Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Polymer International Polym Int 52:337–342 (2003)

DOI: 10.1002/pi.1147

Kinetic and thermodynamic study of


methanolysis of poly(ethylene terephthalate)
waste powder
S Mishra* and AS Goje
Department of Chemical Technology, North Maharashtra University, PO Box 80, Jalgaon 425 001, Maharashtra, India

Abstract: Depolymerization of poly(ethylene terephthalate) waste (PETW) was carried out by


methanolysis using zinc acetate in the presence of lead acetate as the catalyst at 120–140 °C in a closed
batch reactor. The particle size ranging from 50 to 512.5 mm and the reaction time 60 to 150 min
required for methanolysis of PETW were optimized. Optimal percentage conversion of PETW into
dimethyl terephthalate (DMT) and ethylene glycol (EG) was 97.8% (at 120 °C) and 100% (at 130 and
140 °C) for the optimal reaction time of 120 min. Yields of DMT and EG were almost equal to PET
conversion. EG and DMT were analyzed qualitatively and quantitatively. To avoid oxidation/carboni-
zation during the reaction, methanolysis reactions were carried out below 150 °C. A kinetic model is
developed and the experimental data show good agreement with the kinetic model. Rate constants,
equilibrium constant, Gibbs free energy, enthalpy and entropy of reaction are also evaluated at 120, 130
and 140 °C. The methanolysis rate constant of the reaction at 140 °C (10.3 atm) was 1.4  103 g PET
mol1 min1. The activation energy and the frequency factor for methanolysis of PETW were
95.31 kJ mol1 and 107.1 g PET mol1 min1, respectively.
# 2003 Society of Chemical Industry

Keywords: poly(ethylene terephthalate); methanolysis; kinetics; particle size; depolymerization; rate constant

INTRODUCTION from PETW.4 There are solutions proposing a


Methods for the methanolysis of poly(ethylene tereph- combination of high temperature PET methanolysis
thalate) (PET) take account of basic parameters, such with esterification of TPA or with products of p-xylene
as pressure and temperature.1–4 Polymer depolymeri- oxidation.11 It has been observed that the yield of
zation with methanol takes place with the release of DMT from PET methanolysis processes does not
dimethyl terephthalate (DMT) and ethylene glycol exceed 90%.2 Researchers12 have presented a method
(EG).5 Catalysts such as zinc acetate, magnesium of methanolysis of polymer blends. This allows
acetate, cobalt acetate, and lead dioxide enhance the recovery of dimethyl esters of the corresponding
reaction; however, the most commonly used catalyst is dicarboxylic acid and glycol. A method was developed
zinc acetate.6,7 There are examples of using arylsul- in which residue from EG rectification is utilized in
fonic acid salts as catalysts for methanolytic degrada- PETW methanolysis processes in the presence of a
tion of PET.8 Both continuous and batch methods can catalyst.13–16
be applied for methanolysis. Installation for batch In this study, optimal reaction time, reactant par-
methods of methanolysis has been investigated.2 A ticle size, percentage recovery of monomeric products,
two-stage, continuous process was developed by methanolysis rate constant, condensation rate con-
Hoechst,9 in which molten PET was fed in a reactor stant, equilibrium constant, Gibbs free energy, en-
at a ratio of 1:4, respectively. In accordance with thalpy and entropy of reaction were determined for
continuous depolymerization technology, PET waste methanolysis of PETW, using zinc acetate in the
(PETW) is introduced into the reactor in the form of presence of lead acetate as the catalyst at 120, 130 and
an aerosol with an inert gas (ie nitrogen) and methanol 140 °C. A kinetic model was developed and experi-
vapour.1 The Kodak Co possesses a patent describing mental data were fitted with it to validate the model.
a process of PET methanolysis and its optimum The activation energy and the frequency factor for
properties.10 Another approach to methanolysis was methanolysis was 95.31 kJ mol1 and 107.1 g PET
presented in a patent concerning a continuous, two- mol1 min1 respectively. A low temperature (below
stage process for terephthalic acid (TPA) reclamation 150 °C) heterogeneous process was selected and used

* Correspondence to: S Mishra, Department of Chemical Technology, North Maharashtra University, PO Box 80, Jalgaon 425 001,
Maharashtra, India
E-mail: profsm@rediffmail.com
(Received 12 April 2002; revised version received 7 October 2002; accepted 7 October 2002)

# 2003 Society of Chemical Industry. Polym Int 0959–8103/2003/$30.00 337


S Mishra, AS Goje

to avoid oxidation/carbonization of the reactants/ following ways:


products during depolymerization, and simulta-
Depolymerization of PETWð%Þ
neously to save energy. In this work, the simplest
separation techniques for monomers were used and ¼ fðWPETW;i  WPET;R Þ=WPETW;i g  100 ð1Þ
optimized. EG was recovered completely by using a
salting-out method.17 These separation techniques Yield of DMTð%Þ ¼ fmDMT;O =mPETW;i g  100 ð2Þ
allow the complete recovery of monomers (DMT and
Yield of EGð%Þ ¼ ðmEG;O =mPETW;i Þ  100 ð3Þ
EG), and are found most economical and of industrial
significance. where WPETW,i is the initial weight of PETW,
WPETW,R is the weight of unreacted PETW, mDMT,O
is the number of moles of DMT, mEG,O is the
number of moles of EG, and mPETW,i, is the initial
EXPERIMENTAL number of moles of PETW monomers.
Materials In an identical manner the methanolysis was
The PETW used was procured from Garaware undertaken at 130 °C (8.5–7.5 atm), and 140 °C
Polyesters, Aurangabad, MS, (India). Other materials (10.3–8.5 atm) for 110, 120 and 130 min of reaction
used were neutral water, methanol, ethanol, sulphuric time.
acid, hydrochloric acid, calcium oxide, potassium
permanganate, zinc acetate, lead acetate and sodium
Optimization of EG recovery method
sulfate, etc, obtained from SD Fine Chemicals
In order to recover EG from the resulting liquid
(India). These reagents were used without further
(which contained mainly a mixture of EG with excess
purification.
methanol), a batch distillation assembly heated to
60 °C was used Sodium sulfate was introduced into the
Methanolysis of PETW
liquid mixture with constant stirring so as to obtain a
Methanolysis reactions were carried out in a closed
saturated solution17 which was heated slowly up to
batch reactor at 120 °C (7–5 atm). A closed batch
170 °C (below the boiling point of EG, 197 °C) until
reactor ‘Amar Engineering, Mumbai (India) make’ of
the excess methanol was almost distilled away. As a
500-ml capacity equipped with a magnetic drive
result, EG salting-out took place and a separate solid
(stirrer), resistance temperature detector (RTD)
layer of the disodium salt of EG (EG salt) was
sensor, pressure gauge, heater, internal cooling coils
obtained. The batch distillation round-bottom flask
and proportional integral derivative (PID) controller
was cooled to room temperature (26 °C) and the solid
was used to avoid losses of vapour during reaction.
layer was separated from the liquid phase (mainly
Reactions were carried out under pressure to obtain
H2SO4). EG was recovered from EG salt using a
higher conversion of PET at a comparatively lower
stoichiometric amount of HCl and analyzed qualita-
reaction time. Initially various batches of 10 g of PET
tively and quantitatively. During the production of EG
with varying amount of methanol (5–50 ml) were
salt the H2SO4 formed was checked by analysis.
used. For higher optimal conversion of PET, the
H2SO4 was recovered in the form of CaSO4 by adding
amount of methanol was optimized at 30 ml and, for
a stoichiometric amount of CaO. CaSO4 was filtered
further batches, 30 ml of methanol were used. Reac-
from the liquid phase (mainly water).
tion was carried out by taking 10 g of PET waste in
30 ml of methanol using 0.002 mol catalyst (ie
0.001 mol each of lead acetate and zinc acetate) for Determination of molecular weights
different periods of time, ranging from 60 to 150 min. The viscosity-average molecular weight of PETW was
Different particle sizes, ranging from 50 to 512.5 mm, determined by the viscosity method. The solvent used
of PETW were used. Unreacted PETW was sepa- was a mixture of phenol and tetra-chloroethane in the
rated out from the reaction mixture, dried and proportions of 3:5 (v/v). Using the values18 of
weighed. The filtrate was cooled to 12 °C to crystal- k = 22.9  103 and / = 0.73 the molecular weight of
lize the product (DMT) completely, that was then PETW was determined by the relationship18 Zsp/C = k
washed with neutral water and ethanol. The solubi- M/ as equal to 16 700 g mol1. Molecular weights of
lity of DMT in EG–methanol mixture is complete at the residual polyester were also determined by the
120–140 °C but becomes zero at 12 °C. Hence, the same technique and the results ranged from
liquid phase was cooled to 12 °C. DMT was dried in 16 700 g mol1 to 16 665 g mol1, that is very close to
an oven at 80 °C to constant weight and the weight of 16 700 g mol1. This shows that the depolymerization
DMT was measured. From the remaining liquid of PETW in methanol took place only at the PETW
phase, EG was separated by a salting-out method17 surface.
by introducing sodium sulfate. Both monomeric
products (DMT and EG) were analyzed qualitatively Process reaction of depolymerization
and quantitatively. The percentage depolymerization PETW powder was depolymerized using methanol at
of PETW, the yield of DMT, and the yield of EG 120–140 °C with zinc acetate in the presence of lead
were determined by gravimetry and defined in the acetate. The chemical reaction is given in Scheme 1.

338 Polym Int 52:337–342 (2003)


Methanolysis of poly(ethylene terephthalate) waste

Scheme 1. Depolymerization of PETW


powder with methanol in the presence
of catalyst.

Chemical reactions involved during EG recovery merization of PETW was analyzed by determining its
operations density (1.113 g ml 1), boiling point (197 °C), mol-
ecular weight (62 g mol1). These results are close to
HOH2 C CH2 OH þ Na2 SO4 !
the values for ethylene glycol.
NaOH2 C CH2 ONa þ H2 SO4 The solid monomeric product melts at 141 °C, boils
at 284 °C, has an acid value of 0.03, and a molecular
NaOH2 C CH2 ONa þ 2 HCl ! weight of 194 g mol1. These results are close to the
values for DMT. These results clearly indicate that EG
HOH2 C CH2 OH þ 2 NaCl
and DMT are the liquid and solid products respec-
H2 SO4 þ CaO ! CaSO4 þ H2 O tively, obtained from the depolymerization of the
waste polyester.

Analysis of depolymerized products of waste


polyester Kinetic model and determination of rate and
Liquid and solid products obtained from depolymer- equilibrium constants
ization of PETW were analyzed by determining their The rate of appearance of carboxymethyl groups
various physical properties (melting point, boiling (—COOCH3) can be expressed as follows:
point, molecular weight, acid value, etc) to confirm  rA ¼ dCCOOCH3 =dt ¼ kCEL CMEOH  k0 CCOOCH3 COH
the stoichiometry with the polyester used. ð5Þ
where CEL and COH represent the concentration of
RESULTS AND DISCUSSION ester linkages and EG. k and k’ refer to rate constants
Analysis of depolymerized products of waste for the methanolysis and condensation reactions,
polyesters respectively. The kinetic model for the methanolysis
The liquid monomeric product obtained from depoly- is given in eqn (6).
fð1=CMEOH Þ ln½1=ð1  xÞg ¼ ðktÞ þ c: ð6Þ
The value of k, can be determined from a plot of
(1/CMEOH) ln [1/(1  x)] versus t. This plot (Fig 1) gives
a value of 6.76  104 g PET mol1 min1 for k (at
120 °C) and y-intercept (c = 0.072). The value of k’,
can be calculated from the equilibrium constant
(Table 1). Vanecso-Szmercsanyi et al 19 calculated Ke
and found that it was independent of the acid used.
The values of Ke presented in Table 1 are seen to be
quite sensitive to temperature. The values of k and k’

Table 1. Equilibrium concentrations of products and reactants,


and equilibrium constant for an initial reactor charge ratio of
2.79 g MeOH/g PET (at 120-min reaction time)

Temp ( °C) P (atm) CMeOH CEL COH Ke


120 7 25.90 0.832 0.89 0.0343
130 8.5 23.98 0.894 0.77 0.0250
Figure 1. Fitting the kinetic model eqn (6) with experimental data at 120 °C
140 10.3 22.10 0.965 0.65 0.0174
(7 atm) for 127.5-mm particle sizes as a function of time.

Polym Int 52:337–342 (2003) 339


S Mishra, AS Goje

Temp ( °C) P (atm) CMeOH (mol/g PET) k (g PET mol1 min1) k’ (g PET mol1 min1)
Table 2. Methanolysis and condensation 120 7 25.90 0.000676 0.01971
rate constants from initial rate data for a 130 8.5 23.98 0.000899 0.03596
reactor charge ratio of 2.79 g MeOH/g PET
140 10.3 22.10 0.001392 0.08000
(at 120 min)

effect of temperature on Ke, DG, DH and DS is shown


in Table 3.

Optimization of particle size of PET waste by


methanolysis
Results of optimization of particle size of PET waste
for methanolysis are shown in Table 4. Particle sizes
ranging from 50 mm to 512.5 mm were used and
illustrate the decrement in percentage conversion from
97.8 to 26.8% with increasing particle size. The
percentage conversion of polyester waste remains
constant and maximum (97.8%) up to 127.5 mm of
particle size. Further increase in the particle size of
PETW causes a decrease in depolymerization. Hence
127.5 mm was recorded as optimal particle size.

Optimization of reaction time for methanolysis


Figure 2. Arrhenius plot for an initial charge ratio of 2.79 g MeOH/g PET.
Maintaining the particle size of PET waste at
127.5 mm, the reaction time was varied from 60 to
150 min for depolymerization of polyester waste by the
Table 3. Variation of Ke, DG, DH, and DS with temperature in methanolysis. methanolysis method at 120 °C (Fig 3). The minimum
Temp (K) Ke DG J mol1 DH J mol1 DS J K1 mol1 percentage conversion (43.6%) was obtained for
60 min and the maximum (97.8%) recorded for
393 0.0343 11 000 38 600 126
120 min reaction time. Subsequently was no incre-
403 0.0250 12 300 60 600 180
ment noticed in the conversion of PET waste with
413 0.0174 13 900 60 900 182
time. Complete PET conversion was observed at 130
and 140 °C for 120 min, which is therefore the optimal
reaction time.
are also temperature dependent (Table 2). Data in
Table 2 are fitted to the Arrhenius equation in Figure EG recovery method
2. A frequency factor of 107.1 g PET mol1 min1 was Recovery techniques recorded 97.8% yield of EG at
obtained. An activation energy (Ea) of 95.31 kJ mol1 120 °C. Yield of EG is consistent with PET conver-
is obtained, that is slightly smaller than the value sion. The white precipitation test, colour stability of
obtained by Davies et al 20 and McMahon et al 21 of the calcium compound test, and EG recovery results
109–113 kJ mol1, but larger than the value of clearly indicated that EG does not undergo carboniza-
75.4–92.1 kJ mol1 obtained by Ravens.22 Therefore, tion or oxidation under the conditions used. In
methanolysis is rate determining on the PET surface. another test, a calcium compound was placed in
At 100 °C, Marathe et al 2 obtained a rate constant of sulfuric acid solution with diluted KMnO4. The colour
3.0  107 min1. We recorded a value of k as 6.44  was stable. This also confirms the absence of the
104 g PET mol1 min1 at 120 °C and 120 min. The oxalate compound in the calcium compound. The

PETW size (mm) PET conv (%) DMT yield (%) EG yield (%) rA (g mol L1 min1)
50 97.8 97.64 97.62 0.00019
64 97.8 97.64 97.62 0.00019
90 97.8 97.64 97.62 0.00019
127.5 97.8 97.64 97.62 0.00019
181 81.6 81.54 81.51 0.00014
256 62.4 62.36 62.33 0.00011
362.5 37.1 36.86 36.82 0.0006
Table 4. Kinetics of particle size of PET waste
512.5 26.8 26.72 26.69 0.0002
in methanolysis for 120-min reaction time.

340 Polym Int 52:337–342 (2003)


Methanolysis of poly(ethylene terephthalate) waste

CONCLUSION
Depolymerization of poly(ethylene terephtalate)
(PETW) was carried out in the presence of catalysts
at different temperatures, different reaction times and
varying the particle sizes of PETW.
The optimal percentage conversion of PETW
obtained was 97.8% at 120 °C, and 100% at 130
and 140 °C, for the optimal reaction time of 120 min,
keeping the optimal particle size as 127.5 mm. Yields of
dimethyl terephthalate (DMT) and ethylene glycol
(EG) were almost equal to the PET conversion. The
methanolysis rate constant, condensation rate con-
stant, equilibrium constant, activation energy, fre-
quency factor, Gibbs free energy, enthalpy and
entropy of reaction were determined for methanolysis
of depolymerization, using zinc acetate in the presence
of lead acetate as the catalyst at 120, 130 and 140 °C.
The reaction rate constant at 120 °C (7 atm) for
this reaction was 6.76  104 g PET mol1 min1. A
kinetic model was developed and the experimental
Figure 3. Effect of time on conversion at 120 °C (7 atm) for 127.5-mm data fitted with it to validate the model; experimental
particle size. data showed good agreement with it. Activation energy
and frequency factor for methanolysis of PETW equal
to 95.31 kJ mol1 and 107.1 g PET mol1 min1 were
remaining waste aqueous phase was concentrated by recorded, respectively.
evaporation and re-used in the depolymerization pro- A low temperature (below 150 °C) heterogeneous
cess. process was selected and used to avoid oxidation/
carbonization of reactants/products during de-
polymerization, and simultaneously to save energy.
Kinetics of methanolysis of PETW Simplest separation techniques for monomers (DMT
The variation in reaction rate with particle size is also and EG) were optimized and used. EG was recovered
shown in Table 4. Effect of time on rate of reaction is completely by using the salting-out method by
shown in Table 5: it decreases with time. Initially it introducing sodium sulfate in the resulting liquid
remains constant for particle sizes of 50, 64, 90 and phase. These techniques recover completely the
127.5 mm and later it slightly decreases with increasing monomers, and are found most economical, which
particle size. This may be due to greater particle size or has industrial significance.
experimental error. Reactions were mostly carried out
for differing periods of time ie 60–150 min, maintain-
ing the temperature constant (120 °C). The rate of
reaction was determined on the basis of the degree of REFERENCES
methanolysis (X) of PETW for that period of time 1 Lotz R, Wick G and Neuhaus C, Process for the recovery of
(Table 5). The concentration of PETW is found to dimethyl terephthalate from polyethylene terephthalate, US
decrease with increasing reaction time up to 120 min, Patent 3 321 510 (1967).
2 Marathe MN, Dabholkar DA and Jain MK, Process for the
and later remains constant at 120 °C. Concentration of recovery of dimethyl terephthalate from PET, GB Patent 2 041
reactant (PETW) and rate of reaction are decreasing 916 (1980).
with time (Table 5). 3 Michel RE, Recovery of methyl esters of aromatic acids and

Time (min) PET conv (%) Xa DMT yield (%) EG yield (%) rA (g mol L1 min1)
60 43.6 0.436 43.52 43.48 0.00035
70 55.7 0.557 55.64 55.61 0.00032
80 67.3 0.673 67.28 67.25 0.00030
90 75.2 0.752 75.17 75.14 0.00027
100 86.2 0.862 86.16 86.12 0.00024
110 95.2 0.952 95.16 95.13 0.00021
120 97.8 0.978 97.76 97.72 0.00019
130 97.8 0.978 97.76 97.72 0.00018
140 97.8 0.978 97.76 97.72 0.00017
Table 5. Kinetics of methanolysis for 150 97.8 0.978 97.76 97.72 0.00015
127.5mm particle size of PET waste as a
a
function of time. X is the degree of methanolysis of PETW.

Polym Int 52:337–342 (2003) 341


S Mishra, AS Goje

glycols from thermoplastic polyester scrap using methanol carboxylate and dimethyl terephthalate for recycle, Japan
vapour, European Patent 484 963 (1992). Patent 07 196 578 (1995); Chem Abstr 123:257777h.
4 Socrate C and Vosa R, Continuous process for the recovery of 13 Mikolajczyk B, Lubawy A, Deerskin M, Smoczynski P, Pozniak
terephthalic acid from waste or used polyalkylene terephthalate A and Boebel H, Separation of dimethyl terephthalate ethylene
polymers, European Patent 662 466 (1995); Chem Abstr glycol from residues from rectification of waste ethylene glycol,
123:257959u. Polish Patent 126 009 (1985).
5 Chandra R and Adab A, Rubber and Plastic Waste: Recycling, Reuse 14 Mishra S, Goje AS and Zope VS, Proc Internat Conf Plastic
and Future Demand, 1st edn, CBS Publishers & Distributors, Waste Manag Environment, Mar 15–16, New Delhi, p 163
New Delhi (1994). (2001).
6 Improvements in the preparation of high quality dimethyl 15 Kint D and Munoz-Guerra S, Polym Int 48:346 (1999).
terephthalate, GB Patent 748 248 (1957). 16 Ostrysz R, Kicko-Walczak E, Milunski P, Jakubas T, Mental Z,
Okorowski J and Koscielski K, Method of preparation of
7 Dimov K and Terlemezyan E, Catalytic action of calcium acetate
unsaturated terephthalic polyester resins, Polish Patent 150
and the manganese acetate–sodium acetate mixture in the pre-
126 (1990).
esterification of dimethyl terephthalate with EG. J Polym Sci,
17 Paszun D and Spychaj T, Chemical recycling of poly(ethylene
Part I 10:3133 (1972).
terephthalate). Ind Engng Chem Res 36:1373 (1997).
8 Process for the conversion of polyethylene terephthalate into
18 Gowariker VR, Visawanathan NV and Sreedhar Jayadev, Polymer
dimethyl terephthalate, GB Patent 806 269 (1958). Science, 3rd reprint, Wiley Eastern Ltd, New Delhi. p 356
9 Gruschke H, Hammerschick W and Medem W, Process for (1990).
depolymerizing polyethylene terephthalate to terephthalic acid 19 Vanesco-Szmercsanyi E, Makay-Body E et al, Kinetics and
dimethyl ester, US Patent 3 403 115 (1968). Mechanism of Polyreactions, Vol 1, IUPAC Symposium Pre-
10 Debruin BR, Naujokas AA and Gamble WJ, Process for prints, Budapest, 103 pp (1969).
recovering components from polyester resins, US Patent 5 20 Davies T, Goldsmith PL, Ravens DAS and Ward IM, The
432 203 (1995); Chem Abstr 123:229404r. kinetics of hydrolysis of PET films. J Phys Chem 66:175 (1962).
11 Pitat J, Holcik V and Bacak MA, Method of processing waste of 21 McMahon W, Birdsall HA, Johnson GR and Canilli CT,
polyethylene terephthalate by hydrolysis, GB Patent 822 834 Degradation studies of PET. J Chem Engng Data 4:457 (1959).
(1959). 22 Ravens DAS, The Chemical reactivity of PET. Heterogeneous
12 Sato K and Sumitani K, Recovery of dimethyl naphthalenedi- Hydrolysis by HCl. Polymer 1:375 (1960).

342 Polym Int 52:337–342 (2003)

You might also like