Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of Analytical and Applied Pyrolysis

45 (1998) 153 – 169

Kinetic study of scrap tyre pyrolysis and combustion


D.Y.C. Leung *, C.L. Wang
Uni6ersity of Hong Kong, Department of Mechanical Engineering, 7 /F Haking Wong Building,
Pokfulam Road, Hong Kong, Hong Kong
Received 12 March 1997; accepted 16 February 1998

Abstract

This paper investigates the kinetic of pyrolysis and combustion of scrap tyre using
thermogravimetric and derivative thermogravimetric analysis method. Three materials,
namely tyre rubber powder, tyre fiber and wood powder were studied and compared with
each other. The process parameters show that these three materials exhibit different thermal
degradation patterns during pyrolysis and combustion process. Thermal degradation models
were proposed to derive the kinetic parameters. It was found that the process and kinetic
parameters vary with heating rates but are less dependent on the powder sizes. The
simulations by the proposed models agreed well with experimental data. © 1998 Elsevier
Science B.V. All rights reserved.

Keywords: Pyrolysis; Combustion; Tyre powder; Tyre fiber; Wood powder; Char; Kinetic
parameters; TGA; DTG; TG

1. Introduction

The disposal of used automotive tyres has caused many environmental and
economical problems to most countries. In the US, 750 million to 2 billion used
tyres have been stockpiled, which are increasing at a rate of 280 million per year [1].
Also, 15 million tons of tyre are scraped every year in the European Union, 2.5 in
North America, 2.4 in UK and 0.5 in Japan [2]. Most of the scrap tyres are dumped
in open or landfill sites. As known, tyre is made of rubbery materials in the form
of Cx Hy with some fibrous materials. It has high volatile and fixed carbon contents
with heating value greater than that of coal. This makes it a good material for

* Corresponding author. Tel.: + 852 2859 7911; fax: + 852 2858 5415; e-mail: ycleung@hkucc.hku.hk

0165-2370/98/$19.00 © 1998 Elsevier Science B.V. All rights reserved.


PII S0165-2370(98)00065-5
154 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

pyrolysis and combustion [3]. On the other hand, scrap tyre is bulky and does not
degrade in landfills. Therefore, open dumping of scrap tyre not only occupies a
large space, presents an eyesore, causes potential health and environmental haz-
ards, but also illustrates wastage of valuable energy resource. How to recover the
energy efficiently from tyre is an acute and imperative problem to energy re-
searchers.
Pyrolysis, incineration, and gasification processes are considered to be more
attractive and practicable methods for recovering energy from scrap tyre and
biomass. Pyrolysis of carbonaceous materials can be interpreted as incomplete
thermal degradation, generally in the absence of air, resulting in char, condensable
liquids or tars, and trace amount of gaseous products. Gasification refers to
pyrolysis followed by higher temperature reactions of the char, tars, and primary
gases to yield mainly low molecular weight gaseous products.
Several studies have been conducted to investigate the pyrolysis of scrap tyre
and biomass in both laboratory and industrial scale. Boukadir et al. [4] found
that the mechanism of rubber degradation is a two-step reaction with reaction
order 11.5 for the first step and 3 for the second under isothermal conditions.
However, they did not find a correct activation energy. Bouvier et al. [5] reported
that rubber degradation is a one-step mechanism and proposed to be a first order
reaction. The apparent activation energy and frequency factor were found to be
125.5 kJ mol − 1 and 1.08 ×109 min − 1 using TGA. Kim et al. [6] proposed that
each constituent of tyre might undergo an irreversible first-order degradation
independently. Chen and Yeh [7] investigated the Styrene-Butadiene Rubber using
nitrogen as purge gas with different oxygen contents. They obtained apparent
activation energy of 211 and 153 kJ mol − 1, frequency factor of 1.32×1014 and
5.75 × 108 min − 1, and reaction order of 0.6 and 0.48 when the purge gas contains
0 and 20% oxygen, respectively. Xu and Wu [8] put forward a three-stage
mechanism for the isothermal pyrolysis of wood powder and found that the
reaction order, the apparent activation energy, and the frequency factor are
0.6–0.78, 8.5 – 39.8 kcal mol − 1 and 40.8–2.66× 109 min − 1, respectively at a
temperature range of 710 – 900°C. Urvan and Antal [9] studied the pyrolysis of
sewage sludge, which produced volatile gas and solid residue. They modelled their
data with two competitive reactions at heating rates from 0.03–1.92°C min − 1.
High reaction orders (10 and 15, respectively) were found. Furthermore, it is
noted that their model was not able to fit the data for different heating rates
adequately. Ralf et al. [10] also studied the pyrolysis of sewage sludge by TGA
and found lower reaction order (2 and 4, respectively) using a four-stage estima-
tion. Other researchers have paid much attention to the weight loss characteristics
during the pyrolysis of tyre or biomass [11–13] while the effects of powder size
have not been studied in details. This paper aims to study the kinetic of the
pyrolysis of different tyre powders and compare with that of wood powder, a
good biomass material. The kinetic of the combustion of tyre rubber char and
tyre fiber char was also studied.
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 155

2. Experimental details

The physical, chemical and thermal characteristics of the samples are essential
for the understanding of the sample behaviour in the pyrolysis process. The tyre
powder and tyre fiber used in this study was supplied by the Recycling Plant of
Guang Zhou in China while the wood powder by San Ya Timber Mill in China.
The tyre powder sample was produced from granulating scrap tyre into four
different sizes: 1.18 – 2.36 mm (8 – 16 mesh), 1.0–1.18 mm (16 mesh), 0.5–0.6 mm
(30 mesh) and 0.355 – 0.425 mm (40 mesh). The type fiber is extracted from the
scrap tyre during the production of the tyre powder. The size of wood powder
used is 0.075 – 0.088 mm (180 mesh). The proximate analysis of the tyre powder,
tyre fiber and wood powder is shown in Table 1 which indicates that their
compositions are quite different from one to another. The samples of tyre rubber
char and tyre fiber char were prepared by heating the tyre rubber and tyre fiber
in a stream of nitrogen gas with a temperature range of 20–900°C at a heating
rate of 5°C min − 1 and soaking for 1 h at 900°C in a Stanton Redcroft STA1500
Simultaneous Thermal Analyzer.
The pyrolysis experiments of various samples and the combustion of tyre
rubber char and tyre fiber char were conducted using the above mentioned
thermal analyzer. About 8 mg of sample was placed in a platinum pan and
heated in an inert atmosphere of nitrogen gas over a temperature range of
20–600°C at controlled heating rates of 10, 30, 45 and 60°C min − 1 for pyrolysis.
In the case of combustion test, it was conducted in a dry air atmosphere over a
temperature range 20 – 900°C at a heating rate of 10°C min − 1. The choice of the
initial sample weight is based on the optimum kinetic rate controlled conditions.
The sample weight loss percent (SWLP), sample temperature and heating effect
were continuously recorded as a function of heating time. The heating effect was
taken into account when recording sample temperature. So the SWLP can be a
function of both sample temperature and heating time. From the SWLP, the
normalized weight loss ratio (a) of sample can be obtained. The normalized
weight loss rate (NWLR) of sample can be obtained by differentiating (8 ) with
respect to time, which is a function of both sample temperature and heating
time.

Table 1
Composition of scrap tyre powder, tyre fiber and wood powder

Proximate analysis (%) Tyre rubber Tyre fiber Wood powder

Volatile 64.2 80.0 68.1


Fixed C 27.8 5.2 22.1
Ash 7.0 12.3 0.8
Moisture 1.0 2.5 9.0
156 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

3. Kinetic interpretation of experimental data

The pyrolysis and combustion reactions could be represented by the following


solid degradation equation derived by Vachuska and Voboril [14]:
n n
daT dai
=% = % Ki (1 − ai ) (1)
dt i = 1 dt i=1

Since the degradation can be activated at a temperature lower than 800°C, which is
within the temperature range of the kinetic reaction, the rate constant Ki can be
obtained from Arrehenius’ Law:

Ki =Ai exp ( −Ei /RT) (2)

Using Eqs. (1) and (2), the kinetic parameters could be obtained from the
thermogravimetry (TG) and derivative thermogravimetry (DTG) curves obtained
through experiment. At the same time, the effect of reaction heat was introduced,
which deviates from the sample heating rate by a constant. Thus the sample
temperature changes with the actual heating rate.
Substituting Eq. (2) into Eq. (1) and taking natural logarithm yields

ln
daT
= ln %
n
dai n
E 
=ln % Ai exp − i (1−ai )
 n (3)
dt i = 1 dt i=1 RT
If there are two reactions occurring at different temperature regions, which
corresponds to two compositional components degradation, the kinetic descriptions
can be shown as:

ln
daT
= ln
 
da1 da2
+

    
dt dt dt

=ln A1exp −
E1 E
(1 − a1)+ A2exp − 2 (1−a2)
n (4)
RT RT
If only one reaction happens in a specific region, say reaction 1, then da2/dt = 0,
a2 =0 and a1 =aT. Eq. (4) can thus be simplified to

ln
 da1 n E
/(1 − a1) =ln A1 − 1 (5)
dt RT
Eq. (5) gives a straight line of slope E/R and a Y intercept of ln A when the left
hand side of the equation is plotted against 1/T. The value of a, da/dt and T at any
times could be obtained from the experimental TG and DTG curves. Therefore, the
kinetic parameters, i.e., the apparent activation energy E and the frequency factor
A, could be determined from the above method.
After that, the value of a2 and da2/dt could be obtained by subtracting a1 from
aT and da1/dt from daT/dt. Using the same method given above, the kinetic
parameters of reaction 2 can be obtained.
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 157

4. Theoretical prediction of a and da/dt

If the kinetic parameters are known in addition to the temperature range and
heating rate b, the normalized weight loss ratio a could be calculated theoretically

& &  
by integrating Eq. (1) for a temperature range of 0 to T (K) as follows:
a T
da A E
= exp − dT (6)
0 (1 − a) b 0 RT
The right-hand side of the above equation has no exact integral, but by using the

&    
relation of Coats and Redfern [15], it can be approximated as:
T
A E A RT 2 E (− 2)k
exp − dT = exp − %
b 0 RT b E RT k = 0 (E/RT)
k

Hence, a can be expressed as:

a = 1 −exp −
 A RT 2
exp −
E 
%
(− 2)k n (7)
b E RT k = 0 (E/RT)
k

From the above equations, the pyrolysis process parameters, such as a and da/dt,
could be predicted theoretically.

5. Results and discussion

5.1. Tyre rubber pyrolysis and tyre rubber char combustion

Fig. 1a and Fig. 2 show the typical DTG curves for the pyrolysis of tyre rubber
of size 40 mesh at various heating rates and of various mesh sizes at a heating rate
of 10°C min − 1, respectively. Table 2 shows the process parameters of tyre rubber
pyrolysis according to different degradation temperature regions. It is interesting to
note that the DTG curve exhibits three different NWLR regions over a temperature
range of 150 – 600°C. These characteristics may be due to the fact that the main
constituents of tyre used in the present study are either nature rubber (NR), styrene
butadiene rubber (SBR), butadiene rubber (BR) or their combination with mois-
ture, oil, plasticizer and additives as minor constituents. All these constituents lose
their weight at different rates and at different temperatures. According to the
evaporating characteristics of individual constituents, it can be deduced that the
moisture inside the type powder evaporates before the temperature reached 150°C.
At the temperature range of 150 – 350°C, the oil, plasticizer and additives are lost.
The loss of NR, SBR and BR or their combination at the temperature range of
340–550°C gives two peaks in the NWLR curve at about 380 and 450°C.
Liu et al. [16] investigated the pyrolysis of NR, BR and SBR and found that the
maximum weight loss rate of NR occurs at a temperature of 373°C, BR at 372°C
and 460°C, SBR at 372°C and 429 – 460°C. Their results indicated that the three
rubber materials contribute their weight loss together over the temperature range of
370 to 460°C, which matches with the present result.
158

Table 2
Analyzed results of tyre powder, tyre fiber and wood powder pyrolysis

Sample name Pyrolysis process parameters

Heating rate Sample size Starting Temp. at Temp. at Finishing Reaction NWLR at NWLR at Weight loss Weight loss Total weight
(°C min−1) (mesh no.) temp. peak one peak two temp. time peak one peak two at peak one at peak two loss
(°C) (°C) (°C) (°C) (min) (min−1) (min−1) (%) (%) (%)

Tyre powder 10 816 185 383 450 500 29.9 0.1076 0.0621 28.15 56.91 68.71
16 185 380 458 500 29.9 0.0932 0.0742 22.84 55.39 65.84
30 185 379 458 500 29.9 0.1020 0.0668 23.36 56.69 65.26
40 185 381 450 500 29.9 0.1078 0.0606 25.78 53.34 63.74
30 816 250 400 480 520 8.5 0.3045 0.1949 20.51 53.92 64.40
16 250 404 479 520 8.5 0.3141 0.1853 24.16 56.39 66.73
30 250 401 478 520 8.5 0.3107 0.1819 21.86 53.92 63.71
40 250 404 471 520 8.5 0.3299 0.1700 22.92 50.17 61.41
45 816 260 410 — 545 5.5 0.5873 — 25.21 — 66.45
16 260 411 491 545 5.5 0.4882 0.2968 20.12 54.63 65.22
30 260 409 485 545 5.5 0.4942 0.2736 19.42 52.23 63.62
40 260 414 484 525 5.5 0.5216 0.2487 23.10 51.58 61.73
60 816 260 410 492 545 4.2 0.8185 0.4125 17.20 52.41 64.15
16 260 410 485 545 4.2 0.7300 0.3744 19.49 51.57 63.13
30 260 416 490 545 4.2 0.7729 0.3485 22.25 53.32 63.42
40 260 413 484 545 4.2 0.7804 0.3313 20.68 50.16 60.52
Tyre fiber 10 295 412 — 510 20.3 0.0167 — 42.10 — 82.59
30 — 310 445 — 525 6.7 0.0169 — 42.49 — 83.95
45 300 447 — 535 4.8 0.0152 — 34.76 — 80.00
60 300 450 — 535 3.5 0.0219 — 34.58 — 81.88
Wood pow- 10 180 347 69a 400 21.5 0.0089 0.0018a 50.28 4.81a 76.51
der
30 — 200 368 84a 420 6.8 0.0082 0.0017a 50.77 4.29a 74.40
45 200 369 88a 420 4.3 0.0080 0.0019a 50.01 4.34a 74.20
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

60 220 381 92a 440 3.2 0.0084 0.0015a 49.55 3.94a 72.29
a
Representing the date of water moisture separated out from the wood powder
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 159

Fig. 1. Normalized weight loss rate against temperature for pyrolysis of different materials at different
heating rates. (a) Tyre powder; (b) tyre fiber; and (c) wood powder.
160 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

Fig. 1. (Continued)

Yang et al. [17] also studied the pyrolysis behaviour of these three rubbers and
their mixture. The curves they reported at a heating rate of 10°C min − 1 showed
that two peaks of the NWLR are observed for each rubber. The first peak, which
has a maximum weight loss rate at 245°C for NR and BR, and 267°C for SBR, is
attributed to the volatilization of the processing oil, plasticizer, moisture, and any
other low boiling point components. The second peak, which has a maximum
weight loss rate at 377°C for NR, 465°C for BR, and 444°C for SBR, is attributed
to the three straight elastomers thermal decomposition. In addition, they obtained
several DTG curves for the pyrolysis of the mixture with different compositions,
which are similar to those obtained presently.
In addition, Brazier and Nickel [18] also reported a similar DTG curve of tyre
rubber pyrolysis, consisting of only two main components of NR and BR.
Kim et al. [6] investigated the pyrolysis process of sidewall and tread rubber
derived from waste tyre at different heating rates using TGA. Their DTG curves
are different from the present curves, which may be due to the difference in the
sample composition. Nevertheless, they also divided the degradation process into
three regions with different temperature ranges in explaining the pyrolysis process.
It is well understood that many complex reactions are involved in the pyrolysis
process of tyre rubber. Therefore, it is impossible to develop a precise kinetic model
for determining various kinetic parameters from thermogravimetric data alone. It is
generally accepted that the important parameters in pyrolysis are temperature,
sample weight and its loss rate, time, and heating rates. Based on these parameters
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 161

Fig. 2. Normalized weight loss rate against temperature for pyrolysis of different sizes of tyre rubber at
a heating rate of 10°C min − 1.

a model is proposed to predict the weight and weight loss rate profiles of tyre
rubber. The model assumes two reactions in the tyre rubber pyrolysis, correspond-
ing to the main components (NR, BR and SBR) degradation. Each component
contributes to the decomposition at different temperature regions forming volatiles
and char. The pyrolysis rate is considered to be the sum of the two reaction rates.
A schematic diagram of the model is shown below:

It is further assumed that reactions 1 and 2 follow the Arrehenius’ Law for
ascertaining the values of the rate constant and undergo an irreversible first order
degradation independently. Furthermore, reaction 1 mainly occurs at a lower
temperature region than reaction 2 at a higher temperature region and both
reactions occur at the intermediate temperature region.
Fig. 4 shows the typical DTG curves of tyre rubber char of different mesh sizes
at a heating rate of 10°C min − 1. It can be found that the char degrades in one
temperature region only. Using the same approach as mentioned above to investi-
gate the behaviour of tyre rubber char combustion, and assuming that only one
162

Table 3
Kinetic parameters of the pyrolysis of tyre powder, tyre fiber and wood powder

Sample Heating rate Lower temperature stage Higher temperature stage


(°C min−1)

Temperature Apparent activa- Frequency factor Temperature Apparent activa- Frequency factor
range (°C) tion energy (min−1) range (°C) tion energy (min−1)
(kJ mol−1) (kJ mol−1)

Tyre powder 10 300420 164.5 6.29×1013 350500 136.1 2.31×109


(40 mesh)
30 310440 180.9 1.32×1014 370510 133.6 2.09×109
45 320470 203.4 7.58×1015 400540 107.0 3.34×107
60 320480 218.7 1.13×1017 410540 99.1 1.02×107
Wood powder 10 150340 70.0 1.39×106 240400 112.7 1.23×109
30 150350 82.8 7.46×108 270410 119.1 6.50×109
45 150350 102.7 4.47×109 270410 134.8 1.53×1011
60 160360 104.9 5.75×109 280420 141.4 5.56×1011
Tyre fiber 10 320500 152.0 4.81×1011
30 340520 160.7 6.91×1011
45 340520 163.8 6.91×1011
60 350520 201.1 5.11×1014
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 163

reaction occurs during char combustion, we propose the following model to


describe the combustion process:
K1
Tyre Rubber Char “ Ash + Gases
In this model, the reaction is treated as one component undergoing an irre-
versible first order degradation, which follows the Arrehenius’ Law for ascertaining
the values of the rate constant.
According to the above model, the kinetic parameters of tyre rubber pyrolysis
and tyre rubber char combustion are obtained and shown in Tables 3 and 4. Yang
et al. [15] reported the kinetic parameters of NR, BR and SBR pyrolysis at a
heating rate of 10°C min − 1. The activation energy is 207, 215 and 152 kJ mol − 1
and the frequency factor is 2.36 × 1016, 6.32 × 1014 and 4.15 × 1010 min − 1 for NR,
BR and SBR, respectively, which are slightly higher than those obtained from our
model (Table 3). This may be due to the fact that the kinetic parameters obtained
from our model are the combined effect of these three rubbers.
In addition, Kim et al. [6] reported the kinetic parameters of sidewall and tread
rubber pyrolysis according to three temperature regions. For sidewall pyrolysis, the
average apparent activation energy and frequency factor were 204 kJ mol − 1 and
2.04 × 1014 min − 1 for components decomposed at higher temperature, 195 kJ
mol − 1 and 2.08 ×1015 min − 1 for components decomposed at medium temperature,
42 kJ mol − 1 and 1436 min − 1 for components decomposed at lower temperature.
The average apparent activation energy and frequency factor were 127, 209, 39 kJ
mol − 1 and 8.75 ×108, 3.78 × 1016 and 934.5 min − 1 for components of tread rubber
decomposed at higher, medium and lower temperature. These values are only
slightly different from the present result.
The simulation curves are shown in Fig. 3a and Fig. 4. Scrutinizing the results
obtained, it can be observed that the total weight loss during the pyrolysis process
is  60– 69%, among them only  7% is due to moisture, oil, plasticizer and
additives,  44 – 53% and 40 – 49% is lost in reactions 1 and 2, respectively.
During the tyre rubber char combustion process, the total weight loss is 84–93%.
Furthermore, the following points were noted after analyzing the results:
“ The sizes of tyre powder have little effect on the pyrolysis and combustion
process;
“ The heating rate affects the pyrolysis significantly;

Table 4
Kinetic parameters of the combustion of tyre rubber char and tyre fiber char

Sample name Sample size Temperature Apparent activation energy Frequency factor
(mesh) range (°C) (kJ mol−1) (min−1)

Tyre powder 8–16 450610 145.4 2.89×108


16 450620 148.1 3.84×108
40 450620 161.2 3.93×109
Tyre fiber 480570 237.3 9.43×1014
164 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

Fig. 3. Theoretical and experimental normalized weight loss rate against temperature for the pyrolysis of
different materials at a heating rate of 10°C min − 1. (a) Tyre powder; (b) tyre fiber; and (c) wood
powder.
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 165

Fig. 3. (Continued)

“ As the heating rate rises, (a) the reaction shifts to higher temperature range, i.e.
the starting and ending temperature increase; (b) the temperature corresponding
to the peak value of NWLR increases; (c) the NWLR increases and the reaction
time decreases dramatically.
“ During pyrolysis the apparent activation energy and frequency factor of reaction
1 increase with heating rate, indicating that the degradation is more difficult. In
contrast, the apparent activation energy and frequency factor of pyrolysis
reaction 2 decrease, making it easier to degrade.
“ The apparent activation energy and frequency factor of combustion of tyre
rubber char increases as the size of char increases. However, the difference is not
significant.
“ The pyrolysis of tyre rubber and the combustion of its char occurs at tempera-
ture ranges of about 150 – 500°C and 460–620°C, respectively. There is some
overlap in the temperature range, so it can be assumed that these two reactions
follow smoothly.

5.2. Tyre fiber pyrolysis and tyre fiber char combustion

Fig. 1b and Fig. 3b show the typical DTG curves for the tyre fiber pyrolysis at
various heating rates and tyre fiber char combustion at a heating rate of 10°C
min − 1, respectively. Table 2 shows the detailed process parameters of tyre fiber
pyrolysis. It can be found that the degradation occurs at the temperature region of
350–520°C for pyrolysis and 500 – 570°C for combustion. This behaviour is similar
166 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

Fig. 4. Comparison of theoretical and experimental normalized weight loss rate against temperature for
the combustion of tyre rubber char.

to that of tyre rubber char combustion, so the same model as mentioned in tyre
rubber char combustion is proposed. Using this model, the kinetic parameters are
obtained and shown in Tables 3 and 4. The simulation curves are shown in Fig. 3b.
It can be found that the model simulation fits the experimental data well.
The following points were observed from the experimental results.
“ The heating rate affects the pyrolysis significantly;
“ During pyrolysis, as the heating rate rises, (a) the reaction shifts to higher
temperature range; (b) the temperature corresponding to peak value of NWLR
increases; (c) the NWLR increases and the reaction time decreases dramatically;
(d) the apparent activation energy and frequency factor increase, which leads to
a more difficult degradation.
“ The total weight loss is  80 – 84% for tyre fiber pyrolysis and  71% for tyre
fiber char combustion.
“ The pyrolysis of tyre fiber and the combustion of its char occurs at different
temperature ranges which overlap slightly so it can be assumed that these two
reactions are carried out continuously. For the same time period, about 82% of
total weight loss occurred in pyrolysis and only about 12% in combustion.
Furthermore, the apparent activation energy and frequency factor is larger in
combustion than in pyrolysis process. This indicates that attention must be paid
to tyre fiber char combustion.
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 167

5.3. Wood powder pyrolysis

Fig. 1c shows typical DTG curves for wood powder pyrolysis at various
heating rates. Table 2 shows the detailed process parameters according to differ-
ent degradation temperature regions. Comparing the TG and DTG curves with
that of tyre rubber in Fig. 1a, it was found that the curve also shows three
different NWLR regions over a temperature range of 20–420°C. It is well known
that wood is consisted of three main constituents: moisture, lignin and holocellu-
lose (cellulose and hemicellulose). The characteristics of wood degradation at
different temperature regions are made up of these three components. For the
wood powder pyrolysis at a heating rate of 10°C min − 1, the moisture inside the
wood powder evaporates at the temperature range of 20–150°C. At the tempera-
ture range of 200 – 420°C, the loses of lignin and holocellulose give two peaks in
the NWLR curve at about 300°C and 340°C. Also, it can be found that the
temperature corresponding to the peaks value of NWLR are affected by heating
rates.
There are several papers investigating the pyrolysis of cellulose, biomass and
wood powder [8,19 – 21] and different models were proposed to simulate the
pyrolysis process.
From the present results, we used a similar model as that used in tyre rubber
pyrolysis to explain the wood powder pyrolysis which is given below:

From the above model, the kinetic parameters and simulation curves are ob-
tained and shown in Table 3 and Fig. 3c, respectively. It can be seen that for the
total weight loss, only 9% is due to moisture, and 17–22%, 69–74% is due to
reaction 1 and reaction 2, respectively. In addition, the apparent activation energy
and frequency factor for moisture evaporation are 31, 47 kJ mol − 1 and 2.02×
104, 5.99 × 106 min − 1 at the heating rate of 10 and 60°C min − 1, respectively. In
general, as the heating rate increases, the apparent activation energy and fre-
quency factor increase significantly.
It could be found that the kinetic parameters obtained match well with those of
Xu and Wu [8] despite the fact that the tests were carried out under different
conditions.
It should be emphasized that although it is difficult to compare our results with
those in literature directly due to the different tyre samples and experimental
conditions, the good matching of the model simulation with experimental data
indicates that the models can be used to describe the pyrolysis and combustion
process of the tyre rubber and tyre fibre.
168 D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169

6. Conclusions

The thermogravimetric and derivative thermogravimetric analysis conducted in


this study provided valuable information on the pyrolysis kinetics and mechanisms
of heterogeneous materials like tyre rubber powder, tyre fiber and wood powder.
The pyrolysis of tyre rubber exhibits three obvious weight loss regions occurring at
temperature ranges of about 150 – 350°C, 330–450°C and 420–520°C, respectively
while the combustion of tyre rubber char has one obvious weight loss region
occurring at a temperature range of 450–620°C. The tyre fiber pyrolysis and its char
combustion have one obvious weight loss region, which occurs at a temperature
range of about 350 – 520°C and 500 – 570°C, respectively. Similar to the case of tyre
rubber, the pyrolysis of wood powder has three obvious weight loss regions
occurring at temperature ranges of about 20–150°C, 150–360°C and 240–420°C,
respectively. Due to these observations, pyrolysis and combustion models were
proposed to study the kinetic of the processes, which were found to fit the
experimental data well.
It was also found that the heating rate has a significant effect on the pyrolysis and
combustion process. With increasing heating rate the weight loss regions shift to a
higher temperature range and the weight loss rate is increased. The reaction time
shortened quickly but the total weight loss has no obvious change. The apparent
activation energy and frequency factor also increase with increasing in heating rate,
hence increasing the difficulties of the pyrolysis reaction. From the activation energy
obtained, the pyrolysis of tyre rubber was found to be easier than tyre fiber but more
difficult than wood powder. Moreover, the combustion of tyre rubber char is found
to be easier than tyre fiber char. The size of tyre rubber and its char was found to
produce no significant effect on the pyrolysis and combustion process.

Acknowledgements

The authors wish to acknowledge the Honk Kong Research Grant Council and
the CRCG of the University of Hong Kong for supporting this project.

Appendix A. Nomenclature

A frequency factor (min−1)


E apparent activation energy (kJ mol−1)
K rate constant (min−1)
R gas constant, R =8.314×10−3 (kJ mol−1 · K−1)
t reaction time (min)
T sample temperature (K)
W0 sample weight at start time (mg)
W sample weight at time t (mg)
W sample weight at end time (mg)
D.Y.C. Leung, C.L. Wang / J. Anal. Appl. Pyrolysis 45 (1998) 153–169 169

a normalized weight loss ratio, a= (W0−W)/(W0−W )


da/dt normalized weight loss rate (min−1)

References

[1] J.S. Wu, R. Vallabhapuram, Energy Recovery From Scrap Tyres. Hazardous Waste Management
Handbook, Prentice-Hall, Englewood Cliffs, NJ, 1994, pp. 170 – 179.
[2] P.T. Williams, S. Besler, D.T. Taylor, R.P. Bottrill, The pyrolysis of automotive tyre waste. J.
Institute Energy 68 (1995) 11–21.
[3] D.Y.C. Leung, G.C.K. Lam, Prospects for energy recovery from scrap tyres in Hong Kong.
Proceedings of the Third Asian-Pacific International Symposium on Combustion and Energy
Utilization, 1995, pp. 11–15.
[4] D. Boukadir, J.C. David, R. Granger, J. Vergnaud, Preparation of a convenient filler for
thermoplastics by pyrolysis of rubber powder recovered from tyres, J. Anal. Appl. Pyrolysis 3
(1981) 83–89.
[5] J.M. Bouvier, F. Charbel, M. Gelus, Gas-solid pyrolysis of tyre wastes: kinetics and material
balances of batch pyrolysis of used tyres, Resources Conserv. 15 (3) (1987) 205 – 214.
[6] S.D. Kim, J.K. Park, H.D. Chun, Pyrolysis kinetics of scrap tyre rubbers. I: Using DTG and TGA.,
J. Environ. Eng. 121 (1995) 507–514.
[7] K.S. Chen, R.Z. Yen, Kinetics of thermal decomposition of styrene butadiene rubber at low heating
rates in nitrogen and oxygen, Combustion and Flame 180 (1997) 408 – 418.
[8] B.Y. Xu, C.Z. Wu, Kinetic study in biomass gasification, Solar Energy 49 (3) (1992) 199 – 204.
[9] D.L. Urvan, M.J. Antal, Study of the kinetics of sewage sludge pyrolysis using DSC and TGA, Fuel
61 (1982) 799–806.
[10] D. Ralf, R. Werner, R.S. Marc, Kinetic studies of the pyrolysis of sewage sludge by TGA and
comparison with fluidized beds, Can. J. Chem. Eng. 69 (1991) 953 – 963.
[11] F.N.H. Ani, R. Zailani, Liquid fuel from fast pyrolysis of scrap tyres, Proceedings of the
Asia-Pacific conference on sustainable energy and environmental technology, 1996, pp. 487 – 493.
[12] O. Menis, H. Rook, P. Garn, The State-of-the-Art of Thermal Analysis, US Government Printing
Office, Washington, 1980.
[13] D.Y.L. Tzan, C.I. Juch, Fluidized co-combustion of waste tyre and industrial sludge in a circulation
fluidized-bed incinerator. 3rd Asian-Pacific International Symposium on Combustion and Energy
Utilization, 1995, pp. 115–121.
[14] W.W. Wendlandt, (Author), D.D. Chen, (Translator), Thermal Analysis (Chinese) Min 8, Behair
Hall, Taibei, 1992.
[15] A.W. Coats, J.P. Redfern, Kinetic parameters from thermogravimetric data, Nature 201 (1964)
68 – 69.
[16] Z.R. Liu, H.Y. Tang, Y.L. Zheng, Handbook of Rubber Industry, Chemical Industry Publishing
House, China, 1992, pp. 477–478.
[17] J. Yang, S. Kaliaguine, C. Roy, Improved quantitative determination of elastomers in tyre rubber
by kinetic simulation of DTG curves, Rubber Chem. Technol. 66 (1993) 213 – 229.
[18] D.W. Brazier, G.H. Nickel, Thermoanalytical methods in vulcanizate analysis II. Derivative
thermogravimetric analysis, Rubber Chem. Technol. 48 (1975) 661 – 677.
[19] G.W.B. Allan, S. Yoshio, S. Fred, A kinetic model for pyrolysis of Cellulose, J. Appl. Polymer Sci.
23 (1979) 3271–3280.
[20] C.A. Koufopanos, G. Maschio, A. Lucchesi, Kinetic modeling of the pyrolysis of biomass and
biomass components, Can. J. Chem. Eng. 67 (1989) 75 – 84.
[21] V.K. Srivastava, R.K. Jalan, Predictions of concentration in the pyrolysis of biomass materials-I,
Energy Convers. Manage. 12 (1994) 1031– 1040.

You might also like