Vous êtes sur la page 1sur 15

European Polymer Journal 61 (2014) 285299

Contents lists available at ScienceDirect

European Polymer Journal


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

Comparative assessment of miscibility and degradability on


PET/PLA and PET/chitosan blends
A.M. Torres-Huerta a,, D. Palma-Ramrez b, M.A. Domnguez-Crespo a, D. Del Angel-Lpez a,
D. de la Fuente c
a

Instituto Politcnico Nacional, CICATA-Altamira, Km. 14.5 Carretera Tampico-Puerto Industrial Altamira, 89600 Altamira, Tamps, Mexico
PMTA of CICATA-Altamira, IPN, Km. 14.5 Carretera Tampico-Puerto Industrial Altamira, 89600 Altamira, Tamps, Mexico
c
Centro Nacional de Investigaciones Metalrgicas, CENIM (CSIC), Av. Gregorio del Amo 8, 28040 Madrid, Spain
b

a r t i c l e

i n f o

Article history:
Received 4 August 2014
Received in revised form 18 October 2014
Accepted 23 October 2014
Available online 6 November 2014
Keywords:
PET/PLA
PET/chitosan
Blends
Miscibility
Degradability

a b s t r a c t
This work reports the synthesis and miscibility of PET/PLA and PET/chitosan blends as well
as their degradation in real soil environment (6 months) and in accelerated weathering
(1200 h). For this purpose, commercial polyethylene terephthalate (PET) and recycled
PET (R-PET) were used as polymer matrixes and extruded with different amounts of polylactic acid (5, 10 and 15 wt-%) or chitosan (1, 2.5 and 5 wt-%) to form laments. Different
characterization techniques such as X-ray diffraction (XRD), Fourier transform infrared
spectroscopy (FTIR), differential scanning calorimetry/thermogravimetric analysis (DSC/
TGA) and scanning electron microscopy (SEM) were used before and after degradation process. The results indicate weak interactions between blend components suggesting secondary bonds by hydrogen bridges or by electrostatic forces. The miscibility of chitosan in both
PET matrixes is lower in comparison with PLA; the saturation of PLA into polymer matrixes
was reached up to an amount of 10 wt-% whereas longer amounts of 5 wt-% of chitosan
become rigid and brittle. The best performance in the miscibility and degradation process
was found for PET/chitosan (95/5) which is comparable with commercial bottles of BioPET
under similar experimental conditions.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
The long-lasting petroleum polymers have been widely
used provoking that the waste of this kind of polymers
takes a very long time to be broken down. Nowadays, this
indiscriminate use of petroleum-based polymers has
caused a big pollution problem [1,2]. To reduce this
problem, it has been used biodegradable polymers from
renewable sources like collagen, keratin, gluten, milk proteins, soy proteins, polysaccharides like starch, cellulose
derivatives, chitosan, alginate, carrageenan, pectins. These
Corresponding author. Tel.: +52 8332600125x87510; fax: +52
8332600125x87521.
E-mail
addresses:
atorresh@ipn.mx,
atohuer@hotmail.com
(A.M. Torres-Huerta).
http://dx.doi.org/10.1016/j.eurpolymj.2014.10.016
0014-3057/ 2014 Elsevier Ltd. All rights reserved.

biodegradable polymers have a short-lifetime because of


are ideal for short-time applications such as; disposable
packages, agricultural mulches, horticultural pots, etc
[36]. They are also naturally degradable when disposed
in the environment. Despite its advantages, many of these
kinds of polymers exhibit poor thermal stability, low steam
and gas barrier and low mechanical properties, making
them unsuitable for other applications [7,8]. Therefore,
the general trend is to combine the mechanical, barrier
and thermal properties of petroleum based polymers with
the biodegradability properties of renewable polymers,
resulting in the production of polymeric materials with
controlled lifetime. The designed materials must be resistant during their use and must have short time degradation
at the end of their useful life [4,9].

286

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

The most favorable packaging material for disposable


soft drink bottles is polyethylene terephthalate (PET), a
kind of semicrystalline, thermoplastic polyester with high
strength and transparency properties as well as excellent
barrier properties. Unfortunately, most of these beverage
bottles are used only once and then discarded, which inevitably generates serious environmental problems (white
pollution) [10]. Therefore, recycling the discarded PET
polymer along with obtaining biodegradable PET-based
blends are efcient approaches to reduce the resources
consumption and to protect the environment at the same
time [11]. The recycling of post-consumer packaging materials into direct food contact packaging applications was
not possible, because of the lack of knowledge about the
contamination of packaging polymers during rst use or
recollection. However, for PET the situation is much favorable: due to its inert character, recycling technologies have
been developed to establish a bottle-to-bottle recycling of
post-consumer PET bottles [12].
On the other hand, most biodegradable polymers are
thermoplastics (e.g. poly(lactic acid), poly(hydroxyalkanoate), poly(vinyl alcohol)) [9]. Among them, poly(lactic acid)
(PLA) is a bio-based polymer obtained from renewable
sources mainly from corn and starch [13]. PLA is an aromatic polyester and has several applications, for example,
is used for lms, extrusion-thermoformed containers and
medical applications for tissue engineering, bone reconstruction and controlled delivery systems [14]. The use of
PLA in beverage bottles is limited due to its poor oxygen
barrier and low mechanical properties [15]. Another interesting biodegradable polymer is chitosan, a biopolymer
derived from chitin, a natural compound from crustacean
shells; chitosan has the ability to form semipermeable
lms and, in recent years, the efforts have been intensied
to develop chitosan lms and its application in food packing [16]. Biodegradable copolymers of PET and aliphatic
polyesters have been synthesized, such as poly(lactic acid),
poly(b-hydroxyalkanoate), poly( e -caprolactone) and
poly(butylene succinate) in order to obtain a degradable
polymer with a faster degradation rate [17,18].
Additionally, physical mixtures of conventional and
biodegradable phases have been studied [19]. To our best
knowledge, few researches are focused in determine the
effects on physicochemical, structural and morphological
properties as well as degradation time of blends PET/PLA
or PET/Chitosan. In this work, the issue has been investigated from different perspectives and the results are
discussed in the terms of the quantities of biodegradable
polymers that were added in two different matrixes commercial PET and recycled PET during extrusion process.

2. Experimental
2.1. Materials and processing
During the rst set of experiments, it was used a commercial PET (CLEARTUF-MAX2, lot no. 1008-03219) provided as pellets by M&G Polymers Company whereas in
the second step recycled PET (R-PET) was obtained from
discarded bottles after they were washed, dried and cut

into akes. The polylactic acid pellets, PLA-2002D (containing 4.4 wt-% in average of isomer D), (batch: YA0828b131)
and chitosan (low molecular weight, with a deacetylation
degree P75%) were purchased from NatureWorks LLC,
USA and Sigma Aldrich, respectively. Before processing,
the raw materials were dried at 60 C during 24 h in an
oven (Thermolyne). Different amounts of PLA (5, 10 and
15 wt-%) or chitosan (1, 2.5 and 5 wt-%) and commercial
PET or R-PET were hand mixed previous to extrusion process. It is important to mention that, initially, we tried to
add into the polymer matrix the same quantities of chitosan or PLA (5, 10 and 15 wt-%), unfortunately, we observed
that synthesized samples with chitosan become rigid and
brittle, regarding the chitosan polymer is a brittleness
material, however, this characteristic depend if the material is derived of fungal biomass, crustacean shell and
insect cuticles [20,21]. Thus, after several experiments we
found that an optimal percentage to evaluate chitosan is
less than 5 wt-%. Blends with a laments shape (1 mm in
diam.  200 cm length) were obtained in a single-screw
extruder (C.W. Brabender) with L/D ratio of 25:1 and four
heating zones: feeding (225 C), compression (237.5 C),
distribution (260 C), and nally, the extrusion die
(225 C).
2.2. Characterization
Structural characterization of the polymer blends was
carried out using a Bruker D8 Advance diffractometer from
2h = 560 (Cu Ka, k = 0.154 nm) and a rate of 1.5 /min.
The Fourier Transform Infrared Spectroscopy (FT-IR)
spectra were recorded with a Nicolet FT-IR spectrometer
(Magna System 550) equipped with an attenuated total
reectance (ATR) accessory between 2000 and 650 cm1
at an optical resolution of 4 cm1 (40 scans).
Simultaneous thermal analysis was carried out in a
Labsys Evo, Setaram instrument, which was used in the
DSC/TGA conguration with the sample and reference crucibles made of a-Al2O3. Sample amount in the crucible was
about 10 mg. The samples were rstly heated at 40 C and
hold for 2 min and subsequently, the measurements were
carried out in the range of 40300 C to evaluate thermal
degradation under argon atmosphere with a heating rate
of 10 C/min. Then, the samples were hold at 300 C for
2 min followed by a cooling with the same rate. It is well
known that peak temperature is inuenced by the scan
rate; however, in this case such displacement was considered neglected. It is also important to mention that DSC
heating was intentionally evaluated from 40 C to 300 C
in order to observe the total degradation of these samples.
After accelerating weathering tests, DSC/TGA analysis were
also conducted on the samples with similar procedure and
heating rate but using a temperature range from 40 C to
500 C.
In order to dissolve the biopolymer phase and to evaluate the dispersion in the different matrixes (PET and R-PET)
as well as evaluate the morphology of the as-prepared
samples, the polymer blends were solubilized in chloroform and acetic acid for PLA and chitosan, respectively.
The dissolved samples were observed by SEM to analyze
the cross section of the blends; before characterization,

287

Since, the nal properties of the materials strongly


depend on the induced microstructure which can be governed by the complex thermo-mechanical history; it is of
prime interest to observe and to understand the development of the crystalline phase. Fig. 1ac shows X-Ray diffraction patterns of the polymer blends with chitosan or
PLA starting from commercial PET and recycled PET at different weight ratios. As a reference, it is also presented the
diffractograms of raw materials. The commercial PET
(Fig. 1a) displays a well-dened triclinic structure; in
agreement with PDF # 00-050-2275 le, the strong peaks
are observed at 2h degrees of 16.19, 17.2, 21.3, 22.36
 1; 0 1 0; 1
 1 1; 1
 10
and 25.5 which correspond to 0 1
and (1 0 0) planes, respectively. On the other hand, R-PET
shows a diffraction pattern with a wide peak. In general,
the diffraction peaks of semicrystalline polymers are broad
and made up of the amorphous phase and reections of the
crystalline planes. The diffraction pattern of R-PET reveals
a semicrystalline structure with a peak characteristic
reection corresponding to the plane with a Miller index
(1 0 0) at 2h = 25.5 [23].
Selected XRD patterns of chitosan shows two peaks, at
8.14 and 18.58, which are commonly refer as a semicrys-

(100)

(110)

(002)

(100)

(210)

(102)

(300)

(112)

(111)

(201)

(203)

(110)

PLA

20

30

40

50

60

(100)

(110)

(111)

(011)

(b)

(010)

2 (degrees)

PET/PLA 95/5
PET/PLA 90/10
PET/PLA 85/15
R-PET/PLA 95/5
R-PET/PLA 90/10
R-PET/PLA 85/15

10

20

30

40

50

60

(100)

(110)

(111)

(c)

(010)

2 (degrees)

Intensity (a.u.)

3.1. Characterization of as-prepared polymer blends after


extrusion

{(200)

10

The polymer blends were buried in a commercial


compost and exposed to the environment in Altamira,
Tamaulipas Mxico for 6 months where the average
temperatures varies from 30 C to 40 C with an average
of relative humidity of 80%. To analyze the effect on the degradation by adding the biopolymers, DSC/TGA analysis
were recorded to compare with the results of the as-prepared blends.
3. Results and discussion

Chitosan

(011)

2.4. Degradation in soil

R-PET

Intensity (a.u.)

Accelerated weathering tests were carried out in a QUV


accelerated weathering chamber Model QUV/Se which
uses UV-B lamps (313 nm and 0.63 W/m2) as radiation
source. During these experiments, PET, R-PET, commercial
bottles of BioPET (B-PET), PLA and all the as-prepared
blend samples were exposed to the UV radiation using an
exposure time of 1200 h with UV/Condensation cycles of
8/4 h at 60 C/50 C for their further characterization.
Additionally, the average weight loss mass was evaluated
during 300 h, 600 h, 900 h and 1200 h. General process to
obtain commercial bottles of BioPET (B-PET) considers
the reaction of bio-terephthalic acid with pretroleumderive ethylene glycol or bio-terephthalic acid in combination with bio-ethylene glycol [22].

PET

Intensity (a.u.)

2.3. Accelerated weathering

(a)

(111)

the samples were dried at 40 C for 24 h and then coated


with an AuPd thin lm. SEM micrographs were acquired
using SEM JEOL 6300 series apparatus operating at 15 kV.

(011)
(010)

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

PET/Chitosan 99/1
PET/Chitosan 97.5/2.5
PET/Chitosan 95/5
R-PET/Chitosan 99/1

R-PET/Chitosan 97.5/2.5
R-PET/Chitosan 95/5

10

20

30

40

50

60

2 (degrees)
Fig. 1. X-ray diffraction patterns for (a) Raw materials (b) PET/and R-PET/
PLA (c) PET/and R-PET/chitosan.

talline phase. Different works reported only one peak for


chitosan at 2h ca. 20; however, the crystallinity or more
packing in the main chain of our commercial chitosan
was modied using a deacetylation process. In such case,
the rst reection is associated with two different kind of
crystals; according to Hwang et. al. [24] the rst peak at
510 has a unit cell characterized by a = 7.76, b = 10391,
c = 10.30 and b = 90 whereas the second peak at
1825 was characterized by a = 4.4, b = 10.0, c = 10.30
and b = 90 [24,25]. PLA samples display an orthorhombic
structure with a strong peak of two overlapped signals
ca. 16.5 as well as lower intensity peaks around, 19 and
22. The intensities match with the (1 1 0), (2 0 0), (2 0 3),

288

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

(2 0 1), (1 1 1) and (1 0 2), (2 1 0) planes, respectively; PDF #


00-054-1917 le.
The XRD patterns of polymer blends with PLA or chitosan were also analyzed and the results are shown in Fig. 1b
and c. From the X-Ray diffraction patterns, it is seen that
the spectra are quite similar resembling the XRD PET pattern. The observed reections could indicate a weak interaction when the chitosan or PLA are added to the PET
matrixes. It can be also observed that the signal of main
peak at 25.5 (1 0 0) shows a slightly decrease with the
relation of PLA or chitosan; the same trend is detected
 1 and (0 1 0) planes. Similar results
for the intensity of (0 1
have been previously reported; i.e. the increasing of the
amount of biopolymer reduces the intensity of the diffraction peaks [25], characterizing the spectra as semicrystalline phases.
The most important point to highlight is that during
extrusion process a recrystallization of R-PET blends was
obtained. It is well known that the semycristalline polymers have a metastable nanophase structure, where the
various nanophases can be occur like crystal, liquid, glass
or mesophases. The structure is determined by the selforganization, crystallization and vitrication and it is
established during the material processing depending of
its thermal and mechanical history [26,27]. Thus, the
blends crystallized after the extrusion process as a result
of the absence of the cooling system in the extraction zone
of the die, which stimulates the formation of polymeric
crystals.
FT-IR spectra were recorded in order to investigate the
chemical structure of the raw materials and to study the
characteristic signals of the blends after extrusion.
Fig. 2ac shows the FT-IR spectra of raw materials and
as-obtained polymer blends at the proposed compositions.
FT-IR spectra show typical absorption bands of amorphous
or semi-crystalline PET similar to those reported in the literature, except for the variation of the background due to
the lm thickness dependent optical effects caused by
reections from polymer surface and polymer/substrate
interface. FT-IR spectra of PET and R-PET (Fig. 2a) show
the main absorption bands as follows: at 3626 cm1 is
the OAH stretching, 3432 OAH stretching of ethylene glycol end group, 3060 cm1 due to the aromatic CAH stretching, 29652906 cm1 is correlated with the aliphatic CAH
stretching, 25581961 cm1 aromatic summation band,
1723 cm1 (C@O stretch), 16191510 cm1 aromatic skeleton stretching band, 14601341 CH2 deformation band,
12661102 cm1 C(O)AO stretching of ester group,
1018 cm1 (1,4 aromatic substitution), 963 cm1 OACH2
stretching of ethylene glycol segment in PET, 869 cm1
CAH deformation of two adjacent coupled hydrogens on
an aromatic ring and nally, at 730 cm1 is presented the
band associated with the out of plane deformation of the
two carbonyl substituents on the aromatic ring.
On the other hand, the FTIR spectrum of PLA was classied into ve regions, which corresponds to the following
peaks band assignment: ACAHA stretching (2994.2 and
2985 cm1), AC@O ester carbonyl (1746.2 cm1), ACAHA
deformation (1451.3, 1383 and 1358 cm1), ACAOA
stretching (1266, 1178, 1127, 1077 and 1039 cm1) and
ACACA stretching (867 cm1). In chitosan, the NAH

stretching occurs in the 3438 cm1 region overlapping the


OH stretch from the carbohydrate ring. The other peaks
band assignment corresponds to ACAHA stretching in
2884 cm1, amide I, amide II and NAH bending vibrations
of amine and amide II (1656, 1597 cm1), ACH2 bending
(1425 cm1), ACH3 symmetrical deformation (1386 cm1),
CAN stretching vibrations overlap the vibrations of carbohydrate ring ACAOACA (1154 cm1) and skeletal vibration
of CAO stretching (1029 cm1) [13,2834].
FTIR spectra of PET/PLA and R-PET/PLA blends are
shown in Fig. 2b; where it is seen a strong absorption band
between 1768 cm1 and 1670 cm1, attributed to the
superposition of the carbonyl stretching of PET and PLA.
It is well known that a reaction between these materials
resulted in signicant alterations in the bands, normally
associated with the formation of new bonds; in this
respect, some researchers have been reported the shift of
the bands as result of esterication in polymers blends of
Poly carbonate (PC)/PET, Bisphenol A-PC /Poly trimethylene terephthalate and in blends of poly vinylphenol
(PVPh)/poly vinylpyrrolidone (PVP) [3537]. The nondisplacement of the commented bands can be related with
another kind of interactions between polymers, probably
hydrogen bonds interaction.
In the combinations produced with PET/chitosan or
R-PET/chitosan, it was not possible to observe the chitosan
phase or an important displacement in the wavenumber or
bonds (Fig. 2c). This might be due to the used of low chitosan amounts which are under the limit detection of the
equipment (65%). Recent reports demonstrated that the
hydrophilic character of chitosan can provoke a strong
interaction between the components forming the polymer
blend [38,39]. However, in our case FT-IR spectra did not
show appreciable modications in the evaluated range
(Fig. 2a), suggesting that chitosan is physically integrated
to the polymer matrix.
One of the most common methods used to estimate
polymerpolymer compatibility is to determine the glass
transition temperature (Tg) of the blend and compare with
the Tg of the component polymers. If one of the components is crystalline, the reduction in melting temperature
(Tm) can be used to investigate the blend compatibility
[4043]. Then, estimating the changes in the Tm of the
blends, it is possible to study the miscibility of the blends
if one of the components is crystalline in nature.
Table 1 displays the results of DSC thermograms of the
raw materials and polymer blends with different quantities of PLA or chitosan. Specically, PET is characterized
by a glass transition temperature (65140 C) and melting
temperature (240265 C) [44,45]. However, the commercial PET used in this work displayed a glass transition temperature about 82 C and two melting temperatures at
126 C and 243 C. The rst Tm is a result of the additives
that are commonly used during PET processing, which it
is expected do not affect the degree of crystallinity,
whereas the second Tm matched well with the value for
PET [44].
From these results, it can be noticeable that Tg cannot
be detected in all compositions, which can be explained
by considering that PET can be prevented from crystallizing by very fast cooling, and so obtained in completely

289

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

(b)
(a)

C-C CO 1,4 CH

C=O

C-H

Absorbance (%)

PET
C-C C-C CH CO1,4
2
CH
C=O
CH

C-H

R-PET
C=O

C-H CO

C-O

C-C

C-H

PET/PLA 95/5

Absorbance (%)

= Phenyl group

PET/PLA 90/10

PET/PLA 85/15

R-PET/PLA 95/5

PLA
chitosan

O-H and N-H


C-H

Amide I and II

N-H
CH2

C-N

C-O

C=O PLA

3500

3000

2500

2000

1500

C=O PET

PET signals

R-PET/PLA 85/15

COC

4000

R-PET/PLA 90/10

3000

1000

2500

2000

1500

1000

-1)

-1)

Wavenumber (cm

Wavenumber (cm

(c)

Absorbance (%)

PET/chitosan 99/1

PET/chitosan 97.5/2.5

PET/chitosan 95/5

R-PET/chitosan 99/1

R-PET/chitosan 97.5/2.5

R-PET/chitosan 95/5

3000

2500

2000

1500

1000

Wavenumber ( cm -1)
Fig. 2. ATR- FTIR for (a) Raw materials (b) PET/and R-PET/PLA (c) PET/and R-PET/chitosan.

Table 1
Thermal transitions obtained for as-prepared selected blends using PLA and
chitosan with PET and R-PET.
Sample

Tg (C)

Tm (C)

Tc (C)

PET
R-PET
PLA
Chitosan
PET/PLA 95/5
PET/PLA 90/10
PET/PLA 85/15
R-PET/PLA 95/5
R-PET/PLA 90/10
R-PET/PLA 85/15
PET/chitosan 99/1
PET/chitosan 97.5/2.5
PET/chitosan 95/5
R-PET/chitosan 99/1
R-PET/chitosan 97.5/2.5
R-PET/chitosan 95/5

82

69
88

126, 243
245
161
188
250
250
250,158
250
250
250,158
248
249
251
248
250
250

201
205
126

200
202
205
198
201
202
197
197
203
208
207
208

amorphous form [45]. However, in our laboratory setup,


the cooling was carried out at room temperature which
allows the formation of more crystalline regions.
Crystallinity of PET is usually induced by thermal crystallization and/or by stress or strain induced crystallization.
Specically, as it is the case, thermally induced crystallization occurs when polymer is heated above Tg and not
quenched rapidly enough. In this condition, the polymer
turns opaque due to spherulitic structure generated by
thermal crystallization aggregates of non-oriented polymers [44].
A slightly increase in the melting point of PET or R-PET
can be obtained by adding both biopolymers and remain
fairly constant ca. 250 C when the biopolymers are miscible in the matrix. Interestingly, the saturation of PLA into
polymer matrixes was reached up to an amount of
10 wt-%, longer quantities than this value provoke two
phases in the polymer blend which in turn show a second
melting temperature (158 C). Similar works highlighted

290

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

the importance of the crystallinity on polymers blends,


indicating that stable values in the melting point can be
reached for higher quantities of PLA in PET/PLA amorphous
blends produce from casting solution [46]. However, in our
study miscible phases are only obtained with amounts
under 15 wt-% using semi-crystalline biopolymer (PLA).
The endothermic peak observed at 126 C for commercial
PET is missed in the DSC spectra of blends extruded at
260 C, corroborating the presence of small quantities of
additives.
Table 1 also shows that by increasing the amount of PLA
into the PET or R-PET matrixes slightly increases the crystallization temperature (Tc) from 198 C to 205 C. Similar
tendency can also see for PET/Chitosan or R-PET/Chitosan
polymer blends with a range of Tc between 197 C and
208 C. Comparing the crystallization temperatures values
in the polymer blends, it can be mentioned that all the
samples have an afnity to preserve Tc value of the PET
matrix (PET or R-PET) and the small differences in the thermal properties can be associated with a slightly decreased
of the molecular weight of PLA during the processing blend
or to the capacity of chitosan to act as nucleating agent
promoting a faster crystallization of PET in combination
with the crystallinity of the materials.
Dissolution of selected phases in a polymer blend with
suitable solvents is an etching method for morphology
evaluation of polymer blends [47]. Chloroform and acetic
acid etching allows the removal of PLA and chitosan to
improve the analysis and reveal the morphology of dispersed biodegradable polymer. Fig. 3a-l shows typical
SEM observations of PET/PLA, R-PET/PLA, PET/chitosan
and R-PET/chitosan. By comparison, SEM micrographs of
the different compositions of PET with PLA are also shown
(Fig. 3bg). All compositions exhibit a droplet (holes)matrix with heterogeneous structure; i.e. homogenously
dispersed PLA domains are found in PET matrix. The mean
diameter of the PLA agglomerates is ranged from 0.02 lm
to 1.9 lm (in PET) and 0.52.1 lm (in R-PET). In contrast
to PET/PLA blends, the behavior of PET/chitosan blends
was quite different. From Fig. 3hl, cavities were not
observed in ratios of 99/1 and 97.5/2.5. Additionally, PET/
Chitosan 95/5 was observed to have some agglomerates
of chitosan which were not dissolved by acetic acid.
Thus, green polymers were integrated in PET and R-PET
matrixes uniformly but the size of agglomerates that form
the laments varies with the amount of biopolymer. These
holes after etching demonstrate that such integration is
established of semi-spherical shape with a size uctuating
between 0.4 and 4.31 lm. Similar morphology has also
been observed with previous reports in different polymer
blends using chitosan as biodegradable component [48].
3.2. Polymer blends degradation under accelerated
weathering
It has been well recognized that the weight loss of polymeric blends can indicate the resistance to degradation
process in a material. Thus, weight loss percentage has
been used to estimate the degradation of as obtained polymers. Total weight loss% was calculated from the following
equation:

Weight loss%

Wi  Wr
Wi

where Wi and Wr represent the initial and residual weights


of the specimens, respectively.
Fig. 4 shows the trend of the weight loss% of as-prepared samples using PLA or chitosan and monitored during
1200 h under accelerated weathering. For comparison, the
raw materials (except chitosan powders) and an additional
sample that consist of a commercial BioPET specimen
(B-PET) were evaluated.
The evolution of weight loss indicate that only 600 h of
UV irradiation were required to initiate the degradation of
PLA, at this time, the green polymer became brittle and it
was difcult to handle as a consequence of being fragmented and degraded by UV light from lamps [49]. In general at 600 h, the blends showed slightly more weight loss
percent in comparison with the raw materials.
Intervals between 1.12 and 2.23 wt-% of weight loss in
the blends were found depending on the amount of biopolymer and kind of matrix (commercial or recycled PET). The
results can divided in two tendencies; (i) the degradation
process of polymer blends augment as the biopolymer concentration increases independently of the matrix; and (ii)
the highest weight loss of polymer blends was obtained
when it is used a recycled PET matrix, which may be associated to the loss properties during the reprocessing. The
polymer reprocessing provoked that the blend displayed
weaker and fragile material more susceptible to degradation than commercial PET [5052]. Additionally, it is widely
known that amorphous regions in polymers degrade more
easily than crystalline zones [53,54], i.e. in semicrystalline
polyesters, degradation rst occurs in the amorphous
domains and subsequent in the crystalline regions [55].
The weight loss of the amorphous R-PET was approximately
1.40% after 1200 hours which was higher than that obtained
for commercial PET (ca. 0.97%).
On the other hand, it is noteworthy that there were not
found important differences between the weight loss of RPET (1.4%) and B-PET (1.5%), in spite of the last one is
obtained from natural resources.
Weight loss percentages were used to calculate the degradation rate according to Eq. (2)

Degradation rate

KW
ATD

where K is a constant of conversion units to mm/year, W is


the weight loss (g), A is the area exposed (cm2), T is the
time (h) and D is the density of the material (g cm3).
The degradation rate of PET/PLA, PET/chitosan, R-PET/PLA
and R-PET/chitosan blends are shown in Fig. 5 corroborating the weight loss and FT-IR analyses.
By comparing a six year natural weathering study with a
20,000 h accelerated weathering report, it has been shown
that there is a rough relationship of approximately 1000 h
of accelerated QUV weathering being equal to one year of
natural exposure. On the other hand, a rough correlation
used in the paint and coatings industry is 5001500 h of
accelerated exposure equaling approximately 1 year of real
life exposure [56]. Thus, a relationship of 1000 h

291

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

(a)

PET
10 m

(b)

(d)

(c)

PET/PLA 95/5

5 m

(e)

PET/PLA 90/10

5 m

(f)

R-PET/PLA 95/5

5 m

(h)

5 m

R-PET/PLA 90/10

(k)

PET/chitosan
97.5/2.5

5 m

R-PET/chitosan
97.5/2.5

R-PET/PLA 85/15

5 m

(j)

5 m

PET/chitosan 95/5

5 m

(m)

(l)

R-PET/chitosan 99/1 5 m

5 m

(g)

(i)

PET/chitosan 99/1

PET/PLA 85/15

5 m

R-PET/chitosan 95/5 5 m

Fig. 3. Morphological features of selected blends using PLA and chitosan with PET and R-PET.

accelerated weathering equaling to a year of natural weathering is used as a conservative data in this work [57].
Table 2 indicates the lifetime prediction of the synthesized materials studied in this work. The results estimate
that commercial PET will totally decompose in about
125 years whereas R-PET and B-PET decomposes in 76
and 86 years, respectively. As previously mentioned, this
is consistent with the strong dependency of being
semicrystalline B-PET and amorphous R-PET. From overall
samples, R-PET blends with chitosan (95/5) and PLA (85/
15) were found to degrade faster than the others compositions after 45 and 54 years, respectively.

3.2.1. Characterization of degraded polymer blends subjected


to weathering chamber
It is well known that environmental factors, such as
temperature, UV radiation and humidity are the main
causes of the degradation in polymer materials [58].
Generally, the degradation is noticed as changes in the
physical integrity and the loss of the structural properties due to molecular bond scission [59]. Structural
changes in all polymer blends were monitored by analyzing the region of absorption bands (between
2000 cm1 and 650 cm1) where susceptible bonds to
degradation are located.

292

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

3,0
B-PET
R-PET
PET
PET/PLA
R-PET/PLA
PET/chitosan
R-PET/chitosan
PLA

Weight loss (%)

2,5

2,0

1,5

95/5
97.5/2.5

85/15
85/15
99/1
90/10
95/5
B-PET
R-PET
95/5
97.5/2.5
90/10
95/5
PET

1,0

99/1

0,5
PLA

0,0
0

200

400

600

800

1000

1200

1400

Time (h)
Fig. 4. Weight loss% of PET, R-PET, B-PET, PLA and blends under
accelerated weathering.

R-PET/chitosan 95/5
R-PET/chitosan 97.5/2.5
R-PET/chitosan 99/1
PET/chitosan 95/5

Blends

PET/chitosan 97.5/2.5
PET/chitosan 99/1
R-PET/PLA 85/15
R-PET/PLA90/10
R-PET/PLA 95/5
PET/PLA 85/15
PET/PLA90/10
PET/PLA 95/5

0,0

1,0x10

-3

2,0x10

-3

3,0x10

-3

4,0x10

-3

Degradation velocity (mm/year)


Fig. 5. Degradation rate of blends under accelerated weathering.

Table 2
Lifetime prediction of raw materials and blends.
Sample

Lifetime prediction (year)

PET
R-PET
B-PET
PET/PLA 95/5
PET/PLA 90/10
PET/PLA 85/15
R-PET/PLA 95/5
R-PET/PLA 90/10
R-PET/PLA 85/15
PET/Chitosan 99/1
PET/Chitosan 97.5/2.5
PET/Chitosan 95/5
R-PET/Chitosan 99/1
R-PET/Chitosan 97.5/2.5
R-PET/Chitosan 95/5

125
86
76
107
103
76
91
58
54
143
93
60
56
55
45

Fig. 6 shows the FTIR spectra of (a) PET/PLA, (b) R-PET/


PLA, (c) PET/chitosan and (d) R-PET/chitosan blends after
being exposed to accelerate weathering for 1200 h. Using
commercial PET or R-PET as matrix in the polymer blends
and PLA in different weight ratios (Fig. 6a and b), the following observations can be drawn: (i) it is seen that C@O
bands were divided into multiple peaks; probably, due to
the characteristics of each matrix, i.e. the vibrations are
more intense in the commercial PET than that observed
for R-PET. This splitting is commonly used to distinguish
between the crystalline (lower frequency) and amorphous
(higher frequency) phase [60]; (ii) the broader signal
observed around 1700 cm1 indicates the beginning of
photo-oxidation and degradation in the amorphous polyester backbones [61]. The peak in 1747 cm1 at higher
wavenumber is encountered more dened than the band
in the amorphous phase; therefore it can be inferred that
the crystalline phase related to the C@O prevails after
accelerated weathering. Amorphous phase in the polymer
blends is less densely packed and therefore more susceptible to degradation than crystalline regions [62].
The C@O stretching band of the amorphous part of PLA
after the treatment also appears at higher wavenumber
(1760 cm1) [63]. In carbonyl containing compounds the
(C@O) peaks are well known to shift to higher wavenumbers, as the electron-withdrawing effect of the /-substituent is larger [64]. This may explain the observations where
crystalline structure has shorter bonds length than in typical amorphous structures. Characteristic ester bonds
appear at 1318 cm1 and 1180 cm1 for PET and PLA,
respectively; which become broad and weak after degradation process due to the chain scission of the CAO bonds
caused by the UV light and humidity. In the same way,
the common CH2 bands (vibration of ethylene units) that
initially appeared at 1410 cm1 and 1340 cm1 are joined
to form a new broad peak after degradation process
between 1470 and 1420 cm1 [30,65]. Our results seems
to be in good agreement with Liang and coworkers whose
indicate that amorphous CH2 bending mode is found at
1455 cm1 corroborating that the degradation process
has begun [66]. Similar behavior was obtained with polyester bands between 1120 cm1 and 1100 cm1 characteristic of CAO bonds [67]; i.e. after 1200 h of exposure, the
new band is a combination of crystalline and amorphous
phases due to the breakdown of the main bonds of
polyester.
On the other hand, FT-IR spectra of degraded PET/chitosan blends (Fig. 6c) show very small bands, indicating the
typical splitting of the C@O signals with a shift of the signals to higher wavenumbers [68,69]. A noticeable change
in the band intensities is also observed, particularly,
ethylene vibrations (CH2) in the wavenumber range from
1478 to 1400 cm1 and CAO bonds in the range of
1198954 cm1. Finally, the splitting bands correlated
with crystalline-amorphous phase are missing for R-PET/
chitosan (Fig. 6d) polymer blends.
Table 3 shows the thermal properties obtained by DSC
analyses of polymer blends after degradation. Commercial
PET displays a decrease of Tc from 201 C to 170 C with a
Tm that remains fairly constant. R-PET presented a similar
behavior with a decrease of Tc from 205 C to 156 C, a new

293

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

(a)

(b)
Absorbance (%)

Absorbance (%)

PET-PLA 95/5

PET-PLA 90/10

C=O(c)
1747

C-O

C=O(a)
1700

PET-PLA 85/15

2000

1800

R-PET/PLA 95/5

R-PET/PLA 90/10

R-PET/PLA 85/15

1600

1400

1200

1000

800

2000

1800

1600

(c)

1000

800

(d)

PET/chistosan 99/1

Absorbance (%)

Absorbance (%)

1200

Wavenumber (cm )

Wavenumber (cm )

PET/chistosan 97.5/2.5

C=O (c)
1746

1800

R-PET/Chitosan 99/1

R-PET/Chitosan 97.5/2.5

1715
C=O (a)
1700
CH2

PET/chistosan 95/5

2000

1400

-1

-1

1600

1400

C-O

1200

1000

R-PET/Chitosan 95/5

800

2000

1800

1600

1400

1200

1000

800

Wavenumber (cm-1 )

-1

Wavenumber (cm )

Fig. 6. FTIR spectra of: (a) PET/PLA, (b) R-PET/PLA, (c) PET/chitosan and (d) R-PET/chitosan blends after being exposed to accelerated weathering for 1200 h.

Table 3
Thermal properties of raw materials and blends after being subjected to accelerated weathering.

Sample

Tm(PLA) (C)

Tg(chitosan) (C)

Tc (C)

Tm(PET) (C)

Td(chitosan) (C)

Td(PLA) (C)

Td(PET) (C)

PET
R-PET
PLA
Chitosan
PET/PLA 95/5
PET/PLA 90/10
PET/PLA 85/15
R-PET/PLA 95/5
R-PET/PLA 90/10
R-PET/PLA 85/15
PET/Chitosan 99/1
PET/Chitosan 97.5/2.5
PET/Chitosan 95/5
R-PET/Chitosan 99/1
R-PET/Chitosan 97.5/2.5
R-PET/Chitosan 95/5

170
156
120

248
227,245
150

152
152
155
152
152
152

115,
117, 165
117, 171

205
202
200
194
180
180
184
191
190

243
242
244
241
243
239
240
242
249
241

240

340
343
344
340

370
370
370
370
370
370

424,436
424
421
422
422
422
420
421

422

422

Since chitosan was used as powder, DSC data after accelerated weathering was not included.

peak appear at about 227 C and a Tm that remains constant. The crystallization temperature of the PET and
R-PET was found lower than before the weather chamber
tests. The observed shifts in the Tcs were directly corre-

lated with the disentangled and break of the molecules in


the amorphous phase as well as some crystalline regions.
An unexpected feature was observed in blends or PET/
PLA 95/5 and 90/10 at 152 C where the existence of the

294

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

melting point corresponding to PLA as well as the endothermic peak at 370 C corroborates that the degradation
of PLA is taking place even with low PLA concentrations
[70].
In Chitosan, there are several glass transition temperature values, this variability could be attributed to different
factors like physical state, molecular weight and de-acetylation degree [71]. From the Table 3, when chitosan was
added to polymer matrixes two glass transition temperatures appear, one at 115 and 117 C from the plasticized
chitosan as a result of the water supplied in the weathering
test, and the second one at 165 and 171 C attributed to
increase of the molecular movement due to dissociation
of hydrogen bonds and starting of molecular scissions
[72]. The DSC studies of the blends with chitosan also
showed a degradation temperature at 340 C. This
endothermic peak can be attributed to the degradation of
chitosan [68,73] and its effect into the PET matrix. Thus,
both PLA and chitosan biopolymers induce the scission of
PET bonds, producing shorter fragments and therefore,
provoking the PET degradation at 420 C [74].
Thermal stability of the polymer blends after
accelerated weathering and weight loss as a function of
temperature was analyzed by TGA (Fig. 6ad).

(a)

80

PET

PET/PLA (95/5)
PET/PLA (90/10)
PET/PLA (85/15)

60

100

Weight loss (%)

Weight loss (%)

100

From Fig. 7a and c, it is seen that commercial PET before


weathering decompose at 360 C and at 500 C presents
14% of residual weight. Similar results were obtained for
R-PET before weathering (Fig. 7b and d) which decomposes
at 356 C leaving 13.8% of residual weight (500 C). The
TGA analyses showed that the chain scission of the ester
bonds (degradation of the polymer backbone) occurred in
this interval of temperatures [75]. In Fig. 7a and b, it can
be seen that the initial step of PLA degradation starts at
lower temperatures than PET, at around of 307 C leaving
negligible residue (2-wt%) above 397 C. This is explained
due to when PLA is exposed to elevated temperatures
(370 C), undergoes thermal degradation, leading to the
formation of lactide monomers [76].
From the blends TGA studies, is clearly shown that PLA
has strong effect to increase the degradation on both
matrixes (PET and R-PET) (Fig. 7a and b). The initial decomposition temperature of PET/PLA (Fig. 7 a) is reduced as the
content of PLA is increased in the blends, showing values of
287 C, 250 C and 150 C with a residual weight of 17, 16.5
and 13-% at 500 C, for the PET/PLA weight ratios of 95/5,
90/10 and 85/15, respectively. On the contrary, R-PET/PLA
blends (Fig. 7 b) show an increase in the total mass percentage in the range of 250327 C which was correlated with

PLA

40
20

(b)

80 R-PET/PLA (95/5)
R-PET/PLA (90/10)

60 R-PET/PLA (85/15)
PLA

40
20

0
100

200

300

400

50 100 150 200 250 300 350 400 450 500

500

Temperature (C)

Temperature (C)
(c)
100

Weight loss (%)

Weight loss (%)

100

R-PET

80
60
40
20
0
100

PET/Chitosan (95/5)
PET/Chitosan (97.5/2.5)
PET/Chitosan (99/1)
PET

200

300

400

Temperature (C)

(d)

80
60
40
20

500

0
100

R-PET/Chitosan (95/5)
R-PET/Chitosan (97.5/2.5)
R-PET/Chitosan (99/1)
R-PET

200

300

400

500

Temperature (C)

Fig. 7. TGA thermograms of: (a) PET/PLA, (b) R-PET/PLA, (c) PET/chitosan and (d) R-PET/chitosan blends after being exposed to accelerated weathering for
1200 h.

295

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

100 m

(a)

PET/PLA 95/5

(b)

10 m

(d)

R-PET/PLA 95/5

(c)

PET/PLA 90/10

10 m

(g)

PET/chitosan 99/1 10 m

(j)

R-PET/chitosan 99/1 10 m

R-PET/PLA 90/10

10 m

10 m

10 m

PET/chitosan 95/5 10 m

(l)

(k)

R-PET/chitosan
97.5/2.5

R-PET/PLA 85/15

(i)

(h)

PET/chitosan
97.5/2.5

10 m

(f)

(e)

10 m

PET/PLA 85/15

10 m

R-PET/chitosan
95/5

10 m

Fig. 8. PET/PLA, R-PET/PLA, PET/chitosan and R-PET/chitosan blends with different weight ratios after 900 h of accelerated weathering.

the evaporation of additives used during the production of


R-PET. The rst stage of degradation changed to 300 C,
327 C and 50 C, with a residual weight of 19%, 20% at
500 C and 100-% at 466 C, for the weight ratios of 95/5,
90/10 and 85/15, respectively. The TGA analyses evidenced
that the incorporation of PLA and the accelerated weathering conditions decreases signicantly the thermal stability
of the blends.
In the case of blends with chitosan (Fig. 7c and d), nonappreciable changes were observed in the temperature
where degradation starts, however, a reduction in the mass

loss can be observed as the amount of chitosan is


increased. The rst degradation step varies with the weight
ratios of 99/1, 97.5/2.5 and 95/5 at 276 C, 305 C and
311 C with a nal weight of 16, 19 and 20.5 wt-%, respectively. Similar trend is again observed for R-PET/chitosan
polymer blends obtaining degradation temperatures about
367 C, 299 C and 273 C with a nal weight of 16.5, 18.6
and 20.6 wt-%, respectively, under the same weight ratios.
Then, even though chitosan enhances slightly the thermal
stability, reduces the degradation process in comparison
with PLA.

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

3.3. Degradation of polymer blends in soil


From studies of PET degradation, kinetic models have
reported, based on accelerated experiments of hydrolytic
degradation and the life time of PET was found in the range
from 16 to 48 years [80,81]. Unfortunately, there are few
studies about the degradation in soil of PET [82,83]. The

0.004

R-PET/Chitosan (95/5)

R-PET/Chitosan (97.5/2.5)

R-PET/Chitosan (99/1)

PET/Chitosan (95/5)

R-PET/PLA (85/15)

R-PET/PLA (95/5)

R-PET/PLA (90/10)

PET/Chitosan (99/1)

0.000

PET/PLA (85/15)

0.001

PET/PLA (90/10)

0.002

PET/Chitosan (97.5/2.5)

0.003

PET/PLA (95/5)

Fig. 8al shows micrographs of representative weathered blends after 1200 h. The micrographs of the weathered laments evidenced rough surfaces caused by the
effects of the UV light and water that simultaneously cause
surface microcracks; initially, microcracks or ssures were
formed due to the breaking of the backbone CAO bonds
and thereafter are dispersed in different points of the surface. The morphology was quit affected by the PLA amount,
producing further degradation. For damage accumulation,
some regions are so weakened that cracking formation
starts microcracks and expanding over all surface of laments (inset of Fig. 8c). Thus, in presence of PLA and after
this biodegradable polymer begins to degrade, PET starts
the disintegration process in ACAOA and AC@O bonds
[77], which are produced from cleavage at CAC bonds of
the main polymer chain [78]. In Fig. 9, it is shown a proposed model for photo-oxidative degradation of polymer
blends where are illustrated the chemical alterations and
crack formation during weathering test.
By comparison of the morphologies of PLA and chitosan,
it is shown that chitosan blends presented small quantity
of microcracks (Fig. 8gl); however, some deep holes
formed by the swelling of chitosan powders were also evident. The swelling may be caused by the water used in the
QUV camera, leaving away the chitosan particles which
produce the microcracks [79]. In this case, the photo-oxidation process of PET/chitosan blends occurs due to alterations such as cross-linking and chains scission; i.e. the
formation of carboxylic groups increasing the amount of
polar groups on the surface during and after irradiation
time point out both the photo-oxidation of chitosan and
changes of its structure. Additional groups such as OH,
OOH, CO can also contribute to the rapid degradation of
chitosan with the subsequent swelling and cracks formation and in consequence, eventually, the PET disintegration
[80].

Weight Loss (%)

296

Blends after 6 months


Fig. 10. Degradation trends for PET/PLA, R-PET/PLA, PET/Chitosan, R-PET/
chitosan blends in soil.

degradation of the blends was measured by monitoring


the plastic weight loss during 6 months under real eld
conditions (Fig. 10). It can be noticed that the semicrystalline (commercial) and amorphous (recycled) PET as well as
PET from biodegradable bottle did not show important
variations in their weight after six months. It means that
degradation takes a long time for some samples even
though with some amounts seems to be that the material
disintegration has started. For example, PET/PLA 95/5 and
PET/Chitosan 99/1 blends remained without changes during all the time, whereas R-PET/PLA (90/10), R-PET/PLA
(85/15), PET/Chitosan (95/5) and R-PET/Chitosan (95/5)
blends showed certain degradation. It is also seen that
R-PET/chitosan (95/5) blends exhibited the highest loss
weight in comparison with the other polymer blends.
Hence, degradation in soil is favored when higher amounts
of chitosan or PLA are added to PET matrixes and from
these the degradation is accelerated with R-PET.
Accelerated QUV weathering tests were found to be consistent with those obtained under real eld conditions. To
prove these observations thermogravimetric analyses of
the samples were also evaluated (Fig. 11ad). The decomposition temperature (Td) is shifted to low values as PLA
content is increased. Similarly, PET/chitosan with a 95/5
weight ratio showed a displacement up to reach a Td about

Fig. 9. Proposed model for photo-oxidative degradation of PET/PLA polymer blend.

297

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

(a)

80

100

PET
PET/PLA (95/5)
PET/PLA (90/10)
PET/PLA (85/15)

60

PLA

40
20
0
100

200

(b)

300

400

R-PET/PLA (95/5)

60 R-PET/PLA (90/10)
40 R-PET/PLA (85/15)
PLA

20
0
100

500

200

Temperature (C)
(c)

100

80
60
PET

40
PET/Chitosan (99/1)

20

200

300

400

400

500

500

Temperature (C)

(d)

80
60
40

R-PET
R-PET/Chitosan (99/1)

20

PET/Chitosan (97.5/2.5)
PET/Chitosan (95/5)

0
100

300

Temperature (C)

Weight loss (%)

Weight loss (%)

100

R-PET

80

Weight loss (%)

Weight loss (%)

100

R-PET/Chitosan (97.5/2.5)
R-PET/Chitosan (95/5)

0
100 150 200 250 300 350 400 450 500

Temperature (C)

Fig. 11. Thermograms for (a) PET/PLA, (b) R-PET /PLA, (c) PET/chitosan and (d) R-PET/chitosan in composting for 6 months.

384 C whereas with other ratios the decomposition temperature remains fairly constant, corroborating the loss
weight in soil.
According to the above results, the addition of higher
amounts of PLA or chitosan on both PET matrixes can
modied its degradation mechanism. PLA degradation at
temperature between 3040 C and 80-% of humidity
could be due to the chemical hydrolysis that provokes that
eventually the PET/PLA blends cleavage and degrades; similarly chitosan is degraded by the oxygen presence due to
the interaction in the water drop/surface of the blend samples interface.
4. Conclusions
Blends of commercial and recycled PET/PLA and -/
Chitosan were prepared by extrusion process at 250 C to
analyze the miscibility and degradability as a function of
biopolymer content. The following conclusions can be
drawn:
After extrusion process, XRD patterns and infrared spectra analysis showed a weak interaction between PET
matrixes and biodegradable polymers suggesting type secondary bonds by hydrogen bridges or by electrostatic
forces. In independence of the amount of biodegradable
polymers, their semi-crystalline character helps to the
recrystallization of R-PET during the extrusion reprocess-

ing, which in turn is affected by two factors: quenched


temperature and mobility of the biopolymer fraction. The
miscibility of the polymer-biopolymer blends is independent of the PET matrix and the saturation with PLA is
reached with 10 wt-%, whereas for chitosan all the compositions display an adequate miscibility. SEM analysis
showed that the green polymers were uniformly distributed into the polymer matrixes, but the mean diameter
of the agglomerates and their dissolution varied with the
amount and kind of the biopolymer. The accelerating
weathering tests indicate that the interaction between
PET and chitosan favors more the degradation rate in comparison with PET/PLA blends, which it is thermally more
stable. The best performance was obtained for PET/chitosan polymer blend with a 95/5 weight ratio where an estimate time of about 45 years is required for its degradation.
Finally, all the results obtained from various analyses indicate that the differences in the thermal stability and degradation rate suggest a certain degree of interaction between
blend components which are comparable to commercial
bottles of BioPET, which uses higher amounts of biopolymer materials.
Acknowledgements
D. Palma-Ramrez is grateful for her postgraduate
fellowship to CONACYT, COFAA and SIP-IPN. The authors

298

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299

are also grateful for the nancial support provided by


CONACYT through the CB2009-132660 and CB2009133618 projects and to IPN through SIP 2014-0164 and
2014-0992 projects and SNI-CONACYT. The authors also
thank M.E.A.E. Rodrguez-Salazar and ROMFER SA CV
industries for their technical support.
Appendix A. Supplementary material
Supplementary data associated with this article can be
found, in the online version, at http://dx.doi.org/10.1016/
j.eurpolymj.2014.10.016.
References
[1] Ivar do Sul JA, Costa MF. The present and future of microplastic
pollution
in
the
marine
environment.
Environ
Pollut
2014;185:35264.
[2] Tian H, Gao J, Hao J, Lu L, Zhu C, Qiu P. Atmospheric pollution
problems and control proposals associated with solid waste
management in China: a review. J Hazard Mater 2013;252
253:14254.
[3] Kasetaite S, Ostrauskaite J, Grazuleviciene V, Svediene, Bridziuviene
D. Camelina oil- and linseed oil-based polymers with
bisphosphonate crosslinks. J Appl Polym Sci 2014;131:18.
[4] Rubio-Anaya M, Guerrero-Beltrn JA. Polmeros utilizados para la
elaboracin de pelculas biodegradables. Temas Selectos de
Ingeniera de Alimentos 2012;6:17381.
[5] Garrison TF, Kessler MR, Larock RC. Effects of unsaturation and
different ring-opening methods on the properties of vegetable oilbased polyurethane coatings. Polymer 2014;4:100411.
[6] Lpez OV, Castillo LA, Garca MA, Villar MA, Barbosa SE. Food
packaging bags based on thermoplastic corn starch reinforced with
talc nanoparticles. Food Hydrocolloids 2014. http://dx.doi.org/
10.1016/j.foodhyd.2014.04.021.
[7] Petersson L, Kvien I, Oksman K. Structure and thermal properties of
poly(lactic acid)/cellulose whiskers nanocomposite materials.
Compos Sci Technol 2007;67:253544.
[8] Qu P, Gao Y, Wu G-F, Zhang L-P. Nanocomposites of poly(lactic acid)
reinforced with cellulose nanobrils. Bioresources 2010;5:181123.
[9] Choudhary P, Mohanty S, Nayak SK, Unnikrishnan L. Poly(L-lactide)/
polypropylene blends, evaluation of mechanical, thermal, and
morphological characteristics. J Appl Polym Sci 2011;121:322337.
[10] Zhu M, Li S, Li Z, Lu X, Zhang S. Investigation of solid catalysts for
glycolysis
of
polyethylene
terephthalate.
Chem
Eng
J
2012;185:16877.
[11] Girija BG, Sailaja RRN, Madras G. Thermal degradation and
mechanical properties of PET blends. Polym Degrad Stab
2005;90:14753.
[12] Welle F. Twenty years of PET bottle to bottle recycling an
overview. Resour Conserv Recycl 2011;55:86575.
[13] Garlotta D. A literature review of poly (lactic acid). J Polym Environ
2001;9:6384.
[14] Tokiwa Y, Calabia B. Biodegradability and biodegradation of
poly(lactide). Appl Microbiol Biotechnol 2006;72:24451.
[15] Auras RA, Harte B, Selke S, Hernndez R. Mechanical, physical, and
barrier properties of poly(lactide) lms. J Plastic Film Sheeting
2003;19:12335.
[16] Bautista-Baos S, Hernndez-Lauzardo AN, Velzquez del Valle MG,
Hernndez-Lpez M, Ait-Barka E, Bosquez-Molina E, et al. Chitosan
as a potential natural compound to control pre and postharvest
diseases of horticultural commodities. Crop Prot 2006;25:10818.
[17] Kint D, Muoz-Guerra S. A review on the potential biodegradability
of poly(ethylene terephthalate). Polym Int 1999;48:34652.
[18] Siracusa V, Rocculi P, Romani S, Dalla Rosa M. Biodegradable
polymers for food packaging: a review. Trends Food Sci Technol
2008;19:63443.
[19] Zou H, Yi C, Wang L, Xu W. Crystallization, hydrolytic degradation,
and mechanical properties of poly (trimethylene terephthalate)/
poly(lactic acid) blends. Polym Bull 2010;64:47181.
[20] Malgorzata J, Kensuke S, Pierre G, Eric G. Inuence of chitosan
characteristics on polymer properties, I: crystallographic properties.
Polym Int 2003;52:198205.

[21] Hua H, Changsheng L, Wenjing W. Preparation and characterization


of chitosan/PEG/gelatin composites for tissue engineering. J Appl
Polym Sci 2009;114:12205.
[22] US Patent 2013037546. Methods of preparing para-xylene from
biomass, Search International and National Patent Collections.
[23] Xavier Nunes RA, Costa VC, De Araujo Calado VM, Tavares Branco JR.
Wear, friction, and microhardness of a thermal sprayed PET poly
(ethylene terephthalate) coating. Mater Res 2009;12:1215.
[24] Hwang KT, Kim JT, Jung ST, Cho GS, Park HJ. Properties of chitosanbased biopolymer lms with various degrees of deacetylation and
molecular weights. J Appl Polym Sci 2003;89:347684.
[25] Samuels RJ. Solid state characterization of the structure of chitosan
lms. J Polym Sci 1981;19:1081105.
[26] Righetti MC, Tombari E, Angiuli M, Di Lorenzo ML. Enthalpy-based
determination of crystalline, mobile amorphous and rigid
amorphous fractions in semicrystalline polymers poly(ethylene
terephthalate). Thermochim Acta 2007;462:1524.
[27] Wunderlich B. Reversible crystallization and the rigidamorphous
phase in semicrystalline macromolecules. Prog Polym Sci
2003;28:383450.
[28] Holland BJ, Hay JN. The thermal degradation of PET and analogous
polyesters measured by thermal analysis-Fourier transform infrared
spectroscopy. Polymer 2002;43:183547.
[29] Prasad SG, De A, De U. Structural and optical investigations of
radiation damage in transparent PET polymer lms. Int J Spectrosc
2011;1:17.
[30] Drobota M, Aori M, Barboiu V. Protein immobilization on
Poly(ethylene terephthalate) lms modied by plasma and
chemical treatments. Digest J Nanomater Biostruct (DJNB)
2010;5:3542.
[31] Boerio FJ, Bahl SK, McGraw GE. Vibrational analysis of polyethylene
terephthalate and its deuterated derivatives. J Polym Sci Polym Phys
Ed 1976;14:102946.
[32] Miyake A. The infrared spectrum of polyethylene terephthalate. I.
The effect of crystallization. J Polym Sci 1959;38:47995.
[33] Cristina B, Xavier D, Serge E. Characterization of poly(ethylene
terephthalate) used in commercial bottled water. IOP Conf Series:
Mater Sci Eng 2009;5:15.
[34] Zhao Q, Jia Z, Li X, Ye Z. Surface degradation of unsaturated polyester
resin in Xe articial weathering environment. Mater Des
2010;31:445760.
[35] Al-Jabareen A, Illesca S, Maspotch MLI, Santana OO. Effects of
composition and transesterication catalysts on the physicochemical and dynamic properties of PC/PET blends rich in PC. J
Mater Sci 2010;45(24):662333.
[36] Na SK, Kong BG, Choi C, Jang MK, Nah JW. Transesterication and
compatibilization in the blends of bisphenol-A polycarbonate and
poly(trimethylene terephthalate). Macromol Res 2005;13(2):
8895.
[37] Kuo SW, Chang FC. Signicant thermal property and hydrogen
bonding strength increase in poly(vinylphenol-co-vinylpyrrolidone)
copolymer. Polymer 2003;44(10):302130.
[38] Shieh YT, Yang YF. Signicant improvements in mechanical property
and water stability of chitosan by carbon nanotubes. Eur Polym J
2006;42:316270.
[39] Rueda DR, Secall T, Bayer RK. Differences in the interaction of water
with starch and chitosan lms as revealed by infrared spectroscopy
and differential scanning calorimetry. Carbohydr Polym
1999;40:4956.
[40] Adoor SG, Manjeshwar LS, Krishna Rao KSV, Naidu B, Aminabhavi
TM. Solution and solid-state blend compatibility of poly(vinyl
alcohol) and poly(methyl methacrylate). J Appl Polym Sci
2006;100:241521.
[41] Jawalkar SS, Adoor SG, Sairam M, Nadagouda MN, Aminabhavi TM.
Molecular modeling on the binary blend compatibility of poly(vinyl
alcohol) and poly(methyl methacrylate): an atomistic simulation
and thermodynamic approach. J Phys Chem B 2005;109:1561120.
[42] Feng L, Bian X, Li G, Chen Z, Cui Y, Chen X. Determination of ultralow glass transition temperature via differential scanning
calorimetry. Polym Testing 2013;32:136872.
[43] MacKnight WJ, Karasz FE, Fried JR. Solid state transition behavior of
blends. Chapter 5. Polym Blends 1978:185242.
[44] Demirel B, Yaras AH, Elcicek BF. Crystallization behavior of PET
materials. Bil Enst Dergisi Cilt 2011;13:2635.
[45] Strobl G. The Physics of Polymers: Concepts for Understanding Their
Structures and Behavior. Berlin, Germany: Springer; 1997.
[46] Chen H, Pyda M, Cebe P. Non-isothermal crystallization of PET/PLA
blends. Thermochim Acta 2009;492:616.

A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285299


[47] Nisha P, Moghe N, Riccobono O, Tyagi S, Rajagopalan S. Novel
approaches in microstructure evaluation of polymer blends. Microsc
Microanal 2005;11(suppl. S02):21101.
[48] Correlo VM, Boesel LF, Bhattacharya M, Mano JF, Neves NM, Reis RL.
Properties of melt processed chitosan and aliphatic polyester blends.
Mater Sci Eng: A 2005;403:5768.
[49] Muthukumar T, Aravinthan A, Mukesh D. Effect of environment on
the degradation of starch and pro-oxidant blended polyolens.
Polym Degrad Stab 2010;95:198893.
[50] Krcalk M, Pospisil L, Slouf M, Mikesova J, Sikora A, Simonik J, et al.
Effect of glass bers on rheology, thermal and mechanical properties
of recycled PET. Polym Compos 2008;29:91521.
[51] Pawlak A, Plutaa MB, Morawieca J, Galeskia A, Pracellab M.
Characterization of scrap poly(ethylene terephthalate). Eur Polym J
2000;36:187584.
[52] Spinac MAS, De Paoli MA. Characterization of poly(ethylene
terephtalate) after multiple processing cycles. J Appl Polym Sci
2001;80:205.
[53] Tudorachi N, Cascabal CN, Rusu M, Pruteanu M. Testing of polyvinyl
alcohol and starch mixtures as biodegradable polymeric materials.
Polym Testing 2000;19:78599.
[54] Huang SJ, Ho LH, Huang MT, Koenig MF, Cameron JA. Similarities and
differences
between
biodegradation
and
non-enzymatic
degradation. Stud Polym Sci 1994;12:310.
[55] Anderson JM, Shive MS. Biodegradation and biocompatibility of PLA
and PLGA microspheres. Adv Drug Deliv Rev 1997;28(1):524.
[56] Martin D, Eng P. Advanced thin lm geomembrane technology for
biocell liners and covers (available from Layeld Geosynthetics and
Industrial Fabrics Ltd) (2005).
[57] Wagner N, Ramsey B. QUV accelerated weathering study: analysis of
polyethylene lm and sheet samples. Technical Document by GSE
Lining Technology, Inc., 2003. Houston, TX, USA
[58] Andrady AL, Hamid HS, Torikai A. Effects of climate change and UV-B
on materials. Photochem Photobiol Sci 2003;2:6872.
[59] Fischer HR, Semprimosching C, Mooney C, Rohr T, Ernst RH,
Verkuijlen MHW. Degradation mechanism of silicone glues under
UV irradiation and options for designing materials with increased
stability. Polym Degrad Stab 2013;98:7206.
[60] Andanson JM, Kazarian SG. In situ ATR-FTIR spectroscopy of
poly(ethylene terephthalate) subjected to high-temperature
methanol. Macromol Symp 2008;265:195204.
[61] Zhang WR, Hinder SJ, Smith R, Lowe C, Watts JF. An investigation of
the effect of pigment on the degradation of a naturally weathered
polyester coating. J Coat Technol Res 2011;8:32942.
[62] Mai F, Habibi Y, Raquez JM, Dubois P, Feller JF, Peijs T, et al.
Poly(lactic acid)/carbon nanotube nanocomposites with integrated
degradation sensing. Polymer 2013;54:681823.
[63] Zhang J, Sato H, Tsuji H, Noda I, Ozaki Y. Differences in the
CH3. . .O@C interactions among poly(l-lactide), poly(l-lactide)/
poly(d-lactide) stereocomplex, and poly(3-hydroxybutyrate)
studied by infrared spectroscopy. J Mol Struct 2005;735:24957.
[64] Shimizu T, Ohkawa Tanaka Y, Kutsumizu S, Yano S. Structural
Studies of Poly(1H,1H-uoroalkyl alpha uoroacrylate)s by infrared
spectroscopic analysis. Macromolecules 1996;29:35404.

299

[65] Dieval F, Khof F, Mir R, Chaouch W, Le Nouen D, Chakfe N, et al.


Long-term biostability of pet vascular prostheses. Int J Polym Sci
2012;2012:114.
[66] Liang CY, Krimm S. Infrared spectra of high polymers: Part IX.
Polyethylene terephthalate. J Mol Spectrosc 1959;3:55474.
[67] Donelli I, Freddi G, Nierstrasz VA, Taddei P. Surface structure and
properties of poly-(ethylene terephthalate) hydrolyzed by alkali and
cutinase. Polym Degrad Stab 2010;95:154250.
[68] De Britto D, Campana-Filho SP. Kinetics of the thermal degradation
of chitosan. Thermochim Acta 2007;465:7382.
[69] Pierg M, Ostrowska-Czubenko J, Gierszewska-Druzynska M.
Thermal degradation of double crosslinked chitosan membranes.
Prog Chem Appl Chitin Derivatives XVII 2012:6774.
[70] Arrieta MP, Lpez J, Ferrndiz S, Peltzer A. Characterization of PLAlimonene blends for food packaging applications. Polym Testing
2013;32:7608.
[71] Zhang L, Kosaraju SL. Biopolymeric delivery system for controlled
release of polyphenolic antioxi-dants. Eur Polym J 2007;43:
295666.
[72] Mucha M, Pawlak A. Thermal analysis of chitosan and its blends.
Thermochim Acta 2005;427:6976.
[73] Fernandes LL, Resende CX, Tavares DS, Soares GA, Castro LO, Ganjeiro
JM. Cytocompatibility of chitosan and collagen-chitosan scaffolds for
tissue engineering. Polmeros 2011;21(1):16.
[74] Chen DQ, Wang YZ, Hu XP, Wang DY, Hai M, Yang B. Flame-retardant
and anti-dripping effects of a novel char-forming ame retardant for
the treatment of poly(ethylene terephthalate) fabrics. Polym Degrad
Stab 2005;88:34956.
[75] Turnbull L, Liggat JJ, MacDonald WA. Thermal degradation chemistry
of poly(ethylene naphthalate) a study by thermal volatilisation
analysis. Polym Degrad Stab 2013;98:224458.
[76] Lim LT, Auras R, Rubino M. Processing technologies for poly(lactic
acid). Prog Polym Sci 2008;33:82052.
[77] Babanalbandi A, Hill DJT, ODonnell JH, Pomery PJ, Whittaker A. An
electron spin resonance study on gamma-irradiated poly(L-lactic
acid) and poly(D, L-lactic acid). Polym Degrad Stab
1995;50:297304.
[78] Copinet A, Bertrand C, Govindin S, Coma V, Couturier Y. Effects of
ultraviolet light (315 nm), temperature and relative humidity on the
degradation of polylactic acid plastic lms. Chemosphere
2004;55:76373.
[79] Matuana LM, Jin S, Stark NM. Ultraviolet weathering of HDPE/woodour composites coextruded with a clear HDPE cap layer. Polym
Degrad Stab 2011;96:97106.
[80] Eubeler JP, Bernhard M, Knepper TP. Environmental biodegradation
of synthetic polymers II. Biodegradation of different polymer groups.
TrAC Trends Anal Chem 2010;29:84100.
[81] Mller RJ, Kleeberg I, Deckwer WD. Biodegradation of polyesters
containing aromatic constituents. J Biotechnol 2001;86:8795.
[82] Radulovic J. Degradation of polyethylene terephthalate in natural
conditions. Sci Techn Rev LVI 2006:4551.
[83] Platt DK, Rapra L. Technology: biodegradable polymers: Market
Report Rapra Technology (2006).

Vous aimerez peut-être aussi