Sie sind auf Seite 1von 10

COMPATIBILISED LDPE/LLDPE/NANOCLAY

NANOCOMPOSITES: I. STRUCTURAL, MECHANICAL, AND


THERMAL PROPERTIES
Farkhondeh Hemati and Hamid Garmabi*
Polymer Engineering Department, Amirkabir University of Technology, No: 242, Hafez Ave., Tehran, Iran

Nanocomposites of LDPE/LLDPE/nanoclay have been prepared using a lab-scale co-rotating twin screw extruder. Using XRD, tensile testing, AFM,
TGA, effects of some material properties and one processing parameter on mechanical and thermal properties of the prepared nanocomposites
were evaluated. Tensile properties indicated that all the prepared nanocomposites exhibited a signicant improvement in elastic modulus and
toughness compared to pristine LDPE/LLDPE blends of the same composition. Thermal stability of nanocomposites in the air and nitrogen
atmosphere was improved. XRD patterns and AFM micrographs showed semi-exfoliated and intercalated microstructures for the prepared
nanocomposites with different orders of mixing.
On a prepare des nanocomposites de polyethyl`ene basse densite/polyethyl`ene basse densite lineaire/nanoargile a` laide dune extrudeuse a` deux
vis a` co-rotation a` lechelle laboratoire. A` laide de la diffraction des rayons X, de lessai de traction, de la microscopie a` force atomique et de lanalyse
gravimetrique par procede thermique, on a e value les effets de certaines proprietes de materiaux et dun param`etre de traitement sur les proprietes
mecaniques et thermiques des nanocomposites prepares. Les proprietes de traction ont indique que lensemble des nanocomposites prepares ont
afche une amelioration importante du module delasticite et de la tenacite comparativement au melange polyethyl`ene basse densite/polyethyl`ene
basse densite lineaire original de meme composition. La stabilite thermique des nanocomposites dans lair et de latmosph`ere dazote a e te
amelioree. Les mod`eles de diffraction des rayons X et les micrographies de microscopie a` force atomique ont demontre des microstructures
semi-exfoliees et intercalaires pour les nanocomposites prepares avec differents ordres de melange.
Keywords: nanocomposites, mechanical properties, thermal properties, AFM, XRD

INTRODUCTION

lends of linear low density polyethylene (LLDPE) and


low density polyethylene (LDPE) have gained signicant
importance in industrial applications. LDPE/LLDPE lms
are often used for packaging lm applications because of their
good processability and excellent mechanical properties. LLDPE
is added to LDPE owing to its superior mechanical properties, for
example, higher tensile strength, elongation at break and impact
properties. The LDPE/LLDPE lms are characterised by reduced
haze and better bubble stability. On the other hand, blending
LLDPE with LDPE enables manufacturers to use the conventional
LDPE lm-blowing apparatus without modication (Yamaguchi
and Abe, 1999; Lu and Sue, 2002).
In spite of the advantages of LDPE/LLDPE blends in lm
applications, one further point that must be considered is
blend miscibility. The miscibility of LDPE/LLDPE blend has
obvious implications on their thermodynamics, rheology and

VOLUME 89, FEBRUARY 2011

nal properties. Schlund and Utracki (1987) found that a blend


of LDPE/LLDPE was immiscible; however, other blends of
LDPE/LLDPE were reported to be partially miscible (Hussein
and Williams, 2004; Fang et al., 2005). Molecular parameters
such as molecular weight (Mw ) and molecular weight distribution
(MWD) and details of molecular structure like long-chain branching content and composition distribution of the branches were
found to have strong inuence on miscibility of LDPE/LLDPE
blends (Hussein et al., 2003). The effect of molecular structure
on the miscibility was investigated by Delgadillo-Velazquez

Author to whom correspondence may be addressed.


E-mail address: garmabi@aut.ac.ir
Can. J. Chem. Eng. 89:187196, 2011
2010 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.20377
Published online 04 August 2010 in Wiley Online Library
(wileyonlinelibrary.com).

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

187

et al. (2008). The authors have found that ZieglerNatta (ZN)


LLDPE/LDPE blends are miscible at low LDPE compositions and
become immiscible at higher. But m-LLDPE (Metallocene)/LDPE
blend has shown miscibility even at rich-LDPE concentrations. For
the special blends with low LLDPE concentrations studied here,
differential scanning calorimetry (DSC) thermograms have shown
three overlapping peak with different peak temperatures, possibly
because of the existence of cocrystallisation of LDPE and LLDPE
indicating partial miscibility (data not shown here). Although the
miscibility of blend plays an important role in nal morphology
and consequent properties of product, the study of it is not our
priority in this work.
As mentioned, LDPE and LLDPE resins are among the most
versatile polymers, but their uses are restricted due to several
drawbacks, including low strength, stiffness and poor heat resistance. To overcome these drawbacks and to prepare materials with
enhanced properties, during past years polyethylene nanocomposites using various inorganic nanollers were prepared (Zhao
et al., 2005; Durmus et al., 2007). The most attractive feature of
nanocomposites is signicant improvement in properties such as
thermal stability, mechanical and barrier properties at low levels
of nanollers loading. It is due to the structure of nanollers with
high surface area and at least one dimension in the nanometer
range (Wang et al., 2002).
There are different methods to prepare nanocomposites, that
is by polymerising a monomer in the presence of clay (in situ
polymerization) or by dispersing the nanoclay into the molten
polymer. However, melt mixing appears to be the most used and
the most important from an industrial point of view. Regarding this, it can be said that twin-screw extrusion has emerged
as the most effective way to prepare polymer-layered silicate
nanocomposites with better improvement in nal properties in
melt blending (Chavarria et al., 2007; Villanueva et al., 2009).
To achieve the expected improvement in properties, preparation of nanocomposites requires extensive delamination of layered
clay structure and uniform dispersion of the resulting platelets
throughout polymer matrix. This is readily achieved with high
surface energy polymers such as polyamides. However, hydrophobic polymers like polyethylene interact only weekly with mineral
surface of nanoclay platelets, making the synthesis of polyethylene nanocomposites by melt compounding considerably more
difcult (Gopakumar et al., 2002). That is because the pristine structure of the silicate clay is highly polar and contains
metal cations (Na+ , K+ ,. . . ) in the interlayer region. Therefore,
organic modication of the clay is necessary in order to increase
their compatibility with the polymer and making the hydrophilic
clay surface more hydrophobic, by replacing the metal ions with
bulky organic ones (such as alkylammonium or alkylphosphonium cations), containing C12C18 chains (Malucelli et al., 2007;
Golebiewski et al., 2008). However, organically modied clay
does not disperse well in the nonpolar polyethylene. It is believed
that adding a polar material to the non-polar polyethylene matrix
like maleic anhydride modied polyethylene (PEMA) as a compatibiliser facilitates the clay dispersion (Chrissopoulu et al.,
2005). Wang et al. (2001, 2002) stated that clay exfoliation occurs
in LLDPEchemically modied silicate nanocomposites when
the maleic anhydride content in the compatibilisers was above
0.1 wt%.
In this study, LDPE/LLDPE/Nanoclay hybrid nanocomposite
was prepared in a lab-scale co-rotating twin screw extruder from
the blend of LDPE/LLDPE, a compatibiliser PEMA and organically
modied bentonite. The objective of this study is to investigate the
inuence of some material properties and one processing param-

188

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

Table 1. Density and MFI of LDPE, LLDPE, and PEMA


Material

Density (, g/cm3 )a

MFI (g/10 min)b

0.9190.9225
0.9190.921
0.912

0.9
1
1.2

LDPE
LLDPE
PEMA
a
b

Reported by the companies.


Measured at 190 C and 2.16 kg.

eter on structural and mechanical properties and thermal stability


of nanocomposites.

EXPERIMENTAL
Materials
A low density polyethylene (LDPE, LH0075) and linear low
density polyethylene (LLDPE, 0209AA) were prepared by
Bandar Emam Petrochemical Co. (Bandar Imam, Iran) and
Arak Petrochemical Co. (Arak, Iran), respectively. A maleic
anhydride modied linear low density polyethylene (PEMA,
Orevac 18302N) with a maleic anhydride content of 0.22.0 wt%
was supplied by Arkema to be used as a compatibiliser. The
density () and the melt ow index (MFI) of these three resins
are listed in Table 1. The nanoclay (Nanol SE3000) prepared
from pristine Na+ -bentonite by ion exchange reaction using
distearyl dimethyl ammonium chloride was kindly provided by

Sud-chemie
Company (Munich, Germany). All these chemicals
were used as received.

Preparation
LDPE/LLDPE/Nanoclay nanocomposites were prepared by a labscale co-rotating twin-screw extruder manufactured by Coperion
Co. (Stuttgart, Germany) (ZSK25, L/D = 40) at temperature prole
of 160190 C at a screw speed of 600 rpm (without addition any
commercial heat stabiliser). Two different orders of mixing were
used: for preparing type I series, after drying, all the components
were fed into the hopper of twin-screw extruder simultaneously.
The second type of nanocomposites were obtained by preparing a 50 wt% nanoclay/PEMA masterbatch at rst step and then
diluting it by LLDPE and LDPE resins due to pre-determined compositions. Extruded strands were chopped up into pellets by a
plastic granulator and dried at 65 C for 15 h. The obtained pellets were injected molded into dumbbell-shaped test specimens
at temperature prole of 140170 C. Blends of LDPE/LLDPE were
prepared using the same conditions to be used as reference compounds.

Characterization
X-ray diffraction technique (XRD)
XRD patterns were obtained by using a Poland Philips Xpert X-ray
diffractometer with Cu K generator, scanning from 1 to 30 at a
step value of 0.02 at room temperature. The injection molded
bars with thickness of 4 mm were used for XRD studies.

Tensile testing
Engineering stress-strain curves were prepared from uniaxial tension tests (ASTM D638) on dumbbell-shaped tensile bars of type
IV using a Galdabini Sun2500 tensile tester (Galdabini, Italy). The
tensile tests were carried out at crosshead speed of 50 mm/min.

VOLUME 89, FEBRUARY 2011

Thermogravimetery analyser (TGA)


TGA were conducted on a Japan Shimadzu TGA-50 thermoanalyser instrument in nitrogen atmosphere at a ow rate of 0.5 mL/s.
The mass and temperature reproducibility of the instrument was
less than 0.5 mg and 1 C. The specimens were heated from ambient temperature to 595 C at a ramp of 10 C/min using an alumina
sample pan. Thermo-oxidation process of samples was determined under the air atmosphere in the same way, as well. All the
thermogravimetric analyses ran two times. The mentioned results
are the average of repeated runs.

Atomic force microscopy (AFM)


Tapping mode AFM micrographs were prepared using a
Dualscope, DME Atomic Force Microscope (Denmark) equipped
with a DS 95-50-E scanner and an AC probe. One of the developments in tapping mode AFM is the use of the changes in phase
angle of the cantilever probe to produce a second image, called a
phase image or phase contrast image. This image often provides
signicantly more contrast than the topographic image and has
been shown to be sensitive to material surface properties, such
as stiffness, viscoelasticity, and chemical composition. In general,
changes in phase angle during scanning are related to energy dissipation during tip-sample interaction and can be due to changes
in topography, tip-sample molecular interactions and deformation at the tip-sample contact (Raghavan et al., 2000; Jiang et
al., 2003). In this study, the phase images of pure blend and its
nanocomposite have been used to study polymer crystallisation
and nanoclay dispersion, as well.
In order to obtain the height and phase AFM micrographs,
specimens were prepared as compression-molded sheets with a
thickness of less than 1 mm. Measurements were carried out in
air at ambient conditions. Height and phase images were recorded
simultaneously.

Figure 1. XRD patterns for organoclay and nanocomposites with


LDPE/LLDPE matrix.

a quadratic model was used by RSM to approximate the responses


continuously.
In the following parts, samples will be introduced as described
below:
1. The code of sample (LLDPE wt%/clay wt%/PEMA wt% /order
of mixing).
2. The content of LDPE (wt%) is equal to 100 minus the sum of
LLDPE wt%, clay wt% and PEMA wt%. In other words, LLDPE
(wt%) = 100 (LLDPE wt% + clay wt% + PEMA wt%).
3. A sample with a code R16 (40:5:6:2) includes 40 wt% of
LLDPE, 5 wt% of nanoclay, 6 wt% of PEMA, 49 wt% LDPE and
has been produced using the second order of mixing.

RESULTS AND DISCUSSION

Experimental Design

X-Ray Diffraction

Using response surface methodology (RSM) of experimental


design, the effect of some material properties and a processing
parameter on tensile properties have been investigated. RSM is
generally a collection of mathematical and statistical techniques
useful for the modelling and analysis of problems in which a
response of interest is inuenced by several variables (Wang et
al., 2001).
The main idea of this methodology is to use a sequence of
designed experiments to obtain an optimal response. In order to
obtain the optimal response, it is sufcient to determine which
explanatory variables have an impact on the response variable(s)
of interest. Once it is suspected that only signicant explanatory
variables are left, then a rather complicated design, such as a central composite design can be implemented to estimate a second
degree polynomial model, which is still only an approximation
at best. The second-degree model has been used to optimize
the tensile properties as response variables of interest in this
study.
The studied variables consisted of three different material
properties and one processing parameter. These factors include:
nanoclay loading, LLDPE content, PEMA level and order of
mixing. Three different levels and two different strategies were
considered. The tensile properties used as responses include: elastic modulus, stress at break, strain at break and dissipated energy
determined using the area under the stress-strain curve.
Analysing different responses, RSM provided graphs representing responses as a function of assumed variables. As mentioned,

The X-ray diffraction (XRD) patterns of nanoclay and its


LDPE/LLDPE nanocomposites are shown in Figure 1. The organically modied clay containing long alkyl groups in the intercalant
agent displays a strong peak at 2.46 corresponding to a d-spacing
of 3.58 nm. Increasing the d-spacing of clay stacks, the 001 diffraction peak shifts to smaller 2 and its intensity diminishes. As it
can be seen in Figure 1, the XRD patterns of nanocomposites (R2
and R4) illustrate that upon mixing with LDPE/LLDPE the 001
peak of nanocomposites completely disappears or shifts to larger
2 indicating a reduction in the d-spacing. According to Zhu et al.
(2001); and Dintcheva et al. (2009), one of the drawbacks of using
organically modied nanoclay is the limit on the processing temperature. Organoclay is often compounded at temperatures below
200 C to prevent degradation and loss of intercalant agent from
clay galleries. At high melt processing temperatures, the alkyl
ammonium salt molecules are removed from galleries causing a
diminution in d-spacing (Xie et al., 2006). It is believed that due
to high shear viscous heat of high viscosity resins used in this
study, the melt temperature could exceed this critical temperature
in some samples and causes this unwanted phenomenon.

VOLUME 89, FEBRUARY 2011

Atomic Force Microscopy


The height and phase images of neat LDPE/LLDPE (60:40 wt%)
blend and sample R16 (40:5:6:2) are illustrated in Figure 2a and b.
Spherulitic morphology of the LDPE/LLDPE blend can be detected
in Figure 2a. In the phase micrograph, the average z-value of
LDPE/LLDPE crystals which appear as bright brown features is

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

189

Figure 2. AFM images of (a) LDPE/LLDPE (60:40 wt%) blend, (b) R16 (40:5:6:2), right images-phase mode, left images-height mode, z-axes for phase
and height micrographs are on the right and left, respectively.

about 2 V (about 60 nm in the height image). These features


can be observed as dark brown areas in Figure 2b.
As it can be seen, phase image of R16 sample exhibit three shade
differences: black for amorphous phase of matrix, brown (light
to dark) for the crystalline phase and matrix near the nanoclay
particles (interface region), and yellowish white for the nanoclay
phase. This micrograph clearly shows that the nanoclay particles
are conned to particular portions and do not uniformly distribute
all over the matrix.
In order to visualise particle size and distribution of the nanoclay particles, these samples were subjected to an extra surface
analysis (Maiti and Bhowmick, 2006). Figure 3ae shows the
section analysis of the neat LDPE/LLDPE blend and R16 samples. Cross-sectional lines were drawn over the scanned images
to observe the size of crystals and nanoclay particles (Figure 3a
and d). To achieve better resolution, larger magnication of R16
was applied (Figure 3d). The cross-sectional proles along these
reference lines are shown as spectra (Figure 3b and e). These line
spectra show some irregularities which indicate the variation in
stiffness along the lines. The peaks of the neat LDPE/LLDPE blend
spectrum in Figure 3a with y-values around 2 V are related to the
crystalline lamellas. But in the other line spectrum (Figure 3e), the
peaks have height about 56 V representing nanoclay particles. As
it can be found in Figure 3b the nano-particles size values differ
in the range of 970 nm.
Although the 001 peak in the XRD pattern of R16 (40:5:6:2)
sample completely disappears (Figure 4), the bimodal peak in
the line spectra of the nanocomposite (indicated in a circle in
Figure 3e) conrms the presence of intercalated stacks in the
bulk reected the fact that the nanocomposite is not completely
exfoliated. In other words, the d-spacing is about 15 nm corre-

190

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

sponded to 2 less than 1 which are not included in the XRD


patterns.

Tensile Testing
Figure 5 shows engineering stress-strain curves for the pure
LDPE/LLDPE (60:40 wt%) blend and LDPE/LLDPE/Nanoclay
nanocomposies with or without compatibiliser. The selected
curves demonstrate the average stress-strain behaviour of the
materials. It can be seen that the addition of nanoclay increased
both modulus and elongation at break and slightly diminished
stress at break. The diminution in ultimate stress can be as a result
of nanoclay hindrance to polymer stress induced crystallisation.
Maximum enhancement in modulus obtained for R16 (40:5:6:2)
is 58%. The tensile properties of the LDPE/LLDPE blends (LLDPE
20 wt% and 40 wt%) and some of the nanocomposites are summarised in Table 2.
The most attractive feature about tensile properties of the studied nanocompositses is the increase in ductility (i.e., increased
elongation at break or increased lost energy determined from the
area under the stress-strain curve). The increase in ductility as
a result of nanoclay addition has been reported by other groups
(Morawiec et al., 2005; Kausch and Michler, 2007). The maximum
increase observed in elongation at break is 40%. Tensile properties of these nanocomposites are a function of how well clay and
polymer interact with each other.
In this study, both nonpolar matrix and hydrophilic clay platelet
surface were modied to improve their interfacial interaction.
Providing a strong interface and achieving a good dispersion
of nanoclay particles in the matrix results in an increase in
input energy dissipation. As Kausch and Michler (2007) declared
a modied crazing mechanism happens in nanocomposites. In

VOLUME 89, FEBRUARY 2011

Figure 3. (a) AFM image of LDPE/LLDPE (60:40 wt%) blend, (b) the cross-sectional spectrum of LDPE/LLDPE (60:40 wt%) blend, (c) AFM image of R16
(40:5:6:2), (d) the larger magnication of R16 AFM micrograph and, (e) the cross-sectional spectrum of R16 sample.

Table 2. Mechanical properties of LDPE/LLDPE blends and the nanocomposites


Specimen
Blend (LLDPE 40 wt %)
R2 (40:1:6:1)
R4 (40:5:6:1)
R6 (40:3:3:1)
R14 (40:1:6:2)
R16 (40:5:6:2)
R20 (40:3:0:1)
Blend (LLDPE 20 wt%)
R1 (20:1:6:1)
R3 (20:5:6:1)

VOLUME 89, FEBRUARY 2011

Stress at break (MPa)

Strain at break (%)

Modulus (MPa)

Dissipated energy (J)

18.7 0.29
17.57 0.36
17.16 1
17.1 0.5
17.23 0.72
18.175 0.5
18.35 0.29
19.5 0.3
18.1 0.3
17.37 0.31

210.75 8.6
181 8.27
194.6 7.34
287.7 8.75
232.8 5.3
253.76 6.5
252.3 0.5
134.7 0.2
163.8 5
162.8 1.5

159.2 4.6
181 2.6
206.56 6.8
182 5.8
196.2 4.5
243.4 5.8
187.3 0.2
167.1 6
190 3
205.1 13

92.8 7
77.5 4.85
80.27 4.5
121.7 0.7
106.45 5
112.9 2.6
108.9 1.3
58.7 0.5
69.2 1.7
70.1 2.2

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

191

Figure 4. XRD patterns of nanoclay, R4 (40:5:6:1) and R16 (40:5:6:2).

this mechanism particle agglomerates are broken up and/or are


debonded from the matrix, creating smaller or larger voids. The
breaking up or debonding stress determines the crazing stress
which is lower than that in unlled matrix. Therefore, slight
agglomeration facilitates crazing and more crazes can be created
with a positive effect on energy dissipation and toughness.
Although the authors emphasised on agglomerates breaking
up and void nucleation as dominate toughness mechanisms,
another group (Sun et al., 2009) mentioned other toughness
mechanisms which have been reported in studies on nanoclay reinforced polymer matrix nanocomposites including crack
deection, nanoller debonding or pull out, matrix deformation,
bridging, and microcracks. However, these energy absorption
mechanisms are signicant as well and could occur in the studied
nanocomposites here. Increasing clay loading causes the formation of larger agglomerates and tends to initiate voids acting as
nuclei (micro-cracking). It can be observed in Figure 6 that maximum dissipation energy occurs in an optimal amount of nanoclay.
As the particle size increases, the total particle/matrix interfacial
surface area available for energy dissipation decrease, but the critical stress for particle/matrix debonding also decreases and this
could be the reason for the peak in this gure. This trend has been
reported by other groups (Chen et al., 2007).
Another important point represented in Figure 6 is slight diminishing of relative dissipated energy (the dissipated energy of

Figure 5. Engineering stress-strain curves of LDPE/LLDPE blend and its


nanocomposites.

192

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

Figure 6. relative dissipated energy (the dissipated energy of


nanocomposite/the one of pristine blend with the same composition)
versus clay loading (wt%) and LLDPE content (wt%).

nanocomposite/the one of pristine blend with the same composition) with increasing LLDPE content. The shear forces in the
melt-compounding process cause break-up of large-sized agglomerates. Further exfoliation of mineral layers is caused by polymer
chain intercalation between mineral layers. The extent of further
exfoliation of the mineral layers is determined by the interaction between the polymer matrix and the mineral layers and
the amount of entropy reduction of matrix macromolecules during melt intercalation. Being more elastic than LDPE, LLDPE has
greater segmental motion. As a result, LLDPE chains have to eliminate larger entropy reduction barrier to intercalate clay galleries.
Therefore, at higher LLDPE loading nanoclay intercalation is suppressed and d-spacing reduces. This can be veried in Figure
7 where the XRD patterns of the nanocomposites R3 (20:5:6:1)
and R4 (40:5:6:1) show the shifting of 001 peak to larger 2 as a
consequence of increasing LLDPE content.
As can be found in Table 2 and Figure 8, the tensile properties of nanocomposites prepared with the second order of mixing
are better than the type I samples. The same trend has been
found by other authors (Zhu and Xanthos, 2004; Etelaaho et al.,
2009). In feeding strategy of type II, better dispersion and distribution of nanoclay particles are achieved. According to different
nanocompositess XRD patterns with dissimilar feeding strategy

Figure 7. XRD patterns of nanoclay, R3 (20:5:6:1) and R4 (40:5:6:1).

VOLUME 89, FEBRUARY 2011

shown in Figure 4, the 001 diffraction peak of nanocomposite


prepared with second feeding strategy completely disappeared.
The AFM micrographs of R16 (40:5:6:2) and R4 (40:5:6:1) conrm the XRD results. The aforementioned AFM result (Figure 3)
has shown that the morphology of R16 is semi-exfoliated. But
as one can see in Figure 9, the AFM height and phase image of
R4, nanoclay aggregates (yellowish white features) are clearly evident. Higher magnication of phase image (Figure 9c) shows that
nanoclay stacks are intercalated. The line mapping spectrum of
this image (Figure 9d) can help in discerning nanoclay intercalation. Peaks with the height more than 5v representing nanoclay
platelets are located next to each other with short distances
indicating matrix chains diffused into nanoclay galleries. These
results can be explained by the following benets of masterbatchbased nanocomposite production: longer exposure to shear forces
and longer interaction time between the nanoclay, compatibiliser
and matrix through two mixing cycles. A longer residence time in
extruder gives more effective time for the shear forces to exfoliate
clay particles and also enables more thorough polymer penetration between the nanoclay platelets as well as better activity of
the compatibiliser.
As it is reported, to achieve the exfoliated and homogenously
dispersed PE/clay nanocomposites, PEMA plays a very important
role (Morawiec et al., 2005; Xie et al., 2006). It is observed that
increasing PEMA level can improve the dispersion and exfoliation
of nanoclay and thus, it can enhance the load transfer capability
and reinforcing efciency of nano-particles. It seems this inuence has a certain limit. One can see in Figure 10 that after this
limit the relative modulus (the modulus of nanocomposite/the
one of pristine blend with the same composition) of nanocomposites diminishes. According to Sheng et al. (2004), exfoliated
nanocomposites do not always represent the highest amount of

Figure 8. relative modulus (the modulus of nanocomposite/the one of


pristine blend with the same composition) versus order of mixing for two
different clay loadings (wt%).

modulus experimentally. The authors stated that fully exfoliated


single silicate layers and two-layer particles have a greater tendency to be curved, compared to particles with more layers. This
increased bending compliance and resulting angular misorientation will further reduce the efciency of fully exfoliated particles
in enhancing the composite modulus. Therefore, the exfoliated

Figure 9. (a) AFM height image of R4 (40:5:6:1), (b) AFM phase image of R4, (c) the larger magnication of R4 AFM phase micrograph and, (d) the
cross-sectional spectrum of R4.

VOLUME 89, FEBRUARY 2011

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

193

Figure 12. TGA curves of LDPE/LLDPE blend and its nanocomposites in


air atmosphere.
Figure 10. relative modulus (the modulus of nanocomposite/the one of
pristine blend with the same composition) of nanocomposites versus clay
loading (wt%) and PEMA level (wt%).

nanocomposites obtained at high level of PEMA may not be the


optimal choice to achieve the concerned mechanical properties.
As Figure 10 shows addition of PEMA higher than 6 wt% can
reduce the elastic modulus.

Thermogravimetric Analysis
Thermogravimetric analysis providing information in course of
thermal degradation of the neat LDPE/LLDPE blend and its
nanocomposites were performed in the atmosphere of air and
nitrogen. Figure 11 shows the TGA prole of LDPE/LLDPE
(80:20 wt%) blend and its nanocomposites including R1
(20:1:6:1), R3 (20:5:6:1), R16 (40:5:6:2), and R24 (40:5:0:1) samples. This gure demonstrated that thermal stability of these
materials is enhanced compared to that of the virgin blend. Table
3 shows the data for the thermal degradation of all these containing: Tonset , a measure of the onset of the degradation, T50% , the
midpoint of the degradation, the temperature at which 50% mass
lost occurs, Tendset , another measure of thermal stability shows
the end point of degradation, and LM200450 C , the amount of lost
mass at temperature range from 200 to 450 C.

Figure 11. TGA curves of LDPE/LLDPE blend and its nanocomposites in


nitrogen atmosphere.

194

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

As seen in Table 3, there is a 55% reduction in mass loss for


R3 (20:5:6:1) in comparison with pure LDPE/LLDPE blend at
the temperature range from 200 to 450 C. The remarkable reduction in mass loss of this nanocomposite result from the fact that
during the combustion the silicate layers with high aspect ratios
are able to migrate toward the material surface, followed by a
char barrier layer formation which could sustain high temperatures and hinder the heat and mass transfer efciently. It can
be observed in Table 3 that at the initial stage of the degradation, some of the LDPE/LLDPE nanocomposites degrade faster
than pure blend. As Haurie et al. (2007) stated Nanol Se3000
experiences a mass loss of 30 wt% due to the decomposition of
organic modier indicating the grafting level of this modier on
nanoclay platelet surface. Owing to the alkyl ammonium modier
decomposition following the Hoffmann elimination reaction, the
degradation of nanocomposite could start at lower temperatures.
After decomposition of organomodier, created acid sites would
catalyst the thermal degradation of polymer matrix (Zhao et al.,
2005; Dintcheva et al., 2009). One can see in Figure 11 that R1
(20:1:6:1) and R16 (40:5:6:2) with better exfoliation (according
to disappearing the 001 peak of XRD patterns of the nanocomposites) initially decompose at lower temperatures in comparison
with R3 (20:5:6:1), which means more alkylammonium and clay
stacks are present on the surface of nanocomposites. Increased
organomodier and clay content on the surface of sample lead to
a more pronounced catalytic effect. In nanocomposites without
compatibiliser, as a result having poor dispersion of clay like R24
(40:5:0:1), a continuous protective layer is not formed depending on less amount of char which is unable to link together the
clay platelets. Hence, they collapse to form an insulator and mass
transport barrier (Zanetti et al., 2004). As demonstrated in Figure
11 the thermal stability of this nanocomposite does not improve.
Temperature dependence of weight loss and its rst derivative with respect to temperature (DTG) during heating in air for
LDPE/LLDPE (80:20 wt%) blend and its nanocomposites including R1 (20:1:6:1) and R3 (20:5:6:1) are shown in Figures 12 and
13, respectively. The results of TGA of these samples in air atmosphere are collected in Table 4.
As can be seen in Figure 12, thermal-oxidative stability of
nanocomposites is enhanced. This enhancement is more significant than the one in nitrogen atmosphere as consequence of
the higher rate of the formation a clay-riched barrier layer which
can hinder the diffusion of atmospheric oxygen into the material.

VOLUME 89, FEBRUARY 2011

Table 3. TGA data (average), in nitrogen, for LDPE/LLDPE blend and its nanocomposites
Specimen
Blend 1 (LLDPE 20 wt%)
R1 (20:1:6:1)
R3 (20:5:6:1)
R16 (40:5:6:2)
R24 (40:5:0:1)

Tonset ( C)

Tendset ( C)

T50% midpoint ( C)

LM200450 C (%)

462
458
472
460
452

495
500
505
505
498

475
476
485
481
474

24
24.2
10.8
21.4
24.8

Table 4. TGA data (average), in air, for LDPE/LLDPE (80:20) and its nanocomposites
Specimen
Blend 1 (LLDPE20 wt%)
R1 (20:1:6:1)
R3 (20:5:6:1)

Tonset ( C)

Tendset ( C)

T50% midpoint ( C)

LM200450 C (%)

370
367
394

447
480
487

406
423
436

86.4
67
61.6

exfoliated and intercalated morphologies were obtained in the


prepared nanocomposite with different feeding protocols. XRD
results revealed the alkylammonium salt degradation during melt
intercalation. The enhanced tensile properties, most notably,
improved ductility and modulus were obtained as a result of
good interfacial interaction. During the thermal degradation of
nanocomposites, improved thermal stability was observed. This
enhancement is mainly based on the protective charred layer
formed by organoclay catalysis dehydrogenation of PE molecules,
this layer is a mass transport and thermal barrier. Due to the
higher rate of the protective layer formation and hindrance of
oxygen diffusion, the improvement of thermal properties is more
signicant in air atmosphere in comparison with the one in nitrogen atmosphere.
Figure 13. DTG curves of LDPE/LLDPE (80:20) and its nanocomposites
in air atmosphere.

Under nitrogen, the decomposition of the organomodier makes


the silicate layers less organophilic and PEMA deacetylaton
reduces the polarity of the matrix. So, polymer-nanoclay compatibility is reduced and segregation of the two phases occur leading
to a further decrease in the interlayer distance. In air, partial
oxidation of polymer chains triggers an increase in matrix polarity thereby enhancing polymer-nanoclay compatibility, improving
the interaction of the layered silicate and facilitating the clayriched barrier layer formation (Laoutid et al., 2009).
However, the general trend of degradation in the air is the
same as what observed in the nitrogen atmosphere. Although the
decomposition of R1 (20:1:6:1) with better exfoliation initiates
in lower temperature because of the roll of clay and its intercalant agent as catalysts, the exfoliation of clay stacks leads to
a reduced mass loss rate shown in Figure 13. As it is demonstrated in Figure 13, the weight loss intensity of the main peak is
dramatically decreased in R1 indicating the accelerated process of
the formation of clay protective char. One can observe in Table 4
that the maximum increase in Tonset and Tendset are 24 and 40 C,
respectively.

CONCLUSIONS
Compatibilised LDPE/LLDPE/Nanoclay nanocomposites with different compositions and order of mixing were prepared using a
co-rotating twin-screw extruder. According to AFM images semi-

VOLUME 89, FEBRUARY 2011

ACKNOWLEDGEMENTS

The authors would like to acknowledge to Sud-chemie


and
Arkema companies for supplying the nanoclay and PEMA, respectively and also, the authors would like to thank S. Salehinia and
M. Bastani for their experimental support.

REFERENCES
Chavarria, F., R. K. Shah, D. L. Hunter and D. R. Paul, Effect of
Melt Processing Conditions on the Morphology and
Properties of Nylon 6 Nanocomposites, Polym. Eng. Sci. 47,
18471864 (2007).
Chen, J., Z. Huazg and J. Zhu, Size Effect of Particles on the
Damage Dissipation in Nanocomposites, Composite Sci.
Technol. 67, 29902996 (2007).
Chrissopoulu, K., I. Altintzi, S. H. Anastasiadis, E. P. Giannelis,
M. Pitsikalis and N. Hadjichristidis, Controlling the
Miscibility of Polyethylene/Layered Silicate Nanocomposites
by Altering the Polymer/Surface Interactions, Polymer 46,
1244012451 (2005).
Delgadillo-Velazquez, O., S. G. Hatzikiriakos and M. Sentmanat,
Thermorheological Properties of LLDPE/LDPE Blends:
Effects of Production Technology of LLDPE, J. Polym. Sci.
Part B: Polym. Phys. 46, 16691683 (2008).
Dintcheva, N. T., S. Al-Malaika and F. P. La Mantia, Effect of
Extrusion and Photo-Oxidation on Polyethylene/Clay
Nanocomposites, Polym. Degrad. Stabil. 94, 15711588
(2009).

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

195

Durmus, A., A. Kasgoz and C. W. Mascosko, Linear Low


Density polyethylene (LLDPE)/Clay Nanocomposites, Part I:
Structural Characterization and Quantifying Clay Dispersion
by Melt Rheology, Polymer 48, 44924502 (2007).
Etelaaho, P., K. Nevalainen, R. Suihkonen, J. Vuorinen, K.
Hanhi and P. Jarvela, Effects of Direct Melt Compounding
and Masterbatch Dilution on the Structure and Properties of
Nanoclay-Filled Polyolens, Polym. Eng. Sci. 49, 14381446
(2009).
Fang, Y., P. J. Carreau and P. G. Laeur, Thermal and
Rheological Properties of mLLDPE/LDPE Blends, Polym.
Eng. Sci. 45, 12541269 (2005).
Golebiewski, J., A. Rozanski, J. Dzwonkowski and A. Galeski,
Low Density Polyethylene-Montmorillonite Nanocomposites
for Film Blowing, Eur. Polym. J. 44, 270286 (2008).
Gopakumar, T. G., J. A. Lee, M. Kontopoulou and J. S. Parent,
Inuence of Clay Exfoliation on the Physical Properties of
Monmorillonite/Polyethylene Composites, Polymer 43,
54835491 (2002).
Haurie, A., A. I. Fernandez, J. I. Velasco, J. M. Chimenos, J. L.
Cuesta and F. Espiell, Thermal Stability and Flame
Retardancy of LDPE/EVA Blends Filled With Synthetic
Hydromagnesite/Aluminium Hydroxide/Montmorillonite and
Magnesium Hydroxide/Aluminium
Hydroxide/Montmorillonite Mixtures, Polym. Degrad. Stabil.
92, 10821087 (2007).
Hussein, I. A. and M. C. Williams, Rheological Study of the
Inuence of Branch Content on the Miscibility of Octene
m-LLDPE and ZN-LLDPE in LDPE, Polym. Eng. Sci. 44,
660672 (2004).
Hussein, I. A., T. Hameed, B. F. Abu Sharkh and K. Mezghani,
Miscibility of Hexene-LLDPE and LDPE Blends: Inuence of
Branch Content and Composition Distribution, Polymer 44,
46654672 (2003).
Jiang, Y., J. J. Zhou and L. Li, Surface Properties of
Poly(3-Hydroxybutyrate-co-3-Hydroxyvalerate) Banded
Spherulites Studied by Atomic Force Microscopy and
Time-of-Flight Secondary Ion Mass Spectrometry, Langmuir
19, 74177422 (2003).
Kausch, H. H. and G. H. Michler, Effect of Nanoparticle Size
and Size-Distribution on Mechanical Behavior of Filled
Amorphous Thermoplastic Polymers, J. Appl. Polym. Sci.
105, 25772587 (2007).
Laoutid, F., L. Bonnaud, M. Alexandre, J. M. Lopez-Cuesta and
P. H. Dubois, New Prospects in Flame Retardant Polymer
Materials: From Fundamentals to Nanocomposites, Mater.
Sci. Eng. 63, 100125 (2009).
Lu, J. and H. J. Sue, Morphology and Mechanical Properties of
Blown Films of Low Density Polyethylene/Linear Low
Density Polyethylene Blend, J. Polym. Sci. Part B: Polym.
Phys. 40, 507518 (2002).
Maiti, M. and A. K. Bhowmick, New Insights Into Rubber-Clay
Nanocomposites by AFM Imaging, Polymer 47, 61566166
(2006).
Malucelli, G., S. Ronchetti, N. Lak, A. Priola, N. T. Dintcheva
and F. P. La Mantia, Intercalation Effects in
LDPE/O-montmorillonites Nanocomposites, Eur. Polym. J.
43, 328335 (2007).
Morawiec, J., A. Pawalak, M. Slouf, A. Galeski, E. Piorkowska
and N. Krasnikowa, Preparation and Properties of
Compatibilized LDPE/Organo-Modied Montmorillonite
Nanocomposites, Eur. Polym. J. 41, 11151122 (2005).

196

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

Raghavan, D., X. Gu, T. Nguyen, M. VanLandingham and A.


Karim, Mapping Polymer Heterogeneity Using Atomic Force
Microscopy Phase Imaging and Nanoscale Indentation,
Macromolecules 33, 25732583 (2000).
Schlund, B. and L. A. Utracki, Linear Low Density Polyethylene
and Their Blends: Part 3. Extensional Flow of LLDPEs,
Polym. Eng. Sci. 27, 380386 (1987).
Sheng, N., M. C. Boyce, D. M. Parks, G. C. Rutledge, J. I. Abes
and R. E. Cohen, Multiscale Micromechanics Modeling of
Polymer/Clay Nanocomposites and the Effective Clay
Particle, Polymer 45, 487506 (2004).
Sun, L., R. F. Gibson, F. Godaninejad and J. Suhr, Energy
Absorption Capability of Nanocomposites: A Review,
Composite Sci. Technol. 69, 23922409 (2009).
Villanueva, M. P., L. Gabedo, E. Gimenez, J. M. Lagaron, P. D.
Coates and A. L. Kelly, Study of the Dispersion of Nanoclays
in a LDPE Matrix Using Microscopy and In-Process Ultrasonic
Monitoring, Polym. Test. 28, 277287 (2009).
Wang, K. H., M. H. Choi, C. M. Koo, Y. S. Choi and I. J. Chung,
Synthetic and Characterization of Maleated
Polyethylene/Clay Nanocomposites, Polymer 42, 98199826
(2001).
Wang, K. H., M. H. Choi, C. M. Koo, M. Z. Xu, I. J. Chung and
M. C. Jang, Morphology and Physical Properties of
Polyethylene/Silicate Nanocomposites Prepared by Melt
Intercalation, J. Polym. Sci. Part B: Polym. Phys. 40,
14541463 (2002).
Xie, Y., D. Yu, J. Kong, X. Fan and W. Qiao, Study on
Morphology, Crystallization Behaviors of Highly Filled
Maleated Polyethylene-Layered Silicate Nanocomposites, J.
Appl. Polym. Sci. 100, 40044011 (2006).
Yamaguchi, M. and S. Abe, Mechanical, Thermal and
Flammability Properties of Polyethylene/Clay
Nanocomposites, Appl. Polym. Sci. 74, 31353159 (1999).
Zanetti, M., P. Bracco and L. Costa, Thermal Degradation
Behavior of PE/Clay Nanocomposites, Polym. Degrad. Stabil.
85, 657665 (2004).
Zhao, C., H. Qin, F. Gong, M. Feng, S. Zhang and M. Yang,
Mechanical, Thermal and Flammability Properties of
Polyethylene/Clay Nanocomposites, Polym. Degrad. Stabil.
87, 183189 (2005).
Zhu, L. and M. Xanthos, Effects of Process Conditions and
Mixing Protocols on Structure of Extruded Polypropylene
Nanocomposites, J. Appl. Polym. Sci. 93, 18911899 (2004).
Zhu, J., F. M. Uhl, A. B. Morgan and C. A. Wilkie, Studies on
the Mechanism by Which the Formation of Nanocomposites
Enhances Thermal Stability, Chem. Mater. 13, 46494654
(2001).

Manuscript received November 30, 2009; revised manuscript


received March 10, 2010; accepted for publication March 22, 2010.

VOLUME 89, FEBRUARY 2011

Das könnte Ihnen auch gefallen