Sie sind auf Seite 1von 11

Materials Science and Engineering A 393 (2005) 1–11

Review

Experimental trends in polymer nanocomposites—a review


Jeffrey Jordana , Karl I. Jacobb , Rina Tannenbaumc ,
Mohammed A. Sharafb , Iwona Jasiukd,∗
a School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405, USA
b School of Polymer, Textile and Fiber Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0295, USA
c School of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0245, USA
d Department of Mechanical and Industrial Engineering, Concordia University, Montreal, Que. H3G IM8, Canada

Received 17 November 2003; received in revised form 24 September 2004; accepted 27 September 2004

Abstract

A review of the recent work on polymer matrix nanocomposites is presented. This review is not intended to be comprehensive, but provides
an overview of the processing techniques and trends in the mechanical behavior and morphology of nanocomposites. A number of composite
systems with amorphous and/or crystalline polymer matrices and different nano-sized filler materials are discussed.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Nanocomposites; Polymer matrix composites; Mechanical properties

Contents

1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Experimental procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.1. Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.2. Viscoelastic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3. Crystallinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.4. Density/volume change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1. Introduction nature. However, polymers have lower modulus and strength


as compared to metals and ceramics. One way to improve
Polymer systems are widely used due to their unique at- their mechanical properties is to reinforce polymers with in-
tributes: ease of production, light weight, and often ductile clusions (fibers, whiskers, platelets, or particles). The embed-
ding of inclusions in a host matrix to make composites, which
gives material properties not achieved by either phase alone,
∗ Corresponding author. Tel.: +1 514 848 2424x3143; has been a common practice for many years. Using this ap-
fax: +1 514 848 3175. proach, polymer properties can be improved while maintain-
E-mail address: ijasiuk@me.concordia.ca (I. Jasiuk). ing their light weight and ductile nature [1–6]. Improvements

0921-5093/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2004.09.044
2 J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11

in properties can often be found even at relatively low filler volume fraction of particles ranged from 0 to 25%. Sample
content [4,7,8]. preparation consisted of dissolving polymers in a polar sol-
Traditionally, composites were reinforced with micron- vent and mixing in the beads for several hours. The mixture
sized inclusions. Recently, processing techniques have been was then poured over a large surface to allow the solvent to
developed to allow the size of inclusions to go down to evaporate and it was subsequently dried under a vacuum at
nanoscale. For this work, the nano-sized inclusions are de- 100 ◦ C. This mixture involved 30% volume fraction of parti-
fined as those that have at least one dimension in the range cles. Pure polymer was then added to samples to achieve the
1–100 nm. Experiments have shown that nanoscale reinforce- desired particle volume fractions.
ment brings new phenomena, which contribute to material Chan et al. [11] made nanocomposites with polypropylene
properties. In this review, the interest is in how mechanical (PP) matrix and calcium carbonate (CaCO3 ) through melt
properties of polymer matrix composites are altered by intro- mixing of the components. First the components were dried
ducing nano-sized versus micron-sized reinforcement, and in an oven at 120 ◦ C and then cooled to a room temperature.
what additional factors contribute to the material response of The polypropylene was mixed first with an anti-oxidant. The
nanocomposites. CaCO3 nanoparticles, which were 44 nm in diameter, were
Current micromechanics theories rely on the idea that the then added slowly and mixing continued for a fixed time after
effective properties of composite materials, such as Young’s all particles were added. This technique produced good, rea-
modulus, are functions of properties of constituents, volume sonably well-dispersed samples at lower filler volume frac-
fraction of components, shape and arrangement of inclusions, tions, 4.8% and 9.2%, but aggregation was found at a higher
and matrix-inclusion interface. These theories, therefore, pre- volume fraction, 13.2%. Other researchers have also used a
dict that the properties of composite materials are indepen- form of melt-mixing to produce nanocomposites [17].
dent of the size of inclusions. In general, this is correct for Polyurethane–silica nanocomposites have been made by
systems with micron size reinforcement, but, as mentioned Pétrovic et al. [10] by first mixing the silica with polyol. The
above may not be true for nanocomposite systems. mixture was then cured with diisocyanate at 100 ◦ C for 16 h
With the recent developments in the nanoscience and nan- in presence of 0.1% catalyst Cocure 55 (from CasChem) and
otechnology fields, the correlation of material properties with subsequently poured into a mold. The particles were spherical
filler size has become a point of great interest. As a result, with an average diameter of 12 nm and had narrow size dis-
much of the work is still ongoing and there is yet to be a tribution (10–20 nm). Good samples were produced at 10%,
definite conclusion on the effect of nano-sized inclusions on 20%, 30%, and 40% filler weight fraction, but the 50% filler
polymer systems. In the following sections, a review is pre- weight fraction sample did not cure completely.
sented on the processing, experimental results, and possible It is believed that one of the main issues in preparing good
interpretations of those results for polymer matrix nanocom- polymer matrix nanocomposite samples is the good disper-
posites. sion of the nanoparticles in a polymer matrix. Rong et al. [12]
grafted monomers made of styrene to surround the particles
to produce better dispersion. They used isotactic polypropy-
2. Experimental procedures lene as the matrix and SiO2 as inclusions which were ap-
proximately 7 nm in size. The particles were heated first to
Making good samples of polymer matrix nanocompos- remove any water absorbed on the surface. Then, they were
ites is a challenging area that draws considerable effort. Re- mixed with one of the monomers and solvent. This mixture
searchers have tried a variety of processing techniques [9–16] was then irradiated and the solvent was removed. Samples
to make polymer matrix nanocomposites. These include melt were then made by adding polypropylene, tumble mixing,
mixing, in situ polymerization, and other approaches. Creat- compounding and extruding. This technique produced sam-
ing one universal technique for making polymer nanocom- ples without aggregation and, in addition, greatly increased
posites is difficult due to the physical and chemical differ- the particle–polymer matrix interfacial interaction.
ences between each system and various types of equipment In situ polymerization is another technique that has been
available to researchers. Each polymer system requires a spe- used to make polymer matrix nanocomposites [13–15,18,19].
cial set of processing conditions to be formed, based on the It is a method in which particles are dispersed first in
processing efficiency and desired product properties. The dif- monomer and then the mixture is polymerized. In Yang et
ferent processing techniques in general do not yield equiva- al. [13], nanocomposites with polyamide-6 matrix and silica
lent results [17]. inclusions were prepared by first drying the particles to re-
Vollenberg and Heikens [9] were able to produce good move any water absorbed on the surface. Then, the particles
nanocomposite samples by thoroughly mixing filler particles were mixed with ␧-caproamide and concurrently a suitable
with polymer matrix. The polymer matrices used in these polymerization initiator was added. The mixture was then
experiments were polystyrene (PS), styrene–acrylonitrile polymerized at a high temperature under nitrogen [13]. This
copolymer (SAN), polycarbonate (PC) and polypropylene technique produced well-dispersed samples when the inclu-
(PP). The inclusions were alumina beads 35 nm and 400 nm sions were around 50 nm in size, but aggregation occurred for
in size and glass beads 4, 30, or 100 ␮m in diameter. The smaller particles around 12 nm in size [15]. This was most
J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11 3

likely due to the increased surface energy for the smaller par- the challenges in processing of nanocomposites, lack of sys-
ticles, which, in the absence of a stabilizer, favored further tematic experimental results, and scarcity of theoretical stud-
particle segregation. ies. Moreover, some material properties have been studied
Ash et al. [20,21] and Siegel et al. [6] used yet another more in-depth than other, leaving gaps in the knowledge on
technique and produced nanocomposites with polymethyl- nanocomposite behavior. The following sections will outline
methacrylate (PMMA) matrix and alumina inclusions, which some of the experimental results that are available to-date and
were well dispersed throughout the matrix. The particles identify the trends that can be obtained from these results. The
were added to methylmethacrylate (MMA) monomer and focus of this summary will be on particles and matrix mor-
dispersed through sonication in the low-viscosity solution. phology (degree of crystallinity, particle size and arrange-
An initiator and a chain transfer agent were added later. The ment) and the mechanical behavior of the nanocomposite
mixture was then polymerized under nitrogen, broken into materials (elastic modulus, strength, yield stress, strain-to-
smaller pieces, and dried in a vacuum. Tensile specimens failure, fracture toughness, and viscoelastic properties).
were made through compression molding and cooled to room
temperature under pressure. 3.1. Mechanical properties
Li et al. [16] used a unique approach to prepare high-
density polyethylene (HDPE)–polypropylene nanocompos- Vollenberg and Heikens [9] performed a wide range of
ites. Seventy-five percent weight HDPE and 25% polypropy- tests on composites with matrices made of polystyrene (PS),
lene were melt-mixed and extruded into tapes. These tapes styrene–acrylonitrile copolymer (SAN), polycarbonate (PC)
were then cut into shorter pieces and melt-processed by and polypropylene (PP) with micro- as well as nano-sized
either extrusion or compression molding. This produced a glass or alumina inclusions. The PS and PP were observed to
nanocomposite with HDPE as the matrix and polypropylene have a very poor interaction with the particles due to their non-
fibrils ranging from 30 to 150 nm in diameter. polar character while the interfacial interaction was much
For clay nanocomposites, the specific choice of processing more noticeable for the SAN and PC. For composites with
steps depends on the final morphology required in the com- PS, PC and PP matrix and glass or alumina particles the mod-
posite, i.e., exfoliated or intercalated form. In the intercalated ulus increased with decreasing size of particles. In order to
form, matrix polymer molecules are introduced between the investigate the effect of interface adhesion separate experi-
ordered layers of clay resulting in an increase in the inter- ments were done on SAN and PC matrix composites in which
layer spacing, but still maintaining the order. On the other particles were pretreated with siloxane DF 1040 (General
hand, in an exfoliated form, clay layers are separated and are Electric Plastics) to decrease the filler–matrix adhesion. No
distributed within the matrix. Intercalated nanocomposites size effect on elastic modulus was found for composites with
are generally formed by melt blending or by in situ poly- SAN and PC matrix with either excellent adhesion or very
merization. Exfoliation ability depends on the nature of clay, poor adhesion due to treated particles. However, these exper-
blending process, and the agents used for curing. The final iments were done for three different particle sizes which were
structure of clay composite has a wide range of variations, in micron range only, not nanosize range. For all systems the
depending on the degree of intercalation and exfoliation. A elastic modulus increased with increasing volume fraction of
number of experimental techniques were used to character- the inclusions. It is necessary to point out that in these ex-
ize intercalated or exfoliated nanocomposites, such as X-ray periments, while there were several different sizes within the
diffraction, transmission electron microscopy (TEM), DSC, micron range, only one particle size can be considered in the
and thermogravimetric analysis (TGA). nanosize regime by the definition provided earlier.
Thus, in general, there is no single procedure that is fol- Similar conclusions were obtained by Chan et al. [11] for
lowed for making polymer matrix nanocomposites. A com- composites with polypropylene (PP) matrix and calcium car-
bination of melt mixing, extrusion, and compression molding bonate (CaCO3 ) nanoparticles. In their system the CaCO3 in-
seems to be the method of choice for many researchers. As clusions had an average size of 44 nm and a strong interaction
seen from the above mentioned summary the most important with the polymer matrix. The addition of CaCO3 nanoparti-
factor to consider in deciding between different processing cles to a PP matrix produced an increase in the elastic mod-
techniques is the requirement of good particle dispersion in ulus compared to the pure matrix as shown in Fig. 1. The
the polymer matrix. This dispersion will play a major role in increase in modulus coincided with an increase in nanopar-
the results reviewed below. ticle volume fraction. The reverse effect was found for the
yield stress and the tensile strength of the composites; both
of these quantities were highest for the pure polypropylene
3. Experimental results and decreased as the volume fraction of CaCO3 increased.
The strain-to-failure did not change much between the vari-
While there has been much experimental work done in ous systems.
the area of polymer matrix nanocomposites, there is yet to Clay-reinforced nanocomposites have received consid-
be a consensus on how nano-sized inclusions affect material erable attention in recent years (more than 100 articles
properties. This is partly due to the novelty of the area, and have been published in the literature on clay composites
4 J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11

the higher degree of intercalation in the structure resulted


in higher tensile strength and other mechanical properties.
In another work [28] on polystyrene–clay nanocomposites,
prepared by free-radical polymerization, the authors found
that exfoliated nanocomposites had better thermal stabil-
ity and better mechanical properties than pure polystyrene.
With clay–polyurethane nanocomposites, Tortora et al. [29]
showed that exfoliation occurred for low montmorillonite
content but when the clay content was increased more interca-
lation was observed associated with higher elastic modulus
and yield strength. However, breaking stress and breaking
strain decreased with increasing clay content.
Fig. 1. Polypropylene with CaCO3 (from [11]). Masenelli-Varlot et. al [30] studied the effects of
polyamide-6 with intercalated and exfoliated montmoril-
in the past three years). A number of polymers, such as lonite. The intercalated samples ranged from 1.7 to 2.7% clay
PC, PAN, PP, etc. were used as the matrix. Shelley et al. while the exfoliated samples ranged from 2.4 to 4.2%. Under
[22] examined a polyamide-6 system with clay platelets. The low-deformation tests, Young’s modulus increased linearly
platelets constituted 2% and 5% weight fraction and were with filler content with no major difference between interca-
1 nm × 10 nm × 10 nm in size [23]. Good interaction was lated and exfoliated. Under compression tests, the exfoliated
found between the matrix and inclusions. With this setup, the samples had one direction that was higher than the rest while
elastic modulus was found to improve for both the 2% and 5% the intercalated samples did not. The intercalated samples are
samples. For the smaller weight fraction (2%), the increase likely weaker due to the weaker mechanical coupling. The
in effective elastic modulus was 40% over the modulus of ultimate stress and strain-to-failure were lower for all filler
the pure polymer system. The larger weight fraction (5%) concentrations than pure polyamide-6. With the same level of
improved the effective modulus by a factor of two as com- dispersion and filler content, the exfoliated samples exhibited
pared to that of pure polymer. These results were for tensile greater stress at failure than the intercalated samples while
specimens cut in both longitudinal and transverse directions. the intercalated samples had a greater strain-to-failure [30].
In addition, the yield stress also improved for both weight Few micromechanical models were developed which are
fractions, with the greatest improvement found for the higher in agreement with the behavior of some nanocomposites
concentration of inclusions. The other quantity studied was [31–35]. A recent theoretical study [34] shows that when the
the strain-to-failure. The 2% system was found to give higher structure changes from intercalated to an exfoliated struc-
strain-to-failure than the pure system in the longitudinal di- ture, the morphological change is accompanied by a mod-
rection but close to that in the pure system in the transverse erate change in modulus rather than an abrupt change. This
direction. The higher filler content resulted in a decline in theoretical prediction was supported by experimental results
strain-to-failure from the pure system in both directions [22]. in [36]. There is a general agreement in the literature that
Properties of nanocomposites are highly related to theirs exfoliated systems have better mechanical properties, par-
microstructure [24]. For epoxy–silicate clay nanocomposites, ticularly higher modulus, than intercalated nanocomposites
Luo and Daniel [24] found that the ideal case for maximiz- [37]. Also, studies on nylon 6–clay nanocomposites showed
ing stiffness and thermal properties is through full exfolia- that morphology or physical properties are not significantly
tion and dispersion, which is not achieved usually, but often reduced from reactions or polymer molecular degradation
one obtains partial exfoliation and intercalation. Similar re- during the processing stage [38].
sults were presented by Mark and co-workers [25] for clay Polyamide-6 nanocomposites with silica inclusions have
reinforced natural and epoxidized rubber. For sodium mont- been examined (e.g. [15]). In contrast to the polyamide–clay
morillonite silicate clay the degree of its dispersion deter- system described in [22], the nanoparticles were 17, 30, or
mined its reinforcing effect in rubber. Zhang et al. [26] further 80 nm in diameter. The elastic modulus was slightly higher
demonstrated the similar behavior for clay–polypropylene for the nanocomposites than for the pure system, but there was
nanocomposites. They also showed that the reinforcing ef- little difference between composites with different nanopar-
fects demonstrated by storage modulus, thermal stability, ticle sizes. This can be seen in Fig. 2 where S refers to the
etc., are directly related to the dispersion of clay particles smallest particle size and L to the largest particle size. As in
in the matrix. Park et al. [27] fabricated clay nanocomposites the polyamide–clay system, the yield stress increased with
where organophilic clay particles were embedded in a syn- increasing filler content and also increased slightly with the
diotactic polystyrene with melt intercalation. They found that decreased size of the particles. The strain-to-failure followed
stepwise mixing method (where styrene polymer is blended the opposite pattern as it decreased greatly for an increase
with the clay first to enable intercalation followed by blend- in volume fraction and decrease in particle size. In addition,
ing with polystyrene matrix) resulted in a higher degree of there existed a sharp decrease in the stress immediately be-
intercalation (than a simultaneous mixing method) and that fore failure in all systems [15]. A separate study involving
J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11 5

Fig. 3. Stress strain curve for PMMA and alumina (from [20]).

sion of particles, namely the change from the uniform disper-


sion to clustering resulting in microcomposite agglomeration
[12]. In another study on palladium–PMMA nanocomposites
showed that addition of Pd in PMMA increased the oxidative
Fig. 2. Polyamide-6 with small (S) and large (L) inclusions (from [15]). thermal stability of the system. Improvement in thermal sta-
bility was linearly related to the volume fraction of palladium
these same materials examined the influence of a silane pre- inclusions [35], although the elastic modulus decreased as a
treatment, which improved the matrix–filler interface. For the function of the Pd content.
treated system, there was an improvement in strength as well Very interesting results were obtained by Ash et al.
as toughness, which was in contrast with the untreated system [20] who performed a series of tests on PMMA–alumina
in which strength improved as toughness decreased [19]. nanocomposites. The alumina nanoparticles were around
Another study centered on polyurethane–silica compos- 40 nm and had very little interfacial interaction with the poly-
ites [10]. Silica nanoparticles have relatively strong interac- mer matrix. As filler content increased, there was a sharp ini-
tion with polyurethane matrix. Polyurethane–silica compos- tial drop in Young’s modulus followed by a steady increase.
ites were formed with inclusions of 12 nm and 1.4 ␮m at Even at the highest filler content, however, the effective elas-
10–50% weight fraction of inclusions. The tensile strength tic modulus was lower than the pure system. In addition, the
varied little between the micron and nano-sized particles, up yield stress and tensile strength of the pure matrix was higher
to a filler composition of 20% weight fraction. Above 20%, than for the composite as shown in Fig. 3. These results for the
the tensile strength was higher for the nanocomposite as com- strength are very different from the other systems that have
pared to pure polyurethane polymer. In the nanocomposite, been examined. In addition, the strain-to-failure increased by
the strain-to-failure increased by 500% over the pure system around 800% over that for the pure system.
while for composites with larger inclusions the increase was Elastic properties were found to increase for a nanocom-
only by 100% [10]. posite system made 75 wt% of high-density polyethelyne
A unique set of experiments was performed by Rong et (HDPE) matrix and 25 wt% polypropylene (PP) [16]. During
al. [12] with SiO2 nanoparticles in a polypropylene matrix processing the compound was fibrillated and PP nanofibers
with and without the addition of grafting polymers (either PS were created. The fibers were cylindrical in shape, 30–150 nm
or PMMA), which have the ability to improve dispersion of in diameter. The elastic modulus, yield stress, and tensile
nanoparticles and interfacial interactions between the parti- strength were greater for the nanocomposite system than for
cles and the matrix. The elastic modulus increased with in- samples of pure HDPE [16]. However, the strain-to-failure
creasing volume fraction in both systems but was much higher was lower for nanocomposite systems.
in the untreated composite systems (without grafting poly- From the above discussion, it is possible to extract a few
mers). The modulus for the composites treated with grafting trends for the behavior of polymer matrix nanocomposites
polymer only improved minimally as a function of volume based on the nature of the polymer matrix, particularly crys-
fraction of SiO2 particles. The opposite effect was found for talline or amorphous nature of the polymer, and the inter-
the tensile strength where the treated composites showed in- action between the filler and matrix. The elastic modulus
creased strength while the strength changed very little in the tends to increase with the volume fraction of inclusions in
untreated composites as a function of volume fraction of par- every case. In some systems, there is a critical volume frac-
ticles. Interestingly, the results for strength were insensitive tion at which aggregation occurs and the modulus goes down
to change in volume fraction of particles. The strain-to-failure [1–4,8]. In general, there is also an increase in modulus as
was lower in the untreated systems than the pure polypropy- the size of the particle decreases. Interaction between ma-
lene samples. In the treated system the strain-to-failure in- trix and filler may play an important role in the effects of
creased for volume fraction of particles up to 2% and then the nanoparticles [1–3] on composite properties. For polymer
started decreasing. This change in behavior at approximately systems capable of having a higher degree of crystallinity, the
2% volume fraction was attributed to the change in disper- increase in modulus with decreasing particle size is found to
6 J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11

be greater in systems with poor interaction between filler and weight fraction of the filler is as large as 40 or 50%. However,
matrix as opposed to those with good interaction. However, if the weight fraction is below this level, the microcomposite
the overall trend of the modulus of polymer nanocomposites has the higher modulus [10].
is not found to be greatly dependent upon the nature of the In general, viscoelastic properties tend to be higher in
matrix nor the interaction between filler and matrix. nanocomposites than in pure polymer systems. When there is
An examination of the yield stress gives a different trend good filler–matrix interaction, the storage modulus generally
than that of the elastic modulus. For composites with good in- increases with increasing volume fraction. The modulus also
teraction between filler and matrix, the yield stress tends to in- seems to increase as the particle size decreases. However,
crease with increasing volume fraction and decreasing parti- there is little experimental work in this area for composites
cle size, similarly to the increase in modulus under same con- with poor filler–matrix interactions. Overall, the storage mod-
ditions. The pattern changes when there is poor interaction ulus tends to increase with the presence of nanoparticles in a
between the matrix and particles. The addition of nanopar- composite system.
ticles with poor interaction with the matrix causes the yield Morphological details, such as exfoliation, intercalation,
stress to decrease, compared to the neat matrix, regardless or cross-linked matrix versus uncross-linked matrix, have a
of the filler concentration or size. The ultimate stress fol- significant effect on the viscoelastic properties of nanocom-
lows a similar pattern as that observed for the yield stress. posites. In epoxy–clay nanocomposite system forces from
It generally increases in polymer systems (both crystalline cross-linked epoxy molecules cause intercalated clay gal-
and amorphous) with good filler–matrix interaction and it leries to exfoliate, associated with a gradual increase in
increases in general as particle size decreases. There is no viscosity and a relatively fast increase in storage modulus
uniform trend with respect to the volume fraction of particles [40]. In maleated polypropylene–layered silicate nanocom-
for the ultimate stress. A poor filler–matrix interaction leads posites, exfoliated nanocomposite systems were reported to
to a decrease in the ultimate and yield stress as compared to have higher shear and complex viscosities, while in a shear
the pure matrix system. flow exhibiting the largest drop in complex viscosity due to
The change in strain-to-failure behavior for nanocompos- alignment of clay layers [41]. Dynamic storage moduli also
ites, like the yield and ultimate stress, is different depending show a similar behavior [41]. Experiments with clay–nylon-
on the system. In this case, however, the change is with re- 12 nanocomposite systems have shown that melt processing
spect to the nature of the polymer matrix as opposed to the under low shear improves intercalation, while larger shear
filler–matrix interaction. In general, the addition of nanopar- enhances exfoliation associated with a decrease in melt vis-
ticles to a mostly crystalline or semi-crystalline polymer, re- cosity [42].
gardless of the filler–matrix interaction, reduces the maxi-
mum strain. The opposite trend is found in amorphous poly-
mers with the increase in strain-to-failure coinciding with a 3.3. Crystallinity
decrease in particle size.
The degree of crystallinity of a polypropylene system
3.2. Viscoelastic properties was found to change very little with the addition of CaCO3
nanoparticles. On the other hand, the peak crystallization
Viscoelastic properties were found to improve with the ad- temperature was increased by around 10 ◦ C with the addition
dition of particles in a polyamide-6 matrix with clay platelet of the CaCO3 inclusions. This increase was found to be nearly
inclusions. More specifically, the storage modulus was found the same for all filler volume fractions. Moreover, the size of
to be higher for the composite systems than for the pure the crystalline domain spherulites was found to change dra-
polyamide-6 system. The increase in modulus coincided with matically in different systems. Scanning electron microscopy
the increase in filler content [22,30]. However, experimental (SEM) showed that the spherulites in the pure system had an
results also showed that the onset of the decrease in mod- average size of around 40 ␮m, but no spherulites could be
ulus occurred at lower temperatures as the filler content in- seen when CaCO3 nanoparticles were added to the system
creased [22]. The storage modulus was also found to increase [11]. It is possible that the spherulites were too small and
with increasing volume fraction for polyamide-6 nanocom- they were not detected by SEM.
posite with silica nanoparticles rather than clay platelets The addition of nano-sized montmorillonite clay platelets
[15]. A similar increase in storage modulus was found for to a polyamide-6 matrix was found to have no effect on the de-
poly(vinylidene fluoride)–clay nanocomposites [39]. gree of crystallinity of the polymer with weight fraction up to
Viscoelastic properties were also found to improve in other 5%. However, the nanoparticles did have an effect on the size
systems, including polyurethane–silica composites. The stor- of the crystallites. The size of crystallites in the nanocom-
age modulus increased for such composites with both micron posites was nearly an order of magnitude smaller than the
and nano-sized particles. This modulus also increased as the size of spherulites in the pure matrix system [22]. Results
weight fraction of the particles was increased to 50%. It is for a polyamide-6 matrix composite with silica nanoparticles
also noteworthy to mention that in the rubbery state of the sys- showed that neither the size nor the filler content had any
tem, the modulus is highest for the nanocomposite when the effect on the crystallinity of the system [15].
J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11 7

Fig. 4. Young’s modulus and glass transition temperature vs. filler weight fraction (from [20,21]).

In a separate experiment with clay–polyamide nanocom- polymers is not affected very much by the addition of
posites, the glass transition temperature of the system in- nanoparticles. There may be some changes in particular
creased with increasing clay nanoparticle content. This in- nanocomposite systems, but overall no major differences
crease was attributed to either the effect of the clay layers in crystallinity of nanocomposites versus neat polymers
retarding the polyamide molecule main-chain motion or a were observed in any of the systems examined. On the other
slightly higher molecular weight of the polymer due to the hand, the glass transition temperature was influenced by
reactive organoclay used [43]. Another study, however, found the addition of particles. When there is good filler–particle
a 10 ◦ C drop in the glass transition temperature for all clay interaction, the glass transition temperature tends to increase
filler contents. This drop was attributed to the presence of with a decrease in the size of particles for amorphous poly-
surfactants [30]. mers. For crystalline polymers, the transition temperature
Ash et al. [21] studied the glass transition temperature of decreases with an increase in particle concentration. For an
PMMA–alumina nanocomposites. They found that for filler amorphous system with poor filler–polymer interfacial in-
weight fractions less than 0.5%, the glass transition tempera- teraction, the glass transition temperature decreased overall.
ture did not change, but for filler weight fraction above 0.5% Thus, while the degree of crystallinity is not significantly
the glass transition temperature dropped by more than 20 ◦ C affected by the presence of particles, the glass transition
and stabilized at a new temperature at weight fractions above temperature is very dependent upon this factor.
1% as shown in Fig. 4b. This drop occurred at the same weight
fraction as the drop in the elastic modulus, as shown in Fig. 4a, 3.4. Density/volume change
and as mentioned in previous discussion about this system.
A different result was found for the glass transition tem- It is natural for the density of a composite to be higher than
perature of a polyurethane matrix with silica inclusions. In the density for a pure polymer system. The inclusions typi-
this nanocomposite system, the glass transition temperature cally have much higher density than the polymer leading to
was found to increase as filler concentration increased. Dif- a higher density for the composite. This was found to be true
ferent techniques produced differing results but the consen- in a polyurethane matrix with micron and nano-sized silica
sus was that the glass transition temperature increased by inclusions. The density increased, as expected, as the filler
around 10 ◦ C for the nanocomposite above the neat resin as content increased. However, the density of the composites
the weight fraction of the filler was increased to 50%. The with the micron-sized inclusions was higher than the density
transition temperature also increased with the use of micron- for the nano-sized inclusions [10].
sized inclusions, but the increase was not as great as that for The volume change of a composite undergoing tensile
the nanocomposite [10]. elongation was examined by Reynaud et al. [15] using a
The pattern that the crystallinity was little affected polyamide-6 composite with silica inclusions of various sizes
by the addition of particles was also observed in a within the nanometer range. Reynaud et al. [15] reported a
polypropylene–SiO2 composite that was treated with graft- volume increase for the pure polyamide-6 but larger increases
ing polymers. The particles were found to have a small nu- for the composite systems. The volume change was highest
cleation effect on the polypropylene matrix, but, overall, the for composites containing the smallest particles, which pro-
crystallinity was not greatly influenced by the addition of the vides support for their debonding model that will be discussed
nanoparticles [12]. later in this article.
Park et al. [37] found that for (syndiotactic) Investigations examining the change in the density or vol-
polystyrene–organoclay nanocomposites, dispersed clay ume of polymer nanocomposites are limited. Nevertheless,
layers act as nucleating agent competing with crystal growth the trend is that nano-sized inclusions affect the volume of
in polymers. Thus, exfoliated clay nanocomposites have a composite more than micron-sized inclusions do. The vol-
lower degree of crystallinity with faster crystallization rate. ume for the composite is higher for a nanocomposite than
However, the crystallinity of crystalline and semi-crystalline a microcomposite with the same filler weight and the same
8 J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11

polymer weight. It is important to point out that these results An increase in yield and tensile strength and modulus in
are only for composites with relatively good filler–polymer nanocomposite systems as compared to microcomposites can
interaction. be partially explained on the basis of the interaction between
the filler and the matrix. It has been found that a greater adhe-
sion between the matrix and inclusion causes less debonding
4. Discussion when a stress is applied and, consequently the elastic modu-
lus and strength are improved [45]. Vollenberg and Heikens
From the experimental results presented above, we ob- [9] explained that if there is a strong interaction between the
serve that different composite systems can lead to very dif- polymer and the particle, the polymer layer in the immediate
ferent results. One important observation is that composites proximity of the particle will have a higher density. For most
with nano-sized inclusions generally have different proper- systems, density is proportional to elastic modulus, so the
ties than composites with larger scale inclusions. The specific region directly surrounding the inclusions will be a region of
reasons why the polymer matrix composites with nano-sized high modulus. The polymer right outside this high modulus
reinforcement have different properties than composites with region will have a lower density due to the polymer chains
micron-sized reinforcement are not fully understood, but sev- that are moved towards the particle. For large particles, the
eral theories have been introduced to explain some of the size of the low density region will be relatively large, and the
changes in material morphology and behavior that are seen contribution of the high modulus filler will be diminished.
at the nano-scale. It is important to point out, however, that For nanoparticles, the number of particles for a given volume
most of these theories were developed to explain particular fraction is much larger, thus the particles will be much closer
results and, therefore, are not necessarily applicable to a large to one another. If the particles are densely packed, then the
number of polymer nanocomposites. boundary layer of polymer at the interface will comprise a
Chan et al. [11] proposed that properties such as elas- large percentage of the matrix and can create a system where
tic modulus, tensile strength, and yield strength decrease there is no space for a low modulus region to form. This re-
in nanocomposites with polypropylene matrix due to the sults in the elastic moduli of composites with smaller particle
change in nucleation caused by the nanoparticles (Fig. 5). The size (nano) being greater than the moduli of composites with
nanoparticles produce a much larger number of nucleating larger inclusions [9,10].
sites but, in turn, greatly reduce the size of these spherulites. The small interparticle distance in nanocomposites was
In their experimental work, no spherulites were found in the used as another parameter to explain the changes in the elas-
nanocomposites by SEM indicating that either none were tic modulus and strength of these materials when compared
present or they were reduced to such a small size that SEM with the composites with micron-sized particles. The same
could not detect them. It was further proposed that there was parameter also plays a role in the glass transition tempera-
another mechanism which was causing these same properties ture changes observed in nanocomposites versus composites
to increase. The increase occurred when there was a strong with micron-sized reinforcement. Ash et al. [21] found that
interaction between the polymer and filler. This interaction for their system the glass transition temperature was constant
had larger impact in nanocomposites due to the large inter- until around 0.5% weight fraction of particles, then had a
facial area between the filler particles and the matrix. Other sharp drop, and then it remained constant for weight frac-
investigators have suggested that this interaction leads to a tions above 1%. When there is little or no interfacial inter-
layer of polymer that is directly adsorbed and bound to the action between the filler and matrix and the interparticle dis-
particles [9,10,20,21]. Experimental work that has been per- tance is small enough, the polymer between two particles
formed on a polystyrene–cobalt nanocomposite with cobalt acts as a thin film. For a thin film, the glass transition tem-
nanoparticles with an average size of 21 nm has shown that perature decreases as film thickness decreases. The distance
the polymer layer was about 24 nm, and varied non-linearly between particles in a composite with the filler weight frac-
with molecular weight [44]. tion below 0.5% is relatively large, and hence, in this case the

Fig. 5. (a) Pure polypropylene and (b) polypropylene with 9.2% volume filler (from [11]).
J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11 9

perature increases, so the drop in glass transition temperature


is correlated with the drop in modulus [20,21].
Reynaud et al. [15] found that during tensile testing, the
volume of polymer nanocomposites increased, with the great-
est increase occurring in systems with the smallest particles.
To explain this, the debonding process of the polymer next
to the inclusions was examined, as shown in Fig. 6. It was
proposed that the smallest particles tend to aggregate and
debonding occurs around each individual particle. As a re-
sult, the large clusters of small particles act as larger soft
particles. On the other hand, the larger filler particles do not
aggregate and each particle undergoes a single debonding
process.
Due to the different results obtained and the different na-
ture of the various polymer nanocomposite systems, there
Fig. 6. Debonding around 50 and 12 nm particles (from [15]). is no observed universal trend that can be modeled and ex-
plained. There are, however, observations that show the be-
havior of nanocomposites different from composites with
polymer between each particle is not considered to belong to larger scale inclusions. The particle size and the polymer
the thin film regime. As the filler concentration increases, and particle morphology tend to play a very important role.
the interparticle distance and the resulting thickness of the In addition, the nature of dispersion and aggregation of par-
film, decrease. This theory, however, does not explain why ticles can affect the properties of composites significantly.
the glass transition temperature levels off rather than contin- Filler–matrix interaction is another factor that influences the
ues to drop as a function of increasing weight fraction of the properties. The strength of the interaction plays a role in the
filler. A drop-off in Young’s modulus was found for the same thickness and density of the interphase, which consists of a
filler weight fraction as the drop in glass transition tempera- layer of high density polymer around the particle. The ef-
ture. It was proposed that as the glass transition temperature fects of the interface on the behavior of a composite depend
decreased, the relative testing temperature increased. Also, upon the interparticle distance. For constant filler content,
the elastic modulus of the matrix, PMMA, decreases as tem- with reduction in particle size, number of filler particles in-

Table 1
Summary of polymer nanocomposite trends
Crystalline Amorphous
Elastic modulus Increase w/volume fraction Increase w/volume fraction Good interaction
Increase or no change with decrease of size Increase w/decrease size
Increase w/volume fraction Increase w/volume fraction Poor interaction
Increase w/decrease size Increase w/decrease size
Greater increase than for good interaction
Yield stress/strain Increase w/volume fraction N/A Good Interaction
Increase w/decrease size
Decrease with addition of particles Decrease with addition of particles Poor interaction
Ultimate stress/strain Increase w/decrease size Nano>micro after 20%weight Good interaction
No unified result for change inVf
Lower than pure for small volume fractions Decrease with addition of particles Poor interaction
Density/volume Increased volume as size decreases Increased volume as size decreases Good interaction
N/A N/A Poor interaction
Strain-to-failure Decrease with addition of particles Increase with addition of particles Good interaction
Increase w/decrease size
Decrease with addition of particles Increase with addition of particles Poor interaction
Tg Decrease with addition of particles Increase w/decrease size Good interaction
N/A Level until 0.5%, drops off level from 1–10% Poor interaction
Crystallinity No major effect N/A Good interaction
No major effect N/A Poor interaction
Viscoelastic Increase w/volume fraction Increase w/volume fraction nano less regular Good interaction
Increase w/decrease size
N/A Decrease with addition of particles—drop at 1% with rise following Poor interaction
10 J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11

creases, bringing the particles closer to one another. Thus, the [6] R.W. Siegel, S.K. Chang, B.J. Ash, J. Stone, P.M. Ajayan, R.W.
interface layers from adjacent particles overlap, altering the Doremus, L.S. Schadler, Scr. Mater. 44 (2001) 2061–2064.
bulk properties significantly. These issues play a major role [7] C.L. Wu, M.Q. Zhang, M.Z. Rong, K. Friedrich, Compos. Sci. Tech-
nol. 62 (2002) 1327–1340.
in the effect of nano-sized inclusions in a polymer matrix. [8] R. Magaraphan, W. Lilayuthalert, A. Sirivat, J.W. Schwank, Compos.
For nanoparticles, any configuration changes in the matrix Sci. Tech. 61 (2001) 1253–1264.
will have a significant effect when the characteristic radius of [9] P.H.T. Vollenberg, D. Heikens, Polymer 30 (1989) 1656–1662.
polymer chains is of the same order as the inclusions [45,46]. [10] Z.S. Pétrovic, I. Javni, A. Waddon, G. Bánhegyi, J. Appl. Polym.
There are other areas addressed in literature. For exam- Sci. 76 (2000) 133–151.
[11] C.-M. Chan, J. Wu, J.-X. Li, Y.-K. Cheung, Polymer 43 (2002)
ple, López et al. [47] examined the processing and thermal 2981–2992.
and mechanical properties of magnetic nanocomposites. In [12] M.Z. Rong, M.Q. Zhang, Y.X. Zheng, H.M. Zeng, R. Walter, K.
another work the mechanical properties of clay nanocompos- Friedrich, Polymer 42 (2001) 167–183.
ites were analyzed as a function of filler loading and orien- [13] F. Yang, Y. Ou, Z. Yu, J. Appl. Polym. Sci. 69 (1998) 355–361.
tation [30]. Zhang et al. [48] provided a look at matrix–filler [14] Y. Ou, F. Yang, Z.-Z. Yu, J. Polym. Sci. B: Polym. Phys. 36 (1998)
789–795.
interfacial properties. Effect of matrix on the polymer matrix [15] E. Reynaud, T. Jouen, C. Gautheir, G. Vigier, Polymer 42 (2001)
composites were examined by Friedlander et al. [49]. 8759–8768.
[16] J.-X. Li, J. Wu, C.M. Chan, Polymer 41 (2000) 6935–6937.
[17] C.I. Park, O.O. Park, J.G. Lim, H.J. Kim, Polymer 42 (2001)
7465–7475.
5. Conclusions [18] C. Zeng, L.J. Lee, Macromolecules 34 (2001) 4098–4103.
[19] Y. Li, J. Yu, Z.-X. Guo, J. Appl. Polym. Sci. 84 (2002) 827–834.
Currently, a significant amount of work is being published [20] B.J. Ash, J. Stone, D.F. Rogers, L.S. Schadler, R.W. Siegel, B.C.
on polymer nanocomposites. In an attempt to further under- Benicewicz, T. Apple, Mater. Res. Soc. Symp. Proc. 661 (2000).
[21] B.J. Ash, L.S. Schadler, R.W. Siegel, Mater. Lett. 55 (2002) 83–87.
stand the synthesis, processing, and properties analysis of
[22] J.S. Shelley, P.T. Mather, K.L. DeVries, Polymer 42 (2001)
polymer nanocomposites, a selection of representative and 5849–5858.
recent literature was chosen to highlight some of the issues re- [23] A. Okada, T. Usuki, O. Kurauchi, Kamigaito, in: J.E. Mark, C.Y.-C.
lated to the preparation and mechanical behavior of compos- Lee, P.A. Branconi (Eds.), ACS Symposium, 1995, pp. 55–65.
ites with nano-sized reinforcement in comparison with com- [24] J.-J. Luo, I.M. Daniel, Compos. Sci. Technol. 63 (2003) 1607–1616.
[25] Y.T. Vu, J.E. Mark, L.H. Pham, M. Engelhardt, J. Appl. Polym. Sci.
posites with larger micron-sized inclusions. As was discussed
82 (2001) 1392–1403.
above, some trends are observed but no universal patterns for [26] Y.-Q. Zhang, L.-H. Lee, H.-J. Jang, C.-W. Nah, Compos. Part B:
the behavior of polymer nanocomposites can be deduced in Eng. 35 (2004) 133–138.
general. These observed trends are summarized in Table 1. In [27] C.I.L. Park, O.O. Park, J.G. Lim, H.J. Kim, Polymer 42 (2001)
general, the material properties of polymer nanocomposites 7465–7475.
[28] C.-R. Tseng, J.-Y. Wu, H.-Y. Lee, F.-C. Chang, J. Appl. Polym. Sci.
are superior to the pure polymer matrix or composites with
85 (2002) 1370–1377.
larger sized inclusions. The effects of the nanoparticles are [29] M. Tortora, G. Gorrasi, V. Vittoria, G. Galli, S. Ritrovati, E. Chiellini,
dependent on many variables but especially upon the relative Polymer 43 (2002) 6147–6157.
crystalline or amorphous nature of the polymer matrix as well [30] K. Masenelli-Varlot, E. Reynaud, G. Vigier, J. Varlet, J. Polym. Sci.
as the interaction between the filler and matrix. B: Polym. Phys. 40 (2002) 272–283.
[31] P. Uribe-Arocha, C. Mehler, J.E. Puskas, V. Altstat, Polymer 44
(2003) 2441–2446.
[32] T. Tzianetopoulou, M.C. Boyce, Materials Research Society Sym-
Acknowledgments posium Proceedings, vol. 788, Continuous Nanophase Nanostruct.
Mater. (2003) 601–607.
[33] I.M. Daniel, H. Miyagawa, E.E. Gdoutos, J.J. Luo, Exp. Mech. 43
I.J. acknowledges the support by the National Science (2003) 348–354.
Foundation (Grant CMS-0085137) and the Air Force Office [34] N. Sheng, M.C. Boyce, D.M. Parks, G.C. Rutledge, J.I. Abes, R.E.
of Scientific Research (Dr. Thomas Hahn, monitor). Cohen, Polymer 45 (2004) 487–506.
[35] C. Aymonier, D. Bortzmeyer, R. Thomann, R. Mulhaupt, Chem.
Mater. 15 (2003) 4874–4878.
[36] C.-R. Tseng, J.-Y. Wu, H.-Y. Lee, F.-C. Chang, J. Appl. Polym. Sci.
References 85 (2002) 1370–1377.
[37] C.I.L. Park, W.W. Choi, M.H.O. Kim, O.O.K. Park, J. Polym. Sci.
[1] H. Akita, T. Hattori, J. Polym. Sci. B: Polym. Phys. 37 (1999) B: Polym. Phys. 42 (2004) 1685–1693.
189–197. [38] T.D. Fornes, P.J. Yoon, D.R. Paul, Polymer 44 (2003) 7545–
[2] H. Akita, H. Kobayashi, J. Polym. Sci. B: Polym. Phys. 37 (1999) 7556.
209–218. [39] L. Priya, J.P. Jog, J. Polym. Sci. B: Polym. Phys. 40 (2002)
[3] H. Akita, H. Kobayashi, T. Hattori, K. Kagawa, J. Polym. Sci. B: 1682–1689.
Polym. Phys. 37 (1999) 199–207. [40] J.H. Park, S.C. Jana, Macromolecules 36 (2003) 2758–2768.
[4] J.-H. Chang, Y.U. An, J. Polym. Sci. B: Polym. Phys. 40 (2002) [41] C.M. Koo, M.J. Kim, H. Min, S.O. Kim, I.J. Chung, J. Appl. Polym.
670–677. Sci. 88 (2003) 1526–1535.
[5] S.A. Zavyalov, A.N. Pivkina, J. Schoonman, Solid State Ionics 147 [42] C.Y. Lew, W.R. Murphy, G.M. McNally, Dev. Chem. Eng. Miner.
(2002) 415–419. Proc. 12 (2004) 149–158.
J. Jordan et al. / Materials Science and Engineering A 393 (2005) 1–11 11

[43] H.-L. Tyan, K.-H. Wei, T.-E. Hsieh, J. Polym. Sci. B: Polym. Phys. [47] D. López, I. Cendoya, F. Torres, J. Tejada, C. Mijangos, J. Appl.
38 (2000) 2873–2878. Polym. Sci. 82 (2001) 3215–3222.
[44] E. Tadd, A. Zeno, M. Zubris, N. Dan, R. Tannenbaum, Macro- [48] Y. Zhang, S. Ge, B. Tang, M.H. Rafailovich, J.C. Sokolov, D.G.
molecules 36 (2003) 6497–6502. Peiffer, Z. Li, A.J. Dias, K.O. McElrath, S.K. Satija, M.Y. Lin, D.
[45] K.H. Wang, I.J. Chung, M.C. Jang, J.K. Keum, H.H. Song, Macro- Nguyen, Langmuir 17 (2001) 4437–4442.
molecules 35 (2002) 5529–5535. [49] S.K. Friedlander, K. Ogawa, M. Ullmann, J. Polym. Sci. B: Polym.
[46] A. Ranade, N.A. D’Souza, B. Gnade, Polymer 43 (2002) 3759–3766. Phys. 38 (2000) 2658–2665.

Das könnte Ihnen auch gefallen