MMT NR HTTP WWW - Sciencedirect.com Science Montimorilonita NR

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

European Polymer Journal 44 (2008) 3108–3115

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Macromolecular Nanotechnology

Effect of montmorillonite intercalant structure on the cure parameters


of natural rubber
Felipe Avalos a, José Carlos Ortiz a, Roberto Zitzumbo b, Miguel Angel López-Manchado c,
Raquel Verdejo c, Miguel Arroyo c,*
a
Universidad Autonóma de Coahuila, Saltillo, Mexico
b
CIATEC, Omega, 3, León Guanajuato, Mexico
c
Instituto de Ciencia y Tecnologia de Polimeros, CSIC, Juan de la Cierva, 3, 28006 Madrid, Spain
MACROMOLECULAR NANOTECHNOLOGY

a r t i c l e i n f o a b s t r a c t

Article history: Two quaternary phosphonium salts (aromatic and aliphatic) have been used as intercalants
Received 26 March 2008 for Na-montmorillonite and the effect of intercalant structure on clay morphology and nat-
Received in revised form 7 July 2008 ural rubber vulcanization kinetics was investigated. Due to its lower rigid structure the ali-
Accepted 8 July 2008
phatic salt was easier to intercalate into the clay galleries giving rise to a higher interlayer
Available online 23 July 2008
distance and facilitating the rubber intercalation obtaining an exfoliated structure in the
nanocomposite. The vulcanization process was sensibly accelerated by this organoclay
and a higher crosslinking degree was observed in the nanocomposite which gave rise to
Keywords:
Organoclays
materials with improved processing and physical characteristics.
Natural rubber Ó 2008 Elsevier Ltd. All rights reserved.
Vulcanization kinetics
Nanocomposites

1. Introduction The interlayer cations of the montmorillonite can be


easily ion-exchanged with other positive charged atoms
Clays and clay minerals such as montmorillonite, sapo- or organic ions such as quaternary ammonium or phospho-
nite, hectorite, etc., were widely used as filler for rubbers nium salts. The introduction of organic ions into the inter-
and plastics for many years, for saving polymer consump- layer spacing not only render more organophylic
tion and reducing the cost [1]. At present, nanocomposites phyllosilicates but also results in a larger interlayer spacing
based on polymer matrices and organo-layered silicates depending on functionality, packing density and length of
represent an interesting opportunity for the design of high organic molecule [9]. So, organoclays improve the compat-
characteristic materials mainly from the mechanical prop- ibility with a polymer matrix and the intercalation of the
erties and thermal stability points of view [2–5]. polymer within the galleries. The organic chains into the
Two specific characteristics of layered silicates are useful galleries may lay either parallel to the silicate layers, form-
for nanocomposites development. The first one is based on ing mono or bilayers or to radiate away from the surface,
the possibility of these high aspect ratio (L/D relationship) forming mono or even bimolecular ‘‘paraffinic” arrange-
layers [6] to be individually dispersed and the second one ment [10,11], in this case with a higher interlayer spacing.
on their chemical capacity to be modified through cation As above mentioned, quaternary ammonium or phos-
exchange reactions [7]. So, not only their reinforcing effect phonium salts can be used as intercalants for layered sili-
may become apparent but their chemical compatibility with cates. The advantage of using phosphonium instead of
polymer may be considerably improved [8]. ammonium salts is that they give rive to an increase in
the degradation temperature of the organoclay from
250–300 to 350 °C [12]. However, only a few references
* Corresponding author. Tel.: +34 915622900; fax: +34 915644853.
E-mail address: marroyo@ictp.csic.es (M. Arroyo). on these phosphonium salts can be found in the literature.

0014-3057/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.eurpolymj.2008.07.020
F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115 3109

Okamoto et al. [13] used a saponite organically modified heated at 80 °C, was added under strong stirring The pro-
with a commercial quaternary phosphonium salt. Salahud- tonation of the salt was carried out by addition of hydro-
din and Akelah [14] studied the synthesis of phosphonium chloric acid (HCl/quaternary salt molar ratio 1:1). After
salts of styrene–chloromethylstyrene-maleic anhydride 2 h at this temperature and under vigorous stirring, the
terpolymer and their intercalation with montmorillonite. suspension was filtered and washed several times with
Recently, Arroyo et al. [15] synthesized triphenyl vinylben- hot water until complete elimination of chloride anions,
zyl phosphonium chloride and described the treatment of a as was checked by silver nitrate test. Finally the organoclay
bentonite with this salt and its effect on the bentonite was washed with ethanol, filtered and dried at 60 °C over-
morphology. night. The dried organoclay was grinded in a ball mill.
On the other hand, rubber nanocomposites based on nat- Thermogravimetric analysis showed a 27% b.w. modifica-
ural rubber (NR) with layered silicates have received tion of this organoclay named aromatic-MMT.
increasing attention in the last years [16–31]. In fact, inter- The modification of the Na-MMT with tetraoctylphos-
calated and partially exfoliated structures have been suc- phonium bromide was carried out following the above de-
cessfully prepared by several processing techniques, such scribed procedure. In this case, the phosphonium salt was
as vulcanization curing process [17,28,32–38], solution previously dissolved in ethanol and then added to the Na-
blending [39,40] or latex compounding [23,27,41–43]. A MMT suspension under vigorous stirring. Finally, the
natural rubber–organoclay nanocomposite with a fully organoclay was washed several times with water until
exfoliated structure was firstly reported by López-Manch- complete elimination of bromide anions, washed with eth-
ado et al. [35,36]. The nanocomposite was successfully pre- anol, filtered, dried and grinded. Thermogravimetric analy-
pared via vulcanization process. The cure characteristics of sis showed a 31% b.w. modification of this organoclay,
pristine natural rubber (NR) were affected by the organo- named as aliphatic-MMT.

MACROMOLECULAR NANOTECHNOLOGY
clay. So, the optimum vulcanization time was sharply short-
ened with the incorporation of the organoclay which was 2.3. Preparation of NR/clay nanocomposites
attributed to amine groups coming from the organophiliza-
tion of the clay. The addition of the organoclay also resulted Rubber nanocompounds were prepared in an open two-
in a sensible increase in the torque value as compared to the roll mill, at room temperature. The rolls operated at a
pristine NR which indicates that a higher number of cross- speed ratio of 1:1.4. The vulcanization ingredients were
links were formed. This statement was supported by swell- incorporated to the rubber before the addition of the clay,
ing measurements and DSC isothermal studies. and finally, the sulfur was added to the blend. The recipes
Having in mind the above, the main goal of this study is of the studied compounds are compiled in Table 1.
to investigate the effect of the organic phosphonium salt Vulcanization parameters of the nanocompounds were
structure (aliphatic and aromatic) on the morphology, vul- calculated by means of a Rubber Process Analyzer
canization characteristics, thermal stability and tensile (RPA2000, Alpha Technologies), at 170 °C. A frequency of
properties of natural rubber nanocomposites and to ana- 10.49 rad/s and a strain of 2.79% were used.
lyze the structure/behavior relationship of the obtained Vulcanizates were prepared at 170 °C, in a thermofluid
materials. heated press. The vulcanization time of the nanocom-
pounds correspond to the optimum cure time (t97) ob-
tained from the curing curves.
2. Experimental part
2.4. Characterization of the obtained nanocomposites
2.1. Materials
Wide-angle X-ray diffraction (XRD) was used to charac-
Natural rubber was kindly supplied by Malaysian Rub- terize the clays and to study the nature and extent of the
ber under the trade name CV-60 (Mooney viscosity: ML dispersions of the clays in the composites. XRD patterns
(1 + 4), 100 °C: 60). Na-montmorillonite (Na-MMT) with a were collected using a Phillips Diffractometer at the wave
cation exchange capacity (CEC) of 92 meq/100 g was pro-
vided by Southern Clay Products (Gonzalez TX). Triphenyl
vinylbenzyl phosphonium chloride (TVBP), synthesized in Table 1
our laboratories [15] and tetraoctyl phosphonium bromide Recipes of the studied rubber compounds
(TOP) supplied by Aldrich, were used as intercalants. NR NR/pristine- NR/aliphatic- NR/aromatic-
MMT MMT MMT
2.2. Preparation of organoclays Natural rubber 100 100 100 100
Zinc oxide 5 5 5 5
The intercalation of the aliphatic phosphonium salt into Stearic acid 1 1 1 1
the pristine-MMT galleries was carried out through a cat- Sulfur 2.5 2.5 2.5 2.5
MBTSa 1 1 1 1
ionic exchange reaction, following the next procedure:
PBNb 1 1 1 1
100 g of Na-MMT were dispersed in 2000 ml hot water, un- Pristine-MMT – 5 – –
der continuous stirring and maintained over night. Then Aliphatic-MMT – – 5 –
the suspension was heated at 80 °C and a solution (clay/ Aromatic-MMT – – – 5
phosphonium salt molar ratio 1:1.2) of the protonated a
Benzothiazyl disulfide.
quaternary phosphonium salt in 500 ml water, previously b
Phenyl betanaphthylamine.
3110 F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115

0
length, k CuKa = 1.54 Å A, a tube voltage of 40 kV and tube silicate galleries due to the shear stresses that the materi-
current of 25 mA. Bragg’s Law was used to calculate the als undergo during compounding. So, the diffraction peaks
crystallographic spacing (d) for clay and organoclay. of aliphatic-MMT and pristine-MMT shift to lower values
Differential scanning calorimeter measurements and of 2h which can be related with a better dispersion of these
thermal gravimetric analysis were carried out on a Mettler clays in the polymer matrix. In this case, the aliphatic-
Toledo DSC 822, and a TGA/SDTA 851, respectively, at a MMT 2h peak moves from 4.15 to 3.31 and the
heating rate of 10 °C under nitrogen atmosphere. pristine-MMT peak from 7.48 to 7.06. These displacements
The number of active network chain segments per unit correspond to intergallery spacings of 26.7 and 12.5 Å,
of volume (crosslinking density) was determined on the respectively. On the other side, the diffraction peak of the
basis of the rapid solvent-swelling measurements by appli- aromatic-MMT remains practically constant.
cation of the Flory–Rehner equation [44]. The samples The small peak about 2.2 is attributed to the presence of
(approximately 1 g) were swollen in toluene at 30 °C for intercalated rubber into the clay galleries. In the case of the
three days until equilibrium swelling was reached. NR/aliphatic-MMT nanocomposite the peak of the organo-
Maximum tensile strength was measured according to clay shifts towards low angles giving rise to two new peaks
ISO 37-1977 standard specifications with type 2 test spec- which are related with the different arrangement of the
imens, and Shore A Hardness was also measured. intercalant into the galleries, for instance as bilayer and
Transmission electron microscopy (TEM) was used to monolayer arrangements. The presence of this peak, even
evaluate the dispersion and the degree of intercalation in the pristine-MMT, confirms that it may be exclusively
and exfoliation of the clays in the rubber matrix. Samples due to the shear stresses that the composite undergo dur-
were obtained by means of a cryoultramicrotome ing processing, although in the case of the organoclays this
equipped with a diamond knife, and observed on a Tec- peak is more pronounced.
MACROMOLECULAR NANOTECHNOLOGY

nai-20 Microscope using an acceleration voltage of 200 kV.


3.1.2. TGA results
3. Results and discussion The TGA and DTA curves of the pristine and functional-
ized montmorillonites are compiled in Fig. 2. As can be
3.1. Characterization seen in the DTA curve of the pristine-MMT the weight loss
as a function of temperature is given in two steps with
3.1.1. XRD analysis maxima at 300 and 650 °C which correspond to 1% and
The XRD patterns of the different clays used in the pres- 5% weight loss, respectively. According to the structural
ent study are graphically represented in Fig. 1. So, peaks at organization of MMT, it can be assessed that the first step
2h values of 7.48, 4.78 and 4.15 for the pristine-MMT, aro- (200–410 °C) corresponds to the beginning of the dehydra-
matic-MMT and aliphatic-MMT, respectively, can be ob- tion process related to the dehydroxylation of lateral
served, which correspond to intergallery spacing values groups in the MMT whereas the second step (410–
0
of 11.5, 18.5 and 21.2 ÅA, respectively. As can be also de- 710 °C) corresponds to the end of dehydroxylation and
duced from Fig. 1, an additional increase in intergallery bears a change in crystalline structure of the material
spacing of the clays is observed in the nanocomposites, due to collapse of the interlayer spacing [45,46]. On the
which is attributed to some intercalation of NR into the

Fig. 2. TGA (A) and DTA (B) curves of pristine-MMT and


Fig. 1. XRD patterns of clays and composites. organomontmorillonites.
F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115 3111

other hand, the weight loss in the aliphatic-MMT also ap-


pears in two well defined steps with maxima at 430 and
490 °C, in the following intervals: from 244 to 458 °C and
from 458 to 650 °C, respectively. Both decreases corre-
spond to weight losses of 16% and 15%, respectively. Taking
in mind that the TOP+ ion adsorbed on the MMT surface is
spatially symmetric (since it posses four identical organic
substituent, –C8H7, linked to the phosphorous atom) it
can be thought that the mass loss in one or two steps will
be related with the type of bonds originated between the
pristine-MMT and the TOPBr salt during the organophylli-
zation reaction. So, it is clear that the first mass loss
(410 °C) corresponds to the organic material adsorbed on
the external interlamellar oxygen atoms of the MMT sur-
face while the weight loss in the second step (450 °C) is re-
lated with the organic ionic groups jointed to the hydroxyl
groups of the tetra and octahedral layers of the MMT and
to the Lewis acid groups. So, according to several authors
Fig. 3. Vulcanization curves, at 170 °C, of pristine NR and NR/MMTs
[47–49] it can be stated that hydroxyl ends close to the composites with 5 phr of clay.
surface show a typical Brönsted acid character (H+ donor)
while interlamellar hydroxyls show both Brönsted and Le-

MACROMOLECULAR NANOTECHNOLOGY
wis (electron acceptor) acid characters.
with aliphatic-MMT show a maximum torque value of
On the other hand, the first derivative of the aromatic-
13.5 dNm. However, the unfilled natural rubber, and the
MMT thermogram represented in this figure shows that
pristine-MMT and aromatic-MMT composites show very
weight losses of 17%, 5% and 5% take place in three steps,
low torque values, all of them at about 6 dNm. These re-
at 380, 451 and 555 °C, respectively, in the ranges of 250–
sults show that a higher crosslinking degree is obtained
375, 375–650 °C. Taking in mind the existence of hydroxyl
with the aliphatic-MMT clay, and according to the XRD
groups with acid character in the MMT then it is evident that
patterns this clay is better dispersed in the NR matrix than
in the first step the weight loss is due to the organic phase of
the other clays. So, a higher torque is necessary for this
TVBP+ which is absorbed in the Brönsted acid groups while
composite to flow.
the weight loss in the third step is due to the organic content
Vulcanization studies have also been carried out on a
of the TPB+ absorbed on the Lewis acid groups. So, the weight
RPA2000 Rubber Processing Analyzer which enables the
loss in the intermediate step would correspond to the organ-
analysis of both the elastic (S0 ) and viscous (S00 ) compo-
ic ions absorbed on both Brönsted and Lewis acid groups
nents of the torque before the starting of cure reaction.
originally present on the MMT surface. In fact, in the first
The obtained results are graphically represented in Fig. 4.
derivative curve of the aliphatic-MMT, a small shoulder near
So, it has been observed that the viscous component of
the de-absorption peak of the organic ions bonded to the Le-
the aliphatic-MMT composite shows a sudden increase at
wis acid sites is observed.
3.6 min which indicates that vulcanization process is
The weight loss, at lower temperatures than the TOP+, is
attributed to the organic groups of the TVBP+ absorbed on
the MMT surface due to the steric hindrance of the aro-
matic groups bonded to the phosphorous. These groups
are larger than the octyl groups and require less energy
to be eliminated from the MMT surface.
Taking in consideration that functionalization of MMT
with both aromatic and aliphatic quaternary phosphonium
salts was carried out under similar reaction conditions and
that the total weight losses as a function of temperature
were 27% and 31%, respectively, it is evident that the
TOP+ adsorption is more effective than the TVBP+ adsorp-
tion, which can be attributed to electric field or steric ef-
fects of the aromatic groups of the TVBP+. These groups
would delay the adsorption of this quaternary phospho-
nium cation on the MMT surface in comparison with the
TOP+. So, at the same time of reaction, in the case of the
TVBPCl some Lewis acid groups remain unreacted.

3.2. Vulcanization characteristics

The vulcanization curves of the composites are graphi- Fig. 4. Vulcanization parameters, of pristine NR and NR/MMTs
cally represented in Fig. 3. As can be observed, samples composites with 5 phr of clay.
3112 F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115

starting and suggests than aliphatic-MMT behaves as a cat- about 100 °C correspond to water. In the case of the
alyst for NR vulcanization. The vulcanization process of the pristine-MMT this peak is above 100 °C and could be
studied materials starts following the next order: attributed to interlaminar water. However, in the
aliphatic- and aromatic-MMTs this peak is just at 100 °C
NR=aliphatic-MMT > NR=aromatic-MMT
or below that temperature which can be due to the higher
> NR=pristine-MMT P NR hydrophobic character of the organoclays.
The specific times to start vulcanization are 3.6, 4, 4.2
It can be also observed that, at 170 °C, at the beginning
and 4.2 min for NR/aliphatic-MMT, NR/aromatic-MMT,
of the test, the NR/aliphatic-MMT composite shows the
NR/pristine-MMT and NR, respectively. Even though vulca-
lower values of both the viscous and elastic components.
nization process starts earlier in the presence of the orga-
However, 3.6 min. later these values were the highest
noclays, the final torque of the NR/aliphatic-MMT and
among the studied materials. This is attributed to the fact
NR/aromatic-MMT differs by up to 130%. This difference
that at this temperature the organic portion in the ali-
may be attributed to these two factors:
phatic-MMT starts to melt as observed by DSC analysis
(Fig. 5). This melted portion acts as a lubricant, but once
the vulcanization process starts, the elastic behavior of
– A lower intergallery spacing in Ar-MMT, as observed in
the crosslinked rubber chains becomes apparent, so it is
XRD study which can be related with a poorer disper-
evident that this nanoclay is catalyzing the vulcanization
sion of the clay into the rubber matrix.
process. In fact, phosphonium salts are used to accelerate
the crosslinking reactions of epoxidized resins [50,51]. – The rigid structure of the tetraphenyl phosphonium in
As can be seen in Fig. 5, the small shoulder at about comparison to tetraoctyl phosphonium salt, due to the
50 °C in the aliphatic-MMT corresponds to the melting presence of large benzene groups, makes more difficult
MACROMOLECULAR NANOTECHNOLOGY

point of the aliphatic phosphonium salt, according to the the catalytic effect during rubber vulcanization.
supplier. The endothermic peaks of the different clays at
This would explain the differences that can be observed
at the end of the rubber crosslinking process between the
aliphatic and the aromatic phosphonium salt modified
montmorillonite in the composites. Moreover, it is well
known that the latest exerts both a higher steric hindrance
and electric field effects during vulcanization.
Having in mind the above results and considerations, it
was decided to investigate the crosslinking kinetics of
these materials.
In order to obtain the kinetic parameters of vulcani-
zation, the curing torque was measured at different
temperatures. Considering that vulcanization, starting
after the induction time (ti), is a function of curing time,
then
dX=dt ¼ KðnÞð1  XÞn ðfor t > t i Þ ð1Þ
where X is the torque value, t is the time, K is the kinetic
constant. The progress of the reaction is directly related
to the increase in torque value as a function of time:
ð1  XÞ ¼ f ðtÞ ð2Þ
then, when plotting Ln (1  X) vs t (min), a straight line
with negative slope will be obtained and from this slope
the value of Kn can be calculated (Fig. 6).
If cure is carried out at different temperatures then the
activation energy of the curing process can be calculated
by applying the Arrhenius equation, so

KðnÞ ¼ K 0 ðnÞeEa =RT ð3Þ


Ln KðnÞ ¼ Ln K 0 þ Ea =RT ð4Þ

Then, when plotting Ln Kn vs 1/T, the activation energy,


Ea, for rubber vulcanization in the presence of the differ-
ent clays used in this study, can be calculated (Fig. 7). So,
the next Ea values were obtained: 26.69; 24.01; 20.68
and 11.97 cal/mol K, for pristine rubber and Na-MMT/
NR, Ar-MMT/NR and Al-MMT/NR nanocomposites,
Fig. 5. DSC curves of pristine-MMT and organomontmorillonites. respectively.
F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115 3113

can be estimated by application of Flory–Rehner equation


[44]:

½Lnð1  Ur Þ þ Ur þ vU2r  ¼ 0 n½ Ur1=3  ðUr =2Þ ð5Þ


where Ur is the volume fraction of polymer in the swollen
mass, V0 is the molar volume of the solvent (106.2 for tol-
uene) and v is the Flory–Huggins polymer–solvent dimen-
sionless interaction term (the value of v is 0.393 for
toluene) The volume fraction of polymer, Ur, is calculated
by the following expression [52]:

1=Ur ¼ 1 þ ½ðW s qr Þ=ðW i qs Þ ð6Þ


where Wi and Ws are the weights of the rubber samples in
air and in the swollen state, respectively, and qs and qr are
the densities of the solvent (0.87 g cm3 for toluene) and
the rubber (0.92 g cm3), respectively.
The obtained results of the crosslinking density, n, for
Fig. 6. Vulcanization kinetic of aliphatic-MMT/NR nanocomposite.
the studied materials are reported in Table 2. As can be ob-
served, the highest value is obtained in the presence of the
aliphatic-MMT. In this case, the crosslinking density is one
order of magnitude higher than in the other investigated

MACROMOLECULAR NANOTECHNOLOGY
nanocomposites.

3.4. Morphological characterization

TEM micrographs of the studied materials are


shown in Fig. 8. As can be seen, the pristine montmo-
rillonite is heterogeneously dispersed into the natural
rubber matrix. In the NR/Ar-MMT nanocomposite the
Ar-MMT particles are dispersed as tactoids with partial
intercalation of rubber into the clay galleries. Some
broken and folded silicate lamina can be seen what
is attributed to shear stresses during the preparation
of the compounds. However, the NR/Al-MMT nanocom-
posite shows an intercalated and exfoliated structure
with small tactoids, constituted by a few silicate layers
where the rubber is intercalated, and individual silicate
Fig. 7. Activation energy for rubber vulcanization in the composites
layers homogeneously dispersed into the rubber
through Arrhenius equation.
matrix.

As can be seen, the Al-MMT/NR system shows the low- 3.5. Mechanical behavior
est activation energy value which confirms the catalytic ef-
fect of the Al-MMT on the curing of NR. As an example, some of the most important mechanical
properties of the prepared composites are compiled in Ta-
3.3. Crosslinking density ble 2. So, as can be seen, the Al-MMT nanocomposite
shows the best mechanical properties according to its
The effect of the incorporation of the clay and organo- morphology. A slight increase in strength and elongation
clay on the crosslinking density (n) of the natural rubber, is obtained with the Ar-MMT nanoclay.

Table 2
Some physical and morphological characteristics of the studied materials

Filler (phr), Matrix No Filler, NR Na-MMT(5), NR Ar-MMT(5), NR Al-MMT(5), NR


0
d-spacing, A
Å – 12.5 18.5 26.7a
n (mol cm3) 7.6825  105 7.8739  105 7.1723  105 1.7639  104
Max. strength, MPa 9.5 ± .5 7.7 ± .4 7.9 ± .4 27.3 ± 1.0
Max. elongation, % 1116 ± 35 955 ± 30 1066 ± 35 1120 ± 30
Shore A hardness 30.5 ± 1 30.5 ± 1.2 30.0 ± 1 48.2 ± 1.3

n, crosslinking density.
phr, parts per hundred of resin.
a
Partially exfoliated (see Fig. 8).
3114 F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115

4. Conclusions obtained with the tetraoctyl phosphonium salt due to


the more effective adsorption of this salt by the clay
From the above, the next conclusions can be deduced: which is attributed to the flexibility of the organic
substituent.
 An increase in the MMT intergallery distance is obtained  Under the same reaction conditions, a lower degree of
by cationic exchange reaction with the quaternary phos- intercalation is obtained with the aromatic phospho-
phonium salts used in this study. The highest increase is nium salt due to steric effects of the aromatic rings.
 A further increase in the intergallery distance is
obtained in the rubber nanocomposites due to NR inter-
calation during the compounding and vulcanization
processes.
 Vulcanization process is sensibly accelerated by the ali-
phatic salt. This effect is not so clear with the aromatic
salt which may be due to its rigid structure.
 The higher torque value after vulcanization in the Al-
MMT/NR nanocomposite is attributed to the higher
crosslinking density and better dispersion of this clay
in the rubber matrix.
 The Al-MMT/NR nanocomposite shows the lower Ea
value and the higher crosslinking degree which not only
will shorten the vulcanization process but will give rise
MACROMOLECULAR NANOTECHNOLOGY

to improved physical properties in the NR nanocompos-


ite as shown along this study.

Acknowledgements

The authors thank the CICyT, MAT 2004-00825, for


financial support and the Ministerio de Educación y Ciencia
for the concession of a Professor Mobility Fellowship to Dr.
Felipe Avalos.

References

[1] Vaia RA, Jandt KD, Kramer EJ, Giannelis EP. Macromolecules
1995;28(24):8080–5.
[2] Gilman JW. Appl Clay Sci 1999;15(1–2):31–49.
[3] LeBaron PC, Wang Z, Pinnavaia TJ. Appl Clay Sci 1999;15(1–
2):11–29.
[4] Zhu J, Morgan AB, Lamelas FJ, Wilkie CA. Chem Mater
2001;13(10):3774–80.
[5] Ray SS, Okamoto M. Prog Polym Sci 2003;28(11):1539–641.
[6] Giannelis EP, Krishnamoorti R, Manias E. Polymers in Confined
Environments Adv Polym Sci 1999;138:107–47.
[7] Theng BKG. The chemistry of clay-organic reactions. New
York: Wiley; 1974.
[8] Krishnamoorti RVRAGEP. Chem Mater 1996;8:1728–34.
[9] Lagaly G, Beneke K. Colloid Polym Sci 1991;269(12):1198–211.
[10] Lagaly G. Solid State Ionics 1986;22(1):43–51.
[11] Vaia RA, Teukolsky RK, Giannelis EP. Chem Mater
1994;6(7):1017–22.
[12] Hartwig A, Putz D, Schartel B, Bartholmai M, Wendschuh-Josties M.
Macromol Chem Phys 2003;204(18):2247–57.
[13] Okamoto K, Ray SS, Okamoto M. J Polym Sci B Polym Phys
2003;41(24):3160–72.
[14] Salahuddin N, Akelah A. Polym Adv Technol 2002;13(5):339–45.
[15] Arroyo M, Suarez RV, Lopez-Manchado MA, Fernandez JF. J Nanosci
Nanotechnol 2006;6(7):2151–4.
[16] Wang YZ, Zhang LQ, Tang CH, Yu DS. J Appl Polym Sci
2000;78(11):1879–83.
[17] Arroyo M, Lopez-Manchado MA, Herrero B. Polymer
2003;44(8):2447–53.
[18] Sun YH, Luo YF, Jia DM. J Appl Polym Sci 2008;107:2786–92.
[19] Psarras GC, Gatos KG, Karger-Kocsis J. J Appl Polym Sci
2007;106(2):1405–11.
[20] Jurkowska B, Jurkowski B, Oczkowski M, Pesetskii SS, Koval V,
Fig. 8. TEM micrographs of NR/pristine-MMT (top), NR/Ar-MMT (center) Olkhov YA. J Appl Polym Sci 2007;106:360–71.
and NR/Al-MMT (bottom). [21] Li B, Liu L, Luo HX, Luo YF, Jia DM. Acta Polym Sin 2007;5:456–61.
F. Avalos et al. / European Polymer Journal 44 (2008) 3108–3115 3115

[22] Alex R, Nah C. J Appl Polym Sci 2006;102(4):3277–85. [36] Lopez-Manchado MA, Herrero B, Arroyo M. Polym Int
[23] Stephen R, Varghese S, Joseph K, Oommen Z, Thomas S. J Membr Sci 2003;52(7):1070–7.
2006;282(1–2):162–70. [37] Varghese S, Karger-Kocsis J. J Appl Polym Sci 2004;91(2):813–9.
[24] Zukas W, Sennett M, Welsh E, Rodriguez A, Ziegler D, and Touchet P. [38] Wu YP, Ma Y, Wang YQ, Zhang LQ. Macromol Mater Eng
Pentagon Report. 2004. Available from: <http://www. 2004;289(10):890–4.
stormingmedia.us/28/2853/A285334.html>. [39] Lopez-Manchado MA, Herrero B, Arroyo M. Polym Int
[25] Kim W, Kim SK, Kang JH, Choe Y, Chang YW. Macromol Res 2004;53:1766–72.
2006;14(2):187–93. [40] Magaraphan R, Thaijaroen W, Lim-Ochakun R. Rubber Chem Technol
[26] Li SD, Peng Z, Kong LX, Zhong JP. J Nanosci Nanotechnol 2003;76(2):406–18.
2006;6(2):541–6. [41] Valadares LF, Leite CAP, Galembeck F. Polymer 2006;47(2):672–8.
[27] Liu L, Luo YF, Jia DM, Fu WW, Guo BC. J Elastom Plast [42] Varghese S, Gatos KG, Apostolov AA, Karger-Kocsis J. J Appl Polym Sci
2006;38(2):147–61. 2004;92(1):543–51.
[28] Teh PL, Ishak ZAM, Hashim AS, Karger-Kocsis J, Ishiaku US. J Appl [43] Stephen R, Alex R, Cherian T, Varghese S, Joseph K, Thomas S. J Appl
Polym Sci 2006;100(2):1083–92. Polym Sci 2006;101(4):2355–62.
[29] Wu YP, Wang YQ, Zhang HF, Wang YZ, Yu DS, Zhang LQ, Yang J. [44] Flory PJ. Principles of polymer chemistry. Ithaca, New York: Cornell
Compos Sci Technol 2005;65(7–8):1195–202. University Press; 1953.
[30] Kohjiya S, Katoh A, Shimanuki J, Hasegawa T, Ikeda Y. Polymer [45] Bray HJ, Redferm SAT, Clark SM. Mineral Mag 1998;62:647–56.
2005;46(12):4440–6. [46] Xie W, Xie R, Pan WP, Hunter D, Koene B, Tan LS, Vaia R. Chem Mater
[31] Sharif J, Yunus W, Dahlan K, Ahmad MH. Polym Test 2002;14(11):4837–45.
2005;24(2):211–7. [47] Billingham J, Breen C, Yarwood J. Clay Miner 1996;31(4):513–22.
[32] Arroyo M, Lopez-Manchado MA, Valentin JL, Carretero J. Compos Sci [48] Kou MRS, Mendioroz S, Munoz V. Clay Clay Miner
Technol 2007;67(7–8):1330–9. 2000;48(5):528–36.
[33] Bala P, Samantaray BK, Srivastava SK, Nando GB. J Appl Polym Sci [49] Li QB, Hunter KC, East ALL. J Phys Chem A 2005;109(28):6223–31.
2004;92(6):3583–92. [50] Fainleib A, Galy J, Pascault JP, Sue HJ. J Appl Polym Sci
[34] Jia QX, Wu YP, Xiang P, Ye X, Wang YQ, Zhang LQ. Polym Polym 2001;80(4):580–91.
Compos 2005;13(7):709–19. [51] Smith JDB. J Appl Polym Sci 1979;23(5):1385–96.
[35] Lopez-Manchado MA, Arroyo M, Herrero B, Biagiotti J. J Appl Polym [52] Bae YH, Okano T, Kim SW. J Polym Sci B Polym Phys
Sci 2003;89(1):1–15. 1990;28(6):923–36.

MACROMOLECULAR NANOTECHNOLOGY

You might also like