You are on page 1of 14

This article was downloaded by: [University of Saskatchewan Library]

On: 12 March 2013, At: 22:24


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office:
Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Polymer Reviews
Publication details, including instructions for authors and subscription
information:
http://www.tandfonline.com/loi/lmsc20

Polymer Nanocomposites: The “Nano” Effect


on Mechanical Properties
a a
Alfred J. Crosby & Jong‐Young Lee
a
Polymer Science & Engineering Department, University of Massachusetts
Amherst, Amherst, MA
Version of record first published: 26 Apr 2007.

To cite this article: Alfred J. Crosby & Jong‐Young Lee (2007): Polymer Nanocomposites: The “Nano” Effect
on Mechanical Properties, Polymer Reviews, 47:2, 217-229

To link to this article: http://dx.doi.org/10.1080/15583720701271278

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial
or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the
contents will be complete or accurate or up to date. The accuracy of any instructions, formulae,
and drug doses should be independently verified with primary sources. The publisher shall not
be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or
howsoever caused arising directly or indirectly in connection with or arising out of the use of this
material.
Polymer Reviews, 47:217–229, 2007
Copyright # Taylor & Francis Group, LLC
ISSN 1558-3724 print/1558-3716 online
DOI: 10.1080/15583720701271278

Polymer Nanocomposites: The “Nano” Effect


on Mechanical Properties

ALFRED J. CROSBY AND JONG-YOUNG LEE


Polymer Science & Engineering Department, University of Massachusetts
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

Amherst, Amherst, MA

Polymer nanocomposites offer significant potential in the development of advanced


materials for numerous applications. These novel materials benefit from the synergy
between filler particles and polymer chains that are on similar length scales and the
large quantity of interfacial area relative to the volume of the material. Although
enhanced properties of these materials have been demonstrated by numerous research-
ers, our fundamental knowledge of the “nano” effect in terms of mechanical properties
is not fully developed. In this article, we discuss the important properties of three com-
ponents in a general polymer nanocomposite: the polymer matrix, the nanoscale filler,
and the interfacial region. We highlight theory and experimental observations from
several different fields to help guide the future research and development of under-
standing in this critical field.

Keywords nanocomposites, polymer, nanoparticle, mechanical properties, glass


transition

1. Introduction
Polymer nanocomposites have been proposed or are being used for numerous appli-
cations, ranging from car bumpers to advanced optoelectronic devices. Understanding
the impact of nanofillers on the composite mechanical properties is critical to the
success of all of these applications. Consequently, a large number of research groups
are focused on developing a general framework for predicting, or at least, understanding
how the chemistry and morphology of the polymer matrix synergizes with the surface
chemistry, the size, and the shape of a nanoscale filler to define mechanical properties.
Within this general framework, the underlying mechanisms lie at the intersection of
chemistry, physics, materials science, and continuum mechanics. Therefore, the
researchers involved in this critical area of science have an equally varied assortment
of backgrounds and approaches.
In recent years, several review papers on the mechanical properties of polymer
nanocomposites have been written.1 – 4 These papers provide a summary of the current
observations from both experimental and theoretical research and capture the challenge

Received 15 January 2007; Accepted 2 February 2007.


Address correspondence to Alfred J. Crosby, Polymer Science & Engineering Department, Uni-
versity of Massachusetts Amherst, Amherst 01003 MA. E-mail: crosby@mail.pse.umass.edu

217
218 A. J. Crosby and J.-Y. Lee

of defining a general framework for this field. Therefore, the intent of this review paper is
not to provide a comprehensive survey of the current results, but rather introduce new
researchers to materials factors that must be considered in order to contribute to this field.
At the heart of this field of research lies the motivating question: what is the “nano”
effect on macroscopic mechanical properties? Although this question appears simple and
focused, it holds great depth and complexity. First, what is “nano”? In the field of polymer
nanocomposites, it is generally accepted that “nano” refers to the smallest length scale
associated with the filler. Typically, nanocomposite fillers have length scales ranging
from 0.1 nm to 100 nm. On these length scales, the filler objects approach the length
scale of a single polymer coil and develop unique interactions that give rise to optimal
control of mechanical properties, such as the failure properties (Fig. 1).5 To understand
the “nano” effect on mechanical properties, we are interested in defining mechanisms
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

that develop on nanometer length scales, influence the microscale, and impact the macro-
scale. These mechanisms can include the definition of new microstructures due to the
existence of nanofillers, the enhancement in stress transfer between materials with comp-
lementary properties, or an increase in the density of shear deformation events. The
dominance of these or other mechanisms is dictated by the combined impact of the
material components in the nanocomposite.
In this paper, we provide a review of the three primary “ingredients” in polymer nano-
composites: the polymer matrix, the nanoscale fillers, and the interfacial materials that are
used to enhance dispersion. The “nano” effect is defined by all three of these components;
therefore, it is helpful to consider how each component can affect the interpretation of
nanocomposite mechanical properties. In addition to the nanocomposite “ingredients”,
we also provide a short discussion on novel measurement methods for mechanical prop-
erties of polymer nanocomposites.

Figure 1. Polymer nanocomposites take advantage of the synergy of filler particles and polymer
chains on similar length scales. An example of this synergy is the enhancement in mechanical
properties, such as the strain to failure.5
Polymer Nanocomposites 219

2. Polymer Matrices
The primary component by volume in a polymer nanocomposite is the polymer matrix.
Polymers have a wide range of properties that are heavily dependent upon their
chemistry, molecular architecture, and processing history. Accordingly, all of these
factors influence the definition of a general mechanism for polymer nanocomposite mech-
anical properties.

2.1 Linear Polymers


Linear polymers represent the fundamental building block for developing a general
framework of polymer nanocomposites mechanical properties. In considering the proper-
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

ties of solid-like polymer nanocomposites, linear polymer chains can be classified into two
categories: semi-crystalline polymers and amorphous linear chains. For semi-crystalline
polymer matrices, crystallinity and morphology can be altered by the presence of nanofil-
lers.6 – 12 Nanoscale fillers also can interact with multiple length scales related to the
lamellae and spherulitic structures. Due to these complexities and the great breadth of
this component of polymer nanocomposites, we will not address this materials class in
this review, but refer the interested reader to several references.6 – 12
For amorphous linear chains, the nanoparticle-polymer interaction is both entropic
and enthalpic in nature. This interaction is similar in many ways to a blend between
two polymer components.2,3 In a polymer blend system, the enthalpic interaction
between two polymers (xAB) and translational entropy of polymer segments define
the free energy of the blend. However, additional terms must be considered for a
polymer blended with nanoparticles. These terms include the entropic interaction
among nanofillers, including excluded volume interactions, and the depletion interaction
between polymer chains and nanofillers.2,3 Polymer chains in the vicinity of nanofillers are
stretched, thus decreasing the conformational entropy of chains. To maximize confor-
mational entropy, nanofillers are repelled by polymer chains. This depletion interaction
can lead to a low density region of chain segments around the rigid particle surface.
Recent research by Balazs et al. has presented a modified Flory-Huggins theory to
account for these interactions. These theories can be used in describing polymer/nanofiller
phase diagrams and their associated impact on mechanical properties.2,13 – 15
Due to the strong dependence on the entropic contribution for most polymer-based
materials, the processing history of the polymer matrix is critical in defining the free
energy and associated mechanical properties. The processing of the polymer matrix
begins with the synthesis of the polymer. Synthesis can be conducted in situ with the nano-
particles or prior to blending with the nanofiller objects. In either case, the specific
synthesis method (e.g. anionic polymerization, free radical, condensation, atom transfer
radical polymerization) can greatly impact the resulting material properties. Among the
different factors that can affect material properties from different synthetic routes are
residual initiator or transfer agents, different chain terminating chemistries, and differ-
ences in polydispersity. The differences in polydispersity from various synthetic routes
can be especially important for the interpretation of polymer nanocomposite mechanical
properties. Even in small quantities, the existence of low molecular weight chains in
the polymer can lead to a significant decrease in the elastic modulus and strength of the
polymer matrix. This effect may be maximized in polymer nanocomposites since the
presence of a rigid interface (i.e. nanoparticle surface) may drive the local segregation
of lower molecular weight chains distributing them throughout the polymer matrix.
220 A. J. Crosby and J.-Y. Lee

Additionally, for in situ polymerization routes, the presence of the nanoparticle


surface can impact the resulting properties of the synthesized polymer matrix. For
example, in free radical polymerization, the presence of an interface may impact the
rate of chain transfer and chain termination. This alteration of the local polymerization
will lead to differences in the properties of polymer molecules near nanofiller interfaces
compared to the bulk. Therefore, although in situ polymerization can lead to enhanced dis-
persion for nanoparticles, it may ultimately define a new material, which cannot be effec-
tively compared to a baseline material. This impact on materials properties is very
interesting, but the “nano” effect on resulting materials properties is greatly convoluted
by the “nano” effect on the polymerization of the matrix.
As in situ polymerization may make control experiments with baseline materials
difficult, the blending of the polymer nanocomposite ingredients can also convolute the
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

mechanisms related to the “nano” effect on mechanical properties. In blending, polymer


molecules are synthesized prior to interaction with the nanoscale fillers and the individual
components are then combined by either melt mixing or blending of dilution solutions. In
melt mixing, it is widely known that the dispersion of nanofillers is difficult; therefore, the
nanofiller clusters, aggregates, or coalesces.16,17 These processes impact the effective size
of the nanofiller, thus altering the targeted advantages of the nanoscale filler. In the
blending of dilute solutions, a third component, i.e. the solvent, is introduced into the pro-
cessing history. This additional component dictates that the resulting materials structure
will depend not only on the polymer/nanofiller interactions, but also on the polymer/
solvent and nanofiller/solvent interactions. Additionally, the interactions of any of
these components with a bounding substrate and the rate and directionality of the
solvent evaporation must be considered. Residual solvent can also serve as a plasticizer
in the polymer matrix and greatly impact the thermo-mechanical properties of the
nanocomposite.
The geometric confinement of polymer chains is also known to greatly impact
thermo-mechanical properties. These effects have been and are currently a great interest
in the polymer science community and will greatly impact the development of mechanical
properties in polymer nanocomposites. For example, the mobility and/or glass transition
temperature (Tg) of a polymer has been shown to be different in thin films compared to
bulk samples.18,19 This confinement effect on the mobility has been related to the
depression of Tg in amorphous polymer nanocomposites.20 – 22 (Fig. 2) The analogy is
that when the volume fraction of nanofillers is high, the inter-filler distance is small and
the polymer chains between filler objects are confined. In a related manner, experimental
evidence also suggests that the entanglement density of linear polymers changes in
confined geometries.23 This effect has been qualitatively attributed to influencing failure
properties in high volume fraction polymer nanocomposites,5 and it will be an
important consideration for future nanocomposite research and design.

2.2 Polymer Networks


Polymer networks are formed through the chemical and/or physical crosslinking of
polymer chains. These materials represent an important class of polymer matrices for
polymer nanocomposites since they are used in numerous structural applications due to
their long-term resistance to mechanical loads, elasticity, and temperature resistant attri-
butes. The most commonly used polymer networks include epoxies, polyurethanes, and
thermoplastic elastomers. In many ways, the presence of nanoparticles in a polymer
network can influence the polymer matrix in a similar manner to linear polymer
Polymer Nanocomposites 221
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

Figure 2. Demonstrated depression in glass transition temperature (Tg) for polystyrene thin films
and polystyrene-based nanocomposites as a function of average interparticle spacing. Figure used
with permission from Bansal et al.20

matrices. The nanoparticles can influence the development of local network structure
during polymerization and crosslinking reactions, and the excluded volume of the nano-
particles can impact the local dynamics of chains between crosslinks. In some polymer
networks below Tg, the nanofillers can behave as antiplasticizers and depress the short-
scale cooperative motions, eventually decreasing the sub-Tg relaxation (b relaxation).
Additionally, for polymer networks above Tg, the addition of nanofillers can lead to a
secondary network due to the adhesion of the polymer segments to the filler surface.
This secondary network controls the mechanical properties at high deformation.

3. Nanoscale Filler
A wide range of nanoscale fillers have been explored over the past two decades in conjuc-
tion with polymer nanocomposites, but in this review we will focus on the three primary
attributes of the nanoscale filler that impact the development of mechanical properties:
chemistry, size, and shape.

3.1 Nanofiller Chemistry


The chemistry of the nanoscale filler influences two primary properties related to the
polymer/nanofiller interaction. First, the chemistry of the nanofiller contributes to the
enthalpic interaction with the polymer chain. The enthalpic interactions play a large
role in the efficiency of stress transfer across the nanofiller/polymer interface. These
enthalpic interactions can be defined by the van der Waals interactions between the nano-
particle and polymer chains, or they can be related to specific interactions, such as covalent
bonds. The strength of these interactions strongly influences the morphology of a polymer
nancomposite. As mentioned earlier, the equilibrium morphology of a polymer nanocom-
posite is largely described by the polymer blend theory.13 As in the polymer blend theory,
222 A. J. Crosby and J.-Y. Lee

the interaction parameter between particle and polymer can shift the miscibility bound-
aries for given size ratios of the filler and the polymer.13
The van der Waals interaction between neighboring filler particles is also strongly
influenced by the filler chemistry, which is often described in terms of the Hamaker
constant for the material. These interparticle interactions are important at high
volume fractions of the filler in the nanocomposite and in the strength of the filler
aggregates. In terms of this latter effect, nanofillers can phase separate and form
domains that are rich in the nanofiller and poor in the polymer. In these domains, if
the inter-particle interactions are highly attractive, the aggregate will behave as a
large filler particle rather than independent nanoscale fillers. Conversely, if the inter-
particle attractions are weak, then the deformation processes related to the aggregate
may play a large role in the storage and loss of applied mechanical energy. These
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

deformation processes will greatly influence the development of mechanical properties


in the polymer nanocomposite. One example of these inter-filler interactions and their
impact on mechanical properties is within the field of polymer-clay nanocomposites.9
Here, many research groups have demonstrated the impact of intercalated and exfo-
liated morphologies on the mechanical properties. In the exfoliated morphology, the
filler particles are well-dispersed and act independently in context with the polymer
matrix. Here, the length scale of the exfoliated filler is small, but the geometric
stiffness of the filler particle is also small. In intercalated morphologies, the clay
fillers are well-dispersed, but the clay filler is comprised of many layers of clay
sheets that are strongly attracted to each other. These strong attractions cause the inter-
calated clay fillers to behave similarly to larger filler particles with smaller surface-to-
volume ratios. If additional molecules are added to weaken the inter-sheet interactions,
the mechanical properties of intercalated clay nanocomposites are altered.
The mechanical properties of the filler are also defined by its chemistry. Similar to
conventional composites, the mechanical properties of the filler relative to the mechan-
ical properties of the polymer matrix define the local constraint. The impact of the
mechanical properties of the filler on the overall mechanical properties of the nanocom-
posite has been explored by several research groups, experimentally and theoretically.
Liu and Brinson provide a recent contribution that summarizes much of this work from
the continuum mechanics viewpoint.24

3.2 Nanofiller Size and Shape


The size of the filler is at the heart of the “nano” effect on mechanical properties in nano-
composites. Closely coupled with the size is the shape of the nanofiller. In general, we can
consider the filler size and shape as dictating two important contributions to the overall
polymer nanocomposite properties:
1. surface-to-volume ratio of the fillers and
2. excluded volume interactions.
The surface-to-volume ratio of the filler is the primary motivation for the development of
polymer nanocomposites. This ratio indicates the quantity of interfacial region compared
to bulk material in the composite. Among the many reasons for its importance, the inter-
facial region controls new structural arrangements on the molecular scale and is respon-
sible for the efficient transfer of stress across the composite components. Therefore,
maximizing the quantity of interfacial region maximizes the potential of defining new
material properties.
Polymer Nanocomposites 223

For spherical particles with radius r, the surface to volume ratio of the filler is simply:
As 4p r2 3
¼ 3
¼ ð1Þ
Vs 4=3p r r

This relationship demonstrates that the available surface area (As) per volume (Vs) of filler
increases as the spherical radius decreases. The magnitude of property change or control
by the interfacial region is dictated by the total quantity of the interfacial area within a
nanocomposite. In a nanocomposite with filler volume fraction f, the total interfacial
area-to-volume ratio scales directly with f, such that:
Ai;total 3f
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

¼ ð2Þ
Vtotal r

This relationship has two important implications. First, if f is held constant, then decreas-
ing the filler radius increases the available surface area for interfacial interactions. If we
assume in a linear polymer nanocomposite that stress is transferred through adsorbed
chain segments at the filler interface, then each chain/filler interaction will occupy a
given interfacial area. Therefore, increasing the available interfacial area increases the
quantity of chain/filler interactions. This effect increases the efficiency of stress transfer
between the fill and polymer matrix. Additionally, this relationship demonstrates that
two composites with identical interfacial areas can be fabricated with different filler
sizes, if f is controlled independently. This approach can aid in the decoupling of size
effects from other mechanisms related to the “nano” effect.
Shape also plays a role in the design of surface-to-volume ratios for nanocomposites.
For a cylindrical filler of radius r and length L, the surface to volume ratio of the filler is:
Ac 2p r2 þ 2p rL 2 2
¼ ¼ þ ð3Þ
Vc p r2 L r L

If we compare spherical and cylindrical fillers at equal volume fractions, we find that the
surface-to-volume (SV) ratio for the sphere relative to the cylinder scales as:
SVs 3
¼ ð4Þ
SVc 2ð1 þ r=LÞ

This relationship demonstrates that for all plates (r . L) and short rods (L , 2.0r) the
surface to volume ratio for cylindrical fillers is greater than the ratio for a spherical
particle. (Fig. 3) However, spherical fillers do have greater surface to volume ratios
compared to long fibers (L . 2.0r), but the maximum increase is only 50%. Therefore,
if the surface-to-volume ratio is the primary design factor, it is clear that plate geometries,
similar to clay sheets, have a significant advantage at equivalent volume fraction loadings.
Unfortunately, other design factors are critical for the design of mechanical properties in
filled composites, and shape must be equally considered for these factors. For example,
rigid cylindrical fillers are difficult to disperse at high volume fractions in an isotropic
manner. Therefore, the mechanical properties of non-isotropically dispersed fillers will
likely be anisotropic in nature. This design limitation may be advantageous for some appli-
cations, but not all.
224 A. J. Crosby and J.-Y. Lee
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

Figure 3. Surface to volume ratio for spherical particles (SVS) compared to surface to volume ratio
for cylindrical particles (SVC) as a function of particle radius (r) and length (L).

The difficulty of dispersing non-spherical fillers isotropically has been explored in


the field of colloidal dispersions and is largely attributed to the excluded volume
effect of rigid fillers and the ease of packing non-spherical objects.25 The excluded
volume contribution is derived from the fact that two hard filler objects cannot occupy
the same space at the same time; therefore, a repulsive force develops due to the
entropy of the filler objects. Straightforward discussions of the impact of excluded
volume contributions can be found in text books.25 An important conclusion from
excluded volume contributions is that high-aspect ratio objects, such as long cylinders,
are increasingly difficult to disperse isotropically as the aspect ratio, L/r, increases. In
other words, ordered phases, similar to liquid crystalline transitions, will develop at
low volume fractions of fillers.
The excluded volume of the nanoscale fillers coupled with the configurational entropy
of the polymer chain has been shown to play an important role in the spatial distribution of
nanofillers in polymer matrices. This control has been demonstrated with many non-mech-
anical investigations, such as the control of nanoparticles within ordered block copolymer
domains.26 – 33 Balazs et al. has demonstrated with molecular-based simulations, and
Russell et al. has experimentally demonstrated that the excluded volume and configura-
tional entropy of the polymer chains can drive nanoscale fillers to defects in surface
coatings, such as cracks.31,34 Similar to these effects, Crosby et al. has demonstrated
that these nanoscale mechanisms play a large role in altering the deformation and
failure mechanisms of linear glassy polymers.5 Specifically, they have shown that the con-
figurational entropy of the polymer chains can “push” fillers, which are on similar length
scales to the polymer coils. This mobility leads to altered morphologies of crazes in the
glassy polymers (Fig. 4) and ultimately to an increased strain-to-failure for these
materials (Fig. 1). Other researchers have also addressed the importance or role of nano-
filler mobility in the mechanical properties.6 – 21 These contributions of “nano” scale
effects are interesting since they point to the importance of nanofiller mobility, i.e.
weak interfacial interactions, in the development or control of mechanical properties.
Although strong interfacial interactions may be required for balancing other mechanical
properties, the importance and novel use of mobility should not be overlooked.
Polymer Nanocomposites 225
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

Figure 4. Alignment of CdSe nanoparticles in the precraze region of a polystyrene-CdSe nanocom-


posite. Scale bar is 25 nm.5

4. Interfacial Region
To maximize the potential of nanoscale fillers in polymer nanocomposites, the dispersion
of the filler is often an important goal. A common strategy to increase the probability of
isotropic dispersion is to modify the surface of the nanoscale filler to increase favorable
interactions with the polymer matrix and minimize inter-filler attraction. Numerous modi-
fication strategies for the nanofiller surface have been demonstrated, but a common
approach is the binding of small molecules or polymer ligands to the surface of the
nanoscale filler. Although this approach can significantly increase the probability of dis-
persion, the properties of the ligand layer must be considered in the development of mech-
anical properties in polymer nanocomposites.
One example is the impact of the surface bound molecules on the glass transition temp-
erature of a nanocomposite. Consider the use of short polystyrene chains attached to nano-
particles to disperse the particles in a polystyrene matrix. The glass transition temperature of
polystyrene decreases significantly for molecular weights less than 10,000 g/mol. For
example, the Tg of polystyrene MW ¼ 1000 g/mol is approximately 208C. If we assume
that a simple Fox relationship35 holds true and the volume of the polystyrene ligands is rep-
resented by a fraction of the nanoparticle volume fraction fN, then the predicted change in
the Tg of the nanocomposite compared to the bulk polymer Tg (fTg) is:

 
1
DTg ¼ TgB 1  ð5Þ
1 þ ð3fN =1  fN Þð1L =rN ÞðTgB =TgL  1Þ

Where TgB is the bulk polymer Tg, lL is the polystyrene ligand thickness, rN is the
nanoparticle radius, and TgL is the bulk Tg of the non-bound ligand molecules. This
relationship is plotted in Fig. 5 and demonstrates the impact of small molecule dispersing
agents on the thermo-mechanical properties of a nanocomposite due to the high surface to
volume ratio for these materials.
226 A. J. Crosby and J.-Y. Lee
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

Figure 5. Predicted depression of glass transition temperature (DTg) as a function of nanoparticle


volume fraction. Plotted for different ratios of polystyrene ligand coating thickness to nanoparti-
cle radius (l/r). Polystyrene ligand assumed to be 1000 g/mol with Tg of 208C. Modified
nanoparticles blended with bulk polystyrene.

5. Measurement of Mechanical Properties


In addition to conventional methods for the measurement of mechanical properties, two
alternative techniques are convenient for quantifying the mechanical properties of
polymer nanocomposites. One technique is a classical technique, commonly referred to
as the copper grid technique, which provides morphological and energetic information
related to the deformation and failure properties of glassy polymers. The second
technique, SIEBIMM (Strain-Induced Elastic Buckling Instability for Mechanical
Measurements),36,37 was introduced about three years ago and has been gaining utility
in the investigation of low strain properties of advanced materials. The advantage for
both of these techniques in terms of polymer nanocomposites is that they require very
small sample volumes. This advantage is significant when well-defined model nanoparti-
cles and model polymer matrices are used to isolate specific contributions to the “nano”
effect on mechanical properties.
The copper grid technique was generally introduced in 1979 by Lauterwasser and
Kramer,38 and has been a primary tool for several researchers in the investigation of
crazing and local deformation mechanisms in linear polymer films.39 – 41 More recently,
researchers such as Crosby and Yang have adapted this technique for investigating defor-
mation and failure properties in polymer nanocomposites (example data provided in Figs.
1 and 4).5,42 The general procedure begins with the preparation of thin film on a solid
substrate. This film is transferred via a floating technique to the surface of a copper
grid, whose bars have been previously coated with a thin polymer coating for
enhancing adhesion. A strain stage is then used to apply a fixed strain to the copper
grid and sample film. As a function of strain, the copper grid assembly is removed from
the strain stage, while maintaining the primary fraction of applied strain, to perform
chosen microscopy techniques, such as TEM, AFM, fluorescent confocal microscopy,
or optical microscopy, to determine the morphology and failure properties of the film.
In addition to the advantage of working with small sample volumes and obtaining high
Polymer Nanocomposites 227

resolution morphological information under an applied strain, a secondary advantage is


that one copper grad sample may contain over 100 individual grid spans, which behave
independently to the applied strain. These “individual samples” are all prepared and
tested at the same time; therefore, results are statistical and can incorporate gradients to
increase throughput in covering a given parameter space.43
SIEBIMM (Strain-Induced Elastic Buckling Instability for Mechanical Measure-
ments) involves the floating of a thin sample film (e.g. polymer nanocomposite) onto a
thick, crosslinked elastomer substrate.36,37 The elastomer substrate is subsequently
strained, thus transferring strain to the sample film if good adhesion is maintained. Due
to the large difference in moduli between the sample film and the substrate, a net compres-
sive stress develops and buckles the nanocomposite thin film. Above a critical stress, the
compressive stress induces the onset of an elastic instability, and surface wrinkling of a
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

given wavelength is observed on the sample film surface. From the buckling wavelength
(l), the modulus of sample film (Ep) can be obtained by:36,37
 
ð1  v2p Þ l 3
Ep ¼  3E S  ð6Þ
ð1  v2S Þ 2ph

ES is the elastic modulus of substrate. Poisson’s ratio of the substrate and sample film are
nS and np respectively. The thickness of the sample film is h, and l is obtained via Fast
Fourier Transform (FFT) of optical images or by light scattering. This technique has
been used by several research groups recently to provide interesting insight into the mech-
anical properties of polymer thin films, and its impact on our understanding of model
polymer nanocomposites will continue to grow due to the ease, efficiency, and versatility
of the technique.36,37,44 – 46

6. Summary
The combination of particles and polymer chains that are on similar length scales offers
numerous advantages and considerable excitement for the possibility of intelligently
designing predictable, optimized material structures. These materials will undoubtedly
provide an enhanced function for a wide range of applications, but fundamental
knowledge of the mechanisms responsible for these enhanced properties is essential for
using them to their fullest potential. In this paper, we have outlined a few critical properties
of the main “ingredients” in polymer nanocomposites. We have not attempted to provide a
comprehensive review of this broad and growing field, but we have highlighted several
references to other research areas, past and current, within the field of polymer science
upon which our understanding of the “nano” effect in polymer nanocomposites can be built.

Acknowledgements
The authors gratefully acknowledge the financial support of the Petroleum Research Fund
of the American Chemical Society and the NSF-MRSEC at the University of Massachu-
setts Amherst.

References
1. Tjong, S. C. “Structural and mechanical properties of polymer nanocomposites”, Materials
Science & Engineering R-Reports 2006, 53 (3 – 4), 73 –197.
228 A. J. Crosby and J.-Y. Lee

2. Balazs, A. C.; Emrick, T.; Russell, T. P. “Nanoparticle polymer composites: Where two small
worlds meet”, Science 2006, 314 (5802), 1107– 1110.
3. Mayes, A. M. “Nanocomposites—softer at the boundary”, Nature Materials 2005, 4 (9),
651– 652.
4. Ahmed, S.; Jones, F. R. “A review of particulate reinforcement theories for polymer compo-
sites”, Journal of Materials Science 1990, 25 (12), 4933– 4942.
5. Lee, J. Y.; Zhang, Q. L.; Emrick, T.; Crosby, A. J. “Nanoparticle alignment and repulsion during
failure of glassy polymer nanocomposites”, Macromolecules 2006, 39 (21), 7392–7396.
6. Shah, D.; Maiti, P.; Jiang, D. D.; Batt, C. A.; Giannelis, E. P. “Effect of nanoparticle mobility on
toughness of polymer nanocomposites”, Advanced Materials 2005, 17 (5), 525.
7. Krikorian, V.; Pochan, D. J. “Unusual crystallization behavior of organoclay reinforced poly(L -
lactic acid) nanocomposites”, Macromolecules 2004, 37 (17), 6480–6491.
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

8. Bikiaris, D. N.; Papageorgiou, G. Z.; Pavlidou, E.; Vouroutzis, N.; Palatzoglou, P.;
Karayannidis, G. P. “Preparation bv melt mixing and characterization of isotactic polypropy-
lene/SiO2 nanocomposites containing untreated and surface-treated nanoparticles”, Journal
of Applied Polymer Science 2006, 100 (4), 2684– 2696.
9. Sheng, N.; Boyce, M. C.; Parks, D. M.; Rutledge, G. C.; Abes, J. I.; Cohen, R. E. “Multiscale
micromechanical modeling of polymer/clay nanocomposites and the effective clay particle”,
Polymer 2004, 45 (2), 487– 506.
10. Varlot, K.; Reynaud, E.; Kloppfer, M. H.; Vigier, G.; Varlet, J. “Clay-reinforced polyamide:
Preferential orientation of the montmorillonite sheets and the polyamide crystalline
lamellae”, Journal of Polymer Science Part B-Polymer Physics 2001, 39 (12), 1360– 1370.
11. Saujanya, C.; Radhakrishnan, S. “Structure development and crystallization behaviour of PP/
nanoparticulate composite”, Polymer 2001, 42 (16), 6723–6731.
12. Chen, T. K.; Tien, Y. I.; Wei, K. H. “Synthesis and characterization of novel segmented poly-
urethane clay nanocomposite via poly(1-caprolactone)/clay”, Journal of Polymer Science Part
a-Polymer Chemistry 1999, 37 (13), 2225– 2233.
13. He, G.; Ginzburg, V. V.; Balazs, A. C. “Determining the phase behavior of nanoparticle-filled
binary blends”, Journal of Polymer Science Part B-Polymer Physics 2006, 44 (17), 2389– 2403.
14. Smith, K. A.; Tyagi, S.; Balazs, A. C. “Healing surface defects with nanoparticle-filled polymer
coatings: Effect of particle geometry”, Macromolecules 2005, 38 (24), 10138– 10147.
15. Tyagi, S.; Lee, J. Y.; Buxton, G. A.; Balazs, A. C. “Using nanocomposite coatings to heal
surface defects”, Macromolecules 2004, 37 (24), 9160– 9168.
16. Fornes, T. D.; Yoon, P. J.; Paul, D. R. “Polymer matrix degradation and color formation in melt
processed nylon 6/clay nanocomposites”, Polymer 2003, 44 (24), 7545– 7556.
17. Ishida, H.; Campbell, S.; Blackwell, J. “General approach to nanocomposite preparation”,
Chemistry of Materials 2000, 12 (5), 1260– 1267.
18. Keddie, J. L.; Jones, R. A. L.; Cory, R. A. “Size-dependent depression of the glass-transition
temperature in polymer-films”, Europhysics Letters 1994, 27 (1), 59 – 64.
19. Kawana, S.; Jones, R. A. L. “Character of the glass transition in thin supported polymer films”,
Physical Review E 2001, 6302.
20. Bansal, A.; Yang, H. C.; Li, C. Z.; Cho, K. W.; Benicewicz, B. C.; Kumar, S. K.; Schadler, L. S.
“Quantitative equivalence between polymer nanocomposites and thin polymer films”, Nature
Materials 2005, 4 (9), 693– 698.
21. Ash, B. J.; Rogers, D. F.; Wiegand, C. J.; Schadler, L. S.; Siegel, R. W.; Benicewicz, B. C.;
Apple, T. “Mechanical properties of Al2O3/polymethylmethacrylate nanocomposites”,
Polymer Composites 2002, 23 (6), 1014– 1025.
22. Starr, F. W.; Schroder, T. B.; Glotzer, S. C. “Molecular dynamics simulation of a polymer melt
with a nanoscopic particle”, Macromolecules 2002, 35 (11), 4481– 4492.
23. Si, L.; Massa, M. V.; Dalnoki-Veress, K.; Brown, H. R.; Jones, R. A. L. “Chain entanglement in
thin freestanding polymer films”, Physical Review Letters 2005, 94.
Polymer Nanocomposites 229

24. Liu, H.; Brinson, L. C. “A hybrid numerical-analytical method for modeling the viscoelastic
properties of polymer nanocomposites”, Journal of Applied Mechanics-Transactions of the
Asme 2006, 73 (5), 758– 768.
25. Jones, R. A. L. Soft Condensed Matter; Oxford University Press: New York, 2002.
26. Zhang, Q. L.; Gupta, S.; Emrick, T.; Russell, T. P. “Surface-functionalized CdSe nanorods for
assembly in diblock copolymer templates”, Journal of the American Chemical Society 2006,
128 (12), 3898– 3899.
27. Yoon, J.; Lee, W.; Thomas, E. L. “Self-assembly of block copolymers for photonic-bandgap
materials”, Mrs Bulletin 2005, 30 (10), 721– 726.
28. Bockstaller, M. R.; Mickiewicz, R. A.; Thomas, E. L. “Block copolymer nanocomposites: Per-
spectives for tailored functional materials”, Advanced Materials 2005, 17 (11), 1331– 1349.
29. Schultz, A. J.; Hall, C. K.; Genzer, J. “Computer simulation of block copolymer/nanoparticle
composites”, Macromolecules 2005, 38 (7), 3007–3016.
Downloaded by [University of Saskatchewan Library] at 22:24 12 March 2013

30. Lin, Y.; Boker, A.; He, J. B.; Sill, K.; Xiang, H. Q.; Abetz, C.; Li, X. F.; Wang, J.; Emrick, T.;
Long, S.; Wang, Q.; Balazs, A.; Russell, T. P. “Self-directed self-assembly of nanoparticle/
copolymer mixtures”, Nature 2005, 434 (7029), 55 – 59.
31. Lee, J. Y.; Buxton, G. A.; Balazs, A. C. “Using nanoparticles to create self-healing composites”,
Journal of Chemical Physics 2004, 121 (11), 5531– 5540.
32. Buxton, G. A.; Balazs, A. C. “Simulating the morphology and mechanical properties of filled
diblock copolymers”, Physical Review E 2003, 67.
33. Huh, J.; Ginzburg, V. V.; Balazs, A. C. “Thermodynamic behavior of particle/diblock
copolymer mixtures: Simulation and theory”, Macromolecules 2000, 33 (21), 8085– 8096.
34. Gupta, S.; Zhang, Q. L.; Emrick, T.; Balazs, A. C.; Russell, T. P. “Entropy-driven segregation of
nanoparticles to cracks in multilayered composite polymer structures”, Nature Materials 2006,
5 (3), 229– 233.
35. Fox, T. G. Bulletin of the American Physical Society 1956, 2 (2), 123.
36. Stafford, C. M.; Guo, S.; Harrison, C.; Chiang, M. Y. M. “Combinatorial and high-throughput
measurements of the modulus of thin polymer films”, Review of Scientific Instruments 2005, 76.
37. Stafford, C. M.; Harrison, C.; Beers, K. L.; Karim, A.; Amis, E. J.; Vanlandingham, M. R.;
Kim, H. C.; Volksen, W.; Miller, R. D.; Simonyi, E. E. “A buckling-based metrology for
measuring the elastic moduli of polymeric thin films”, Nature Materials 2004, 3 (8), 545– 550.
38. Lauterwasser, B. D.; Kramer, E. J. “Microscopic mechanisms and mechanics of craze growth
and fracture”, Philosophical Magazine a-Physics of Condensed Matter Structure Defects and
Mechanical Properties 1979, 39 (4), 469– 495.
39. Creton, C.; Kramer, E. J.; Hui, C. Y.; Brown, H. R. “Failure mechanisms of polymer interfaces
reinforced with block copolymers”, Macromolecules 1992, 25 (12), 3075– 3088.
40. Donald, A. M.; Kramer, E. J. “Effect of molecular entanglements on craze microstructure in
glassy-polymers”, Journal of Polymer Science Part B-Polymer Physics 1982, 20 (5), 899– 909.
41. Lee, J. Y.; Crosby, A. J. “Crazing in glassy block copolymer thin films”, Macromolecules 2005,
38 (23), 9711– 9717.
42. Hsiao, C. C.; Lin, T. S.; Cheng, L. Y.; Ma, C. C. M.; Yang, A. C. M. “Nanomechanical prop-
erties of polystyrene thin films embedded with surface-grafted multiwalled carbon nanotubes”,
Macromolecules 2005, 38 (11), 4811– 4818.
43. Crosby, A. J.; Fasolka, M. J.; Beers, K. L. “High-throughput craze studies in gradient thin films
using ductile copper grids”, Macromolecules 2004, 37 (26), 9968– 9974.
44. Nolte, A. J.; Cohen, R. E.; Rubner, M. F. “A two-plate buckling technique for thin film modulus
measurements: Applications to polyelectrolyte multilayers”, Macromolecules 2006, 39 (14),
4841– 4847.
45. Stafford, C. M.; Vogt, B. D.; Harrison, C.; Julthongpiput, D.; Huang, R. “Elastic moduli of
ultrathin amorphous polymer films”, Macromolecules 2006, 39 (15), 5095– 5099.
46. Wilder, E. A.; Guo, S.; Lin-Gibson, S.; Fasolka, M. J.; Stafford, C. M. “Measuring the modulus
of soft polymer networks via a buckling-based metrology”, Macromolecules 2006, 39 (12),
4138– 4143.

You might also like