, M., Dubois, P.: Polymer-layered silicate nanocomposites: preparation,
properties and uses of a new class of materials. Mater. Sci. Eng. 28, 1–63 (2000)
37. Goodier, J.N.: Concentration of stress around spherical and cylindrical inclusions and flaws.
J. Appl. Mech. 55, 39–44 (1933)
38. Webb, A.R., Yang, J., Ameer, G.A.: Biodegradable polyester elastomers in tissue
engineering. Expert Opin. Biol. Ther. 4(6), 801–812 (2004)
39. Adams, R.K., Hoeschele, G.K., Witsiepe, W.K.: Thermoplastic polyether ester elastomers.
In: Holden, G., Kricheldorf, H.R., Quirk, R.P. (eds.) Thermoplastic Elastomers. Hanser
Verlag, Munich (2004)
40. Wang, Y., Ameer, G.A., Sheppard, B.J., Langer, R.: A tough biodegradable elastomer. Nat.
Biotechnol 20(6), 602–606 (2002)
41. Lim, S.T., Hyun, Y.H., Lee, C.H., Choi, H.J.: Preparation and characterization of microbial
biodegradable poly(3-hydroxybutyrate)/organoclay nanocomposite. J. Mater. Sci. Lett. 22,
299–302 (2003)
42. Martin, D.P., Williams, S.F.: Medical applications of poly-4-hydroxybutyrate: a strong
flexible absorbable biomaterial. Biochem. Eng. J. 16, 97–105 (2003)
43. Younes, H., Bravo-Grimaldo, E., Amsden, B.: Synthesis, characterization and in vitro
degradation of a biodegradable elastomer. Biomaterials 25(22), 5261–5269 (2004)
44. Amsden, B., Wang, S., Wyss, U.: Synthesis and characterization of thermoset biodegradable
elastomers based on star-poly(E-caprolactone-co-D, L-lactide). Biomacromolecules 5,
1399–1404 (2004)
45. Wainstein, M., Anderson, J., Elder, J.S.: Comparison of effects of suture materials on
wound healing in a rabbit pyeloplasty model. Urology 49(2), 261–264 (1997)
46. Kluin, O.S., van der Mei, H.C., Busscher, H.J., Neut, D.: A surface-eroding antibiotic
delivery system based on poly-(trimethylene carbonate). Biomaterials 30(27), 4738–4742
(2009)
47. Zhang, Y., Zhuo, R.-X.: Synthesis and drug release behavior of poly (trimethylene
carbonate)–poly (ethylene glycol)–poly (trimethylene carbonate) nanoparticles.
Biomaterials 26, 2089–2094 (2005)
48. Redenti, S., Neeley, W.L., Rompani, S., et al.: Engineering retinal progenitor cell and
scrollable poly(glycerol-sebacate) composites for expansion and subretinal transplantation.
Biomaterials 30(20), 3405–3414 (2009)
49. Chen, Q.-Z., Bismarck, A., Hansen, U., et al.: Characterization of a soft elastomer
poly(glycerol sebacate) designed to match the mechanical properties of myocardial tissue.
Biomaterials 29, 47–57 (2008)
50. Yang, J., Webb, A.R., Ameer, G.A.: Novel citric acid-based biodegradable elastomers for
tissue engineering. Adv. Mater. 16(6), 511–516 (2004)
51. Yang, J., Webb, A.R., Pickerill, S.J., et al.: Synthesis and evaluation of poly(dio citrate)
biodegradable elastomers. Biomaterials 27, 1889–1898 (2006)
52. Philip, S., Keshavarz, T., Roy, I.: Polyhydroxyalkanoates: biodegradable polymers with a
range of applications. J. Chem. Technol. Biotechnol. 82, 233–247 (2007)
278 N. Fong et al.

53. Chen, D.Z., Tang, C.Y., Chan, K.C., et al.: Dynamic mechanical properties and in vitro
bioactivity of PHBHV/HA nanocomposite. Compos. Sci. Technol. 67, 1617–1626 (2007)
54. Wouter, J.E., Habraken, M., Zhang, Z., et al.: Introduction of enzymatically degradable
poly(trimethylene carbonate) microspheres into an injectable calcium phosphate cement.
Biomaterials 29(16), 2464–2476 (2008)
55. de Vos, S.V.N., Koopmans, S.A., Hooymans, J.M.M., et al.: Poly(1, 3-trimethylene
carbonate) networks as a resorbable scleral buckle. J. Controlled Release 132, e51–e52
(2008)
56. Seal, B.L., Otero, T.C., Panitch, A.: Review of polymeric biomaterials for tissue and organ
regeneration. Mater. Sci. Eng., R 34, 147 (2001)
57. Pego, A.P., Van Luyn, M.J.A., Brouwer, L.A. et al.: In vivo behavior of poly(1,
3-trimethylene carbonate) and copolymers of 1,3-trimethylene carbonate or e-caprolactone:
degradation and tissue response. J. Biomed. Mater. Res., Part A 67A(3), 1044–1054 (2003)
58. Pego, A.P., Poot, A.A., Grijpma, D.W., Feijen, J.: In vitro degradation of trimethylene
carbonate based (co)polymers. Macromol. Biosci. 2(9), 411–419 (2003)
59. Ribeiro, R., Ganguly, P., Darensbourg, D., et al.: Biomimetic study of a polymeric
composite material for joint repair applications. J. Mater. Res. 22(6), 1632–1639 (2007)
60. Di, Y., Iannac, S., Sanguigno, L., Nicolais, L.: Barrier and mechanical properties of
poly(caprolactone)/organoclay nanocomposites. Macromol. Symp. 228, 115–124 (2005)
61. Kim, H.-W.: Biomedical nanocomposites of hydroxyapatite/polycaprolactone obtained by
surfactant mediation. J. Biomed. Mater. Res., Part A 83A(1), 169–177
62. Fuchs, S., Jiang, X., Gotman, I., Makarov, C., Schmidt, H., Gutmanas, E.Y., Kirkpatrick,
C.J. Influence of polymer content in Ca-deficient hydroxyapatite–polycaprolactone
nanocomposites on the formation of microvessel-like structures. Acta. Biomater. 6(8),
3169–3177 (2010)
63. Webb, A.R., Kuma, V.A., Ameer, G.A.: Biodegradable poly(diol citrate) nanocomposite
elastomers for soft tissue eingineering. J. Mater. Chem. 17, 900–906 (2007)
64. Gunatillake, P.A., Martin, D.J., Meijs, G.F., et al.: Designing biostable polyurethane
elastomers for biomedical implants. Aust. J. Chem. 56, 545–557 (2003)
65. Boretos, J.W., Pierce, W.S.: Segmented polyurethane: a new elastomer for biomedical
applications. Science 158, 1481–1482 (1967)
66. Boretos, J.W., Pierce, W.S., Baier, R.E., et al.: Surface and bulk characterization of a
polyether urethane for artificial hearts. J. Biomed. Mater. Res. 9, 327–340 (1975)
67. Simmons, A., Hyvarinen, J., Odell, R.A., et al.: Long-term in vivo biostability of
poly(dimethylsilozane)/poly(hexamethylene oxide) mixed macrodiol-based polyurethane
elastomers. Biomaterials 25, 4887–4900 (2004)
68. Hung, H.S., Hsu, S.-H.: Biological performances of poly(ether)urethane–silver
nanocomposites. Nanotechnology 18, 1–9 (2007)
69. Jörg, M., Schierholz, H., Steinhauser, A., et al.: Controlled release of antibiotics from
biomedical polyurethanes: morphological and structural features. Biomaterials 18(12),
839–844 (1997)
70. Pinchuk, L.: A review of the biostability and carcinogenicity of polyurethanes in medicine
and the new generation of ‘biostable’ polyurethanes. J. Biomater. Sci., Polym. Ed. 6,
225–267 (1994)
71. Anderson, J.M., Hiltner, A., Wiggins, M.J., et al.: Recent advances in biomedical
polyurethane biostability and biodegradation. Polym. Int. 46, 163–171 (1998)
72. Poole-Warren, L.A., Farrugia, B., Fong, N., et al.: Controlling cell-material interactions
with polymer nanocomposites by use of surface modifying additives. Appl. Surf. Sci. 255,
519–522 (2008)
73. Fong, N., Simmons, A., Poole-Warren, L.A.: Antibacterial polyurethane nanocomposites
using chlorhexidine diacetate as an organic modifier. Acta Biomater. 6(7), 2554–2561
(2010)
74. Styan, K., Abrahamian, M., Hume, E., Poole-Warren, L.A.: Antibacterial polyurethane
organosilicate nanocomposites. Key Eng. Mater. 342–343, 757–760 (2007)
Elastomeric Nanocomposites for Biomedical Applications 279

75. Da Silva, G.S., Ayres, E., Orefice, R.L., et al.: Controlled release of dexamethasone acetate
from biodegradable and biocompatible polyurethane and polyurethane nanocomposite.
J. Drug Targeting 17(5), 374–383 (2009)
76. Shaikh, S., Birdi, A., Qutubuddin, S., et al.: Controlled release in transdermal pressure
sensitive adhesives using organosilicate nanocomposites. Ann. Biomed. Eng. 35(12),
2130–2137 (2007)
77. Rupp, M.E., Archer, G.L.: Coagulase-negative staphylococci: pathogens associated with
medical progress. Clin. Infect. Dis. 19, 231–245 (1994)
78. Xu, R., Manias, E., Snyder, A.J., Runt, J.: Low permeability biomedical polyurethane
nanocomposites. J. Biomed. Mater. Res. 64A, 114–119 (2003)
79. Williams, C.M., Nash, M.A., Poole-Warren, L.A.: Electrically conductive polyurethanes for
biomedical applications. In: Nicolau, D.V. (ed.) Biomedical Applications of Micro- and
Nanoengineering II, Proceedings of SPIE, 5651, 254 (2005)
80. Colas, A., Curtis, J.: Silicone biomaterials: history and chemistry and medical applications
of silicones. In: Ratner, B.D., Hoffman, A.S., Schoen, F.J., Lemons, J.E. (eds.) Biomaterials
Science, 2nd edn. Elsevier, Amsterdam (2004)
81. Sutinen, R., Laasanen, V., Paronen, P., Urtti, A.: pH-controlled silicone microspheres for
controlled drug delivery. J. Control. Release 33, 163–171 (1995)
82. Yabuta, T., Bescher, E.P., Mackenzie, J.D., et al.: Synthesis of PDMS-based porous
materials for biomedical applications. J. Sol-Gel. Sci. Technol. 26, 1219–1222 (2003)
83. Keohan, F., Wei, X.F., Wongsarnpigoon, A., et al.: Fabrication and evaluation of conductive
elastomer electrodes for neural stimulation. J. Biomater. Sci. Polym. Ed. 18(8), 1057–1073
(2007)
84. Vondracek, P., Dolezel, B.: Biostability of medical elastomers: a review. Biomaterials 5,
209–214 (1984)
85. Meng, N., Zhou, N.-L., Zhang, S.-Q., Shen, J.: Synthesis and antifugal activities of polymer/
montmorillonite–terbinafine hydrochloride nanocomposite films. Appl. Clay Sci. 46,
136–140 (2009)
86. Meng, N., Zhou, N.-L., Zhang, S.-Q., Shen, J.: Synthesis and antimicrobial activities of
polymer/montmorillonite–chlorhexidine acetate nanocomposite films. Appl. Clay Sci. 42,
667–670 (2009)
87. Thein-Han, W.-W., Shah, J., Misra, R.D.K.: Superior in vitro biological response and
mechanical properties of an implantable nanostructured biomaterial: nanohydroxyapatite–
silicone rubber composite. Acta Biomater. 5, 2668–2679 (2009)
88. Depan, D., Kumar, B., Singh, R.P.: Preparation and characterization of novel hybrid of
chitosan-g-PDMS and sodium montmorillonite. Appl. Biomater. 84B, 184–190 (2008)
89. Zhou, N.-L., Liu, Y., Li, L., et al.: A new nanocomposite biomedical material of polymer/
Clay–Cts–Ag nanocomposites. Curr. Appl. Phys. 7S1:e58–e62 (2007)
90. Zhou, N.-L., Fang, S., Xu, D., et al.: Montmorillonite–phosphatidyl choline/PDMS films: a
novel antithrombogenic material. Appl. Clay Sci. 46, 401–403 (2009)
91. Bishop, E.T., O’Neill, W.P.: Block copolymers for use in blood pumps and oxygenators:
preparation and characterization. In: Hasting, F.W., Harminson, C.T. (eds.) Artificial Heart
Program Proceeding (1969)
92. Kaszas, G., Puskas, J.E., Hager, W.G., Kennedy, J.P., Chen, C.C.: Electron pair donors in
carbocationic polymerizations. III. Carbonation stabilization by external electron pair
donors in isobutylene polymerization. J. Macromol. Sci., Pure Appl. Chem. 25, 1099 (1989)
93. Pinchuk, L., Kahn, J., Martin, J.B., Wilson, G.J.: Transactions of the Sixth World Biomater
Congress, p. 1452 (2001)
94. Pinchuk, L., Nott, S., Schwarz, M., Kamath, K.: US Patent 6 545 097 (2003)
95. Pinchuk, L., Nott, S., Schwarz, M., Kamath, K.: US Patent 20030171496 (2003)
96. Silber, S.: Paclitaxel-eluting stents: are they all equal?: An analysis of six randomized
controlled trials in De Novo Lesions of 3,319 patients. J. Intervent. Cardiol. 16(6), 485–490
(2003)
280 N. Fong et al.

97. Pinchuk, L., Kahn, J., Martin, J.B., Wilson, G.J.: Transactions of the World Biomaterials
Congress, p. 1452 (2001)
98. Feng, D., Higashihara, T., Cheng, G., Cho, J.C., Faust, R.: Block copolymers by the
combination of cationic and anionic polymerizations for biomedical applications.
Macromol. Symp. 245–246(1), 14–21 (2006)
99. Ranade, S.V., Richard, R.E., Helmus, M.N.: Styrenic block copolymers for biomaterial and
drug delivery applications. Acta Biomater. 1, 137–144 (2005)
100. Hasegawa, N., Usuki, A.: Arranged microdomain structure induced by clay silicate layers in
block copolymer-clay nanocomposites. Polym. Bull. 51, 77–83 (2003)
101. Ganguly, A., De Sarkar, M., Bhowmick, A.K.: Thermoplastic elastomeric nanocomposites
from poly [styrene-(ethylene-co-butylene)-styrene] triblock copolymer and clay:
preparation and characterization. J. Appl. Polym. Sci. 100, 2040–2052 (2006)
102. Styan, K.E., Martin, D.J., Poole-Warren, L.A.: In vitro fibroblast response to polyurethane
organosilicate nanocomposites. J. Biomed. Mater. Res. 86A, 571–582 (2008)
103. Da Silva, G.R., Da Silva-Cunha, A., Behar-Cohen Jr., F., et al.: Biodegradation of
polyurethanes and nanocomposites to non-cytotoxic degradation products. Polym. Degrad.
Stab. 95, 491–499 (2010)
104. Dutta, S., Karak, N., Saikia, J.P., Konwar, B.K.: Biocompatible epoxy modified bio-based
polyurethane nanocomposites: mechanical property, cytotoxicity and biodegradation.
Bioresour. Technol. 100, 6391–6397 (2009)
105. Lanone, S., Boczkowski, J.: Biomedical applications and potential health risks of
nanomaterials: molecular mechanisms. Curr. Mol. Med. 6, 651–663 (2006)
106. Fischer, H.C., Chan, W.C.W.: Nanotoxicity: the growing need for in vivo study. Curr. Opin.
Biotechnol. 18, 565–571 (2007)
107. Smart, S., Cassady, A.I., Lu, G.Q., Martin, D.J.: The biocompatibility of carbon nanotubes.
Carbon 44, 1034–1047 (2006)
108. Lam, C.-W., James, J.R., McCluskey, R., Hunter, R.L.: Pulmonary toxicity of single-wall
carbon nanotubes in mice 7 and 90 days after intratracheal instillation. Toxicol. Sci. 77,
126–134 (2004)
109. Warheit, D.B., Laurence, B.R., Reed, K.L., et al.: Comparative pulmonary toxicity
assessment of single-wall carbon nanotubes in rats. Toxicol. Sci. 77(1), 117–25 (2004)
110. Muller, J., Huaux, F., Moreau, N., et al.: Respiratory toxicity of multi-wall carbon
nanotubes. Toxicol. Appl. Pharmacol. 207(3), 221–31 (2005)
Actuators and Energy Harvesters Based
on Electrostrictive Elastomeric
Nanocomposites

Kaori Yuse, Pierre-Jean Cottinet and Daniel Guyomar

Abstract Research and development efforts devoted to electro active polymers


(EAPs) are being actively undertaken today due to the numerous advantages of
these materials. Moreover, from the viewpoint of world-wide ecological tenden-
cies, renewable and clean energy sources turn the heads of not only researchers.
The harvesting or scavenging of ambient energy constitutes an important alter-
native stage. Among various specifics of EAPs, the point of a high strain level is
being considered in order to render EAPs promising materials for actuators or
energy harvesters. Piezo devices are usually used for such applications, but when
comparing their strain level to that of EAPs, i.e., around 0.2% for piezo elements
and easily more than 20% for EAPs, the replacement is highly interesting.
Although the Young modulus is smaller than that of ferroelectric materials, the
potential storage energy remains higher. EAPs thus show much promise, but so far,
the application of high electrical voltages is required and the utilization with other
electrical components is limited. To overcome these drawbacks, an intermediate
material was developed, exhibiting a large strain at reasonable levels of applied
voltage. An investigation of the power harvesting with EAPs was also carried out.

 
Keywords Electroactive polymer (EAP) Dielectric polymer Conductive filler 

Nano-carbon Low-powered actuator

K. Yuse (&), P.-J. Cottinet and D. Guyomar


Université de Lyon, INSA-LGEF, 8 rue de la Physique, 69621 Villeurbanne, France
e-mail: kaori.yuse@insa-lyon.fr

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 281


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_11,
Ó Springer-Verlag Berlin Heidelberg 2011
282 K. Yuse et al.

1 Introduction

1.1 EAPs as Actuator Materials

Today, there are several active studies on polymers capable of being actuated by
induction of electrical fields. These polymers are called electro active polymers
(EAPs), and their most interesting characteristic is their ability to be largely
deformed. EAPs can achieve strains up to several hundred percent. They are thus
of interest to be used as sensors/actuators where large deformations are neces-
sary. As sensor/actuator materials, many kinds of transitions are possible, and
principal types include mechanical to electric, magnetic or thermal energy. These
transitions can be obtained by a large variety of materials, e.g., piezo-elements,
optical fibers, or shape memory alloys (SMAs). They can be used in a wide
range of applications.
As materials for the transition between mechanical and electrical energy, the
piezo element is the most frequent and widely commercialized material due to its
numerous advantages, such as its high precision, high electro-mechanical
responses, and its elevated response speeds. PZT (Pb(Zrx, Ti1-x)O3) appeared in
the 1950s. This is a ceramic material that can obtain much bigger deformations
than those of the already existing piezo element, BaTiO3. Soon after its appear-
ance, it took the place as main sensor/actuator material. Many other components
have also been investigated. With the ecological tendencies of today, lead-free
piezoelectric materials are in demand. Their deformation level is, however, cur-
rently not high enough to replace the ceramics comprising lead. Polymer materials,
such as PVDF, also exist.
Nevertheless, their low strain level around 0.2% is regrettable. With a crystal
structure, the strain level can be made to increase but no further than to approx.
1.7%. For applications of micro-structures, PVDF is an ideal material but not for
applications of, for example, artificial human muscles where an extra large strain is
required.
The shape memory alloy (SMA) is famous for being a large-displacement
sensor/actuator material. Still, its maximum strain is limited to around 8%, and
because of its rapid fatigue, the expected strain is generally lower than this when
the material is employed under cyclic loading. Despite this, SMA has another
specific advantage: it can generate a large force. The response speed is quite small
but it can be salved by miniaturelization. Nonetheless, the difficulty of temperature
control and its large energy requirement need to be improved or solved in order for
it to be discussed for the same application fields as PZT.
EAPs then appeared which can generate extra large deformations. The studies
on EAPs started mainly in robotics and especially for the medical field. The human
muscle can generate around 20% of deformation. An air pump method has been
often used for the replacement but a smooth movement was difficult to create.
EAPs can realize such smooth and flexible actuation with the strain level easily
exceeding 20%. The advantage of EAPs is not only their amazing strain level:
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 283

they are also light-weight, inexpensive, with a convenient power source, fracture
tolerance, softness, and ease of processing and forming. A small margin com-
pensation is another weak point for ceramic piezo elements but EAPs can efface
this factor. The stored potential energy is described by YS2/2. The Young modulus
Y of piezo element is significant, 90–100 GPa (PZT), but with its low strain
S level, 0.2%, the energy ends very small. Instead, the Young modulus of the
polymer is generally low, up to some hundreds MPa by PU [1]. However, its large
strain compensates for the low energy. Compared to shape-memory alloys
(SMAs), EAPs demonstrate a higher response speed, lower density, and easier
control mechanism. Instead of magnetic energy transition, the loss is quite low.
However, the scope of practical applications of EAPs is limited by a low actuation
force, a low mechanical energy density, and a low robustness. Progress toward
actuators being used in robotic applications with performances comparable to
those of biological systems will lead to great benefits [2]. The application field
expands day by day, and there are many domains which have awaited such a
promising large-deformation actuator/sensor material.

1.2 Energy Harvesting and EAP

This chapter presents the use of EAPs as energy harvesting materials. The concept
of energy harvesting is briefly introduced here.
The energy issue has always been a potential threat to society ever since
industrialization. During the past 10 years, much research has been performed on
renewable and clean energy sources in order to gradually eliminate fossil fuels
from megawatt power plants. While megawatt power plants generate electricity
with increasing dimensions, the proliferation of portable or wearable devices are
going toward the other end of the scale. Moore forecasted a doubling in the
number of transistors in a given area of silicon approximately every 2 years. This
prediction, now known as Moore’s Law, has held true for approximately four
decades. As Moore’s Law indicates, a brand new generation of digital devices has
followed the improvements in semiconductor technology. Small size scales,
compact storage densities, minimized energy consumptions and shorter processing
times are all direct results of that evolution. Consequently, the recent progresses in
ultra low-power electronics allow the powering of complex systems using either
batteries or environmental energy harvesting.
The harvesting or scavenging of ambient energy provides an important alter-
native for the paradigm of primary electricity generation strategies and storage-
based supply. Significant research efforts and industrial development have led to
energy harvesting based on piezoelectric materials. This is one of the most
promising solutions for direct power supply and energy storage for low-power
wearable devices.
Initially, the electrical stimulation of polymers produced relatively small strains
or harvesting energies, restricting their practical use. But nowadays, polymers
284 K. Yuse et al.

exhibiting large strains have been elaborated and show great potential and capa-
bilities for the development of practical applications. Active polymers which
respond to electric stimuli, i.e., EAPs, exhibit deformations two to three orders of
magnitude higher than the striction-limited, rigid and fragile electroactive ceramics
(EACs). From a technological point of view, the generator mode of electronic EAPs
is potentially as important as the actuator mode. Actuators are indeed pervasive in
modern technology, yet the critical need for new energy systems, such as genera-
tors, may be of larger significance than the sheer number of possible applications.

2 Variety and Principles of EAP Actuators

Mainly, EAPs used for harvesting energy belong to one of the following two types:
(1) dielectric elastomers and (2) electrostrictive polymers. This section explains
their principles one by one, which have been thoroughly reviewed in the cited
references. Also, some of the most recent developments for certain polymers are
presented, and various applications of active polymers are given.

2.1 Dielectric Elastomer

Figure 1 illustrates the generator mode of dielectric elastomers. The dielectric


elastomer is stretched by an outside mechanical force, as shown in Fig. 1a. In the
stretched state, opposite charges are induced on the opposing compliant electrodes
using a voltage difference. The dielectric elastomer is now allowed to contract
using elastic restoring forces. Note that in the contracted state, shown in Fig. 1b,
the dielectric elastomer is thicker and has a smaller area relative to the stretched
state. Assuming that the charge is constant during the contraction process, the
change in geometry has caused a separation of opposite charges by a greater
distance (the increase in thickness by contraction), and has compressed like
charges into a smaller area (the decrease in area by contraction). Both changes
increased the electrical energy, and hence the stored electrical energy. The

Fig. 1 The principle of dielectric elastomer generators: a stretched (charge at low voltage) and
b contracted (charge at high voltage)
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 285

mechanical work has been converted to a greater amount of stored electrical


energy, and the material has operated as a generator transducer.
The qualitative physics analysis of the generator mode, described above, can be
analyzed electrically in quantitative terms using conventional macroscopic lumped
parameters. Although electrical lumped parameters such as capacitance are more
easily manipulated in the engineering sense, the physics of the dielectric elastomer
generator mode can provide greater insights into more complex phenomena, some
of which are described below.
Parameter models of dielectric elastomer generators typically start, as they do
for the actuator mode, by considering the capacitance of the electroded film. The
capacitance in the stretched or contracted (relaxed) states, can be written as
e0  er  A
C¼ ð1Þ
t
where e0 is the permittivity of free space, er is the relative permittivity of the
material, A is the stretched or contracted area where the opposite electrodes
overlap (called the ‘active area’), and t is the polymer thickness. The electrical
energy of a capacitor We with charge Q can be written according to the following
well-known formula;

1 Q2 1 t  Q2
We ¼  ¼ ð2Þ
2 C 2 ðe0  er  A Þ
The polymer can be described by the constant volume approximation [3]; i.e.,
A  t ¼ Vol ¼ constant ð3Þ

where Vol is the volume of the elastomer. The change in electrical energy dWe,
due to an incremental change in state can be then be written as

Q We
dWe ¼ dQ  2  dA ðconstant volumeÞ ð4Þ
C A

or equivalently,
Q We
dWe ¼ dQ þ 2  dt ðconstant volumeÞ ð5Þ
C t
The term (Q/C)dQ is merely the incremental change in the electrical energy
due to charge flowing onto the film at voltage V = Q/C. This term also includes
any leakage of charge through the film. Equations 4 and 5 are very general,
given the constant volume assumption, but further insight into the generator
behavior and the important parameters can be obtained by examining simplified
cases. Consider a constant charge system with charge Q. Using Eq. 2 directly for
the electrical energy stored in the film, the difference in electrical energy
between stretched and contracted states, i.e., the generated energy Wg, can be
written as
286 K. Yuse et al.

 
1 2 1 1
Wg ¼ Wcontracted  Wstreched ¼ Q  ð6Þ
2 Ccontracted Cstreched
Equation 6 can be rewritten as
     
1 Q2 Cstreched Cstreched
Wg ¼    1 ¼ Wstreched 1 ð7Þ
2 Cstreched Ccontracted Ccontracted
Summarizing the basics of the dielectric elastomer generator mode, an initial
charge is placed on a dielectric elastomer film that is stretched with regard to its area
[4]. The film is then allowed to contract, which in turn further separates opposite
charges and compresses like charges to a smaller area. From a macroscopic lumped
parameter perspective, contraction reduces the capacitance. Using either viewpoint,
if the charge is constant on the system, the contraction causes an increase in the
stored electrical energy by raising the voltage of the charge. This increase in
electrical energy is due to the fact that mechanical forces work against the electric
field pressure. With a suitable electrical circuitry, the increase in stored electrical
energy can be harvested to produce a net generated electrical energy output from the
system. It is implied in this description that the system typically cycles. In other
words, the film is stretched again by mechanical forces, an initial charge and energy
is placed on the film, and this is followed by a new round of contraction.
Dielectric elastomer leakage is also a practical consideration in power genera-
tion. Leakage is a direct loss to the system. The importance of leakage phenomena
depends very much on the dielectric elastomer resistivity and the frequency
operation. Lower frequency operation requires lower leakage losses and higher
polymer resistivities. This is because lower frequencies allow more time for leakage
losses to accumulate relative to the energy-producing cycle. Leakage losses may
influence the choice of polymer, but a good dielectric elastomer can generally be
identified to address this issue for most applications.
Perhaps the biggest advantage of dielectric elastomers is the point that, in
addition to a good dynamic range, they operate best at relatively long strokes and
modest forces. It is difficult to address this part of the design space using con-
ventional smart materials such as piezoelectrics. Moreover, it is a very common
mechanical input available from a number of sources in the environment, such as
human motion and wave power.

2.2 Electrostrictive Polymers

Piezoelectric ceramics (such as PZT) have long been used for mechanical-
to-electrical energy harvesting [5]. However, these materials tend to be stiff and
display limitations in mechanical strain abilities. Consequently, they cannot be
used for many applications requiring low frequencies and large stroke mechanical
excitations (such as human movement). Organic materials, however, are softer and
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 287

more flexible; therefore, the input mechanical energy is considerably higher under
the same mechanical force. Piezoelectric polymers, such as PVDF, unfortunately
have a much lower piezoelectric coefficient compared to piezoelectric ceramic
materials. A study has shown that the energy harvesting is lower than with pie-
zoelectric ceramic bimorphs [6]. Electrostatic-based systems, such as dielectric
elastomers, usually require a very high electrical field intensity (50–120 MV/m) to
achieve a significant energy harvesting [4]. Recent research has shown that
polyurethane, an electrostrictive polymer, is capable of generating strains above
10% under a moderate electrical field (20 MV/m), thus leading to them being
considered as potential actuators. Furthermore, these materials are lightweight,
very flexible, have low manufacturing costs and are easily molded into any desired
shape. Although not very well-known, these materials can also be used for
mechanical-to-electrical energy harvesting.
Electrostriction is generally defined as a quadratic coupling between strain and
electrical field. Assuming a linear relationship between the polarization and the
electrical field, the strain Sij and the electric flux density Di are expressed as
independent variables of the electrical field intensity Ek, El and stress Tkl by the
constitutive relations presented below [7].
(
Sij ¼ Mijkl Ek  El þ sEijkl  Tkl
ð8Þ
Di ¼ eTik  Ek þ 2  Mijkl  El  Tkl

Here, sEijkl is the elastic compliance, Mijkl is the electrical-field-related electro-


striction coefficient, and eTik is the linear dielectric permittivity.
There exist two methods for harvesting energy using an electrostrictive mate-
rial. The first one consists in realizing cycles and the second involves working with
the pseudo-piezoelectric behavior.

2.2.1 Energy harvesting cycles

The method proposed by Liu et al [8] was inspired by the approach for harvesting
energy in the case of dielectric elastomers. The mechanical-to-electrical energy
harvesting in electrostrictive materials can be illustrated by for instance the
mechanical stress/strain and electric field/flux density plots, such as those shown in
Fig. 2. Initially the material presented in Fig. 2 had no applied stress, but then a
stress was applied and the state traveled along path A. The applied stress was

Fig. 2 Energy harvesting


cycle
288 K. Yuse et al.

Fig. 3 Energy harvesting


cycles under constant electri-
cal field conditions as the
material is stressed and
unstressed

subsequently reduced. Due to the change in the electrical boundary conditions, the
contraction path did not follow path A but path B. Both in the mechanical and
electrical planes, the material state traversed a closed loop. In the mechanical
plane, the rotation designated that the clear energy flow went from the mechanical
to the electrical. The area enclosed in the loop of the mechanical and electrical
planes was the same and corresponded to the converted energy density in units of
J/m3.
Ideally the energy harvesting cycle consists of the largest possible loop,
bounded only by the limitations of the material. Liu et al. have analyzed electrical
boundary conditions that can be applied for optimizing the energy harvesting.
They demonstrated that electrostrictive materials have significant electric energy
densities that can be harvested. With the electrical boundary conditions investi-
gated in their work, the best energy harvesting density occurred when the electrical
field in the material was increased from zero to its maximum value at a maximum
stress, and then returned to zero at a minimum stress (Fig. 3). A constant electrical
field E0 existed from state 1 to state 2 as the stress was increased to Tmax. From
state 2 to state 3, the electrical field was increased from E0 to E1, then kept
constant until the stress was reduced from Tmax to 0 from state 3 to state 4. At zero
stress, the electrical field was reduced to E0, returning to state 1. In the dielectric-
field plot, the paths 1–4 and 2–3 were not parallel, which was due to the stress
dependence of the dielectric constant.
Ren et al. [7] investigated a means of using this method for harvesting energy.
An experiment was carried out under quasistatic conditions of 1 Hz. The electric
parameters were E0 = 46 MV/m and E1 = 67 MV/m, and 22.4 mJ/cm3 of energy
was harvested for a transverse strain of 2%. If the piezopolymers or piezoceramics
were used with a conventional energy harvesting scheme, the energy harvesting
was below 5 mJ/cm3 [9].

2.2.2 Energy Harvesting with Pseudo Piezoelectric Behavior

Another way for harvesting energy, using electrostrictive polymers, consists in


working with a pseudo piezoelectric behavior. For this, the electrostrictive poly-
mer was subjected to a DC biased electrical field. As the polymer was not
piezoelectric, it was necessary to induce polarization with a DC bias in order to
obtain the desired pseudo piezoelectric behavior [1, 10, 11].
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 289

Fig. 4 The mechanical configuration of an electrostrictive polymer

An isotropic electrostrictive polymer film contracts along the thickness direc-


tion and expands along the film direction when an electrical field is applied across
the thickness, assuming that only a nonzero stress is applied along the length of the
film (Fig. 4). The constitutive relation, expressed by Eq. 8, can thus be simplified
according to
(
S1 ¼ M31  E32 þ sE11  T1
ð9Þ
D3 ¼ eT33  E3 þ 2  M31  E3  T1

The current induced by the transverse vibration is measured as


Z
oD3
I¼ dA ð10Þ
ot
A

where A corresponds to the area of the EPC. The current produced by the polymer
can thereby be related to the strain and electrical field by
Z "   #
oE3 T 2
2  M31  S1  6  M31  E32 2  M31  oSot1  E3
I¼  e þ þ  dA ð11Þ
ot 33 sE11 sE11
A

where oE3=ot and oS1=ot are the time derivates of the electrical field and strain,
respectively. Since a DC electrical field (Edc) is applied on the sample, so that
oE3 =ot ¼ 0; the short circuit current is therefore given by:
290 K. Yuse et al.

Z
 oS1
I0 ¼ 2  M31  Y  Edc   dA ð12Þ
ot
A

with 1 sE11 ¼ Y: Here, Y is the Young modulus. Assuming a constant strain, the
relation between the displacement and strain S1 in the polymer can be expressed by
Eq. 13.
DL u
S1 ¼ ¼ ð13Þ
L0 L0
Here, DL = u is the amplitude of the transverse displacement, and L0 is the initial
value of the length.
The electric impedance of a polymer vibrating at a given frequency could be
modeled by an equivalent electric circuit. Figure 5 displays the most commonly
adopted form of an electrical scheme, where Cp is the capacitance of the clamped
polymer and Rp(x) is a resistance representing the dielectric losses [11]. Both of
these factors are functions of the frequency caused by the relaxation phenomenon.
1
tan d ¼ ð14Þ
Rp ðxÞ  Cp  x

The first branch from the left part of Fig. 5 is the motional branch, modeled by the
current source I0, given in Eq. 12. This equation can be used to model the
harvested current from vibrations.
A previous study has demonstrated that it was possible to neglect Rp [11]. The
dynamic model of the current can thus be simplified to
ou oV
Ih ¼ aðEdc Þ  þ Cp  ð15Þ
ot ot
with aðEdc Þ ¼ 2M31LYAE
0
dc
and where u is the displacement. This expression is
similar to the typical system of equations in the case of piezoelectricity [12].
This last remark is very interesting. In fact, the nonlinear approach of piezo-
electric energy conversion leads to three different high-performance techniques.
Badel et al. [13] presented one technique with an approach based on a theoretical
model and experimental results. They showed that the harvested power may be

Fig. 5 The equivalent elec-


tric circuit of an electrostric-
tive polymer
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 291

increased by a factor of 8 as compared to the most frequently used techniques.


Ongoing work aims at extending this nonlinear approach to electrostrictive
materials.
Between dielectric elastomers and electrostrictive polymers, the difference in
energy harvesting is not very important. The great advantage of a dielectric
elastomer is that it is possible of have a strain of more than 300%, to be compared
to that of an electrostrictive polymer which is 100%. Nevertheless, these materials
require higher electrical fields, causing an increase in the size of the device as well
as in the loss due to the conversion of the high voltage.

3 Energy Harvesting Applications using EAPs

One of the most important trends in the electronic equipment technology from its
origins has been the reduction in size and the increase in functionality. Nowadays,
small, handheld, though very powerful devices are commercially available and
allow the user to play music, to communicate wirelessly or to compute practically
everywhere or, in other words, ubiquitously. In the next years, new products will
become available, providing vision and other extended functions to the user. The
size of such devices is becoming so small that they are becoming referred to as
wearable instead of portable. They can be integrated in objects for everyday use,
such as watches, glasses, clothes, etc.
EAPs are well suited for harvesting energy from human motion. Natural
muscles, the driving force for human motion, are typically of low frequency and
intrinsically linear; two characteristics where EAPs offer advantages. Human
walking is a good example of an application of energy harvesting using human
motion. Proof-of-principle heel-strike generators have been built using dielectric
elastomer devices. With more development, it will likely become feasible to obtain
up to approximately 1–2 J per step. The upper limit for energy harvesting without
causing discomfort to the walker is estimated at 2–4 W for two shoes, assuming a
typical 1-Hz step frequency (Fig. 6) [14]. This amount of power is adequate for
portable applications such as cell phone battery charging, and emergency locators
for soldiers, as illustrated in Fig. 7. The figure also provides the orders of

Fig. 6 A dielectric elastomer


heel strike generator [14]
292 K. Yuse et al.

Fig. 7 Typical power consumptions [15]

magnitude of the powers consumed by various CMOS (complementary metal


oxide semiconductor) electronic equipment that could be powered by energy
harvesting devices [15].
EAPs have demonstrated excellent performances, and numerous applications
appear feasible, but challenges remain. EAPs appear most advantageous for
applications requiring low or variable frequencies, low cost, and large areas.
Currently, piezoelectric (PZT) materials are the most popular for harvesting
mechanical energy because of their compact configuration and compatibility with
MEMSs (micro-electro-mechanical systems). Nevertheless, there are inherent
limitations such as aging, depolarization, and brittleness. In order to overcome
these limitations, the use of EAPs seems promising and exciting.

4 Examples of Material Development

4.1 Introduction

The development of one type of electrostrictive polymer with conductive fillers is


currently ongoing. The aim is to fabricate an intermediate material between PZT
and actual EAPs, i.e., a novel EAP capable of generating strains larger than those
of PZT (2%) and, at the same time, requiring low electric driving fields
(10–20 MV/m). The present section presents the ongoing investigation as one
example of material development within this field.
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 293

4.2 Drawbacks and Solution

Two problems were considered: (1) a high driving energy and (2) aggregation. As
mentioned above, the strain level of EAPs is quite impressive and it is true that it
has significantly increased the potential of these materials. Nevertheless, one
should note that these values have been obtained under a high applied electric
field, i.e., 120–150 MV/m. Many factors have to be considered when selecting the
actuators; e.g., the response speed, the endurance limit or the generated force. It is
clear that the advantages and drawbacks can depend on the applications. For the
case when the EAPs are intended to be utilized with other electric components, the
high electric field requirement becomes a big obstacle. Since PZT needs only a few
MV/m at maximum, its replacement by EAPs is limited. Consequently, a reduction
of the induced electric field is expected.
As an alternative to using the pure polymer material by itself, some micro- or
nano-sized fillers can be added to the polymer matrix in order to increase its
capacity. Consider the incorporation of ferroelectric fillers, such as PZT particles,
into a polymeric matrix. The relative permittivity er of the matrix is small enough
to be ignored as compared to that of the ferroelectric fillers. Thus, the strain of
such a composite has a direct relation to the permittivity of the fillers. The per-
mittivity of the composite has an exponential relation to the volume of the fillers.
In order for the composite to generate a large strain, a high volume percentage of
fillers is necessary. Here, a notorious problem appears: aggregation. As the volume
percentage of filler particles is raised, aggregation occurs with increasing ease.
Also in the case where the diameter of the fillers is lowered, aggregates form more
easily due to surface effects becoming increasingly significant. The existence of
large aggregates raises the risk of electric break down. Percolation threshold must
thus be considered. Numerous researchers have attempted to solve this problem by
both chemical and mechanical approaches, such as sol-gel processing, melt-mixing
methods, or ultrasonic activation. Although some of these alternatives are quite
effective, they often cause the process to become long and complicated.
To overcome these problems, a first attempt involved reducing the amount of
fillers. This was not possible with ferroelectric fillers but with conductive fillers,
and the concept differs from that of ferroelectric fillers. In the case of the con-
ductive fillers, each particle is believed to become a dipole when an electric field is
induced on the composite film, and a high volume percentage is essential in order
to obtain a reasonable strain. Instead, in the case of conductive filler particles, a
small volume percentage is assumed to be sufficient for generating a reasonable
strain. It is believed that the electrons remain on the surface of each conductive
particle during electric field induction on the composite film since the electric field
is null inside. Consequently, a smaller electric field should be enough to drive such
composite films with conductive fillers. With a low volume percentage of fillers,
the risk of electric short circuiting caused by aggregates should decrease.
In the ongoing study, carbon black was chosen as the filler material. Moreover,
the micellar form of carbon black particles was selected in order to more efficiently
294 K. Yuse et al.

avoid the aggregation problem. These micelles were obtained directly in the form of
a carbon black ink.

4.3 Experimental Setup

The chosen materials were the following: commercially available polyurethane


(PU) granules, and a polyether-type thermoplastic TPU5888 from Estane, used as
the matrix. There were two types of carbon black (CB), of which the first was CB
in micellar form, denoted ‘CB-ink’ here. The diameter of the CB micelles was
around from 15 to 30 nm as obtained from transmission electron microscope
(TEM) observation [16]. For the sake of comparison, simple nano-sized CB par-
ticles, denoted ‘CBP’, were also employed. These CBPs were amorphous carbon
nanopowders obtained from Aldrich, with an average particle size of 30 nm.
N,N-dimethylformamide (DMF) was used as a solvent for PU.
The composite films were prepared by a simple solution casting method
[17–19]. PU granules were completely dissolved in DMF at around 75°C. Subse-
quently, either CB-ink or CBP was added to the solution under non-magnetic
stirring. The volume percentage of fillers did not exceed 2% in either of the two
cases. At a constant temperature, the stirring time varied between 1 and 2 h. To
visualize the dispersion effect, an ultrasonic treatment was carried out on certain
solutions. The application time ranged from 2 to 4 h. The mixed solution was then
poured onto a glass plate and dried at 60°C under air during 15 h. The films, after
being removed from the glass plate, were dried again at 130°C under air. The film
thicknesses varied between 7 and 143 lm. Samples without any additives, denoted
‘Pure PU’, were also prepared for comparison. The films were cut into discs with a
diameter of 25 mm. For electromechanical characterization, gold electrodes (6 mm
in diameter and approximately 20 nm thick) were sputtered onto the two surfaces.
The field-induced thickness strain was measured with a double beam laser
interferometer (Agilent 10889B) with a precision on the order of 10 nm. Micro-
order strain could be measured with more precision with two beams as opposed to
just a single one. The corresponding bipolar voltage was delivered by an Agilent
33250A function generator, amplified by a factor 1000 through a high-voltage
lock-in Trek 10/10B amplifier.

4.4 Results

4.4.1 Dispersion

The pure PU sample was completely transparent. If care was not taken, numerous
aggregates were easily formed with the CBP addition. This film was basically
transparent but many black aggregates could be visualized with the naked eye, as
shown in the photographs of Fig. 8a. After ultrasonic activation, large aggregates
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 295

Fig. 8 Photographs of CB-


ink/PU samples: a without
aggregates and b with
aggregates

were no longer observed. The well-mixed specimen was completely black and the
color deepened as the percentage of CB increased. Figure 8b shows a photograph
of the CB-ink/PU sample. In this case, even without ultrasonic treatment, no large
aggregations could be observed. It was thus possible to readily avoid large
aggregates by using CB-ink in micellar form.
Figure 9 shows SEM micrographs of the CB-ink/PU samples: (a) and (b)
display the samples with 1.7 vol. % of CB-ink, and (c) and (d) present specimens with
0.1 vol. %. The micrographs in Fig. 9a, c demonstrated very good dispersions (some
specks of dusts were present on the surfaces), with the exception of a few blobs as the
one encircled in Fig. 9a. Figure 9b shows a close-up of this micaceous aggregate. As
can be seen in Fig. 9c, the existence of these blobs was very rare in the 0.1-vol. %
specimens. Nevertheless, Fig. 9d shows a close-up a one of these rare blobs.
These micaceous aggregates were formed with small particle-like spots.
Although difficult to properly visualize in the printed photos, the surfaces of these
spots did not seem attached to each other. Rather, interfaces were observed. This

Fig. 9 SEM micrographs of


CB-ink/PU samples: a and
b samples with 1.7 vol. % of
CB-ink; c and d samples with
0.1 vol. % CB-ink
296 K. Yuse et al.

signifies that they were some kind of primly particles. The aggregates here were
different from those that are often found in polymer composites with particles, as
shown in Fig. 8b. In this case, the particles were attached to each other. These
micaceous aggregations with CB-ink were denoted ‘‘clusters’’ in order to differ-
entiate between them. It is however unclear whether the electrons remained on each
individual surface of the particles in these clusters or on the surface of the whole
cluster. Since the electric break down did not occur easily, it was determined that
proper electrical contacts could not be made at such low volume percentages.

4.4.2 Overview of Typical Electrical Field-Inducted Measurement Results

The results of the pure PU specimens and those with addition of CB showed the
same tendency. First, typical results are shown after which the differences will be
discussed. Some figures also display simulated data for the sake of comparison.
This is discussed in the next section.
The induced electrical field corresponded to 2 cycles of a triangle-shaped wave
of 0.1 Hz. Each field started from zero and the maximum was varied as 1, 2, 3, 4,
5, 7, 8, 9, 10, 15 and 20 MV/mm. Figure 10 shows the typical result of each strain
S behavior versus time. An image of the electrical field is presented for compar-
ison. Most of the time, the more the maximum induced electrical field was raised,
the more the maximum strain increased, but not always. The convergence of the
maximum strain appeared with certain values of the electrical field. Figure 11
displays the maximum strain as a function of the maximum applied electrical field.
The one shows an example of convergence and the other displays non-conver-
gence. The convergence and decrease of the strain can be clearly seen. If the
amplitude of the electrical field is increased, it is estimated that also these spec-
imens would demonstrate the convergence. The reason for this will be explained
later on. Figure 12 illustrates the typical result of the polarization P as a function
of the electric field. The curves are closed and present an almond shape. The
appearance of a hysteresis can be seen.

Fig. 10 Typical strain


behaviors versus time under 2
cycles of a triangle-shaped
electrical field
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 297

Fig. 11 Typical results of


the maximum strain versus
the maximum applied elec-
trical field

Fig. 12 A typical result of


polarization versus the
applied electrical field

The typical change of the current and polarization when increasing the
maximum electrical field is shown in Fig. 13. The data was obtained at (a)
1 MV/m, (b) at 5 MV/m and (c) at 20 MV/m of the maximum electrical field
induction.
Figure 14 displays the results of the maximum strain vs. the maximum elec-
trical field for several types of samples: 1.7 and 0.1 vol. % of CB-ink/PU, 2 and 1
vol. % of CBP/PU and 2 samples of pure PU. The conversion of strain for the pure
PU appeared to be less than 10 MV/m. Consequently, the strain level was limited
to 10%. The samples to which CBP had been added also displayed convergence.
These values of maximum strain were slightly higher than 10%. It is clear that the
CB-ink/PU samples presented higher strain levels than their pure PU and CBP/PU
counterparts. The samples with 1.7 and 0.1 vol. % exhibited maximum strains
around 24 and 12%, respectively. It was quite remarkable that even the addition of
merely 0.1 vol. % could give rise to higher strain levels than the 2-vol. % CBP/PU
sample.
Another significant result is that there was a dependence of the maximum strain
on the thickness. Figure 15 shows this dependence for the specimen with 1.7 vol.
% of CB-ink/PU.
298 K. Yuse et al.

Fig. 13 The change of the


form of the current and
polarization with a maximum
electrical field of a 1 MV/m,
b 5 MV/m and c 20 MV/m
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 299

Fig. 14 The maximum strain as a function of the maximum applied electrical field

Fig. 15 The dependence of


the maximum strain on the
thickness

4.5 Discussion

There are two possible explanations for the convergence of strain: a mechanical
and an electrical explanation. The former corresponds to an effect of the mea-
surement. When a triangular electrical field is induced, whether positive or neg-
ative, there occurs a deformation of the thickness in the direction of compression.
The strain in the present study was based on the compressed deformation.
Simultaneously, the elongation occurred in the direction of the diameter of the
disk-shaped specimen. This elongation, stimulated by the electric field, was
however limited in the elastic domain. It differed from the force application
measurement, and it is thus natural that the conversion occurred also with regard to
the thickness deformation. Still, a convergence around 10% of strain (pure PU)
seemed quite low when caused by such a phenomenon.
Another explanation could be an electrical reason, i.e., the change in material
characteristics caused by the induced electrical field. The basic constitutive
equations are the following:
300 K. Yuse et al.

S3 ¼ s33 s3 þ a33 P23


ð16Þ
E ¼ bP þ 2asP

Under no-stress condition, taking P ^ eE at a low electrical field regime, the


relation between strain and electrical field is always found to be quadratic
according to

S33 ’ ae2 E2 ; ð17Þ

where a is an electrostrictive coefficient. Besides, a polymeric composite can be


considered as a capacitive material. The current is expressed as

_
I ¼ C V; ð18Þ

where C is the capacitance. With a triangle-shaped electrical induction, the current


becomes rectangular as shown in Fig. 13a. The polarization takes on the triangular
shape..
Taking into account electrical losses, the current can be written
V
I ¼ CV_ þ ð19Þ
R
The electrical loss appears
R as the inner surface of Fig. 12 since the polarization
P is calculated by P ¼ A1 Idt: A represents the conductive surface area and its
change is neglected here. Equation 19 does not present the experimental evalua-
tion of Fig. 13. To overcome this problem, polarization was considered. The
evolution of polarization is difficult to observe experimentally because of the
prevention by the loss, see in Fig. 12. Here, the polarization undergoes saturation
according to the following equation:
 
E
P ¼ eEsat  tanh ð20Þ
Esat

where Esat is the convergence value of the electrical field. This assumption of
polarization was confirmed by the following two conditions: (1) when the electric
field is small compared to Esat, P = eE and (2) when it is much bigger, i.e.,
E  Esat, P = eEsat.
The strain can be simply expressed by the equation of polarization as S = aP2.
Figure 10b is the simulation curves. The form change of strain from a triangular
shape at a low electrical field to a round shape at a high electrical field is accu-
rately described with this simulation. Since the loss is not yet concerned, the strain
returns to zero at each E = 0 in Fig. 10b.
From Q = PA, the current can be also obtained in equation. The simulation
resulting current has a nonlinear relation with the electrical field. The simulation
curves at both a low and high electrical regime show good correspondence with
experimental data.
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 301

5 Example Study of Energy Harvesting using EAPs

5.1 Principle of Measurement of Harvested Power

Figure 16 provides a schematic representation of the setup developed for char-


acterizing the power harvested by the polymer film. The mechanical system
consisted of a shaker and a capacitive sensor. The shaker produced the vibration
force in sinusoidal form, causing the sample to undergo a transverse vibration. The
capacitive sensor (Fogale MC 940) measured the transverse displacement of the
sample from which the strain S1 was calculated. The electrostrictive polymer was
subjected to a DC biased electrical field, produced by a function generator and
amplified by the Trek Model 10/10 High-Voltage Power Amplifier. As the poly-
mer was not piezoelectric, it was necessary to induce a polarization with a DC bias
in order to obtain a pseudo-piezoelectric behavior. The EPC was excited both
electrically and mechanically, in order for its expansion and contraction to induce
a current measured by the current amplifier (Keithley 617), thus giving rise to ‘‘an
image’’ of the power harvesting by the polymer, due to electrical resistance (Rc). In
this setup, the current was chosen as it is known to be less sensitive to the noise
from the electrical network (50 Hz) and in order to avoid problems of impedance
adaptation. All the data was monitored by an oscilloscope (Agilent DS0 6054A
Mega zoom).

Fig. 16 A schematic of the experimental setup for the energy harvesting measurements
302 K. Yuse et al.

5.2 Materials

Two types of commercially available polymers were used: polyurethane and


nylon. Moreover, two composites were synthesized specifically for the study.
These polyurethane composites were prepared in the laboratory, using a thermo-
plastic polyurethane, the 58887 TPU elastomer (Estane), as the matrix. Neat
polyurethane (PU) films as well as their filled counterparts were prepared by a
solution casting method as mentioned in the previous section. Two types of
inorganic fillers, a carbon black nanopowder (CB) and silicon carbide (SiC)
nanowires, were chosen. Before the mixing, they were individually and ultra-
sonically dispersed in DMF.

5.3 Results

Figure 17 presents the power versus electrical field, for a constant electrical load
and strain (S1 = 0.25%), in the case of nylon. The RMS power harvested on the
load was derived using P = Rc.I2h. As expected from the model, a quadratic
dependence between the power and the static electrical field strain was observed.
There was a good agreement between the experimental results and the model, thus
validating the developed model for evaluation of the harvested power.
Table 1 gives the harvested power (Pharvested_AC = Rc.I2h) measured on the
various materials (with a transverse strain of S1 = 0.25% at 5 V/lm and 100 Hz)
for a matched load equal to Rc ¼ Cp1x ¼ Zoptimal : In fact, according to [11], there
should be an optimal load resistance for which the conversion power is at a
maximum. The power density for the polymer and composite was very low.
Although this could be considered a disappointing result, it should be kept in mind

Fig. 17 Power harvested as


a function of the electrical
field for a constant strain of
0.25% at 100 Hz, and for an
electric load of 2.5 MX
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 303

Table 1 Material characteristics and harvested power density at 100 Hz for a static electrical
field of 5 V/lm and a transverse strain of 0.25%
Type er Y (MPa) Power harvested (lW/cm3)
Neat PU 4.8 40 0.2
PU 0.5 wt%SiC 5.6 70 0.4
PU Vol1%CB 8.2 40 1.0
Nylon 12 2800 2.3

that in the electrostrictive case, the power was proportional to the square of the
bias field, evaluated in Table 1 for a relatively low bias value. For example,
doubling the value of the bias field to 10 V/lm (which is still quite low) would
result in a power that was 4 times larger. The nylon provided the higher harvested
power density. According to [11, 20–21], the M31 coefficient was proportional to
e0 ðer  1Þ2 =ðY  er Þ: In other words, M31.Y was roughly proportional to er and in
order to increase the energy harvesting of the polymer, it would be necessary to
employ a composite with a high dielectric permittivity. The incorporation of fillers
provides extreme interests. As shown in Table 1, the addition of conductive
nanofillers rendered it possible to increase the permittivity, which in turn led to an
increase in the harvested power.

6 Conclusions

This chapter has described the use of EAPs for energy harvesting actuator pur-
poses. After an introduction of some EAPs adapted for energy harvesters, two
experimental investigations were presented. Elastrostrictive polymers were mainly
introduced with our development and analysis study. One of the problems of EAPs
today is their high electric requirement.
In order to overcome this problem, the first study was devoted to fabricating an
intermediate material between a piezo element and actual EAPs, i.e., an EAP that
can exhibit a large strain with a reasonably low electrical field. Carbon black was
chosen as a conductive filler. In order to avoid aggregation, a micellar type of
carbon black, directly obtained from ink, was used. As a result, more than 20%
strain was easily obtained at less than 20 MV/m. The conductive mechanism was
investigated. The achievement of large-displacement EAPs requiring a supply
energy as low as that of peizo elements should lead the way to a large variety of
applications.
Furthermore, a second study involving energy harvesting through the use of
EAPs was also presented. The obtained results are not yet gratifying, but there is
hope that they will be improved soon.
EAPs is a very interesting and promising material class, and it is believed that
the amount of research devoted to it, and consequently the applications derived
from it, will expand from one day to the next.
304 K. Yuse et al.

References

1. Lebrun, L., Guyomar, D., Guiffard, B., Cottinet, P.-J., Putson, C.: The characterisation of the
harvesting capabilities of an electrostrictive polymer composite. Sens. Actuators A Phys. 153,
251–257 (2009)
2. Bar-Cohen, Y.: Electroactive Polymer (EAP) Actuators as Artificial Muscles (reality,
potential, and challenges). SPIE Press, Bellingham (2001)
3. Kofod, G., Kornbluh, R., Pelrine, R., et al.: Actuation response of polyacrylate dielectric
elastomers. In: Bar-Cohen, Y. (ed.) Smart Structures and Materials: Electroactive Polymer
Actuators and Devices. Proc. SPIE, vol. 4329 (2001)
4. Pelrine, R., Kornbluh, R., Eckerle, J., et al.: Dielectric elastomer: generator mode fundamentals
and applications. In: Bar-Cohen, Y. (ed.) Smart Structures and Materials: Electroactive Polymer
Actuators and Devices. Proc. SPIE, vol. 4329 (2001)
5. Lefeuvre, E., Badel, A., Richard, C., Guyomar, D.: High efficiency piezoelectric vibration
energy reclamation. In: Proc. SPIE SSM Conference, San Diego (2005)
6. Liu, Y., Ren, K., Hofmann, H.F., Zhang, Q.M.: Electrostrictive polymer for mechanical
energy harvesting. In: Proc. SPIE, Int. Soc. Opt. Eng., vol. 5385, pp. 17–28 (2004)
7. Ren, K., Liu, Y., Hofmann, H., Zhang, Q.M.: An active energy harvesting scheme with an
electroactive polymer. Appl. Phys. Lett. 91, 132910 (2007)
8. Liu, Y., Ren, K.L., Hofmann, F., et al.: Investigation of electrostrictive polymers for energy
harvesting. IEEE UFFC 52(12), 2411–2417 (2005)
9. Poulin, G., Sarraute, E., Costa, F., et al.: Generation of electrical energy for portable devices:
comparative study of an electromagnetic and piezoelectric system. Sens. Actuators A Phys.
116(3), 461–471 (2004)
10. Guyomar, D., Lebrun, L., Putson, C., Cottinet, P.-J., Guiffard, B., Muensit, S.: Elec-
trostrictive energy conversion in polyurethane nanocomposites. J. Appl. Phys. 106, 014910
(2009)
11. Cottinet, P.-J., Guyomar, D., Guiffard, B., Putson, C., Lebrun, L.: Modeling and
experimentation on an electrostrictive polymer composite for energy harvesting. IEEE
Trans. Ultrason. Ferroelectr. Freq. Control 57(4), 774–784 (2010)
12. Badel, A., Guyomar, D., Lefeuvre, E., et al.: Efficiency enhancement of a piezoelectric
energy harvesting device in pulsed operation by synchronous charge inversion. J. Intell.
Mater. Syst. Struct. 16, 889 (2005)
13. Badel, A., Benayad, A., Lefeuvre, E., Lebrun, L., Richard, C., Guyomar, D.: Single crystals
and nonlinear process for outstanding vibration-powered electrical generators. IEEE Trans.
Ultrason. Ferroelectr. Freq. Control 53, 673–684 (2006)
14. Prahald, H., Kornbluh, R., Pelrine, R., et al.: Dielectric elastomers and their applications in
distributed actuation and power generator. In: Proceedings of ISSS 2005. International
conference on Smart Materials Structures and Systems, India (2005)
15. Sebald, G., Lefeuvre, E., Guyomar, D.: Pyroelectric energy conversion: optimization
principles. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 55(3), 538–551 (2008)
16. Kanda, M., Yuse, K., Guiffard, B., Guyomar, D.: Large field induced strain in carbon nano-
filled composite polyurethane (PU). In: Proc. Smart’09, 493 pp (2009)
17. Guiffard, B., Seveyrat, L., Sebaid, G., Guyomar, D.: Enhanced electric field-induced strain in
non-percolative carbon nanopowder/polyurethane composites. J. Phys. D 39, 3053–3057
(2006)
18. Petit, L., Guiffard, B., Seveyrat, L., et al.: Actuating abilities of electroactive carbon
nanopowder/polyurethane composite films. Sens. Actuators A Phys. 148, 105–110
(2008)
19. Guyomar, D., Matei, D.F., Guiffard, B., Le, Q., Belouadah, R.: Magnetoelectricity in
polyurethane films loaded with different magnetic particles. Mater. Lett. 63, 611–613
(2009)
Actuators and Energy Harvesters Based on Electrostrictive Elastomeric Nanocomposites 305

20. Lebrun, L., Guyomar, D., Guiffard, B., Cottinet, P.-J., Putson, C.: The characterization of the
harvesting capabilities of an electrostrictive polymer composite. Sens. Actuators A Phys. 153,
251–257 (2009)
21. Guyomar, D., Lebrun, L., Putson, C., Guiffard, B., Cottinet, P.-J., Muensit, S.:
Electrostrictive energy conversion in polyurethane nanocomposites. J. Appl. Phys. (2009).
doi:10.1063/1.3159900
Elastomeric Nanocomposites
for Aerospace Applications

James Njuguna, Krzysztof Pielichowski and Agnieszka Leszczyńska

Abstract Nanotechnology adds a new capability to the design of materials


resulting in significant performance improvements at macroscopic level. In par-
ticular, nanostructured materials open a new paradigm wherein fibres and matrix
resins can be tailored to enhance composite properties of interest; just as fibres
and ply orientation is currently used to tailor current advanced composites
without increasing the mass. This chapter focuses on these new materials with
improved electrical, thermal, and mechanical properties. Special attention is given
on elastomeric nanocomposites from polyamides, polyurethanes, polyaniline and
polyethylene terephthalate due to their diversified applications. Polyimide, poly-
arylacetylene poly(aryl–ether–ether–ketone) and poly(p-phenylene benzbisoxaz-
ole) nanocomposites are also covered for high performance and temperature
aerospace applications. Exploitation of these new elastomers’ intrinsic properties
relies on the ability to systematically synthesize, characterize, and integrate
standardized materials into actual commercial products.

1 Introduction

The uses of polymer nanocomposites in structures have several predictable


impacts on aerospace design and applications, primarily by providing a safer,

J. Njuguna (&)
Department of Sustainable Systems, Cranfield University, Bedfordshire MK43 0AL, UK
e-mail: j.njuguna@cranfield.ac.uk
K. Pielichowski and A. Leszczyńska
Department of Chemistry and Technology of Polymers, Cracow University of
Technology, ul. Warszawska 24, 31-155 Kraków, Poland

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 307


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_12,
Ó Springer-Verlag Berlin Heidelberg 2011
308 J. Njuguna et al.

faster, and eventually cheaper transportation in the future. The most obvious of
which is significant airframe weight reduction stemming from polymer nano-
composites low density and complemented by its high strength and modulus.
Proper engagement of the nanoparticles in composite systems depends strongly on
the ability to homogeneously disperse them throughout the matrix without
destroying their integrity. The properties of the matrix, the distribution and
properties of the filler as well as the nature of their interface control the behaviour
of a typical composite material [1–3]. Thus, the nanoparticles often strongly
influence the properties of the composites at very low volume fractions. This is
mainly due to their small interparticle distances and the transformation of a large
fraction of the polymer matrix near their surfaces into an interphase of different
properties as well as to the consequent change in morphology [4]. As a result, the
desired properties are usually reached at low filler volume fraction, which allows
the nanocomposites to retain the macroscopic homogeneity and low density of the
polymer. Besides, the geometrical shape of the particles plays an important role in
determining the properties of the composites. Consequently, nanocomposites have
attracted much scientific and industrial interest in recent times.
So far, most of the scientific work has been focussing on the synthesis of
polymer nanocomposites and on the study of their physical and mechanical
properties. The use of these nanocomposites as matrix in fibre-reinforced
composites is in its infancy. Nano-composites from polymeric matrix materials
(thermoplasts or thermosets) reinforced with nano-sized fillers (see Fig. 1) such as
carbon nano-tubes, nano-sized metal and metal oxide particles or lamellar inor-
ganic or carbonaceous nanoparticles are an active area of research [5, 6].
The primary aim is to improve the deficiencies of existing composities or to
extend their functional limits. For example, the improved mechanical performance
of polymeric nanocomposites is macroscopically depicted in the increase in for
instance, fracture energy values. Nanofillers, due to their size, can be significantly
present in the plastic zone deformation while in the case of micro-particles only
few of them participate in this small zone. In this way, nanofillers can lead to
increased fracture properties of the brittle matrix. On the other hand, the matrix
material is pointed out to be a key parameter for [2, 3] mode I delamination
resistance (critical strain energy release rate for delamination in an opening mode)
of fibre reinforced polymers. Therefore, enhancement in the matrix fracture
toughness can lead to an overall advanced fracture behaviour. Additionally, the
influence of the nanofibres has also been extended to the reinforcing fibres by
making more nanofibres to be involved during the delamination process and thus
increasing the fracture toughness. Further, by invoking the properties of nano-
structures it may be possible to control the shock wave propagation in the material
and thus, significantly enhance the impact energy dissipation; however, only
limited research works to date concern about damping composite structures and
shock resistant nanomaterials. Adsorption of vibration energy by mechanical
damping is significant problem in many engineering designs; high damping
materials are used to reduce vibration in aircraft and other machinery. The benefits
of damping treatment are advanced durability, reliability and service life of
Elastomeric Nanocomposites for Aerospace Applications 309

Fig. 1 a Comparison of conventional fiber composite and nanocomposites. b Logarithmic iso-


lines of interfacial (surface) area/volume of particles (lm-1 = m2/ml) with respect to the aspect
ratio, a = H/R, and largest dimension of particle (R = radius, H = height, length) based on
approximating particles as cylinders (area/volume = 1/H ? 1/R). Aspect ratios greater than one
correspond to rods (length/diameter) and less than one to plates (height/diameter) (reproduced
from Refs. [5, 6])

components, reductions in weight, noise and costs. It should be noted, however,


that damping and integrity often have opposite requirements (see Fig. 2) [7];
viscoelastic polymers are good for damping, but stiffness is decreased at higher
temperatures.

2 Polymer Nanocomposites

A key advantage of the use of nanocomposite instead of other microparticulate


fillers to improve the fibre composite properties is that the properties can be
improved without any change in the processing conditions. The most promising
current approaches towards increasing the orientation of nanoscale reinforcements
within a matrix include optimisation of the extrusion die and stretching the
composite melt to form films and fibres. One complication is that the micro-
structure of semicrystalline polymer matrices is influenced not only by the pro-
cessing history but also by the presence of nanoparticles. The addition of various
types of nanofillers to polymers has already been observed to influence the
310 J. Njuguna et al.

Fig. 2 Stiffness-damping map for nanoparticle-reinforced and conventional composites (repro-


duced from Ref. [7])

crystallisation kinetics and resulting morphology. Such changes in matrix mor-


phology need to be considered when evaluating the nanocomposite performance
with regard to the intrinsic filler properties. The effects of carbon nanotubes or
nanofibres on such oriented polymer systems, although significant, have not yet
been fully established. Finally, it should be noted that the presence of additives
such as colouring pigments has been shown to influence matrix morphology during
fibre spinning, whilst there is the whole technology of nucleating agents which are
deliberately added to influence crystalline microstructure. Nanoparticle rein-
forcement of fibre composites has been shown to be a possibility, but much work
remains to be performed in order to understand how nanoreinforcement results in
major changes in material properties. The understanding of these phenomena will
facilitate their extension to the reinforcement of more complicated anisotropic
structures and advanced polymeric composite systems. The property and perfor-
mance enhancements made possible by nanoparticle reinforcement may be of great
utility for fibre-reinforced composites that are applied in aerospace industry.

2.1 Polyamide (PA) Nanocomposites

In the early 1990s of twentieth century, Toyota Research group synthesized PA-6-
based clay nanocomposites that demonstrated the first use of nanoclays as rein-
forcement of polymer systems. They concluded that nanoclays not only influenced
Elastomeric Nanocomposites for Aerospace Applications 311

the crystallization process, but that they were also responsible for morphological
changes. Recognizing these benefits, many researchers, using a variety of clays
and polymeric matrices, have produced nanocomposites with improved properties.
For instance, Liu et al. [8] reported that there was an increase in storage elastic
modulus of 100% when clay content was up to 8 wt% in comparison with net PA-
11. Usuki et al. [9] polymerized e-caprolactam in the interlayer of an organoclay to
form a nanocomposite. This material contained only 4.2 wt% clay and had a 50%
increase in strength, an increase in the heat distortion temperature of 80°C, a 100%
increase in tensile modulus, and a 20% increase in impact resistance.
The electrospun nanocomposite fibres have great potential for the applications
where both high surface-to-volume ratio and strong mechanical properties are
required such as the high-performance filters and fibre reinforcement materials.
Since the mechanical properties of fibres in general improve substantially with
decreasing fibre diameter, there is considerable interest in the development of
continuous electrospun polymer nanofibres. In this respect, Lincoln et al. [10]
reported that the degree of crystallinity of PA-6 annealed at 205°C increased
substantially with the addition of montmorillonite (MMT). This implied that the
silicate layers could act as nucleating agents and/or growth accelerators. In con-
trast, the study of Fong et al. [11] showed a very similar overall degree of crys-
tallinity for electrospun PA-6 and PA-6/Cloisite-30B nanocomposite fibres
containing 7.5 wt% of organically-modified MMT (OMMT) layers. Fornes and
Paul [12] have found that OMMT layers could serve as nucleating agents at 3%
concentration in PA-6/OMMT nanocomposite but retarded the crystallization of
PA-6 at a higher concentration of around 7%. In addition, the differences in the
molecular weight of PA-6 and the solvent used for electrospinning were also
expected to have different impacts on the mobility of PA-6 macro chains and the
interactions between them and OMMT layers, which may also affect the crystal-
lization behaviour of polymer during the electrospinning. Li et al. [13] manufac-
tured PA-6 fibres and nanocomposite fibres with average diameters around 100 nm
by electrospinning using 88% aqueous formic acid as the solvent. The addition of
OMMT layers in the PA-6 solution increased the solution viscosity significantly
and changed the resulting fibre morphology and sizes. Transmission electron
microscopy (TEM) images of the nanocomposite fibres, including ultra-thin fibre
sections, and wide angle X-ray diffraction (WAXD) results showed that OMMT
layers were well exfoliated inside the nanocomposite fibres and oriented along the
fibre axial direction. The degree of crystallinity and crystallite size were both
increased for the nanocomposite fibres and more significant for the fibres elec-
trospun from 15% nanocomposite solution, which exhibited the finest average
fibres size. As a result, the tensile properties of electrospun nanocomposites were
greatly improved. The Young’s modulus and ultimate strength of electrospun
nanocomposite fibrous mats were improved up to 70 and 30%, respectively, when
compared with PA-6 electrospun mats. However, the ultimate strength of the
nanocomposite fibrous mats electrospun from 20% nanocomposites solution was
decreased by about 20% due to their larger fibre sizes. The Young’s modulus of
PA-6 electrospun single fibres with a diameter around 80 nm was almost double
312 J. Njuguna et al.

the highest value that had been reported for the conventional PA-6 fibres and could
be improved by about 100% for the electrospun nanocomposite single fibres of
similar diameters.
In another interesting study [14], a range of polymer matrices were examined
including poly(vinyl alcohol), poly(9-vinyl carbazole) and polyamide. To compare
production methods, polymer composite films and fibres were produced. It was
found that by adding various mass fractions of nanofillers, both the Young’s
modulus and hardness increased significantly for both films and fibres. In addition,
the thermal behaviour was seen to be strongly dependent on the nanofillers added
to the polymer matrices. Wu et al. [15] prepared carbon fibre and glass fibre
reinforced PA-6 and PA-6/clay nanocomposites. The fabrication method involved
first mechanically mixing PA-6 and PA-6/clay with E-glass short fibre (6-mm
long) and carbon fibre (6-mm long), separately. A twin-screw extruder at a rota-
tional speed 20 rpm extruded the fibres. The temperature profiles of the barrel
were 190–210–230–220°C from the hopper to the die. The extrudate was pellet-
ized, dried, and injection moulded into standard test samples for mechanical
properties test. The injection-moulding temperature and pressure were 230°C and
13.5 MPa, respectively. The research found out that the tensile strength of PA-6/
clay containing 30 wt% glass fibres was 11% higher than that of PA-6 containing
30 wt% glass fibre, while the tensile modulus of nanocomposite increased by 42%.
Flexural strength and flexural modulus of PA-6/clay were found similar to PA-6
reinforced with 20 wt% glass fibres. It was eluded that the effect of nanoscale clay
on toughness was more significant than that of the fibre. Heat distortion temper-
atures of PA-6/clay and PA-6 were 112 and 62°C, respectively. Consequently, the
heat distortion temperature of fibre reinforced PA-6/clay system was almost 20°C
higher than that of fibre reinforced PA-6 system. Notched Izod impact strength of
the composites decreased with the addition of the fibre. The scanning electron
microscopy (SEM) microphotographs showed that the wet-out of glass fibre was
better than carbon fibre. The study concluded that the mechanical and thermal
properties of the PA 6/clay nanocomposites were superior to those of PA-6
composite in terms of the heat distortion temperature, tensile and flexural strength
and modulus without sacrificing their impact strength. This was attributed to the
nanoscale effects, and the strong interaction force existing between the PA-6
matrix and the clay interface.
In case of short fibres, Akkapeddi [16] prepared PA-6 based nanocomposites
using chopped glass fibres. In a typical experiment, a commercial grade PA-6 of
MW = 30,000 and specially designed functional organo-quaternary ammonium-
clay complexes based on MMT or hectorite type clays. Freshly dried PA-6
(moisture \ 0.05%) was blended with 3–5 wt% of a selected organoclay powder
and extruded at 260°C in a single step, under high shear mixing conditions.
Alternatively, the organoclay was master-batched first into PA-6 (at 25 wt%
loading and then re-extruded in a second step with more PA-6 to dilute the clay
content to B5 wt%. Conventional chopped glass fibre with 10 lm diameter and
about 3 mm length was then added, as an optional reinforcement through a
downstream feed port of the twin screw extruder. The glass fibre was compounded
Elastomeric Nanocomposites for Aerospace Applications 313

Fig. 3 Flexural modulus


against glass fibre density
(GF), PA-6/nanoclay (PA-6/
NC) versus PA-6 moulding
resins (reproduced from Ref.
[9])

with the molten, premixed PA-6 nanocomposite either as a one-step extrusion


process or in a second extrusion step. The extrudate was quenched in a water bath
and pelletized. The pellets were dried under vacuum at 85°C, and injection
moulded into standard ASTM test specimens. As shown in Fig. 3, significant
improvements in modulus were achievable in both the dry and the moisture
conditioned state for PA-6 nanocomposites compared to standard PA-6, at any
given level of glass fibre reinforcement.
In particular, a small amount (3–4 wt%) of nanometer scale dispersed layered
silicate was capable of replacing up to 40 wt% of a standard mineral filler or
10–15 wt% of glass fibre to give equivalent stiffness at a lower density. In addi-
tion, improved moisture resistance, permeation barrier and fast crystallization/
mould cycle time contribute to the usefulness of such composites.
Vlasveld et al. [17, 18] investigated fibre–matrix adhesion in glass-fibre rein-
forced PA-6 silicate nanocomposites. The main reinforcing phase consisted of
continuous E-glass fibres, whereas the PA-6 based matrix was a nanocomposite
reinforced with platelets of exfoliated layered silicate. Two different types of
nanocomposite were used with different degrees of exfoliation of the silicate
layers: one with non-modified silicate and one with an organically modified sili-
cate. They developed nanocomposite laminates by sol–gel and modified dia-
phragm methods. The route for the preparation of PA-6 nanocomposites consisted
of melt-compounding polymer (AkulonÒ K122D) with organically modified
fluorine-containing hectorite (SomasifÒ MEE and SomasifÒ ME-100) by means of
a co-rotating twin-screw extruder at 240°C. For the SomasifÒ MEE nanocomposite
materials, first an 11 wt% MEE master batch was compounded. To obtain the
various concentrations of the MEE nanocomposite, the master batch was extruded
for a second time without dilution for the 11 wt% nanocomposite, or diluted with
PA-6 to concentrations of 6.1 and 2.7 wt%. The 2.5 wt% SomasifÒ ME-100
314 J. Njuguna et al.

nanocomposite material was produced by diluting a 10% organoclay master batch


with polymer in the extruder. All mentioned percentages are weight percentages
silicate as measured with a thermogravimetric analyser after heating for 40 min at
800°C in air. Two demands for the preparation of the single fibre fragmentation
specimen had to be met: the fibre had to lie straight in the centre of the specimen,
and the matrix material of the specimen had to be thin enough to be transparent,
since the fibre fragments were examined and measured using an optical micro-
scope. A Fontijne hot plate press heated to 240°C was used to produce the films
necessary for the single fibre fragmentation test specimen preparation. Single
fibres were carefully extracted from a fibre bundle and placed with a distance of
approximately 2 cm parallel to each other between the PA or nanocomposite films.
The hot plate press at the same temperature was used to melt the polymer films and
a pressure of 0.8 N/mm2 was applied for 30 s to provide the necessary bonding
with the fibre. After cooling between cold metal plates, tensile test specimens were
prepared. It was observed that the ultimate strength and stiffness increased by
adding 1% SiO2 nanoparticles, while little improvement in fatigue behaviour was
found. It was concluded that the failure mechanism was by interfacial de-bonding
and that both the addition of nanoparticles and moisture conditioning had a neg-
ative effect on the bonding between the matrix and the glass fibres. In addition, the
researchers noted that in the formed composites the adhesion between the nano-
composites and the carbon fibres (Fig. 4) was probably worse than between the
unfilled PA-6, reducing the potentially positive influence of the increased matrix
modulus.
In a parallel research, Vlasveld et al. [18] developed three-phase thermoplastic
composite, consisting of a PA-6 nanocomposite matrix and a main reinforcing
phase of woven glass or carbon fibres. The nanocomposite used in this research
had moduli that were much higher than those of unfilled PA-6, also above Tg and
in moisture conditioned samples. Flexural tests on commercial PA-6 fibre

Fig. 4 Flexural strength of


carbon fibre composites with
PA6, a commercially avail-
able PA6 nanocomposite
(Unitika M1030D from Uni-
tika) and nanocomposite
matrix as a function of the
matrix modulus (dry and
moisture conditioned)
(reproduced from Ref. [18])
Elastomeric Nanocomposites for Aerospace Applications 315

composites showed decrease of the flexural strength upon increasing temperature.


The researchers claimed that the strength of glass fibre composite can be increased
by more than 40% at elevated temperatures and the temperature range at which a
certain minimum strength is present can be increased by 40–50°C. Carbon fibre
composites also showed significant improvements at elevated temperatures,
although not at room temperature. Based on flexural tests on PA-6 based glass and
carbon fibre composites over a large temperature range up to near the melting
point, it became clear that for these fibre composites it is important to have a
reasonably high matrix modulus: both glass and carbon composites were very
sensitive to a decrease of the matrix modulus below values around 1 GPa. At
higher moduli, carbon fibre composites are more sensitive to the matrix modulus
than glass fibre composites. The modulus of unfilled PA-6 decreased below the
(arbitrary) 1 GPa level just above Tg; it is noteworthy that the nanocomposites
used in this research had moduli that were much higher and stayed above the
1 GPa level up to 160°C, which was more than 80°C higher than for unfilled PA-6.
The nanocomposites also showed much higher moduli in moisture conditioned
samples, and even in moisture conditioned samples tested at 80°C the modulus
was much higher than that of the dry unfilled PA-6, again well above 1 GPa.
Dynamic mechanical analysis (DMA) measurements showed that the nanocom-
posites did not show a change of Tg in the dry state, and that the reduction of the
modulus upon absorption of moisture was due to the Tg decrease.
An assessment of reactively processed anionic polyamide-6 (APA-6) for use as
matrix material in fibre composites was conducted by van Rijswijk et al. [19], who
also compared it with melt processed PA-6 and PA-6 nanocomposites. A special
designed lab-scale mixing unit was used to prepare two liquid material formula-
tions at 110°C under a nitrogen atmosphere: a monomer/activator-mixture in tank
A and a monomer/initiator-mixture in tank B, as shown on Fig. 5.
After individually degassing both tanks (15 min at 100 mbar), the two material
feeds were mixed by using a heated (110°C) static mixer and dispensed (1:1 ratio)
into a heated (110°C) buffer vessel with nitrogen protective environment. Stainless
steel infusion mould (Fig. 3) was used together with a 3 mm thick stainless steel
cover plate (not shown) to manufacture neat APA-6 panels (250 9 250 9 2 mm).
Homogeneous heating of the mould was obtained by placing it in a vertically
positioned hot flat plate press. A silicon tube connected the resin inlet of the mould
with the buffer vessel and the resin outlet with a vacuum pump. Infusion from
bottom to top was necessary to prevent entrapment of air. A pressure control
system was used to precisely set the infusion and curing pressure (absolute pres-
sure in the mould cavity). Loss of control over the pressure in the mould cavity due
to solidification of resin in the unheated outlet tube had to be prevented. To avoid
this, a buffer cavity was machined in the mould near the outlet to slow down the
infusion, hence giving ample time to stop the resin flow before it was able to exit
the mould. For every infusion pressure, the infusion time to reach the buffer cavity
was determined visually by replacing the steel cover plate by a glass one. Addi-
tionally, a resin trap and a cold trap were placed directly after the mould to protect
the vacuum pump. The mechanical properties of APA-6 (commercial grade of
316 J. Njuguna et al.

Fig. 5 Infusion equipment from left to right: Mini Mixing Unit ‘‘MMU-TU Delft’’, resin res-
ervoir, stainless steel infusion mould, resin trap, cold trap and vacuum pump (reproduced from
Ref. [12])

PA6, AkulonÒ K222D) and HPA-6 (low MW injection-moulding grade hydro-


lytically polymerised PA-6) nanocomposites were compared with injection
moulded neat HPA-6. As expected, the HPA-6 nanocomposite had the highest
modulus over the entire range of temperatures (20–160°C) and moisture contents
(0–10 wt%) tested. However, APA-6 came close and had the highest maximum
strength due to its characteristic crystal morphology, which was directly linked to
the reactive type of processing used. The same morphology, it was claimed, also
made APA-6 slightly less ductile compared to melt processed HPA-6. Compared
to the melt processed HPA-6, APA-6 polymerised at 150°C and the HPA-6
nanocomposite offered a higher modulus at similar temperature, or similar mod-
ulus at a higher temperature (40–80°C increase). It is noteworthy that such an
increase in maximum use temperature, related to the heat distortion temperature,
can seriously expand the application field of PA-6 and PA-6 composites. For all
PAs, temperature and moisture absorption reduced the modulus and the strength
and increased the maximum strain, which was directly related to the glass tran-
sition temperature. Whereas with increasing testing temperature at a certain
moment the Tg of the dry polymer was exceeded, moisture absorption reduced the
Tg at a certain point below the testing temperature. However, the effect of both was
in essence the same. Retention of mechanical properties of APA-6 after condi-
tioning at 70°C for 500 h and subsequent drying was demonstrated. Conditioning
submersed in water at the same temperature, however, resulted in a brittle material
with surface cracks, as is common to most polyamides due to continued crystal-
lization and removal of unreacted monomer. Given the fact that submersion at
elevated temperatures is usually not an environment in which PA-6 and its com-
posites are applied, the encountered property reduction was therefore not
Elastomeric Nanocomposites for Aerospace Applications 317

detrimental for application of these materials. The overall conclusion of the


comparative study for application of the polyamides as matrix material in fibre
composites was that both APA-6 and the HPA-6 nanocomposites outperformed
the melt processed HPA-6 in terms of modulus and maximum strength. Therefore, the
researchers concluded that both special grade PAs may be expected to enhance the
matrix dominated composite properties like compressive and flexural strength,
provided that a strong fibre-to-matrix interphase is obtained.
The presence of multi wall carbon nanotubes (MWNTs) improves the thermal
stability of PA-6 under air obviously, but has little effect on the thermal degra-
dation behaviour of PA-6 under nitrogen atmosphere. The thermal degradation
mechanism of PA-6 has been proposed by Levchik et al. [20]. A comparative study
was conducted by Sandler et al. [21] on melt spun PA-12 fibres reinforced with
carbon nanotubes and nanofibres. A range of multi-walled carbon nanotubes
(MWNT) and carbon nanofibres (CNF) were mixed with a PA-12 matrix using a
twin-screw microextruder, and the resulting blends spun to produce a series of
reinforced polymer fibres. The work aimed to compare the dispersion and resulting
mechanical properties achieved for nanotubes produced by the electric arc and a
variety of chemical vapour deposition techniques. A high quality of dispersion was
achieved for all the catalytically-grown materials and the greatest improvements in
stiffness were observed using aligned, substrate-grown, carbon nanotubes. The use
of entangled MWNT led to the most pronounced increase in yield stress, most
likely as result of increased constraint of the polymer matrix due to their relatively
high surface area. The degrees of polymer and nanofiller alignment and the
morphology of the polymer matrix were assessed using X-ray diffraction (XRD)
and differential scanning calorimetry (DSC). The carbon nanotubes were found to
act as nucleation sites under slow cooling conditions, the effect scaling with
effective surface area. Nevertheless, no significant variations in polymer mor-
phology as a function of nanoscale filler type and loading fraction were observed
under the melt spinning conditions applied. A simple rule-of-mixture evaluation of
the nanocomposite stiffness revealed a higher effective modulus for the MWNT
compared to the CNF, as a result of fewer imperfections in graphitic crystall lattice
of MWNT. In addition, this approach allowed a general comparison of the
effective nanotube modulus with those of nanoclays as well as common short glass
and carbon fibre fillers in melt-blended PA composites. The experimental results
further highlighted the fact that the intrinsic crystalline qualities, as well as the
straightness of the embedded nanotubes, were significant factors influencing the
reinforcement capability.
In an intumescent ethylene-co-vinyl acetate (EVA)-based formulation, using
PA-6 clay nanocomposite instead of pure PA-6 (carbonization agent) has been
shown to improve the fire properties of the intumescent blend. Using clay as
‘‘classical’’ filler enabled the same level of flame retardation performance to be
obtained in the first step of the combustion as when directly using exfoliated clay
in PA6. But in the second half of the combustion, the clay destabilizes the system
and increases the flammability. Moreover, a kinetic modeling of the degradation of
the EVA-based formulations shows that adding clay to the blend enables same
318 J. Njuguna et al.

mode of degradation and the same invariant parameters as for the polyamide-6
clay nanocomposite containing intumescent blend. The increase in the flamma-
bility in the second half of the combustion shows the advantages of using nanoclay
rather than micronclay in an intumescent system [20].
Organically modified clay-reinforced polyamide-6 was subjected to accelerated
heat aging to estimate its long-term thermo-oxidative stability and useful lifetime
compared to the virgin material [22]. Changes in molecular weight and thermal
and mechanical properties were monitored and connected to the polymer modi-
fication encountered during aging. Generally, the strong interaction between the
matrix and the clay filler renders the polymer chains, mainly that adjacent to
silicates that are highly restrained mechanically, enabling a significant portion of
an applied force to be transferred to the higher modulus silicates. This mechanism
explains the enhancement of tensile modulus that the non-aged clay-reinforced
PA6 exhibited (1,320 MPa) with regard to the neat polymer (1,190 MPa), as
shown in Fig. 6.

2.2 Polyurethane (PU) Nanocomposites

The physical properties of PU are derived from their molecular structure caused by
interactions between the polymer chains. The segmental flexibility, the chain
entanglement, and the cross-linking effects are all factors that influence on the
properties and determine the use of the end-products [23]. Thermoplastic PU
elastomers (TPUs) are linear block copolymers consisting of alternating hard and
soft segments. The hard segments are composed of alternating diisocyanate and
chain extender molecules (i.e. diol or diamine), while the soft segments are formed
from a linear, long-chain diol. Phase separation occurs in TPUs because of the
thermodynamic incompatibility of the hard and soft segments. The segments
aggregate into microdomains resulting in a structure consisting of glassy (hard
domains) and rubbery (soft domains) states that are below and above their glass
transition temperatures at room temperature, respectively. The hard domains gain
their rigidity through physical crosslinking (hydrogen bonding between hard
segments), and provide filler-like reinforcement to the soft segments [24, 25].
To improve their mechanical properties PU/layered silicate nanocomposites
have received increasing interest. For instance, Ma et al. [26] studies on elasto-
meric PU/clay nanocomposite based on poly(propylene glycol), glycerol prop-
oxylate and toluene diisocyanate (TDI) showed that the gallery distance of the clay
in the hybrid was enlarged from 1.9 to 4.5 nm or more. Introducing clay in the PU
matrix resulted in an increase in both the tensile strength and elongation at break.
When the clay content reached about 8%, the tensile strength and elongation at
break were two times and five times higher to that of the pure PU, respectively. In
Han et al. [27] work, the tensile properties of the PU/MMT nanocomposites
prepared by solution intercalation displayed higher enhancement relative to PU
matrix. Xu et al. [28] proved that the silicate layer spacing in the nanocomposites
Elastomeric Nanocomposites for Aerospace Applications 319

Fig. 6 Tensile properties of PA6 and PA6 nanocomposite, oven-aged at a 120°C and b 150°C
(reproduced from Ref. [28])

increased significantly compared with the neat organically layered silicate (OLS),
signifying the formation of intercalated PU-urea/OLS structures. The nanocom-
posite materials exhibited increased modulus with increasing OLS content,
while maintaining polymer strength and ductility. Zilg et al. [29] prepared PU
320 J. Njuguna et al.

nanocomposites from modified reactive fluoromica clay. The dried organophilic


mica was dispersed by means of a high shear mixer in trihydroxy-terminated oligo-
propyleneoxide (average MW = 3,800 g/mol). Stable and transparent polyol dis-
persions were obtained and then cured with 4,40 -diisocyanatophenylmethane
(MDI) and accelerated with 0.6 wt% N,N-dimethyl-benzyl-amine. The final
nanocomposite PUs showed an increase in tensile strength and elongation in
conjunction with a slight decrease in Young’s modulus, whereas non-modified
mica gave increased stiffness and only a small increase in tensile strength. Seg-
mented PU/clay nanocomposites have also been developed by Chen et al. [30, 31].
The first report by this group described the initial synthesis of a polycaprolactone/
clay (PCL/clay) nanocomposite by in situ polymerisation and its subsequent
incorporation into the solution synthesis of a segmented PU as a partial replace-
ment of the chain extender (1,4-butanediol) in the hard block [30]. Later on, the
same group developed PU/clay nanocomposites by directly mixing organoclay
with PU using the technique of intercalation from dimethylacetamide solution
[32]. In both cases the PU/clay nanocomposites developed by Chen et al. showed
enhancements in properties such as modulus and strength. Li [33] proved that a
complete and effective entry of monomers into the OMMT layers is very difficult
when OMMT content is higher. However, Song et al. [34] reported that the tensile
strength of PU/OMMT nanocomposites with 4.7 wt% OMMT content was double
compared to that of bulk PU, while Tortora et al. [35] used X-ray analysis to show
that exfoliation occurred for low MMT content, whereas for higher contents the
intercalated clay rearranged to a minor extent. However, either exfoliated or
delaminated samples showed an improvement in the elastic modulus and yield
stress but a decrease in the stress and strain at breaking on increasing the clay
content.
Song et al. [36] synthesised high performance nanocomposites comprising a PU
elastomer, based on poly(propylene glycol), 4,40 -methylene bis(cyclohexyl iso-
cyanate) and 1,4-butandiol, and an organically modified layered silicate. The
tensile strength and strain at break for these PU elastomer nanocomposites
increased more than 150%, but the hardness remained unchanged. The fatigue
properties were significantly improved too, particularly for the nanocomposites
with 3 wt% organoclay. The effects of the isocyanate index on the mechanical
properties of the PU elastomer nanocomposites were subjected to investigations
and found that an isoyanate index of 1.1 resulted in the best improvement in stress
and elongation at break.
In another development, nanocomposites with different concentrations of soft
segments (50 and 70%) and prepared by adding nanosilica (up to 30%) to the
single-phase PU matrix displayed higher strength and elongation at break but
lower density, modulus and hardness than the corresponding micron size silica-
filled PUs [37]. Corresponding results have been presented elsewhere, whereby the
addition of a small amount of nanosilica increased the hardness, abrasion resis-
tance and tensile properties of the polymer films, but similarly the properties
worsened at higher nanosilica contents [38]. In a parallel work, a two fold increase
in the tensile strength has been reported for 1% PU/benzidine-MMT as compared
Elastomeric Nanocomposites for Aerospace Applications 321

to that of pure PU [31]. Mishra et al. [39] DMA results indicated a significant
increase in storage modulus and tensile strength with increased organoclay loading
in millable PU/organoclay nanocomposites. In the study, clay treated with methyl
tallow bis(2-hydroxyethyl) quaternary ammonium chloride was used as an
organoclay for nanocomposite preparation. The mechanical properties improve-
ments were in good agreement with those reported by Kim et al. [40] on nano-
composites synthesized using amphiphilic urethane precursor (APU)/Na+-MMT
emulsions with microphase-separated structure that had greater tensile strength
than those prepared with melt-mixed APU/Na+-MMT mixtures. The modulus,
tensile strength and hardness were enhanced by the reinforcing effect of interca-
lated or exfoliated organoclay. It was claimed that the reinforcing effect was more
evident when the elasticity of the PU matrix was increased at the temperature range
above the Tg of hard segments and in the test where the deformation was large [41].
In Wang et al. [42], Fourier transform infrared spectroscopy (FTIR), WAXD
and TEM results verified the incorporation of the OMMT (the MMT modified with
was N-octadecyl N,N,N-trimethyl ammonium chloride (DK1) and N-12-di-
hydroksyethyldodecyl-N,N,N-trimethylammonium chloride (DK2)) into EPU
matrix and revealed that the degree of basal-spacing expansion was largely
increased. The enhanced mechanical and physical properties demonstrated effi-
cient reinforcing and better thermal stability properties of the latter OMMT. The
tensile strength of 3% EPU/DK2 was about 350% higher than that of pure EPU
and the elongation at break also showed a remarkable enhancement compared with
that of pure EPU. When the amount of MMT increased higher than 3%, the tensile
strength decreased, which was blamed to the aggregates of modified MMT in the
composites as reported elsewhere by Younghoon and James [43]. Nevertheless,
this is in contrast to the work results of Ni et al. [44] where OMMT content [3%
reportedly resulted to increase in tensile strength and elongation at break. The
work synthesized a novel polyether PU/clay nanocomposite was using poly(tet-
ramethylene glycol), 4,40 -diphenylmethane diisocyanate (MDI), 1,6-hexamethy-
lenediamine, and modified Na+-MMT [44]. Here, OMMT was formed by applying
1,6-hexamethylenediamine as a swelling agent to treat the Na+-MMT. The X-ray
analysis showed that exfoliation occurred for the higher OMMT content (40 wt%)
in the polymer matrix. The mechanical analysis indicated that, when the OMMT
was used as a chain extender to replace a part of the 1,2-diaminopropane to form
PU/clay nanocomposites, the strength and strain at break of the polymer was
enhanced when increasing the content of OMMT in the matrix. When the OMMT
content reached about 5%, the tensile strength and elongation at break were over
two times that of the pure PU. The thermal stability and the glass transition of the
PU/OMMT nanocomposites also increased with increasing OMMT content.
Jin et al. [45] investigation results on the viscoelasticity of PU-organoclay
nanocomposites suggested that the addition of organoclay resulted in the increase
in the elasticity of PU and the decrease in damping property, and significant
improvement of the thermal stability of PU. However, it was also found that the
addition of organoclay did not enhance the modulus of PU significantly, and that at
high contents of soft segments the modulus of PU decreased with the addition of
322 J. Njuguna et al.

organoclay. Hysteresis results indicated that energy dissipation increased with


increasing organoclay; at small deformation of 50 or 100%, the difference of
dissipated energy between the two cycles was not affected by the organoclay. To
the researchers, this meant that the part of the domain deformation energy was the
same for the PU nanocomposites at different organoclay contents. It demonstrated
that the hard domain deformation was not disturbed by the orientation of the
organoclay platelets at small deformation. At the larger deformation of 200% a
mutual influence between the domain deformation and the orientation of the
organoclay platelets coexisted. Therefore, during the larger deformation the extra
energy was required to overcome the effect. The more organoclay incorporated in
the PU, the more the dissipated energy was. However, the researchers noted that
the intercalation characteristic of organoclay by PU chains and the dispersion of
organoclay in PU matrix could be key factors affecting the viscoelasticity of the
composites.
When inorganic fillers are dispersed in polymers, elastic modulus is typically
increased with a concurrent sacrifice of the ultimate strength and elongation. Not
surprisingly, the nanodispersed silicates resulted in a significant increase in
modulus by a factor of two to three at 20 wt% organic layered silicate (OLS)
content [28]. However, the ultimate strength remained comparable to that of the
neat PU. These changes were also accompanied by retention of the native PU
ductility and even an increase in strain to failure at relatively low OLS levels.
Although in the bulk state silicates behaved as relatively rigid materials, TEM
micrographs of nanocomposite morphology showed that individual silicate layers,
because of their high aspect ratio and nanometer thickness, exhibited a measurable
flexibility. Studies by Zhou et al. [46] found the Young’s modulus of PU coatings
with nanosilica and with fumed silica was about the same, while the Young’s
modulus of coatings with micron size silica was distinguishably smaller than both.
The researchers attributed the findings to the reaction of isocyanate with hydroxyl
groups on the surfaces of nanosilica resulting in a higher crosslinking degree for
nanosilica embedded film than for micro-structured silica contained film. Another
reason given was that there ought to have been greater interaction strength
between nanosilica and organic matrix than micron size silica and organic matrix,
since the former had greater specific surface area than the latter; fumed silica has
to some extent similar particle characteristic to nanosilica. As concentration of
microsilica increased to 10 wt%, the Young’s modulus of the coatings showed an
apparent increase, especially under the high normal loads. It was concluded that
the tensile strength and Young’s modulus were enhanced with the increasing
content of nanosilica as displayed on Fig. 7; however, the elongation at break
decreased as nanosilica content was increasing.
Another report on PU/nanosilica composites showed a modulus of elasticity of
4.8 MPa while the elongation at break was measured to be 580% in comparison to
that of pure PU; however, the elongation at break decreased as nanosilica content
increased [46]. The UV absorbance in the wavelength of 290–400 nm increased as
nano-SiO2 content increased. In contrast, for the PU coatings with fumed silica or
micro-sized silica embedded, only hardness and abrasion resistance showed some
Elastomeric Nanocomposites for Aerospace Applications 323

Fig. 7 The tensile properties


of PU films with different
silica content (a) and
Young’s modulus (b) of PU
films containing different
types of silica (reproduced
from Ref. [56])

increase. In still developed experiment, nanotube sensors have been implemented


in PU matrix for quantitative stress mapping in an elastic deformation situation
[47]. This has been demonstrated by measuring the stress field in the vicinity of
holes in polymer films; the experimental data were found to be in good agreement
with the classical theory of Inglis as noted elsewhere [48]. The random dispersed
nanotubes are more useful in practice because different stress components can be
measured in the same sample. Further, Zhou et al. [38] prepared polyester-based
PUs with embedded nanosilica particles. It was reported that addition of a small
amount of nanosilica increased the hardness, abrasion resistance and tensile
properties of the polymer films although these mechanical properties worsened at
higher nanosilica contents. Also, according to the UV–Vis analysis it was noted
that the absorbance and reflection of ultraviolet–visible light by the PU films
increased as the nano-SiO2 content increased, especially at wavelengths of 290–
400 nm. The work further noted that the viscosity of the PU resin increased as the
nanosilica content increased. Chiang et al. [49] report showed a significant
advantages in the utilization of fullerenols as hyper-cross-linkers for the synthesis
of PU nanoelastomers resulting in a substantial increase of polymer tensile
strength for an order of magnitude from that of its linear polymer analog to a
324 J. Njuguna et al.

Fig. 8 Nominal stress-nomi-


nal strain curves for exfoli-
ated second pseudo-
generation HBP/Na+-MMT-
based PU nanocomposites
and intercalated second
pseudo-generation HBP/
organically modified MMT
composites for the overall
compositions indicated (27°C
and 4 mm/min for a 50-mm
gauge length) (reproduced
from Ref. [50])

maximum of 7.2 MPa. In Plummer et al. [50] report, the strength and stiffness in
the exfoliated composites increased with respect to the neat matrix, however, there
was a decrease in strain at break, as seen in on Fig. 8.
The PU nanocomposites containing exfoliated OMS showed an increase in
both strength and strain at break in comparison to nanosystems containing
intercalated OMS. These results are in agreement with those of Tien and Wei
[51], who reported a 34% increase in Young’s modulus. In mechanical properties
of a novel segmented PU/clay nanocomposite based on PCL, MDI, butanediol,
and PCL/clay prepolymer, about 1.4% PCL/clay in PU/clay resulted in a large
increase in the elongation of PU/clay. However, when the amount of PCL/clay
was 4.2%, the elongation of PU/clay was reduced drastically. This behaviour
indicated that PU/clay could be transformed from an elastomer to a thermoplastic
material as the amount of PCL/clay in PU/clay increased. Additionally, the lap
shear stress of PU/clay was at least three times that of neat PU as a result of the
influence of PCL/clay component [30]. A two fold increase in the tensile strength
and a three fold increase in the elongation were found for PU composite con-
taining 1% of benzidine-MMT (BZD-MMT) as compared to that of pure PU.
Furthermore, both PU-based nanocomposites containing 1% of MMT modified
with 12-aminolauric acid (12COOH-MMT/PU) and 1% BZD-MMT/PU exhibited
lower water absorption properties than that of pure PU. In addition, both 1%
12COOH-MMT/PU and 1% BZD-MMT/PU exhibited lower water absorption
properties than that of pure PU [31]. Scheme 1 shows the schematic drawing of
proposed intermolecular interaction between PU molecules and (a) 12 aminol-
auric acid-MMT, (b) benzidine-MMT.
Polymer sandwich composites, including PU-based composites, are widely
used in aircraft engine nacelle, on wings for fuel tanks protection, tailplane panels
for protection of stones and pebbles on take-off and landing, naval ships, human
Elastomeric Nanocomposites for Aerospace Applications 325

Scheme 1 The schematic drawing of possible intermolecular interaction between PU molecules


and a 12 aminolauric acid-montmorillonite, b benzidine-montmorillonite (reproduced from Ref.
[31])

vests and helmets for ballistic protection, automotive for collision and heat pro-
tection. In all these applications, the intrinsic properties of light weight and rigidity
are used. These sandwiches, composed of a core of cellular material and two
composite skins, are light (since their constituents are of low density), rigid in
traction and compression (the composite materials have good mechanical prop-
erties) but also in bending since the foam core thickens the structure (and thus
increases its quadratic moment while limiting its weight) and supports high
bending moments. These properties are particularly interesting for producing
functional structures that must sustain high stresses under normal conditions.
During severe or extreme loadings (crashes or accidents), these structures must
deform plastically and absorb the impact energy to protect either the rest of the
structure or the vehicle occupants. A possible way of improving the properties of
foam materials is through the inclusion of small amounts of nanoparticles (carbon
nanotubes and nanofibres, TiO2, nanoclay, etc.) to improve the foam density and
modulus properties. Up to now, MMT nanoclays have been the best candidate for
foam reinforcement due to ease of processing, enhanced thermal–mechanical
properties, wide availability and cost [52, 53]. Likewise, polyurethanes (PU) are
core materials of choice due to their tailorable and versatile physical properties,
ease of manufacture and their low costs. The use of epoxy resins filled with
nanoparticles to construct either laminates or foams is relatively new. Moreover,
326 J. Njuguna et al.

the use of nanoparticles in such laminates [54], or foams in sandwich composite


construction, is in its infancy but has been found to be both realistic and beneficial
[4]. For instance, by using less than 5% by weight of nanoclay loadings, significant
improvement in foam failure strength and energy absorption has been realised with
over a 50% increase in the impact load carrying capacity when compared to a neat
foam sandwich [55, 56]. Njuguna et al. [57] have recently conducted low velocity
impact studies on nanophased polyurethane cores in sandwich structures. The
investigation observed that nanophased sandwich structures are capable of taking
higher peak loads than those made of neat polyurethane cores when subject to low-
velocity impact. It was found that the incorporation of MMT resulted in higher
number of PU cells with smaller dimensions and higher anisotropy index (cross-
sections RI and RII). The obtained materials exhibited improved parameters in
terms of thermal insulation properties. The results also show that nanophased
sandwich structures are capable of withstanding higher peak loads than those made
of neat polyurethane foam cores when subject to low-velocity impact despite lower
density than that of neat PU foams. This is especially significant for multi-impact
recurrences within the threshold loads and energies studied as shown on Fig. 9.
Cao et al. [58, 59] claimed that nanoparticle loading caused significant changes
in the cellular structures of the foam with cell dimension almost doubling with the
inclusion of 3 wt% of TiO2 nanoparticles. On an average, the increment in strength
and stiffness was 30 and 62%, respectively, over the neat system. Gain in strength

Fig. 9 Load versus time graph obtained from second impact test on sandwich structures using
aluminium faceplates and PU/MMT nanofoams
Elastomeric Nanocomposites for Aerospace Applications 327

was attributed to the delay in the formation of initial cracks during loading. It was
believed that nanoparticles embedded in the cell walls and edges and the associ-
ated interfaces surrounding the nanoparticles resisted crack formation/coalescence
at the earlier stage of the loading. Accordingly, this allowed higher sustenance of
load, but with the infusion of higher percentages of nanoparticles such as SiC, both
thermal and mechanical properties began to deteriorate. During the flexural tests it
was observed that failure always initiated on the tension side of the specimen.
Cracks first appeared along the width of the specimen and then began to propagate
through the thickness towards the compression side. Once the crack completed
traversing the entire width and proceeded towards the thickness, the failure was
very quick. Modes of crack propagation were almost identical with both neat and
nanophased foams. However, there was a significant delay in the first appearance
of the cracks in case of nanophased foams since they began to appear at a much
higher load. It was proposed that nanoparticles embedded in the cell structures
resisted the initial crack growth at the earlier stage of the loading that eventually
contributed to higher failure load.
SWNTs are considered to be the ideal reinforcing agents due to their excep-
tional mechanical properties, low density and high aspect ratio. Theoretical and
experimental studies have shown that SWNTs have extremely high Young mod-
ulus, similar to that of graphite in-plane (*1,000 GPa). Of special interest is the
study presented by Sen et al. [60], who fabricated SWNT-reinforced composite
nanofibres and membranes using the electrospinning process. The membranes
showed a nonlinear elastic behavior in the low stress region (0–2 MPa) and plastic
deformation at higher stress. Additionally, the maximum stress at break was the
tensile strength that increased for PU membranes on SWNT incorporation
(Fig. 10).
Further, compared to pure PU membranes, the tensile strength of as prepared
PU-SWNT nanocomposites increased by 46% from 7.02 to 10.26 MPa—this
enhancement in the mechanical properties was related to efficient load transfer to

Fig. 10 Tensile stress–strain


curves for PU membranes
containing as-prepared (AP)
and ester (EST) functional-
ized SWNTs. The SWNT-to-
PU weight ratio is 1:100 in
the composite membranes
(b, c) (reproduced from
Ref. [53])
328 J. Njuguna et al.

the SWNTs in the composite material. On the other hand, nanocomposites fabri-
cated with ester functionalized SWNT showed an increase of 104% in tensile
strength when compared to electrospun pure PU. Better mechanical properties for
the ester functionalized SWNT composites was attributed to improved dispersion
of the SWNTs, but it was suggested that this could also have been a response of the
polar functionalities in the modified SWNT to the opportunities offered by
hydrogen bonding sites in the polymer matrix, or to reactions between free amine
in the PU and the ester groups in the functionalized SWNTs.
Generally, nanocomposites are materials whose components interact at a
nanoscale level, a characteristic which confers great mechanical strength to these
materials due to the quantum-scale interaction between these molecules which also
gives it anti-fracture properties. The reason for this according to Griffin is that
when a composite has fillers below a critical length, their strength even if cracked,
is virtually equivalent to that of a solid crystal. This holds true for nanocomposites
and explains the relative immunity of nanocomposites to fracture. Along this line
of interest, one work [61] reported tests that showed significant improvements of
impact strength of 27% for the phenolic resin/linear-PU system and 54% for the
phenolic resin/C60-PU system, respectively, both with 3 wt% of linear-PU and
C60-PU content. These results may be related to the intensive intermolecular
hydrogen bonding that exists between phenolic resin and C60-PU as evident on
FTIR data. Significant improvement in the toughness of the phenolic resin/C60-PU
nanocomposite was also observed. However, Zilg et al. [29] reported both Shore A
hardness and Young’s modulus, which reflected polymer stiffness, decreased
slightly in modified PU/mica nanocomposites, whereas non-modified mica gave
increased stiffness and only small increases of tensile strength with increasing
filler content. In addition, the follow-up work [62] reported that interfacial cou-
pling can be controlled in the nanostructure PU formation as a function of orga-
nophilic modification thus improving selected properties, such as toughness/
stiffness balance, heat distortion temperature and flame retardancy. Finnigan et al.
[63] observed large improvements in stiffness on silicate addition, particularly for
the soft PU elastomer of Shore Hardness 80A, which contained a higher fraction of
soft segments. At a 7 wt% loading of organically modified layered silicate, a 3.2-
fold increase in Young’s modulus was noted. The improvement in stiffness was
attributed to the good dispersion and delamination achieved, in addition to the
strong interaction between PU matrix and OLS. Surprisingly, the addition of OLS
resulted in a decrease in tensile strength and elongation; a fact worth further
investigations. Nevertheless, the researchers attributed these effects to altered
microphase morphology, excessive polymer-filler interaction, and an inhibition of
the usual morphological changes that accompanied deformation in segmented PU,
i.e. hard domain rotation, interchain slippage, fibrillation and soft segment crys-
tallisation. The addition of layered silicates with high aspect ratios was observed to
increase the hysteresis and permanent set of these PU elastomers.
The enhancements of hardness properties in PU nanocomposites have been
witnessed elsewhere too [46, 64]. In one report [41], this reinforcing effect was
more evident when the elasticity of the PU matrix was increased at the temperature
Elastomeric Nanocomposites for Aerospace Applications 329

range above the Tg of hard segments and also in the situation where the defor-
mation was large. Kuan et al. [65] work results on waterborne PU and polysilicic
acid nanoparticles (PU/PSA nanocomposites) indicated that the wear index
declined as the PSA particles content increased (below 15 wt%) over 500 and
1,000 wear cycles; the wear index was proportional to the reciprocal of the wear-
resistance. SEM revealed that the particles were well dispersed in the polymer
matrix, suggesting that the nanocomposites exhibit good miscibility between
organic and inorganic phases, so the wear resistance was enhanced by wear-
endurable PSA nanoparticles. The wear-resistance of the nanocomposites was
greatest at 15 wt% PSA nanoparticle content. Adding polysilicic acid nanoparti-
cles to waterborne PU increased the initial tensile modulus of the composites. The
tensile modulus of the polysilicic acid nanoparticles/waterborne PU nanocom-
posite with 20 wt% silica increased from 20 to 41 MPa. Thus, the polysilicic acid
nanoparticles play an important role in strengthening the composites by effectively
transferring the stress between the silicon-containing nanoparticles and the PU
matrix.

2.3 Polyaniline Nanocomposites

Polyaniline (PANI) is a conducting polymer and its properties are strongly


dependent on synthetic procedures, type of dopant, morphology, and other vari-
ables. PANI being electrically conducting in nature can be used for conductive
adhesive, conductive ink, conductive paint, antistatic textile, and electrostatic
discharge (ESD) materials. Applications for such composites are wide spread,
these are used for interconnections, printed circuit boards, encapsulations, die
attach, heat sinks, conducting adhesives, electromagnetic interference (EMI)
shielding, ESD, and aerospace engineering. PANI-inorganic nanocomposites have
also been proven to possess a wide range of properties such as electrical,
mechanical, and structural properties because of synergistic effect owing to the
intimate mixing between organic components in molecular level. The degradation
behaviour of polyaniline nanocomposite is yet to be clearly understood and only a
handful published work is available [66].
Liu et al. [67] synthesized Fe2O3 magnetic nanoparticles with size range of 50–
100 nm. The initial decomposition temperature of the composite coating was
250°C. The solar absorptivity (as) of the composite coatings was as high as 90%,
and its emission rate (en) was reduced to 56%, while the relative efficiency of
light–heat transition (as/en) was about 1.6. The nanocomposite coatings exhibited
excellent abrasion resistance, weatherability and water resistance due to the for-
mation of a three-dimensional network structure during the thermal curing process.
The results indicate that the nanocomposite material could be used as a solar light–
heat transition coating that could be employed in solar hot-water collection. For
photodegradable packaging materials, nano-TiO2 can be used. The PANI–TiO2
nanocomposite powders showed highly enhanced photodegradation and the
330 J. Njuguna et al.

photodegradation increased with decreasing ratios of PANI:TiO2. A weight loss of


about 6.8% was found for the PANI–TiO2 (1:3) nanocomposite; however, the
weight loss of the PANI–HCl powder was only 0.3% after being irradiated for 60 h
under air.
The photocatalytic degradation of the nanocomposite powders accompanied the
peak intensity decrease in the FTIR spectra at 1,235 cm-1, attributed to C–N
stretching mode for benzenoid unit, and the depigmentation of the powders due to
the visible light scattering from growing cavities. The elemental analysis and
X-ray photon spectroscopy (XPS) analysis of the composite showed that the bulk
and surface concentration of N decreased with irradiation [68]. Lanthanum (La)-
doped Fe3O4 magnetic nanoparticles were prepared in aqueous solution at room
temperature, then La-doped Fe3O4–polyaniline (PANI) nanocomposites containing
a dispersion of La-doped Fe3O4 nanoparticles were synthesized via in situ poly-
merization of aniline monomer. The La-doped Fe3O4–PANI composite presented
core–shell structures; polyaniline covered the La-doped Fe3O4 completely. The
specific saturated magnetization of La-doped Fe3O4–PANI depended on the
starting material of La-doped Fe3O4.
Lee and Char [69] have found that the PANI/Na-MMT nanocomposites were
more thermally stable than the physical mixture of PANI and Na-MMT. Polyan-
iline undergoes a three-step thermal decomposition. The weight loss in the third
step, which is attributed to polyaniline backbone decomposition, was found to be
maximum at 530°C for pure PANI and this was shifted 25°C more for PANI/Na-
MMT nanocomposite. From the XRD investigation after TG analysis, it can be
concluded that the PANI chains residing outside the silicate layers decomposed
mostly, whereby chains residing inside the layers to a small extent. So the
shielding effect of intercalation into the layers imparts the thermal stability to
polymeric materials.

2.4 Poly(Ethylene Terephthalate) Nanocomposites

Polyethylene terephthalate (PET) exists both as an amorphous (transparent) and a


semi-crystalline (opaque and white) thermoplastic, and can be made into either as
resin, fibre or film. The semi-crystalline PET has good strength, ductility, stiffness
and hardness while the amorphous PET has better ductility. PETs are well suited
for flexible laminates, customized for specific performance parameters, are of
interest to aerospace programs such as lighter-than-air vehicles, balloon systems,
decelerator systems, flexible inflatable structures and pressure vessels, inflatable
structures and pressure vessels. The material requirements for these applications
include high strength-to-weight ratio and modulus, low gas permeability, pressure
retention and the capability to survive in harsh atmospheric, marine and/or
stratospheric environments for extended periods of time. These flexible laminates
achieve a significant weight savings over woven fabrics of similar strengths by
eliminating strength and modulus loss and other structural deficiencies caused by
Elastomeric Nanocomposites for Aerospace Applications 331

crimping of yarns during the weaving process. The absence of crimp in non-woven
fabrics results in a linear elastic response that allows for ease in predicting material
properties and simplification of structural models. These flexible composites afford
the ability to specify structural properties, oriented to meet any design require-
ment. Parts can be manufactured with complex.
Zeng et al. [70] prepared poly(ethylene terephthalate) (PET)-based nanocom-
posite fibres by melt spinning three types of PET/polyhedral oligomeric sils-
esquioxane (POSS) composites. These composites were made by either melt
blending POSS with PET at 5 wt% loading level (non-reactive POSS and silanol
POSS) or by in situ polymerization with 2.5 wt% reactive POSS. Significant
increases in tensile modulus and tensile strengths were achieved in PET fibres with
non-reactive POSS at room temperature. The high temperature modulus retention
was found to be much better for PET/silanol POSS fibre when compared to that of
control PET. Although other PET/POSS nanocomposite fibres tested did not show
this high retention of modulus at elevated temperatures, PET/isooctylPOSS
nanocomposite fibres did show increased modulus at elevated temperature com-
pared to that of PET. Higher compressive strengths, compared to PET fibres, were
observed for all three nanocomposite fibres. Gel permeation chromatography
measurement suggested that there was no significant change in molecular weight
during preparation of PET/POSS nanocomposites. SEM observations suggested
that there was no obvious phase separation in any of the three PET/POSS systems.
The fibre spinning and mechanical performance with 10 and 20 wt% of tris-
ilanolisooctyl POSS were also investigated. It was noted that the nanocomposites
with higher concentrations of this nanofiller can be spun without any difficulty. At
room temperature, the fibre tensile modulus increased steadily with the POSS
concentration while fibre tensile strength showed no significant change. The
elongation at break decreased significantly with increasing of POSS concentration.
The high-temperature moduli of PET/POSS nanocomposite fibres were found to be
rather variable, likely due to the modest compatibility between filler and polymers,
which lead to structural anisotropy within the composite.

2.5 Polyimide Nanocomposites

Fiber reinforced PMR polyimides are finding increased acceptance as engineering


materials for high performance structural applications. Prepreg materials based on
this novel class of highly processable, high temperature resistant polyimides [71].
Applications of polyimide foams in aerospace vehicles are include thermal-insu-
lating systems for cryogenic tank, honeycomb sandwich structures, radomes and
chairs. Ogasawara et al. [72] directed their investigations toward improvement of
heat resistance of phenylethynyl terminated imide oligomer (Tri-A PI) by loading
of MWNT. They fabricated the Tri-A PI/MWNT nanocomposites containing 0,
3.3, 7.7, and 14.3 wt% MWNT using a mechanical blender without any solution
(dry condition) for several minutes. The volume fraction of MWNT were
332 J. Njuguna et al.

Fig. 11 Effect of the MW-


CNT concentration on
Young’s modulus of the
composites (reproduced from
Ref. [65])

calculated to be 2.3, 5.4, 10.3 vol% from the density of the MWNT (1.9 g/cm3)
and the cured polyimide (1.3 g/cm3). Scanning electron micrographs showed the
particle size of the imide oligomers to be in the range of 0.1–10 lm, and MWNT
were not dispersed uniformly in the mixture. The loss of aspect ratio during the
mechanical blending was not significant, therefore the MWNT were flexible for
mechanical blend process with the imide oligomers. The preparation of the
nanocomposite involved the melt mixing of imide oligomer/MWNT at 320°C for
10 min on a steel plate in a hot press, and then curing at 370°C for 1 h under
0.2 MPa of pressure with PTFE spacer (thickness 1 mm). Tensile tests on the
composites showed an increase in the elastic modulus and the yield strength, and
decrease in the failure strain. Figure 11 shows the effect of the MWNT concen-
tration on Young’s modulus of the composites.
Dynamic mechanical analysis showed an increase in the glass transition tem-
perature with incorporation of the carbon nanotubes. The experimental results
suggested that the carbon nanotubes were acting as macroscopic crosslinks, and
were further immobilizing the polyimide chains at elevated temperature. As to the
reason why dispersed MWNT increased the heat distortion temperature, the
researchers explained that the dispersed MWNT impedes the molecular motion in
polyimide network at elevated temperature. The other property improvements in
this material are that MWNT showed some potential for controlling electric
conductivity and electro-magnetic wave absorbability. Although static properties
were obtained, discussions were not given, and it is evident that more research
work would be required to prove that the suggested phenomenon is a true cause of
higher glass transition temperature.

2.6 Polyarylacetylene Nanocomposites

Polyarylacetylene (PAA) is going through increasing development in the field of


advanced heat resistant composites owing to its outstanding heat resistance and
excellent ablative properties. PAA is developed increasingly as the matrix for high
Elastomeric Nanocomposites for Aerospace Applications 333

temperature composites of next generation in carbon–carbon composites, ablation


materials and structural materials owing to its outstanding heat resistance and
excellent process properties. Resulting mainly from the diethynylbenzene (DEB),
PAA is a highly cross-linked aromatic polymer that contains only carbon and
hydrogen when it is cured by means of addition polymerization, which makes it
own excellent thermal stability and oxidation. When the PAA is heated to high
temperatures in an inert environment, only about 10 wt% is volatilized, while the
remaining 90 wt% is carbon char, which means far less volatile material is gen-
erated and minimal shrinkage is associated with pyrolysis. Furthermore, PAA resin
is liquid or solvable and fusible solid at room temperature so that it provides better
processing flexibility, being applicable for conventional curing processes like
compression molding process, vacuum bagging method, resin transfer molding.
However, the bad wettability between carbon fibre and PAA resin from the non-
polar structure of PAA and chemical inertness of carbon fiber causes the weak
interfacial adhesion between fiber and non-polar PAA resin.
The main potential applications of PAA resin are in conventional resin matrix
composites with ultra-low moisture outgassing characteristics and improved
dimensional stability suitable for spacecraft structures, as an ablative insulator for
solid rocket motors, and as a precursor for carbon–carbon composites. Carbon
fibre reinforced PAA composites (carbon fibre/PAA) undoubtedly play a very
important role in all these fields. Unfortunately, the mechanical properties of the
carbon fibre/PAA material are not yet sufficiently satisfactory to replace the widely
used heat resistant composites such as carbon or graphite reinforced phenolic
resins. The mechanical properties of carbon fibre reinforced resin matrix com-
posites depend on the properties of carbon fibre and matrix, especially on the
effectiveness of the interfacial adhesion between carbon fibre and matrix.
Polyarylacetylene has high content of benzene rings and hence a highly cross-
linked network structure, which render the material brittle. Moreover, the chemical
inert characteristics of the carbon fibre surface lead to weak interfacial adhesion
between fibres and non-polar PAA resin. To ensure that the material could be used
safely in complicated environmental conditions and to exploit the excellent heat
resistant and ablative properties more effectively, it is necessary to improve the
mechanical properties of the carbon fibre/PAA composites. To achieve this pur-
pose, two kinds of methods can be used. One method is to improve the properties
of PAA resin by structural modification or by co-mixing other resins, such as
phenolic resin. The other is treatment of carbon fibre surface. The latter method
has been studied for a long time and several methods, such as heat treatment, wet
chemical or electrochemical oxidation, plasma treatment, gas-phase oxidation, and
high-energy radiation techniques have been demonstrated to be effective in the
modification of the mechanical interfacial properties of advanced composites for
engineering applications. In Zhang et al. [73] investigations, for instance, carbon
fibres were subjected to oxidation–reduction processes followed by vinyltrimeth-
oxysilanes–silsesquioxane (VMS–SSO) preparation method to improve the inter-
facial mechanical properties of the carbon fibre/PAA composites. The carbon fibre
surface treatment process is shown in Fig. 12.
334 J. Njuguna et al.

Fig. 12 Schematic diagrams of carbon fibre treatment process (sample 1 is oxygen plasma
oxidation, sample 2 is LiAlH4 reduction, sample 3 is VMS–SSO coating) (reproduced from Ref.
[66])

Polar functional groups, including carboxyl and hydroxyl, were implemented


on carbon fibre surface after the oxygen plasma oxidation treatment. The quantity
of carboxyl groups on carbon fibre surface was decreased and that of hydroxyl
moieties on carbon fibre surface was increased after the LiAlH4 reduction treat-
ment [74]. The VMS–SSO coating was grafted onto the carbon fibre surface by the
reaction of the hydroxyl groups in VMS–SSO and that on carbon fibre surface. The
VMS–SSO coating concentrations and treatment time were decided according to
Zhang et al. [75] who had optimized VMS–SSO coating treatment parameters. The
investigation found out that interlaminar shear strength of the PAA/carbon fibre
composites was increased by 59.3% at the end of treatment [73]. This kind of
modification method could be widely used in different resin matrix composites by
changing the functional groups on silsesquioxanes according to that on the resin.

2.7 Poly(ether ether ketone) PEEK Nanocomposites

Poly(aryl–ether–ether–ketone) (PEEK) is a high performance semicrystalline


thermoplastic with wide applications in aerospace, automotive, coating, electrical
insulating material and oil fields. Highly crystalline PEEK materials display
excellent mechanical, thermal and chemical resistance properties which allow
filled or unfilled PEEK to be used in various hostile environments. Even though
PEEK materials have been shown to perform satisfactorily up to temperatures of
200°C, there are a few current applications that require materials with even
higher upper temperature limits which are targets for improvement using
nanofillers.
Jen et al. [76] manufactured PEEK/silica nano-composite laminates and also
studied their mechanical responses. The experimental procedure were as follows:
firstly, the nanoparticles were diluted in alcohol (50 ml alcohol:2 g SiO2) and
stirred uniformly, then 16 plies of [0/90]4s cross-ply and [0/±45/90]2s quasi-iso-
tropic prepregs were cut, SiO2 solution was then spread on the prepreg in a
temperature-controlled box, and later weighed the nanoparticles after evaporation
Elastomeric Nanocomposites for Aerospace Applications 335

Fig. 13 Pressure–temperature profile of the curing process of AS-4/PEEK APC-2 nanocom-


posites (reproduced from Ref. [76])

of alcohol in the range of 111–148 mg/ply. A repeat on spreading for 5, 8, 10,


15 plies was the next step followed by curing (the curing process is shown in
Fig. 13) the stacked plies in a hot press to form a laminate of 2 mm thick.
Next, the laminates were cut into specimens and tested according to ASTM
D3039M. The tensile tests were repeated at 50, 75, 100, 125, 150°C to receive
respective stress–strain curves, strength and stiffness, and the obtained data
compared with the original PEEK laminate (no SiO2 nanoparticles) to find the
optimal SiO2 content. From tensile tests it was found out that the optimal content
of nanoparticles (SiO2) was 1% by total weight. The ultimate strength increased by
about 12.48% and elastic modulus by 19.93% in quasi-isotropic nano-laminates,
whilst, the improvement of cross-ply nano-composite laminates was less than that
of quasi-isotropic laminates. At elevated temperatures the ultimate strength
decreased slightly below 75°C and the elastic modulus reduced slightly below
125°C, however, both properties degraded highly at 150°C (&Tg) for the two lay-
ups. Finally, after the constant stress amplitude tension–tension (T–T) cyclic
testing, it was found that both the stress-cycles (S–N) curves were very close
below 104 cycles for cross-ply laminates with or without nanoparticles, and the
S–N curve of nano-laminates slightly bent down after 105 cycles.
Sandler et al. [77] produced poly(ether ether ketone) nanocomposites con-
taining vapour-grown CNF using standard polymer processing techniques.
Macroscopic PEEK nanocomposite master batches containing up to 15 wt%
vapour grown CNF were prepared using a co-rotating twin-screw extruder with a
length-to-diameter ratio of 33. The processing temperatures were set to about
380°C. The strand leaving the extruder was quenched in a water bath, air dried
and then regranulated followed by drying at 150°C for 4 h. Tensile bars according
to the ISO 179A standard were manufactured on an injection moulding machine at
336 J. Njuguna et al.

processing temperatures of 390°C, with the mould temperature set to 150°C. Prior
to mechanical testing, all samples were heat treated at 200°C for 30 min followed
by 4 h at 220°C in an attempt to ensure a similar degree of crystallinity of the
polymer matrix. Macroscopic tensile tests were performed at room temperature—
the cross-head speed was set to 0.5 mm/min in the 0–0.25% strain range and was
then increased to 10 mm/min until specimen fracture occurred. Evaluation of the
mechanical composite properties revealed a linear increase in tensile stiffness and
strength with nanofibre loading fractions up to 15 wt% while matrix ductility was
maintained up to 10 wt%. Electron microscopy confirmed the homogeneous dis-
persion and alignment of nanofibres. An interpretation of the composite perfor-
mance by short-fibre theory resulted in rather low intrinsic stiffness properties of
the vapour-grown carbon nanofibre. DSC showed that an interaction between
matrix and the nanoscale filler could occur during processing. However, such
changes in polymer morphology due to the presence of nanoscale filler need to be
considered when evaluating the mechanical properties of such nanocomposites.
Schmidt [78] investigation involved multifunctional inorganic–organic com-
posite sol–gel coatings for glass surfaces. The sol–gel process allowed the fabri-
cation of ceramic colloidal particles in the presence of organo alkoxy silanes
carrying various functions and the synthesis of multi-functional transparent inor-
ganic–organic composites. The report claimed that, in addition, these composites
can be used as controlled release systems or designed as gradient systems. Using
this approach, a coating with a very low surface free energy (antisoiling properties)
and temperature stability up to 350°C, a controlled release system for permanent
wettability (anti-fogging) and systems containing metal colloids for optical effects
were developed. Lin [79] and Wang et al. [80] studied the effect on wear and
friction by adding SiC nanoparticles in PEEK. The latter studied the effect of the
synergism between nanometer SiC and PTFE on the wear of PEEK. The PEEK
fine powders (ICI grade 450P, g = 0.62) in a diameter of approximately 100 lm,
were prepared. The nanometer SiC used as filler had the size smaller than 80 nm.
The PTFE powders (diameter 25 lm), nanometer SiC and PEEK were fully mixed
ultrasonically, dispersed in alcohol for *15 min. Then the mixture was dried at
110°C for 6 h to remove the alcohol and moisture. Finally, the mixture was
moulded into the block specimens by compression moulding, in which the mixture
was heated at a rate of 10°C min-1 to 340°C, held there for 8 min, and then cooled
in the mould to 100°C. After releasing from the mould, the resultant block
specimen was prepared for friction and wear tests. A tribological study found that
the incorporation of PTFE into 3.3 vol% nanometer SiC filled PEEK had a det-
rimental effect on the tribological properties of SiC–PTFE–PEEK composite. The
morphologies of worn surfaces and the properties of transfer films deteriorated,
while the load-carrying capacity of the SiC–PTFE–PEEK composite was also
adversely affected. The researchers claimed the reason for this was due to SiFx,
which was formed on the original surface and worn surface during the compres-
sion moulding process and sliding friction process as a result of the chemical
reaction between nanometer SiC and PTFE. The chemical reaction and the for-
mation of SiFx dominated the tribological behaviour of the SiC–PTFE–PEEK
Elastomeric Nanocomposites for Aerospace Applications 337

composites filled with various contents of PTFE and 3.3 vol% nanometeric SiC.
When the PTFE volume percentage was low, then the SiFx caused the friction and
wear of the SiC–PTFE–PEEK composite to rise. However, at high volume per-
cents the low friction PTFE dominated the friction and wear behaviours and the
friction decreased as the percentage of PTFE increased. The chemical reaction and
the formation of SiFx led to changes in the worn surface morphologies and det-
rimental effect on the characteristics of the transfer films.

2.8 Poly(p-phenylene benzbisoxazole) (PBO)


Nanocomposites

Poly(p-phenylene benzbisoxazole), a rigid-rod polymer, is characterized by high


tensile strength, high stiffness, and high thermal stability. Kumar et al. [81] found
out that PBO/CNT reinforced fibres exhibited twice the energy absorbing capa-
bility than the plain PBO fibres. The nanocomposites were prepared as follows:
into a 250 ml glass flask, equipped with a mechanical stirrer and a nitrogen inlet/
outlet, were placed *4.3 g (0.02 mol) of 1,4-diaminoresorcinol dihydrochloride,
*4 g (0.02 mol) of terephthaloyl chloride, and *12 g of phosphoric acid (85%).
The resulting mixture was dehydrochlorinated under a nitrogen atmosphere at
65°C for 16 h and subsequently at 80°C for 4 h. At this stage, 0.234 g of purified
and vacuum-dried HiPco nanotubes was added to the reaction flask. The mixture
was heated to 100°C for 16 h while stirring and then cooled to room temperature.
P2O5 (8.04 g) was added to the mixture to generate poly(phosphoric acid) (77%
P2O5). The mixture was stirred for 2 h at 80°C and then cooled to room temper-
ature. Further P2O5 (7.15 g) was then added to the mixture to bring the P2O5
concentration to 83% and the polymer concentration to 14 wt%. The mixture was
heated at 160°C for 16 h with constant stirring. Stir opalescence was observed
during this step. The mixture was finally heated to 190°C for an additional 4 h
while stirring. An aliquot of the polymer solution was precipitated, washed in
water, and dried under vacuum at 100°C for 24 h. An intrinsic viscosity of 14 dl/g
was determined in methanesulfonic acid at 30°C. A control polymerization of pure
PBO was also carried out under the same conditions without adding SWNT. For
PBO/SWNT (90/10) composition, 0.47 g of purified HiPco tubes (SWNT) was
added to the mixture. The sequence of steps and polymerization conditions
remained the same as those for PBO/SWNT (95/5) composition. Intrinsic viscosity
values of PBO and PBO/SWNT (90/10) were 12 and 14 dl/g, respectively. Single-
walled nanotubes were well dispersed during PBO synthesis in PPA. PBO/SWNT
composite fibres were successfully spun from the liquid crystalline solutions using
dry-jet wet spinning. The addition of 10 wt% SWNT increased PBO fibre tensile
strength by about 50% and reduced shrinkage and high-temperature creep. The
existence of SWNT in the spun PBO/SWNT fibres was evidenced by the
1,590 cm-1 Raman absorption band.
338 J. Njuguna et al.

3 Concluding Remarks

Advancements in the nanotechnology industry promise to offer improvements in


capabilities across a spectrum of applications. This is of immense strategic
importance to the aerospace sector which has historically leveraged technological
advances. In particular, nanomaterials represent a class of exciting new materials
with extraordinary electrical, thermal, and mechanical properties. Taking advan-
tage of these intrinsic properties relies on the ability to systematically synthesize,
characterize, and integrate standardized materials into actual commercial products.
Appropriate characterization of the properties (diameter, length, purity, surface
area, bundling, etc.) has been highlighted with microscopic, spectroscopic, ther-
mal, and surface area analysis, showing the importance of these techniques in
moving towards material utilization.
A few points can be emphasized here: (a) nanocomposites differ from con-
ventional composites in that the mixing of phases occurs over a much smaller
length scale in comparison to the micrometer length scale of conventional com-
posites, resulting into nanocomposite materials with properties substantially dif-
ferent from those of the parent end members, (b) proper engagement of the
nanoparticles in composite systems depends strongly on the ability to homoge-
neously disperse them throughout the matrix without destroying their integrity, and
(c) the properties of the matrix, the distribution and properties of the filler as well
as the nature of their interface control the behaviour of a typical composite
material.
In these terms, the nanoparticles often strongly influence the properties of the
composites at very low volume fractions. This is mainly due to their small
interparticle distances and the conversion of a large fraction of the polymer matrix
near their surfaces into an interphase of different properties in addition to the
consequent change in morphology. As a result, the desired properties are usually
reached at low filler volume fraction, which allows the nanocomposites to retain
the macroscopic homogeneity and low density of the polymer. Besides, the geo-
metrical shape of the particles plays an important role in determining the prop-
erties of the composites. Consequently, nanocomposites have attracted much
scientific and industrial interest in recent times.
It is emphasized that polymers are subjected to destructive factors such as
mechanical stress, the presence of different chemicals, UV-light, ablation and high
temperatures throughout shelf- and service-life. These factors cause degradation
and ultimately affect performance and lifetime of the polymers that are sometimes
stored for long periods of time. Therefore it is important to know how long and
under which conditions the polymers may best be stored with minimum deterio-
ration of the properties. According to the lifetime stages of polymers, the relevant
processes are classified as melt degradation; long-term heat ageing and weathering
based on the mechanisms involved, i.e. thermomechanical, thermal, catalytic and
radiation-induced oxidations and environmental biodegradation. The commercial
importance of polymers has been driving intense applications in the form of
Elastomeric Nanocomposites for Aerospace Applications 339

composites in aerospace needs. Performance during use is a key feature of any


composite material, which decides the real fate of products during use in aerospace
indoor and outdoor applications. Whatever the application, there is often a natural
concern regarding the durability of polymeric materials partly because of their
useful lifetime, maintenance and replacement. The deterioration of these materials
depends on the duration and the extent of interaction with the environment. Intense
research is required in this area to allow real nano enhanced products applications.

References

1. Njuguna, J., Pielichowski, K.: Polymer nanocomposites for aerospace applications:


properties. Adv. Eng. Mater. 5, 769–778 (2003)
2. Njuguna, J., Pielichowski, K.: Polymer nanocomposites for aerospace applications:
fabrication. Adv. Eng. Mater. 6, 193–203 (2004)
3. Njuguna, J., Pielichowski, K.: Polymer nanocomposites for aerospace applications:
characterisation. Adv. Eng. Mater. 6, 204–210 (2004)
4. Njuguna, J., Pielichowski, K., Desai, D.: Nanofiller-reinforced polymer nanocomposites.
Polym. Adv. Technol. 19, 947–959 (2008)
5. Kumar, A.P., Depan, D., Singh, T.N., et al.: Nanoscale particles for polymer degradation and
stabilization—trends and future perspectives. Progr. Polym. Sci. 34, 479–515 (2009)
6. Vaia, R.A., Wagner, H.D.: Framework for nanocomposites. Mater. Today 7, 32–37 (2004)
7. Kireitseu, M., Hui, D., Tomlinson, G.: Advanced shock-resistant and vibration damping of
nanoparticle-reinforced composite material. Compos. B Eng. 39, 128–138 (2008)
8. Liu, T., Ping Lim, K., Chauhari Tjiu, W., et al.: Preparation and characterization of nylon 11/
organoclay nanocomposites. Polymer 44, 3529–3535 (2003)
9. Usuki, A., Kawasumi, M., Kojima, Y., et al.: Swelling behavior of montmorillonite cation
exchanged for x-amino acids by e-caprolactam. J. Mater. Res. 8, 1174–1178 (1993)
10. Lincoln, D.M., Vaia, R.A., Wang, Z., et al.: Secondary structure and elevated temperature
crystallite morphology of nylon-6/layered silicate nanocomposites. Polymer 42, 1621–1631
(2001)
11. Fong, H., Liu, W., Wang, C.S., et al.: Generation of electrospun fibers of nylon 6 and nylon
6-montmorillonite nanocomposite. Polymer 43, 775–780 (2002)
12. Fornes, T.D., Paul, D.R.: Crystallization behavior of nylon 6 nanocomposites. Polymer 44,
3945–3961 (2003)
13. Li, L., Bellan, L.M., Craighead, H.G., et al.: Formation and properties of nylon-6 and nylon-
6/montmorillonite composite nanofibers. Polymer 47, 6208–6217 (2006)
14. Cadek, M., Le Foulgoc, B., Coleman, J.N., et al.: Mechanical and thermal properties of CNT and
CNF reinforced polymer composites. In: AIP Conference Proceedings, pp. 562–565 (2002)
15. Wu, S., Wang, F., Ma, C.M., et al.: Mechanical, thermal and morphological properties of
glass fiber and carbon fiber reinforced polyamide-6 and polyamide-6/clay nanocomposites.
Mater. Lett. 49, 327–333 (2001)
16. Akkapeddi, M.K.: Glass fiber reinforced polyamide-6 nanocomposites. Polym. Compos. 21,
576–585 (2000)
17. Vlasveld, D.P.N., Parlevliet, P.P., Bersee, H.E.N., et al.: Fibre–matrix adhesion in glass-fibre
reinforced polyamide-6 silicate nanocomposites. Compos. A 36, 1–11 (2005)
18. Vlasveld, D.P.N., Bersee, H.E.N., Picken, S.J.: Nanocomposite matrix for increased fibre
composite strength. Polymer 46, 10269–10278 (2005)
19. van Rijswijk, K., Lindstedt, S., Vlasveld, D.P.N., et al.: Reactive processing of anionic
polyamide-6 for application in fiber composites: A comparative study with melt processed
polyamides and nanocomposites. Polym. Test. 25, 873–887 (2006)
340 J. Njuguna et al.

20. Levchik, S.V., Weil, E.D., Lewin, M.: Thermal decomposition of aliphatic nylons. Polym.
Int. 48, 532–557 (1999)
21. Sandler, J.K.W., Pegel, S., Cadek, M., et al.: A comparative study of melt spun polyamide-12
fibres reinforced with carbon nanotubes and nanofibres. Polymer 45, 2001–2015 (2004)
22. Kiliaris, P., Papaspyrides, C.D., Pfaendner, R.: Influence of accelerated aging on clay-
reinforced polyamide 6. Polym. Degrad. Stab. 94, 389–396 (2009)
23. Lesiak, T., Maciejewski, L., Pielichowski, J.: Pro-ecological synthesis of intermediates for
the manufacture of polyurethanes: Part 1. Preparation of organic diisocyanates by addition of
isocyanic acid to unsaturated ethers and diethers. Inter. Polym. Sci. Technol. 22, 94 (1995)
24. Pielichowski, J., Vig, A., Huczkowski, P.: Badanie odpornosci na starzenie elastomerow
poliuretanowych; resistance to aging of polyurethane elastomers. Polimery 32, 469–471
(1987)
25. Pielichowski, K., Leszczynska, A.: Structure–property relationships in polyoxymethylene/
thermoplastic polyurethane elastomer blends. J. Polym. Eng. 25, 359–373 (2005)
26. Ma, C.-M., Kuan, Hsu-Chiang, Chang, Wei-Ping, et al.: Synthesis, thermal, mechanical and
rheological properties of multiwall carbon nanotube/waterborne polyurethane
nanocomposite. Compos. Sci. Technol. 65, 1703–1710 (2005)
27. Han, B., Cheng, A., Ji, G., et al.: Effect of organophilic montmorillonite on polyurethane/
montmorillonite nanocomposites. J. Appl. Polym. Sci. 91, 2536–2542 (2004)
28. Xu, R., Manias, E., Snyder, A.J., et al.: Low permeability biomedical polyurethane
nanocomposites. J. Biomed. Mater. Res. A 64, 114–119 (2003)
29. Zilg, C., Thomann, R., Muelhaupt, R., et al.: Polyurethane nanocomposites containing
laminated anisotropic nanoparticles derived from organophilic layered silicates. Adv. Mater.
11, 49–52 (1999)
30. Chen, T.K., Tien, Y.I., Wei, K.H.: Synthesis and characterization of novel segmented
polyurethane/clay nanocomposite via poly(ε-caprolactone)/clay. J. Polym. Sci. Part
A 37, 2225–2233 (1999)
31. Chen, T.K., Tien, Y.I., Wei, K.H.: Synthesis and characterization of novel segmented
polyurethane/clay nanocomposites. Polymer 41, 1345–1353 (2000)
32. Chen, Y., Shaw, D.T., Guo, L.: Field emission of different oriented carbon nanotubes. Appl.
Phys. Lett. 76, 2469–2471 (2000)
33. Li, J.: High performance epoxy resin nanocomposites containing both organic
montmorillonite and castor oil-polyurethane. Polym. Bull. 56, 377–384 (2006)
34. Song, L.R., Hu, Y.R., Li, B.R., et al.: A study on the synthesis and properties of polyurethane/
Clay nanocomposites. Int. J. Polym. Anal. Character 8, 317–326 (2003)
35. Tortora, M., Gorrasi, G., Vittoria, V.A., et al.: Structural characterization and transport
properties of organically modified montmorillonite/polyurethane nanocomposites. Polymer
43, 6147–6157 (2002)
36. Song, M., Hourston, D.J., Yao, K.J., et al.: High performance nanocomposites of
polyurethane elastomer and organically modified layered silicate. J. Appl. Polym. Sci. 90,
3239–3243 (2003)
37. Petrovic, Z.S., Javni, I., Waddon, A., et al.: Structure and properties of polyurethane-silica
nanocomposites. J. Appl. Polym. Sci. 76, 133–151 (2000)
38. Zhou, S.X., Wu, L.M., Sun, J., et al.: Effect of nanosilica on the properties of polyester-based
polyurethane. J. Appl. Polym. Sci. 88, 189–193 (2003)
39. Mishra, J.K., Kim, I., Ha, C.: New millable polyurethane/organoclay nanocomposite:
Preparation, characterization and properties. Macromol. Rapid Commun. 24, 671–675 (2003)
40. Kim, J., Jung, W., Park, K., et al.: Synthesis of Na+-montmorillonite/amphiphilic
polyurethane nanocomposite via bulk and coalescence emulsion polymerization. J. Appl.
Polym. Sci. 89, 3130–3136 (2003)
41. Kim, B.K., Seo, J.W., Jeong, H.M.: Morphology and properties of waterborne polyurethane/
clay nanocomposites. Eur. Polym. J. 39, 85–91 (2003)
42. Wang, J., Chen, Y., Wang, J.: Preparation and properties of a novel elastomeric polyurethane/
organic montmorillonite nanocomposite. J. Appl. Polym. Sci. 99, 3578–3585 (2006)
Elastomeric Nanocomposites for Aerospace Applications 341

43. Kim, Y., White, J.L.: Melt-intercalation nanocomposites with chlorinated polymers. J. Appl.
Polym. Sci. 90, 1581–1588 (2003)
44. Ni, P., Wang, Q., Li, J., et al.: Novel polyether polyurethane/clay nanocomposites
synthesized with organically modified montmorillonite as chain extenders. J. Appl. Polym.
Sci. 99, 6–13 (2006)
45. Jin, J., Song, M., Yao, K.J., et al.: A study on viscoelasticity of polyurethane-organoclay
nanocomposites. J. Appl. Polym. Sci. 99, 3677–3683 (2006)
46. Zhou, S., Wu, L., Sun, J., et al.: The change of the properties of acrylic-based polyurethane
via addition of nano-silica. Progr. Org. Coat. 45, 33–42 (2002)
47. Zhao, B., Liu, C.S., Zhu, X.S., Xu, K.W.: Research on the vibration cutting performance of
particle reinforced metallic matrix composites SiCp/Al. In: 10th International Manufacturing
Conference in China (IMCC 2002), Oct 11 2002, pp. 380–384. Elsevier Science Ltd, Fujian
(2002)
48. Hirai, T., Miyamoto, M., Watanabe, T., et al.: Effects of thiols on photocatalytic properties of
nano-CdS-polythiourethane composite particles. J. Chem. Eng. Japan 31, 1003–1006 (1998)
49. Chiang, L.Y., Wang, L.Y., Kuo, C.: Polyhydroxylated C60 cross-linked polyurethanes.
Macromolecules 28, 7574–7576 (1995)
50. Plummer, C.J.G., Garamszegi, L., Leterrier, Y., et al.: Hyperbranched polymer layered
silicate nanocomposites. Chem. Mater. 14, 486–488 (2002)
51. Tien, Y.I., Wei, K.H.: High-tensile-property layered silicates/polyurethane nanocomposites
by using reactive silicates as pseudo chain extenders. Macromolecules 34, 9045–9052 (2001)
52. Leszczyńska, A., Njuguna, J., Pielichowski, K., et al.: Polymer/montmorillonite
nanocomposites with improved thermal properties: Part I. Factors influencing thermal
stability and mechanisms of thermal stability improvement. Thermochim. Acta 453, 75–96
(2007)
53. Leszczyńska, A., Njuguna, J., Pielichowski, K., et al.: Polymer/montmorillonite
nanocomposites with improved thermal properties: Part II. Thermal stability of
montmorillonite nanocomposites based on different polymeric matrixes. Thermochim. Acta
454, 1–22 (2007)
54. Njuguna, J., Pielichowski, K., Alcock, J.R.: Epoxy-based fibre reinforced nanocomposites.
Adv. Eng. Mat. 9, 835–847 (2007)
55. Mohammed, A.A., Hosur, M.V., Jeelani, S.: Processing and characterization of nanophased
polyurethane foams. Cell Polym. 25, 293–306 (2006)
56. Uddin, M.F., Mahfuz, H., Zainuddin, S., Jeelani, S.: Infusion of spherical and acicular
nanoparticles into polyurethane foam and their influences on dynamic performances. In: 2005
SEM Annual Conference and Exposition on Experimental and Applied Mechanics, June 7–9
2005, pp. 2005–2011. Society for Experimental Mechanics Inc., Bethel, CT 06801, United
States, Portland (2005)
57. Njuguna, J., Pielichowski, K.: Ageing and performance predictions of polymer
nanocomposites for exterior defence and aerospace applications. In: Smithers Rapra
Technology, 2010, 11–12 March 2010. Hamburg, Germany
58. Cao, X., James Lee, L., Widya, T., et al.: Polyurethane/clay nanocomposites foams:
Processing, structure and properties. Polymer 46, 775–783 (2005)
59. Cao, X., Lee, L.J., Widya, T., Macosko, C.: Structure and properties of polyurethane/clay
nanocomposites and foams. In: ANTEC 2004, Annual Technical Conference Proceedings,
May 16–20, 2004, pp. 1896–1900. Society of Plastics Engineers, Chicago (2004)
60. Sen, R., Zhao, Bin, Perea, D., et al.: Preparation of single-walled carbon nanotube reinforced
polystyrene and polyurethane nanofibers and membranes by electrospinning. Nano Lett. 4,
459–464 (2004)
61. Ma, C.-M., Sung, S.C., Wang, F.Y., et al.: Thermal, mechanical, and morphological
properties of novolac-type phenolic resin blended with fullerenol polyurethane and linear
polyurethane. J. Polym. Sci. B 39, 2436–2443 (2001)
62. Zilg, C., Dietsche, F., Hoffmann, B., et al.: Nanofillers based upon organophilic layered
silicates. Die Makromolekulare Chemie. Macromol. Symp. 169, 65–77 (2001)
342 J. Njuguna et al.

63. Finnigan, B., Martin, D., Halley, P., et al.: Morphology and properties of thermoplastic
polyurethane nanocomposites incorporating hydrophilic layered silicates. Polymer 45, 2249–
2260 (2004)
64. Barna, E., Bommer, B., Kursteiner, J., et al.: Innovative, scratch proof nanocomposites for
clear coatings. Compos. Part A 36, 473–480 (2005)
65. Kuan, H., Chuang, W., Ma, C.M., et al.: Synthesis and characterization of a clay/waterborne
polyurethane nanocomposite. J. Mater. Sci. 40, 179–185 (2005)
66. Pielichowski, K.: Kinetic analysis of the thermal decomposition of polyaniline. Solid State
Ionics 104, 123–132 (1997)
67. Liu, G., Li, Y., Zhou, L., et al.: Photo-heat transition of coatings derived from furfural resin
and Fe2O3 nanoparticles. High Perform. Polym. 17, 469–481 (2005)
68. Zhang, L., Liu, P., Su, Z.: Preparation of PANI–TiO2 nanocomposites and their solid-phase
photocatalytic degradation. Polym. Degrad. Stab. 91, 2213–2219 (2006)
69. Lee, D., Char, K.: Thermal degradation behavior of polyaniline in polyaniline/Na+-
montmorillonite nanocomposites. Polym. Degrad. Stab. 75, 555–560 (2002)
70. Zeng, J., Kumar, S., Iyer, S., et al.: Reinforcement of poly(ethylene terephthalate) fibers with
polyhedral oligomeric silsesquioxanes (POSS). High Perform. Polym. 17, 403–424 (2005)
71. Weiser, E.S., Johnson, T.F., St Clair, T.L., et al.: Polyimide foams for aerospace vehicles.
High Perform. Polym. 12, 1–12 (2000)
72. Ogasawara, T., Ishida, Y., Ishikawa, T., et al.: Characterization of multi-walled carbon
nanotube/phenylethynyl terminated polyimide composites. Compos. A 35, 67–74 (2004)
73. Zhang, X., Huang, Y., Wang, T., et al.: Influence of fibre surface oxidation–reduction
followed by silsesquioxane coating treatment on interfacial mechanical properties of carbon
fibre/polyarylacetylene composites. Compos. A 38, 936–944 (2007)
74. Lin, Z., Ye, W., Du, K., et al.: Homogenization of functional groups on surface of carbon
fiber and its surface energy. J. Huaqiao Univ. 22, 261–263 (2001)
75. Zhang, X., Huang, Y., Wang, T., et al.: Influence of oligomeric silsesquioxane coatings
treatment on the interfacial property of CF/PAA composites. Acta Mater. Compos. Sin. 23,
105–111 (2006)
76. Jen, M.R., Tseng, Y., Wu, C.: Manufacturing and mechanical response of nanocomposite
laminates. Compos. Sci. Technol. 65, 775–779 (2005)
77. Sandler, J., Werner, P., Shaffer, M.S.P., et al.: Carbon-nanofibre-reinforced poly(ether ether
ketone) composites. Compos. A 33, 1033–1039 (2002)
78. Schmidt, H.: Multifunctional inorganic-organic composite sol-gel coatings for glass surfaces.
J. Non Crystal. Solids 178, 302–312 (1994)
79. Lin, J.: In situ syntheses and phase behavior investigations of inorganic materials in orgamic
polymer solid matrices. PhD dissertation, Department of Chemistry, The Pennsylvania State
University, (1992)
80. Wang, Q., Xue, Q., Liu, W., et al.: The friction and wear characteristics of nanometer SiC and
polytetrafluoroethylene filled polyetheretherketone. Wear 243, 140–146 (2000)
81. Kumar, S., Dang, T.D., Arnold, F.E., et al.: Synthesis, structure, and properties of PBO/
SWNT composites. Macromolecules 35, 9039–9043 (2002)
Friction and Wear of Rubber
Nanocomposites Containing Layered
Silicates and Carbon Nanotubes

D. Felhös and J. Karger-Kocsis

Abstract This chapter gives a survey on the tribological performance of or-


ganoclay and carbon nanotube reinforced rubbers of both conventional (thermoset)
and thermoplastic versions. The unlubricated friction and wear of rubbers were
grouped in abrasion-, sliding- and rolling-types in order to support the overview. It
was highlighted that the coefficient of friction and specific wear rate strongly
depend on the configuration and testing parameters of the tribotests used. It was
demonstrated that the incorporation of the above nanofillers is not always asso-
ciated with improved resistance to wear and reduced coefficient of friction. Further
experimental studies, data mining through proper statistical techniques, and
extensive modeling works are needed to realize the potential of the above
nanofillers in tribological applications, and to deduce relationships between wear
and other characteristics (e.g. network-, mechanical response-related) of rubbers.

D. Felhös (&)
Department of Polymers Engineering, Faculty of Materials Science and Engineering,
University of Miskolc, 3515 Miskolc, Hungary
e-mail: kohfelda@uni-miskolc.hu
J. Karger-Kocsis
Department of Polymer Technology, Faculty of Engineering and Built Environment,
Tshwane University of Technology, Pretoria 0001, Republic of South Africa
J. Karger-Kocsis
Department of Polymer Engineering, Faculty of Mechanical Engineering, Budapest
University of Technology and Economics, 1111 Budapest, Hungary

V. Mittal et al. (eds.), Recent Advances in Elastomeric Nanocomposites, 343


Advanced Structured Materials, 9, DOI: 10.1007/978-3-642-15787-5_13,
 Springer-Verlag Berlin Heidelberg 2011
344 D. Felhös and J. Karger-Kocsis

1 Introduction

1.1 Nanotubes

Nowadays considerable research interest is focused on carbon nanotubes (CNT)


and their use in structural and functional polymer composites. CNT-reinforced
elastomers are no exceptions in this respect. Already a large body of work is
devoted to investigate the applicability of this novel filler in various rubber
systems. The most exciting properties of CNTs are their high strength and
extremely high modulus (*1 TPa), excellent electrical and thermal conductivities,
high specific surface area and very high aspect ratio (ratio of length to diameter)
([1] and references therein). These properties of CNTs are very promising also for
rubber products. The ongoing research and development works are dealing with
property improvements and generation of novel functional properties.
The commonly used CNT types are single-walled (SWCNT) and multi-walled
(MWCNT) versions. However, one can find more special types like radial single-
walled and nanohorn-like CNTs. The dispersion quality of the nano-sized fillers in
the matrix material controls the properties of the corresponding nanocomposites.
The larger is the active filler surface in contact with the matrix, the larger the
reinforcing effect is. It is obvious that avoiding the agglomeration of nanofillers
with proper dispersion techniques has paramount importance. Inside the agglom-
erates there are not enough polymer matrix molecules to guarantee the stress
transfer towards the reinforcing nanofillers. As a consequence the filler cannot bear
the mechanical load. In addition, the non-wetted inner part of the agglomerate has
stress concentration effect and thus it causes premature failure during loading.
Moreover, good nanofiller dispersion is more economical as the same filler-matrix
contact surface is reached with less filler.
Good dispersion of the nanotubes can be reached by applying dispersion-,
solution- or melt-blending methods. The greatest challenge is to overcome the van
der Waals forces between the CNTs in order to disintegrate their agglomerates. To
decrease the van der Waals interactions between the nanotubes, the CNTs are
usually surface treated for example in acid bath, adding coupling agents, or
applying plasma-, thermal- and laser-treatments [2–6]. The agglomerates can be
broken in suitable dispersion media (aqueous solutions, solvents) mechanically.
This occurs by ball milling [7, 8], sonication [9–12], or applying high rotation
speeds in a dissolver [13, 14]. The surface treated and deagglomerated nanotubes
can be mixed with the rubber in a kneader or on an open mill. It was observed, that
the common use carbon black (CB) or silica with CNT helps to disperse the
nanotubes homogeneously [15–17]. Other possibility is the in situ polymerization
of elastomer in presence of CNT [18–20] which results in a homogeneous
dispersion of the latter.
Similar to CB, CNTs are also active fillers. This means that their incorporation
in elastomers leads to an increase in the ‘‘apparent’’ crosslink density. As a con-
sequence, the hardness, stiffness and strength of the rubbers increase with
Friction and Wear of Rubber Nanocomposites 345

increasing filler content [21–28]. On the other hand, when CNTs are incorporated
in large amount in the rubber matrix, they easily reagglomerate. Though the
stiffness and the strength increase with filler loading, even in case of reagglom-
eration, the ultimate strain to failure of the related rubber nanocomposite is
strongly reduced [11, 24, 25, 33]. Similar results are obtained when CNTs are
added in thermoplastic elastomers [21, 33]. This scenario may change, however,
when semicrystalline polymer phase is present in the thermoplastic elastomer, as
in PP/EPDM blends [29]. In this case the CNTs are incorporated in the thermo-
plastic phase, influencing heavily the crystallization process of the thermoplast.
Because of this feature, only a small amount of CNT filler should be introduced
(\0.5%) to improve the mechanical properties of the corresponding thermoplastic
elastomer. Note that similar conclusion was deduced in a study in which carbon
nanofiber was mixed with an olefinic thermoplastic dynamic vulcanizate [30].
Orientation of the nanotubes in tension direction was observed during uniaxial
stretching of CNT reinforced elastomers [24]. Similar to traditional composites,
the orientation of the nanotubes before curing influences largely the mechanical
properties of the of the CNT-elastomer nanocomposites [1].
Swelling tests showed similar results for CNT-filled rubbers, namely the
equilibrium swollen volume of the nanocomposites decreased with increasing filler
loading, confirming a strong interaction between the filler and the matrix [13, 31].
Dynamic-mechanical thermal analysis (DMTA) indicated that the maximum of
the mechanical loss (tan d) peak decreases with increasing CNT filler content.
Interestingly it is not clear yet, how the incorporated CNT affect the glass
transition temperature (Tg). In the literature contradictory results are reported
related this question. Slight increments or unchanged Tg values were reported in
Refs. [7, 8, 26, 28, 31]. On the other hand, a significant Tg decrease was observed
in Ref. [25].
Further open questions are linked with the explanation of the widely investi-
gated Mullins and Payne effects [24, 25]. The former may be caused by the
relaxation of the molecule chains, or through molecule-detachments from the filler
surface. The Payne effect is commonly explained through the breakdown of the
filler particle agglomerates due to large strains imposed on the material [16].
For the producers of rubber goods it is of great interest how the curing
parameters are affected by the CNT filler. It turned out that the necessary vulca-
nization temperature increases with the incorporation of CNTs [7, 26, 28], and
parallel to that the curing time decreases [8].
The electrical conductivity of rubber compounds is an important feature for tire
compounds. The CB containing rubber compounds have relatively small electric
resistance, but applying novel fillers such as silica changes this situation [32].
Discharge phenomena after static charging are inconvenient for the passengers and
may damage the sensitive electronic equipments of modern cars, as well. To avoid
static charging the electric conductivity of polymeric materials should be
enhanced. General observation is that incorporation of CNTs increases the con-
ductivity both of elastomers [16, 17, 33] and thermoplastics [34, 35]. The rapid
increase in the electric conductivity due to increasing CNT concentration is
346 D. Felhös and J. Karger-Kocsis

explained by the aspect ratio of the nanotubes ([100), which is higher than that of
short fibers (\30) or of CB (*1) [36]. Due to the high aspect ratio of CNTs more
electrical contacts are created in the nanocomposites and the electric percolation
threshold reached at very low filler loading. Similar to the mechanical properties of
fiber reinforced composites, Kim et al. [1] reported, that the orientation of the
CNTs enhances also the electrical properties of the nanotube-filled elastomer
composite.

1.2 Organoclays

Layered silicates in their pristine (crude) form are no active fillers for polymeric
materials. Their special layered structure can be exploited for reinforcement
only after chemical treatment. During this treatment the hydrophilic alkali metal
cations at the surface of the silicate layers are replaced by bulky organophilic
cations bearing usually long apolar chains. Through this cation exchange the
sheet-to-sheet (interlayer) distance increases from the original ca. 1.5 nm to ca.
2.5 nm in the corresponding organoclays (OCs) [37]. This organophilic treat-
ment renders the silicate layers hydrophobic and the interlayer van der Walls
forces are dramatically decreased. This supports the intrusion of the rubber
molecules between the layers during in situ polymerization [38–40], or melt
mixing [41–43]. Depending on the compatibility of the organoclay with the
polymer molecules, an intercalated or fully exfoliated clay structure is the final
outcome.
The modulus and the strength of the nanocomposites can highly be increased
when the clay layers are fully exfoliated (delaminated) in the elastomer matrix [39,
41]. The larger is the distance between the modified clay layers initially the easier
the generation of a well dispersed nanocomposite structure is [37]. Due to this, the
physical properties of the nanocomposites are largely influenced trough the
treatment of the nanoparticles [41, 44, 45]. In contrast to intercalated structures,
the stiffness and elongation to break values increase monotonously with filler
content for fully exfoliated nanocomposites [41]. Similar observations were
reported by Yang [46], who pinpointed also the change in the viscoelastic
properties between exfoliated and intercalated nanocomposites. The excellent
properties of nanocomposites with exfoliated clay are underlined by Yousoh [39].
On the other hand, several studies suggested that the OC layers are energetically
more stable when they are ranged close to each other. This can lead to de-exfo-
liation or de-intercalation processes of the exfoliated or intercalated structures
during the relaxation of the polymer molecules [41, 47].
Contrary to increasing tensile strength of elastomeric organoclay (OC) nano-
composites, decreasing tear strength [43] and decreasing fracture toughness [48]
were observed for epoxidized natural rubber (ENR)/OC and for polyamide (PA-6)/
maleinated polyolefin elastomer/OC composites, respectively, with increasing OC
content.
Friction and Wear of Rubber Nanocomposites 347

Special properties of the elastomer/OC nanocomposites are the following:


increased thermal stability [38, 40, 43, 49–53], increased flame retardancy
[53, 54], improved barrier properties to gases and liquids [40, 51, 53, 55].
The increased thermal stability due to organoclay incorporation, however, did not
affect the recyclability via remelting as shown on example of thermoplastic
elastomers. This was traced to the fact that the nanofiller was located in the
thermoplastic phase [56].
In many applications the optical properties of the elastomers are important.
Wang [42] reported that the OC particles tend to agglomerate with their increasing
content. This decreases the transparency of the initially transparent elastomeric
matrix. Other studies suggested that the critical percolation filler content can be
detected by considering changes in the linear viscoelastic material properties [57]
or in the initial modulus of composites [58].

2 Friction and Wear of Elastomers

2.1 Abrasion-Type

Dominant part of the tribological investigations on elastomers was done for the
car tire industry. Cardinal questions in this respect are the wear and frictional
properties of the tires. Note that these properties are highly valued as they affect
the safety of the car passengers, as well. Because of abrasive wear controls the
performance of car tires, this is the most widely studied wear phenomenon.
To compare the abrasion resistance of different rubber mixtures, without field
trials, several standard laboratory abrasive test methods were developed, such as
the Taber Abrader method [59]. Simplified laboratory test methods provide only
comparative test results on the investigated rubber goods. To observe and ana-
lyze the wear mechanisms under abrasive conditions specific methods were
developed. In the corresponding tests a razor blade or a needle abrade the rubber
surface, simulating a single roughness asperity of the abrasive surface. Detailed
description of the latter test methods is disclosed in Refs. [74–77; blade on
rubber plate test or blade abrader], and in Refs. [69, 78, 79; scratch or needle on
rubber plate test]. Applying these tests one can mimic and investigate the effect
of one single roughness peak of an abrader. The dimensions of the rubber
tongues, angles of the crack propagation, crack density, etc… were determined
and different mechanical models proposed to explain the abrasion properties of
the investigated rubbers [67, 74, 82–84]. Generally the abrasion is a very harsh
action which produces a rough surface. For the heavy surface destruction hardly
any schematic interpretation can be given. A typical unidirectional abraded
surface structure is depicted in Fig. 1 showing abrasive grooves along with wear
debris. This kind of wear can be ‘‘produced’’ also by using the laboratory
abrasion test rig shown in Fig. 2.
348 D. Felhös and J. Karger-Kocsis

Fig. 1 Unidirectional
abraded surface of
conventional, thermoset
elastomers

Fig. 2 Abrasive Pin on Plate (POP) tests for TPO based thermoplastic elastomer nanocompos-
ites. Note: the wear of the specimens was collated based on the diameter of the wear tracks

2.1.1 Conventional Rubbers

Extended investigations were performed by Faulkner et al. [15] who studied the
mechanical and tribological properties of CNT reinforced fluororubber (FKM)
and hydrogenated acrylonitrile–butadiene (HNBR). The deduced abrasive wear
mechanisms suggested that the abrasion resistance of the investigated rubbers
depends on their tear strength directly. However, the increase of the tear strength
was not accompanied by a monotonous increase in the abrasion resistance. This
means that also other mechanical parameters play an important role in the wear
resistance of the material. For bisphenol cured FKM higher abrasion resistance
was noticed at 10 phr CNT than at 30 phr N-990 CB content (DIN abrasion test
(ASTM D5963-04) results are 11.2 and 18.1 mg, respectively). In this case the
change in the tear strengths was also remarkable: 55.3 and 18.4 kN/m for the
10 phr CNT and 30 phr N-990 CB fillers, respectively. Slightly different abrasive
wear resistance was found for the FKM when instead of bisphenol, peroxide
curing was used. In this case the DIN abrasion tests resulted in 7.5 and 7.7 mg
material losses for 10 phr CNT and 35 phr N-990 CB contents, respectively. On
the other hand, the tear strength values were 66.3 and 26.3 kN/m for the CNT
and N-990 CB containing compounds. Almost three times increase in the tear
strength yielded only a negligible improvement in the abrasion resistance. The
same authors [15] reported successful improvement in the abrasion resistance of
a medium crosslinked HNBR elastomer when both MWCNT and CB were at the
Friction and Wear of Rubber Nanocomposites 349

same time incorporated in the matrix (50 phr N-550 type CB). When only
nanotubes were added to the HNBR mixture, without the presence of the CB, the
tear strength of the nanocomposite was more increased. On the other hand, the
abrasion resistance was less when the HNBR mixture contained only nanotubes
without CB.
Pal et al. [43] prepared and tested natural rubber (NR)/styrene–butadiene rubber
(SBR) (with high styrene contents) compounds). The mixes contained 40 phr CB
(type of the CB was varied: N-231 and N-774) both with 10 phr ENR epoxidized
natural rubber/Cloisite 20A (ratio 1:1) compounds. DIN (ASTM 5963-04),
Du-Pont (ASTM D394) abrasion tests were performed on standard devices and on
a home built machine, where the rotating rubber wheels (mounted with aluminum
core) were pressed against different rock materials (sandstone, concrete and
granite) with constant normal loads. The authors observed that for the standard
DIN and Du-Pont wear tests the OC containing rubber mixtures showed better
wear resistance than the traditional compounds without OC fillers. Against
different rock materials the weight loss trend of the investigated rubbers differed
from that received in standard abrasion tests. Against sandstone and concrete the
organoclay containing rubbers were more wear resistant than those with traditional
fillers. On the other hand, against granite, all the rubber compounds exhibited
similar wear resistances.

2.1.2 Thermoplastic Rubbers

A polyolefin elastomer, namely poly(ethylene–octene) copolymer (Engage 8401;


referred further as TPO thermoplastic elastomer), was filled with OC and
MWCNT, respectively. The coding, nanofiller content and the mechanical
properties of the related nanocomposites is disclosed in Sect. 2.2.2.
Abrasive pin(rubber)-on-plate (abrasive) type (POP) tests (cf. Fig. 2) were
performed to test the abrasive wear behavior of the TPO-based nanocomposites
containing OC and MWCNT fillers. The 1 mm thick elastomer composite speci-
mens were glued to the surface of a fixed ceramic ball (Ø * 8.77 mm). The so
prepared specimens were pressed against the plate counterbody which was covered
with sandpaper (P 180, average grind size 82 lm). The radius of the circular path
was 15 mm, the sliding speed was 20 cm/s, the normal load was 1 N and the
sliding distance was 100 m. During the tests the COF was online registered and
the loss volumes of the different materials were compared with each other based on
the related diameter of the wear paths (cf. Fig. 2). The worn surface of the
specimens was inspected by optical microscopy. The wear behavior of the TPO-
based nanocomposites is collated in Fig. 3. Considering the diameter of the wear
track the OC containing nanocomposites exhibited increased wear resistance in
contrast to the MWCNT filled ones the wear of which was enhanced. With
increasing OC content the wear resistivity of the composites decreased. So the
lowest wear was observed for the TPO-2 (containing 2 wt% OC) composite.
Among the MWCNT-containing materials, the lowest wear was measured for the
350 D. Felhös and J. Karger-Kocsis

Fig. 3 Diameter distribution


of the wear tracks after
abrasive POP tests for the
TPO based nanocomposites
containing OC and MWCNT
in different amounts.
Designations: TPO-2, -5, -10
contain 2, 5 and 10 wt% OC
whereas TPO-C05, -C1, -C2
contain 0.5, 1 and 2 wt%
MWCNT, respectively

TPO-C1 (containing 1 wt% MWCNT). These results suggest that the nanofiller
content may have an optimum value, which should be determined accordingly.
The time dependence of the coefficient of friction (COF) values during the
abrasive POP tests shows a fast increase initially, and a pretty stable value in the
steady state (cf. Fig. 4). The fast COF increase may be caused by the temperature
rise at the contact surface. The OC nanofiller does not influence the COF value,
whereas with varying MWCNT content the COF varies, as well. The smallest COF
value, and at the same time the smallest wear for the TPO/MWCNT series, were
measured for the TPO-C1 (containing 1 wt% MWCNT). Compared to this
nanocomposite, the steady state COF value and the wear increased for both
TPO-C05 and TPO-C2 composites.
The abrasive wear patterns of the TPO-based nanocomposites are very similar
to each other. Figure 5 displays a light microscopic picture which was taken from
the TPO-5 (containing 5 wt% OC) nanocomposite. Here a typical unidirectional
abraded pattern can be resolved. Compared to the worn surface of conventional
thermoset rubbers in this thermoplastic rubber rather plastic deformation
dominates instead of coarse edged cracking. The surface is covered by grooves and
large size wear particles (cf. Figs. 5, 6). During sliding large amount of heat is
generated and accumulated by the friction supporting the plastic deformation of
the material under adiabatic conditions. This may cause the melting of the TPO,
reflected by enhanced irreversible deformations. Accordingly, due to the ther-
moplastic behavior of TPO its worn surface differs from that in Fig. 1, which is
more characteristic for conventional thermoset rubbers.

2.2 Sliding-Type

Sliding wear gained importance nowadays, because efforts are made to avoid the
lubrication between the sliding rubber and other elements. This trend is reasoned
by economical and environmental issues. Therefore, investigations on the sliding
type wear under dry conditions are very important. Recall that for sealing and
Friction and Wear of Rubber Nanocomposites 351

Fig. 4 Coefficient of friction (COF) as a function of sliding time for the TPO/OC (left side) and
TPO/MWCNT (right side) nanocomposites during the abrasive POP tribotest. For designations
cf. Fig. 3

related applications the elastic and viscoelastic properties of the rubbers are
exploited. In these applications usually high adhesion arises between the rubber
and the counterbody owing to their smooth surfaces. Under relative displacements
between the contacting pairs adhesive sliding and wear takes place. To meet the
industrial needs new rubber formulations, novel coatings and rubber–polymer
sliding pairs have to be developed.
For sliding wear tests the ball-on-prism configuration is often favored [60, 61],
as in the related device allows to test many specimens at the same time. In this
ball(counterbody)-on-prism(rubber) test usually a revolving steel ball is pressed
against two rubber plates forming a prism. Disadvantage of this technique is that
no friction force can be detected. Other test possibility is the roller-on-plate (ROP;
also termed to shaft-on-plate (SOP) test) configuration, where a rotating shaft is
352 D. Felhös and J. Karger-Kocsis

pressed against a rubber plate, whereby the normal and friction forces are detected
(see Fig. 7). Generally it is not always easy to distinguish between the different
methods being differently classified in respect to the wear type. The simplified
laboratory POP test (see Fig. 7) can be labeled as adhesive type sliding tribotest if

Fig. 5 Light microscopic


picture of the worn surface of
the TPO-5 (containing 5 wt%
OC) nanocomposite. Note:
sliding direction is downward

Fig. 6 Wear mechanism


concluded for the
thermoplastic polyolefin
elastomer containing OC and
MWCNT nanofillers

Fig. 7 Schematic set-up of the tribotesting devices used. Designations: pin(steel)-on-plate(rub-


ber) (POP), roller(steel)-on-plate(rubber) (ROP)
Friction and Wear of Rubber Nanocomposites 353

the surface of the abrader counterbody is smooth enough. However, there is no


straightforward definition on how to differentiate between abrasive-like and
‘‘smooth’’ surfaces, and thus between abrasive and sliding wears.

2.2.1 Conventional Rubbers

The sliding, friction and wear behaviors of peroxide cured hydrogenated nitrile
rubber (HNNBR) with 10 and 30 phr MWCNT were investigated [62]. Mechan-
ical properties (hardness, tensile modulus, ultimate tensile strength and strain, tear
strength) of the rubbers were determined. Dynamic mechanical thermal analysis
(DMTA) was also performed and the apparent crosslink density estimated.
The related mechanical and material related data are summarized in Table 1.
Friction and wear characteristics were determined in pin(steel)-on-plate(rubber)
(POP) configuration in which a steel pin (100Cr6; arithmetical roughness, Ra, less
than 1 lm) with a hemispherical tip of 10 mm diameter rotated along a circular
path, see Fig. 7. The pin was pushed against the rubber plate with a given load.
The following parameters were selected for this configuration––normal load: 2 N,
sliding speed: 250 mm/s, duration: 1.5 h (h). Measuring both the normal and the
friction force components via a torque load cell the COF values were calculated
and monitored during the test. To study the sliding wear a further test, viz. ROP,
was also used, see Fig. 7. A rotating steel roller (100Cr6, diameter: 10 mm, width:
20 mm, Ra & 0.9 lm) pressed against a rubber strip of 9 mm width in the related
tribotester. The frictional force induced by the torque was measured online and
thus the COF was registered during the test. The test parameters were––load: 2 N;
sliding speed: 250 mm/s; duration: max. 1.5 h. It is noteworthy that the selected
test parameters were adjusted to the praxis by considering the contact pressure and
sliding speed.

Table 1 Network-related and mechanical data for the HNBR rubbers containing MWCNT as
filler. For the description of the test procedure see reference [62]
HNBR HNBR ? 10phr MWCNT HNBR ? 30phr MWCNT
Mc [g/mol] 1931 1113 150
nc [10+26 9 m-3] 3.3 5.7 45.3
tan d at Tg [-] 1.36 1.1 0.5
Density [g/cm3] 1.057 1.057 1.135
Shore A [] 42 56 82
Martens hardness [MPa] 1.45 2.16 10.2
M-0.01 DMTA [MPa] 3.9 6.7 52.4
M-100 [MPa] 1.27 2.4 12.7
M-200 [MPa] 1.3 2.58 12.4
Tensile strength [MPa] 4.4 4.9 16.5
Tensile strain [%] 280 179 132
Tear strength [kN/m] 4.2 8.0 22.0
354 D. Felhös and J. Karger-Kocsis

Figure 8 shows the effect of MWCNT on the COF. One can recognize that the
COFs are lower in POP tests, while they are markedly higher for the ROP
configuration. This is due to the related test set-up favoring substantial heating
under ROP. For the latter case even rubber against rubber type friction may be
observed (when the roller is coated with rubber). Moreover, the change in the COF
with the MWCNT content is different when various test devices are used. For the
POP configuration the COF is increasing with increasing MWCNT content
compared to the neat HNBR. Using the ROP device the measured COF is smaller
when MWCNT is added to the HNBR, but with increasing MWCNT content the
COF is increasing.
Comparing the specific wear rate results one can see, that their differences may
cover some order of magnitudes (cf. Fig. 8). Higher wear is caused by the POP
than by the ROP tests. Using MWCNT additives the wear performance of the
related HNBR is always better. With increasing filler content the resistance to wear
of the HNBR compound is enhanced. It should be born in mind that the specific
wear rate is usually decreasing with increasing active filler content as demon-
strated for CB-containing ethylene–propylene–diene (EPDM) rubbers [63, 64].
Characteristic scanning electron microscopic (SEM) pictures taken of the worn
tracks of the HNBR stocks are given in Figs. 9, 10 for the POP and ROP tests,
respectively.

Fig. 8 Measured specific


wear rate (Ws). a and steady-
state coefficient of friction
(COF) values and b for the
POP and ROP tribotests
Friction and Wear of Rubber Nanocomposites 355

Fig. 9 SEM pictures taken from the wear tracks after POP tests. a pure HNBR,
b HNBR ? 10phr MWCNT, and c HNBR ? 30phr MWCNT. Note: sliding direction is
downwards [62]. (Reprinted from: Felhös, D. Karger-Kocsis, J. Xu, D.: Tribological testing of
peroxide cured HNBR with different MWCNT and silica contents under dry sliding and rolling
conditions against steel. Journal of Applied Polymer Science, 2008, 108, pp. 2840–2851,
Copyright 2010, with permission from John Wiley and Sons)

POP tests: The surface of the evaluated wear track on the pure HNBR shows
that quite large particles were chipped off from the surface and a crater-like pattern
appeared (Fig. 9a). At 10 phr MWCNT content, however, the so-called Schalla-
mach waviness [65, 66] becomes recognizable (Fig. 9b). With further increase of
the MWCNT content the Schallamach waves become less resolved and a band-like
pattern appeared (Fig. 9c). The reason of the latter may be linked with the
MWCNT fragments induced reinforcing effect. Nevertheless, the smooth worn
surface suggests low wear, that was found in fact (cf. Figs. 8a, 9c). Interestingly,
the smooth surface appearance was not accompanied with low COF (cf. Fig. 8b).
ROP tests: On the worn surface of the pure HNBR chipped off debris, roll
formation and spherical particles are seen (Fig. 9a). This suggests high COFs
(cf. Fig. 8b) and high wear rates (cf. Fig. 8a), which were found, in fact. By
reinforcing the HNBR with 10 phr MWCNT the worn surface become smeared
(Fig. 10b) indicating for lower COF and specific wear rate (Ws) in accordance with
the experimental findings. With increasing MWCNT content (30 phr) the surface
of the wear track become ever smoother, and practically no wear debris or cracks
356 D. Felhös and J. Karger-Kocsis

Fig. 10 SEM pictures taken from the wear tracks after ROP tests. a pure HNBR,
b HNBR ? 10phr MWCNT, and c HNBR ? 30phr MWCNT. Note: sliding direction is
downwards [62]. (Reprinted from: Felhös, D. Karger-Kocsis, J. Xu, D.: Tribological testing of
peroxide cured HNBR with different MWCNT and silica contents under dry sliding and rolling
conditions against steel. Journal of Applied Polymer Science, 2008, 108, pp. 2840–2851,
Copyright 2010, with permission from John Wiley and Sons)

could be found (Fig. 10c). In that case the MWCNT worked probably as a solid
lubricant.
Recall that values of both the COF and specific wear rate strongly depend on
the testing rigs. This is in line with the usual claim that friction and wear are
‘‘system properties’’. Increasing MWCNT content generally enhances the
resistance to wear. This finding is similar what was reported on the effect of CB
fillers in rubbers. Between the mechanical and tribological properties of the
investigated materials no definite correlations can be found. The complexity of the
wear behavior requires a large data base in respect with the thermo-mechanical
and network-related properties of rubbers based on which their effects on wear
may be traced.
Xu and Karger-Kocsis published a well founded study on the tribological
properties of organophilic layered nanosilicate (OLS) filled HNBR [67]. As clay
modifier a quaternary amine with hydroxile and double bonds was used.
The properties of the OLS filled nanocomposites were compared to the pure
Friction and Wear of Rubber Nanocomposites 357

HNBR material. Dry sliding tests were performed applying different testing
configurations. It was found, that for all the executed dry sliding tests, the wear
resistance was negatively influenced by the intercalated OLS filler. On the other
hand, the coefficient of friction was slightly increased or dramatically decreased
depending on the testing configurations.
Gatos et al. [68] performed an extensive wear study on montmorillonite (MMT
with and without organophilic treatment) filled HNBR and EPDM rubber grafted
with maleic anhydride (EPDM–MA) elastomers. The MMT modifiers were the
following: octadecyl amine (ODA; primary amine), octadecyl trimethylammonium
(ODTMA) and methyl-tallow-bis(2-hydroxy-ethyl) quaternary ammonium salts
(MTH). It was found that for dry sliding, all the MMT additives resulted in an
increased coefficient of friction. The measured wear was higher for the HNBR/
MMT nanocomposites than for the neat HNBR rubber, except the unmodified
MMT containing HNBR. For the EPDM–MA with MMT–ODA and MMT–
ODTMA both the coefficient of friction and the measured specific wear were
reduced under POP condition.

2.2.2 Thermoplastic Rubbers

Two thermoplastic rubber compounds were investigated in respect to their dry


sliding friction and wear characteristics. One of them is a polyester-based linear
thermoplastic polyurethane (Elastollan C60A W; referred further as TPU), which
according to the datasheet contains 10 wt% plasticizer, has a nominal Shore A
hardness of 64, and the melting temperature of hard segments is at about 185C.
The second one was a polyolefin-based version, viz. TPO. These two thermoplastic
rubbers were filled with unmodified MWCNT (NC7000 of Nanocyl, Sambreville,
Belgium) and organoclay (OC, in this case: Nanofil 5, Süd-Chemie AG, München,
Germany). The surfactant of the latter was distearyl dimethyl ammonium chloride.
The OC containing TPU and TPO thermoplastic elastomers are denoted further as
TPU-2, -5, -10 and TPO-2, -5, -10, where the last number means the wt% of the
OC filler. The MWCNT containing TPUs and TPOs are designated as TPU-C05,
-C1, -C2 and TPO-C05, -C1, -C2, where C stays for CNT and the last number
denotes the wt% of the MWCNT filler.
The composites were prepared on an open mill at 145C (TPO) and at 185C
(TPU). 1 mm and 4 mm thick sheets were hot pressed at 150C (TPO) and at
190C (TPU). Standard ISO527-2 specimens were cut from the 1 mm thick sheets
for tensile tests. The Shore A hardness was tested on the 4 mm thick samples. The
mechanical properties of the thermoplastic elastomers and their nanocomposites
are summarized in Table 2. In Table 2a the mechanical properties of the TPO
based nanocomposites are listed. The Shore A hardness of the TPO increased with
incorporation of both the OC and MWCNT additives. The tensile strength of the
TPO nanocomposites decreased with increasing MWCNT and OC. The decrease
in the tensile strength is larger for MWCNT than for OC. Interestingly the change
in the strain to failure value is minor for the OC additive. In contrary to this,
358 D. Felhös and J. Karger-Kocsis

Table 2 Mechanical properties of the thermoplastic elastomers and their nanocomposites


Shore-A Tensile strength [MPa] Strain to failure [%] Density [g/cm3]
a
TPO-0 85 9.61 829 0.884
TPO-2 86 8.61 790 0.890
TPO-5 88 8.69 823 0.902
TPO-10 89 8.78 820 0.931
TPO-C05 87 7.89 702 0.884
TPO-C1 88 6.68 554 0.891
TPO-C2 89 8.35 777 0.893
b
TPU-0 74 29.85 1003 1.163
TPU-2 74 28.63 967 1.168
TPU-5 75 22.98 897 1.176
TPU-10 74 13.56 739 1.192
TPU-C05 75 21.57 817 1.164
TPU-C1 77 17.70 651 1.165
TPU-C2 79 22.83 864 1.171

the strain to failure diminishes markedly with the incorporation of MWCNT.


The density of the nanocomposites increases monotonously with increasing
additive content. In Table 2b the mechanical properties of the TPU-based
nanocomposites are collated. Note that the OC filler slight increases the Shore A
hardness. In contrast, the MWCNT filler is a more ‘‘active’’ one, which increases
the Shore A hardness with its increasing content. The tensile strength and the strain
to failure values both decrease with incorporation of OC and MWCNT. Decreasing
tendency in the strength and elongation properties is to observe with increasing OC
and MWCNT contents. The density of the nanocomposites increases monoto-
nously with increasing additive content.
Sliding friction tests were made on a POP-type tribometer (see Fig. 7)
(THT-0000 type from CSM, Peseux, Switzerland). For the sliding the 4 mm thick
samples were used. The TPU-based nanocomposites were tested with ceramic
(Al2O3) ball of 8.77 mm diameter, applying 15 N normal load (FN) and a sliding
speed of 25 cm/s with a distance of 400 m. The radius of the circular path was
4.99 mm. The TPO-based nanocomposites were tested using Cr6 bearing ball of
6 mm diameter applying 5 N load 10 cm/s and 200 m. The radius of the circular
path was 4.99 mm. The loss volume of the specimens after POP tests was esti-
mated based on the half with and on the depth of the wear grooves cross section,
which was considered to be a half ellipse. To get the loss volume the estimated
area of the half ellipse was multiplied with the circumference of the orbital rev-
olution of the pin (r = 4.99 mm).
Loss volume (wear) results after POP tests are shown for the TPO/OC and TPO/
MWCNT nanocomposites in Fig. 11. Though the results underlay a large scatter, it
can be established, that both OC and MWCNT additives improve the wear
resistance. For the TPO/OC and TPO/MWCNT nanocomposites the lowest wear
Friction and Wear of Rubber Nanocomposites 359

Fig. 11 Loss volume (wear)


results in POP tests for the
TPO-based nanocomposites
containing OC and MWCNT

was measured at 2 and 1 wt% additive, respectively. This suggests again that
enhanced sliding wear resistance is probably linked with an optimum amount of
nanofiller.
Considering the development of COF as a function of time (cf. Fig. 12) some
preliminary conclusions can be drawn on the wear process. At the very beginning
of the tests there is a large drop in the COF values. Recall that this is often the case
for POP tests. The increasing heat and the damage of the specimen surface along
with the appearance of wear debris are the main reasons for the COF decrease in
the initiation phase (t \ 100 s). In the following time period (steady state) the
COF changes marginally. The second drop in the COF versus time curve may
represent the total destruction of the material structure at the surface. At this stage
large sized cracks appear at the surface and the final wear groove pattern develops.
One exception is the TPO-C2 composite (see diagram (b) in Fig. 12), the COF of
which is steeply reduced shortly after the initial phase. After the total surface
destruction, the cracks propagate fast at the surface and the COF values tend to
level off at ca. 0.5. The length of the steady state correlates with the wear resis-
tance of the material, as the related time intervals are in accord with the loss
volume results. The longest steady state period (as well as the lowest wear) among
the TPO/OC and TPO/MWCNT nanocomposites were found at the 2 wt% OC-
and 1 wt% MWCNT-containing samples (TPO-2 and TPO-C1, respectively).
Similar observations were reported in Ref. [69] where the COF and wear rate of
the investigated thermoplastic elastomers were decreased with increasing hardness
of the material. In [30] this observation was limited for the POP configurations.
In Fig. 13 one can see the worn surface of the TPO based nanocomposites after
POP tests. These photographs display cracks across the wear path which are
opened in the sliding direction. It can be established that the dominant wear
mechanism does not change, but the density of the cracks increases with the
addition of nanofillers.
Based on the optical inspection the following wear mechanism occurs in POP
test (see Fig. 14). First, Schallamach type waves form at the surface due to the
high adhesion and low modulus of the elastomer (cf. Fig. 14a). At the wave fronts,
associated with large deformations, cracks initiate (cf. Fig. 14b). The initial cracks
360 D. Felhös and J. Karger-Kocsis

Fig. 12 Coefficient of friction (COF) during POP tests for the TPO-based nanocomposites
containing OC and MWCNT

Fig. 13 Light microscopic pictures taken of the worn surfaces of TPO/OC and TPO/MWCNT
nanocomposites after POP tests. a plain TPO, b TPO-5, and c TPO C2. Note: sliding direction is
downward
Friction and Wear of Rubber Nanocomposites 361

at the surface are plastically deformed owing to ‘‘over-sliding’’ of the rigid


counterbody (cf. Fig. 14c). During the repeated ‘‘over-sliding’’ the cracks open in
the sliding direction and the cracks undergo severe plastic deformation. This
process results in the formation ‘‘rubber tongues’’ which is typical for sliding wear
patterns in rubbers (cf. Fig. 14d).
In Fig. 15 the loss volume results for the TPU/OC and TPU/MWCNT
nanocomposites are collected. One can see that the incorporation of the nanofillers
reduced the wear resistance of the composites. The incorporation of MWCNT

Fig. 14 Wear mechanism of the TPO nanocomposites for POP test. a formation of Schallamach
waves, b initiation of cracks due to the Schallamach waviness, c deformation and growth of the
initial cracks, d formation of rubber tongues at the surface

Fig. 15 Loss volume (wear)


results after POP tests for
TPU/OC and TPU/MWCNT
nanocomposites
362 D. Felhös and J. Karger-Kocsis

seems to be less ‘‘harmful’’ than OC from the view point of the wear. Interestingly
the same OC and MWCNT amounts (2 and 1 wt%, respectively) yield the most
wear resistant composites similar to the TPO-based ones. On the other hand, all of
the nanofiller containing composites are less wear resistant than the neat TPU
matrix.
The course of COF as a function of sliding time is depicted in Fig. 16 for the TPU
based nanocomposites. Again, a large drop in the COF value can be observed in the
initial phase of the sliding friction tests. After this fast reduction the COF increases
with time (and thus sliding distance). The smallest COF during the tests in the steady
state was observed for the neat TPU matrix material. The highest steady state COF
values were observed for the TPU-5 (cf. red curve in diagram (a) in Fig. 16)

Fig. 16 COF as a function of sliding time during POP test for TPU/OC (a) and TPU/MWCNT
nanocomposites (b)
Friction and Wear of Rubber Nanocomposites 363

and for the TPU-C05 (cf. blue curve in diagram (b) in Fig. 16). Nanocomposites
with the highest COFs exhibited the largest loss volumes, as well. For the TPU based
composites, particularly for the MWCNT filled ones, the COF becomes rather stable
for a given period after the initial wear-in phase. The length of the stable period
correlates again with the wear resistance. The lowest wear among the MWCNT
filled nanocomposites was measured for the TPU-C1 and this sample possessed also
the longest time interval with stable COF. This may imply that the onset of surface
cracking occurred after more wear cycles in this material than in the others.
Figure 17 shows the light microscopic pictures taken of the worn tracks of the
TPU (cf. Fig. 17a, b) and TPU/OC (cf. Fig. 17c, d) and TPU/MWCNT
nanocomposites after POP tests (cf. Fig. 17e, f). Similar wear mechanisms can be
recognized on these figures for all of the composites. One of these wear formations
is the Schallamach-type waviness which is characteristic for elastomeric materials,
especially at low hardness (cf. Fig. 14). The other group of wear formations
belongs to the plastic deformations. These are namely the ‘‘ploughing’’ and
the extrusion- or ‘‘ironing’’-type deformations, which are characteristic for ther-
moplastic materials (cf. Fig. 18a, b). The plastic deformations are partly generated
by surface roughness peaks of the counterbody, partly due to the high adhesion
between counterbody and rubber substrate. Interestingly, multiplied Schallamach
waviness appears on the surface after the POP tests. This means, that between the
large sized Schallamach-type waves smaller ones are also present. Plastic
deformations are also present in the wear tracks, namely ploughing, extrusion and
‘‘ironing’’ (cf. Fig. 19). The Schallamach waviness and the clearly recognizable
plastic deformations suggest that the investigated elastomer has a dual (rubber- and
thermoplastic-like) feature at the same time.

Fig. 17 Light microscopic pictures taken of the worn surfaces of TPU and TPU/OC and TPU/
MWCNT nanocomposites after POP test. Designations: a, b plain TPU-0 matrix; c, d TPU-5, and
e, f TPU-C2
364 D. Felhös and J. Karger-Kocsis

Fig. 18 Plastic micro-deformations on the worn surface schematically. Designations: a plough-


ing and b ironing, extrusion

Fig. 19 Combined wear mechanisms (dual waviness and plastic deformation) in the TPU based
nanocomposites after POP test

2.3 Rolling-Type

Pioneering works in the rolling friction for rubbery materials have to be credited to
Tabor, Greenwood and coworkers in 1950s [70–76]. They all attributed the rolling
friction principally to mechanical hysteresis in the substrate when the latter is
deformed and the deformation is released by the passage of the rolling body [77].
When a spherical indenter is pressed into rubber a certain amount of elastic work is
performed. As the indenter moves forward elastic work is needed to deform the
rubber in front of the indenter whilst it is recovered from the rear. Assuming ideal
elastic behavior, the rubber behind the indenter would yield identical amount of
work and no net kinetic energy would be lost. But in fact, a constant fraction of the
input elastic energy is lost as a result of viscoelastic hysteresis in the rubber.
During rolling this is the primary source of the frictional work [71]. In addition,
the adhesion phenomenon is superposed to the hysteretic losses. In special cases
both effects can be comparable with each other. In Fig. 20 the adhesion
Friction and Wear of Rubber Nanocomposites 365

Fig. 20 The adhesion effect in case of a rolling cylinder on elastomeric plane [79] (a). Wear
mechanism (b) and wear formation (c) in case of rolling [80] tribotests due to repeated stresses
(fatigue)

phenomenon between a rolling cylinder and an elastomeric plane under unlubri-


cated conditions is depicted. The distance j between the rigid body and the
elastomeric plane represents the molecular roughness of the elastomer. One can
assume, that the polymer chains are fixed in the bulk at their one end
(via crosslinking) while their other is free. During sliding the free molecule ends
touch the surface of the rigid body and adhere to it. The detaching process needs
more energy than the adhering process. As a consequence, the molecules at the
detaching (peeling) site of the contact region will be extended causing additional
friction forces [78]. Based on this adhesion the same process causes friction even if
the rigid body in Fig. 20 is a surface asperity in micron- or in nano-range in sliding
movement [79].
Different wear mechanisms can be observed, when the wear of rubbery
materials is tested under rolling. In case of lubricated rolling contact, the typical
wear process is of fatigue type. During fatigue-like tests, if the surfaces are
lubricated, peculiar subsurface crack initiation and propagation may take place. It
is known, that in Hertzian-like contacts the maximum stresses and strains appear in
the bulk near to the surface. So, the probability of crack initiation is the highest
there, where the arising strains reach their maximum. Thus cracks arise below the
surface and the cracks propagate during the fatigue process. This crack propaga-
tion advances as long the cracks grow together and wear particles break out of the
surface (see Fig. 20b). This phenomenon is called pitting or gouging. In Fig. 20c
one can see the schematic picture of the gouging effect, the onset of which can be
observed in fretting or in rolling tests.
The rolling friction and wear properties of elastomeric materials can be
determined in different testing configurations, such as the orbital rolling ball-on-
plate (O-RBOP) test. For this test, however, one needs to clarify the kinematics of
the rolling ball, which qualitatively determinates the wear and friction during the
tests. In Fig. 21 one can see the schematic sketch of the O-RBOP test configuration
366 D. Felhös and J. Karger-Kocsis

Fig. 21 Schematic sketch of


the orbital rolling ball(steel)-
on-plate(rubber) test
configuration (O-RBOP)

Fig. 22 Superposition of the sliding velocities in the contact area during orbital rolling [62].
(Reprinted from: Felhös, D. Karger-Kocsis, J. Xu, D.: Tribological testing of peroxide cured
HNBR with different MWCNT and silica contents under dry sliding and rolling conditions
against steel. Journal of Applied Polymer Science, 2008, 108, pp. 2840–2851, Copyright 2010,
with permission from John Wiley and Sons)

wherein a steel ball, which is pressed with a constant normal load into the rubber
specimen, rolls along an orbital path on the rubber specimen (cf. Fig. 21).
The investigation of the kinematical process of the rolling tribotests is inevi-
table, due to the fact, that under dry rolling conditions, beneath the fatigue process,
additional sliding wear occurs via microslipping. This significant microslipping
(sliding of the non-adhering counter surfaces) can take place because of the
differences in the elastic moduli of the rolling pairs, and due to superposition of
different rolling phenomena. Accordingly, the dry rolling is always accompanied
with sliding wear. Note that depending on the surface quality of the counterbody
the sliding wear may turn into abrasive one, similar to the wear of car tires.
In Fig. 22 one can see the superposition of different sliding processes during
O-RBOP test in the wear track. The different sliding processes are originated from
the different components (drilling, forward rolling) of the orbital rolling method.
The detailed description of the rolling–sliding processes during the O-RBOP tests
is given in Ref. [62].

2.3.1 Conventional Rubbers

Different rubber blends based on HNBR and FKM with and without CB/MWCNT
fillers were tested with the O-RBOP tribotest configuration under dry rolling
Friction and Wear of Rubber Nanocomposites 367

conditions. The peroxide curable HNBR was mixed with 20 phr CB (N550), 10/
20/30 phr MWCNT (Baytubes C 150 P from Bayer MaterialScience, Leverku-
sen, Germany), respectively. The MWCNT filler in the composites will be des-
ignated with the letter ‘‘C’’. The samples prepared are referred to as HNBR,
HNBR-20CB, HNBR-10/20/30C, respectively. The digits in the designations
represent the contents of fillers in phr. HNBR with hybrid fillers (i.e. CB and
MWCNT) was also produced. The corresponding composition contained 20 phr
CB and 10 or 20 phr MWCNT, and denoted as HNBR-CB-C.
Being both peroxide-curable HNBR and FKM (Viton GF-600S of DuPont
Performance Elastomers Geneva, Switzerland) were combined in the following
HNBR/FKM ratios: 100/0, 100/33.3 and 100/100 (designations: HF-1:0, HF-1:1/3
and HF-1:1 respectively). To the above mixtures 10 phr MWCNT was incorpo-
rated on a laboratory mill. Note that the cure recipe was not adjusted, i.e. the
ready-to-cure HNBR was just diluted by FKM. This means that a direct com-
parison can only be made between rubber compounds with and without MWCNT
filler when their HNBR/FKM ratio is the same. The rubbers involved are further on
referred to as HNBR, HNBR-10C, HF-1:1/3, HF-1:1/3-10C, HF-1:1, and HF-1:
1-10C, respectively. Their network-related parameters, density and hardness
values are summarized in Table 3.
From the point of the physical properties it is an important question how the
CNTs are dispersed in the matrix. To investigate the distribution of the nanotubes,
transmission electron microscopic (TEM) pictures were taken from the prepared
HNBR/FKM/MWCNT composites. The TEM inspection showed that the
MWCNT filler is preferentially located in the continuous HNBR phase (see

Fig. 23 TEM picture of the


HNBR/FKM/MWCNT (HF-
1:1-10C) nanocomposite.
Note that MWCNT is located
only in the HNBR phase
368 D. Felhös and J. Karger-Kocsis

Fig. 23). Attention should be paid to the fact that the HNBR is the continuous
phase even in the HF-1: 1mixture.
In the home-made O-RBOP test configuration (cf. Fig. 21), the rubber sheet is
worn by a steel ball (100Cr6, diameter: 14 mm, arithmetical roughness Ra: 1 lm),
rolling along a circular path diameter: 33 mm), pushed by a defined normal load
against the rubber sheet. The normal load (FN) was 90 N, the revolution number is
280 rpm, and the test duration was 3 h. The specific wear rate was computed
according to Eq. 1. The loss volume (DV) (cf. Eq. 1) was calculated by measuring
the depth and width of the wear track and estimating it as a half ellipse. The cross
section of the wear track was assessed by a white light profilometer. The wear
degree can be assessed by specific wear rate (Ws):
 
DV mm3
Ws ¼ ð1Þ
FN  L N  m

where DV [mm3] is the volume loss, FN [N] is the normal load, L [m] is the overall
rolling distance.
Table 4 summarizes the COF and Ws of HNBR with and without fillers
measured in O-RBOP tests. The COF changes marginally with the increasing
content of MWCNT (cf. Table 4). Using MWCNT the wear resistance of the
related HNBR mixes was strongly enhanced, compared to that of the pure HNBR.
The results suggest that incorporation of CB, lowered the Ws and slightly enhanced
the COF. CB proved to be better in improving the rolling wear resistance of HNBR
compared to MWCNT. The Ws values of HNBR–FKM hybrids with and without
MWCNT measured in O-RBOP show a great improvement in the wear resistance.
Most prominent is the enhancement of the wear resistance for the HNBR/FKM
100-33.3 MWCNT hybrid (HF-1:1/3-10C; here the Ws is 30 times lower than that
is for the neat HNBR). Incorporation of MWCNT into HNBR/FKM decreased the
Ws and increased the COF. The COF values of the HNBR/FKM hybrids are higher
compared to the neat HNBR.
The SEM photos in Fig. 24 were taken from the wear track of the HNBR, after
O-RBOP tests. For the pure HNBR rubber Schallamach type pattern with roll head
can be found on the both outer and inner sides of the wear track (cf. Fig. 24).
The wave fronts in the two regions are adverse to each other. This reflects that the
direction of the ball movement in the two regions is opposite (cf. Fig. 22).
The waves are especially well-developed in the outer side, confirming a larger
speed in the outer compared to the inner region (cf. Fig. 22). In the middle section
large agglomerates can be observed (cf. Fig. 24). Note that the effect of the spin of
the ball is negligible in the centre region. The debris produced in the outer and
inner regions were likely swept into the centre region by the spin of the ball
favoring the formation of such agglomerates. Based on the SEM pictures one can
draw the following conclusions: for CB and MWCNT, the Schallamach-type
pattern completely disappeared and the debris fragmented surface in the centre
region of the wear track result in a lower Ws and higher COF compared to the pure
HNBR, respectively (see Ref. [62]).
Table 3 Network-related properties and hardness of the HNBR/FKM–CB/MWCNT nanocomposites and hybrids
HNBR HNBR-10C HNBR-20C HNBR-30C HNBR- HNBR- HF-1:1/3 HF-1:1/3- HF-1:1 HF-1:1-10C
20CB 20CB-20C 10C
Mc [g/mol] 2013 1176 407 161 1775 407 3088 1068 3715 1182
mc [mol/m3] 525 899 2710 7052 615 2710 371 1213 352 1119
Friction and Wear of Rubber Nanocomposites

tan d at Tg 1.37 1.11 0.68 0.50 1.25 0.68 1.16/0.31 0.36/0.56 0.56/0.64 0.35/0.58
Density [g/ 1.057 1.057 1.102 1.135 1.091 1.102 1.146 1.296 1.309 1.322
cm3]
Martens 1.02 ± 0.56 2.20 ± 0.24 3.37 ± 0.71 10.23 ± 0.84 0.66 ± 0.29 3.37 ± 0.71 1.19 ± 0.09 2.21 ± 1.08 1.12 ± 0.06 2.98 ± 0.75
hardness
[MPa]
369
370

Table 4 Wear and coefficient of friction (COF) results for the HNBR/FKM–CB/MWCNT nanocomposites and hybrids
HNBR HNBR- HNBR- HNBR- HNBR- HNBR- HF-1:1/3 HF-1:1/3-10C HF-1:1 HF-1:1-10C
10C 20C 30C 20CB 20CB-20C
Ws [10-4 mm3/ 2.95 ± 0.839 2.6 ± 0.739 1.05 ± 0.122 0.917± 0.613 ± 0.201 1.05 ± 0.122 0.283 ± 0.0512 0.066 ± 0.0279 0.788 ± 0.430 0.117 ± 0.0154
Nm]
COF[-] 0.042 ± 0.01 0.038 ± 0.00181 0.043 ± 0.0015 0.041± 0.0497 ± 0.0004 0.043 ± 0.0015 0.049 ± 0.0004 0.053 ± 0.0004 0.055 ± 0.0016 0.055 ± 0.0026
D. Felhös and J. Karger-Kocsis
Friction and Wear of Rubber Nanocomposites 371

Fig. 24 SEM photos taken from the rolling wear track after Orbital-RBOP test of the HNBR-
Pure. Picture to the left is the outer region, the picture in the middle is the centre region and the
picture to the right was taken from the inner region. Note: rolling direction is downward the
arrows indicate the local sliding direction [62]. (Reprinted from: Felhös, D. Karger-Kocsis, J. Xu,
D.: Tribological testing of peroxide cured HNBR with different MWCNT and silica contents
under dry sliding and rolling conditions against steel. Journal of Applied Polymer Science, 2008,
108, pp. 2840–2851, Copyright 2010, with permission from John Wiley and Sons)

It is noticeable that the wear loss of the filler reinforced HNBR system
correlates with the occurrence of the Schallamach-type pattern. The dramatically
prohibited waviness corresponds to the decreased Ws for CB and MWCNT
reinforcement. The introduction of active fillers (MWCNT and CB) is the reason
for the less developed or lacking waves, as often observed in rubbery systems with
active fillers (e.g. [64, 81, 82]). One may also observe that incorporation of 10 phr
MWCNT enhances only slightly the wear resistance of the HNBR. However, when
20 phr fillers were added, the specific wear rate was strongly reduced. Interest-
ingly, only very slight further decrease of Ws was found when the compounds
containing higher amounts of filler.
SEM photos taken from the rolling wear tracks of HF-1:1/3 with and without
MWCNT after Orbital-RBOP tests are shown in Fig. 25. In the outer region of
HF-1:1/3, Schallamach type pattern appears with some fragments (cf. Fig. 25a).
Clusters from debris (agglomerates) were formed in the centre and inner regions
(cf. Fig. 25c). Incorporation of MWCNT changes the basic wear mechanisms in
the outer region (cf. Fig. 25b). Fatigued-induced formation of holes becomes the
main wear mechanism instead of the Schallamach wavy pattern in HF-1:1/3. Hole
development and debris clustering were found in the centre region of HF-1:1/3
10C (cf. Fig. 25b). Fatigue-induced pittings and small ‘‘ironed’’ particles are seen
in the inner region of the HF-1:1/3 10C (cf. Fig. 25d).
The wear mechanisms for the HF-1:1 (MWCNT) did not change practically
when compared to HF-1:1/3 (MWCNT)––cf. Fig. 26. However, the inner region of
HF-1:1 is full with ‘‘ironed’’ rolls, debris (cf. Fig. 26c). By comparing the worn
surfaces of HNBR–FKM with and without MWCNT, one can notice that the
Schallamach wavy pattern disappears after MWCNT introduction. This change in
the wear mechanism reflects a reduction in the specific wear rate, which was found
also for other active fillers, e.g. CB. It is worth noting that the Schallamach-type
wear disappears with increasing reinforcement of a rubber independent of the type
of the active filler used [67, 83].
372 D. Felhös and J. Karger-Kocsis

Fig. 25 SEM photos taken from the rolling wear tracks of HF-1:1/3 without (a, c) and with (b,
d) MWCNT (HF-1:1/3-10C) after Orbital-RBOP tests. Designation: a outer region, b outer and
centre regions, c centre and inner regions, and d inner region. Note: rolling direction is downward
[84]. (Reprinted from: D. Xu, J. Karger-Kocsis, Z. Major and R. Thomann: Unlubricated rolling
wear of HNBR/FKM/MWCNT compounds against steel. Journal of Applied Polymer Science,
112, (2009), pp. 1461–1470, Copyright 2010, with permission from John Wiley and Sons)

2.3.2 Thermoplastic Rubbers

The TPU rubbers used were as follows: a polyetherdiol-based TPU (Estane 58300,
Shore A = 80) with and without 5 wt% organoclay (Nanomer I30P), and a
polyesterdiol-based TPU (Estane 58142, Shore D = 65) with and without 5 wt%
Nanomer I30P. The surfactant (organophilic modifier) of Nanomer I30P was
octadecyl amine. Nanocomposites were produced by incorporating the OC in the
TPUs via extrusion melt compounding. The four materials used are further on
referred to as TPU–L, TPU–L/OC, TPU–H and TPU–H/OC in the text. Note that L
and H stand for the low and high hardness TPU grades, respectively.
The wear topography in the outer region of TPU–L/OC 150 N sample is very
unique (cf. Fig. 27b). It is of Schallamach type with rolls, but some ‘‘waves’’ are
locally aligned parallel to the rolling direction. So, Fig. 27b shows the onset of a
two-dimensional wavy network, mesh elements of which are oriented along the
Friction and Wear of Rubber Nanocomposites 373

Fig. 26 SEM photos taken from the rolling wear tracks of HF-1:1 without (a, c) and with (b, d)
MWCNT (HF-1:1-10C) after Orbital-RBOP tests. Designation: (a, b) outer and centre regions, (c,
d) inner region. Note: rolling direction is downward [84]. (Reprinted from: D. Xu, J. Karger-Kocsis,
Z. Major and R. Thomann: Unlubricated rolling wear of HNBR/FKM/MWCNT compounds against
steel. Journal of Applied Polymer Science, 112, (2009), pp. 1461–1470, Copyright 2010, with
permission from John Wiley and Sons)

rolling direction. This peculiar Schallamach-type pattern may probably explain the
lower specific wear rate of this OC loaded sample compared to that of the neat
TPU–L (cf. Fig. 27a and b; Table 5). SEM pictures in Fig. 27a support why higher
specific wear rate and COF were found for the plane TPU–L sample. Schallamach
pattern with rolls (‘‘tongues’’) are well discernible in the outer and inner regions.

Table 5 Specific wear results for the TPU–L/OC and TPU–H/OC nanocomposites after
O-RBOP tests
Normal load/test duration Specific wear rate (mm3/N m)
TPU–L TPU–L/OC TPU–H TPU–H/OC
60 N/3 h 1.01E–04 1.45E–05 6.26E–06 5.89E–06
90 N/3 h 1.12E–04 3.94E–05 1.39E–05 4.71E–06
120 N/3 h 1.20E–04 6.53E–05 1.67E–05 1.49E–05
150 N/3 h 1.69E–04 1.06E–04 2.88E–05 2.66E–05
374 D. Felhös and J. Karger-Kocsis

Fig. 27 SEM pictures taken from the rolling wear track of TPU–L (a) and TPU–L/OC (b) after
O-RBOP test performed at 150 N normal load for 3 h [85]. (Reprinted from: D. Xu, J. Karger-
Kocsis and A. K. Schlarb: Rolling friction and wear of organoclay-modified thermoplastic
polyurethane rubbers against steel. Kautschuk, Gummi, Kunststoffe, 61 (2008), pp. 98–106,
Copyright 2010, with permission from Hüthig GmbH)

In the first approximation, one can state that the better developed the Schallamach-
waves are, the higher the COF (cf. Table 6) and specific wear rate are
(cf. Table 5). In the outer region of TPU–H an embryonic wavy pattern appears
after dry rolling test (cf. Fig. 28a). This is believed to be the primary stage of the
Schallamach pattern formation. No track edges and wavy pattern could be resolved
in the worn surface of TPU-/OC (cf. Fig. 28b). On the other hand, the wear track
becomes different from the unfilled version when OC is incorporated.
The Schallamach pattern is the more pronounced the lower the hardness of the
related rubbers is. Note that this claim holds also for traditional rubbers ([64] and
references therein). The OC modification improved the resistance to dry rolling
wear (cf. Table 5). The improvement was the higher the lower the hardness of the
parent TPU was. The COF was slightly reduced by OC incorporation when tested
under the same conditions for softer TPUs. However, the COF was increased in
presence of OC for the harder TPUs (cf. Table 6). It is noteworthy that no such
clear tendencies could be concluded for the dry sliding wear of OC-modified
rubbers [68]. The COF was hardly affected by the test duration. By contrast the
specific wear rate was reduced with the time. This can be explained by a reduction

Table 6 Coefficient of friction (COF) results for the TPU–L/OC and TPU–H/OC
nanocomposites
Normal load/test duration COF (-)
TPU–L TPU–L/OC TPU–H TPU–H/OC
60 N/3 h 0.0366 0.0288 0.0172 0.0193
90 N/3 h 0.0347 0.0315 0.0240 0.0265
120 N/3 h 0.0375 0.0352 0.0259 0.0306
150 N/3 h 0.0438 0.0383 0.0341 0.0371
Friction and Wear of Rubber Nanocomposites 375

Fig. 28 SEM pictures taken from the rolling wear track of TPU–H (a) and TPU–H/OC (b) after
O-RBOP test performed at 150 N normal load for 3 h [85]. (Reprinted from: D. Xu, J. Karger-
Kocsis and A. K. Schlarb: Rolling friction and wear of organoclay-modified thermoplastic
polyurethane rubbers against steel. Kautschuk, Gummi, Kunststoffe, 61 (2008), pp. 98–106,
Copyright 2010, with permission from Hüthig GmbH)

in the real contact surface owing to debris formation. Both COF and specific wear
rate increased with increasing normal load (cf. Tables 5, 6). Recall that the wear
mechanisms in Orbital-RBOP tests are complex due to the superimposed rolling
and spinning.

3 Outlook and Future Trends

Novel nanofillers, such as organophilic-modified clays (OC) and carbon nanotubes


(CNT) are promising additives to improve the tribological performance, targeting
both reduced coefficient of friction and wear rate, of rubbers. Further work is
needed, however, to realize this potential. The results reported above suggest that
OC and CNT are less suited for rubber/rubber and rubber/thermoplastic blends
because of their selective embedding into one component of the compositions.
CNT, when not properly functionalized and thus dispersed, may be far less
effective than expected. In this respect attention should be paid to the fact that the
quantitative assessment of the nanofillers’ dispersion is not solved. Hybrid rein-
forcing of rubbers, i.e. the combined use of traditional and novel fillers, may be the
right research strategy for the next future. Less information is available on how
these nanofillers affect the lubricated wear of rubbers. Very interesting is the
tribological (both dry and lubricated) performance of nanofilled rubbers under
high-pressure conditions. According to the authors’ opinion both OC and CNT
may be suitable additives to improve the resistance to rapid gas decompression of
rubbers. Further impetus to the friction and wear of rubbers is expected from
modeling studies considering the wear types of rubbers accordingly. Though
considerable efforts are devoted to this issue, an in depth understanding is still
376 D. Felhös and J. Karger-Kocsis

lacking. Note that model predictions would be very helpful for further material
(recipe) development and ‘‘refining’’ of tribological tests.

Acknowledgments Dr. Dávid Felhös is very thankful to Mr. György Szabó and his family
their selfless and friendly advocacy to take a fresh start in Miskolc. The authors express their
thanks to Dr. Dan Xu for the performed tests and for the results presented in Sects. 2.3.1 and
2.3.2 and to Dr. Kálmán Marossy for the preparation of the TPU and TPO based
nanocomposites.

References

1. Kim, Y.A., Hayashi, T., Endo, M., Gotoh, Y., Wada, N., Seiyama, J.: Fabrication of aligned
carbon nanotube-filled rubber composite. Scripta Mater. 54, 31–35 (2006)
2. Vast, L., Philippin, G., Destrée, A., Moreau, N., Fonseca, A., Nagy, J.B., Delhalle, J.,
Mekhalif, Z.: Chemical functionalization by a fluorinated trichlorosilane of multi-walled
carbon nanotubes. Nanotechnology 15, 781–785 (2004)
3. Ma, P.C., Kim, J.-K., Tang, B.Z.: Functionalization of carbon nanotubes using a silane
coupling agent. Carbon 44, 3232–3238 (2006)
4. Kokaia, F., Koshioa, A., Shiraishia, M., Matsutaa, T., Shimodaa, S., Ishiharab, M., Kogab,
Y., Denoc, H.: Modification of carbon nanotubes by laser ablation. Diam. Relat. Mater. 14,
724–728 (2005)
5. Breton, Y., Delpeux, S., Benoit, R., Salvetat, J.P., Sinturel, C., Beguin, F., Bonnamy, S.:
Functionalization of multiwall carbon nanotubes: properties of nanotubes-epoxy composites.
Mol. Cryst. Liq. Cryst. 387, 135–140 (2002)
6. Hill, D.E., Lin, Y., Rao, A.M., Allard, L.F., Sun, Y.-P.: Functionalization of carbon
nanotubes with polystyrene. Macromolecules 35, 9466–9471 (2002)
7. Sui, G., Zhong, W.H., Yang, X.P., Yu, Y.H.: Curing kinetics and mechanical behavior of
natural rubber reinforced with pretreated carbon nanotubes. Mater. Sci. Eng. A-Struct. 485,
524–531 (2008)
8. Sui, G., Zhong, W.H., Yang, X.P., Yu, Y.H., Zhao, S.H.: Preparation and properties of natural
rubber composites reinforced with pretreated carbon nanotubes. Polym. Adv. Technol. 19,
1543–1549 (2008)
9. Falco, A.D., Goyanes, S., Rubiolo, G.H., Mondragon, I., Marzocca, A.: Carbon nanotubes as
reinforcement of styrene–butadiene rubber. Appl. Surf. Sci. 254, 262–265 (2007)
10. Das, A., Stöckelhuber, K.W., Jurk, R., Fritzsche, J., Klüppel, M., Heinrich, G.: Coupling
activity of ionic liquids between diene elastomers and multi-walled carbon nanotubes.
Carbon 47, 3313–3321 (2009)
11. Atieh, M.A., Girun, N., Ahmadun, F.-R., Guan, C.T., Mahdi, E.-S., Baik, D.R.: Multi-wall
carbon nanotubes/natural rubber nanocomposite. J. Nanotechnol. Online (2005). doi:
10.2240/azonjo0106
12. Luo, Y., Wang, C., Li, Z.: Preparation, fabrication and response behavior of HTBN/TDI/
MWCNT composite sensing film by in situ dispersed polymerization. Synth. Met. 157, 390–
400 (2007)
13. Bokobza, L., Kolodziej, M.: On the use of carbon nanotubes as reinforcing fillers for
elastomeric materials. Polym. Int. 55, 1090–1098 (2006)
14. Bokobza, L., Belin, C.: Effect of strain on the properties of a styrene–butadiene rubber filled
with multiwall carbon nanotubes. J. Appl. Polym. Sci. 105, 2054–2061 (2007)
15. Faulkner, R.W., Mumby, K.J., Fisherm A., Jozokos, T., Zhou, S.: Multiwall carbon nanotube
reinforcement of HNBR and FKM. http://rethink-technologies.com/static/Hyperion_Fibrils_
in_FKM_and_HNBR_ACS_Presentation-KM_001e.pdf. (2009)
Friction and Wear of Rubber Nanocomposites 377

16. Fritzsche, J., Lorenz, H., Klüppel, M.: CNT based elastomer-hybrid-nanocomposites with
promising mechanical and electrical properties. Macromol. Mater. Eng. 294, 551–560 (2009)
17. Bokobza, L., Rhamani, M., Belin, C., Bruneel, J.-C., Bounia, N.-E.E.: Blends of carbon
blacks and multiwall carbon nanotubes as reinforcing fillers for hydrocarbon rubbers.
J. Polym. Sci. Pol. Phys. 46, 1939–1951 (2008)
18. Zhu, J., Kim, J.D., Peng, H., Margrave, J.L., Khabashesku, V.N., Barrera, E.V.: Improving
the dispersion and integration of single-walled carbon nanotubes in epoxy composites
through functionalization. Nano. Lett. 3, 1107–1113 (2003)
19. Jung, Y.C., Sahoo, N.G., Cho, J.W.: Polymeric nanocomposites of polyurethane block
copolymers and functionalized multi-walled carbon nanotubes as crosslinkers. Macromol.
Rapid Comm. 27, 126–131 (2006)
20. Xia, H., Song, M.: Preparation and characterisation of polyurethane grafted single-walled
carbon nanotubes and derived polyurethane nanocomposites. J. Mater. Chem. 16, 1843–1851
(2006)
21. Coleman, J.N., Cadek, M., Blake, R., Nicolsi, V., Ryan, K.P., Belton, C., Fonseca, A., Nagy,
J.B., Gun’ko, Y.K., Blau, W.J.: High-performance nanotube-reinforced plastics:
understanding the mechanism of strength increase. Adv. Funct. Mater. 14, 791–798 (2004)
22. Claes, M., Duoin, G., Luizi, F.: Latest developments in carbon nanotubes based
nanocomposites. Rubber World 239, 28–34 (2009)
23. Sato, Y., Hasegawa, K., Nodasaka, Y., Motomiya, K., Namura, M., Ito, N., Jeyadevan, B.,
Tohji, K.: Reinforcement of rubber using single-walled nanotube soot and its shock
dampening properties. Carbon 46, 1506–1517 (2008)
24. Bokobza, L.: Multiwall carbon nanotube elastomeric composites: A review. Polymer 48,
4907–4920 (2007)
25. Bhattacharyya, S., Sinturel, C., Bahloul, O., Saboungi, M.-L., Thomas, S., Salvetat, J.-P.:
Improving reinforcement of natural rubber by networking of activated carbon nanotubes.
Carbon 46, 1037–1045 (2008)
26. López-Manchado, M.A., Biagiotti, J., Valentini, L., Kenny, J.M.: Dynamic mechanical and
Raman spectroscopy studies on interaction between single-walled carbon nanotubes and
natural rubber. J. Appl. Polym. Sci. 92, 3394–3400 (2004)
27. Lu, L., Zhou, Z., Zhang, Y., Wang, S., Zhang, Y.: Reinforcement of styrene–butadiene-
styrene tri-block copolymer by multi-walled carbon nanotubes via melt mixing. Carbon 45,
2621–2627 (2007)
28. Lu, L., Zhai, Y., Zhang, Y., Ong, C., Guo, S.: Reinforcement of hydrogenated nitrile-
butadiene rubber by multi-walled carbon nanotubes. Appl. Surf. Sci. 255, 2162–2166 (2008)
29. Valentini, L., Biagiotti, J., Kenny, J.M., Manchado, M.A.L.: Physical and mechanical
behavior of single-walled carbon nanotube/polypropylene/ethylene-propylene-diene rubber
nanocomposites. J. Appl. Polym. Sci. 89, 2657–2663 (2003)
30. Karger-Kocsis, J., Felhös, D., Thomann, R.: Tribological behavior of a carbon-nanofiber-
modified Santoprene thermoplastic elastomer under dry sliding and fretting conditions
against steel. J. Appl. Polym. Sci. 108, 724–730 (2008)
31. Perez, L.D., Zuluaga, M.A., Kyu, T., Mark, J.E., Lopez, B.L.: Preparation, characterization,
and physical properties of multiwall carbon nanotube/elastomers composites. Polym. Eng.
Sci. 49, 866–874 (2009)
32. Byres, J.T.: Fillers for balancing passenger tire tread properties. Rubber Chem. Technol. 75,
527–547 (2002)
33. Paglicawan, M.A., Kim, J.K., Bang, D.-S.: Dispersion of multiwalled carbon nanotubes in
thermoplastic elastomer gels: morphological, rheological, and electrical properties. Polym.
Compos. 31, 210–217 (2009)
34. Meincke, O., Kaempfer, D., Weickmann, H., Friedrich, C., Vathauer, M., Warth, H.:
Mechanical properties and electrical conductivity of carbon-nanotube filled polyamide-6 and
its blends with acrylonitrile/butadiene/styrene. Polymer 45, 739–748 (2007)
35. Li, Q., Xue, Q.Z., Gao, X.L., Zheng, Q.B.: Temperature dependence of the electrical
properties of the carbon nanotube/polymer composites. Exp. Polym. Lett. 3, 769–777 (2009)
378 D. Felhös and J. Karger-Kocsis

36. Pötschke, P., Dudkin, S.M., Alig, I.: Dielectric spectroscopy on melt processed polycarbonate
multiwalled carbon nanotube composites. Polymer 44, 5023–5030 (2003)
37. Elsworth, M.W.: Organoclay-polymer composites. US Patent, 5962553, 1999
38. Chattopadhyay, D.K., Mishra, A.K., Sreedhar, B., Raju, K.V.S.N.: Thermal and viscoelastic
properties of polyurethane-imide/clay hybrid coatings. Polym. Deg. Stab. 91, 1837–1849 (2006)
39. Yusoh, K., Jin, J., Song, M.: Subsurface mechanical properties of polyurethane/organoclay
nanocomposite thin films studied by nanoindentation. Prog. Org. Coat. 67, 220–224 (2010)
40. Choi, W.J., Kim, S.H., Kim, Y.J., Kim, S.C.: Synthesis of a chain-extended organifier and
properties of polyurethane/clay nanocomposites. Polymer 45, 6045–6057 (2004)
41. Gatos, K.G., Karger-Kocsis, J.: Effects of primary and quarternary intercalants on the
organoclay dispersion in a sulfur-cured EPDM rubber. Polymer 46, 3069–3076 (2005)
42. Wang, S., Zhang, Y., Ren, W., Zhang, Y., Lin, H.: Morphology, mechanical and optical
properties of transparent BR/clay nanocomposites. Polym. Test. 24, 766–774 (2005)
43. Pal, K., Rajasekar, R., Kang, D.J., Zhang, Z.X., Kim, J.K., Das, C.K.: Effect of epoxidized
natural rubber-organoclay nanocomposites on NR/high styrene rubber blends with fillers.
Mater. Design 30, 4035–4042 (2009)
44. Sharif, J., Yunus, W.M.Z.W., Dahlan, K.Z.H.M., Ahmad, M.H.: Preparation and properties of
radiation crosslinked natural rubber/clay nanocomposites. Polym. Test. 24, 211–217 (2005)
45. Carretero-González, J., Valentín, J.-L., Arroyo, M., Saalwächter, K., Lopez-Manchado,
M.A.: Natural rubber/clay nanocomposites: Influence of poly(ethylene glycol) on the silicate
dispersion and local chain order network. Eur. Polym. J. 44, 3493–3500 (2008)
46. Yang, I.-K., Tsai, P.-H.: Intercalation and viscoelasticity of poly(ether-block-amide)
copolymer/montmorillonite nanocomposites: Effect of surfactant. Polymer 47, 5131–5140
(2006)
47. Wang, X.-P., Huang, A.-M., Jia, D.-M., Li, Y.-M.: From exfoliation to intercalation-changes
in morphology of HNBR/organoclay nanocomposites. Eur. Polym. J. 44, 2784–2789 (2008)
48. Lim, S.-H., Dasari, A., Yu, Z.-Z., Mai, Y.-W., Liu, S., Yong, M.S.: Fracture toughness of
nylon 6/organoclay/elastomers nanocomposites. Compos. Sci. Technol. 67, 2914–2923
(2007)
49. Sun, T., Dong, X., Du, K., Wang, K., Fu, Q., Han, C.C.: Structural and thermal stabilization of
isotactic polypropylene/organoclay/montmorillonite/poly(ethylene-co-octene) nanocomposites
by an elastomer component. Polymer 49, 588–598 (2008)
50. Chiu, F.-C., Lai, S.-M., Chen, Y.-L., Lee, T.-H.: Investigation of the polyamide 6/organoclay
nanocomposites with or without a maleated polyolefin elastomers as a toughener. Polymer
46, 11600–11609 (2005)
51. Kim, B.K., Seo, J.W., Jeong, H.M.: Morphology and properties of waterborne polyurethane/
clay nanocomposites. Eur. Polym. J. 39, 85–91 (2003)
52. Zhu, L., Wool, R.P.: Nanoclay reinforced bio-based elastomers: synthesis and
characterization. Polymer 47, 8106–8115 (2006)
53. Alexandre, M., Dubois, P.: Polymer-layered silicate nanocomposites: preparation, properties
and uses of a new class of materials. Mater. Sci. Eng. 28, 1–63 (2000)
54. Beyer, G.: Flame retardancy of nanocomposites based on organoclays and carbon nanotubes
with aluminium trihydrate. Polym. Adv. Technol. 17, 218–225 (2006)
55. Frounchia, M., Dadbinb, S., Salehpoura, Z., Noferestia, M.: Gas barrier properties of PP/
EPDM blend nanocomposites. J. Memb. Sci. 282, 142–148 (2006)
56. Thompson, M.R., Yeung, K.K.: Recyclability of a layered silicate-thermoplastic olefin
elastomers nanocomposite. Polym. Deg. Stab. 91, 2396–2407 (2006)
57. Chung, J.W., Han, S.J., Kwak, S.-Y.: Application of strain-time correspondence as a tool for
structural analysis of acrylonitrile–butadiene copolymer nanocomposites with various
organoclay loadings. Eur. Polym. J. 45, 79–87 (2009)
58. Ramorino, G., Bignotti, F., Pandini, S., Riccó, T.: Mechanical reinforcement in natural
rubber/organoclay nanocomposites. Compos. Sci. Technol. 69, 1206–1211 (2009)
59. TABER Rotary Platform Abraser: http://www.taberindustries.com/Products/abraser/index.
asp?ct=1&sc=1
Friction and Wear of Rubber Nanocomposites 379

60. Jacobs, O., Jaskulkaa, R., Yangb, F., Wub, W.: Sliding wear of epoxy compounds against
different counterparts under dry and aqueous conditions. Wear 256, 9–15 (2004)
61. Jacobs, O., Xu, W., Schädel, B., Wu, W.: Wear behaviour of carbon nanotube reinforced
epoxy resin composites. Tribol. Lett. 23, 65–75 (2006)
62. Felhös, D., Karger-Kocsis, J., Xu, D.: Tribological testing of peroxide cured HNBR with
different MWCNT and silica contents under dry sliding and rolling conditions against steel.
J. Appl. Polym. Sci. 108, 2840–2851 (2008)
63. El-Tayeb, N.S.M., RMd, Nasir: Effect of soft carbon black on tribology of deproteinised and
polyisoprene rubbers. Wear 262, 350–361 (2007)
64. Karger-Kocsis, J., Mousa, A., Major, Z., Békési, N.: Dry friction and sliding wear of EPDM
rubbers against steel as a function of carbon black content. Wear 264, 357–365 (2008)
65. Schallamach, A.: Friction and abrasion of rubber. Wear 1, 384–417 (1958)
66. Schallamach, A.: How does rubber slide? Wear 17, 301–3012 (1971)
67. Xu, D., Karger-Kocsis, J.: Dry rolling and sliding friction and wear of organophilic layered
silicate/hydrogenated nitrile rubber nanocomposite. J. Mater. Sci. 45, 1293–1298 (2010)
68. Gatos, K.G., Kameo, K., Karger-Kocsis, J.: On the friction and sliding wear of rubber/layered
silicate nanocomposites. Exp. Polym. Lett. 1, 27–31 (2007)
69. Karger-Kocsis, J., Felhös, D., Xu, D., Schlarb, A.K.: Unlubricted sliding and rolling wear of
thermoplastic dynamic vulcanizates (Santoprene) against steel. Wear 265, 292–300 (2008)
70. Greenwood, J.A., Tabor, D.: Deformation properties of friction junctions. Proc. Phys. Soc.
Lond. B 68, 609–619 (1955)
71. Greenwood, J.A., Tabor, D.: The friction of hard sliders on lubricated rubber: the importance
of deformation losses. Proc. Phys. Soc. 71, 989–1001 (1958)
72. Greenwood, J.A., Minshall, H., Tabor, D.: Hysteresis losses in rolling and sliding friction.
Proc. Phys. Soc. Lon. A Math. Phys. Sci 259, 480–507 (1961)
73. Tabor, D.: The mechanism of rolling friction. 2. The elastic range. Proc. Phys. Soc. Lon.
A Math. Phys. Sci 229, 198–220 (1955)
74. Evans, I.: The rolling resistance of a wheel with a solid rubber tyre. Br. J. Appl. Phys. 5, 187–
188 (1954)
75. May, W.D., Morris, E.L., Atack, D.: Rolling friction of a hard cylinder over a viscoelastic
material. J. Appl. Phys. 30, 1713–1724 (1959)
76. Flom, D.G., Bueche, A.M.: Theory of rolling friction for spheres. J. Appl. Phys. 30, 1725–
1730 (1959)
77. Gent, A.N., Henry, R.L.: Rolling friction on viscoelastic substrates. T. Soc. Rheol. 13, 255–
271 (1969)
78. Zaghzi, N., Carre, A., Shanahan, M.E.R., Papirer, E., Schultz, J.: A study of spontaneous
rubber/metal adhesion. I. The rolling cylinder test. J. Polym. Sci. B Polym. Phys. 25, 2393–
2402 (1987)
79. Chernyak, Y.B., Leonov, A.I.: On the theory of the adhesive friction of elastomers. Wear
108, 105–138 (1986)
80. Glaeser, W.A.: Wear debris classification. In: Bhushan, B. (ed.) Modern Tribology
Handbook, vol. 1. CRC Press, Boca Raton (2001)
81. Xu, D., Karger-Kocsis, J., Schlarb, A.K.: Rolling wear of EPDM and SBR rubbers as a
function of carbon black contents: correlation with microhardness. J. Mater. Sci. 43, 4330–
4339 (2008)
82. Felhös, D., Karger-Kocsis, J.: Tribological testing of peroxide-cured EPDM rubbers with
different carbon black contents under dry sliding conditions against steel. Tribol. Int. 41, 404–
415 (2008)
83. Xu, D., Karger-Kocsis, J., Schlarb, A.K.: Friction and wear of HNBR with different fillers
under dry sliding and rolling conditions. Exp. Polym. Lett. 3, 126–136 (2009)
84. Xu, D., Karger-Kocsis, J., Major, Z., Thomann, R.: Unlubricated rolling wear of HNBR/
FKM/MWCNT compounds against steel. J. Appl. Polym. Sci. 112, 1461–1470 (2009)
85. Xu, D., Karger-Kocsis, J., Schlarb, A.K.: Rolling friction and wear of organoclay-modified
thermoplastic polyurethane rubbers against steel. Kaut. Gummi. Kunstst. 61, 98–106 (2008)
Index

A carbon black, 5–8, 13, 29, 38, 39, 50, 52,


abrasion, 4–6, 13, 58, 180, 201–203, 54, 55, 58, 69, 83, 90, 91,
205–208, 213, 218, 226–231, 113–116, 120, 136, 139, 140,
320, 322, 329, 343, 347–349, 379 141, 153, 179, 180–184, 189,
abrasion resistance, 4, 58, 180, 202, 205–208, 190, 192, 193, 195, 198, 201,
213, 218, 224, 226, 322, 324, 329, 203, 206–208, 210, 214–217,
347–349 220–223, 226–228, 260, 293,
actuators, 281, 283, 287, 293, 304, 305 294, 302, 304, 344, 379
additive, 58, 61, 106, 113, 155, 157, 161, Carbon nanofibres, 158, 163, 317
162, 164, 165, 167–175, 186, Carbon nanotubes, 5, 7, 8, 12–17, 23, 50,
203, 208, 209, 215, 223, 270, 53, 54, 90, 96, 97, 115, 117,
278, 294, 310, 354, 357–359, 375 120, 130, 131, 136, 158, 163,
aggregation, 23, 57, 58, 65–67, 70, 71, 73, 164, 172, 175–177, 184, 187,
81, 97, 142, 220, 293–296, 304 188, 193–195, 198, 210, 267,
alkylammonium, 180 271, 281, 310, 317, 332, 340,
aspect ratio, 6–10, 12, 14, 29, 48, 52, 53, 89, 343, 344, 375–378
99, 120, 140, 144, 145, 151, 220, catalytic effect, 158, 167, 171
234, 242, 243, 247, 253–256, 260, chain movement, 135
262, 264, 309, 322, 327, 328, 332, chemical architecture, 238, 252
344, 346 chemical modifications, 235
cloisite, 166, 201, 215, 238, 242, 271, 274,
311, 349
B coefficient of friction, 205, 267, 343, 350,
barrier effect, 49, 158, 162, 163, 170, 173 351, 354, 357, 360, 370,
barrier function, 180 374, 375
barrier performance, 23, 234, 236, 248 compatibilizer, 104, 117, 166–168, 180,
basal plane spacing, 236, 240–242, 247, 248 185–187, 196, 213
blending, 11, 96, 103, 104, 116, 156, 163, compression molding, 217, 218, 333
208, 209, 213, 227, 228, 331, Coulombic interaction, 121, 124
332, 344 crack growth resistance, 180
broadband dielectric spectroscopy, 89, 94, cross-linked, 32, 61–65, 71, 75, 76,
114, 116 333, 341
crystallisation kinetics, 310
curing, 49, 70, 72, 77, 79, 91, 115,
C 140, 201, 248, 315, 329,
capacitance, 33–35, 113, 285, 286, 290, 300 332, 333, 335, 345,
carbohydrates, 136 348, 376

381
382 Index

D electrospun, 212, 311, 312, 328, 339


damping treatment, 308 energy harvesters, 281, 303
degradation, 13, 48–50, 54, 65, 156–159, ENR, 49, 70, 71, 74, 76, 90, 103–105, 176,
161–163, 165, 170, 172, 177, 201, 213–215, 217, 219, 220, 222,
179, 180, 185–188, 190, 191, 224, 226, 346, 349, 379
195–197, 206, 217, 221, 223, entropy modulus, 135
226, 233, 247, 250, 251, 255, epoxy, 16, 52–54, 70, 74, 78, 79, 84, 114,
265–268, 271, 274, 275, 277, 176, 177, 233, 234, 239–252,
278, 280, 317, 318, 329, 330, 254–256, 276, 277, 280, 325,
338, 339, 342 310, 341, 376–379
degree of bonding, 92 equilibrium state, 37, 90
dielectric, 17, 31, 33–35, 54, 67, 89, 94, 95, 99, exfoliation, 10, 12, 23, 110, 131, 165, 236,
101, 102, 106, 108–114, 116–118, 241–243, 247, 248, 253, 255, 256,
190, 281, 284–288, 290, 291, 262, 276, 313, 320, 321, 344, 376
303–305, 377 extrusion, 120, 309, 313, 363, 364, 372
dielectric elastomers, 284, 286, 287, 291, 305
dielectric loss peaks, 95
DIN abrasion, 218, 224, 225, 248 F
disordered solids, 91 fibres, 82, 158, 163, 197, 212, 260, 307–314,
dispersion, 3, 5, 9–19, 21, 23, 26, 48, 52–54, 317, 325, 327, 331, 333, 336,
58, 61, 68–71, 73, 75, 77, 78, 92, 337, 340
98, 102, 103, 117, 118, 130, 136, filler shape factor, 58
142, 146, 155, 156, 158, 159, finite element approach, 254
162, 165, 167–171, 174–177, fire hazards, 156
180, 189, 193, 195, 211–213, fire regulations, 155
217, 218, 222, 226, 240, 256, flame inhibition, 157
259–265, 270, 271, 294, 295, flame retardancy, 155–157, 159, 161, 162, 164,
322, 328, 330, 344, 375, 376 169, 170, 172–178, 180, 189, 328,
dispersion state, 155, 159, 165, 168, 174, 347, 378
175, 211 flame retardant, 155–163, 165, 167–178
Du-Pont abrasion, 218, 224 flammability, 155, 156, 159–162, 170, 171,
durability, 4, 185, 308, 339 174–177, 190, 197, 256, 317
dynamic mechanical, 5, 17, 31, 37, 38, 40, 43, fractography, 29
51, 54, 55, 68, 71–74, 77, 80, 89, 90, fracture properties, 308, 328
93, 95, 96, 102, 104, 112, 114, 115, frequency, 32, 33, 36, 44–46, 54, 93–96,
255, 267, 274, 278, 315, 332, 345, 98–101, 103, 106–111, 172,
353, 377 290, 291
dynamic mechanical analysis, 31, 68, 71, 73,
77, 80, 89, 90, 95, 102, 104, 112,
114, 115, 315, 332 G
gas permeability, 136, 255, 276, 330
green rubber, 59, 60
E grinding, 4, 11, 14
ecological tendencies, 282
elastomeric nanocomposites as biomaterials,
264 H
elastomeric nanocomposites for biomedical harvesting, 281, 283, 284, 286–288, 291, 292,
applications, 259 301–305
electrical behaviour, 89, 109, 110, 131 hybrid, 14, 43, 44, 51, 52, 57, 59, 62, 71, 74,
electrical circuitry, 286 77–85, 103, 116, 117, 131, 189,
electrical devices, 155 190, 207, 234, 276, 279, 318,
electrical field, 89, 282, 287–289, 291, 365–368, 373, 374, 376
296–304 hydrolytic sol-gel process, 59, 81
electro active polymers, 281, 282 hydrophilicity, 77, 243, 272
Index 383

I molecular dynamic effects, 89


Impedance, 3, 32–34, 54, 94, 271, 290, 302 montmorillonite, 9, 18, 51, 52, 90, 98,
in situ, 10, 55, 57, 58, 60–66, 68–85, 103, 116, 115–117, 120, 136, 151, 157,
131, 138, 154, 162, 188, 190, 213, 162–165, 167, 172–178, 180, 185,
261, 267, 272, 320, 330, 331, 342, 190, 195–197, 214, 233, 235, 236,
344, 346, 376 238, 239, 241–245, 247–253,
incompatibility, 58, 209, 233, 234, 238, 242, 255–257, 260, 275, 276, 279, 311,
247, 318 325, 339–342, 357, 378
injection molding, 120 Moore’s Law, 283
inorganic oxides, 57, 58, 60, 80, 81 multi-walled carbon nanotube, 96, 115, 117,
interatomic, 119, 120, 121, 127 135, 141, 142, 158, 163, 193, 194,
intercalation, 10, 32, 51, 99, 111, 131, 166, 317, 376, 377
167, 177, 213, 219, 226, 233, 235,
242, 248, 260–262, 272, 276, 318,
320, 322, 330, 341, 346, 378 N
interface, 12, 13, 16, 43, 50, 53, 54, 59, 71, nanoclay, 9–11, 19, 99, 115–117, 136,
78, 82, 91, 101–103, 105, 106, 154, 164, 177, 180, 181, 183,
108–110, 113–116, 119, 131, 146, 185–187, 189, 191, 213–215,
148, 191, 201, 206, 209, 233, 234, 217, 219, 222, 224, 310, 313,
238, 250, 264, 271, 275, 295, 308, 317, 318, 325, 326, 378
312, 327, 338 nanocomposite characterization, 262
interface area, 119 nanographite, 3, 7, 10, 11, 17–21, 26, 27, 29,
intumescence, 157 30, 32–36
nanostructured materials, 307
nanotechnology, 91, 131–133, 198, 259, 278,
L 307, 338, 376
layered silicate, 5, 11, 49, 51, 52, 63, 90, 98, natural rubber, 10, 15, 41, 51–53, 55, 66, 70, 82,
99, 114, 116, 117, 131, 157, 158, 83, 90, 95, 114–118, 135–137, 139,
162, 163, 167, 175, 185, 196, 197, 149, 152–154, 160, 176, 189–192,
213, 255, 256, 260, 271, 276, 277, 201, 206–208, 213–215, 219, 220,
313, 318–320, 322, 328, 339–342, 223, 227, 228, 346, 349, 376–378
346, 378 Nyquist plots, 33, 35
Lennard-Jones function, 124
lifecycle, 179
loss modulus, 38–40, 45, 47, 93 O
loss tangent, 34, 37–39, 42, 44, 97, 98, 103 oligomer, 59, 74, 90, 103, 331, 332, 342
opacity, 64
optimisation, 309
M organoclay, 51, 52, 98, 99, 102, 104, 105,
magnetic force microscopy (MFM), 23 115–117, 154, 177, 190, 201,
mechanical milling, 137, 138 212, 213, 226, 255, 256, 276–278,
Mechanical performance, 97, 103, 109, 191, 311, 312, 314, 320–322, 339–341,
233–235, 238, 308, 331 343, 346, 347, 357, 372, 374,
melt flow, 120 375, 378, 379
miniaturelization, 280 orientation, 12, 18, 24, 30, 41, 50, 54, 106,
modification, 15, 61, 75, 84, 98, 108, 110, 108, 113, 128, 132, 135, 136, 145,
131, 136, 146, 155, 157, 161, 146, 148, 151, 153, 210, 211, 235,
162, 167, 171, 174, 190, 196, 262, 307, 309, 322, 345, 346
233–235, 237–254, 256, 272,
333, 335, 374, 376
mold geometry, 120 P
molecular and continuum modeling, 119 packaging, 155, 190, 192, 197, 233–236, 238,
molecular chain entanglement, 73, 122, 142, 239, 243, 247, 259, 329
221, 318 particulate releases, 179
384 Index

P (cont.) Schallamach wave, 355, 361


percolation, 3, 35, 37, 39, 54, 63, service life, 205, 308, 338
291, 344, 345 shape memory alloys, 282, 283
phase changes, 89, 94, 112, 113 sliding wear, 205, 350, 351, 353, 359, 361,
piezo devices, 281 366, 374, 378, 379
piezoelectric materials, 282, 283 sol-gel, 53, 57, 59–62, 65, 66, 70, 71,
poly(aryl-ether–ether–ketone), 307, 334 73–75, 77–85, 189, 277, 293,
polydimethylsiloxane, 160, 186, 197, 313, 336, 342
258, 271 solid state physical properties, 119
poly(p-phenylene benzbisoxazole), 337 solvent intercalation, 10, 32, 51, 99, 111,
polyaniline, 307, 329, 330, 342 131, 166, 167, 177, 213, 219,
polyarylacetylene, 332, 333, 342 226, 233, 235, 242, 248, 260–262,
Polymer Nanocomposite Synthesis, 261 272, 276, 318, 320, 322, 330,
polyurethane, 9, 49, 51, 80, 85, 90, 103, 114, 341, 346, 378
116, 117, 131, 146, 160, 161, 164, specific wear rate, 343, 354–356, 368, 371,
176, 187, 189, 196, 197, 233–240, 373–375
242, 243, 254, 255, 257, 258, 260, steric effect, 158
265, 269, 276–280, 287, 294, 302, stiffness, 5, 11, 12, 33, 58, 67, 121–123, 180,
304, 305, 307, 318, 325, 326, 223, 256, 258, 265, 309, 310, 313,
340–342, 357, 374, 375, 377–379 314, 316, 320, 324, 326, 328, 330,
polyurethane nanocomposites, 235–239, 242, 335–337, 344–346
243, 254, 255, 269, 278, 280, 304, storage modulus, 16, 38–47, 71–73, 77, 80, 93,
305, 340–342, 377 97–99, 103–106, 267, 321
potential energy, 119–121, 123, 124, 127, strain-induced crystallization behavior, 153
133, 283 stress-strain curves, 36, 67, 69, 73, 74, 76,
potential energy functions, 119, 121, 123 138, 141, 142, 144, 147, 149,
precursors, 8, 60, 65, 74, 81 151, 327, 335
processing conditions, 234, 309 structure-property relationship, 71, 82, 83,
processing technique, 58, 335 89, 340
protective barrier, 158, 159 styrene-butadiene rubber, 51, 73, 84, 115, 116,
proteins, 136, 137 131, 135, 139, 170, 206, 349, 376
pseudo bilayer, 250 substrate, 62, 98, 115, 158, 164, 169, 170,
pultrusion, 120 207, 235, 247, 249, 317, 363,
pyrolysis, 50, 157, 182, 187, 188, 191, 192, 364, 379
197, 333 surfactant, 16, 17, 54, 73, 98, 136, 165, 167,
174, 260, 278, 357, 372, 378
suspensions, 16, 91, 241, 247, 248
R swelling, 9, 17, 49, 51, 60, 61, 65, 67, 71–76,
radius of gyration, 10, 12 215, 216, 260, 275, 321, 339, 345
recycling, 179–189, 191, 192, 195–197 synergies, 155, 171, 175
relaxation phenomena, 89, 91, 93–95, 98, 99,
102, 106, 110, 111, 113, 114
reliability, 308 T
renewable, 281, 283 tear strength, 5, 58, 79, 180, 182, 220, 221,
resource cascading, 179, 181, 183–186, 223, 346, 348, 349, 353
188, 191 tensile mode, 93, 98
retreading, 182, 183, 212 tensile tests, 217, 264, 269, 332, 335,
rolling wear, 368, 371–375, 379 336, 357
Rubbing, 14, 207 thermal degradation, 48, 65, 156, 158, 159,
163, 197, 223, 250, 251, 255, 275,
317, 342
S thermal resistance, 156, 160, 164, 217
scanning electron microscopy, 24, 217, 218 thermally stable, 155, 156, 161, 168,
scavenging, 281, 283 174, 330
Index 385

thermoforming, 120 W
thermomechanical, 60, 89, 255, 276, water vapor permeation, 236–238, 242, 243,
277, 338 249
titania, 8, 58, 60, 64–66, 80–82, 180, 194 Wear, 49, 96, 109, 114, 115, 186, 188, 189,
torsional angle, 121, 123 201–208, 213, 214, 226–229, 283,
TPE gels, 3, 17, 26, 28, 29, 43, 45–49 291, 329, 336, 337, 342, 343,
Transmission Electron Microscopy (TEM), 347–359, 361–366, 368–373,
165, 217 378, 379
transitions, 39, 93, 282 wide-angle X-ray diffraction, 137
twin-screw extruder, 312, 313, 335
tyres, 179–184, 186, 187, 189, 191, 192, 198,
201–208, 212, 213 X
X-ray Diffraction (XRD), 167, 262, 317
X-ray scattering, 57, 60, 64, 65, 153
V XRD, 18, 80, 90, 99, 110, 167, 216, 219, 220,
van der Waals interactions, 25, 344 222, 226, 257, 262, 263, 266, 274,
vermiculite, 233, 245, 247–249, 256 317, 330
viscoelastic, 31, 32, 43, 55, 93, 96, 132, 309,
321, 322, 341, 346, 347, 351, 364,
378, 379 Z
vulcanization time, 58 zirconia, 58, 60, 64, 66, 81, 82

You might also like