You are on page 1of 7

Energy Conversion and Management 51 (2010) 1363–1369

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Thermal degradation behaviors of polyethylene and polypropylene. Part I: Pyrolysis


kinetics and mechanisms
A. Aboulkas a,b,*, K. El harfi a,b, A. El Bouadili b
a
Laboratoire de Recherche sur la Réactivités des Matériaux et l’Optimisation des Procédés «REMATOP», Département de chimie, Faculté des Sciences Semlalia,
Université Cadi Ayyad, BP 2390, 40001 Marrakech, Morocco
b
Laboratoire Interdisciplinaire de Recherche en Sciences et Techniques, Faculté polydisciplinaire de Béni-Mellal, Université Sultan Moulay Slimane, BP 592,
23000 Béni-Mellal, Morocco

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

Article history: Study of the decomposition kinetics is an important tool for the development of polymer recycling in
Received 24 July 2008 industrial scale. In this work, the activation energy and the reaction model of the pyrolysis of high den-
Received in revised form 18 May 2009 sity polyethylene (HDPE), low density polyethylene (LDPE) and polypropylene (PP) have been estimated
Accepted 13 December 2009
from non-isothermal kinetic results. Firstly, the activation energy values obtained by Friedman, Kissing-
Available online 20 January 2010
er–Akahira–Sunose and Flynn–Wall–Ozawa isoconversional methods, are 238–247 kJ/mol for HDPE,
215–221 kJ/mol for LDPE and 179–188 kJ/mol for PP. Secondly, the appropriate conversion model of
Keywords:
the process was determined by Coats–Redfern and Criado methods. The pyrolysis reaction models of
HDPE
LDPE
HDPE and LDPE are accounted for by ‘‘Contracting Sphere” model, whereas that of PP by ‘‘Contracting
PP Cylinder” model.
Thermal analysis Ó 2009 Elsevier Ltd. All rights reserved.
Isoconversional methods
Master-plot method

1. Introduction contributes to the solution of environmental problems [5]. The


determination of the parameters of the thermal decomposition
Polyethylene and polypropylene are a major component of plas- process by means of thermogravimetric techniques allows the
tic waste from domestic refuse. Its efficient reutilization has a development of the recycling process of these materials in an
growing importance these years due to the increased demand for industrial scale.
resource recycling and environmental protection. In general, plas- Thermal behavior of plastics can be improved by knowing ther-
tic waste has been mainly disposed of by landfill or incineration, mal degradation kinetics. Many studies on pyrolysis kinetics of
but these processes are not fully acceptable under current interna- plastic wastes have been carried out, and most of these studies
tional policy, which focuses on efficient recovery of raw material have been developed on the assumption that the reaction can be
and energy. Pyrolysis and gasification processes are promising described by a nth order reaction model [3,6–10]. The assumption
routes for optimal upgrading from waste. Moreover, pyrolysis of of a nth order reaction model would result in the Arrhenius param-
plastic, based on the decomposition of polymers at different tem- eters deviating from the real ones. Few attempts, however, have
peratures, allows the treatment of polymers with simultaneous been made to identify the reaction model of the polyethylene
decomposition and separation [1,2]. These processes allow the and polypropylene pyrolysis [11,12]. Isothermal reduced-time
obtainment of combustible, gases and/or energy, with the reduc- plots have been applied to determine the reaction model of the
tion of landfilling as an added advantage [3,4]. The first step for a thermal decomposition of solids [13–17]. Kim et al. [11,12,15–
suitable design of any pyrolysis reactor is knowledge of the 17] confirmed the applicability of the isothermal RTP to identify
kinetics. the pyrolysis reaction models of polyethylene [12,15,16] and poly-
Thermal degradation of polymers has great interest as an alter- propylene [11,17]. However, any one have been attempted to iden-
native source of energy or chemical raw materials, as well as it tify reaction models using Coats–Redfern and Criado methods
under non-isothermal conditions that were new trials.
* Corresponding author. Address: Laboratoire Interdisciplinaire de Recherche en A thermogravimetric analysis (TGA) technique is an excellent
Sciences et Techniques, Faculté polydisciplinaire de Béni-Mellal, Université Sultan way for studying the kinetics of thermal degradation [18–20]. It
Moulay Slimane, BP 592, 23000 Béni-Mellal, Morocco. Tel.: +212 23 485112/22/82;
provides information on activation energy and kinetic model.
fax: +212 23 485201.
E-mail address: a.aboulkas@yahoo.fr (A. Aboulkas). Two materials in significant quantities in municipal plastic wastes

0196-8904/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2009.12.017
1364 A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369

is polyethylene (LDPE and HDPE) and polypropylene wastes. Thus, w0  w


x¼ ð2Þ
the objective of this work was to study the thermal decomposition w  wf
kinetics of these kinds of plastic wastes which have the highest
where w is the weight of the sample at a given time t, w0 and wf,
portion (70%) of municipal and industrial plastic wastes in Moroc-
refer to values at the beginning and the end of the weight loss event
co [21].
of interest. f(x) and K(T) are functions of conversion and tempera-
The aim of this work is specifically to show the utility of a ki-
ture, respectively.
netic analysis method based on the complementary use of isocon-
K(T) the temperature dependence of the rate of weight loss, is
versional, Coats–Redfern and master-plot methods from non-
often modeled successfully by the Arrhenius equation
isothermal thermogravimetric data. These are new attempts for  
thermoplastics in view of applying Coats–Redfern and Criado E
KðTÞ ¼ A exp  ð3Þ
methods to identify the pyrolysis reaction model, possibly contrib- RT
uting to the research field of pyrolysis kinetics of thermoplastic.
where E is the activation energy, A the pre-exponential factor and R
In this work, the thermogravimetric study of HDPE, LDPE and PP
the gas constant.
was realized using non-isothermal method in order to determine
By combining the Eqs. (1) and (3), the reaction rate can be writ-
the apparent activation energy and the pyrolysis reaction model.
ten in the form:
 
2. Experimental dx E
b ¼ A exp  f ðxÞ ð4Þ
dT RT
2.1. Plastics investigated
3.1. Friedman method (FR) [22]
Three plastic types (HDPE, LDPE and PP) were used in this work.
The plastic were obtained from Plador plastic, Marrakech, Morocco.
This method is a differential isoconversional method, and it di-
The results of characterization of these plastic are given in Table 1.
rectly based on Eq. (4) whose logarithm is
   
2.2. TGA procedure dx dx E
ln ¼ ln b ¼ ln½Af ðxÞ  ð5Þ
dt dT RT
Plastic samples were subjected to thermogravimetric analysis
From this equation, it is easy to obtain values for E over a wide
(TGA) in an inert atmosphere of nitrogen. Rheometrix Scientific  dx 
range of conversions by plotting ln b dT against 1T for a constant x
STA 1500 TGA analyzer was used to measure and record the sam-
value.
ple mass change with temperature over the course of the pyrolysis
reaction. Thermogravimetric curves were obtained at four different
3.2. Kissinger–Akahira–Sunose method (KAS) [23,24]
heating rates (2, 10, 20 and 50 K/min) between 300 K and 975 K;
the precision of reported temperatures was estimated to be
The standard Eq. (4) can be shown as follows:
±2 °C. Nitrogen gas was used as an inert purge gas to displace air
 
in the pyrolysis zone, thus avoiding unwanted oxidation of the dx A E
sample. A flow rate of around 60 ml min1 was fed to the system
¼ exp  dT ð6Þ
f ðxÞ b RT
from a point below the sample and a purge time of 60 min (to be
sure the air was eliminated from the system and the atmosphere which is integrated with the initial condition of x = 0 at T = T0 to ob-
is inert). The balance can hold a maximum of 45 mg; therefore, tain the following expression:
all sample amounts used in this study averaged approximately Z x Z T    
dx A E AE E
20 mg. gðxÞ ¼ ¼ exp  dT  p ð7Þ
The reproducibility of the experiments is acceptable and the 0 f ðxÞ b T0 RT bR RT
experimental data presented in this paper corresponding to the
Essentially the technique assumes that the A, f(x) and E are
different operating conditions are the mean values of runs carried
independent of T while A and E are independent of x, then Eq. (7)
out two or three times.
may be integrated to give the following equation in logarithmic
form:
3. Theoretical consideration    
AE E
ln gðxÞ ¼ ln  ln b þ ln p ð8Þ
Generally for polymer degradation, it is assumed that the rates R RT
of conversion are proportional to the concentration of reacted
The KAS method is based on the Coats–Redfern approximation
material. The rate of conversion can be expressed by the following
[25] according to which:
basic rate equation  
 
E exp  E
dx dx p ffi  2RT ð9Þ
¼b ¼ KðTÞf ðxÞ ð1Þ RT E
dt dT RT

where x is the degree of advance defined by From relationships (7) and (9) it follows that:

Table 1
Some characteristics of HDPE, LDPE and PP.

Sample Proximate analysis (wt.%) Elemental analysis (wt.% dry ash free)
Volatile matter Ash C H N S
HDPE 99.4 0.6 85.5 14.2 – 0.3
LDPE 99.6 0.4 85.5 14.3 – 0.2
PP 99.1 0.9 85.1 14.4 – 0.5
A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369 1365

b AR E 4. Results and discussion


ln ¼ ln  ð10Þ
T2 EgðxÞ RT
4.1. Thermal degradation

3.3. Flynn–Wall–Ozawa method (FWO) [26,27] Figs. 1–3 show the TG/DTG curves for HDPE, LDPE and PP at dif-
ferent heating rates (2, 10, 20 and 50 K/min), respectively. Maxi-
This method is derived from integral isoconversional method. mum degradation temperatures for HDPE, LDPE and PP are
Using Doyle’s approximation [28] for the integral which allows summarized in Table 2. The weight loss curves have almost the
 E E
 same trends; this indicates that they have the same pyrolysis
ln p RT ffi 5:331  1:052 RT , Eq. (8) now can be simplified as
behaviors due to similar chemical bonds in their molecular struc-
AE E tures. The weight losses show that degradation of plastic occurs al-
ln b ¼ ln  5:331  1:052 ð11Þ most totally in one-step process as can be concluded by the
RgðxÞ RT
presence of only one peak in DTG. The plastic thermal degradation
Thus, for x = constant, the plot ln b vs. 1T , obtained from thermo- starts at 577 K and is almost complete at approximately 850 K. As
grams recorded at several heating rates, should be a straight line the heating rate is increased, Table 2 shows that there was a lateral
whose slope can be used to evaluate the activation energy. shift to higher temperatures for Tmax. The lateral shift is also illus-
trated in Figs. 1–3. The rate of weight loss also reflects the lateral
shift with an increase in the rate as the heating rate was increased
3.4. Coats–Redfern method [25] from 2 K/min to 50 K/min.
The displacement of TG curves with heating rate has been de-
Coats–Redfern method is also an integral method, and it in- scribed in the literature by different researchers [7,10,30]. There
volves the thermal degradation mechanism. Using an asymptotic are various arguments to explain these displacements. Many
approximation for the resolution of Eq. (7) (2RT=E  1), the follow- authors consider that a change in the mechanism of the reaction
ing equation can be obtained: can be produced when there is a change in the heating rate. In con-
trast inefficient heat transfer from furnace to sample may produce
gðxÞ AR E large differences between the temperature of the furnace and that
ln ¼ ln  ð12Þ
T2 bE RT of the sample, which obviously increase with heating rate. How-
ever, the displacement observed can also be explained by the
mathematical form of the kinetic laws, which may lead to displace-
3.5. Criado method [29] ments of the curves as the heating rate increases [31].

The kinetic model of the process can be determined by this


4.2. Calculation of the activation energy
method. Combining the Eq. (4) with Eq. (12), the following equa-
tion is obtained:
For the calculation of the activation energies all heating rates
 2 have been used and they were estimated using the Friedman, Kis-
ZðxÞ f ðxÞgðxÞ Tx ðdx=dtÞx singer–Akahira–Sunose and Flynn–Wall–Ozawa methods for com-
¼ ¼ ð13Þ
Zð0:5Þ f ð0:5Þgð0:5Þ T 0:5 ðdx=dtÞ0:5 parison reasons.
Firstly, the isoconversional Friedman method was used to calcu-
where 0.5 refers to the conversion in x = 0.5. late the activation energy for different conversion values by plot-
f ðxÞgðxÞ
The left side of Eq. (13) f ð0:5Þgð0:5Þ is a reduced theoretical curve, ting of ln(dx/dT) against 1/T for a constant a value and calculated
which is characteristic of each reaction mechanism, whereas the the activation energy. The results for HDPE, LDPE and PP are shown
right side of the equation associated with the reduced rate can in Tables 3–5, respectively. The mean values of activation energy
be obtained from experimental data. A comparison of both sides were 247, 221 and 188 kJ/mol for HDPE, LDPE and PP, respectively.
of Eq. (13) tells us which kinetic model describes an experimental Secondly, the isoconversional Kissinger–Akahira–Sunose meth-
reactive process. Table 2 indicates the algebraic expressions of f(x) od was used to calculate the activation energy for different conver-
and g(x) for the kinetic models used. sion values by plotting ln(b/T2) against 1/T and the results are

Table 2
Algebraic expressions of functions of the most common reaction mechanisms.

Mechanism f(x) g(x)


Power law (P2) 2x1/2 x1/2
Power law (P3) 3x2/3 x1/3
Power law (P4) 4x3/4 x1/4
Avarami-Erofe’ve (A2) 2(1  x)[ln(1  x)]1/2 [ln(1  x)]1/2
Avarami-Erofe’ve (A3) 3(1  x)[ln(1  x)]2/3 [ln(1  x)]1/3
Avarami-Erofe’ve (A4) 4(1  x)[ln(1  x)]3/4 [ln(1  x)]1/4
Contracting Sphere (R2) 2(1  x)1/2 [1  (1  x)1/2]
Contracting Cylinder (R3) 3(1  x)2/3 [1  (1  x)1/3]
One-dimensional diffusion (D1) 1/2x x2
Two-dimensional diffusion (D2) [ln(1  x)]1 [(1  x) ln(1  x)] + x
Three-dimensional diffusion, Jander (D3) 3(1  x)2/3/[2(1  (1  x)1/3)] [1  (1  x)1/3]2
Ginstling-Brounshtein (D4) 3/2((1  x)1/3  1) 1  (2x/3)  (1x)2/3
First-order (F1) (1  x) ln(1  x)
Second-order (F2) (1  x)2 (1  x)1  1
Third-order (F3) (1  x)3 [(1  x)2  1]/2
1366 A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369

100 Table 4
2 K/min Activation energies of LDPE obtained by FR, KAS and FWO methods.
10 K/min
Weight left (%)
80
20 K/min Conversion x FR KAS FWO
50 K/min
60 150
0.1 224 194 208
-dw/dt (%/min)

120
0.2 226 212 211
90
40 0.3 226 217 223
60
0.4 225 223 226
30
20 0.5 223 225 231
0
600 900
0.6 219 227 230
Temperature (K)
0 0.7 217 226 227
300 600 900 0.8 215 214 207
Temperature (K) 0.9 214 197 199
Mean 221 ± 3 215 ± 8 218 ± 7
Fig. 1. TG curves of HDPE at different heating rates. Inset: corresponding DTG
curves.

Table 5
100 Activation energies of PP obtained by FR, KAS and FWO methods.
2 K/min Conversion x FR KAS FWO
Weight left (%)

80 10 K/min
20 K/min 0.1 175 158 163
60 50 K/min 0.2 181 162 170
160
0.3 189 170 173
-dw/dt (%/min)

120
40 0.4 196 179 182
80
0.5 198 185 188
40
20 0.6 197 189 192
0
600
Temperature (K)
900 0.7 191 191 194
0 0.8 186 191 194
300 600 900 0.9 178 186 190
Temperature (K)
Mean 188 ± 6 179 ± 8 183 ± 8
Fig. 2. TG curves of LDPE at different heating rates. Inset: corresponding DTG
curves.
activation energy of HDPE, LDPE and PP are summarized in Tables
3–5, respectively. The mean values of activation energy were 243,
100 218 and 183 kJ/mol for HDPE, LDPE and PP, respectively.
2 K/min
Figs. 4–6 shows the dependence of activation energy on the
Weight left (%)

80 10 K/min
20 K/min conversion for x = 0.1–0.9 for HDPE, LDPE and PP based on three
50 K/min
60 160 isoconversional methods, respectively. It is evident that the values
-dx/dt (%/min)

120
of the apparent activation energies obtained by FR method are
40 80
higher than the values of apparent activation energies obtained
40
20 0
by FWO and KAS methods. Independent of the calculation proce-
600 900
Temperature (K) dure used, the activation energy is practically constant in the
0
0.1–0.9 range. The dependence of activation energy on conversion
300 600 900
Temperature (K) value, as calculated with FR, FWO and KAS methods, presents al-
most the same tendency. Small variations are basically observed
Fig. 3. TG curves of PP at different heating rates. Inset: corresponding DTG curves. at the very initial and final stages, with the Friedman method being
more sensitive in those areas.
Differential method uses the point value of the overall reaction
Table 3
Activation energies of HDPE obtained by FR, KAS and FWO methods. rate, while the integral method uses integrals which describe the
history of the system. The above differences in determined values
Conversion x FR KAS FWO
of activation energy by FR, FWO and KAS methods can be expected,
0.1 247 207 208 because FWO and KAS methods involve a systematic error in acti-
0.2 257 235 236
vation energy that does not appear in the FR method. Thus, the
0.3 256 242 244
0.4 255 247 248
activation energies obtained as a function of the degree of conver-
0.5 252 249 250 sion from the FR methods are more reliable than those obtained
0.6 232 250 252 from the FWO and KAS methods [32].
0.7 235 248 249
0.8 248 249 250
Activation energy (kJ/mol)

0.9 237 245 247 400


FR: 247 kJ/mol
Mean 247 ± 5 238 ± 11 243 ± 11 350 KAS: 238 kJ/mol
FWO: 243 kJ/mol
300
250
shown in Tables 3–5 for HDPE, LDPE and PP, respectively. The 200
mean values of activation energy were 238, 215 and 179 kJ/mol
150
for HDPE, LDPE and PP, respectively.
Secondly, Flynn–Wall–Ozawa method is an integral method, 100
0,0 0,2 0,4 0,6 0,8 1,0
which is also independent of the degradation mechanism. Eq.
Conversion x
(10) was used, and the activation energy of HDPE, LDPE and PP
was obtained from plot of log(b) against 1/T at a fixed conversion Fig. 4. The dependence of activation energy on the conversion (x) for thermal
with the slope of such a line being 0.4567E/RT. The values of degradation of HDPE according to FR, KAS and FWO methods.
A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369 1367

350 and González [34] performed non-isothermal TGA on LDPE and


Activation energy (kJ/mol) FR: 221 kJ/mol
KAS: 215 kJ/mol
PP and found activation energies of 220–259 kJ/mol and 153–
300
FWO: 218 kJ/mol 265 kJ/mol, respectively.
250 The method by Coats–Redfern is one of the most widely used
procedure for the determination of the reaction processes
200 [25,35]. From Eq. (12), proposed by Coats and Redfern, the activa-
tion energy for all g(x) functions listed in Table 2 can be obtained at
150
constant heating rate. In this study, the same conversion values
100 have been used as those used in the isoconvensional methods. Ta-
0,0 0,2 0,4 0,6 0,8 1,0 ble 6 shows activation energy and correlations for HDPE at con-
Conversion x stant heating rate of 10 °C/min. It was found that the thermal
degradation mechanism of HDPE is likely to be of R2 type, because
Fig. 5. The dependence of activation energy on the conversion (x) for thermal
degradation of LDPE according to FR, KAS and FWO methods.
this mechanism presents the activation energy (248 kJ/mol) that is
similar to the value obtained by FR isoconvensional methods
(247 kJ/mol). The type of degradation mechanism is confirmed by
Criado method in the next determination.
300
Activation energy (kJ/mol)

As presented in Table 6, Coats–Redfern method was applied at


FR: 188 kJ/mol
250 KAS: 179 kJ/mol the heating rate of 10 °C/min to determine the average values of
FWO: 183 kJ/mol the activation energy and the degradation mechanism of LDPE. It
200 is clearly shown that the mechanism for LDPE degradation is pro-
posed to be R2 type. The activation energy of this mechanism was
150
around 214 kJ/mol, which was similar to activation energy ob-
100 tained by KAS isoconvensional method (215 kJ/mol).
By using Coats–Redfern method, the calculated activation en-
50 ergy and degradation mechanism of PP at 10 °C/min are presented
0,0 0,2 0,4 0,6 0,8 1,0
Conversion x in Table 6. The thermal degradation mechanism of PP is proposed
to be of R3 type. The activation energy of this mechanism (187 kJ/
Fig. 6. The dependence of activation energy on the conversion (x) for thermal mol) is in agreement with the results in FR isoconvensional meth-
degradation of PP according to FR, KAS and FWO methods. od (188 kJ/mol).

4.3. Determination of the degradation mechanism using Criado


From the Tables 3–5, the activation energy values have been method
found 238–247 kJ/mol for HDPE, 215–221 kJ/mol for LDPE and
179–183 kJ/mol for PP. These values may be compared to litera- The Z(x)/Z(0.5) master curves can be plotted using Eq. (13)
ture. Wu et al. [30] obtained activation energy (Friedman method) according to different reaction mechanisms g(x) shown in Table
of 233, 206 and 184 kJ/mol for HDPE, LDPE and PP, respectively. 1. Here, the used experimental TG data are still from TG curve at
Paik and Kar [33] investigated the kinetics of thermal degradation 10 °C/min heating rates. Figs. 7–9 show the Z(x)/Z(0.5) master
of polypropylene by applying isoconversional methods of FWO and and experimental curves of HDPE, LDPE and PP, respectively. As
Friedman. He reported activation energy of 125–224 kJ/mol and can be seen, the experimental curve of HDPE, LDPE and PP nearly
67–241 kJ/mol for FWO and Friedman methods, respectively. Yang overlaps the master curve Z(R2)/Z(0.5), indicating that the thermal
et al. [7] in the non-isothermal analysis of HDPE, reported value of degradation of HDPE, LDPE belongs to R2 reaction mechanism and
activation energy (Friedman method) of 240–260 kJ/mol. Other PP could possibly be governed by R3 type. The thermal degradation
workers have also reported similar apparent activation energies of HDPE and LDPE is R2 kinetic reaction and PP is R3 mechanism,
for the HDPE degradation, for example, between 273 kJ/mol and which is in accordance with the result obtained by Coats–Redfern
164 kJ/mol for heating rate between 5 and 50 °C/min [3]. Encinar method.

Table 6
Activation energies of HDPE, LDPE and PP obtained by Coats–Redfern method.

Model HDPE LDPE PP


Activation energy (kJ/mol) Correlation coefficient Activation energy Correlation coefficient Activation energy Correlation coefficient
(kJ/mol) (kJ/mol)
P2 94 ± 25 0.98818 90 ± 14 0.98818 73 ± 15 0.99043
P3 61 ± 17 0.98546 59 ± 10 0.98546 45 ± 22 0.99166
P4 42 ± 12 0.98435 48 ± 21 0.98435 31 ± 18 0.99265
A2 98 ± 10 0.99712 95 ± 9 0.99712 105 ± 20 0.98299
A3 97 ± 30 0.99698 69 ± 15 0.99698 66 ± 23 0.98109
A4 72 ± 21 0.99665 70 ± 11 0.99665 47 ± 19 0.97834
R2 248 ± 10 0.99778 214 ± 8 0.99778 201 ± 17 0.99072
R3 270 ± 11 0.99812 265 ± 16 0.99812 187 ± 11 0.99688
D1 406 ± 25 0.99044 395 ± 10 0.99044 320 ± 7 0.99231
D2 450 ± 19 0.99517 444 ± 13 0.99517 335 ± 20 0.99494
D3 520 ± 10 0.99123 487 ± 19 0.99123 342 ± 28 0.99120
D4 480 ± 14 0.99543 436 ± 23 0.99543 383 ± 9 0.99394
F1 317 ± 11 0.99682 312 ± 17 0.99682 222 ± 11 0.98461
F2 497 ± 18 0.97645 457 ± 22 0.97645 313 ± 19 0.95565
F3 500 ± 20 0.95345 489 ± 25 0.95345 376 ± 16 0.93568
1368 A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369

Experimental mechanism of thermal degradation of plastics. The pyrolysis reac-


P2
2,0 tion models of HDPE and LDPE can be described by ‘‘Contracting
P3
P4 Sphere” model, whereas that of PP by ‘‘Contracting Cylinder”
1,6 R2 model.
R3
Z(x)/Z(0.5)

1,2 F1
F2 References
F3
0,8 A2
A3 [1] Marongiu A, Faravelli T, Bozzano G, Dente M, Ranzi E. Thermal degradation of
0,4 A4 poly(vinyl chloride). J Anal Appl Pyrolysis 2003;70:519–53.
D1 [2] Kiran N, Ekinci E, Snape CE. Recyling of plastic wastes via pyrolysis. Resour
D2 Conserv Recycl 2000;29:273–83.
0,0
0,0 0,2 0,4 0,6 0,8 1,0 D3 [3] Ceamanos J, Mastral JF, Millera A, Aldea ME. Kinetics of pyrolysis of high
D4 density polyethylene. Comparison of isothermal and dynamic experiments. J
Conversion x Anal Appl Pyrolysis 2002;63:93–110.
[4] Bockhorn H, Hornung A, Hornung U, Schawaller D. Kinetic study on the
Fig. 7. Masterplots of different kinetic models and experimental data at 10 K/min thermal degradation of polypropylene and polyethylene. J Anal Appl Pyrolysis
calculated by Eq. (13) for HDPE thermal degradation. 1999;48:93–109.
[5] Martin-Gullon I, Esperanza M, Font R. Kinetic model for the pyrolysis and
combustion of poly-(ethylene terephthalate) (PET). J Anal Appl Pyrolysis
2001;58–59:635–50.
Experimental [6] Westerhout RWJ, Waanders J, Kuipers JAM, Van Swaaij WPM. Kinetics of the
P2
2,0 low-temperature pyrolysis of polyethene, polypropene, and polystyrene
P3
modeling, experimental determination, and comparison with literature
P4
1,6 models and data. Ind Eng Chem Res 1997;36:1955–64.
R2
[7] Yang J, Miranda R, Roy C. Using the DTG curve fitting method to determine the
R3
Z(x)/Z(0.5)

apparent kinetic parameters of thermal decomposition of polymers. Polym


1,2 F1
F2 Degrad Stabil 2001;73:455–61.
F3 [8] Gao Z, Amasaki I, Nakada M. A thermogravimetric study on thermal
0,8 degradation of polyethylene. J Anal Appl Pyrolysis 2003;67:1–9.
A2
A3 [9] Gao Z, Kaneko T, Amasaki I, Nakada M. A kinetic study of thermal degradation
0,4 A4 of polypropylene. Polym Degrad Stabil 2003;80:269–74.
D1 [10] Park JW, Oh SC, Lee HP, Kim HT, Yoo KO. A kinetic analysis of thermal
0,0 D2 degradation of polymers using a dynamic method. Polym Degrad Stabil
0,0 0,2 0,4 0,6 0,8 1,0 D3 2000;67:535–40.
Conversion x D4 [11] Kim S, Kavitha D, Yu TU, Jung J, Song J, Lee S, et al. Using isothermal kinetic
results to estimate kinetic triplet of pyrolysis reaction of polypropylene. J Anal
Appl Pyrolysis 2008;81:100–5.
Fig. 8. Masterplots of different kinetic models and experimental data at 10 K/min
[12] Kim S, Kim Y. Using isothermal kinetic results to estimate the kinetic triplet of
calculated by Eq. (13) for LDPE thermal degradation. the pyrolysis of high density polyethylene. J Anal Appl Pyrolysis
2005;73:117–21.
[13] Vyazovkin S, Wight CA. Model-free and model-fitting approaches to kinetic
Experimental analysis of isothermal and nonisothermal data. Thermochim. Acta 1999;340–
P2 341:53–68.
2,0 P3 [14] Sewry JD, Brown ME. ‘‘Model-free” kinetic analysis? Thermochim Acta
P4 2002;390:217–25.
1,6 R2 [15] Kim S, Kim Y, Jang E. Identification of reaction model of pyrolysis reaction of
R3 HDPE. Chem Lett 2004;33:1310–1.
Z(x)/Z(0.5)

F1 [16] Kim S, Kavitha D. Identification of pyrolysis reaction model of linear low


1,2
F2 density polyethylene (LLDPE). Chem Lett 2006;35:446–7.
F3 [17] Kim S, Kim Y, Kim YM, Eom Y. Identification of pyrolysis reaction model of
0,8 A2 polypropylene. Chem Lett 2005;34:1268–9.
A3 [18] Aboulkas A, El Harfi K, El Bouadili A. Non-isothermal kinetic studies on co-
0,4 A4 processing of olive residue and polypropylene. Energy Convers Manage
D1 2008;49:3666–71.
D2 [19] Zhou L, Luo T, Huang Q. Co-pyrolysis characteristics and kinetics of coal and
0,0 D3
0,0 0,2 0,4 0,6 0,8 1,0 plastic blends. Energy Convers Manage 2009;50:705–10.
D4 [20] Zhaosheng Y, Xiaoqian M, Ao L. Thermogravimetric analysis of rice and wheat
Conversion x straw catalytic combustion in air- and oxygen-enriched atmospheres. Energy
Convers Manage 2009;50:561–6.
Fig. 9. Masterplots of different kinetic models and experimental data at 10 K/min [21] Rapport. Association marocaine des polymères et leurs applications; 2006.
calculated by Eq. (13) for PP thermal degradation. [22] Friedman H. Kinetics of thermal degradation of char-forming plastics from
thermogravimetry. Application to a phenolic plastic. J Polym Sci Part C
1964;6:183–95.
[23] Kissinger HE. Reaction kinetics in differential thermal analysis. Anal Chem
1957;29:1702–6.
5. Conclusions [24] Akahira T, Sunose T. Trans joint convention of Four Electrical Institutes, paper
no. 246, 1969 research report. Chiba Institute of Technology Sci Technol, vol.
The kinetics of the thermal degradation of HDPE, LDPE and PP 16; 1971. p. 22–31.
[25] Coats AW, Redfern J. Kinetic parameters from thermogravimetric data. Nature
was accurately determined from a series of experiments at four 1964;201:68–9.
heating rates (2, 10, 20 and 50 K/min). The activation energy was [26] Flynn J, Wall LA. Quick direct method for the determination of activation
calculated by the isoconversional methods (Friedman, Kissinger– energy from thermogravimetric data. Polym Lett 1966;4:323–8.
[27] Ozawa T. A new method of analyzing thermogravimetric data. B Chem Soc Jpn
Akahira–Sunose, Flynn–Wall–Ozawa) without previous assump-
1965;38:1881–6.
tion regarding the conversion model fulfilled by the reaction. The [28] Doyle C. Kinetic analysis of thermogravimetric data. J Appl Polym Sci
activation energy was found practically constant in the 0.1–0.9 1961;5:285–92.
[29] Criado JM. Kinetic analysis of DTG data from master curves. Termochim Acta
conversion range, this suggesting that the pyrolysis was a single-
1978;24:186–9.
step process with an activation energy of 238–247 kJ/mol for [30] Wu CH, Chan CY g, Hor JL, Shih SM, Chen LW, Chang FW. On the thermal
HDPE, 215–221 kJ/mol for LDPE and 179–188 kJ/mol for PP. The treatment of plastic mixtures of MSW: pyrolysis kinetics. Waste Manage
corresponding kinetic parameters were calculated to well interpret 1993;13:221–35.
[31] Conesa JA, Marcilla A, Font R, Caballero JA. Thermogravimetric studies on
the relationship between the plastics. Finally, Coats–Redfern and the thermal decomposition of polyethylene. J Anal Appl Pyrolysis
Criado methods were successfully utilized to predict the reaction 1996;36:1–15.
A. Aboulkas et al. / Energy Conversion and Management 51 (2010) 1363–1369 1369

[32] Vyazovkin S. Modification of the integral isoconversional method to account [34] Encinar JM, González JF. Pyrolysis of synthetic polymers and plastic wastes.
for variation in the activation energy. J Comput Chem 2001;22:178–83. Kinetic study. Fuel Process Technol 2008;89:678–86.
[33] Paik P, Kar KK. Kinetics of thermal degradation and estimation of lifetime for [35] Yang KK, Wang XL, Wang YZ, Wu B, Jin YD, Yang B. Kinetics of thermal
polypropylene particles: effects of particle size. Polym Degrad Stabil degradation and thermal oxidative degradation of poly(p-dioxanone). Eur
2008;93:24–35. Polym J 2003;39:1567–74.

You might also like