Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

THE CHEMICAL RECYCLING OF PET IN THE FRAMEWORK OF

SUSTAINABLE DEVELOPMENT

D. S. ACHILIAS∗ and G. P. KARAYANNIDIS


Laboratory of Organic Chemical Technology, Department of Chemistry, Aristotle University of
Thessaloniki, 541 24 Thessaloniki, Greece
(∗ author for correspondence, e-mail: axilias@chem.auth.gr; Fax: +30 231 997886)

Abstract. In this investigation, all the techniques used in the chemical recycling of polyethylene
terephthalate (PET) are critically reviewed according to the overall benefits together with the envi-
ronmental surcharge that they cause. Those, which are consistent with the principles of sustainable
development, are indicated. Experimental data are presented for the acid hydrolysis of PET and com-
pared with previous results on the alkaline hydrolysis of PET with, or without, the use of a phase
transfer catalyst. Overall material balances are carried out for the hydrolysis of PET. Finally, it can
be postulated that recycling according to the scheme:

is the only one within the framework of sustainable development. Therefore, the recycling of PET does
not only serve as a partial solution to the solid waste problem but also contributes to the conservation
of raw petrochemical products and energy.

Keywords: acid hydrolysis, alkaline hydrolysis, chemical recycling of PET, sustainable development,
terephthalic acid

1. Introduction

Polyethylene terephthalate (PET) is one of the versatile engineering plastics which


is used to manufacture mainly textiles and bottles for packaging. The overall world
consumption of PET amounts to about 13 million tons, of which 9.5 million tons are
processed for the textile industry, 2 million tons are consumed in the manufacture of
audio and video tapes and 1.5 million tons are used in the manufacture of various
types of packaging (mainly bottles) (Paszun and Spychaj, 1997). In Greece, the
main part of mineral water, a great part of olive and other seed-oils and all soft
drinks are offered now in PET bottles.
PET is produced by the polycondensation reaction of a di-acid and a di-alcohol.
The industrial practice since then was to use the dimethyl terephthalate (DMT) and
ethylene glycol (EG) as the two monomers for the production of PET. However,
nowadays the use of the esterification reaction between terephthalic acid (TPA) and
Water, Air, and Soil Pollution: Focus 4: 385–396, 2004.

C 2004 Kluwer Academic Publishers. Printed in the Netherlands.
386 D. S. ACHILIAS AND G. P. KARAYANNIDIS

EG is gaining importance. TPA and EG are used as the two main raw materials for
the industrial production of PET in the only Greek industry which produces PET
(Volos PET Industry).
PET recycling represents one of the most successful and widespread examples
of polymer recycling. PET bottle collection in Europe (European Union member
states plus Norway, Iceland, Switzerland, and all EU candidate countries) is growing
steadily. In 2001, 3 44 000 tones were collected, a 20% increase in comparison
with 2000. By 2006, it is predicted that European PET collection will increase to
700 000 tones (Petcore Report, 2002). The main driving force responsible for this
increased recycling of post-consumer PET is its widespread use, particularly in the
beverage industry. A very important feature of PET, decisive in the choice of its
wide application in the manufacture of packaging for the food industry is that it
does not have any side effects on the human beings. It should be pointed out, that
PET does not create a direct hazard to the environment but, due to its substantial
fraction by volume in the waste stream and its high resistance to the atmospheric
and biological agents, it is seen as a noxious material (Scheirs, 1998). Therefore,
the recycling of PET does not only serve as a partial solution to the solid waste
problem but also contributes to the conservation of raw petrochemical products and
energy. Products made from recycled plastics can result in 50–60% energy saving
as compared to making the same product from virgin resin.
The recycling of waste polymers including PET can be carried out in many ways.
Four main approaches have been proposed (Scheirs, 1998; Paszun and Spychaj,
1997; Papaspyrides and Poulakis, 1996):

1. Primary recycling refers to the ‘in-plant’ recycling of the scrap material of


controlled history. This process remains the most popular as it ensures simplicity
and low cost, dealing however only with the recycling of clean uncontaminated
single-type waste.
2. Mechanical recycling. In this approach, the polymer is separated from its associ-
ated contaminants and it can be readily reprocessed into granules by conventional
melt extrusion. Mechanical recycling includes the sorting and separation of the
wastes, size reduction and melt filtration. The main disadvantage of this type
of recycling is the deterioration of product properties in every cycle. This oc-
curs because the molecular weight of the recycled resin is reduced due to chain
scission reactions caused by the presence of water and trace acidic impurities.
Strategies for maintaining the polymer average molecular weight during repro-
cessing include intensive drying, reprocessing with degassing vacuum, the use
of chain extender compounds, etc. (Scheirs, 1998).
3. Chemical recycling has been defined as the process leading in total depolymer-
ization of PET to the monomers, or partial depolymerization to oligomers and
other chemical substances. Figure 1 summarizes the main methods used for the
chemical recycling of PET. These methods are extensively discussed in the next
session.
CHEMICAL RECYCLING OF PET 387

Figure 1. Chemical recycling processes of PET.

4. Energy recovery refers to the recovery of plastic’s energy content. Incineration


aiming at the recovery of energy is currently the most effective way to reduce
the volume of organic materials. Although polymers are actually high-yielding
energy sources, this method has been widely accused as ecologically unaccept-
able owing to the health risk from air born toxic substances e.g. dioxins (in the
case of chlorine containing polymers).

Among the above recycling techniques, the only one acceptable according to
the principles of Sustainable Development (development that meets the needs of
present generation without compromising the ability of future generations to meet
their needs) is the chemical recycling, because it leads to the formation of the
raw materials (monomers) from which the polymer is made of. In this way the
environment is not surcharged and there is no need for extra resources (monomers)
for the production of PET. The chemical recycling of PET is discussed in detail in
the next section.

2. Chemical Recycling of PET

Chemical recycling processes for PET are divided as follows: (i) Glycolysis, (ii)
methanolysis, (iii) hydrolysis and (iv) other processes (Figure 1). The main depoly-
merization processes that have reached commercial maturity till now are glycolysis
and methanolysis (Scheirs, 1998; Paszun and Spychaj, 1997).
Essentially, glycolysis involves the insertion of ethylene glycol (or diethylene
glycol and propylene glycol) in PET chains to give bis-(hydroxyethyl) terephthalate
(BHET) which is a substrate for PET synthesis and other oligomers. The kinetics
of PET glycolysis have been reviewed by Bisio and Xanthos (1994) and Brandrup
et al. (1996). The process is conducted in a wide range of temperatures from 180–
240 ◦ C, during a time period of 0.5–8 h. Zinc compounds were found to have
a catalytic effect on PET glycolysis below 245 ◦ C. The production of secondary
useful products, such as alkyl resins, obtained from the glycolysis of PET have
been proposed (Suh et al., 2000; Kawamura et al., 2002). The main advantage of
388 D. S. ACHILIAS AND G. P. KARAYANNIDIS

this method is that it can be easily integrated into a conventional PET production
plant and the recovered BHET can be blended in with fresh BHET. However, there
is always the risk of contaminants. On the other hand, the main disadvantage is
that the reaction products are not discrete chemicals but BHET along with higher
oligomers, which are difficult to purify with conventional techniques.
Methanolysis actually is the degradation of PET by methanol at high temper-
atures and high pressures with main products: dimethyl terephthalate (DMT) and
ethylene glycol (EG). Methanolysis of PET flakes is usually performed at tempera-
tures of 180–280 ◦ C and pressures 2–4 MPa (Scheirs, 1998). The resultant DMT is
purified by distillation, which removes all physical contaminants and can be reused
as a raw material for the production of PET. Methanolysis is currently being suc-
cessfully applied to PET scrap. Typical feedstocks for methanolysis include scrap
bottles, fiber waste, used films and plant waste (Paszun and Spychaj, 1997). The
main advantage of this method is that an installation of methanolysis can be located
in the polymer production line, because the DMT produced has a product quality
identical to virgin DMT. Ethylene glycol and methanol can also be easily recovered
and recycled. In this way, waste PET arising in the production cycle is utilized and
the monomers recovered can be reused in the manufacture of a full value polymer.
Disadvantages of the method include the high cost associated with the separation
and refining of the mixture of the reaction products (glycols, alcohols and phthalate
derivatives), if water does perturb the process it poisons the catalyst and forms
various azeotropes. However, the main disadvantage is associated with the trend of
all new PET production processes to use TPA instead of DMT as the raw material.
The conversion of the DMT produced by hydrolysis to TPA adds a considerable
cost in the methanolysis process.
Hydrolysis is the next method of PET waste chemical recycling. The growing
interest in this method is connected with the development of new factories for PET
synthesis directly from TPA and EG. Commercially, hydrolysis is not widely used
to produce food-grade recycled PET, till now, because of the cost associated with
purification of the recycled TPA. Hydrolysis of PET flakes can be carried out in an
acid, alkaline or neutral environment.
Acid hydrolysis is performed most frequently using concentrated sulfuric, nitric
or phosphoric acid. To avoid high pressures and temperatures in the reaction vessel,
a concentrated sulfuric acid (>14.5 M) has been proposed (Pusztaszeri, 1982).
However, the process proves very costly due to the need to recycle large amounts
of concentrated H2 SO4 and the purification of EG from the sulfuric acid. An acid
hydrolysis of waste PET powder in relatively dilute sulfuric acid (<10 M) and
the reuse of the sulfuric acid by recovery methods such as dialysis has also been
proposed (Yoshioka et al., 2001). However, this requires the prolongation of the
reaction time (5 h) and an increase in the reaction temperature (150 ◦ C). Yoshioka
et al. (1998) described a process for the depolymerization of PET powder from
waste bottles using nitric acid (7–13 M) at 70–100 ◦ C for 72 h. TPA and EG were
produced and the resulted EG was simultaneously oxidized to oxalic acid. The
CHEMICAL RECYCLING OF PET 389

proposed method had the advantage to result in value added products such as
oxalic acid, which is more expensive than TPA and EG.
Alkaline hydrolysis of PET is usually carried out with the use of an aqueous
alkaline solution of NaOH, or KOH of a concentration of 4–20 wt% (Paszun and
Spychaj, 1997; Scheirs, 1998). The reaction products are EG and the disodium
terephthalate salt TPA-Na2 . The mixture is heated up to 340 ◦ C to evaporate and
recover the EG by distillation. Pure TPA can be obtained by neutralization of the
reaction mixture with a strong mineral acid (e.g. H2 SO4 ). The process runs for
3–5 h at temperatures of 210–250 ◦ C, under pressure of 1.4–2 MPa. Apart of the
aqueous alkaline hydrolysis of PET, recently, alkali decomposition in non-aqueous
solutions has been reported (Oku et al., 1997). The addition of an ether (such
as dioxane, or tetrahydrofuran) as a mixed solvent with an alcohol (methanol, or
ethanol) accelerated the reaction. The time for complete reaction (>96%) of solid
PET with NaOH in methanol at 60 ◦ C was 40 min with dioxane as a co-solvent and
7 h without dioxane. The main advantage of this method is that it can tolerate highly
contaminated post-consumer PET such as magnetic recording tape, metallized PET
film, or photographic film (X-ray film) (Scheirs, 1998). The process is relatively
simple and less costly than methanolysis.
Neutral hydrolysis is carried out with the use of hot water or steam. The process
usually runs in high-pressure autoclaves at temperatures 200–300 ◦ C and pressure
1–4 MPa and with excess water (Scheirs, 1998; Paszun and Spychaj, 1997). The
products are high purity TPA and EG. The effect of catalysts such as zinc and sodium
salts on the PET hydrolysis was also investigated (Campanelli et al., 1994). Neutral
hydrolysis has the main advantage in comparison to acid and alkaline hydrolysis that
formation of substantial amounts of inorganic salts, which is difficult to dispose
of, is avoided. Therefore, it can be considered as more environmental friendly
and growing interest in this technology is expected. Its main drawback is that all
mechanical impurities present in PET are left in the TPA, thus the product can be
considered of worse purity than the product of acid or alkaline hydrolysis.
Other processes include some newly developed techniques for the depolymer-
ization of PET waste such as aminolysis and ammonolysis (Spychaj and Paszun,
1998).
Aminolysis is the reaction of PET with different amine aqueous solutions to yield
the corresponding diamides of TPA and EG. There are no known reports for the
utilization of this process in a commercial scale. However, partial aminolysis has
found its application in the improvement of PET properties in the manufacture of
fibers with defined processing properties. The amines used include methylamine,
ethylamine and ethanolamine in the temperature range of 20–100 ◦ C. Degrada-
tion with triethanolamines gives different ester products that may constitute poten-
tially raw materials for the synthesis of polyurethanes, mainly rigid polyurethane
foams.
Ammonolysis is the reaction of anhydrous ammonia with PET to produce a
terephthaldiamide. This can be converted to terephthalonitrile and further to other
390 D. S. ACHILIAS AND G. P. KARAYANNIDIS

chemical substances. The reaction was carried out with post-consumer PET bottles
at 120–180 ◦ C and pressure of about 2 MPa for 1–7 h.
From all the aforementioned techniques we focused our attention on a specific
method which should conform to the sustainable development principles and the
new trends for the industrial production of PET. In other words chemical recycling
by hydrolysis leading to their monomers TPA and EG. In this paper the acid hy-
drolysis is examined and results are compared with catalyzed and non-catalyzed
alkaline hydrolysis studied by our group in previous publications (Karayannidis et
al., 2002; Kosmidis et al., 2001).

3. Materials and Methods

3.1. M ATERIALS

PET flakes free from other plastics, were prepared from used clear PET bottles,
from which the crystallized caps and bottom parts, together with the label and the
glue had been removed. The bottles were cut and fed to a rotary cutter with a max
size of 6 mm. The chemicals used were reagent grade.

3.2. D EPOLYMERIZATION OF PET IN A CID SOLUTION

The depolymerization reaction was carried out in a 0.5 L round-bottom flask


equipped with a reflux condenser and a magnetic agitator. The required amount
(25 mL) of the sulfuric acid solution (70–83 wt.%), together with the PET flakes
(5 g) were added into the reactor and heated to the desired reaction temperature
(30–90 ◦ C). The agitation started in order to keep the mixture homogeneous and
the reflux condenser set. The reaction time started and the mixture was allowed to
react for 3 to 5 h. Afterwards, the mixture was filtered to separate the TPA produced
and the unreacted PET. A solution of KOH was added to the solid product. In this
way the TPA reacted to form the dipotassium salt TPAK2 , whereas PET remained
unreacted. Finally, the mixture was filtered again, dried in an oven until constantly
weighted in order to calculate the percent-unreacted PET.

3.3. D EPOLYMERIZATION OF PET IN A LKALINE S OLUTION

In the alkaline hydrolysis of PET two different approaches were followed, namely
non-catalyzed reaction at high temperatures (Karayannidis et al., 2002) and cat-
alyzed, using trioctyl methyl ammonium bromide (TOMAB) at temperatures less
than 100 ◦ C (Kosmidis et al., 2001). The depolymerization reaction was carried
out in a three-neck, 2 L round-bottom reactor equipped with a reflux condenser, a
mechanical agitator and an electric heating mantle. 1.5 L of the sodium hydroxide
CHEMICAL RECYCLING OF PET 391

solution (5–15 wt.%), was added into the reactor and heated to the desired reaction
temperature (80–200 ◦ C for the non-catalyzed reaction and 70–95 ◦ C for the cat-
alyzed). The agitation started in order to keep the mixture homogeneous and the
reflux condenser set. The desired quantity of PET flakes (11.52 g) and the catalyst
(if used) were then added. The reaction time started and the mixture was allowed
to react for 5 to 6 h. The sample was neutralized to pH 7 with H2 SO4 and filtered
through glass filter (G3). The TPA in the mixture was precipitated by the addition
of H2 SO4 until a pH 2.5–3. It was then removed by filtration with a G3 glass filter
and washed with water. The final solid TPA produced was dried in a vacuum oven
at 120 ◦ C and weighed. The final unreacted PET was also measured on filtration
of the final mixture through a G3 glass filter. The solid PET remained was washed
with water, dried in a vacuum oven at 120 ◦ C and weighed.

3.4. A NALYSIS OF THE R ESULTS

The % yield in TPA was calculated using the formula:

NTPA
TPA Yield (%) = 100 (1)
NTPA,0

where, NTPA and NTPA,0 refer to the number of moles weighted and the theoretical
number of TPA moles that will be produced on complete decomposition of PET,
respectively.
The percent degradation of PET was calculated using the following equation:

WPET,0 − WPET, f
PET Degradation (%) = 100 (2)
WPET,0

where, WPET,0 and WPET, f refer to the initial and final weight of PET, respectively.
The purity of the TPA produced was examined using three different methods:

1. 1 H NMR analysis. The TPA separated was dissolved in deuterated dimethyl


sulfoxide and its purity examined by 1 H NMR analysis.
2. Titration with KOH. About 1 g of TPA is weighed into a 250 mL conical flask. To
dissolve the sample 25 mL of analytical grade pyridine is added by pipette and
the suspension is heated with a reflux condenser until a clear solution is obtained.
The condenser is then washed out by the addition of about 5 mL of pure pyridine
through the top and the content of the flask is titrated with approximately 0.5 N
standard potassium hydroxide solution to a phenolphthalein endpoint.
3. Repolymerization of TPA to PET. Terephthalic acid separated from recycled PET
was reacted with ethylene glycol to produce a polyester sample with a white
color according to the experimental details provided in Kosmidis et al., 2001.
392 D. S. ACHILIAS AND G. P. KARAYANNIDIS

4. Results and Discussion

4.1. ACID H YDROLYSIS OF PET

The depolymerization reaction of PET with water in an acid (H2 SO4 ) environment
is happens according to the following reaction:

The effect of the sulfuric acid concentration on the degree of PET degradation
is shown in Figure 2. As it can be seen, in 90 ◦ C almost complete decomposition of
PET to its monomers is achieved only with a very concentrated acid solution (80
wt%). If the acid concentration is less than 76 wt% the depolymerization reaction is
very slow. This means that if one wants to carry out the reaction at temperatures less
than 100 ◦ C acid-corrosion resistant equipments are required. This great increase
in PET degradation in a relatively small sulfuric acid concentration interval means

Figure 2. Effect of H2 SO4 concentration on the degree of PET degradation, in acidic hydrolysis at
90 ◦ C for 3 h.
CHEMICAL RECYCLING OF PET 393

Figure 3. Effect of temperature on the degree of PET degradation, in acidic hydrolysis at 80 wt%
H2 SO4 concentration for 3 h.

that a type of dissolution of PET in H2 SO4 occurs before the hydrolysis takes place.
Thus, if the appropriate acid concentration for the dissolution at the prespesified
temperature is met, then the reaction occurs much more easily.
The effect of depolymerization temperature on the amount of unreacted PET is
shown in Figure 3 for an 80 wt% H2 SO4 solution. It can be seen that the amount of
unreacted PET begins to decrease at a temperature of 60 ◦ C and much more at 70 ◦ C
(27 wt%). Whereas, if the reaction is carried out at temperatures less than 50 ◦ C,
PET degradation does not occur in a great extent. Hence, at this acid concentration,
a temperature of 70 ◦ C would be adequate for the degradation of PET.

4.2. PET H YDROLYSIS IN A LKALINE SOLUTION

The effect of reaction temperature on the alkaline hydrolysis of PET at two de-
polymerization time periods is shown in Figure 4. As expected, an increase in the
reaction temperature leads to increased decomposition of PET. On comparing the
values obtained between the catalyzed and the non-catalyzed reaction it can be
postulated that the temperature required for the complete degradation of PET when
no catalyst is added is much higher (200 ◦ C) than the corresponding value when
the catalyst is used (90 ◦ C). The use of the phase transfer catalyst (trioctyl methyl
ammonium bromide) for the degradation of fibers made of Nylon 66 and PET at
short time periods and temperatures less than 100 ◦ C had been observed by Polk
et al. (1999).
Finally, results on the effect of the amount of catalyst on the PET degradation
are shown in Figure 5. As it can be observed, after 5 h of reaction, that the TPA yield
without the use of catalyst is approximately 7%, compared to 90% in the case of
0.01 mol catalyst/mol PET. Another interesting point is that there exists a limiting
394 D. S. ACHILIAS AND G. P. KARAYANNIDIS

Figure 4. Effect of temperature on the amount of TPA produced during the alkaline hydrolysis of
PET with and without the use of a phase transfer catalyst at two time periods 3 h and 5 h.

Figure 5. Effect of the amount of the phase transfer catalyst (mol cat/mol PET) on the terephthalic
acid produced during the alkaline hydrolysis of PET at 80 ◦ C and 1 h, 3 h and 5 h.

mole fraction of approximately 0.003 mol catalyst/mol PET, beyond which any
further increase in the amount of catalyst does not influence the degradation rate.
The purity of the TPA on the basis of carboxyl content was found to be 99.6%.
The 1 H NMR analysis revealed an admixture of about 2% isophthalic acid used in
the production of PET bottles.
Finally, because this technique (i.e. catalytic alkaline hydrolysis) seems to be
very promising, total material balances are carried out next. All the amounts of
chemical reagents used per 1 g of PET recycled are presented in Table I. The TPA
CHEMICAL RECYCLING OF PET 395

TABLE I
Material balances for the catalyzed alkaline hydrolysis of PET

Reactants Products (monomers)


Byproduct
Material NaOH (pellets) H2 SO4 TPA EG Na2 SO4

Quantity in 0.417 0.51 0.864 (100% conversion) 0.323 0.74


gr/1 gr PET 0.778 (90% conversion)

(monomer) produced is of high purity and the secondary materials used (H2 SO4 and
NaOH) are cheap and not in very large amount. Furthermore, a valuable byproduct
(Na2 SO4 ) is produced that could be fed in a corresponding industry. The amount of
energy required (heating and agitation) is rather low (low reaction temperatures).
Hence, this method could be adequate for the chemical recycling of PET. However,
in order to reach the industrial maturity, a techno-economical analysis should be
performed taking into consideration as well as the cost associated with the collecting
of waste PET.

5. Conclusions

In this investigation, the chemical recycling of PET was examined in the context of
Sustainable Development. Results were presented for the acid hydrolysis of PET
and compared to our previous results on the alkaline hydrolysis of PET with or
without the presence of a phase transfer catalyst. The main objective was to obtain
clear monomers (TPA and EG) from which PET could be reproduced. In this way,
the environment is neither surcharged nor there is not a need for extra resources for
the production of PET. PET can be re-produced directly from its wastes. From all the
analyzed techniques the alkaline hydrolysis using special phase transfer catalysts
seems to be the most promising in order to find future industrial applications.

Aknowledgements

We would like to thank the postgraduate students V. Kosmidis and A. Chatziav-


goustis and the graduate student P. Megalokonomos for carrying out part of the
work presented here.

References

Bisio, A. and Xanthos, M.: 1994, How to Manage Plastics Waste, Hanser Gardner, Cincinatti, OH.
Brandrup, J., Bittner, M., Menges, G. and Michaeli, W.: 1996, Recycling and Recovery of Plastics,
Hanser Gardner, Cincinatti, OH.
396 D. S. ACHILIAS AND G. P. KARAYANNIDIS

Campanelli, J. R., Cooper, D. G. and Kamal, M. R.: 1994, ‘Catalysed hydrolysis of PET melts’,
J. Appl. Polym. Sci. 53, 985–993.
Karayannidis, G., Chatziavgoustis, A. and Achilias, D. S.: 2002, ‘PET recycling and recovery of pure
terephthalic acid by alkaline hydrolysis’, Adv. Polym. Technol. 21, 250–259.
Kawamura, C., Ito, K., Nishida, R., Yoshihara, I. and Numa, N.: 2002 ‘Coating resins synthesized
from recycled PET’, Prog. Org. Coat. 45, 185–181.
Kosmidis, V., Achilias, D. S. and Karayannidis, G.: 2001, ‘Poly(ethylene terephthalate) recycling and
recovery of pure terephthalic acid. Kinetics of a phase transfer catalyzed alkaline hydrolysis’,
Macromol. Mater. Eng. 286, 640–647.
Oku, A., Hu, L. and Yamada, E.: 1997, ‘Alkali decomposition of PET with sodium hydroxide in
non-aqueous ethylene glycol: a study on recycling of terephthalic acid and ethylene glycol’,
J. Appl. Polym. Sci. 63, 589–601.
Papaspyrides, C. and Poulakis, J.: 1996, in J. C. Salamone (ed.), Encyclopedia of Polymeric Materials,
CRC Press, Boca Raton, FL, USA, pp. 7403–7418.
Paszun, D. and Spychaj, T.: 1997, ‘Chemical recycling of poly(ethylene terephthalate)’, Ind. Eng.
Chem. Res. 36, 1373–1383.
PETCORE Report: 2002, 4(4) (Available in www.Petcore.com).
Polk, M., Leboeuf, L., Shah, M., Won, C.-Y., Hu, X. and Ding, W.: 1999, ‘Nylon 66, Nylon 46 and
PET phase-transfer-catalysed alkaline depolymerization at atmospheric pressure’, Polym. Plast.
Technol. Eng. 38(3), 459–470.
Pusztaszeri, S. F.: 1982, ‘Method for Recovery of Terephthalic Acid from Polyester Scrap’, U.S. Patent
4 355 175.
Scheirs, J.: 1998, Polymer Recycling, Wiley, W. Sussex.
Spychaj, T. and Paszun, D.: 1998, ‘New trends in chemical recycling of PET’, Macromol. Symp. 135,
137–145.
Suh, D. J., Park, O. O. and Yoon, K. H.: 2000, ‘The properties of unsaturated polyester based on the
glycolyzed PET with various glycol compositions’, Polymer 41, 461–466.
Yoshioka, T., Motoki, T. and Okuwaki, A.: 2001, ‘Kinetics of hydrolysis of PET powder in sulfuric
acid by a modified shrinking core model’, Ind. Eng. Chem. Res. 40, 75–79.
Yoshioka, T., Okayama, N. and Okuwaki, A.: 1998, ‘Kinetics of hydrolysis of PET powder in nitric
acid by a modified shrinking core model’, Ind. Eng. Chem. Res. 37, 336–340.

You might also like