scispace - formally typeset
Open AccessJournal ArticleDOI

Filler-filler interactions and viscoelastic behavior of polymer nanocomposites

TLDR
In this article, the main results obtained within a project on mechanical properties of polymer-based nanocomposites are presented, together with original experimental results and micro-mechanical modeling.
Abstract
This work presents the main results obtained within a project on mechanical properties of polymer based nanocomposites. The specific point was how to analyze and model the filler–filler interactions in the description of the viscoelastic behavior of these materials. This paper aims at presenting the general strategy used by the different partners to address this question, together with original experimental results and micro-mechanical modeling. Different nanocomposite materials were fabricated using the latex route, leading to random dispersions of rigid submicronic particles (PS = polystyrene, silica) in a flexible polybutylacrylate matrix at various volume fractions. In addition, encapsulated silica particles in a styrene–acrylate copolymer were produced, leading, after film formation, to a limited number of contacts between silica fillers. The processing route of these encapsulated particles was optimized and the resulting morphology was analyzed by TEM experiments. In the case of random mixtures, a strong effect of reinforcement appears in the rubbery field of the soft phase when the filler content is above a critical fraction (percolation threshold). The reinforcement in the rubbery plateau can be still exacerbated in the case of the PS particles if the material undergoes a heat treatment above the main relaxation of the PS phase. These experimental results illustrate the difference between geometrical percolation (when particles are just in contact) and mechanical percolation (with strong interactions between the fillers). The comparison of the results for PS and silica fillers shows once more that the strength of the interactions plays an important role. To account for the whole set of experimental data, two ways of modeling were explored: (i) homogenization methods based on generalized self-consistent schemes and (ii) a discrete model of spheres assembly which explicitly describes the ability of the contacts to transmit efforts.

read more

Content maybe subject to copyright    Report

HAL Id: hal-00111433
https://hal.archives-ouvertes.fr/hal-00111433
Submitted on 31 Oct 2018
HAL is a multi-disciplinary open access
archive for the deposit and dissemination of sci-
entic research documents, whether they are pub-
lished or not. The documents may come from
teaching and research institutions in France or
abroad, or from public or private research centers.
L’archive ouverte pluridisciplinaire HAL, est
destinée au dépôt et à la diusion de documents
scientiques de niveau recherche, publiés ou non,
émanant des établissements d’enseignement et de
recherche français ou étrangers, des laboratoires
publics ou privés.
Filler-ller interactions and viscoelastic behavior of
polymer nanocomposites
Emmanuelle Chabert, Michel Bornert, Elodie Bourgeat-Lami, Jean-Yves
Cavaillé, Remy Dendievel, Catherine Gauthier, Jean-Luc Putaux, André Zaoui
To cite this version:
Emmanuelle Chabert, Michel Bornert, Elodie Bourgeat-Lami, Jean-Yves Cavaillé, Remy Dendievel,
et al.. Filler-ller interactions and viscoelastic behavior of polymer nanocomposites. Materials Sci-
ence and Engineering: A, Elsevier, 2004, 381 (1-2), pp.320-330. �10.1016/j.msea.2004.04.064�. �hal-
00111433�

Filler–filler interactions and viscoelastic behavior
of polymer nanocomposites
E. Chabert
a,b,
, M. Bornert
b
, E. Bourgeat-Lami
c
, J.-Y. Cavaillé
a
,
R. Dendievel
d
, C. Gauthier
a
, J.L. Putaux
e
, A. Zaoui
b
a
GEMPPM, UMR 5510 INSA/CNRS, 69621 Villeurbanne cedex, France
b
LMS, UMR 7649 École Polytechnique/CNRS, 91128 Palaiseau cedex, France
c
LCPP, UMR 140 CPE/CNRS, 69616 Villeurbanne cedex, France
d
GPM2, UMR 5010 INPG/CNRS, BP 46, 38042 St Martin d’Hères cedex, France
e
CERMAV, UPR 5301 CNRS, BP 53, 38041 Grenoble cedex 9, France
This work presents the main results obtained within a project on
mechanical properties of polymer
based
nanocomposites. The
specific
point
was how to analyze and model the filler–filler interactions in the description of the viscoelastic behavior of these materials. This paper
aims at presenting the general strategy used by the different partners to address this question, together with original experimental results and
micro-mechanical modeling. Different nanocomposite materials were fabricated using the latex route, leading to random dispersions of rigid
submicronic particles (PS = polystyrene, silica) in a flexible polybutylacrylate matrix at various volume fractions. In addition, encapsulated
silica particles in a styrene–acrylate copolymer were produced, leading, after film formation, to a limited number of contacts between silica
fillers. The processing route of these encapsulated particles was optimized and the resulting morphology was analyzed by TEM experiments.
In the case of random mixtures, a strong effect of reinforcement appears in the rubbery field of the soft phase when the filler content is above a
critical fraction (percolation threshold). The reinforcement in the rubbery plateau can be still exacerbated in the case of the PS particles if the
material undergoes a heat treatment above the main relaxation of the PS phase. These experimental results illustrate the difference between
geometrical percolation (when particles are just in contact) and mechanical percolation (with strong interactions between the fillers). The
comparison of the results for PS and silica fillers shows once more that the strength of the interactions plays an important role. To account for
the whole set of experimental data, two ways of modeling were explored: (i) homogenization methods based on generalized self-consistent
schemes and (ii) a discrete model of spheres assembly which explicitly describes the ability of the contacts to transmit efforts.
Keywords: Nanocomposites; Mechanical properties; Percolation; Discrete modeling approach; Self-consistent modeling; Cryo TEM
1. Introduction
Among the several reasons to incorporate fillers into poly-
mers (cost reduction, improvement of some physical proper-
ties such as flame retardancy or barrier properties) mechan-
ical reinforcement is expected [1]. Traditional fillers display
average characteristic sizes in the range of several microns.
However, due to the development of nanosized fillers, the
specific influence of the nanometric size in the reinforce-
ment mechanisms has to be addressed. Composite materials
Corresponding author. Tel: +33 169333318; fax:+33 169333026.
E-mail address: chabert@lms.polytechnique.fr (E. Chabert).
based on nano-sized fillers, the so-called nanocomposites,
are presently studied especially because they may have un-
usual combinations of properties [2–6]. These unusual prop-
erties may be a consequence of the extremely large specific
interfacial area (hundreds of m
2
/g). They may be also related
to the very short distances between the reinforcing fillers
(about 10
8
m), that may become close to the characteristic
size of the macromolecular coils. In addition, some years
ago, we showed that drastic reinforcing effects may be ob-
served at very low volume fractions for fillers with very large
aspect ratio, when the percolation of the fillers occurs [7].
However, the processing of composite materials that only
differ by the size of the fillers (keeping constant the disper-
sion state and surface properties) is not a simple task. In or-
1

der to by-pass this difficulty, experimental data may be com-
pared to theoretical predictions that are developed to account
for the mechanical behavior of heterogeneous materials. Pa-
rameters which play a role for the mechanical properties of a
filled polymer (elastic modulus for instance) are the follow-
ing: (i) the (visco)elastic properties of its constitutive phases,
(ii) the volume fraction of filler, (iii) the morphology (i.e.,
shape, aspect ratio, and distribution of the filler within the
polymeric matrix) and (iv) the interactions between fillers
and between filler and matrix. Various models have been
proposed in the literature to understand the complex inter-
play between these parameters and to display a prediction
of the elastic moduli of polymer composites. Moreover, it
is generally accepted that these elastic calculations can be
extended to the description of the viscoelasticity of filled
polymers through the correspondence principle of Hashin
[8]. In fact, the classical models, at least in their former de-
velopments, ignore direct interactions between fillers: fillers
are considered as a homogeneous phase which interacts with
the matrix or, at most, with the effective medium, and thus
these models may be unefficient when the interactions be-
tween the fillers themselves rule the mechanical response
of the composite. Moreover, they cannot simply account for
the percolation of rigid fillers within a soft matrix.
The work presented in the following has been performed
in the framework of a collaboration between several labo-
ratories. The general strategy was first to elaborate model
nanocomposites with controlled morphologies, to character-
ize their mechanical behavior in the linear domain and to
develop mechanical models adapted to describe their behav-
iors. The first part of the paper presents the fabrication route
to process polymers filled with spherical particles with di-
ameters in the nanometer range. In order to understand re-
inforcement mechanisms depending on connectivity or on
the aggregation state, we have focused on spherical fillers
to avoid orientation effects. Materials were produced with
different dispersion states: nanocomposites obtained from
a mixture of hard/soft latex particles and nanocomposites
made of encapsulated hard particles surrounded by a soft
polymer shell. In the first part (Section 2), we provide de-
tails on sample preparation and characterization of the ob-
tained materials. In a second part (Section 3), the viscoelastic
behavior of the nanocomposites, characterized by dynamic
mechanical measurements, is described. These experimen-
tal data illustrate the influence of filler–filler interactions on
the linear mechanical behavior of these materials. At last
(Section 4), different routes to model the (visco)elastic re-
sponses are discussed.
2. Processing of nanocomposite systems
A convenient way to process nanocomposite materials
is based on the mixture of various aqueous suspensions
(colloids) [9–12]. Emulsion polymerization is well known
to provide in a simple way polymer colloidal suspensions
with typical particle size in the range of ten to a few hun-
dred nanometers. When both colloids are film forming, a
co-continuous material can be expected, provided that the
fraction of each component is large enough. Blending of
hard and soft particles would lead, after film formation, to a
random distribution of hard particles in the continuous soft
matrix, with a certain probability to form aggregates de-
pending on the volume fraction of filler. On the contrary, if
the particles are structured, with a stiff core and a soft shell,
then the material should consist, after shell coalescence, of a
soft matrix with regularly dispersed stiff spherical domains,
without contact between each other.
2.1. Nanocomposites obtained from PS–PBA and
silica–PBA latex blends
Different composites, based on the mixture of a film form-
ing latex (matrix) with aqueous suspension of fillers in re-
quired proportions, were prepared. Choice was made of a
poly(butyl acrylate) matrix (PBA, glass transition tempera-
ture T
g
around 47
C) filled with spherical particles of ei-
ther polystyrene (PS, T
g
around 97
C) or silica. Homopoly-
mer latexes of PS or PBA were obtained through batch emul-
sion polymerization process. A typical recipe was as fol-
lows: deionized water = 900 g, monomer = 90 g, NaHCO
3
= 0.75g,initiator = 0.75 g (K
2
S
2
O
8
for PS and (NH
4
)
2
S
2
O
8
for PBA), emulsifier = 2.91 phm of NC12(3-(dimethyl do-
decylammonium)propane-1-sulfonate) for PS, and 0.33 phm
of sodium dodecyl sulfate for PBA. The temperature of
polymerization = 70
C. Latexes were cleaned using ion
exchange resins removing most of the surfactant, in order
to avoid a possible effect on film forming process, film
microstructure or film properties. They were cleaned in a
non-diluted state (around 10 wt.% solid content) until con-
ductivity remains constant. Particle sizes were around 110
nm for PS and 135 nm for the PBA latex (as measured by
dynamic light scattering, Malvern Autosizer Lo-C instru-
ment). Silica nanoparticles have been synthesized according
to the Stöber process [13]. In a typical procedure, absolute
ethanol (1555 g, Acros Organics), de-ionized water (24.9 g)
and ammonia (12.75 M, 96.6 g, Laurylab) were filled into a
5L polypropylene flask. The mixture was stirred at 160 rpm
to be homogenized and tetraethoxysilane (91.65 g, Fluka)
was introduced at once. Reaction occurred at room temper-
ature under continuous stirring for 2 weeks. The alcoholic
silica suspension was dialyzed against water before use. Par-
ticle size was determined to be 125 nm by dynamic light
scattering (DLS). The solid content of the resulting suspen-
sion was determined gravimetrically.
After blending different amounts of filler (PS or silica)
(from 15 to 45vol.%) and PBA latexes, films were made
by evaporation under controlled atmosphere during a de-
lay long enough to achieve a complete maturation, i.e., 2
weeks at 35
C and 90% relative humidity. Although PBA
matrix is transparent after maturation, highly filled com-
posite films were rather opaque, suggesting aggregation of
2

(CH
3
O)
3
Si
O
O
Fig. 1. Chemical formula of the 3-trimethoxysilyl propyl methacrylate
silane coupling agent used in the silica encapsulation reaction.
the hard phase [14], as it has been previously observed on
similar PS/PBA systems by small angle neutron scattering
[15]. Films were tested after the lm formation process (and
called as-dried lms) or after a thermal treatment at 140
C
for 4 h.
2.2. Nanocomposites obtained from encapsulated silica in
P(S–BA) copolymer
The second nanocomposite system was based on encap-
sulated silica in a P(SBA) copolymer. The choice of this
copolymer (with glass transition close to 0
C was moti-
vated by the possibility to test the mechanical behavior. As
a matter of fact, we tried rst with lms issued from latex of
encapsulated silica in PBA, but these lms were very dif-
cult to handle. One reason could be the formation of a high
degree of crosslinking in these systems due to both chain
transfer to polymer and bimolecular termination by recom-
bination [16]. Such a mechanism is expected to signicantly
affect lm formation and mechanical properties.
Since silica is initially hydrophilic, its surface needs to be
modied to make possible anchoring and polymerization of
the hydrophobic butyl acrylate and styrene (co)monomers
through an emulsion polymerization process. This was
achieved by the chemical grafting of an alkoxysilane bear-
ing a polymerizable methacryloyl end group (see Fig. 1).
The grafting was performed as described previously by
direct addition of 3-trimethoxysilyl propyl methacrylate
(MPS, Acros Organics) to an aqueous silica suspension
[17]. The silica sol (30N50), with a mean hydrodynamic
diameter of 68nm as determined by DLS and 32 wt.% solid
content, was kindly supplied by Clariant S.A. (France) and
diluted in de-ionized water before use. A xed amount
of MPS (corresponding to 16 mol/m
2
silica surface) was
introduced in the diluted silica sol (10g L
1
) containing
0.25g L
1
sodium docedyl sulfate surfactant (SDS, Acros
Organics). The reaction was conducted at room tempera-
ture for 1 week. The suspension was concentrated using an
evaporating rotator before use.
The grafted silica particles were further engaged in a
free radical polymerization process to make the polymer
grow from their surface. The emulsion polymerization re-
action was performed in batch at 70
C in a 250 mL double
wall glass reactor tted with a condenser. The reactor was
charged with 100g of the aqueous suspension containing the
grafted silica beads (37g L
1
) and the surfactant (a mixture
of SDS : 1.25g L
1
and poly(oxyethylene) isooctyl cyclo-
hexyl ether : TX-405, 0.75g L
1
, Aldrich). After degassing,
the monomers from Aldrich (styrene : 36 g L
1
and butyl
Fig. 2. Cryo-TEM image of grafted silica particles (dark) encapsulated
with P(S-BA) through emulsion polymerization.
acrylate : 63 g L
1
) and the initiator (potassium persulfate
: KPS, 0.5 g L
1
, Acros Organics) were successively intro-
duced at 70
C under stirring to start polymerization. The
monomer to polymer conversion was of 86% as determined
gravimetrically.
The morphology of the composite particles was char-
acterized using cryo-transmission electron microscopy
(cryo-TEM). Using the method described elsewhere [18,19],
thin liquid lms of the particle suspension were formed on
carbon membranes and rapidly frozen into liquid ethane.
The particles were then observed at low temperature em-
bedded in a preserving lm of vitreous ice, using a Philips
CM200 Cryo microscope operated at an accelerating volt-
age of 80 kV. The TEM image in Fig. 2 clearly attests for a
successful encapsulation of the nanometric silica particles
by the P(SBA) copolymer. Although the silica beads ap-
pear to be uncentered in the polymer shell, they are expected
to be homogeneously distributed in the coalesced latex lm.
However, a limited number of silica/silica contacts may be
present.
For comparison purpose, blends of silica and P(SBA)
copolymer latexes were prepared by mixing together the sil-
ica sol and a latex of P(SBA). This copolymer latex was
obtained from the previous latex of encapsulated silica in
P(SBA), after having eliminated the grafted silica by cen-
trifugation. Silica/P(SBA) nanocomposites were obtained
after maturation of latex blends or encapsulated latex during
2 weeks at 35
C and 90% relative humidity.
3. Viscoelastic behavior
The dynamic shear moduli (G
, G

) of various nanocom-
posites were measured as a function of temperature with
a homemade inverted pendulum working in helium atmo-
sphere [20]. Experiments were performed in the temperature
3

0,00001
0,0001
0,001
0,01
0,1
1
-100 -50 0 50 100 150
T (˚C)
G' (GPa)
PBA
15% PS
20% PS
25% PS
35% PS
45% PS
PS
0,01
0,1
1
10
-100 -50 0 50 100 150
T (˚C)
tan d
PBA
15% PS
20% PS
25% PS
35% PS
45% PS
PS
Fig. 3. Real shear modulus and loss factor vs. temperature for as dried PSPBA nanocomposites with different ller amounts (1 Hz).
range [170 to 150
C] with a heating rate of 1
C/min and
at a xed frequency of 0.1 or 1 Hz.
3.1. PS–PBA nanocomposites
The dynamic mechanical responses of PBA and PS/PBA
as-dried systems (i.e., just after PBA maturation) are
shown in Fig. 3. The properties of a lm of pure PS (ob-
tained by freeze-drying/hot pressing) are also reported. A
signicant reinforcement of the shear modulus is observed
in between PS and PBA main (or α) relaxations (32
C
T 118
C). One can notice that this reinforcement is
not linear: the faster increase is observed around 20 vol.%,
i.e., near the geometric percolation threshold. In the same
temperature range, the height and the width of the tan δ
peak (associated with the PBA glass transition) decrease,
with a slight shift of the relaxation to lower temperature.
Nevertheless, differential scanning calorimetry measure-
0,01
0,1
1
10
-100 0 100
T (˚C)
after annealing
before annealing
after annealing
before annealing
0,0001
0,001
0,01
0,1
1
-100 0 100
T (˚C)
G' (GPa)
tan d
Fig. 4. Real shear modulus and loss factor for PBA samples lled with 35% PS, before and after thermal treatment (1 Hz).
ments performed on these systems showed that the PBA
glass transition temperature remains constant and equal to
47
C (for a heating rate of 10
C/min) whatever the ller
content. Hence, the shift of T
α
to lower temperature as the
ller content increases is probably due to a mechanical
coupling effect. In addition to a higher level of the relaxed
modulus, the temperature dependence of composite mod-
uli differs from that observed for the pure matrix and for
the ller phase. For the composite materials, the relaxed
modulus rst decreases until 70
C and then increases be-
fore falling again above the PS main relaxation. These two
counteracting phenomena lead to the presence of a bump
in the the tan δ curves, located between the peaks associ-
ated with the PBA and PS main relaxations. As discussed
in [21], these two phenomena are a consequence of the
evolution of ller–filler interactions with temperature: in
the low temperature range (T 70
C), the decrease of
relaxed composite moduli can be attributed to a decrease of
4

Citations
More filters
Journal ArticleDOI

Characterization of Polymer Nanocomposite Interphase and Its Impact on Mechanical Properties

TL;DR: In this paper, the structure of the interphase, a region between nanoparticle fillers and the bulk polymer matrix in a particle reinforced composite, was investigated using two different approaches and the results showed that the polymer matrix is more robust than the nanoparticles.
Journal ArticleDOI

Well-Dispersed Fractal Aggregates as Filler in Polymer−Silica Nanocomposites: Long-Range Effects in Rheology

TL;DR: In this article, a new method of processing polystyrene−silica nanocomposites is presented, which results in a very well-defined dispersion of small primary aggregates (assembly of 15 nanoparticles of 10 nm diameter) in the matrix.
Journal ArticleDOI

Concepts and conflicts in nanoparticles reinforcement to polymers beyond hydrodynamics

TL;DR: In this article, a comprehensive survey is presented to report the cluster-cluster aggregation model, and jamming, percolation and soft colloidal dynamics theories and their applications in NPFPs in relation to nanoparticle reinforcement of polymers beyond hydrodynamics.
Journal ArticleDOI

Processing and microstructure of PCL/clay nanocomposites

TL;DR: In this paper, the morphology and mechanical properties of polycaprolactone/clay nanocomposite films prepared by two techniques (casting: exfoliation-adsorption; intensive mixing: melt-intercalation) were studied.
Journal ArticleDOI

Aggregation of colloidal nanoparticles in polymer matrices

TL;DR: It is believed that the fine-tuning of the structure of the filler phase opens new perspectives for systematic studies of the reinforcement effect, by modifying filler-polymer interfacial properties at fixed structure, or by generating original structures.
References
More filters
Journal ArticleDOI

Controlled growth of monodisperse silica spheres in the micron size range

TL;DR: In this article, a system of chemical reactions has been developed which permits the controlled growth of spherical silica particles of uniform size by means of hydrolysis of alkyl silicates and subsequent condensation of silicic acid in alcoholic solutions.
Journal ArticleDOI

A variational approach to the theory of the elastic behaviour of multiphase materials

TL;DR: In this paper, the authors derived upper and lower bounds for the effective elastic moduli of quasi-isotropic and quasi-homogeneous multiphase materials of arbitrary phase geometry.
Journal ArticleDOI

Polymer-layered silicate nanocomposites: an overview

TL;DR: An overview of polymer-clay hybrid nanocomposites is provided with emphasis placed on the use of alkylammonium exchanged smectite clays as the reinforcement phase in selected polymer matrices as discussed by the authors.
Journal ArticleDOI

Analysis of Composite Materials—A Survey

TL;DR: In this paper, the authors review the analysis of composite materials from the applied mechanics and engineering science point of view, including elasticity, thermal expansion, moisture swelling, viscoelasticity, conductivity, static strength, and fatigue failure.
Journal ArticleDOI

Solutions for effective shear properties in three phase sphere and cylinder models

TL;DR: In this paper, the effective shear modulus of two types of composite material models are compared. And the results are found to differ from those of the well-known Kerner and Hermans formulae for the same models.
Related Papers (5)