You are on page 1of 6

D. Spaseska, M.

Civkaroska Journal of the University of Chemical Technology and Metallurgy, 45, 4, 2010, 379-384

ALKALINE HYDROLYSIS OF POLY(ETHYLENE TEREPHTHALATE) RECYCLED FROM THE POSTCONSUMER SOFT-DRINK BOTTLES
D. Spaseska, M. Civkaroska
Faculty of Technology and Metallurgy SS Cyril and Methodius University of Skopje Republic of Macedonia Received 30 June 2010 Accepted 28 September 2010

ABSTRACT The post consumer Poly (ethylene terephthalate)PET) recycling science was initiated by the fact that this polymer was non-degradable in nature, since its molecules were too large to decompose. There has been a growing need, however, for a chemical recycling process as one of the most successful method for post-consumer PET transformation into monomers. The solution of this problem seems to be in the creation of remunerative processes, for which post-consumer PET is used as a source material. This paper presents a review of works that cover PET waste recycling with alkaline hydrolysis of post-consumer bottles for their depolymerization to monomers (terephthalic acid). The process is carried out with sodium hydroxide and trioctyl methyl ammonium bromide (TOMAB) as catalyst within a relatively short time and temperature. A mathematical model of PET recycling has been derived according to the definition of a 23 factorial experimental design. The influence of the main factors on the alkaline hydrolysis conversion of the PET, as expressed by the yield of terepthalic acid (TPA), in ascending order, includes the process temperature and the amount of catalyst, but the hold-up time is negligible. Keywords: PET recycling, hydrolysis conversion, terephthalic acid, factorial experimental design, mathematical model of the process, hydrolysis time, amount of catalyst, hydrolysis temperature.

INTRODUCTION Plastics make up a high proportion of waste the volume and range of which increases dramatically. Although plastics make up between 5 mass % and 15% mass of municipal solid waste it comprises 20-30 % of the volume [1]. Most plastics are non-degradable and take a long time to decompose, possibly up to hundreds of years although no one knows for certain as plastics have not existed for long enough, since they started to be land filled. PET is one of the versatile engineering plastics that are widely used in manufacturing of high strength fibers, audio and video tapes and various types of packaging, mainly soft drink bottles and jars [2-4]. The majority of the world PET production is for synthetic fibers (in excess of 60 %) with bottle production accounting for around 30 % of global demand. One of the main reasons for the widespread use of PET is the pos-

sibility for producing a number of different grades over a broad range of molecular masses in a single multiproduct polymerization plant [5]. 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 [2]. 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. The huge amounts of PET products cause serious environmental pollution, because commonly PET content reaches about 12 % in municipal plastic waste showing a slow rate of natural decomposition [1,6]. PET

379

Journal of the University of Chemical Technology and Metallurgy, 45, 4, 2010

recycling is very important for at least two main reasons: firstly, to reduce the increasing volumes of plastic waste and secondly - to generate value-added materials from low cost sources by converting them into valuable materials. The demand to expand grade uses of recycled PET led to the research into alternative processing methods in order to produce higher value products. Numerous methods for recycling of disposable beverage bottles and other containers made of PET or blends of PET with other materials have been widely reported [7-12]. One of the most applicable methods for PET recycling is chemical recycling defined as the process leading to total depolymerization of PET to the monomers, or partial depolymerization to oligomers and other chemical substances. This means that the process results in the formation of the raw materials (monomers) from which the polymer is made of, and there is no need for extra resources (monomers) for PET production. Glycolysis, methanolysis, hydrolysis are the main processes included in the chemical recycling processes for PET [2, 3]. Hydrolysis is a method of PET waste chemical recycling which attracts a growing interest related to the development of factories for PET synthesis directly from TPA and EG. Commercially, hydrolysis is not widely used to produce food-grade recycled PET, because of the cost associated with purification of the recycled TPA. Hydrolysis as the method of PET waste recycling by the reaction of PET with water in an acid, alkaline or neutral environment, leads to total depolymerization to its monomers - terephthalic acid and ethylene glycol. The reaction time ranges from a few to 30 minutes at high temperature and under high pressure, and requires no additives, such as catalysts or neutralizers [13-20]. 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 mass % [2, 14]. The reaction products are EG and the disodium terephthalate salt TPA-Na2. Pure TPA can be obtained by neutralization of the reaction mixture with a strong mineral acid (e.g. H2SO4). Apart from the aqueous alkaline hydrolysis of PET, alkali decomposition in non-aqueous solutions has been reported [21]. The addition of an ether (such as dioxane, or tetrahydrofuran) as a mixed solvent with an alcohol (methanol, or ethanol) accelerated the reaction. The main advantage of this method is that it can tolerate highly contaminated post-consumer

PET such as magnetic recording tape, metalized PET film, or photographic film (X-ray film) [2] . The process is relatively simple and less costly than methanolysis. Kosmidis et al. [22] have made a comparison between the acid hydrolysis with catalyzed and non-catalyzed alkaline hydrolysis. They point out that alkaline hydrolysis using special phase transfer catalysts seems to be most promising for future industrial applications. The process, named Solid State Shear Pulverization [23, 24] for production of PET powders requires lower experimental temperature and pressure in comparison with using PET flakes. Solid-state shear pulverization is a continuous, one-step process that has been initially conceived as a means to convert pelletized or flaked polymer feedstock into powder. Under certain process conditions, polymer chain scission accompanies powder production, as evidenced by reduction in molecular mass [25] and the production of free radicals [26]. The yield of terephthalic acid was as high as 98 % by pulverization and depolymerization of PET [27]. The alkaline hydrolysis of PET with [22], or without [28] the use of a phase transfer catalyst, compared with the acid hydrolysis, has been studied by D. S. Achilias and G. P. Karayannidis [13]. Overall material balances were carried out for the hydrolysis of PET. Also, it was postulated that recycling according to the scheme:

is the only one within the framework of sustainable development. In the works of G. P. Karayannidis and D. S. Achilias [28, 29] as well as the works of J.Das et al. [30], hydrolysis in alkaline environment was employed in order to recover pure terephthalic acid monomer that could be repolymerized to form the polymer again. Thus, 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. A phase-transfer catalyst was introduced in the alkaline hydrolysis, so that the reaction takes place at atmospheric pressure and in mild experimental conditions. R. Lpez-Fonseca et al. [31] were focused on the identification of the catalytic behavior, if any, of a series of quaternary phosphonium and ammonium salts as phase transfer catalysts for the alkaline hydrolysis of PET, and on the determination of the kinetics of the

380

D. Spaseska, M. Civkaroska

phase transfer catalyzed process. The use of selected phosphonium quaternary salts exhibited a remarkably positive effect on the experimental conditions under which the depolymerisation of poly(ethylene terephthalate) by alkaline hydrolysis can be carried out, especially in terms of low operating temperature. The proposed kinetic model accounted for the unanalyzed and catalyzed reactions and predicted for the reaction rate a linear correlation with the concentration of the quaternary salt. The notable increase in the phase transfer catalyzed reaction rate was related mainly to the greater value of the pre-exponential factor while the value of the activation energy was hardly modified by the presence of the quaternary phosphonium salt, thereby suggesting a similar mechanism for the alkaline hydrolysis with or without phase transfer catalyst. The present study was undertaken with the aim of PET depolymerization at milder conditions than those used generally. Mathematical design of the process is a significant domain of study in the investigated process, because it points out the correlation between the main process parameters and the ultimate parameters (yield or characteristics of the products) [32-34]. In this work a mathematical model for the catalytic alkaline hydrolysis has been derived. The effects of various operating parameters, temperature, hold-up time and the amount of catalyst on the process, have been investigated. The three-factor interaction effect is significantly related to the hydrolysis conversion of the recycled PET bottles. EXPERIMENTAL

tor and an electric heating mantle. 1.5 L of the sodium hydroxide solution (5-15 mass %), were added into the reactor and heated to the desired reaction temperature, 80 and 120C. Agitation was 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 (trioctyl methyl ammonium bromide, TOMAB) were then added. The reaction time started and the mixture was allowed to react for 3 and 5 h. At the end of the reaction, the reaction mass was neutralized to pH 7 with H2SO4 and filtered through a glass filter (G3). The TPA in the mixture was precipitated by the addition of H2SO4 down to 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 120C and weighed. The final unreacted PET was also measured by filtration of the final mixture through a G3 glass filter. The solid PET that remained was washed with water, dried in a vacuum oven at 120C and weighed. The main investigated parameters were as followed: reaction temperature (x1) 80 and 120C catalyst concentration (x2): 0,1 and 2,0 mol/mol PET reaction time (x3): 3 and 5 h. The % yield of TPA was used as a reflection function (y).The percent degradation of PET was calculated using the following equation: PET degradation (%) = (W PET,0 W PET,f) 100 / W PET,0 where, W PET,0 and W PET, f refer to the initial and final mass of PET, respectively. RESULTS AND DISCUSSION

Materials The postconsumer soft-drink PET bottles free from caps and label were used as a starting material for depolymerization. PET bottles were consequently cut to the size of 6 mm. Depolymerization of PET in alkaline solution The alkaline depolymerization was carried out in accordance with the method described in the work of Kosmidis et al. [22]. In order to derive the mathematical relationship for the depolymerization process, the experiments were carried out according to an experimental design [32]. The depolymerization reaction was carried out in a three neck, 2 L round-bottom reactor equipped with a reflux condenser, a mechanical agita-

PET was first hydrolyzed in sodium hydroxide to yield the disodium salt, Na2TPA according to the chemical reaction shown in Scheme 1. The salt Na2TPA was then acidified with 98 % sulphuric acid to precipitate out TPA. Initially the reaction mixture consists of the solid organic phase of PET and the aqueous alkaline solution (NaOH). The phase transfer catalyst treoctyl methyl ammonium bromide (TOMAB) fulfills the requirement of having enough organic character in order to avoid steric hindrance. The cationic part of the catalyst does not carry the hydroxide anion into the organic phase (extraction mechanism) but on its surface (interfacial

381

Journal of the University of Chemical Technology and Metallurgy, 45, 4, 2010

Scheme 1. Chemical reactions involved in the process. mechanism). In this way the PET macromolecules on the surface of the particles can be easily attacked by the OH- and then depolymerized. The terephthalate anion produced returns to the aqueous phase and forms the disodium terephthalate salt with Na+. The reaction proceeds until complete depolymerization of PET to Na2TPA, while the catalyst remains in the aqueous phase [30]. For the purpose of determination of the dependence of yield of TPA on process parameters, like temperature, catalyst concentration and hold-up time, the experiments have been carried out in accordance with an experimental design, resulting in a derived mathematical model for the degradation process [32]. The derivation of the mathematical model of the process is enabled by a minimal number of the performed experiments. As the reflection function, the yield of TPA has been taken. The obtained results Table 1. Yield of TPA in dependence on process parameters.
No. X1 t/ C
0

are presented in Table 1. From the experimentally obtained results, the derived equation with coded variables for the produced TPA is given with the relationship: Y = 55,75 + 3,5 X1 + 37,25 X2 + 3,5 X1X2 + 4 X2X3 (1) 5,75 X1X2X3 The equation with natural variables is given by the following relationship: y = 161,2105 1,2895 x1 11710,5263 x2 36,1974 x3 + 139,4737x 1 x 2 + 0,3178 x 1x3 + 3447,3684 x 2x3 30,2632 x1x2x3 (2) The dependence of the obtained TPA concentration on temperature and catalyst concentration, according to equation (2) are presented in Fig.1 and Fig.2, respectively. According to the relationship (1) and Fig.1

X2 Catalyst conc. mol(x0,01)/mol PET 0,1 0,1 2,0 2,0 0,1 0,1 2,0 2,0 h 3 3 3 3 5 5 5 5

X3 Time /

Y TPA / % w.

1. 2. 3. 4. 5. 6. 7. 8.

80 120 80 120 80 120 80 120

32 15 80 100 5 22 92 100

Fig. 1. Effect of temperature on the amount of TPA produced during alkaline hydrolysis of PET with the amount of catalyst and time fixed at the varied levels.

382

D. Spaseska, M. Civkaroska

REFERENCES 1. P. T. Williams, Waste Treatment and Disposal. 2nd ed. John Wiley & Sons, Weinheim, Germany, 2006. 2. J. Scheirs, Chapter 4: Recycling of PET, Polymer Recycling, Wiley Series in Polymer Science, John Wiley & Sons, Sussex, 1998. 3. D. Paszun, T. Spychaj, Chemical recycling of poly(ethylene terephthalate), Ind. Eng. Chem. Res., 36, 1997, 1373-1383. 4. Y. Sueoka et al., Chemical recycling techniques for plastics, Mitsubishi Juko Giho, 36, 3, 1999, 146149. 5. V.M. Nadkami, Recycling of Polyesters. In: Handbook of Thermoplastic Polyesters. Ed. by S. Fakirov, v. 2, Weinheim, Germany: Wiley-VCH, 2002, pp. 1223-1249 . 6. M. Edge, M. Hayes, M. Mohammadian, N.S. Allen, T.S. Jewitt, K. Brems, K. Jones, Aspects of Poly (Ethylene Terephthalate) Degradation for Archival Life and Environmental Degradation, Polymer Degradation and Stability, 32, 2, 1991, 131-153. 7. D.E. Nikles, M.S. Farahat, New Motivation for the Depolymerization Products Derived from Poly (Ethylene Terephthalate) (PET) Waste: a Review, Macromolecular Materials and Engineering, 290, 2005, 13-30. 8. A. Firas, P. Dumitru, Recycling of PET, European Polymer Journal, 41, 7, 2005, 1453-1698. 9. C. Miller, Polyethylene Terephthalate Waste Age, 33, 5, 2002, 102-106. 10. A. Pawlak, et al., Characterization of Scrap Poly(ethylene Terephthalate), European Polymer Journal, 36, 9, 2000, 1875-1884. 11. A. Oromiehie, A. Mamizadeh, Recycling PET Beverage Bottles and Improving Properties, Polymer International, 53, 2004, 728-732. 12. D.M. Fann, S.K. Huang, J.Y. Lee, Kinetics and Thermal Crystallinity of Recycled Pet. I. Dynamic Cooling Crystallization Studies on Blends Recycled with Engineering PET, Journal of Applied Polymer Science, 63, 8, 1996, 1375-1385. 13. D.S. Achilias, G.P. Karayannidis, The Chemical Recycling of PET in the framework of Sustainable Development, Water, Air, and Soil Pollution: Focus 4: 2004, 385-396. 14. P.M. Subramian, Plastics Recycling and Waste

Fig. 2. Effect of the amount of the phase transfer catalyst (mol cat/mol PET) on the terephthalic acid produced during alkaline hydrolysis of PET.

and Fig. 2 both process parameters, temperature and catalyst concentration, have a positive effect to TPA yield, but the hold-up time, in the investigated interval, has a neglecting effect. Taking into account the coefficients before the individual parameters (equation 1), the most expressed influence on the reflection function (Y) has the parameter catalyst concentration. CONCLUSIONS PET depolymerization in alkaline medium and catalyst presence results in TPA production. Almost complete conversion of PET in relatively mild process conditions - low temperature and considerably low concentration of alkali and catalyst, could be achieved. The alkaline hydrolysis with phase transfer catalysts is one of the most promising methods for PET chemical recycling. The yield of TPA is directly dependent on temperature and catalyst concentration, but the hold-up time as a process parameter in the investigated interval is negligible. The catalyst concentration is the most expressed parameter for the alkali PET Depolymerization, followed by the reaction temperature. The experimental design was successfully used for the determination of the best conditions for carrying out the PET alkaline depolymerization process as well as for the optimal yield of TPA.

383

Journal of the University of Chemical Technology and Metallurgy, 45, 4, 2010

Management in the US, Resources, Conservation and Recycling, 28, 2000, 253-263. 15. G. Guclu, T. Yalcinyuva, S. Ozgumus, M. Orbay, Hydrolysis of Waste Polyethylene Terephthalate and Characterization of Products by Differential Scanning Calorimetry, Thermochimica Acta, 404, 1, 2003, 193-205. 16. G. Grause, W. Kaminsky, G. Fahrbach, Hydrolysis of Poly(Ethylene Terephthalate) in a Fluidised Bed Reactor, Polymer Degradation and Stability, 85, 1, 2004, 571-575. 17. M. Rosmaninho, E. Jardim, M.C.C. Flavia, G. Ferreira, V. Thom, M. Yoshida, M.H. Araujo, R.M. Lago, Surface Hydrolysis of Postconsumer Polyethylene Terephthalate to Produce Adsorbents for Cationic Contaminants, Journal of Applied Polymer Science, 102, 6, 2006, 5284- 5291. 18. US Patent 6649792: Method of Chemical Recycling of Polyethylene Terephthalate Waste, 2003. 19. Patent WO/2001/068581: The Method of Chemical Recycling of Polyethylene Terephthalate Waste, 2001. 20. D. Carta, G. Cao, C.D Angeli, Chemical Recycling of Poly(Ethylene Terephthalate) (PET) by Hydrolysis and Glycolysis, Environmental Science and Pollution Research, 10, 6, 2003, 390-394. 21. A. Oku, L. Hu, E. Yamada, 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, 1997, 589-601. 22. V. Kosmidis, D.S. Achilias, G. Karayannidis, Poly(ethylene terephthalate) recycling and recovery of pure terephthalic acid. Kinetics of a phase transfer catalyzed alkaline hydrolysis, Macromol. Mater. Eng. 286, 2001, 640647. 23. K. Khait, J.M. Torkelson, Solid-state shear pulverization of plastics: a green recycling process, Polym.Plast. Technol. Eng., 38, 3, 1999, 445-457. 24. N. Furgiuele, A.H. Lebovitz, K. Khait, J.M.

Torkelson, Polym. Eng. Sci., 40, 6, 2000, 1447-1468. 25. N. Furgiuele, A.H. Lebovitz, K. Khait, J.M. Torkelson, Macromolecules, 33, 2000, 225-240. 26. D. Ahn, K. Khait, M.A. Petrich, Microstructure Changes in Homopolymers and Polymer Blends by Elastic Strain ulverization, J. Appl. Polym. Sci., 55, 1995, 1431-1440 27. Behzad Shirkavand Hadavand , Hossein Hosseini,Chemical Recycling of PET with Solid State Shear Pulverization Technology, GPEC 2007. 28. G. Karayannidis, A. Chatziavgoustis, D.S. Achilias, PET recycling and recovery of pure terephthalic acid by alkaline hydrolysis, Adv. Polym. Technol., 21, 2002, 250259. 29. G.P. Karayannidis , S.A. Dimitris, Chemical Recycling of Poly(ethylene terephthalate), Macromolecular Materials and Engineering, 292, 2, 2007,128146. 30. J. Das, A.B. Halgeri, V. Sahu, P.A. Parikh, Alkaline hydrolysis of Poly(ethylene terephthalate) in presence of a phase transfer catalyst, Indian Journal of Chemical Technology, 14, 2007, 173-177. 31. R. Lpez-Fonseca, M. P. Gonzlez-Marcos, J. R. Gonzlez-Velasco, J. I. Gutirrez- Ortiz, A kinetic study of the depolymerisation of poly(ethylene terephthalate) by phase transfer catalysed alkaline hydrolysis, Journal of Chemical Technology & Biotechnology, 84,1, 2009, 92-99. 32. S.N. Sautin, Planirovanie eksperimentov v himii i himiceskoi tehnologii, Leningrad, 1973, (in Russian). 33. Cheng-Ho Chen, Chuh-Yean Chen, Yu-Wen Lo, Ching-Feng Mao, Wei-Tung Liao, Studies of glycolysis of poly(ethylene terephthalate) recycled from postconsumer soft-drink bottles. II. Factorial experimental design, Journal of Applied Polymer Science,80, 7, 2001, 956-962. 34. Cheng-Ho Chen, Study of glycolysis of poly(ethylene terephthalate) recycled from postconsumer softdrink bottles. III. Further investigation, Journal of Applied Polymer Science,87,12, 2003, 2004-2010.

384

You might also like