Sei sulla pagina 1di 7

Polymer Degradation and Stability 95 (2010) 1022e1028

Contents lists available at ScienceDirect

Polymer Degradation and Stability


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

Chemical recycling of post-consumer PET wastes by glycolysis


in the presence of metal salts
R. Lpez-Fonseca a, I. Duque-Ingunza a, B. de Rivas a, S. Arnaiz b, J.I. Gutirrez-Ortiz a, *
a

Chemical Technologies for Environmental Sustainability Group, Department of Chemical Engineering, Faculty of Science and Technology,
Universidad del Pas Vasco/EHU, P.O. Box 644, E-48080 Bilbao, Spain
b
GAIKER-IK4, Parque Tecnolgico, Edicio 202, E-48170 Zamudio, Bizkaia, Spain

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 5 November 2009
Received in revised form
3 February 2010
Accepted 10 March 2010
Available online 16 March 2010

Chemical recycling of poly(ethylene terephthalate) (PET) has been the subject of increased interest as
a valuable feedstock for different chemical processes. In this work, glycolysis of PET waste granules was
carried out using excess ethylene glycol in the presence of different simple chemicals acting as catalysts,
namely zinc acetate, sodium carbonate, sodium bicarbonate, sodium sulphate and potassium sulphate.
Comparable high yields (z70%) of the monomer bis(2-hydroxyethyl terephthalate) were obtained with
zinc acetate and sodium carbonate as depolymerisation catalysts at 196  C with a PET:catalyst molar ratio
of 100:1 in the presence of a large excess of glycol. The puried monomer was characterised by elemental
analysis, differential scanning calorimetry, infrared spectroscopy, and nuclear magnetic resonance. These
results revealed that, although the intrinsic activity of zinc acetate was signicantly higher than that of
sodium carbonate, this latter salt could indeed act as an effective, eco-friendly catalyst for glycolysis. Also
an exploratory study on the application of this catalytic recycling technology for complex PET wastes,
namely highly coloured and multi-layered PET, was performed.
2010 Elsevier Ltd. All rights reserved.

Keywords:
Post-consumer PET waste
Chemical recycling
Glycolysis
Zinc acetate
Sodium carbonate

1. Introduction
Increased environmental awareness, legislative measures, and
public demand for environmental sustainability are leading to an
increased interest in plastics recycling. Plastics or polymer recycling
is very important for a wide number of reasons such as conservation of oil, reduction of greenhouse gas emissions, saving of landll
space, conservation of energy, and benets of reuse [1]. Poly
(ethylene terephthalate) (PET), a semi-crystalline thermoplastic
polyester that is extensively used in diverse applications etextiles,
high strength bres, photographic lms, disposable soft-drink
bottles, and otherse is also one of the largest components of the
post-consumer plastics in landlls.
Chemical PET decomposition and conversion into reusable
chemical products of added value is one of the important recycling
strategies for this material. Petcore recently announced that European post-sorting PET collection reached 1.13 million tonnes in
2007, an increase of about 20% on 2006. In 2007 just about 40% of
all PET bottles in the market were collected for recycle [2]. Four
main approaches have been proposed for PET recycling, namely
primary or in-plant recycle of the scrap material of controlled
* Corresponding author. Tel.: 34 94 6012683; fax: 34 94 6015963.
E-mail address: joseignacio.gutierrez@ehu.es (J.I. Gutirrez-Ortiz).
0141-3910/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2010.03.007

history, secondary or mechanical, tertiary or chemical and quaternary or recovery of energy through incineration [3]. Among recycling techniques, the only acceptable one according to the
principles of sustainable development is the tertiary or chemical
recycling since it leads to the formation of the raw materials or
monomers. Chemical recycling processes for PET are divided as
follows: (i) glycolysis, (ii) methanolysis, (iii) hydrolysis and, (iv)
other processes such as aminolysis or ammonolysis [4e6]. Mild
recycling conditions are best preferred to obtain higher recovery of
monomers that can be readily assimilated into the polymerisation
technology, easier separation of products, lower environmental
impact, and simpler or more economical processes in terms of low
energy consumption.
Glycolysis can be described as a molecular depolymerisation
process by transesterication between PET ester groups and a diol,
usually ethylene glycol (EG) in excess, to obtain the monomer bis(2hydroxyethyl terephthalate) (BHET), according to the following
reaction scheme,

PETn n  1EG

!
nBHET

(1)

In this process ester linkages are broken and replaced with


hydroxyl terminals. The main advantage of this strategy is that it
can be easily integrated into a conventional PET production plant

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

and the recovered BHET can be blended with fresh BHET [7].
Further, the monomer (and higher oligomers) thus obtained can be
used as building blocks to synthesise other polymers with higher
economical values such as unsaturated polyesters, polyurethane
foams, polyisocyanurate foams, copolyesters, polyurethane coatings, alkyd resins, low temperature curable resins, or UV curable
resins [8e13]. These numerous applications, in turn, provide an
economical exibility when using the raw recycled monomer for
the most protable option upon demand.
Moreover, glycolysis avoids the use of, supercritical in some
cases, methanol used in methanolysis and limits the environmental
impact of the recycling strategy since strong acids or alkalis
(necessary in PET hydrolysis) are not employed. The reaction is
typically performed under relatively mild conditions, 180e260  C
and atmospheric pressure, but metal catalysts are often required to
increase the rate. Two major drawbacks of this process are, one
hand, the non-biodegradable and toxic nature of heavy metal
catalysts such as Zn salts, which are reported as the most active
glycolysis catalysts [14e17] and, on the other hand, the low selectivity of the reaction towards BHET avoiding the formation of
higher oligomers (dimers and trimers), which are difcult to purify
by conventional separation methods.
The primary goals of this study are the optimisation of the main
operation variables of the glycolysis process (time, temperature, EG:
PET molar ratio, nature of catalyst and concentration) giving a special
attention to searching alternative, eco-friendly active salts as catalysts for the reaction. Hence, an exploratory study of four metal salts
(sodium carbonate, sodium bicarbonate, sodium sulphate and
potassium sulphate) was conducted and results were compared with
those obtained with the most active catalyst reported in the literature, zinc acetate. The chemical structure of the glycolysed products
was analysed as well. Finally some preliminary results on the catalysed glycolysis of several coloured (green, amber, blue, white and
silver), and multi-layered PET wastes were discussed.
2. Experimental
Clear, highly coloured and complex multi-layered postconsumer PET wastes were kindly provided by Ecoembalajes
Espaa, S.A. (ECOEMBES), the Spanish integrated management
system for the light packaging waste. Samples were rst conditioned with an aqueous diluted (1%wt) sodium hydroxide solution
for about 1 h to remove any surface impurities and subsequently
washed with water and dried in an oven at 80  C. PET scraps having
a particle size between 2 and 4 mm were previously cut with
a cryogenic rotary cutter (Retsch ZM 2000) to reduce the particle
diameter to 0.25 mm. Ethylene glycol (Sigma Aldrich), zinc acetate
(Probus), sodium carbonate (Fluka), sodium bicarbonate (Panreac),
sodium sulphate (Probus) and potassium sulphate (Merck) were of
analytical reagent grade.
Glycolysis experiments were carried out in a 500 cm3, a threenecked, at-bottom glass reactor equipped with a thermometer
and a reux condenser operated at atmospheric pressure. A
magnetic stirrer was used to ensure proper mixing and turbulence.
The glycolysis reaction was allowed to proceed up to 8 h. In all runs
30 g of PET, equivalent to 0.16 mol of repeating unit in the PET chain,
were charged to the reactor. Table 1 summarises the experimental
conditions of the runs carried out in this work. The reaction vessel
and the mixture of EG reactant and the catalyst were preheated to
the selected temperature (165e196  C) prior to the addition of the
PET granules in order to minimise the time required to reach the
reaction temperature. The EG:PET molar ratio varied between 5.4:1
and 7.6:1. Zinc acetate, sodium carbonate, sodium bicarbonate,
sodium sulphate and potassium sulphate were used as catalysts.
After a specied time interval at the reaction temperature,

1023

Table 1
Operation conditions of the glycolysis runs carried out in this work.
Runa Type of
PET
waste

Catalyst Amount of catalyst,


Temperature, Amount of

C
EG, g (EG:PET salt
g (PET:catalyst
molar ratio)
molar ratio)

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28

196
196
180
165
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196
196

Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Clear
Green
Amber
Silver
Blue
White
Multi-layered

73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
55.7
36.8
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5
73.5

(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(5.7:1)
(3.8:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)
(7.6:1)

Zn(Ac)2
None
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Na2CO3
NaHCO3
Na2SO4
K2SO4
Na2CO3
Na2CO3
Na2CO3
Na2CO3
Na2CO3
NaHCO3
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2
Zn(Ac)2

0.35 (98:1)
e
0.35 (98:1)
0.35 (98:1)
0.09 (381:1)
0.14 (245:1)
0.035 (980:1)
0.0175 (1960:1)
0.0105 (3300:1)
0.0035 (9800:1)
0.09 (381:1)
0.09 (381:1)
0.045 (381:1)
0.034 (381:1)
0.058 (381:1)
0.071 (381:1)
0.168 (98:1)
0.068 (245:1)
0.033 (503:1)
0.017 (980:1)
0.005 (3300:1)
0.069 (190:1)
0.35 (98:1)
0.35 (98:1)
0.35 (98:1)
0.35 (98:1)
0.35 (98:1)
0.35 (98:1)

30 g of PET (0.25 mm particle size) waste were used in all the experiments.

the reactor vessel was removed from the heating mantle and
quenched quickly in an ice bath. The vessel was subsequently
opened, and the product removed. It is worth pointing out that
during the initial steps of the reaction the mixture was heterogeneous since two phases were clearly distinguishable, solid (PET)
and liquid (ethylene glycol and catalyst). However, after a certain
time interval the solid phase was no longer observed and the
reaction occurred in one single phase.
Hot distilled water (400 g) was then added in excess to the
reaction mixture under vigorous agitation since BHET is known to
be quite soluble in boiling water. Hot water also dissolved the
catalyst employed, and probably to some extent higher oligomers
such as dimers and trimers. While still hot, the suspension was
quickly ltered. The product was separated into solid and aqueous
phases using a sintered glass lter (Whatman glass microber
binder free, grade GF/C-1.2 mm) under vacuum. The glycolysed
product was obtained as a residue after ltration. The ltrate
contained ethylene glycol, BHET, and small quantities of few watersoluble oligomers. This was heated until a clear mixture was
obtained. Afterwards it was ltrated with a sintered glass lter
(Whatman glass microber binder free, grade GF/C-0.7 mm) under
vacuum. This second ltrate was collected and stored in a refrigerator at 5  C for 16 h to provoke the precipitation of white crystalline
BHET akes. After ltration, BHET was dried in an oven at 60  C for
30 h, and weighed on an analytical balance to estimate the BHET
molar yield according to the following equation.

Y%

WBHET;f =MWBHET
 100
WPET;0 =MWPET

(2)

where WPET,0 and WBHET,f refer to the initial weight of PET and the
weight of BHET at a specic reaction time, respectively. MWBHET
and MWPET are the molecular weights of BHET (254 g mol1) and
the PET (192 g mol1) repeating unit, respectively.

1024

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

The puried BHET product was subjected to various characterisation techniques, namely elemental analysis, differential scanning
calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR)
and 1H nuclear magnetic resonance (NMR) were used to characterise the degradation products obtained at various stages of the
process. Elemental analysis was carried out with a EuroVector
Elemental Analyzer apparatus. DSC was performed on a Mettler
Toledo DSC822e instrument (sample size 3 mg) under a purge gas
ow of 15 cm3 min1, and with a heating rate of 10  C min1 in the
range 25e350  C. An FTIR Nicolet Proteg 460 spectrometer was
used in transmission mode to record spectra from degradation
products on KBr wafers (100 mg with a dilution of 1:50, 13 mm in
diameter) in the 4000e500 cm1 range. The instrument was
equipped with a DTGS detector averaging 50 scans with a resolution of 2 cm1. 1H NMR analysis was recorded with a Bruker AV500
spectrometer operating at 500 MHz. The spectra were obtained in
d6-acetone solution. Details of the experimental procedure for
separation and quantitative analysis of the reaction products are
schematically depicted in Fig. 1.
3. Results and discussion
3.1. Effect of reaction conditions and analysis of glycolysed products
Glycolysis involves a partial to full depolymerisation where the
high molecular weight polymer is usually contacted with ethylene
glycol. The glycol causes chain scission by attacking the ester
linkages along the polymer backbone. The objective is to recover
puried low molecular weight oligomers, preferably the single
monomer bis(hydroxyethyl) terephthalate, for repolymerisation.
As a rst approach the efciency of zinc acetate, Zn(Ac)2, as catalyst
for glycolysis and the effect of reaction time were examined (run 1).
For this purpose the following operation conditions were selected,
196  C, a marked excess of ethylene glycol with an EG:PET molar
ratio of 7.6:1, and a PET:Zn(Ac)2 molar ratio of 100:1 (equivalent to
1% based on the weight of PET). Results with increasing reaction
progress at 0.33, 0.66, 1, 2, 4, 5 and 8 h in terms of BHET yield, as
calculated by Equation (2), are shown in Fig. 2. Also an experiment

was carried out in the absence of catalyst with the same amounts
of reactants at the same temperature (run 2). Two observations
were clearly noticeable. Firstly, without adding Zn(Ac)2, a high PET
conversion could only be attained after a reaction time as long as
8 h. Secondly, almost 70% conversion, that corresponding to the
equilibrium under the selected conditions, could be obtained after
1 h. This yield was equivalent to the formation of about 25 g of BHET
from 30 g of PET (83% on a weight basis). This observation clearly
evidenced the catalytic role of the zinc salt in promoting the
glycolytic depolymerisation within a reasonable period of time. As
can be seen in Fig. 2, even at the very initial stage (0.33 h) glycolysis
proceeded at a very high rate. Longer reaction time led to further
but quite moderate increase in BHET yield. Consequently, 1 h was
selected as the optimum reaction interval at which the highest PET
conversion was reached and thereafter it remained virtually
constant. On the other hand, from these results it was noticed that
the glycolysis of PET was a reversible equilibrium reaction, the
reverse reaction being polycondensation. During the initial stages
of depolymerisation, polycondensation proceeded at extremely low
rate and hence it could be ignored. With an increase in time, this
reverse reaction occurred to a signicant extent thereby decreasing
the BHET yield. Additional experiments carried out at lower
temperatures (165, 175 and 185  C) and relatively long reaction
time revealed that equilibrium yield decreased with decreasing
temperature (from 67% at 185  C down to 34% at 165  C), this
suggesting that the studied reaction is endothermic.
Fig. 3 shows the inuence of reaction temperature in the range
of 165e196  C at an EG:PET molar ratio 7.6:1 and PET:Zn(Ac)2 molar
ratio 100:1, after a reaction time interval of 1 h (runs 2e4). Note
that 196  C is the normal boiling point of EG. An obvious dependence of the conversion with temperature was noted although the
difference in conversion between 180 and 196  C was less than 1.5%.
Next an attempt was made to optimise the concentration of the zinc
salt (PET:Zn(Ac)2 molar ratio) varying this parameter from about
10 000:1 to 100:1 (Fig. 4). Up to seven Zn(Ac)2 concentrations were
thus tested (with an amount of catalyst ranging between 0.0035
and 0.35 g). The operation conditions for this set of experiments

100
No catalyst

90
80
70

Yiel d, %

60
50
40
30
20
10
0
0

10 11

Time, h
Fig. 1. Schematic diagram of the reaction and analytical procedure for the glycolysis of
PET.

Fig. 2. Evolution of BHET yield with time in the presence and absence of zinc acetate
(196  C, EG:PET molar ratio 7.6:1, PET: Zn(Ac)2 molar ratio 100:1).

100

100

90

90

80

80

70

70

60

60

Yield, %

Yield, %

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

50
40

1025

50
40

30

30

20

20

10

10

0
160

165

170

175

180

185

190

195

200

Temperature, C
Fig. 3. Effect of reaction temperature on BHET yield (EG:PET molar ratio 7.6:1, PET: Zn
(Ac)2 molar ratio 100:1).

were 196  C, EG:PET molar ratio 7.6:1, after a reaction time interval
of 1 h (runs 2, 5e10). A signicant BHET yield (about 44%) was
obtained with a catalyst concentration as low as 3300:1. However,
a higher concentration was required (above 380:1) to attain the
highest yield, close to 70%. Within the experimental error
increasing the concentration up to 245:1e100:1 did not give an
added benet. With this optimum amount of the zinc salt (0.09 g)
the inuence of EG:PET molar ratio from 3.8:1 to 7.6:1 was analysed
(runs 5, 11 and 12). Fig. 5 indicates that above a certain value, 5.7:1,
the variation in BHET yield (64%) was hardly discernible [18].

EG:PET molar ratio


Fig. 5. Effect of EG:PET molar ratio on BHET yield (196  C, PET:catalyst molar ratio
380:1, 1 h).

In order to check the purity of the recovered BHET after PET


waste glycolysis, this was analysed by means of a number of
analytical techniques, and the results were compared with those
corresponding to a commercial BHET sample provided by the
Aldrich Chemical Co. The results of the elemental analysis,
expressed as the molar percentages of C, H and O, of the puried
BHET product were 56.4% C, 5.4% H and 38.2% O, in fairly good
agreement with 57.1% C, 5.9% H and 37.0% O found for the
commercial sample. As for 1H NMR analysis (Fig. 6), signals at d 8.1,
4.4, 4.2 and 3.9 ppm were noticed. The signal at d 8.1 ppm indicated
the presence of the four aromatic protons of the benzene ring.

100
90
80
70

Yield, %

60

50
40
30
20

Zn(Ac)2
Na2CO3

10

0
0

333:1

167:1

111:1

83:1

PET:Catalyst molar ratio


Fig. 4. Effect of PET:catalyst molar ratio on BHET yield for zinc acetate and sodium
carbonate (196  C, EG:PET molar ratio 7.6:1, 1 h).

Chemical shift, ppm


1

Fig. 6. H NMR spectra of BHET product (a-commercial BHET, b-BHET from glycolysis).

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

Signals at d 4.4 and 3.9 were characteristic of the methylene protons


of COOeCH2 and CH2OH, respectively. On the other hand, the triplet
at d 4.1 ppm was attributed to the protons of the hydroxyl group
[19]. In addition, the FTIR spectra (Fig. 7) of the puried monomer
clearly showed eOH band at 3450 cm1 and 1135 cm1, C]O
stretching at 1715 cm1, alkyl CeH at 2879 and 2954 cm1 and
aromatic CeH at 1411e1504 cm1, present in BHET [20,21]. Further,
the DSC prole also showed a reasonably sharp endothermic peak
at 110  C in agreement with the known melting point of BHET
(Fig. 8). From all these observations it was concluded that the
product from glycolysis was highly pure BHET.
3.2. Behaviour of alternative metal salts as catalysts
and glycolysis of complex PET wastes
The use of various salts as alternative catalysts to zinc acetate for
depolymerisation of PET into BHET by glycolysis was examined.
Four different sodium and potassium salts were selected, namely
sodium carbonate, sodium bicarbonate, sodium sulphate and
potassium sulphate. These compounds are not harmful to the
environment in relative low amounts. The following operation
conditions were employed for comparative purposes, 196  C, EG:
PET molar ratio 7.6:1, PET:catalyst molar ratio 380:1, after a reaction
time interval of 1 h. Activity results are summarised in Fig. 9 (runs 5,
and 13e16), and concluded that none of the salts investigated was
more active than zinc acetate. Hence, sodium carbonate led to
a slightly lower conversion (50%) compared with 64% given by zinc
acetate. In contrast, sodium bicarbonate was able to achieve only
40% yield while conversion values as low as 10 and 1% were
obtained with sodium sulphate and potassium sulphate, respectively. Interestingly, these results revealed a good potential for
sodium carbonate as an eco-friendly alternative since the environmental impact of sodium cation is less severe than that of toxic
zinc. In Fig. 4 the effect of Na2CO3 concentration on PET glycolysis is
examined under the following experimental conditions, 196  C, EG:
PET molar ratio 7.6:1, and 1 h (runs 17e21). It was observed that
although zinc acetate exhibited a higher intrinsic depolymerisation
activity, when a PET:Na2CO3 molar concentration of about 245:1

Endothermic

1026

40

60

80 100 120 140 160 180 200 220

Temperature, C
Fig. 8. DSC prole of BHET product (a-commercial BHET, b-BHET from glycolysis).

was employed, the attained conversion was 65%, close to that


corresponding to the equilibrium. In other words, the highest
achievable yield could be obtained with a PET:catalyst molar ratio
of 380:1 and 245:1 for zinc acetate and sodium carbonate,
respectively. This means that at the cost of increasing the molar
concentration of Na2CO3, the environmental impact of the glycolysis process can be markedly diminished since the use of Zn cation
is avoided. As for sodium and potassium sulphates their low activity
was attributed to its reduced solubility in EG. In contrast both
sodium carbonate and sodium bicarbonate were highly soluble in

100
90
80
70

Yield, %

Absorbance

60
50
40
30

20
10
0

500

1000 1500 2000 2500 3000 3500 4000

Wavenumber, cm-1
Fig. 7. IR spectra of BHET product (a-commercial BHET, b-BHET from glycolysis).

c) 2
(A
Zn

O3
CO 3
HC
a
Na 2
N

SO 4
Na 2

SO 4
K2

Fig. 9. BHET yield in the glycolysis of clear PET catalysed by several metal salts (196  C,
EG:PET molar ratio 7.6:1, PET:catalyst molar ratio 380:1) (solid bars). The conversion
value for a PET:NaHCO3 molar ratio of 190:1 is also included (lined bar).

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

this glycol. The higher activity of sodium carbonate with respect to


sodium bicarbonate was associated with the higher amount of
sodium ions present in the salt. In fact an additional experiment
(run 22) carried out with a PET:NaHCO3 molar ratio of 190:1
(196  C, EG:PET molar ratio 7.6:1, 1 h) resulted in a conversion value
of about 50%, comparable with that obtained with a PET:Na2CO3
molar ratio of 380:1 (run 14). This revealed that, when the same
amount of sodium cation was present in the reaction mixture, the
activity of these two salts was independent of the nature of the
anion (carbonate or bicarbonate).
It is believed that the glycolysis reaction proceeds by the
nucleophilic attack of the hydroxyl group of ethylene glycol upon
ester carbonyl groups present in the polymer. This carbonyl group is
rstly activated by the cation (Zn or Na). The reaction intermediate
is regarded as a complex formed by the coordination of ester
carbonyl groups to the metal species. The coordination lowers the
electron density of the carbonyl atom and facilitates the nucleophilic attack of the hydroxyl group upon this positively polarised
carbon atom, resulting in the cleavage of the polymer chain to give
rise to the monomer [22]. To be effective as a catalyst, the metal ion
must easily form an intermediate complex with the reacting
molecule. Thus it is believed that the strength of the metal-oxygen
bond formed can be a key factor in governing the activity of the
catalyst (metal salt). Since strength and length are inversely related
a short bond length would correspond to a strong metal-oxygen
interaction. In view of the values of this parameter for Zn (2.26 )
[23] and Na (2.41 ) [24] the higher activity of zinc acetate could be
explained as it forms stronger bonds with the carbonyl oxygen.
On the other hand, in an attempt to gain insight into the
macroscopic characterisation of the depolymerisation reaction, the
residues obtained from the rst ltration step after extraction with
boiling water were analysed by DSC. DSC proles corresponding to
the samples obtained in the runs carried out in the presence of
varying amounts of sodium carbonate (1 h, 196  C, EG:PET molar
ratio 7.6, runs 2, 17e21) are shown in Fig. 10. As a general rule, three
main DSC peaks were observed at 110  C, 150e155  C and 250  C,
which were attributed to trace amounts of BHET not properly
extracted, oligomers of different composition, and unconverted PET

1027

[25,26]. It was found that the peak associated with PET notably
decreased with an increase in the extent of the reaction (as a result
of the use of increasing amounts of catalyst), which was accompanied by a concomitant increase in the amount of the dimmer
formed. Hence, the use of sodium carbonate in the presence of
a large excess of ethylene glycol selectively depolymerised PET into
BHET and the dimer limiting the presence of higher oligomers.
When the equilibrium yield (about 70%) was reached it was
noticed that the shape of the DSC prole remained virtually unaltered as revealed by the notable similarity between the DSC proles
recorded at a time interval of 1 and 2 h. This suggested that probably an equilibrium would exist between the BHET product and the
remaining amount of the dimer [27,28], according to this scheme,
very fast

very fast

PETn ! oligomers ! dimer

!
BHET

(3)

On the other hand, a comparison among the conversion of the


various experiments, the shape of the DSC proles of the corresponding residues and a visual inspection of the evolution of the
reaction mixture with time led to the conclusion that the reaction
occurred in a single homogeneous phase above 50e60% BHET yield
whereas solid PET was dispersed into the liquid phase for a lower
yield. It is believed that when the sufcient amount of BHET and
soluble oligomers was formed, the reaction continued to proceed in
the single liquid phase until the equilibrium was reached [29].
As the last step of this study, the glycolysis reactions of clear,
coloured (green, amber, blue, white and silver), and multi-layered
recycled PET were preliminarily compared. In the literature few
reports can be found regarding the management of this type wastes
by glycolysis [30]. This task is of great relevance given the increased
implementation of these commercial plastics, typically associated
with brands and quality perception. Note that in contrast to clear
PET wastes, these fractions are not appropriate for an eventual
mechanical recycling because derived products present serious
limitations mainly due to colour, transparency or intrinsic viscosity,

100
90

100:1/ 2h

80
70

503:1
980:1

60

Y ield, %

Endothermic

100:1/ 1h

50
40
30

3300:1

20

No Catalyst

10

250

300

350

Temperature, C
Fig. 10. DSC proles of the solid residue of the glycolysis reaction with increasing
amounts of sodium carbonate (196  C, EG:PET molar ratio 7.6:1, 1e2 h).

Bl
ue
W
hi
te
M
ul
tila
ye
r

200

Si
lv
er

150

Am
be
r

100

Cl
ea
r
G
re
en

Fig. 11. BHET yield in the glycolysis of highly coloured and multi-layered PET wastes in
the presence of zinc acetate (196  C, EG:PET molar ratio 7.6:1, PET:catalyst molar ratio
100:1, 2 h).

1028

R. Lpez-Fonseca et al. / Polymer Degradation and Stability 95 (2010) 1022e1028

thereby restricting their applications and value. This means that up


to now the management of these materials can be only focused on
incineration or landll.
Since the scope of this study was to evaluate the afnity of
complex PET wastes for glycolysis and establish whether these
materials might give different glycolysed products, the most
favourable operation conditions were used in terms of temperature
(196  C), and large excess of both EG (EG.PET molar ratio 7.6:1) and
active zinc acetate (PET:catalyst molar ratio 100:1). Yield results
were recorded after 2 h (runs 23e28). It could be noticed in Fig. 11
that within experimental errors, these seven types of recycled PET
yielded practically similar glycolysed products with conversion
ranging between 59 and 66%. Expectedly the colouring compounds
did not play any chemical role in the process. However, they tended
to discolour the glycolysed products. As a result a further purication (extraction step is taken to isolate them) process of the
monomers is necessary to extract the pigments.
The BHET product obtained from the various PET wastes was
also characterised. Hence, the DSC, IR and 1H NMR features were
the same as those observed with the BHET obtained from the
glycolysis from clear PET, thereby evidencing its high purity.
4. Conclusions
The effect of various operation parameters on the catalysed
glycolysis of PET waste was investigated in order to efciently recover
its puried monomer, bis(2-hydroxyethyl terephthalate). The yield of
monomer increased with an increase in EG:PET molar ratio,
temperature, amount of catalyst and reaction time until the reaction
reached equilibrium conditions. Hence, a yield of BHET close to 70%
could be attained at 196  C, with an EG:PET molar ratio of 7.6:1 and
using zinc acetate as catalyst with a concentration (PET:catalyst molar
ratio) of 380:1. A detailed study on the properties of the recovered
product revealed its high purity. In an attempt to nd an alternative to
zinc acetate, a highly active catalyst for the process but with
a considerable environmental impact, several eco-friendly simple
salts were examined namely, sodium carbonate, sodium bicarbonate,
sodium sulphate and potassium sulphate. From the comparison of the
efciencies it was found that sodium carbonate and sodium bicarbonate could depolymerise PET wastes almost as efciently as zinc
acetate does at the cost of using a slightly higher concentration in
the reaction mixture. Under equilibrium conditions only BHET and
oligomers (dimer) were present in the reaction mixture.
On the other hand, it was demonstrated that glycolysis with
ethylene glycol appeared to be suitable for chemical recycling of
PET wastes of with a more complex nature and characteristics
(highly coloured and multi-layered PET) than those of clear PET,
used as a reference waste in this study. Very similar depolymerisation yields were obtained although it is necessary to address
improving the quality of the recovered BHET by removing impurities present in the wastes.
Acknowledgements
The nancial support for this work provided by the Spanish
Ministerio de Ciencia e Innovacin (CTQ2008-06868-C02-01 and
CTQ2008-06868-C02-02) and Diputacin Foral de Bizkaia (DIPE08/
13) are gratefully acknowledged. Also the supply of post-consumer
waste samples of PET by ECOEMBES is acknowledged.

References
[1] Al-Salem SM, Lettieri P, Baeyens J. Recycling and recovery routes of plastic
solid waste (PSW): a review. Waste Manag 2009;29(10):2625e43.
[2] PETCORE (PET containers recycling Europe), http://www.petcore.org
[accessed 26.10.09].
[3] Awaja F, Pavel D. Recycling of PET. Eur Polym J 2005;41(7):1453e77.
[4] Paszun D, Spychaj T. Chemical recycling of poly(ethylene terephthalate). Ind
Eng Chem Res 1987;36(4):1373e83.
[5] Lorenzetti C, Manaresi P, Berti C, Barbiroli G. Chemical recovery of useful
chemicals from polyester (PET) waste for resource conservation: a survey of
state of the art. J Polym Environ 2006;14(1):89e101.
[6] Karayannidis GP, Achilias DS. Chemical recycling of poly(ethylene terephthalate). Macromol Mater Eng 2007;292(2):128e46.
[7] Pang K, Kotek R, Tonelli A. Review of conventional and novel polymerization
processes for polyesters. Prog Polym Sci 2006;31(11):1009e37.
[8] Nadkarni VM. Recycling of polyesters. In: Fakirov S, editor. Handbook of
thermoplastic polyesters. Weinheim: Wiley-VCH Verlag GmbH; 2002. p. 1223.
[9] Nikles DE, Farahat MS. New motivation for the depolymerisation products
derived from poly(ethylene terephthalate) (PET) waste: a review. Macromol
Mater Eng 2005;290(1):13e30.
[10] Karayannidis GP, Nikolaidis AK, Sideridou ID, Bikiaris DN, Achilias DS.
Chemical recycling of PET by glycolysis: polymerization and characterization
of the dimethacrylated glycolysate. Macromol Mater Eng 2006;291(11):
1338e47.
[11] Shukla SR, Harad AM, Jawale LS. Recycling of waste PET into useful textile
auxiliaries. Waste Manag 2008;28(1):51e6.
[12] Zahedi AR, Razadeh M, Ghafarian SR. Unsaturated polyester resin via
chemical recycling of off-grade poly(ethylene terephthalate). Polym Int
2009;58(9):1084e91.
[13] Shukla SR, Harad AM, Jawale LS. Chemical recycling of PET waste into
hydrophobic textile dyestuffs. Polym Degrad Stab 2009;94(4):604e9.
[14] Campanelli JR, Kamal MR, Cooper DG. Kinetics of glycolysis of poly(ethylene
terephthalate) melts. J Appl Polym Sci 1994;54(11):1731e40.
[15] Chen J-W, Chen L-W, Cheng W-H. Kinetics of glycolysis of polyethylene
terephthalate with zinc catalyst. Polym Int 1999;48(9):885e8.
[16] Goje AS, Mishra S. Chemical kinetics, simulation, and thermodynamics of
glycolytic depolymerisation of poly(ethylene terephthalate) waste with
catalyst optimization for recycling of value added monomeric products.
Macromol Mater Eng 2003;288(4):326e36.
[17] Shukla S, Harad AM. Glycolysis of polyethylene terephthalate waste bers.
J Appl Polym Sci 2005;97(2):513e7.
[18] Xi G, Li M, Sun C. Study of depolymerization of waste polyethylene terephthalate into monomer of bis(2-hydroxyethyle terephthalate). Polym Degrad
Stab 2005;87(1):117e20.
[19] Ghaemy M, Mossaddegh K. Depolymerisation of poly(ethylene terephthalate)
bre wastes using ethylene glycol. Polym Degrad Stab 2005;90(3):570e6.
[20] Ikladious NE. Recycling of poly(terephthalate): identication of glycolysis
products. J Elastom Plast 2000;32(2):140e51.
[21] Chen C-H, Chen C-Y, Lo Y-W, Mao C-F, Liao W-T. Studies of glycolysis of poly
(ethylene terephthalate) recycled from postconsumer soft-drink bottles. I.
Inuences of glycolysis conditions. J Appl Polym Sci 2001;80(7):943e8.
[22] Santacesaria E, Trulli F, Minervini L, Di Serio M, Tesser R, Contessa S. Kinetic
and catalytic aspects in melt transesterication of dimethyl terephathalate
with ethylene glycol. J Appl Polym Sci 1994;54(9):1371e84.
[23] Wang CS, Sun YM. Studies on the formation of PEN. II. Polymerization of bis
(hydroxyethyl)naphthalate by various metallic catalysts. J Appl Polym Sci
1994;32(7):1305e15.
[24] Wood RM, Palenik GJ. Bond valance sums in coordination chemistry. Sodiumoxygen complexes. Inorg Chem 1999;38(17):3926e30.
[25] Guclu G, Kasgoz A, Ozbudak S, Ozgumus S, Orbay M. Glycolysis of poly
(ethylene terephthalate) wastes in xylene. J Appl Polym Sci 1998;69
(12):2311e9.
[26] Krehula LK, Hrnjak-Murgic Z, Jelencic J, Andricic B. Evaluation of poly
(ethylene-terephthalate) products of chemical recycling by differential scanning calorimetry. J Polym Environ 2009;17(1):20e7.
[27] Chen JY, Ou CF, Hu YC, Lin CC. Depolymerization of poly(ethylene terephthalate) resin under pressure. J Appl Polym Sci 1991;42(6):1501e7.
[28] Wang H, Liu Y, Li Z, Zhang X, Zhang S, Zhang Y. Glycolysis of poly
(ethylene terephthalate) catalyzed by ionic liquids. Eur Polym J 2009;45
(5):1535e44.
[29] Pardal F, Tersac G. Kinetics of poly(terephtalate) glycolysis by diethylene
glycol. I. Evolution of liquid and solid phases. Polym Degrad Stab 2006;91
(12):2840e7.
[30] Baliga S, Wong WT. Depolymerization of poly(ethylene terephthalate) recycled from post-consumer soft-drinks bottles. J Polym Sci A 1989;27
(6):2071e82.

Potrebbero piacerti anche