Sunteți pe pagina 1din 15

European Polymer Journal 61 (2014) 285–299

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-Ramírez b, M.A. Domínguez-Crespo a, D. Del Angel-López a,
D. de la Fuente c
a
Instituto Politécnico Nacional, CICATA-Altamira, Km. 14.5 Carretera Tampico-Puerto Industrial Altamira, 89600 Altamira, Tamps, Mexico
b
PMTA of CICATA-Altamira, IPN, Km. 14.5 Carretera Tampico-Puerto Industrial Altamira, 89600 Altamira, Tamps, Mexico
c
Centro Nacional de Investigaciones Metalúrgicas, CENIM (CSIC), Av. Gregorio del Amo 8, 28040 Madrid, Spain

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

Article history: This work reports the synthesis and miscibility of PET/PLA and PET/chitosan blends as well
Received 4 August 2014 as their degradation in real soil environment (6 months) and in accelerated weathering
Received in revised form 18 October 2014 (1200 h). For this purpose, commercial polyethylene terephthalate (PET) and recycled
Accepted 23 October 2014
PET (R-PET) were used as polymer matrixes and extruded with different amounts of poly-
Available online 6 November 2014
lactic acid (5, 10 and 15 wt-%) or chitosan (1, 2.5 and 5 wt-%) to form filaments. Different
characterization techniques such as X-ray diffraction (XRD), Fourier transform infrared
Keywords:
spectroscopy (FTIR), differential scanning calorimetry/thermogravimetric analysis (DSC/
PET/PLA
PET/chitosan
TGA) and scanning electron microscopy (SEM) were used before and after degradation pro-
Blends cess. The results indicate weak interactions between blend components suggesting second-
Miscibility ary bonds by hydrogen bridges or by electrostatic forces. The miscibility of chitosan in both
Degradability 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.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction biodegradable polymers have a short-lifetime because of


are ideal for short-time applications such as; disposable
The long-lasting petroleum polymers have been widely packages, agricultural mulches, horticultural pots, etc
used provoking that the waste of this kind of polymers [3–6]. They are also naturally degradable when disposed
takes a very long time to be broken down. Nowadays, this in the environment. Despite its advantages, many of these
indiscriminate use of petroleum-based polymers has kinds of polymers exhibit poor thermal stability, low steam
caused a big pollution problem [1,2]. To reduce this and gas barrier and low mechanical properties, making
problem, it has been used biodegradable polymers from them unsuitable for other applications [7,8]. Therefore,
renewable sources like collagen, keratin, gluten, milk pro- the general trend is to combine the mechanical, barrier
teins, soy proteins, polysaccharides like starch, cellulose and thermal properties of petroleum based polymers with
derivatives, chitosan, alginate, carrageenan, pectins. These the biodegradability properties of renewable polymers,
resulting in the production of polymeric materials with
⇑ Corresponding author. Tel.: +52 8332600125x87510; fax: +52 controlled lifetime. The designed materials must be resis-
8332600125x87521. tant during their use and must have short time degradation
E-mail addresses: atorresh@ipn.mx, atohuer@hotmail.com at the end of their useful life [4,9].
(A.M. Torres-Huerta).

http://dx.doi.org/10.1016/j.eurpolymj.2014.10.016
0014-3057/Ó 2014 Elsevier Ltd. All rights reserved.
286 A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299

The most favorable packaging material for disposable into flakes. The polylactic acid pellets, PLA-2002D (contain-
soft drink bottles is polyethylene terephthalate (PET), a ing 4.4 wt-% in average of isomer D), (batch: YA0828b131)
kind of semicrystalline, thermoplastic polyester with high and chitosan (low molecular weight, with a deacetylation
strength and transparency properties as well as excellent degree P75%) were purchased from NatureWorks LLC,
barrier properties. Unfortunately, most of these beverage USA and Sigma Aldrich, respectively. Before processing,
bottles are used only once and then discarded, which inev- the raw materials were dried at 60 °C during 24 h in an
itably generates serious environmental problems (white oven (Thermolyne). Different amounts of PLA (5, 10 and
pollution) [10]. Therefore, recycling the discarded PET 15 wt-%) or chitosan (1, 2.5 and 5 wt-%) and commercial
polymer along with obtaining biodegradable PET-based PET or R-PET were hand mixed previous to extrusion pro-
blends are efficient approaches to reduce the resources cess. It is important to mention that, initially, we tried to
consumption and to protect the environment at the same add into the polymer matrix the same quantities of chito-
time [11]. The recycling of post-consumer packaging mate- san or PLA (5, 10 and 15 wt-%), unfortunately, we observed
rials into direct food contact packaging applications was that synthesized samples with chitosan become rigid and
not possible, because of the lack of knowledge about the brittle, regarding the chitosan polymer is a brittleness
contamination of packaging polymers during first use or material, however, this characteristic depend if the mate-
recollection. However, for PET the situation is much favor- rial is derived of fungal biomass, crustacean shell and
able: due to its inert character, recycling technologies have insect cuticles [20,21]. Thus, after several experiments we
been developed to establish a bottle-to-bottle recycling of found that an optimal percentage to evaluate chitosan is
post-consumer PET bottles [12]. less than 5 wt-%. Blends with a filaments shape (1 mm in
On the other hand, most biodegradable polymers are diam.  200 cm length) were obtained in a single-screw
thermoplastics (e.g. poly(lactic acid), poly(hydroxyalkano- extruder (C.W. Brabender) with L/D ratio of 25:1 and four
ate), poly(vinyl alcohol)) [9]. Among them, poly(lactic acid) heating zones: feeding (225 °C), compression (237.5 °C),
(PLA) is a bio-based polymer obtained from renewable distribution (260 °C), and finally, the extrusion die
sources mainly from corn and starch [13]. PLA is an aro- (225 °C).
matic polyester and has several applications, for example,
is used for films, extrusion-thermoformed containers and 2.2. Characterization
medical applications for tissue engineering, bone recon-
struction and controlled delivery systems [14]. The use of Structural characterization of the polymer blends was
PLA in beverage bottles is limited due to its poor oxygen carried out using a Bruker D8 Advance diffractometer from
barrier and low mechanical properties [15]. Another inter- 2h = 5–60° (Cu Ka, k = 0.154 nm) and a rate of 1.5 °/min.
esting biodegradable polymer is chitosan, a biopolymer The Fourier Transform Infrared Spectroscopy (FT-IR)
derived from chitin, a natural compound from crustacean spectra were recorded with a Nicolet FT-IR spectrometer
shells; chitosan has the ability to form semipermeable (Magna System 550) equipped with an attenuated total
films and, in recent years, the efforts have been intensified reflectance (ATR) accessory between 2000 and 650 cm1
to develop chitosan films and its application in food pack- at an optical resolution of 4 cm1 (40 scans).
ing [16]. Biodegradable copolymers of PET and aliphatic Simultaneous thermal analysis was carried out in a
polyesters have been synthesized, such as poly(lactic acid), Labsys Evo, Setaram instrument, which was used in the
poly(b-hydroxyalkanoate), poly( e -caprolactone) and DSC/TGA configuration with the sample and reference cru-
poly(butylene succinate) in order to obtain a degradable cibles made of a-Al2O3. Sample amount in the crucible was
polymer with a faster degradation rate [17,18]. about 10 mg. The samples were firstly heated at 40 °C and
Additionally, physical mixtures of conventional and hold for 2 min and subsequently, the measurements were
biodegradable phases have been studied [19]. To our best carried out in the range of 40–300 °C to evaluate thermal
knowledge, few researches are focused in determine the degradation under argon atmosphere with a heating rate
effects on physicochemical, structural and morphological of 10 °C/min. Then, the samples were hold at 300 °C for
properties as well as degradation time of blends PET/PLA 2 min followed by a cooling with the same rate. It is well
or PET/Chitosan. In this work, the issue has been investi- known that peak temperature is influenced by the scan
gated from different perspectives and the results are rate; however, in this case such displacement was consid-
discussed in the terms of the quantities of biodegradable ered neglected. It is also important to mention that DSC
polymers that were added in two different matrixes com- heating was intentionally evaluated from 40 °C to 300 °C
mercial PET and recycled PET during extrusion process. 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
2. Experimental heating rate but using a temperature range from 40 °C to
500 °C.
2.1. Materials and processing In order to dissolve the biopolymer phase and to evalu-
ate the dispersion in the different matrixes (PET and R-PET)
During the first set of experiments, it was used a com- as well as evaluate the morphology of the as-prepared
mercial PET (CLEARTUFÒ-MAX2, lot no. 1008-03219) pro- samples, the polymer blends were solubilized in chloro-
vided as pellets by M&G Polymers Company whereas in form and acetic acid for PLA and chitosan, respectively.
the second step recycled PET (R-PET) was obtained from The dissolved samples were observed by SEM to analyze
discarded bottles after they were washed, dried and cut the cross section of the blends; before characterization,
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 287

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

(011)
(010)

(110)
(100)
(111)
(a)
with an Au–Pd thin film. SEM micrographs were acquired
PET
using SEM JEOL 6300 series apparatus operating at 15 kV.

R-PET

Intensity (a.u.)
2.3. Accelerated weathering

(002)
(100)
Accelerated weathering tests were carried out in a QUV Chitosan
accelerated weathering chamber Model QUV/Se which

(201)

(111)
(203)
(110)
{(200)
uses UV-B lamps (313 nm and 0.63 W/m2) as radiation

(112)
{

(300)
(102)

(210)
source. During these experiments, PET, R-PET, commercial

{
PLA
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 10 20 30 40 50 60
8/4 h at 60 °C/50 °C for their further characterization. 2θ (degrees)
Additionally, the average weight loss mass was evaluated
during 300 h, 600 h, 900 h and 1200 h. General process to (b)

(010)
(111)
(110)

(100)
(011)
obtain commercial bottles of BioPET (B-PET) considers PET/PLA 95/5
the reaction of bio-terephthalic acid with pretroleum-

Intensity (a.u.)
derive ethylene glycol or bio-terephthalic acid in combina- PET/PLA 90/10
tion with bio-ethylene glycol [22].
PET/PLA 85/15

2.4. Degradation in soil R-PET/PLA 95/5

The polymer blends were buried in a commercial R-PET/PLA 90/10


compost and exposed to the environment in Altamira, R-PET/PLA 85/15
Tamaulipas México for 6 months where the average
temperatures varies from 30 °C to 40 °C with an average 10 20 30 40 50 60
of relative humidity of 80%. To analyze the effect on the deg-
2 θ (degrees)
radation by adding the biopolymers, DSC/TGA analysis
were recorded to compare with the results of the as-pre-
pared blends. (c)
(010)
(111)
(110)

(100)
(011)

PET/Chitosan 99/1

3. Results and discussion


Intensity (a.u.)

PET/Chitosan 97.5/2.5

3.1. Characterization of as-prepared polymer blends after PET/Chitosan 95/5


extrusion
R-PET/Chitosan 99/1
Since, the final properties of the materials strongly R-PET/Chitosan 97.5/2.5
depend on the induced microstructure which can be gov-
erned by the complex thermo-mechanical history; it is of R-PET/Chitosan 95/5

prime interest to observe and to understand the develop-


10 20 30 40 50 60
ment of the crystalline phase. Fig. 1a–c shows X-Ray dif-
fraction patterns of the polymer blends with chitosan or 2 θ (degrees)
PLA starting from commercial PET and recycled PET at dif-
Fig. 1. X-ray diffraction patterns for (a) Raw materials (b) PET/and R-PET/
ferent weight ratios. As a reference, it is also presented the PLA (c) PET/and R-PET/chitosan.
diffractograms of raw materials. The commercial PET
(Fig. 1a) displays a well-defined triclinic structure; in
agreement with PDF # 00-050-2275 file, the strong peaks talline phase. Different works reported only one peak for
are observed at 2h degrees of 16.19°, 17.2°, 21.3°, 22.36° chitosan at 2h ca. 20°; however, the crystallinity or more
and 25.5° which correspond to ð0 1  1Þ; ð0 1 0Þ; ð1
 1 1Þ; ð1
 10Þ packing in the main chain of our commercial chitosan
and (1 0 0) planes, respectively. On the other hand, R-PET was modified using a deacetylation process. In such case,
shows a diffraction pattern with a wide peak. In general, the first reflection is associated with two different kind of
the diffraction peaks of semicrystalline polymers are broad crystals; according to Hwang et. al. [24] the first peak at
and made up of the amorphous phase and reflections of the 5–10° has a unit cell characterized by a = 7.76, b = 10391,
crystalline planes. The diffraction pattern of R-PET reveals c = 10.30 Å and b = 90° whereas the second peak at
a semicrystalline structure with a peak characteristic 18–25° was characterized by a = 4.4, b = 10.0, c = 10.30 Å
reflection corresponding to the plane with a Miller index and b = 90° [24,25]. PLA samples display an orthorhombic
(1 0 0) at 2h = 25.5° [23]. structure with a strong peak of two overlapped signals
Selected XRD patterns of chitosan shows two peaks, at ca. 16.5° as well as lower intensity peaks around, 19° and
8.14° and 18.58°, which are commonly refer as a semicrys- 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) 285–299

(2 0 1), (1 1 1) and (1 0 2), (2 1 0) planes, respectively; PDF # stretching occurs in the 3438 cm1 region overlapping the
00-054-1917 file. OH stretch from the carbohydrate ring. The other peaks
The XRD patterns of polymer blends with PLA or chito- band assignment corresponds to ACAHA stretching in
san were also analyzed and the results are shown in Fig. 1b 2884 cm1, amide I, amide II and NAH bending vibrations
and c. From the X-Ray diffraction patterns, it is seen that of amine and amide II (1656, 1597 cm1), ACH2 bending
the spectra are quite similar resembling the XRD PET pat- (1425 cm1), ACH3 symmetrical deformation (1386 cm1),
tern. The observed reflections could indicate a weak inter- CAN stretching vibrations overlap the vibrations of carbo-
action when the chitosan or PLA are added to the PET hydrate ring ACAOACA (1154 cm1) and skeletal vibration
matrixes. It can be also observed that the signal of main of CAO stretching (1029 cm1) [13,28–34].
peak at 25.5° (1 0 0) shows a slightly decrease with the FTIR spectra of PET/PLA and R-PET/PLA blends are
relation of PLA or chitosan; the same trend is detected shown in Fig. 2b; where it is seen a strong absorption band
for the intensity of (ð0 1 1Þ and (0 1 0) planes. Similar results between 1768 cm1 and 1670 cm1, attributed to the
have been previously reported; i.e. the increasing of the superposition of the carbonyl stretching of PET and PLA.
amount of biopolymer reduces the intensity of the diffrac- It is well known that a reaction between these materials
tion peaks [25], characterizing the spectra as semicrystal- resulted in significant alterations in the bands, normally
line phases. associated with the formation of new bonds; in this
The most important point to highlight is that during respect, some researchers have been reported the shift of
extrusion process a recrystallization of R-PET blends was the bands as result of esterification in polymers blends of
obtained. It is well known that the semycristalline poly- Poly carbonate (PC)/PET, Bisphenol A-PC /Poly trimethyl-
mers have a metastable nanophase structure, where the ene terephthalate and in blends of poly vinylphenol
various nanophases can be occur like crystal, liquid, glass (PVPh)/poly vinylpyrrolidone (PVP) [35–37]. The non-
or mesophases. The structure is determined by the self- displacement of the commented bands can be related with
organization, crystallization and vitrification and it is another kind of interactions between polymers, probably
established during the material processing depending of hydrogen bonds interaction.
its thermal and mechanical history [26,27]. Thus, the In the combinations produced with PET/chitosan or
blends crystallized after the extrusion process as a result R-PET/chitosan, it was not possible to observe the chitosan
of the absence of the cooling system in the extraction zone phase or an important displacement in the wavenumber or
of the die, which stimulates the formation of polymeric bonds (Fig. 2c). This might be due to the used of low chito-
crystals. san amounts which are under the limit detection of the
FT-IR spectra were recorded in order to investigate the equipment (65%). Recent reports demonstrated that the
chemical structure of the raw materials and to study the hydrophilic character of chitosan can provoke a strong
characteristic signals of the blends after extrusion. interaction between the components forming the polymer
Fig. 2a–c shows the FT-IR spectra of raw materials and blend [38,39]. However, in our case FT-IR spectra did not
as-obtained polymer blends at the proposed compositions. show appreciable modifications in the evaluated range
FT-IR spectra show typical absorption bands of amorphous (Fig. 2a), suggesting that chitosan is physically integrated
or semi-crystalline PET similar to those reported in the lit- to the polymer matrix.
erature, except for the variation of the background due to One of the most common methods used to estimate
the film thickness dependent optical effects caused by polymer–polymer compatibility is to determine the glass
reflections from polymer surface and polymer/substrate transition temperature (Tg) of the blend and compare with
interface. FT-IR spectra of PET and R-PET (Fig. 2a) show the Tg of the component polymers. If one of the compo-
the main absorption bands as follows: at 3626 cm1 is nents is crystalline, the reduction in melting temperature
the OAH stretching, 3432 OAH stretching of ethylene gly- (Tm) can be used to investigate the blend compatibility
col end group, 3060 cm1 due to the aromatic CAH stretch- [40–43]. Then, estimating the changes in the Tm of the
ing, 2965–2906 cm1 is correlated with the aliphatic CAH blends, it is possible to study the miscibility of the blends
stretching, 2558–1961 cm1 aromatic summation band, if one of the components is crystalline in nature.
1723 cm1 (C@O stretch), 1619–1510 cm1 aromatic skel- Table 1 displays the results of DSC thermograms of the
eton stretching band, 1460–1341 CH2 deformation band, raw materials and polymer blends with different quanti-
1266–1102 cm1 C(O)AO stretching of ester group, ties of PLA or chitosan. Specifically, PET is characterized
1018 cm1 (1,4 aromatic substitution), 963 cm1 OACH2 by a glass transition temperature (65–140 °C) and melting
stretching of ethylene glycol segment in PET, 869 cm1 temperature (240–265 °C) [44,45]. However, the commer-
CAH deformation of two adjacent coupled hydrogens on cial PET used in this work displayed a glass transition tem-
an aromatic ring and finally, at 730 cm1 is presented the perature about 82 °C and two melting temperatures at
band associated with the out of plane deformation of the 126 °C and 243 °C. The first Tm is a result of the additives
two carbonyl substituents on the aromatic ring. that are commonly used during PET processing, which it
On the other hand, the FTIR spectrum of PLA was classi- is expected do not affect the degree of crystallinity,
fied into five regions, which corresponds to the following whereas the second Tm matched well with the value for
peaks band assignment: ACAHA stretching (2994.2 and PET [44].
2985 cm1), AC@O ester carbonyl (1746.2 cm1), ACAHA From these results, it can be noticeable that Tg cannot
deformation (1451.3, 1383 and 1358 cm1), ACAOA be detected in all compositions, which can be explained
stretching (1266, 1178, 1127, 1077 and 1039 cm1) and by considering that PET can be prevented from crystalliz-
ACACA stretching (867 cm1). In chitosan, the NAH ing by very fast cooling, and so obtained in completely
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 289

(b)
Φ= Phenyl group C=O C-C CO 1,4Φ CHΦ PET/PLA 95/5
(a) C-H

Absorbance (%)
PET
PET/PLA 90/10
C-C C-C CH CO1,4Φ
Absorbance (%)

C-H C=O CH
2 CHΦ
2
PET/PLA 85/15
R-PET
C-O
C=O C-H CO C-C R-PET/PLA 95/5
C-H
PLA
O-H and N-H C-N R-PET/PLA 90/10
Amide I and II N-H C-O
chitosan CH2 C=O PET
C-H C=O PLA PET signals

R-PET/PLA 85/15
COC
4000 3500 3000 2500 2000 1500 1000 3000 2500 2000 1500 1000
-1) -1)
Wavenumber (cm Wavenumber (cm

(c)
PET/chitosan 99/1

PET/chitosan 97.5/2.5
Absorbance (%)

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


Thermal transitions obtained for as-prepared selected blends using PLA and the cooling was carried out at room temperature which
chitosan with PET and R-PET. allows the formation of more crystalline regions.
Sample Tg (°C) Tm (°C) Tc (°C) Crystallinity of PET is usually induced by thermal crystalli-
zation and/or by stress or strain induced crystallization.
PET 82 126, 243 201
R-PET – 245 205 Specifically, as it is the case, thermally induced crystalliza-
PLA 69 161 126 tion occurs when polymer is heated above Tg and not
Chitosan 88 188 – quenched rapidly enough. In this condition, the polymer
PET/PLA 95/5 – 250 200 turns opaque due to spherulitic structure generated by
PET/PLA 90/10 – 250 202
thermal crystallization aggregates of non-oriented poly-
PET/PLA 85/15 – 250,158 205
R-PET/PLA 95/5 – 250 198 mers [44].
R-PET/PLA 90/10 – 250 201 A slightly increase in the melting point of PET or R-PET
R-PET/PLA 85/15 – 250,158 202 can be obtained by adding both biopolymers and remain
PET/chitosan 99/1 – 248 197
fairly constant ca. 250 °C when the biopolymers are misci-
PET/chitosan 97.5/2.5 – 249 197
PET/chitosan 95/5 – 251 203
ble in the matrix. Interestingly, the saturation of PLA into
R-PET/chitosan 99/1 – 248 208 polymer matrixes was reached up to an amount of
R-PET/chitosan 97.5/2.5 – 250 207 10 wt-%, longer quantities than this value provoke two
R-PET/chitosan 95/5 – 250 208 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) 285–299

the importance of the crystallinity on polymers blends, Wi  Wr


Weight loss% ¼ ð1Þ
indicating that stable values in the melting point can be Wi
reached for higher quantities of PLA in PET/PLA amorphous
blends produce from casting solution [46]. However, in our where Wi and Wr represent the initial and residual weights
study miscible phases are only obtained with amounts of the specimens, respectively.
under 15 wt-% using semi-crystalline biopolymer (PLA). Fig. 4 shows the trend of the weight loss% of as-pre-
The endothermic peak observed at 126 °C for commercial pared samples using PLA or chitosan and monitored during
PET is missed in the DSC spectra of blends extruded at 1200 h under accelerated weathering. For comparison, the
260 °C, corroborating the presence of small quantities of raw materials (except chitosan powders) and an additional
additives. sample that consist of a commercial BioPET specimen
Table 1 also shows that by increasing the amount of PLA (B-PET) were evaluated.
into the PET or R-PET matrixes slightly increases the crys- The evolution of weight loss indicate that only 600 h of
tallization temperature (Tc) from 198 °C to 205 °C. Similar UV irradiation were required to initiate the degradation of
tendency can also see for PET/Chitosan or R-PET/Chitosan PLA, at this time, the green polymer became brittle and it
polymer blends with a range of Tc between 197 °C and was difficult to handle as a consequence of being frag-
208 °C. Comparing the crystallization temperatures values mented and degraded by UV light from lamps [49]. In gen-
in the polymer blends, it can be mentioned that all the eral at 600 h, the blends showed slightly more weight loss
samples have an affinity to preserve Tc value of the PET percent in comparison with the raw materials.
matrix (PET or R-PET) and the small differences in the ther- Intervals between 1.12 and 2.23 wt-% of weight loss in
mal properties can be associated with a slightly decreased the blends were found depending on the amount of biopoly-
of the molecular weight of PLA during the processing blend mer and kind of matrix (commercial or recycled PET). The
or to the capacity of chitosan to act as nucleating agent results can divided in two tendencies; (i) the degradation
promoting a faster crystallization of PET in combination process of polymer blends augment as the biopolymer con-
with the crystallinity of the materials. centration increases independently of the matrix; and (ii)
Dissolution of selected phases in a polymer blend with the highest weight loss of polymer blends was obtained
suitable solvents is an etching method for morphology when it is used a recycled PET matrix, which may be associ-
evaluation of polymer blends [47]. Chloroform and acetic ated to the loss properties during the reprocessing. The
acid etching allows the removal of PLA and chitosan to polymer reprocessing provoked that the blend displayed
improve the analysis and reveal the morphology of dis- weaker and fragile material more susceptible to degrada-
persed biodegradable polymer. Fig. 3a-l shows typical tion than commercial PET [50–52]. Additionally, it is widely
SEM observations of PET/PLA, R-PET/PLA, PET/chitosan known that amorphous regions in polymers degrade more
and R-PET/chitosan. By comparison, SEM micrographs of easily than crystalline zones [53,54], i.e. in semicrystalline
the different compositions of PET with PLA are also shown polyesters, degradation first occurs in the amorphous
(Fig. 3b–g). All compositions exhibit a droplet (holes)- domains and subsequent in the crystalline regions [55].
matrix with heterogeneous structure; i.e. homogenously The weight loss of the amorphous R-PET was approximately
dispersed PLA domains are found in PET matrix. The mean 1.40% after 1200 hours which was higher than that obtained
diameter of the PLA agglomerates is ranged from 0.02 lm for commercial PET (ca. 0.97%).
to 1.9 lm (in PET) and 0.5–2.1 lm (in R-PET). In contrast On the other hand, it is noteworthy that there were not
to PET/PLA blends, the behavior of PET/chitosan blends found important differences between the weight loss of R-
was quite different. From Fig. 3h–l, cavities were not PET (1.4%) and B-PET (1.5%), in spite of the last one is
observed in ratios of 99/1 and 97.5/2.5. Additionally, PET/ obtained from natural resources.
Chitosan 95/5 was observed to have some agglomerates Weight loss percentages were used to calculate the deg-
of chitosan which were not dissolved by acetic acid. radation rate according to Eq. (2)
Thus, green polymers were integrated in PET and R-PET
matrixes uniformly but the size of agglomerates that form KW
Degradation rate ¼ ð2Þ
the filaments varies with the amount of biopolymer. These ATD
holes after etching demonstrate that such integration is
established of semi-spherical shape with a size fluctuating where K is a constant of conversion units to mm/year, W is
between 0.4 and 4.31 lm. Similar morphology has also the weight loss (g), A is the area exposed (cm2), T is the
been observed with previous reports in different polymer time (h) and D is the density of the material (g cm3).
blends using chitosan as biodegradable component [48]. The degradation rate of PET/PLA, PET/chitosan, R-PET/PLA
and R-PET/chitosan blends are shown in Fig. 5 corroborat-
3.2. Polymer blends degradation under accelerated ing the weight loss and FT-IR analyses.
weathering By comparing a six year natural weathering study with a
20,000 h accelerated weathering report, it has been shown
It has been well recognized that the weight loss of poly- that there is a rough relationship of approximately 1000 h
meric blends can indicate the resistance to degradation of accelerated QUV weathering being equal to one year of
process in a material. Thus, weight loss percentage has natural exposure. On the other hand, a rough correlation
been used to estimate the degradation of as obtained poly- used in the paint and coatings industry is 500–1500 h of
mers. Total weight loss% was calculated from the following accelerated exposure equaling approximately 1 year of real
equation: life exposure [56]. Thus, a relationship of 1000 h
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 291

(a)

PET
10 µm

(b) (c) (d)

PET/PLA 95/5 5 µm PET/PLA 90/10 5 µm PET/PLA 85/15 5 µm

(e) (f) (g)

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

(h) (i) (j)

PET/chitosan
PET/chitosan 99/1 5 µm 97.5/2.5 5 µm PET/chitosan 95/5 5 µm

(k) (l) (m)

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

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

accelerated weathering equaling to a year of natural weath- 3.2.1. Characterization of degraded polymer blends subjected
ering is used as a conservative data in this work [57]. to weathering chamber
Table 2 indicates the lifetime prediction of the synthe- It is well known that environmental factors, such as
sized materials studied in this work. The results estimate temperature, UV radiation and humidity are the main
that commercial PET will totally decompose in about causes of the degradation in polymer materials [58].
125 years whereas R-PET and B-PET decomposes in 76 Generally, the degradation is noticed as changes in the
and 86 years, respectively. As previously mentioned, this physical integrity and the loss of the structural proper-
is consistent with the strong dependency of being ties due to molecular bond scission [59]. Structural
semicrystalline B-PET and amorphous R-PET. From overall changes in all polymer blends were monitored by ana-
samples, R-PET blends with chitosan (95/5) and PLA (85/ lyzing the region of absorption bands (between
15) were found to degrade faster than the others composi- 2000 cm1 and 650 cm1) where susceptible bonds to
tions after 45 and 54 years, respectively. degradation are located.
292 A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299

3,0 Fig. 6 shows the FTIR spectra of (a) PET/PLA, (b) R-PET/
B-PET
PLA, (c) PET/chitosan and (d) R-PET/chitosan blends after
R-PET 95/5
2,5 being exposed to accelerate weathering for 1200 h. Using
PET 97.5/2.5 85/15
PET/PLA 85/15
commercial PET or R-PET as matrix in the polymer blends
R-PET/PLA 99/1
90/10
and PLA in different weight ratios (Fig. 6a and b), the fol-
2,0
Weight loss (%)

PET/chitosan 95/5 lowing observations can be drawn: (i) it is seen that C@O
B-PET
R-PET/chitosan bands were divided into multiple peaks; probably, due to
1,5 PLA R-PET
the characteristics of each matrix, i.e. the vibrations are
95/5
97.5/2.5
90/10
more intense in the commercial PET than that observed
95/5
1,0
PET
for R-PET. This splitting is commonly used to distinguish
99/1 between the crystalline (lower frequency) and amorphous
0,5 (higher frequency) phase [60]; (ii) the broader signal
PLA
observed around 1700 cm1 indicates the beginning of
photo-oxidation and degradation in the amorphous poly-
0,0
0 200 400 600 800 1000 1200 1400 ester backbones [61]. The peak in 1747 cm1 at higher
Time (h) wavenumber is encountered more defined than the band
in the amorphous phase; therefore it can be inferred that
Fig. 4. Weight loss% of PET, R-PET, B-PET, PLA and blends under the crystalline phase related to the C@O prevails after
accelerated weathering. accelerated weathering. Amorphous phase in the polymer
blends is less densely packed and therefore more suscepti-
ble 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
R-PET/chitosan 95/5 (1760 cm1) [63]. In carbonyl containing compounds the
R-PET/chitosan 97.5/2.5 (C@O) peaks are well known to shift to higher wavenum-
bers, as the electron-withdrawing effect of the /-substitu-
R-PET/chitosan 99/1
ent is larger [64]. This may explain the observations where
PET/chitosan 95/5
crystalline structure has shorter bonds length than in typ-
PET/chitosan 97.5/2.5
ical amorphous structures. Characteristic ester bonds
PET/chitosan 99/1 appear at 1318 cm1 and 1180 cm1 for PET and PLA,
Blends

R-PET/PLA 85/15 respectively; which become broad and weak after degrada-
R-PET/PLA90/10
tion process due to the chain scission of the CAO bonds
caused by the UV light and humidity. In the same way,
R-PET/PLA 95/5
the common CH2 bands (vibration of ethylene units) that
PET/PLA 85/15
initially appeared at 1410 cm1 and 1340 cm1 are joined
PET/PLA90/10 to form a new broad peak after degradation process
PET/PLA 95/5 between 1470 and 1420 cm1 [30,65]. Our results seems
0,0 -3 -3 -3 -3 to be in good agreement with Liang and coworkers whose
1,0x10 2,0x10 3,0x10 4,0x10
indicate that amorphous CH2 bending mode is found at
Degradation velocity (mm/year) 1455 cm1 corroborating that the degradation process
Fig. 5. Degradation rate of blends under accelerated weathering.
has begun [66]. Similar behavior was obtained with poly-
ester bands between 1120 cm1 and 1100 cm1 character-
istic 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
Table 2
polyester.
Lifetime prediction of raw materials and blends.
On the other hand, FT-IR spectra of degraded PET/chito-
Sample Lifetime prediction (year) san blends (Fig. 6c) show very small bands, indicating the
PET 125 typical splitting of the C@O signals with a shift of the sig-
R-PET 86 nals to higher wavenumbers [68,69]. A noticeable change
B-PET 76
in the band intensities is also observed, particularly,
PET/PLA 95/5 107
PET/PLA 90/10 103 ethylene vibrations (CH2) in the wavenumber range from
PET/PLA 85/15 76 1478 to 1400 cm1 and CAO bonds in the range of
R-PET/PLA 95/5 91 1198–954 cm1. Finally, the splitting bands correlated
R-PET/PLA 90/10 58 with crystalline-amorphous phase are missing for R-PET/
R-PET/PLA 85/15 54
PET/Chitosan 99/1 143
chitosan (Fig. 6d) polymer blends.
PET/Chitosan 97.5/2.5 93 Table 3 shows the thermal properties obtained by DSC
PET/Chitosan 95/5 60 analyses of polymer blends after degradation. Commercial
R-PET/Chitosan 99/1 56 PET displays a decrease of Tc from 201 °C to 170 °C with a
R-PET/Chitosan 97.5/2.5 55
Tm that remains fairly constant. R-PET presented a similar
R-PET/Chitosan 95/5 45
behavior with a decrease of Tc from 205 °C to 156 °C, a new
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 293

(a) (b)
PET-PLA 95/5
R-PET/PLA 95/5

Absorbance (%)
Absorbance (%)

PET-PLA 90/10

R-PET/PLA 90/10
C=O(c)
C=O(a) C-O
1747
1700

PET-PLA 85/15 R-PET/PLA 85/15

2000 1800 1600 1400 1200 1000 800 2000 1800 1600 1400 1200 1000 800
-1 -1
Wavenumber (cm ) Wavenumber (cm )

(c) (d)
R-PET/Chitosan 99/1
PET/chistosan 99/1
Absorbance (%)

Absorbance (%)
PET/chistosan 97.5/2.5
R-PET/Chitosan 97.5/2.5

C=O (c)
1746 1715
PET/chistosan 95/5 C=O (a) C-O R-PET/Chitosan 95/5
1700 CH2

2000 1800 1600 1400 1200 1000 800 2000 1800 1600 1400 1200 1000 800

Wavenumber (cm ) -1 Wavenumber (cm-1 )

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 – – 170 248 – – –
R-PET – – 156 227,245 – – –
PLA – – 120 150 – – –
* * * * * * *
Chitosan
PET/PLA 95/5 152 – 205 243 – 370 424,436
PET/PLA 90/10 152 – 202 242 – 370 424
PET/PLA 85/15 155 – 200 244 – 370 421
R-PET/PLA 95/5 152 – 194 241 – 370 422
R-PET/PLA 90/10 152 – 180 243 – 370 422
R-PET/PLA 85/15 152 – 180 239 – 370 422
PET/Chitosan 99/1 – 115, 184 240 340 – 420
PET/Chitosan 97.5/2.5 – 117, 165 191 242 343 – 421
PET/Chitosan 95/5 – 117, 171 190 249 344 – –
R-PET/Chitosan 99/1 – – – 241 340 – 422
R-PET/Chitosan 97.5/2.5 – – – – – – –
R-PET/Chitosan 95/5 – – – 240 – – 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 con- lated with the disentangled and break of the molecules in
stant. The crystallization temperature of the PET and the amorphous phase as well as some crystalline regions.
R-PET was found lower than before the weather chamber An unexpected feature was observed in blends or PET/
tests. The observed shifts in the Tćs were directly corre- 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) 285–299

melting point corresponding to PLA as well as the endo- From Fig. 7a and c, it is seen that commercial PET before
thermic peak at 370 °C corroborates that the degradation weathering decompose at 360 °C and at 500 °C presents
of PLA is taking place even with low PLA concentrations 14% of residual weight. Similar results were obtained for
[70]. R-PET before weathering (Fig. 7b and d) which decomposes
In Chitosan, there are several glass transition tempera- at 356 °C leaving 13.8% of residual weight (500 °C). The
ture values, this variability could be attributed to different TGA analyses showed that the chain scission of the ester
factors like physical state, molecular weight and de-acety- bonds (degradation of the polymer backbone) occurred in
lation degree [71]. From the Table 3, when chitosan was this interval of temperatures [75]. In Fig. 7a and b, it can
added to polymer matrixes two glass transition tempera- be seen that the initial step of PLA degradation starts at
tures appear, one at 115 and 117 °C from the plasticized lower temperatures than PET, at around of 307 °C leaving
chitosan as a result of the water supplied in the weathering negligible residue (2-wt%) above 397 °C. This is explained
test, and the second one at 165 and 171 °C attributed to due to when PLA is exposed to elevated temperatures
increase of the molecular movement due to dissociation (370 °C), undergoes thermal degradation, leading to the
of hydrogen bonds and starting of molecular scissions formation of lactide monomers [76].
[72]. The DSC studies of the blends with chitosan also From the blends TGA studies, is clearly shown that PLA
showed a degradation temperature at 340 °C. This has strong effect to increase the degradation on both
endothermic peak can be attributed to the degradation of matrixes (PET and R-PET) (Fig. 7a and b). The initial decom-
chitosan [68,73] and its effect into the PET matrix. Thus, position temperature of PET/PLA (Fig. 7 a) is reduced as the
both PLA and chitosan biopolymers induce the scission of content of PLA is increased in the blends, showing values of
PET bonds, producing shorter fragments and therefore, 287 °C, 250 °C and 150 °C with a residual weight of 17, 16.5
provoking the PET degradation at 420 °C [74]. and 13-% at 500 °C, for the PET/PLA weight ratios of 95/5,
Thermal stability of the polymer blends after 90/10 and 85/15, respectively. On the contrary, R-PET/PLA
accelerated weathering and weight loss as a function of blends (Fig. 7 b) show an increase in the total mass percent-
temperature was analyzed by TGA (Fig. 6a–d). age in the range of 250–327 °C which was correlated with

(a) PET (b) R-PET


100 100

80 80 R-PET/PLA (95/5)
Weight loss (%)

PET/PLA (95/5)
Weight loss (%)

PET/PLA (90/10) R-PET/PLA (90/10)


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

40 40

20 20

0 0
100 200 300 400 500 50 100 150 200 250 300 350 400 450 500
Temperature (°C) Temperature (°C)

(c) (d)
100 100
Weight loss (%)

80 80
Weight loss (%)

60 60

40 40

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


20 PET/Chitosan (97.5/2.5) 20 R-PET/Chitosan (97.5/2.5)
PET/Chitosan (99/1) R-PET/Chitosan (99/1)
PET R-PET
0 0
100 200 300 400 500 100 200 300 400 500
Temperature (°C) 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.
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 295

100 µm

(a) (b) (c)

PET/PLA 95/5 10 µm PET/PLA 90/10 10 µm PET/PLA 85/15 10 µm

(d) (e) (f)

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

(g) (h) (i)

PET/chitosan
PET/chitosan 99/1 10 µm 97.5/2.5 10 µm PET/chitosan 95/5 10 µm

(j) (k) (l)

R-PET/chitosan R-PET/chitosan
R-PET/chitosan 99/1 10 µm 97.5/2.5 10 µm 95/5 10 µm

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 loss can be observed as the amount of chitosan is
R-PET. The first stage of degradation changed to 300 °C, increased. The first degradation step varies with the weight
327 °C and 50 °C, with a residual weight of 19%, 20% at ratios of 99/1, 97.5/2.5 and 95/5 at 276 °C, 305 °C and
500 °C and 100-% at 466 °C, for the weight ratios of 95/5, 311 °C with a final weight of 16, 19 and 20.5 wt-%, respec-
90/10 and 85/15, respectively. The TGA analyses evidenced tively. Similar trend is again observed for R-PET/chitosan
that the incorporation of PLA and the accelerated weather- polymer blends obtaining degradation temperatures about
ing conditions decreases significantly the thermal stability 367 °C, 299 °C and 273 °C with a final weight of 16.5, 18.6
of the blends. and 20.6 wt-%, respectively, under the same weight ratios.
In the case of blends with chitosan (Fig. 7c and d), non- Then, even though chitosan enhances slightly the thermal
appreciable changes were observed in the temperature stability, reduces the degradation process in comparison
where degradation starts, however, a reduction in the mass with PLA.
296 A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299

Fig. 8a–l shows micrographs of representative weath- 0.004


ered blends after 1200 h. The micrographs of the weath-
ered filaments evidenced rough surfaces caused by the
effects of the UV light and water that simultaneously cause 0.003

Weight Loss (%)


surface microcracks; initially, microcracks or fissures were
formed due to the breaking of the backbone CAO bonds

R-PET/Chitosan (97.5/2.5)
and thereafter are dispersed in different points of the sur-

PET/Chitosan (97.5/2.5)
0.002

R-PET/Chitosan (99/1)

R-PET/Chitosan (95/5)
face. The morphology was quit affected by the PLA amount,

PET/Chitosan (95/5)
PET/Chitosan (99/1)
R-PET/PLA (85/15)
R-PET/PLA (90/10)
R-PET/PLA (95/5)
producing further degradation. For damage accumulation,

PET/PLA (85/15)
PET/PLA (90/10)
PET/PLA (95/5)
some regions are so weakened that cracking formation
starts microcracks and expanding over all surface of fila- 0.001
ments (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 0.000
[77], which are produced from cleavage at CAC bonds of Blends after 6 months
the main polymer chain [78]. In Fig. 9, it is shown a pro-
Fig. 10. Degradation trends for PET/PLA, R-PET/PLA, PET/Chitosan, R-PET/
posed model for photo-oxidative degradation of polymer chitosan blends in soil.
blends where are illustrated the chemical alterations and
crack formation during weathering test.
By comparison of the morphologies of PLA and chitosan, degradation of the blends was measured by monitoring
it is shown that chitosan blends presented small quantity the plastic weight loss during 6 months under real field
of microcracks (Fig. 8g–l); however, some deep holes conditions (Fig. 10). It can be noticed that the semicrystal-
formed by the swelling of chitosan powders were also evi- line (commercial) and amorphous (recycled) PET as well as
dent. The swelling may be caused by the water used in the PET from biodegradable bottle did not show important
QUV camera, leaving away the chitosan particles which variations in their weight after six months. It means that
produce the microcracks [79]. In this case, the photo-oxi- degradation takes a long time for some samples even
dation process of PET/chitosan blends occurs due to alter- though with some amounts seems to be that the material
ations such as cross-linking and chains scission; i.e. the disintegration has started. For example, PET/PLA 95/5 and
formation of carboxylic groups increasing the amount of PET/Chitosan 99/1 blends remained without changes dur-
polar groups on the surface during and after irradiation ing all the time, whereas R-PET/PLA (90/10), R-PET/PLA
time point out both the photo-oxidation of chitosan and (85/15), PET/Chitosan (95/5) and R-PET/Chitosan (95/5)
changes of its structure. Additional groups such as OH, blends showed certain degradation. It is also seen that
OOH, CO can also contribute to the rapid degradation of R-PET/chitosan (95/5) blends exhibited the highest loss
chitosan with the subsequent swelling and cracks forma- weight in comparison with the other polymer blends.
tion and in consequence, eventually, the PET disintegration Hence, degradation in soil is favored when higher amounts
[80]. of chitosan or PLA are added to PET matrixes and from
these the degradation is accelerated with R-PET.
3.3. Degradation of polymer blends in soil Accelerated QUV weathering tests were found to be consis-
tent with those obtained under real field conditions. To
From studies of PET degradation, kinetic models have prove these observations thermogravimetric analyses of
reported, based on accelerated experiments of hydrolytic the samples were also evaluated (Fig. 11a–d). The decom-
degradation and the life time of PET was found in the range position temperature (Td) is shifted to low values as PLA
from 16 to 48 years [80,81]. Unfortunately, there are few content is increased. Similarly, PET/chitosan with a 95/5
studies about the degradation in soil of PET [82,83]. The weight ratio showed a displacement up to reach a Td about

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


A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 297

(a) (b) R-PET


100 PET 100

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

Weight loss (%)


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

40 40 R-PET/PLA (85/15)
PLA
20 20

0 0
100 200 300 400 500 100 200 300 400 500
Temperature (°C) Temperature (°C)

(c) (d)
100 100

80 80

Weight loss (%)


Weight loss (%)

60 60

PET
40 40
R-PET
PET/Chitosan (99/1) R-PET/Chitosan (99/1)
20 PET/Chitosan (97.5/2.5) 20 R-PET/Chitosan (97.5/2.5)
PET/Chitosan (95/5) R-PET/Chitosan (95/5)

0 0
100 200 300 400 500 100 150 200 250 300 350 400 450 500
Temperature (°C) 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 tem- ing, which in turn is affected by two factors: quenched
perature remains fairly constant, corroborating the loss temperature and mobility of the biopolymer fraction. The
weight in soil. miscibility of the polymer-biopolymer blends is indepen-
According to the above results, the addition of higher dent of the PET matrix and the saturation with PLA is
amounts of PLA or chitosan on both PET matrixes can reached with 10 wt-%, whereas for chitosan all the compo-
modified its degradation mechanism. PLA degradation at sitions display an adequate miscibility. SEM analysis
temperature between 30–40 °C and 80-% of humidity showed that the green polymers were uniformly distrib-
could be due to the chemical hydrolysis that provokes that uted into the polymer matrixes, but the mean diameter
eventually the PET/PLA blends cleavage and degrades; sim- of the agglomerates and their dissolution varied with the
ilarly chitosan is degraded by the oxygen presence due to amount and kind of the biopolymer. The accelerating
the interaction in the water drop/surface of the blend sam- weathering tests indicate that the interaction between
ples interface. PET and chitosan favors more the degradation rate in com-
parison with PET/PLA blends, which it is thermally more
4. Conclusions stable. The best performance was obtained for PET/chito-
san polymer blend with a 95/5 weight ratio where an esti-
Blends of commercial and recycled PET/PLA and -/ mate time of about 45 years is required for its degradation.
Chitosan were prepared by extrusion process at 250 °C to Finally, all the results obtained from various analyses indi-
analyze the miscibility and degradability as a function of cate that the differences in the thermal stability and degra-
biopolymer content. The following conclusions can be dation rate suggest a certain degree of interaction between
drawn: blend components which are comparable to commercial
After extrusion process, XRD patterns and infrared spec- bottles of BioPET, which uses higher amounts of biopoly-
tra analysis showed a weak interaction between PET mer materials.
matrixes and biodegradable polymers suggesting type sec-
ondary bonds by hydrogen bridges or by electrostatic Acknowledgements
forces. In independence of the amount of biodegradable
polymers, their semi-crystalline character helps to the D. Palma-Ramírez is grateful for her postgraduate
recrystallization of R-PET during the extrusion reprocess- fellowship to CONACYT, COFAA and SIP-IPN. The authors
298 A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299

are also grateful for the financial support provided by [21] Hua H, Changsheng L, Wenjing W. Preparation and characterization
of chitosan/PEG/gelatin composites for tissue engineering. J Appl
CONACYT through the CB2009-132660 and CB2009-
Polym Sci 2009;114:1220–5.
133618 projects and to IPN through SIP 2014-0164 and [22] US Patent 2013037546. Methods of preparing para-xylene from
2014-0992 projects and SNI-CONACYT. The authors also biomass, Search International and National Patent Collections.
thank M.E.A.E. Rodríguez-Salazar and ROMFER SA CV [23] Xavier Nunes RA, Costa VC, De Araujo Calado VM, Tavares Branco JR.
Wear, friction, and microhardness of a thermal sprayed PET – poly
industries for their technical support. (ethylene terephthalate) coating. Mater Res 2009;12:121–5.
[24] Hwang KT, Kim JT, Jung ST, Cho GS, Park HJ. Properties of chitosan-
based biopolymer films with various degrees of deacetylation and
Appendix A. Supplementary material molecular weights. J Appl Polym Sci 2003;89:3476–84.
[25] Samuels RJ. Solid state characterization of the structure of chitosan
Supplementary data associated with this article can be films. J Polym Sci 1981;19:1081–105.
[26] Righetti MC, Tombari E, Angiuli M, Di Lorenzo ML. Enthalpy-based
found, in the online version, at http://dx.doi.org/10.1016/ determination of crystalline, mobile amorphous and rigid
j.eurpolymj.2014.10.016. amorphous fractions in semicrystalline polymers poly(ethylene
terephthalate). Thermochim Acta 2007;462:15–24.
[27] Wunderlich B. Reversible crystallization and the rigid–amorphous
References phase in semicrystalline macromolecules. Prog Polym Sci
2003;28:383–450.
[1] Ivar do Sul JA, Costa MF. The present and future of microplastic [28] Holland BJ, Hay JN. The thermal degradation of PET and analogous
pollution in the marine environment. Environ Pollut polyesters measured by thermal analysis-Fourier transform infrared
2014;185:352–64. spectroscopy. Polymer 2002;43:1835–47.
[2] Tian H, Gao J, Hao J, Lu L, Zhu C, Qiu P. Atmospheric pollution [29] Prasad SG, De A, De U. Structural and optical investigations of
problems and control proposals associated with solid waste radiation damage in transparent PET polymer films. Int J Spectrosc
management in China: a review. J Hazard Mater 2013;252– 2011;1:1–7.
253:142–54. [30] Drobota M, Aflori M, Barboiu V. Protein immobilization on
[3] Kasetaite S, Ostrauskaite J, Grazuleviciene V, Svediene, Bridziuviene Poly(ethylene terephthalate) films modified by plasma and
D. Camelina oil- and linseed oil-based polymers with chemical treatments. Digest J Nanomater Biostruct (DJNB)
bisphosphonate crosslinks. J Appl Polym Sci 2014;131:1–8. 2010;5:35–42.
[4] Rubio-Anaya M, Guerrero-Beltrán JA. Polímeros utilizados para la [31] Boerio FJ, Bahl SK, McGraw GE. Vibrational analysis of polyethylene
elaboración de películas biodegradables. Temas Selectos de terephthalate and its deuterated derivatives. J Polym Sci Polym Phys
Ingeniería de Alimentos 2012;6:173–81. Ed 1976;14:1029–46.
[5] Garrison TF, Kessler MR, Larock RC. Effects of unsaturation and [32] Miyake A. The infrared spectrum of polyethylene terephthalate. I.
different ring-opening methods on the properties of vegetable oil- The effect of crystallization. J Polym Sci 1959;38:479–95.
based polyurethane coatings. Polymer 2014;4:1004–11. [33] Cristina B, Xavier D, Serge E. Characterization of poly(ethylene
[6] López OV, Castillo LA, García MA, Villar MA, Barbosa SE. Food terephthalate) used in commercial bottled water. IOP Conf Series:
packaging bags based on thermoplastic corn starch reinforced with Mater Sci Eng 2009;5:1–5.
talc nanoparticles. Food Hydrocolloids 2014. http://dx.doi.org/ [34] Zhao Q, Jia Z, Li X, Ye Z. Surface degradation of unsaturated polyester
10.1016/j.foodhyd.2014.04.021. resin in Xe artificial weathering environment. Mater Des
[7] Petersson L, Kvien I, Oksman K. Structure and thermal properties of 2010;31:4457–60.
poly(lactic acid)/cellulose whiskers nanocomposite materials. [35] Al-Jabareen A, Illesca S, Maspotch MLI, Santana OO. Effects of
Compos Sci Technol 2007;67:2535–44. composition and transesterification catalysts on the physico-
[8] Qu P, Gao Y, Wu G-F, Zhang L-P. Nanocomposites of poly(lactic acid) chemical and dynamic properties of PC/PET blends rich in PC. J
reinforced with cellulose nanofibrils. Bioresources 2010;5:1811–23. Mater Sci 2010;45(24):6623–33.
[9] Choudhary P, Mohanty S, Nayak SK, Unnikrishnan L. Poly(L-lactide)/ [36] Na SK, Kong BG, Choi C, Jang MK, Nah JW. Transesterification and
polypropylene blends, evaluation of mechanical, thermal, and compatibilization in the blends of bisphenol-A polycarbonate and
morphological characteristics. J Appl Polym Sci 2011;121:3223–37. poly(trimethylene terephthalate). Macromol Res 2005;13(2):
[10] Zhu M, Li S, Li Z, Lu X, Zhang S. Investigation of solid catalysts for 88–95.
glycolysis of polyethylene terephthalate. Chem Eng J [37] Kuo SW, Chang FC. Significant thermal property and hydrogen
2012;185:168–77. bonding strength increase in poly(vinylphenol-co-vinylpyrrolidone)
[11] Girija BG, Sailaja RRN, Madras G. Thermal degradation and copolymer. Polymer 2003;44(10):3021–30.
mechanical properties of PET blends. Polym Degrad Stab [38] Shieh YT, Yang YF. Significant improvements in mechanical property
2005;90:147–53. and water stability of chitosan by carbon nanotubes. Eur Polym J
[12] Welle F. Twenty years of PET bottle to bottle recycling – an 2006;42:3162–70.
overview. Resour Conserv Recycl 2011;55:865–75. [39] Rueda DR, Secall T, Bayer RK. Differences in the interaction of water
[13] Garlotta D. A literature review of poly (lactic acid). J Polym Environ with starch and chitosan films as revealed by infrared spectroscopy
2001;9:63–84. and differential scanning calorimetry. Carbohydr Polym
[14] Tokiwa Y, Calabia B. Biodegradability and biodegradation of 1999;40:49–56.
poly(lactide). Appl Microbiol Biotechnol 2006;72:244–51. [40] Adoor SG, Manjeshwar LS, Krishna Rao KSV, Naidu B, Aminabhavi
[15] Auras RA, Harte B, Selke S, Hernández R. Mechanical, physical, and TM. Solution and solid-state blend compatibility of poly(vinyl
barrier properties of poly(lactide) films. J Plastic Film Sheeting alcohol) and poly(methyl methacrylate). J Appl Polym Sci
2003;19:123–35. 2006;100:2415–21.
[16] Bautista-Baños S, Hernández-Lauzardo AN, Velázquez del Valle MG, [41] Jawalkar SS, Adoor SG, Sairam M, Nadagouda MN, Aminabhavi TM.
Hernández-López M, Ait-Barka E, Bosquez-Molina E, et al. Chitosan Molecular modeling on the binary blend compatibility of poly(vinyl
as a potential natural compound to control pre and postharvest alcohol) and poly(methyl methacrylate): an atomistic simulation
diseases of horticultural commodities. Crop Prot 2006;25:108–18. and thermodynamic approach. J Phys Chem B 2005;109:15611–20.
[17] Kint D, Muñoz-Guerra S. A review on the potential biodegradability [42] Feng L, Bian X, Li G, Chen Z, Cui Y, Chen X. Determination of ultra-
of poly(ethylene terephthalate). Polym Int 1999;48:346–52. low glass transition temperature via differential scanning
[18] Siracusa V, Rocculi P, Romani S, Dalla Rosa M. Biodegradable calorimetry. Polym Testing 2013;32:1368–72.
polymers for food packaging: a review. Trends Food Sci Technol [43] MacKnight WJ, Karasz FE, Fried JR. Solid state transition behavior of
2008;19:634–43. blends. Chapter 5. Polym Blends 1978:185–242.
[19] Zou H, Yi C, Wang L, Xu W. Crystallization, hydrolytic degradation, [44] Demirel B, Yaras AH, Elcicek BF. Crystallization behavior of PET
and mechanical properties of poly (trimethylene terephthalate)/ materials. Bil Enst Dergisi Cilt 2011;13:26–35.
poly(lactic acid) blends. Polym Bull 2010;64:471–81. [45] Strobl G. The Physics of Polymers: Concepts for Understanding Their
[20] Malgorzata J, Kensuke S, Pierre G, Eric G. Influence of chitosan Structures and Behavior. Berlin, Germany: Springer; 1997.
characteristics on polymer properties, I: crystallographic properties. [46] Chen H, Pyda M, Cebe P. Non-isothermal crystallization of PET/PLA
Polym Int 2003;52:198–205. blends. Thermochim Acta 2009;492:61–6.
A.M. Torres-Huerta et al. / European Polymer Journal 61 (2014) 285–299 299

[47] Nisha P, Moghe N, Riccobono O, Tyagi S, Rajagopalan S. Novel [65] Dieval F, Khoffi F, Mir R, Chaouch W, Le Nouen D, Chakfe N, et al.
approaches in microstructure evaluation of polymer blends. Microsc Long-term biostability of pet vascular prostheses. Int J Polym Sci
Microanal 2005;11(suppl. S02):2110–1. 2012;2012:1–14.
[48] Correlo VM, Boesel LF, Bhattacharya M, Mano JF, Neves NM, Reis RL. [66] Liang CY, Krimm S. Infrared spectra of high polymers: Part IX.
Properties of melt processed chitosan and aliphatic polyester blends. Polyethylene terephthalate. J Mol Spectrosc 1959;3:554–74.
Mater Sci Eng: A 2005;403:57–68. [67] Donelli I, Freddi G, Nierstrasz VA, Taddei P. Surface structure and
[49] Muthukumar T, Aravinthan A, Mukesh D. Effect of environment on properties of poly-(ethylene terephthalate) hydrolyzed by alkali and
the degradation of starch and pro-oxidant blended polyolefins. cutinase. Polym Degrad Stab 2010;95:1542–50.
Polym Degrad Stab 2010;95:1988–93. [68] De Britto D, Campana-Filho SP. Kinetics of the thermal degradation
[50] Kráčalík M, Pospisil L, Slouf M, Mikesova J, Sikora A, Simonik J, et al. of chitosan. Thermochim Acta 2007;465:73–82.
Effect of glass fibers on rheology, thermal and mechanical properties [69] Pieróg M, Ostrowska-Czubenko J, Gierszewska-Druzynska M.
of recycled PET. Polym Compos 2008;29:915–21. Thermal degradation of double crosslinked chitosan membranes.
[51] Pawlak A, Plutaa MB, Morawieca J, Galeskia A, Pracellab M. Prog Chem Appl Chitin Derivatives XVII 2012:67–74.
Characterization of scrap poly(ethylene terephthalate). Eur Polym J [70] Arrieta MP, López J, Ferrándiz S, Peltzer A. Characterization of PLA-
2000;36:1875–84. limonene blends for food packaging applications. Polym Testing
[52] Spinacé MAS, De Paoli MA. Characterization of poly(ethylene 2013;32:760–8.
terephtalate) after multiple processing cycles. J Appl Polym Sci [71] Zhang L, Kosaraju SL. Biopolymeric delivery system for controlled
2001;80:20–5. release of polyphenolic antioxi-dants. Eur Polym J 2007;43:
[53] Tudorachi N, Cascabal CN, Rusu M, Pruteanu M. Testing of polyvinyl 2956–66.
alcohol and starch mixtures as biodegradable polymeric materials. [72] Mucha M, Pawlak A. Thermal analysis of chitosan and its blends.
Polym Testing 2000;19:785–99. Thermochim Acta 2005;427:69–76.
[54] Huang SJ, Ho LH, Huang MT, Koenig MF, Cameron JA. Similarities and [73] Fernandes LL, Resende CX, Tavares DS, Soares GA, Castro LO, Ganjeiro
differences between biodegradation and non-enzymatic JM. Cytocompatibility of chitosan and collagen-chitosan scaffolds for
degradation. Stud Polym Sci 1994;12:3–10. tissue engineering. Polímeros 2011;21(1):1–6.
[55] Anderson JM, Shive MS. Biodegradation and biocompatibility of PLA [74] Chen DQ, Wang YZ, Hu XP, Wang DY, Hai M, Yang B. Flame-retardant
and PLGA microspheres. Adv Drug Deliv Rev 1997;28(1):5–24. and anti-dripping effects of a novel char-forming flame retardant for
[56] Martin D, Eng P. Advanced thin film geomembrane technology for the treatment of poly(ethylene terephthalate) fabrics. Polym Degrad
biocell liners and covers (available from Layfield Geosynthetics and Stab 2005;88:349–56.
Industrial Fabrics Ltd) (2005). [75] Turnbull L, Liggat JJ, MacDonald WA. Thermal degradation chemistry
[57] Wagner N, Ramsey B. QUV accelerated weathering study: analysis of of poly(ethylene naphthalate) – a study by thermal volatilisation
polyethylene film and sheet samples. Technical Document by GSE analysis. Polym Degrad Stab 2013;98:2244–58.
Lining Technology, Inc., 2003. Houston, TX, USA [76] Lim LT, Auras R, Rubino M. Processing technologies for poly(lactic
[58] Andrady AL, Hamid HS, Torikai A. Effects of climate change and UV-B acid). Prog Polym Sci 2008;33:820–52.
on materials. Photochem Photobiol Sci 2003;2:68–72. [77] Babanalbandi A, Hill DJT, O’Donnell JH, Pomery PJ, Whittaker A. An
[59] Fischer HR, Semprimosching C, Mooney C, Rohr T, Ernst RH, electron spin resonance study on gamma-irradiated poly(L-lactic
Verkuijlen MHW. Degradation mechanism of silicone glues under acid) and poly(D, L-lactic acid). Polym Degrad Stab
UV irradiation and options for designing materials with increased 1995;50:297–304.
stability. Polym Degrad Stab 2013;98:720–6. [78] Copinet A, Bertrand C, Govindin S, Coma V, Couturier Y. Effects of
[60] Andanson JM, Kazarian SG. In situ ATR-FTIR spectroscopy of ultraviolet light (315 nm), temperature and relative humidity on the
poly(ethylene terephthalate) subjected to high-temperature degradation of polylactic acid plastic films. Chemosphere
methanol. Macromol Symp 2008;265:195–204. 2004;55:763–73.
[61] Zhang WR, Hinder SJ, Smith R, Lowe C, Watts JF. An investigation of [79] Matuana LM, Jin S, Stark NM. Ultraviolet weathering of HDPE/wood-
the effect of pigment on the degradation of a naturally weathered flour composites coextruded with a clear HDPE cap layer. Polym
polyester coating. J Coat Technol Res 2011;8:329–42. Degrad Stab 2011;96:97–106.
[62] Mai F, Habibi Y, Raquez JM, Dubois P, Feller JF, Peijs T, et al. [80] Eubeler JP, Bernhard M, Knepper TP. Environmental biodegradation
Poly(lactic acid)/carbon nanotube nanocomposites with integrated of synthetic polymers II. Biodegradation of different polymer groups.
degradation sensing. Polymer 2013;54:6818–23. TrAC Trends Anal Chem 2010;29:84–100.
[63] Zhang J, Sato H, Tsuji H, Noda I, Ozaki Y. Differences in the [81] Müller RJ, Kleeberg I, Deckwer WD. Biodegradation of polyesters
CH3. . .O@C interactions among poly(l-lactide), poly(l-lactide)/ containing aromatic constituents. J Biotechnol 2001;86:87–95.
poly(d-lactide) stereocomplex, and poly(3-hydroxybutyrate) [82] Radulovic J. Degradation of polyethylene terephthalate in natural
studied by infrared spectroscopy. J Mol Struct 2005;735:249–57. conditions. Sci Techn Rev LVI 2006:45–51.
[64] Shimizu T, Ohkawa Tanaka Y, Kutsumizu S, Yano S. Structural [83] Platt DK, Rapra L. Technology: biodegradable polymers: Market
Studies of Poly(1H,1H-fluoroalkyl alpha fluoroacrylate)s by infrared Report Rapra Technology (2006).
spectroscopic analysis. Macromolecules 1996;29:3540–4.

S-ar putea să vă placă și