Sunteți pe pagina 1din 5

Polymer Degradation and Stability 95 (2010) 672e676

Contents lists available at ScienceDirect

Polymer Degradation and Stability


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

Accelerating effect of montmorillonite on oxidative degradation


of polyethylene nanocomposites
T.O. Kumanayaka, R. Parthasarathy*, M. Jollands
Rheology and Materials Processing Centre, School of Civil, Environmental and Chemical Engineering, RMIT University, Melbourne, Victoria 3001, Australia

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 10 November 2009
Accepted 23 November 2009
Available online 26 November 2009

Polyethylene, one of the most widely used packaging materials, can be made biodegradable by blending
it with biopolymers such as starch and/or pro-oxidants which are metal complexes (e.g. cobalt stearate,
cerium stearate). Recent studies on polyethylene degradation have found that addition of nanoclay,
which is used as a ller in polymer composites mainly to enhance their mechanical properties, also
increases their photo-oxidative degradation. The present study aims to investigate the degradation of
low density polyethylene (LDPE) formulated with nanoclay and evaluate the effect of nanoclay
compositions on the overall photo-oxidation process. Photo-oxidative aging of polyethylene and its
nanocomposites were carried out in a QUV weathering tester for a maximum period of two weeks. The
degradation progress was followed by monitoring the chemical changes of the samples using Fourier
transform infrared spectroscopy (FTIR) and gel permeation chromatography (GPC). The results indicate
that the incorporation of nanoclay signicantly enhances the degradation of polyethylene.
2009 Elsevier Ltd. All rights reserved.

Keywords:
Polyethylene
Photo-oxidation
Nanocomposites
UV irradiation

1. Introduction
During the past 25 years, there has been a continuous increase
in the manufacturing of commodity and packaging plastic products
such as polyolens. Polyethylene is one of the fastest growing
commercial thermoplastic polyolen materials due to its low cost
and desirable properties such as high strength, good barrier properties, light weight, water resistance and higher stability [1]. The
continuous use of polyethylene in packaging and agricultural
applications leads to the generation of a large quantity of plastic
waste every year. The accumulation of plastic waste in the environment is leading to long term environmental and waste
management problems. This has led to an increased environmental
concern in the community which has led to the development of
more environmental-friendly polymeric packaging materials. In
recent years, there has been a focus on development of environmentally friendly degradable polymers [2e4].
The main reasons for the poor degradability of polyethylene are
its hydrophobicity and high molecular weight with larger dimensions of its molecule. However recent studies have found that
polyethylene can be degraded by photo-oxidation [5]. Photooxidative degradation is the process of decomposition of materials

* Corresponding author. Tel.: 61 3 99252941; fax: 61 3 99253746.


E-mail address: rchrp@rmit.edu.au (R. Parthasarathy).
0141-3910/$ e see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2009.11.036

in the presence of UV light which is one of the primary sources of


material damage at ambient conditions [6]. The photo degradation
is initiated with the formation of polymer radicals due to the
breakup of chemical bonds in the polymer chain. UV radiation in
the 290e320 nm range is equivalent to the energy required to
dissociate CeC and CeH bonds and produce free radicals [1]. The
free radicals formed by this cleavage act as initiators for polymer
degradation. Photo-oxidation can also be enhanced by other
methods such as introducing chromophores into the materials,
copolymerizing with a small amount of monomers containing
carbonyl groups, and using transition metal compounds such as
metal stearates and dithiocarbamates [2]. The role of pro-oxidants
in the photo-oxidation of polyethylene has been discussed in
a number of papers [7e9].
Inorganic llers are often added to polyethylene to form
composites or nanocomposites in which the ller serves to enhance
the mechanical properties. Nanocomposites are a new class of lled
polymers in which inorganic llers such as clay platelets at the
nanometre scale are dispersed in a polymer matrix. In recent
studies on polyethylene degradation, it has been found that polyethylene nanocomposites exhibit greater degradation rate as
compared to polyethylene [10,11]. Tidjani and Wilkie reported that
the photo-oxidative degradation of polypropylene nanocomposites
is greater than that of pure polypropylene [10]. Similar results were
reported by Qin and co-workers for polyethylene nanocomposites
[11]. They proposed a mechanism in which the main cause for this

T.O. Kumanayaka et al. / Polymer Degradation and Stability 95 (2010) 672e676

enhancement is considered to be the acidic sites created on the clay


surface by thermal decomposition of ammonium ions in the clay. In
spite of these studies, a comprehensive knowledge on the impact of
clay on the photo-oxidation of polyethylene is still lacking.
The present paper reports a study which investigates the effect
of montmorillonite clay, which can be used to improve the
mechanical properties of polyethylene, on the photo-degradation
of polyethylene. The knowledge obtained from this investigation
will be useful in the design of more environmentally friendly
degradable polymeric materials.
2. Experimental
2.1. Materials
The materials used in this investigation were: low density
polyethylene (LDD203 lm grade) supplied by Qenos, Australia,
maleic anhydride-modied polyethylene (PE-gMA, known also as
Fusabond MX110D) supplied by Dupont, and organo-modied clay
(with trade name Cloisite 15A) supplied by Southern Clay Products.
Cloisite 15A is a natural montmorillonite (MMT) exchanged with
quaternary ammonium salt with interlayer spacing of 31.5 and
a cation exchange capacity of 95 meq/100 g clay.
2.2. Preparation of polyethylene nanocomposites
PE/MMT nanocomposites were prepared using a Brabender
twin screw extruder. Initially LDPE was added with 5 wt% Fusabond
and extruded to facilitate compatibility of LDPE with clay. The
operating temperature of the extruder was maintained at 120, 150
and 190  C from hopper to die section of the extruder and the
operating screw speed was maintained at 60 rpm. Maleic anhydride
grafted polyethylene and clay were then mixed thoroughly and
extruded again. Before the melt-mixing, the clay was dried in
a vacuum oven at 80  C for at least 16 h.
The compositions of different nanocomposites prepared in this
study are shown in Table 1. Thin PE and PE nanocomposites lms of
65e75 mm thickness were prepared using blown lm assembly
equipment made by Strand Palst Maskiner, Sweden. Nanocomposites pellets were dried at room temperature for 24 h before
using them in lm preparation. Films were produced at 200  C (die
exit temperature) at a rate of 0.5 kg/h with an outer die of 40 mm
diameter and 1 mm die gap.
2.3. Morphological characterization
Wide angle X-ray diffraction (WAXD) was used to evaluate the
extent of intercalation of polymer between the clay layers. A Philips
X-ray generator using 30 kV accelerating voltage and 30 mA current
was used in the analysis. Ni ltered Cu Ka (l 0.154 nm) radiation
was used to scan the lm samples from 2q 1.5 e40 intensities at
the rate of 8 /min. The scattering intensities were modied by

673

subtracting the background scattering measured using an empty


sample holder.

2.4. Ultraviolet exposure test


Ultraviolet (UV) exposure tests were carried out in an Accelerated QUV Weathering tester (Q-lab Corp., U.S.A.) which uses xenon
lamps. It reproduces the damage that can occur in outdoor environment by several environmental factors such as sunlight,
temperature and rain. Spectral irradiance in the tester was
according to ASTM G154. Nanocomposites lms were cut into
150  75-mm strips and subjected to 0.68 W/m2/nm intensity at
340 nm at the black panel temperature at 60  C. The 12-h weathering cycle used include 8-h continuous radiation followed by 4-h
condensation of de-mineralized water on the surface of the lm at
50  C. Samples were allowed to dry for 1 h after each testing cycle.

2.5. Fourier transform infrared spectroscopy (FTIR)


The photo-oxidation rate was followed by measuring the
carbonyl absorbance of polyethylene samples using FTIR spectroscopy. The FTIR measurements were performed in a PerkineElmer
2000 spectrometer in the wave number range of 400e4000 cm1.
The surface of the lm samples was in contact with a ZneSe crystal
that has a 45 angle of incidence. Interferograms were obtained
from 32 scans. Results from FTIR analysis were used to determine
carbonyl index (C.I.) which is dened as the absorbance ratio of
carbonyl and methylene groups according to equation (1).

C:I:

Absorbance@1713 cm1
Absorbance@1464 cm1

(1)

C.I. is a measurement of the amount of carbonyl compounds


formed during the photo-oxidation.

2.6. Gel permeation chromatography (GPC)


Molecular weight distribution of PE and PE nanocomposites
samples was monitored by high temperature GPC analysis performed at 140  C. The instrument used was Waters Alliance GPCV
2000 chromatographer equipped with differential refractive index
(DRI) and viscometer detectors. 1, 2, 4-trichlorobenzene (TCB) was
used as the solvent at a ow rate of 1.00 ml/min. The system of
three Styragel HT (4, 5 and 6) columns was calibrated with polystyrene standards with average molecular weight ranging from
1000 to 5 000 000. PE and its nanocomposite samples were dissolved in TCB and ltered through a 0.5 mm polytetrauoroethylene
(PTFE) lter to remove the solid particles. Universal calibration was
applied and the chromatograms were processed using Millennium software. Weight average molecular weight (Mw) was
determined from the analysis.

2.7. X-ray uorescence analysis (XRF)


Table 1
Compositions of the samples used in this study.
Sample name

PE (wt%)

PE-gMA (wt%)

MMT (wt%)

PE
PE-2
PE-3
PE-5
PE-7

100
93
92
90
88

5
5
5
5
5

e
2
3
5
7

wt%
wt%
wt%
wt%

clay
clay
clay
clay

(2C)
(3C)
(5C)
(7C)

Major transition elements present in the montmorillonite clay


were analysed using a Bruker AXS S4 pioneer XR uorescence
spectrometer. Sample disks were prepared by pressing the clay
powder. The X-ray tube was operated at 50 kV and 10 mA and the K
lines were used for the determination of elements. LiF (200) single
crystal was used as the analysing crystal. Its angular divergence was
0.23. The analytical lines obtained for elements were registered by
means of a scintillation detector.

674

T.O. Kumanayaka et al. / Polymer Degradation and Stability 95 (2010) 672e676


Table 2
Carbonyl absorption at 1713 cm1 for PE and its nanocomposites.

Fig. 1. XRD patterns of cloisite 15A (MMT) and PE/MMT nanocomposites.

3. Results and discussion


3.1. Film characterization
X-ray diffraction patterns of cloisite 15A and PE nanocomposites
are shown in Fig. 1. Using the results shown in Fig. 1, the d-spacing
for cloisite 15A was determined (using Bragg's Law, nl 2d sin q) to
be 3.17 nm as 2q was 2.78 . Curves for all nanocomposite samples,
except for the sample with 2 wt% clay, display characteristic peaks
at angles that are smaller than that for pure clay indicating the
formation of intercalated morphology. These results indicate that
the silicates were dispersed in the polymer matrix. But the stacked
structure of the pristine clay (MMT) was retained with only a small
increase in the layer spacing from 3.17 to 3.32 nm for nanocomposites with 3, 5 and 7 wt% clay loadings. The absence of the
characteristic ller peak for the sample with 2 wt% clay indicates
that the clay has nearly exfoliated in the polymer matrix.
3.2. FTIR analysis
Photo-oxidation tests were carried out with PE and its nanocomposites with 2, 3, 5 and 7 wt% clay to investigate the effect of
nanoclay on photo-degradation. During the photo-oxidation,
polymer macromolecules degrade leading to the formation of alkyl
radicals. These unstable alkyl radicals then combine with oxygen to
form hydroperoxides, which can then decompose to produce
alkoxy radicals. These alkoxy radicals may either abstract hydrogen
from the polymer backbone or undergo b-scission. As a result of

Sample name

Absorbance at 1713 cm1

PE
2C
3C
5C
7C

0.2073
0.8103
0.9043
1.0122
1.4928

these propagation steps, the polymer matrix is degraded further


leading to the production of a mixture of carbonyl and hydroxyl
species [11]. The formation of carbonyl and hydroxyl groups can be
monitored using FTIR results as shown in Fig. 2. The peaks observed
at 1713 cm1 are due to the formation of carbonyl groups (C]O)
(Table 2). The effect of photo-oxidation is signicant in all nanocomposites whereas it is negligible in polyethylene. It is also
clear that the level of absorbance increases with increase in clay
concentration. The shape of the peaks for all nanocomposites
is fairly broad suggesting that more than one functional groups is
formed within the C]O family. The peak at 1757 cm1 is assigned
to either isolated carboxylic acids or pre-esters [12]. The peak at
1735 cm1 is assigned to carbonyl absorption of esters whereas that
at 1780 cm1 is assigned to the vibration of lactone [13]. The broad
peaks observed around the wave number of 3000e3600 cm1
indicate the formation of hydroxyls. These peaks indicate that clay
concentration has an inuence also on the formation of hydroxyl
groups. The broad peak could be assigned to free OH groups or OH
groups in carboxylic acids or those in MMT.
The above results indicate that the rate of photo-oxidation as
determined by the heights of the carbonyl and hydroxyl peaks
increases with increasing clay content. These results agree well
with the results reported by Qin and co-workers for the photooxidation of polyethylene nanocomposites. Qin et al. suggested that
the mechanism of photo-oxidation of the polymer does not change
due to the presence of clay [11]. On the other hand, Pandey et al.
concluded that clay affects the photo-oxidation mechanism either
by producing or hindering some oxidation products [17].
The inuence of clay on photo-oxidation could be explained by
considering the decomposition of alkyl ammonium ions in the
organically modied clay (MMT) according to the Hofmann
mechanism as shown below [14].
LSLD NR3 eCH2 eCH2 n eCH3  / LSL HD D CH2
]CHeCH2 nL1 eCH3 D NR3

0.5

0
1500

PE
2C
3C
5C
7C

0.4
Absorbance

1.5
Absorbance

0.5

PE
2C
3C
5C
7C

1641cm-1
0.3
0.2
0.1

2000

2500

3000

3500

4000

-1

Wave number (cm )


Fig. 2. FTIR spectra (1500e4000 cm
aging.

1

) of PE and its nanocomposites after 10 days

0
1600

1620

1640

1660

1680

1700

-1

Wave number (cm )


Fig. 3. FTIR spectra (1600e1700 cm1) of PE and its nanocomposites after 10 days of
aging.

0.7

0.8

0.6

0.7

0.5

0.6

675

PE
2C
5C
7C

0.4

dwt/d(Log M)

Carbonyl Index

T.O. Kumanayaka et al. / Polymer Degradation and Stability 95 (2010) 672e676

0.3
0.2

0.5
0.4
0.3

0.1

0.2
0

10

0.1

15

Time (days)
PE

2C

3C

5C

7C

Fig. 4. Carbonyl index of PE and its nanocomposites irradiated for 14 days.

As a result of this reaction, tertiary amine, olen and acidic sites


are formed on the clay layers. The small peaks observed for PE
nanocomposites at 1641 cm1 in the FTIR spectra shown in Fig. 3
are the characteristic peaks for olen and they conrm that olen is
formed in nanocomposites consistent with the Hofmann reaction.
The olen and the acidic sites formed on clay layers tend to enhance
the formation of free radicals which accelerate the photodegradation.
Carbonyl Index (C.I.) values for PE and nanocomposites are
shown in Fig. 4 as a function of irradiation time. It can be seen that
CI values increase with irradiation time for all polymer samples
indicating all of them are oxidised during the photo-oxidation.
However, the rate of increase in CI values for nanocomposites is
substantially greater than that for PE.
Higher CI values for the nanocomposite with 7 wt% clay as
compared to other nanocomposites also indicate that the increase
in clay concentration enhances the rate of photo-oxidation. The fact
that the nanocomposites with 7 wt% clay (which has an intercalated structure) leads to higher CI values as compared to nanocomposites with 2 wt% clay (which has an exfoliated structure)
indicates that photo-oxidation does not depend much on the
morphology.
3.3. XRF analysis
Naturally present iron (Fe3) ions in the MMT could also affect
the rate of degradation of a polymeric matrix by reducing to ferrous
ions (Fe2) by redox catalysis of the hydroperoxide decomposition

reaction since a clay mineral can act as an electron acceptor and/or


electron donor [15]. X-ray uorescence analysis results for organically modied clay are given in Fig. 5. It proves that MMT contains
Fe in its octahedral sheet.
3.4. Evaluation of molecular weight changes
The changes in molecular weight distribution (MWD) and mass
average molecular weight data for irradiated samples of pure PE
and nanocomposites are shown in Figs. 6 and 7, respectively. It can
be seen that the irradiation causes the MWD curves of the nanocomposites to shift towards the low molecular weight scale indicating the formation of low molecular weight compounds in the
polymer during the photo-oxidation (Fig. 6). After 14 days of irradiation, the mass average molecular weight of PE has decreased
by 65% whereas that of 7C nanocomposite has decreased by 94%
(Table 3). It is also clear that the mass average molecular weight of
the polymer decreases rapidly with increasing clay concentration
but the difference in molecular weight of irradiated nanocomposite
samples are not signicant (Fig. 7). This nding is consistent with
the FTIR results which show that there is no signicant difference in
carbonyl index values of 2C, 5C and 7C nanocomposites irradiated
for 14 days (Fig. 4).

Molecular weight (Daltons)

16000
-1

200000

18000

counts s

Fig. 6. Molecular weight distribution irradiated PE and its nanocomposites, Irradiation


period 14 days.

20000
K Fe

14000
12000
10000
8000
6000
4000

K Fe

2000
0

Log M

10

20

30

40

50

60

2
Fig. 5. X-ray uorescence analysis of MMT.

70

80

180000

0 days

160000

7 days
10 days

140000

14 days

120000
100000
80000
60000
40000
20000
0
PE

2C

5C

7C

Fig. 7. Molecular weights changes of PE and its nanocomposites irradiated at 0.68


W/m2/nm.

676

T.O. Kumanayaka et al. / Polymer Degradation and Stability 95 (2010) 672e676

Table 3
Molecular weight changes of PE and its nanocomposites during photo-oxidation.
Sample

UV exposure time
0 days

PE
2C
5C
7C

7 days

10 days

14 days

Mw (Daltons)

Mn

Mw (Daltons)

Mn

Mw (Daltons)

Mn

Mw (Daltons)

Mn

167
181
184
115

31
33
32
21

158
23
22
22

29 350
4260
4110
3962

78 464
20 106
16 865
9258

12 329
3959
3218
1689

62 194
15 938
9249
6291

9755
3140
1825
1185

310
198
007
197

770
299
560
659

802
181
875
405

The GPC results suggest that nanoclay plays an important role in


the degradation of the polymer by enhancing the breakdown of
high molecular weight species into low molecular weight molecules. Since it is easier for the micro-organisms to utilize low
molecular weight species, it can be expected that such molecular
mass reduction occurred during photo-oxidation will facilitate
biodegradation [6].
On the basis of all the above results, it can be stated that MMT
nano-ller affects the photo-oxidation of PE in many different
ways. The cloisite 15A used is a natural montmorillonite which
contains iron which is one of the transition metals responsible for
photochemical degradation of the polymer [16]. Iron can act as
a pro-oxidant and initiate the polymer oxidation by the well known
photo-Fenton reaction by reducing hydroperoxides into hydroxyl
radicals [15]. This mechanism is consistent with the mechanism
proposed for polypropylene nanocomposites in the literature [16].
Another theory postulated on the role of clay on polymer oxidation
is that it interferes with the diffusion of oxygen in the polymer
matrix enabling it to remain longer in the material thereby
enhancing the rate of oxidation [17].
4. Conclusions
The effect of photo-oxidation on pure PE and its nanocomposites
was investigated using an accelerated weathering tester. The rate of
photo-oxidation of PE was enhanced by the addition of nano-ller.
Results obtained from FTIR and GPC analysis show that, upon
exposure to ultraviolet radiation, the polymer matrix undergoes
chain scission and end up with low molecular weight species. The
rate of photo-oxidation increased with an increase in the clay
concentration due to the presence of Fe in MMT. Also the decomposition of alkyl ammonium ions in MMT creates olens and acidic
sites on the clay surface which can accelerate the radical formation
in the polymer matrix. It is also found that the dispersion state of
the clay particles in the polymer matrix has no inuence on the
photo-oxidative degradation of the polymeric matrix.

References
[1] Singh B, Sharma N. Mechanistic implications of plastic degradation. Polymer
Degradation and Stability 2008;93(3):561e84.
[2] Albertsson A-C, Karlsson S. Degradable polymers for the future. Acta Polymer
1995;46:114e23.
[3] Chiellini E, Corti A, D'antone S, Baciu R. Oxo-biodegradable carbon backbone
polymers e oxidative degradation of polyethylene under accelerated test
conditions. Polymer Degradation and Stability 2006;91(11):2739e47.
[4] Koutny M, Lemaire J, Delort A-M. Biodegradation of polyethylene lms with
prooxidant additives. Chemosphere 2006;64(8):1243e52.
[5] Kemp TJ, McIntyre RA. Inuence of transition metal-doped titanium (IV)
dioxide on the photodegradation of polyethylene. Polymer Degradation and
Stability 2006;91(12):3020e5.
[6] Albertsson A-C, Andersson SO, Karlsson S. The mechanism of biodegradation
of polyethylene. Polymer Degradation and Stability 1987;18(1):73e87.
[7] Roy PK, Surekha P, Rajagopal C, Raman R, Choudhary V. Study on the degradation of low-density polyethylene in the presence of cobalt stearate and
benzil. Journal of Applied Polymer Science 2006;99(1):236e43.
[8] Roy PK, Surekha P, Rajagopal C, Choudhary V. Effect of cobalt carboxylates on
the photo-oxidative degradation of low-density polyethylene. Part-I. Polymer
Degradation and Stability 2006;91(9):1980e8.
[9] Vogt NB, Kleppe EA. Oxo-biodegradable polyolens show continued and
increased thermal oxidative degradation after exposure to light. Polymer
Degradation and Stability 2009;94(4):659e63.
[10] Tidjani A, Wilkie CA. Photo-oxidation of polymeric-inorganic nanocomposites:
chemical, thermal stability and re retardancy investigations. Polymer
Degradation and Stability 2001;74(1):33e7.
[11] Qin H, Zhao C, Zhang S, Chen G, Yang M. Photo-oxidative degradation of polyethylene/montmorillonite nanocomposite. Polymer Degradation and Stability
2003;81(3):497e500.
[12] Philippart J-L, Sinturel C, Arnaud R, Gardette J-L. Inuence of the exposure
parameters on the mechanism of photooxidation of polypropylene. Polymer
Degradation and Stability 1999;64(2):213e25.
[13] Tidjani A. Comparison of formation of oxidation products during photo-oxidation
of linear low density polyethylene under different natural and accelerated
weathering conditions. Polymer Degradation and Stability 2000;68(3):465e9.
[14] Qin H, Zhang S, Liu H, Xie S, Yang M, Shen D. Photo-oxidative degradation of
polypropylene/montmorillonite nanocomposites. Polymer 2005;46(9):3149e56.
[15] Solomon DH. Clay minerals as electron acceptors and/or electron donors in
organic reactions. Clay and Clay Minerals 1968;16:31e9.
[16] Qin H, Zhang Z, Feng M, Gong F, Zhang S, Yang M. The inuence of interlayer cations
on the photo-oxidative degradation of polyethylene/montmorillonite composites.
Journal of Polymer Science: Part B: Polymer Physics 2004;42:3006e12.
[17] Pandey JK, Raghunatha Reddy K, Pratheep Kumar A, Singh RP. An overview on
the degradability of polymer nanocomposites. Polymer Degradation and
Stability 2005;88(2):234e50.

S-ar putea să vă placă și