Está en la página 1de 8

INSTITUTE OF PHYSICS PUBLISHING NANOTECHNOLOGY

Nanotechnology 16 (2005) S514–S521 doi:10.1088/0957-4484/16/7/028

Structure and melt rheology of


polystyrene-based layered silicate
nanocomposites
Liang Xu1 , Stacey Reeder2 , Mahesh Thopasridharan2 ,
Jiaxiang Ren1 , Devon A Shipp2,3 and Ramanan Krishnamoorti1,3
1
Department of Chemical Engineering, University of Houston, 4800 Calhoun, Houston,
TX 77204–4004, USA
2
Department of Chemistry, Clarkson University, Potsdam, NY 13699-5810, USA

E-mail: dshipp@clarkson.edu and ramanan@uh.edu

Received 18 January 2005, in final form 27 April 2005


Published 2 June 2005
Online at stacks.iop.org/Nano/16/S514
Abstract
The melt-state viscoelastic properties of exfoliated in situ polymerized and
intercalated solution-blended polystyrene (PS) and organically modified
montmorillonite nanocomposites were investigated and compared. The PS
nanocomposites prepared by nitroxide-mediated polymerization (NMP)
exhibit a stable exfoliated structure whereas the PS nanocomposites
prepared by solution mixing exhibit an intercalated structure. The linear
viscoelastic properties were strongly correlated with the dispersion state of
the nanocomposites. On the other hand, the non-linear oscillatory shear
properties exhibited shear thinning character and were consistent with the
weak interactions between the polymer and the layered silicate.

1. Introduction tethering the polymer chains to the layered silicate surface


and by controlling the rates of intra-gallery and extra-gallery
Significant interest has recently focused on understanding polymerization, the final equilibrium structure of the hybrids
structure–property relations in polymer nanocomposites and can be controlled [1, 4, 5, 13, 14]. Additionally, the rheological
how processing affects both structure and properties [1, 2]. properties of in situ polymerized ionically tethered chains are
In part, this is due to the dramatic improvements in material distinctly different from those of melt-blended nanocomposites
properties of nylon-6-layered silicate-based nanocomposites and presumed to originate from the differences in polymer
that were demonstrated by the Toyota research group on in situ conformation in those hybrids [2].
polymerized hybrids [3, 4]. Subsequently, interest has also In this work, we exploit a recently developed free
focused on trying to reproduce and emulate those results using radical polymerization route termed nitroxide-mediated
melt processing of the components [1, 5]. polymerization (NMP) for the preparation of partially tethered
Polystyrene-layered silicate nanocomposites have been nanocomposites based on polystyrene and compare and
prepared by both in situ bulk, solution or emulsion contrast the structure and rheological properties to those
polymerization [6–10] and melt or solution processing [11, 12]. from the solution-based intercalation approach. By using
The latter typically results in intercalated materials [11], while NMP, the in situ polymerized polystyrene samples have low
the former may give a variety of mixtures of intercalated [7] polydispersity (<1.3), thus making for more convenient and
or exfoliated [8–10] materials, depending on a variety of meaningful comparisons. The melt and solution intercalation
parameters such as mode of polymerization and type of of layered silicates by polystyrene has been extensively studied
silicate modifier. Of particular interest when examining previously and is considered as a model system where only
structure–property relationships is the speculation that by weak thermodynamic interactions are expected [9, 10, 15–18].
performing in situ polymerization, with some possibility of Melt state rheological measurements have been shown to be
an extremely powerful method to understand the mesoscale
3 Authors to whom any correspondence should be addressed.
structure and the strength of polymer–filler interactions in both

0957-4484/05/070514+08$30.00 © 2005 IOP Publishing Ltd Printed in the UK S514


Structure and melt rheology of polystyrene-based layered silicate nanocomposites

macrocomposites and nanocomposites [18–20]. In particular, polymer and clay were separated by refluxing in the sample
we use linear viscoelastic measurements to examine if the 2.5 wt% LiBr in a methanol/THF mixture (1/3 v/v) and
nanoparticles create a network structure that is able to sustain refluxing for 3 h. The solution was then subjected to
stress and therefore understand the mesoscale dispersion of the centrifugation, and the polymer removed from the supernatant
nanoparticles. Additionally, we also examine if the presence liquid by precipitation into methanol.
of a small fraction of tethered chains is able to render the
non-linear viscoelastic properties to exhibit strain-hardening or 2.2. Solution intercalation
whether the shear-thinning character of the matrix dominates
the non-linear response. Solution intercalation samples of PS30K (Pressure Chemicals
Inc., Mw /Mn < 1.04) with a dimethyl dioctadecyl ammonium
modified montmorillonite (2C18M) were prepared with 3,
2. Experimental methods 6.7 and 9 wt% layered silicate by dispersing the silicate
as a dilute dispersion in toluene to which the polymer
2.1. In situ polymerization
was added and subsequently rapidly dried to remove the
2.1.1. Materials. N ,N -dimethyl-n-hexadecylamine (Fluka) solvent. The montmorillonite used here had a charge
was used as received. 4-vinylbenzyl chloride (4VBC; exchange capacity (CEC) of 0.9 eq kg−1 and a nominal
Aldrich) was filtered though alumina to remove the disc diameter of 0.5 µm, as estimated from dilute solution
inhibitor then purified via distillation under pressure. dynamic light scattering and confirmed by quantitative analysis
Styrene (Fisher) was purified by filtration through alumina of several tens of transmission electron micrographs of
followed by distillation under pressure over CaH2 . 2,2 - dispersions in epoxy [18]. The description of the organic
azobis(isobutyronitrile) (AIBN; Kodak) was recrystallized modification and sample preparation is routine and is provided
from methanol. 2,2,6,6-tetramethylpiperidinoxy (TEMPO; elsewhere [18, 21]. Samples were subsequently annealed at
Aldrich) was used as received. Montmorillonite clay (MMT; 100 ◦ C in a vacuum oven for ∼12 h followed by extensive
Cloisite-Na from Southern Clay Products, ion exchange annealing (∼24 h) at 160 ◦ C to remove any remaining solvent
capacity of 92 mequiv/100 g) was used as received. and to facilitate complete polymer intercalation between the
silicate layers.
2.1.2. Synthesis of N ,N -dimethyl-n-hexadecyl-(4-vinylbenzyl)
ammonium chloride (VB16). Synthesis and purification of 2.3. Instrumentation
VB16 were carried out according to the published proce- 2.3.1. Gel permeation chromatography (GPC). Samples
dure [9]. To a round bottom flask 10.80 g of N ,N -dimethyl-n- for GPC analysis were dissolved in a solution of 2.5 wt%
hexadecylamine was added with 4.56 g of 4VBC in 40 ml of LiBr in a methanol/THF mixture (1/3 v/v) and refluxed for
ethyl acetate. The reaction was stirred at 40 ◦ C overnight. A 3 h. The solution was then subjected to centrifugation and
white precipitate was recovered by filtration and recrystallized the supernatant filtered through a 0.2 µm filter. Molecular
using ethyl acetate. 10.34 g of pure product was obtained to weights and molecular weight distributions were determined
give a percentage yield of 71.14%. using gel permeation chromatography (GPC) equipped with
a Waters 717 autosampler, a Waters 515 HPLC pump, three
2.1.3. Intercalation of VB16 into clay. The intercalation was Waters columns and a Viscotek LR40 laser refractometer with
carried out according to the published procedure [9]. 25 g of THF as the eluent. The GPC was calibrated using polystyrene
unmodified MMT was stirred in 1 l of water overnight. 10.34 g standards.
of VB16 was added to 100 ml of water and this solution was
then added dropwise to the clay mixture and stirred for 3 h 2.3.2. X-ray diffraction analysis (XRD). The dispersion state
at temperatures between 0–5 ◦ C. The clay was filtered and of the layered silicates was ascertained by x-ray diffraction and
washed with water several times and vacuum dried over several using Siemens D5000 diffractometers with Cu Kα radiation
days. A portion of the clay was then stirred in petroleum ether (λ = 1.54 Å) generated at 30 mA and 40 kV. Diffraction traces
for 1 h and vacuum dried overnight. were obtained over a 2θ range of 2◦ –10◦ in steps of 0.02◦ and
counting time of 3 s at each angular position.
2.1.4. General synthesis of polystyrene. All solids were
first added to a Schlenk flask. The flask was subjected to 2.3.3. Thermogravimetric analysis (TGA). Thermogravi-
three cycles of alternating vacuum/N2 . Styrene was added via metric analysis (TGA) of dried polymer samples was per-
syringe after being purged with N2 for 30 min. The reaction formed on a Perkin-Elmer TGA7. Measurements were per-
was stirred for 48 h at 120 ◦ C. THF was then added to the formed in flowing air at a heating rate of 10 ◦ C min−1 .
flask, and the sample then precipitated into methanol, filtered
and dried under vacuum. 2.3.4. Melt rheology. Melt rheological measurements
were performed using a melt state ARES rheometer (TA
2.1.5. Separation of free and tethered polystyrene. Instruments) with a torque transducer capable of torque
Free polystyrene was removed through dissolution of the measurements over 0.2–2000 g cm. Dynamic oscillatory shear
nanocomposite in THF and then centrifugation. The measurements (120 ◦ C  T  190 ◦ C, with a N2 purge
supernatant solvent was removed by rotary evaporation and used for all measurements) were performed by applying a
the polystyrene yield determined gravimetrically. The tethered sinusoidal strain of the form γ (t) = γ0 sin(ωt) (where γ0

S515
L Xu et al

H3C(H2C)15 N

NMP
N (CH2)15CH3
Styrene + TEMPO + BPO

H3C(H2C)15 N
O N

N (CH2)15CH3
O O

O O
Polymer-silicate nanocomposite

VB16-MMT

Scheme 1. (This figure is in colour only in the electronic version)

Table 1. Distribution of free and tethered polymer, and recovered


layered silicatea.
PS30K
E-PS-A E-PS-B + 9 wt %
2C18M

Intensity (AU)
Free polymer 87.0 87.4
Tethered polymer 8.8 6.6 PS 30K
+ 6 wt % PS 30K
Layered silicate 3.3 4.0
2C18M + 3 wt %
a
Percentages based on the mass of 2C18M
recovered material relative to the initial
mass of the sample used, and add up to
less than 100 because of losses during E-PS-B
work-up. E-PS-A
2 4 6 8 10 12
2θ (λ = 1.54 A)
is the strain amplitude and ω is the frequency with 0.001 
ω  100 rad s−1 ), and measuring the resultant shear stress Figure 1. X-ray diffraction traces for the two in situ polymerized
σ (t) (=γ0 {G  sin(ωt) + G  cos(ωt)}), with G  and G  being nanocomposites E-PS-A and E-PS-B and the three
the storage and loss modulus, respectively. Discs of samples solution-intercalated nanocomposites based on 2C18M and a model
with thickness ∼1–2 mm and diameter 28 mm were prepared PS30K. All samples were annealed in the melt (160 ◦ C) for at least
24 h prior to XRD measurements. Both E-PS-A and E-PS-B exhibit
by vacuum moulding the samples at 160 ◦ C in a pellet die no peaks and suggest disordered intercalated or exfoliated
and heated and pressed in a Carver press with the application structures, while all three solution-intercalated samples exhibit
of minimal force on the plunger to ensure minimal process- peaks corresponding to 001, 002 and 003 reflections, suggesting a
related orientation in these samples. For the linear viscoelastic well organized intercalated structure.
properties small strain amplitudes (i.e., γ0 < 0.05) were
typically employed and the response was verified to be linear by tethered polystyrene based nanocomposites, samples E-PS-A
changing the strain amplitude by a factor of two and observing and E-PS-B, were prepared using a nitroxide mediated scheme
the invariance of the viscoelastic parameters measured. For as illustrated in scheme 1.
shear-alignment tests, large amplitude oscillatory shear over From these polymerizations it is expected that some
prolonged periods of time was imposed and the resulting shear polymer chains will be grafted (or tethered) to the silicates
stress interpreted in the context of the linear response theory. through copolymerization of the styrene with the VB16 that
However, no effort was made in these studies to verify that the is bound to the silicate layers, while other chains remain
higher order harmonics were indeed negligible. Finally, for the unattached to the silicate. The relative amounts of tethered and
strain sweep measurements, at constant frequency the sample untethered (‘free’) polymer chains were determined for both E-
was subjected to progressively increasing strain amplitudes PS-A and E-PS-B, and are summarized in table 1. These data
and followed by a succession of smaller strains. Again the indicate that most of the chains are not tethered, with only 6–
data were interpreted in terms of the linear parameters and 9% bound to the clay. Su and Wilkie [10] have reported similar
should be construed only as a qualitative representation of the values (approximately 5%) for bound PS in nanocomposites
true viscoelastic response of these materials. based on VB16-MMT synthesized using conventional radical
polymerization of styrene. The clay content of each sample
that was determined experimentally through extraction (3.3
3. Results and discussion
and 4.0 wt% for E-PS-A and E-PS-B, respectively) is similar
3.1. Synthesis and structure of nanocomposites to the clay content calculated from the initial amount of
clay added and the monomer conversion (3.5 and 3.9 wt%,
The VB16 modified layered silicates exhibited a gallery respectively).
height of 1.25 nm (d = 2.2 nm) and on the basis of TGA The development of the polymer molecular weight during
measurements had a total organic content of 27 wt%. The nitroxide-mediated polymerization (NMP) is quite different

S516
Structure and melt rheology of polystyrene-based layered silicate nanocomposites

(a) 8 (b) 8
10 10
PS30K + 2C18M PS30K + 2C18M
7
10 wt % 2C18M 10
7 wt % 2C18M
6
0 0
10 3

bTG'' (dynes/cm2)
bTG' (dynes/cm2)

6 3
6.7 10
6.7
5
10 9 9
5
10 2
4 2.0
10 1
0
3
6.7
1.5 4 9
10

log(aT)
3 0
10 1.0

β
-1
0.5 3
10
2 10 -2
0.0
130 140 150 160 170 180
0 2 4 6 8 10
o
Silicate content, wt% T ( C)
1 2
10 -4 10
-3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
o o
aTω(rad/sec) [To=150 C] aTω(rad/sec) [To=150 C]

(c) 108 (d) 8


10
1.0
PS30K + 2C18M
0.8
7 wt % 2C18M
10 0.6 7
10
α

0.4 0
0.2 3
6 6 6.7
η∗ (poise)

10 η∗ (poise)
0.0
0 2 4 6 8 10
10
Silicate content, wt%
9

5 5
10 10
PS30K + 2C18M
4 wt % 2C18M 4
10 10
0 6.7
3 9
3 3
10 10
-4 -3 -2 -1 0 1 2 3 4 3 4 5 6 7 8
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
o 2
aTω(rad/sec) [To=150 C] G* (dynes/cm )

Figure 2. Time–temperature superposed mastercurves illustrating the frequency dependence of the linear viscoelastic dynamic storage
modulus G  (a), loss modulus G  (b) and complex viscosity η∗ (c) for PS30K and the intercalated nanocomposites of PS30K with 2C18M.
The data for the 6.7 and 9 wt% nanocomposites demonstrate solid-like character and this behaviour is best observed in a cross-plot of the
complex modulus G ∗ versus η∗ (d) and demonstrates that the viscosity for those nanocomposites diverges at a finite and large value of G ∗ .
The low-frequency frequency dependence of G  evaluated by the parameter β(G  ∝ ωβ ) and of η∗ evaluated by the parameter α(η∗ ∝ ωα )
are captured in the insets of parts (a) and (c) respectively, and clearly demonstrate the similarity to the scaling observed for soft glassy
materials with a finite yield stress. The inset in (b) illustrates the temperature dependence of the frequency shift factors (aT ) used to prepare
the viscoelastic mastercurves for the polymer and the nanocomposites. The shift factors, within the errors of the measurements, are
independent of the silicate loading and are adequately represented by the simple WLF equation.

when compared to conventional radical polymerization. In no peak while the three solution-intercalated nanocomposites
NMP, all chains begin growth early in the polymerization, exhibit intercalated structures with well ordered structures
producing oligomers that continue to grow throughout the as suggested by the presence of 001, 002 and 003
reaction. Ideally, the number average molecular weight reflections. Both E-PS-A and E-PS-B when examined by
(Mn ) increases linearly with monomer conversion, and the electron microscopy demonstrate largely exfoliated structure
polydispersity is less than 1.5, and often less than 1.3. The while the solution-intercalated nanocomposites exhibit ∼10%
GPC analysis of the E-PS-A sample demonstrated an increase individual sheets and the rest in tactoids consisting of on
in the number average molecular weight with reaction time average eight to nine layers.
and monomer conversion, with no appreciable change in
the polydispersity index (less than 1.35 throughout). This 3.2. Linear viscoelasticity
indicates that the polymerizations proceeded as expected.
Two samples were synthesized using this method: E-PS-A The linear viscoelastic time–temperature superposed mas-
containing 3.3 wt% silicate, Mn = 12.3 × 103 g mol−1 and tercurves of macroscopically unoriented solution-intercalated
Mw /Mn = 1.19 , and E-PS-B containing 4.0 wt% silicate, PS30K with differing amounts of 2C18M are shown in fig-
Mn = 21.8 × 103 g mol−1 and Mw /Mn = 1.32 . The XRD ure 2. PS30K is a lightly entangled polymer and for reduced
traces for aliquots of samples removed after 5, 24 and 48 h frequencies (aT ω) below 100 rad s−1 exhibits liquid-like be-
all demonstrated no observable peak corresponding to layer— haviour with G  scaling as ω2 and η∗ scaling as ω0 as ex-
layer stacking over a scattering angle 2θ of 2◦ –10◦ . pected from a simple Newtonian liquid [22]. On the other
The structures of melt-annealed nanocomposites were hand, the nanocomposites demonstrate: (i) time–temperature
characterized using XRD and the traces are shown in superposed mastercurves, with frequency shift factors compa-
figure 1. The two nanocomposites prepared by NMP in rable to that of polystyrene and moduli shift factors close to
the presence of the silicate, E-PS-A and E-PS-B, exhibit unity (0.97  bT  1.03); and (ii) a gradual change in the low

S517
L Xu et al

(a) 7 (b)
10 107 108
7 E-PS-A
10 E-PS-B
PS12.3K + 3.5 wt% MMT PS21.8K + 3.9 wt% MMT
106 107
G', G" (dynes/cm2)

G', G" (dynes/cm2)


6
6 10
10

η∗ (Poise)

η∗ (Poise)
5 6
10 10

5 5 G"
10 10
G' 4 5
10 10
G'
10
4 G"
4 3 4
10 10 10
-2 -1 0 1 2 3 -4 -3 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10 10 10 10
aTω (rad/s) [To=130 oC] aTω (rad/s) [To=140 oC]

(c)
E-PS-B
7 o
10 To=140 C
η∗ (Poise)

106

105
E-PS-A
To=130 oC
4
10
103 104 105 106 107
2
G* (dynes/cm )

Figure 3. Time–temperature superposed linear viscoelastic mastercurves of G  , G  and η∗ for (a) E-PS-A (PS12.3K + 3.5 wt% MMT; NMP
exfoliated) and (b) E-PS-B (PS21.8K + 3.9 wt% MMT; NMP exfoliated). The data for both samples demonstrate solid-like melt rheological
properties. The data in part (c) illustrate the solid-like character for both of these nanocomposites with the viscosity diverging at reasonably
high values of the complex modulus.

frequency viscoelastic response from that of a liquid to that of percolation between 3 and 6.7 wt% of the layered silicate for
a solid-like material. Three signatures are worth considering the randomly oriented nanocomposites is consistent with the
for the transition from liquid-like to solid-like melt rheological intercalated nature of the nanocomposites with the presence of
behaviour: <10% (by number) of individualized silicate sheets (obtained
by extensive electron microscopy characterization and not
(a) the power-law exponent describing the low frequency
shown here [18]) and an effective aspect ratio of the order
response of G  (G  ∝ ωβ ), β, decreases from two to a
of ∼30 [21, 25].
value close to zero, with the values for the 6.7 and 9 wt%
The linear oscillatory viscoelastic responses for the
2C18M nanocomposites suggestive of a low frequency
exfoliated tethered nanocomposites prepared by nitroxide-
plateau in G  ;
mediated polymerization are presented in figure 3. The
(b) the power-law exponent describing the low frequency
linear viscoelastic response for both samples is capable of
response of η∗ (η∗ ∝ ωα ), α, increases from near zero
excellent time–temperature superpositioning of all viscoelastic
to a value close to unity, with the high volume fraction
nanocomposites showing large values for α or strong non- functions. These data demonstrate the solid-like character
Newtonian behaviour even at low frequencies; of these materials with a low frequency plateau of G  (β =
(c) a cross-plot of G ∗ versus η∗ reveals the divergence of 0.13 and 0.16 for E-PS-A and E-PS-B respectively) and
η∗ at a finite value of G ∗ for the higher weight fraction G  exceeding G  in the terminal region, low frequency
silicate nanocomposites and suggestive of a yield stress divergence of η∗ (α = 0.75 and 0.68 for E-PS-A and
material [21]. E-PS-B respectively) and the divergence of η∗ at a finite
(and relatively large) value of G ∗ . Clearly, both E-PS
These signatures of transformation from liquid-like to solid- samples are hydrodynamically percolated with the presence
like material are similar to those observed for other soft of a finite yield stress [20, 21, 26, 27]. The differences
glassy materials [23] and in the case of layered silicate in α and β between the two samples suggest that the
nanocomposites have been attributed to the geometrical dispersion state and the development of filler network are
percolation of dispersed nanoclay sheets or aggregates of the much better established for the E-PS-A sample. This is
sheets [18, 20, 21, 24]. The location of the geometrical further confirmed from an examination of the value of G ∗

S518
Structure and melt rheology of polystyrene-based layered silicate nanocomposites

6 5
(a) 3.0 0.7 (b) 10 10
γ o = 0.5 10 rad/sec

η∗ (P), E-PS-A, To=130 oC

η∗ (P), E-PS-B, To=160 oC


10 rad/sec
0.6
G' (dynes/cm2) x 10-4

G' (dynes/cm2) x 10-3


2.5
0.1 rad/sec 0.1 rad/sec
γ o = 0.75
E-PS-A; 130 oC

E-PS-B; 160 oC
0.5
2.0
γ o = 1.0 0.4
1.5
γo = 5
10 10
4
0.3
0.5
1.0
γ o = 1.5 0.2
γ o = 0.75
0.5 γ o = 1.0 γ = 1.5 0.1
o
0.0 4 3
0.0 10 10
0 5 10 15 10-3 10-2 10-1 100
t (sec) x 10
-3 γo

(c)

4
10
E-PS-A, To 130 oC

E-PS-B, To =160 oC
10
G' (dynes/cm2),

G' (dynes/cm2),
10 rad/sec
4
0.1 rad/sec 3
10 10 rad/sec 10
0.1 rad/sec

3
10 2
10

-3 -2 -1 0
10 10 10 10
γo

Figure 4. (a) Time variation of the storage modulus G  for E-PS-A and E-PS-B during the application of prolonged large strain amplitudes
(γ0 ) at a frequency of 0.1 rad s−1 . The strain amplitude was successively raised from 0.5 to 1.5 as noted in the figure. The values of G 
decrease with increasing the amplitudes of strain, indicating the shear-induced alignment of the silicate layers. (b), (c) The strain amplitude
dependence of η∗ (b) and G  (c) at specified frequencies for E-PS-A and E-PS-B obtained immediately after prolonged
large-strain-amplitude measurements described in (a) and with progressively increasing strain amplitude. Similar data with decreasing strain
amplitude with minimal hysteresis was obtained.

at which η∗ diverges—∼3 × 104 dynes cm−2 for E-PS-A as increasing molecular weight of the matrix [18]. Such
compared to ∼2 × 103 dynes cm−2 for E-PS-B. In contrast, improvements in the dispersion with increasing molecular
for the PS30K solution-intercalated samples described above weight have been reported by Fornes and Paul [1, 5] on
the corresponding values of G ∗ are 5 × 103 and 1.3 × melt-processed nylon nanocomposites and by Manias et al
104 dynes cm−2 for the 6.7 and 9 wt% nanocomposites [29] on melt-processed polypropylene nanocomposites. The
respectively. It is clear from the rheological evidence presented presence of a higher fraction of individual silicate sheets in
above that in spite of the lower silicate content in E-PS-A as the PS systems results in the following observations for the
compared to E-PS-B (3.3 versus 4.0 wt%) E-PS-A exhibits rheological measurements of nanocomposites at comparable
a stronger filler network structure, and on the basis of the levels of added silicate: (i) an increase in the G ∗ value
G ∗ values the network structure of the silicate in E-PS-A is at which η∗ diverges with increasing molecular weight and
somewhat superior to that of the 9 wt% intercalated 30 K individualized sheet content; and (ii) an increase in the
sample and clearly demonstrates the excellent dispersion of low-frequency (terminal zone for the matrix) values of η∗
the silicate sheets. (normalized to that of the zero shear viscosity of the matrix)
The most significant difference between E-PS-A and and G  at comparable values of ω/ωrelaxation [18]. On the basis
E-PS-B is the lower molecular weight of E-PS-A. On of these previous experiments, that the results presented here
the basis of the mean-field theory developed by Vaia and suggest that E-PS-B, having a higher molecular weight and
Giannelis [16, 17], no difference in structure would be slightly higher silicate loading, exhibits a weaker silicate layer
anticipated on the basis of the change in molecular weight. network structure compared to E-PS-A is somewhat surprising.
On the other hand, the self-consistent theory developed by Perhaps the tethering of some of the chains to the silicate
Balazs and co-workers [28] would anticipate that increasing and the relatively poor inherent affinity between polystyrene
the molecular weight would lead to poorer dispersions and and the silicates [16] results in the poorer dispersion of the
consistent with the experiments described here. silicates with increasing chain length. However, it is important
Previous work on PS-based intercalated nanocomposites, to stress that, unlike the case of thermoset nanocomposites
prepared by solution processing, indicated that the fraction prepared with epoxies where the dispersion state, which can
of individual layers in the nanocomposites increased with be a function of extent of cure, has been shown to depend

S519
L Xu et al

most critically on the relative ratio of inter- and intra-gallery exhibit shear-thinning character and appear to suggest that the
rates of polymerization and is kinetically trapped [14, 30], large fraction of the matrix free polymer chains dominates the
in these polystyrene nanocomposites the structure formed large strain response. In this respect, it would be interesting to
remains thermodynamically stable throughout its exposure to test the properties of nanocomposites with increasing fraction
the melt state. These complexities point to the need for further of chains tethered to the surface to examine the transformation
work in this area. of matrix-dominated to tethered-dominated rheology. We are
currently examining this issue, in addition to exploring other
3.3. Non-linear viscoelasticity questions raised in this study, such as the structure-property
dependence on molecular weight.
The strain amplitude dependences of the dynamic oscillatory
shear data for the two exfoliated nanocomposites were studied Acknowledgments
to examine if the samples strain softened or if the samples
strain-hardened as observed in other brush systems end- We (JR and RK) gratefully acknowledge the ExxonMobil
tethered to the silicate sheets. Prior to examining the strain Chemical Company for partial financial support of this
amplitude dependence, the samples were pre-conditioned to work. We (LX and RK) also gratefully acknowledge
achieve a high state of orientation and therefore eliminate support in part by the Texas Institute for Intelligent Bio-
the orientation of the silicate layers from the viscoelastic Nano Materials and Structures for Aerospace Vehicles,
response [31, 32]. The pre-conditioning was achieved by funded by NASA Cooperative Agreement No. NCC-1-02038.
subjecting the samples to prolonged large amplitude oscillatory Further, the Center for Advanced Materials Processing at
shear (0.5  γ0  1.5) at low frequencies (ω = 0.1 rad s−1 ) Clarkson University (a New York State Center for Advanced
and the during-shear storage modulus as a function of shearing Technology), and the donors of The Petroleum Research Fund,
time is shown in figure 4(a). Typically, alignment experiments administered by the American Chemical Society, are gratefully
were stopped after all viscoelastic functions had acquired time- acknowledged for their financial support to DAS.
independent values. Previous studies have demonstrated that
such large amplitude oscillatory flows result in the orientation
of the silicate sheets with sheet normals parallel to the velocity
References
gradient direction [31, 32]. [1] Fornes T D and Paul D R 2003 Polymer 44 4993–5013
Immediately following the large-amplitude-strain mea- [2] Krishnamoorti R, Vaia R A and Giannelis E P 1996 Chem.
surements, strain sweep measurements at fixed frequency and Mater. 8 1728
temperature were conducted, first with increasing strain am- Giannelis E P, Krishnamoorti R and Manias E 1999 Adv.
Polym. Sci. 138 107–47
plitude and then followed by a sweep with decreasing strain Alexandre M and Dubois P 2000 Mater. Sci. Eng. 28 1–63
amplitude (figures 4(b) and (c) respectively) [27, 32]. At low [3] Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T,
strain amplitudes i.e., γ0  0.1, both η∗ and G  are indepen- Kamigaito O and Kaji K 1994 J. Polym. Sci. B 32 625–30
dent of strain amplitude. However, progression to higher strain Yano K, Usuki A, Karauchi T and Kamigaito O 1993 J. Polym.
amplitudes (γ0 > 0.1) leads to shear thinning behaviour and Sci. A 31 2493–8
Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T and
results in a decrease in both η∗ and G  . This result is signif- Kamigaito O 1993 J. Polym. Sci. A 31 983–6
icantly different from those obtained on poly(ε-caprolactone) Usuki A, Kato M, Okada A and Kurauchi T 1997 J. Appl.
end-tethered silicate nanocomposites where reversible strain- Polym. Sci. 63 137
hardening was observed and suggested to arise from a coil to [4] Usuki A, Koiwai A, Kojima Y, Kawasumi M, Okada A,
Kurauchi T and Kamigaito O 1995 J. Appl. Polym. Sci. 55
stretch transformation of the polymer chains [32]. In both E-
119–23
PS-A and E-PS-B the fraction of chains tethered is <10% while Kojima Y, Usuki A, Kawasumi M, Okada A, Kurauchi T,
for the PCL-based nanocomposites the fractions of free poly- Kamigaito O and Kaji K 1995 J. Polym. Sci. B 33 1039–45
mers were considerably smaller (∼<10% in all cases). Hence [5] Fornes T D, Yoon P J, Keskkula H and Paul D R 2001 Polymer
in the current measurements the effect of the tethered chains is 42 9929–40
[6] Doh J G and Cho I 1998 Polym. Bull. 41 511–8
not observed and the non-linear viscoelasticity dominated by
Laus M, Cameriani M, Lelli M, Sparnacci K, Sandrolini F and
the presence of free untethered shear-thinning PS chains. Francescangeli O 1998 J. Mater. Sci. 33 2883–8
Moet A and Akelah A 1993 Mater. Lett. 18 97–102
Noh M W and Lee D C 1999 Polym. Bull. 42 619–26
4. Conclusions Zeng C and Lee L J 2001 Macromolecules 34 4098–103
Zhu J, Start P, Mauritz K A and Wilkie C A 2002 J. Polym.
The preparation of in situ polymerized polystyrene nanocom- Sci. A 40 1498–503
posites using nitroxide-mediated polymerization leads to some Zhu J, Uhl F M, Morgan A B and Wilkie C A 2001 Chem.
tethering of the polymer chains and good dispersion of the sil- Mater. 13 4649–54
icate sheets to a reasonably exfoliated state. In comparison, Zhu J and Wilkie C A 2000 Polym. Int. 49 1158–63
[7] Akelah A and Moet A 1996 J. Mater. Sci. 31 3589–96
solution-blended nanocomposites resulted only in intercalated Okamoto M, Morita S, Taguchi H, Kim Y H, Kotaka T and
structures. Further, from melt state linear viscoelastic mea- Tateyama H 2000 Polymer 41 3887–90
surements we infer that in the case of the in situ polymer- [8] Fu X and Qutubuddin S 2000 Mater. Lett. 42 12–5
ized materials the dispersion improves with lowering molecu- Fu X and Qutubuddin S 2001 Polymer 42 807–13
lar weight and consistent with the weak interactions between Weimer M W, Chen H, Giannelis E P and Sogah D Y 1999
J. Am. Chem. Soc. 121 1615–6
the polymer and the silicate. Finally, the non-linear rheolog- [9] Zhu J, Morgan A B, Lamelas F J and Wilkie C A 2001 Chem.
ical properties for these in situ polymerized nanocomposites Mater. 13 3774–80

S520
Structure and melt rheology of polystyrene-based layered silicate nanocomposites

[10] Su S and Wilkie C A 2003 J. Polym. Sci. A 41 1124–35 [23] Sollich P, Lequeux F, Hebraud P and Cates M E 1997 Phys.
[11] Vaia R A, Jandt K D, Kramer E J and Giannelis E P 1995 Rev. Lett. 78 2020–3
Macromolecules 28 8180 Sollich P 1998 Phys. Rev. E 58 738–59
Vaia R A, Ishii H and Giannelis E P 1993 Chem. Mater. 5 1694 Bonn D, Kellay H, Tanaka H, Wegdam G and Meunier J 1999
Vaia R A and Giannelis E P 1997 Macromolecules 30 8000–9 Langmuir 15 7534–6
[12] Beyer F L, Tan N C B, Dasgupta A and Galvin M E 2002 Bonn D, Tanaka J, Wegdam G, Kellay H and Meunier J 1999
Chem. Mater. 14 2983–88 Europhys. Lett. 45 52–7
Kim T H, Lim S T, Lee C H, Choi H J and Jhon M S 2003 Bonn D, Tanase S, Abou B, Tanaka H and Meunier J 2002
J. Appl. Polym. Sci. 87 2106–12 Phys. Rev. Lett. 89 015701
Park C I, Park O O, Lim J G and Kim H J 2001 Polymer 42 [24] Krishnamoorti R and Giannelis E P 1997 Macromolecules 30
7465–75 4097–102
[13] Vaia R A, Lincoln D M, Wang Z-G, Hsiao B S and Krishnamoorti R and Silva A S 2000 Polymer–Clay
Krishnamoorti R 2001 Polymer 42 9975–84 Nanocomposites ed T J Pinnavaia and G W Beall (New
Lincoln D M, Vaia R A and Krishnamoorti R 2004 York: Wiley) pp 315–43
Macromolecules 37 4554–61 Mitchell C A and Krishnamoorti R 2001 Polymer
[14] Wang Z and Pinnavaia T J 1998 Chem. Mater. 10 1820–6 Nanocomposites vol 804, ed R Krishnamoorti and
Wang Z, Lan T and Pinnavaia T J 1996 Chem. Mater. 8 2200–4 R A Vaia (Washington, DC: ACS) pp 159–75
[15] Hoffmann B, Dietrich C, Thomann R, Friedrich C and Mitchell C A and Krishnamoorti R 2002 J. Polym. Sci. B 40
1434–43
Mulhaupt R 2000 Macromol. Rapid Commun. 21 57–61
[25] Garboczi E J, Snyder K A, Douglas J F and Thorpe M F 1995
Zax D B, Yang D K, Santos R A, Hegemann H, Giannelis E P
Phys. Rev. E 52 819–28
and Manias E 2000 J. Chem. Phys. 112 2945–51
[26] Ren J and Krishnamoorti R 2003 Macromolecules 36 4443–51
Wang D Y, Zhu J, Yao Q and Wilkie C A 2002 Chem. Mater.
[27] Krishnamoorti R, Ren J and Silva A S 2001 J. Chem. Phys.
14 3837–43 114 4968–73
[16] Vaia R A and Giannelis E P 1997 Macromolecules 30 8000–8 [28] Ginzburg V V and Balazs A C 1999 Macromolecules 32 5681
[17] Vaia R A and Giannelis E P 1997 Macromolecules 30 7990–9 Balazs A C, Singh C and Zhulina E 1998 Macromolecules 31
[18] Ren J 2002 PhD Thesis Department of Chemical Engineering 8370–81
(Houston: University of Houston) Kuznetsov V D and Balazs C A 2000 J. Chem. Phys. 112
[19] Yurekli K, Krishnamoorti R, Tse M F, McElrath K O, 4365–75
Tsou A H and Wang H-C 2001 J. Polym. Sci. B 39 256–75 Balazs A C, Singh C, Zhulina E and Lyatskaya Y 1999 Acc.
Hsieh A J, Moy P, Beyer F L, Madison P, Napadensky E, Chem. Res. 32 651–7
Ren J and Krishnamoorti R 2004 Polym. Eng. Sci. 44 Balazs A C 2000 Curr. Opin. Colloid Interface Sci. 4 443–8
825–37 [29] Manias E 2003 private communication
[20] Krishnamoorti R and Yurekli K 2001 Curr. Opin. Colloid [30] Wang M S and Pinnavaia T J 1994 Chem. Mater. 6 2216
Interface Sci. 6 464–70 [31] Ren J, Casanueva B F, Mitchell C A and
[21] Ren J, Silva A S and Krishnamoorti R 2000 Macromolecules Krishnamoorti R 2003 Macromolecules 36 4188–94
33 3739–46 [32] Krishnamoorti R and Giannelis E P 2001 Langmuir 17
[22] Graessley W W 1974 Adv. Polym. Sci. 16 1 1448–52

S521

También podría gustarte