You are on page 1of 22

Journal of Analytical and Applied Pyrolysis

44 (1998) 131 152

Composition of oils derived from the batch pyrolysis of


tyres
Adrian M. Cunliffe, Paul T. Williams *
Department of Fuel and Energy, The Uni6ersity of Leeds, Leeds LS2 9JT, UK
Received 9 April 1997; accepted 7 October 1997

Abstract
A nitrogen purged static-bed batch reactor was used to pyrolyse 3 kg batches of shredded
scrap tyres at temperatures between 450 and 600C. The oils were trapped in a series of
condensers and the derived gases analysed off-line by packed column gas chromatography.
The oil yield was found to decrease with increasing final pyrolysis temperature and the yield
of product gases increased. The fuel properties of the condensed oil including, calorific
values, ultimate analyses, flash point, moisture content, fluorine and chlorine contents were
determined. The concentration of polycyclic aromatic hydrocarbons (PAH) and lighter
aromatic hydrocarbons were determined. The results showed that the derived tyre oils had
fuel properties similar to those of a light petroleum fuel oil. The influence of pyrolysis
temperature showed an increase in the aromatic content of the oils with increasing temperature, with a consequent decrease in aliphatic content. The total PAH content of the oils were
found to increase from 1.5 to 3.5 wt.% of the total oil as the pyrolysis temperature was
increased from 450 to 600C. Biologically active compounds such as methylfluorenes, tri- and
tetra-methylphenanthrenes and chrysene were identified in significant concentrations. The
results of gas analysis supported a Diels Alder mechanism of alkane dehydrogenation to
alkenes, followed by cyclisation and aromatisation. Limonene was identified as a major
component of the oils, representing 3.1 wt.% at 450C falling to 2.5 wt.% total oil at 600C.
Significant quantities of light aromatics such as benzene, toluene, xylene and styrene were
also found. 1998 Elsevier Science B.V.
Keywords: Pyrolysis; Tyres; Analysis; Polycyclic aromatic hydrocarbons

* Corresponding author.
0165-2370/98/$19.00 1998 Elsevier Science B.V. All rights reserved.
PII S 0 1 6 5 - 2 3 7 0 ( 9 7 ) 0 0 0 8 5 - 5

132

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

1. Introduction
Current estimates of the annual arisings of scrap tyres within North America is
2.5 million tonnes with 2.0 million tonnes in the European Union and 0.5 million
tonnes per year in Japan [1 3]. The majority of scrap tyres are disposed of in open
or landfill sites. However, landfilling of tyres is declining as a disposal option, since
tyres do not degrade easily in landfills, they are bulky, taking up valuable landfill
space and preventing waste compaction. Open dumping may result in accidental
fires with high pollution emissions and tyres can be a breeding ground for insects
and a home for vermin. Alternative waste management options to landfilling and
open dumping have included, tyre retreading, crumbing to produce rubber for
applications such as carpets, sports surfaces and childrens playgrounds. Incineration of tyres with energy recovery is also a growing option since it utilises the high
calorific value of tyres.
A further alternative waste management option is tyre pyrolysis which is currently receiving renewed attention. Pyrolysis has a number of advantages as a
treatment option since the derived oils may be used directly as fuels or added to
petroleum refinery feed stocks, they may also be an important source for refined
chemicals. The derived gases are also useful as fuel and the solid char may be used
either as smokeless fuel, carbon black or activated carbon [414].
The production of a liquid oil product increases the ease of handling, storage and
transport and hence the product does not have to be used at or near the recycling
plant. There have been several proposed uses for the tyre oil, the main uses being
substitution for conventional fuels and as chemical feedstocks. In order for tyre
pyrolysis oils to compete with petroleum derived fuels the behaviour of the fuel in
comparison to such fuels should be determined. Tyre pyrolysis oils have been found
to have a high calorific value of around 4144 MJ kg 1, which would encourage
their use as replacements for conventional liquid fuels [4,6,8]. However, in addition
the oils have been shown to contain polycyclic aromatic hydrocarbons (PAH). The
determination of the PAH present in a fuel has significance in that the fuel PAH
tend towards particulate formation upon combustion [15,16]. Olsen and Pickens
[15] determined the threshold soot index (TSI), a relative measure of the propensity
of a fuel towards soot formation, of a range of hydrocarbons when combusted in
a premixed flame. PAH were shown to have the greatest TSI values, that is, they
were the most likely to produce soot. The unburned fuel PAH can also be deposited
on soot formed from certain combustion systems and enter the atmosphere [1619].
This presents a hazard because a significant number of these compounds are known
to be, or are suspected of being, carcinogenic. It may be seen that the safety aspects
of handling the oils will also be affected by the presence of significant concentrations of carcinogenic PAH.
In addition to their use as fuels, the oils have been shown to be a potential source
of light aromatics such as benzene, toluene and xylene (BTX), which command a
higher market value than the raw oil [8,13,2023]. Similarly, the oils have been
shown to contain limonene, a high value product used in industrial applications
including formulation of industrial solvents, resins and adhesives, as a dispersing

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

133

agent for pigments, as a fragrance in cleaning products and as an environmentally


acceptable solvent [20,23].
In this work a static-bed batch reactor of 3 kg loading was used to produce tyre
pyrolysis oils which were analysed for their properties as potential fuels and their
composition in terms of PAH and potentially high value light aromatic hydrocarbons in relation to pyrolysis temperature.

2. Experimental

2.1. Tyre samples


Shredded used passenger car tyres in narrow strips approximately 3 cm wide
1.5 cm thick and of between 50 and 150 cm maximum dimension retaining both
steel and fabric cords were kept in dry conditions prior to pyrolysis. The proximate
and ultimate analyses, on a steel- and fabric-free basis, and the calorific value of the
raw tyre tread used are given in Table 1.

2.2. Pyrolysis unit


The tyres were pyrolysed in a static-bed batch reactor of inner diameter 24 cm
and depth 36 cm, with a feedstock loading of 3 kg, as shown in Fig. 1. The reactor
is externally electrically heated, with the heating regime being electronically controlled. The heating rate was approximately 5 K min 1. Purge gas distributor rings
are positioned in the base and upper part of the reactor. The purge gas is preheated
before entering the reactor. Gases are exited via a down tube to a water-cooled
condenser of mesh-screen design and passed to a second water-cooled condenser. A
gas sampling point is positioned just after the second condenser allowing gas
samples to be taken via gas-syringes for off-line analysis. Batches of 3 kg of tyres,
including the steel core, were loaded into the pyrolysis unit and the reactor was
purged with nitrogen in order to minimise secondary reactions in the hot zone. The
maximum residence time in the reactor was calculated as approximately 120 s. An
initial heater ramp temperature of 150C was set in each case, followed by final
temperatures of 450, 475, 500, 525, 560 and 600C. Once the temperature had
Table 1
Typical composition of the scrap tyre feedstock rubber
Elemental composition (%)

Proximate analysis (%)

Gross calorific value (MJ kg1)

C (86.4)
H (8.0)
N (0.5)
S (1.7)
O (3.4)
Ash (2.4)

Volatiles (62.2)
Fixed carbon (29.4)
Ash (7.1)
Moisture (1.3)

40.0

134

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

Fig. 1. Schematic diagram of the tyre pyrolysis reactor.

stabilised near to the final set point, the system was held at this temperature for 90
min. Temperatures throughout the reactor were taken at timed intervals. The oils
collected from the condenser system were centrifuged at 3000 rpm for 15 min to
remove any water and sediment present prior to all analyses, except for those
samples used for water content determinations. The oil samples were stored under
refrigerated conditions prior to analysis. The residual char was removed from the
reactor at the end of each experiment, separated into char and steel using magnetic
separation and weighed.
Gas samples were taken by gas syringe at timed intervals and analysed off-line by
packed column gas chromatography for N2, O2, H2, CO, CO2 and hydrocarbons up
to C4. The results were then corrected for N2 and calculated in terms of the outflow
for each gas per unit mass of reactive tyre and the total yield of each gas for each
experiment. The molecular weight and yield of each analysed gas was used to
calculate the total mass of gas produced during pyrolysis.

2.3. Oil analysis


2.3.1. Fuel properties
The pyrolytic oils were analysed for their fuel properties according to standard
Institute of Petroleum (IP) or American Society for Testing and Materials (ASTM)
tests. The gross calorific value and combustible sulphur content of the tyre oils was

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

135

determined by the IP 12 and IP 61 methods equivalent to ASTM 129-64 described


in the standards. The flashpoint of the pyrolysis oils were determined using the
PenskyMartens closed cup apparatus, IP 34, ASTM D93-73 method. As preliminary tests showed that the flashpoint was in the region 1020C, the samples were
cooled in a freezer prior to testing. The water content of the tyre oils was
determined using the standard IP 74, Dean and Stark method. The ash content of
the tyre oils was determined by the IP 4, ASTM D482-74 method. In view of the
high moisture content of the oils as determined by the Dean and Stark assay, the
IP recommended method of avoiding loss of material caused by excess frothing
upon heating was followed; a 10 cm3 equivolume mixture of benzene and propan-2ol was stirred into the oil and four strips of ashless filter paper were placed into the
oil such that 2 3 cm projected above the oil surface. This procedure was found to
effectively prevent excessive frothing from occurring.
A Perkin-Elmer 240C elemental analyser was used to give the percent composition by mass of carbon, hydrogen and nitrogen. The contribution of sulphur and
ash were taken into account when calculating the oxygen content by difference. The
presence of chlorine, and fluorine were determined in the washings from the bomb
calorimetry experiments of the determination of calorific value. The presence and
content of fluoride and chloride anions in the washings was determined using a
Dionex QIC ion chromatograph fitted with a Dionex Ionpac AS4A 4 mm column
and electrical conductivity detector. The column was flushed with deionised water
between samples to minimise contamination by sulphate and nitrate ions, which
were present in high concentrations in the washings.
The pyrolysis oils were analysed by size exclusion chromatography (SEC) to
determine the molecular weight (Mr) range of the tyre oils. The results were used to
compare the molecular weight range of the oils in comparison to the refined
petroleum oils, gas oil (diesel fuel), light fuel oil and medium fuel oil to indicate
their potential as substitute fuels. The system incorporated two 150 4.6 mm
columns with Polymer Laboratories 5 mm RPSEC 100 A type packing. A third
column of the same material was placed in line between the pump and the injection
valve, to ensure pre-saturation of the solvent with the column packing material and
also to avoid analytical column dissolution and hence loss of performance. The
solvent used for the mobile phase was tetrahydrofuran (THF). The calibration
system used was based on polystyrene samples of low polydispersity in the Mr range
of 800860 000 Da, but also included was benzene for low Mr calibration.
Although benzene is not normally used as a calibration standard, the oils contained
many compounds in this Mr range, as do many refined petroleum fuels and it was
felt the results would give a useful comparison of the fuels. Samples were introduced through a 2 ml loop injection valve. The detection system was a UV detector
from Merck Hitachi. Ultra-violet scanning of the polystyrene Mr fractions in THF
indicated that the maximum absorbance was at 262 nm, and all calibration and
efficiency measurements were taken at this wavelength. The Mr distribution was
determined as number and weight average Mr. Details of the SEC system including
optimisation and calibration have been previously reported [29].

136

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

2.3.2. Oil analysis


2.3.2.1. Volatile hydrocarbons. Samples of the whole oils were made up to standard
volume with n-pentane. The resulting solutions were then centrifuged (3000 rpm, 5
min) to separate out high boiling point, high molecular weight, complex aromatic
compounds which remained out of solution. The volatile hydrocarbons were then
quantified by gas chromatography with flame ionisation detection. Standard solutions of the volatile hydrocarbons were used to determine retention times and for
quantification. The system used was a Carlo-Erba Mega Series HRGC 5300 gas
chromatograph, DB5 capillary column, temperature programme of 50C for 10 min
followed by 5 K min 1 to 280C. The carrier gas was helium and the flow rate was
2 ml min 1 at 280C.
2.3.2.2. Polycyclic aromatic hydrocarbons. The oil composition was analysed in
detail for polycyclic aromatic hydrocarbons by liquid chromatography to separate
the oils into chemical class fractions followed by gas chromatography/mass spectrometry and gas chromatography/flame ionisation detection. The liquid chromatography consisted of 101 cm glass columns packed with 5.0 g silica, Bondesil
(sepralyte) sorbent, pre-treated at 105C for 2 h prior to use. To prevent the
formation of a solid phase with the addition of the pentane mobile phase, and to
improve solvent contact with the oil, the oil (0.25 g) was intimately mixed with
Chromosorb G/AW/DMCS 60 80 support and packed in the column above the
analytical phase. The column was then sequentially eluted under vacuum with 10 ml
sequential elutions of pentane, benzene, ethyl acetate and methanol to produce,
aliphatic, aromatic, hetero-atom and polar fractions respectively. Analysis using gas
chromatography/mass spectrometry and Fourier transform infra-red spectroscopy
confirmed the composition of each fraction. Each fraction was evaporated to
dryness, weighed and the percentage mass in each fraction calculated. The evaporation of the solvent would inevitably lead to some loss of volatile material,
consequently, this step in the analytical procedure was carefully carried out to
minimise losses.
The aromatic fractions were analysed by capillary column gas chromatographymass spectroscopy together with relative retention times using relative retention
indices [24 27] to determine identification of PAH and thereby verify the separation of the PAH fraction. The mass spectrometer used was of the bench top ion
trap detector type manufactured by Finnigan-MAT. The ion trap detector had a
mass range from 20 to 650 u with scan times between 0.125 and 2 s and it was
linked to a computer with a library containing 38 752 mass spectra. Single ion
monitoring (SIM) was also carried out to confirm the identification of compounds
and also to examine the samples for a series of substituted compounds, for example,
phenanthrene and its methyl-, dimethyl-, trimethyl- and tetra-methyl derivatives.
The ethylacetate and methanol fractions may contain heterocyclic PAH such as
nitrogen containing PAH and sulphur containing PAH, which were not determined
in this study. Quantification of the PAH was determined by capillary column gas
chromatography with cold on-column injection and a flame ionisation detector with

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

137

internal PAH standards and data analysis computer package. The capillary
columns used in both gas chromatographs were DB5 and the temperature programme was 50C for 2 min followed by a 5C min 1 heating rate to 270C. A
series of sequential dilutions of standard solutions of PAH, for instance fluorene
and methylated naphthalenes, were analysed to produce straight line plots of
response against concentration, which were then used to quantify the PAH species
identified. The results represent the average of two separate quantifications for each
oil.
The analyses were repeated several times and the analytical method developments
included full statistical error analysis which have been previously reported [4,28,29].

3. Results and discussion

3.1. Product yields


The product yields of oil, char and gases in relation to process conditions in the
batch reactor are shown in Table 2. The results corrected for the percentage mass
of steel core, show that the char yield remains fairly constant with a mean of 37.8
wt.%. Similar results have been reported by Kawakami et al. [6] for tyre pyrolysis
using a rotary kiln reactor, with char yields of between 38 and 40 wt.% over a
temperature range of 540 750C. Only a small influence of temperature on the
yield of char from the pyrolysis of tyres was observed by Teng et al. [11]. Pyrolysing
in a thermo-gravimetic analyser, they found the char yield decreased with increasing
temperature, from 40% at 500C to 36% at 900C at a heating rate of 30C min 1
in a helium atmosphere. Other workers have shown significant decreases in char
yield as the pyrolysis temperature was increased. Tyre char has been shown to
consist mainly of carbon, with some solid hydrocarbons and metal oxides [9].
Williams et al. [4] on a smaller scale fixed bed reactor with shorter gas residence
times in the hot zone found a significant decrease in char yield from 56.6 to 40.2
wt.% as the temperature of pyrolysis was increased from 420 to 600C, for the
heating rate of 5 K min 1. A decrease in char yield was also observed for an
increase in heating rate from 20 to 80 K min 1. The decrease of char yield was
attributed to increasing devolatilisation of solid hydrocarbons in the char. AlternaTable 2
Yield of tyre pyrolysis products
Final temperature (C)

Char yield (wt.%)

Oil yield (wt.%)

Gas yield (wt.%)

450
475
500
525
560
600

37.4
37.3
38.3
37.8
38.1
38.0

58.1
58.2
56.2
56.9
55.4
53.1

4.5
4.5
5.5
5.2
6.5
8.9

138

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

tively, there may be partial gasification of the char. Commercial scale, i.e. 1 tonne
batch pyrolysis of tyres has been shown to give higher yields of char ranging from
59.4 wt.% at 700C to 46.8 wt.% at 950C final temperature of pyrolysis [20].
Competing with char loss reactions are carbonisation reactions of the oil. It is
known that long residence times of oil hydrocarbons in the hot zone can result in
char formation [30]. Kaminsky and Sinn [5] using a fluidised bed pyrolysis unit,
found an increase in char yield as the fluidised bed temperature was increased from
640 to 840C. The formation of char from tyre pyrolysis is clearly linked to a
number of factors, not only the temperature and heating rate but also system
specific parameters such as the size of reactor, the efficiency of heat transfer from
the hot reactor surfaces to and within the tyre mass and gas residence time in the
hot zone.
The yield of oil from the pyrolysis of tyres shown in Table 2 reached a maximum
at 475C with a high yield of 58.2 wt.% and that the yield of oil decreased with
increasing temperature. There was a corresponding increase in the yield of gas from
4.5 wt.% at 450C to 8.9 wt.% at 600C. Similar high yields of oil have been found
by other workers. For example, Roy and Unsworth [1] used a vacuum pyrolysis
unit to pyrolyse tyres at 415C and reported a maximum oil yield of 56.6 wt.%.
Kawakami et al. obtained an oil yield of 53 wt.% [6] from the pyrolysis of tyres in
a rotary kiln at between 540 and 640C. Williams et al. [4], pyrolysing tyres in a
small scale 50 g capacity, nitrogen purged static-bed reactor, found that the yield of
oil reached a maximum of 58.8 wt.%. Lucchesi and Maschio [11] produced 47 wt.%
oil at 500C in a continuously-fed vertical reactor with a rotating grate, purged by
a counter current of nitrogen.
Lower oil yields have been reported by Kaminsky and Sinn [5] who obtained a
40% oil yield in a fluidised bed reactor operated at 640C falling to 27% at 840C,
and Collin [12] who obtained a 23% oil yield using a rotary kiln reactor operated
at 700C. Williams et al. [20] obtained oil yields of a maximum of 32.5 wt.% for a
large scale 1 tonne batch tyre pyrolysis unit. High heating rates with short hot zone
residence times and rapid quenching of the products are regarded as favouring the
formation of liquid products, since the pyrolysis gases and vapours are condensed
before further reaction breaks down the higher molecular weight species into
gaseous products. Therefore, the removal of pyrolysis products from the hot zone
reduces the extent of secondary reactions which are known to increase the yield of
char at the expense of oil formation. At higher temperatures the major product is
gas. The high oil yields in this work from the batch reactor were obtained since the
volatile products were removed from the hot zone by the nitrogen purge gas, before
significant secondary char forming reactions could occur. Similarly the high oil
yields obtained by Roy and Unsworth [2] and Bennalal et al. [22] were obtained by
the rapid removal of the gases by the vacuum process of the reactor. For example,
their 19 kg h 1 pilot unit yielded 50% oil at 510C and 220 kPa, when
uncorrected for steel and fibres, representing approximately 57% when corrected to
allow comparison with the data in this paper. Cypres and Bettens [13] increased the
rate of removal of pyrolysis vapours from scrap tyres from a secondary hot zone by
increasing the flow of nitrogen carrier gas. They found that this increased the yield

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

139

of oil by reducing the secondary reactions. Primary vapours are first produced in
the pyrolysis process, the characteristics of which are most influenced by heating
rate. These primary vapours then degrade to secondary tars and gases, the
proportion and characteristics of which are a function of temperature and time.
The decrease in oil yield with increasing temperature and corresponding increase
in gas yield have been observed by other workers. For example, Williams et al.
[4,20], Kaminsky and Sinn [5] and Lucchesi and Maschio [11] all found significant
decreases in oil yield and increases in gas yield with increasing temperature of
pyrolysis.
Packed column gas chromatography analysis of the gaseous products showed
that the main gases were H2, CH4, C4H6, CO and CO2 and minor concentrations of
other hydrocarbon gases. Butadiene, C4H6, was the main gas evolved, and is
formed from the thermal degradation of the polymerbutadienestyrene rubber
used in the manufacture of tyres. Cypres and Bettens [13] have suggested that
hydrogen and methane are derived from secondary aromatisation reactions. Similarly, Williams and Taylor [28] have shown that hydrogen, methane and ethene
result from secondary aromatisation reactions which produce aromatic hydrocarbons. Roy and Unsworth [2] and Kaminsky and Sinn [5] have also shown the gas
phase to consist mainly of H2, CO, CO2, and hydrocarbons including CH4, C2H6,
C2H4, C3H8, C3H6, C4H10, C4H8 and C4H6. Dodds et al. [9] showed that the rubbers
used in tyre manufacture are characterised by carboncarbon double bonds, and
that thermal decomposition produces highly reactive free radicals, which are often
sub-units of the original rubber molecule. The gases generated have a significant
calorific value, in the order of 40 MJ m 3 and it has been suggested that the gases
may be used to provide the total energy requirement of the pyrolysis plant. Other
workers have reported the presence of other product gases for example, Teng et al.
[11], for instance, detected SO2 and NH3 in the product gases, while Wolfson and
co-workers [8] found H2S.

3.2. Oil analysis


3.2.1. Fuel properties
Table 3 shows the properties of the oils derived from the pyrolysis of the scrap
tyres. The carbon/hydrogen ratio is some what higher than that found for
petroleum derived fuels. The sulphur content is similar to a light fuel oil which are
typically about 1.4 1.5 wt.%. The nitrogen content is however, rather higher than
a gas oil which has a nitrogen content of about 0.05 wt.% but similar to a heavy
fuel oil at a nitrogen content of about 0.4 wt.%. Fluorine and chlorine contents in
the oils are significantly high and may have a consequences on environmental
emissions on combustion. Moisture content is low but still significant. The gross
calorific value is similar to a light fuel oil and reflects the potential of the oils to be
used as liquid fuels. The flash point of a liquid fuel is the temperature at which the
oil begins to evolve vapours in sufficient quantity to form a flammable mixture with
air. The temperature is an indirect measure of volatility and serves as an indication
of the fire hazards associated with storage and application of the fuel. The flash

140

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

Table 3
Properties of tyre derived pyrolysis oils
Property

Carbon (wt.%)
Hydrogen (wt.%)
Nitrogen (wt.%)
Sulphur (wt.%)
Fluorine (ppm)
Chlorine (ppm)
Ash (wt.%)
Oxygen (wt.% by difference)
Moisture content (%vol)
Flash point (C)
Gross calorific value (MJ kg1)

Tyre pyrolysis temperature (C)


450

475

500

525

560

600

85.6
10.4
0.5
1.4
80
150
0.002
2.0
2.8
14
42.6

84.6
11.2
0.5
1.4
100
270
0.002
2.2
3.0
15
42.9

84.9
10.2
0.5
1.3
140
190
0.002
3.1
4.4
15
42.1

87.1
9.1
0.7
1.3
100
210
0.002
1.8
4.2
17
42.4

86.1
10.0
0.6
1.3
140
90
0.002
2.0
4.4
18
42.1

87.9
10.1
0.5
1.3
100
100
0.002
0.1
4.6
18
41.2

nd, not determined.

point of the tyre derived oils was in the range 1418C. The flash point is low when
compared with petroleum refined fuels, for example, kerosene has a required
minimum flash point of 23C, diesel fuel of 75C and light fuel oil 79C. The low
flash points of the tyre oil are not surprising since the oil represents an un-refined
oil with a mixture of components having a wide distillation range.
Fig. 2 shows the molecular weight (Mr) range of tyre derived pyrolysis oil
measured using the size exclusion chromatography system compared with a range
of petroleum derived fuels. Only the 450C derived pyrolysis oil is shown for clarity,
although over the pyrolysis temperature studied there was only a small decrease in
the Mr range. Previous work on the detailed analysis of the Mr range of tyre oils
and their different chemical fractions has shown that increasing pyrolysis temperature up to 712C shows a decrease in the number average and weight average Mr
[29]. The Mr range of oils was from a nominal 50 to over 1500 Da. Comparison of
the molecular weight range of the oils with petroleum derived fuels show that a
light fuel oil and gas oil have a Mr range from 50 to 900 Da. Medium fuel oil has
a Mr range from 50 to approximately 10 000 Da. The Mr range of the various fuels
reflects the range of compounds and boiling points of the different fuels and gives
an indication of the relative distillation ranges of the various fuels. In this respect,
the tyre oils are similar to a light to medium fuel oil and would be expected to have
a similar distillation range.

3.2.2. Chemical class separation


Fractionation of the oils into chemical classes by liquid column chromatography
revealed a decrease in the percentage yield of aliphatics (pentane fraction), accompanied by an increase in the yields of aromatics (benzene) with increasing temperature (Table 4). Lucchesi and Maschio [11] found an increase in oil aromaticity with
temperature by studying the C:H ratio. Similar trends have been reported by

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

141

Fig. 2. Molecular weight range of tyre pyrolysis oil compared with petroleum derived fuels.

Cypres and Bettens [13], Wolfson et al. [8] and Kaminsky and Sinn [5]. A
DielsAlder type reaction mechanism as discussed below was suggested as the
cause of this increase in aromatisation [13,28].
Table 4
Chemical class fractionation of tyre pyrolysis oils in comparison to petroleum derived fuels
Final pyrolysis temperature (C)

Pentane fraction
(wt.%)

Benzene fraction
(wt.%)

Ethyl acetate fraction (wt.%)

Methanol fraction (wt.%)

450
475
500
525
560
600

51.3
49.0
47.2
45.3
43.7
36.1

36.7
42.6
40.5
41.8
41.7
45.6

8.6
6.3
10.7
8.9
7.3
11.6

3.3
1.8
1.6
4.0
7.3
6.6

74.6
83.5
35.7

19.8
11.1
58.4

3.0
2.5
4.8

2.6
2.9
1.2

Petroleum derived fuels


Gas oil
Light fuel oil
Medium fuel oil

142

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

Table 4 also shows the analysis of gas oil, light and medium fuel oils as
representative refined fuels from petroleum using the analytical scheme used to
chemically fractionate the tyre oils. Higher alkane concentrations in the lighter
petroleum derived fuels, the gas oil and light fuel oil are indicated by the high
percentage pentane fraction. The lighter petroleum fuels have lower corresponding
aromatic fractions. The medium fuel oil has chemical class fractions similar to those
of the tyre oils, with a high aromatic content. The ethyl acetate and methanol
fractions representing the more polar chemical classes were low in the petroleum
derived fuels, similar to the tyre oils.

3.2.3. Polycyclic aromatic hydrocarbon analysis


The PAH identified in the tyre pyrolysis oils in relation to pyrolysis temperature
are shown in Table 5. There is some concern over the presence of PAH in the
environment, since amongst chemical groups PAH comprise the largest group of
carcinogens [31]. The PAH found in tyre pyrolysis oil consists largely of alkylated
naphthalenes, fluorenes and phenanthrenes. Some of the PAH shown in Table 5
have been shown to be carcinogenic and/or mutagenic. Lee et al. [31] list the
relative carcinogenicities of certain PAH and show that chrysene, and certain triand tetramethylphenanthrene have been shown to give positive results in carcinogenicity tests. Longwell [32] has also shown that phenanthrene and the
methylphenanthrenes are mutagenic in both human and bacterial cell tests. In
addition, Barfnecht et al. [33] have shown that the methylfluorenes are biologically
active in mutagenicity bioassays.
Table 5 shows that as the temperature of pyrolysis is increased from 450 to 600C
the concentration of PAH in the oils increases. For example, naphthalene, fluorene,
phenanthrene increase from 465 to 1630 ppm, 280605 ppm and 95315 ppm
respectively. Alkylated compounds of naphthalene, fluorene and phenanthrene also
increase with pyrolysis temperature. The total alkylated naphthalenes shown in
Table 5 increase from 5475 to 14 275 ppm, the alkylated fluorenes increase from 790
to 2630 ppm and the alkylated phenanthrenes increase in concentration from 5380
to 8100 ppm as the pyrolysis temperature was increased. Table 5 shows that the
total PAH content of the oils increased from 1.53 wt.% at 450C to 3.43 wt.% total
oil at 600C.
PAH in tyre pyrolysis oils have also been reported by other workers. Kaminsky
and Sinn [5] reported concentrations of 0.85, 0.16, 0.29 and 0.21% by weight of tyre
for naphthalene, fluorene, phenanthrene and pyrene respectively, in the oil derived
from the pyrolysis of tyre pieces at 750C in a fluidised bed reactor. Cypres and
Bettens [13] pyrolysed scrap tyres via a two stage process comprising of an initial
low temperature pyrolysis phase followed by continuous post-cracking of the
volatile material at higher temperatures. They found that concentrations of PAH
increased with increasing post-cracking temperature. Concentrations of individual
PAH included naphthalene at 3.8% and phenanthrene at 1.6% by weight of tyre.
Other PAH detected were fluorene, acenaphthene, anthracene and methylnaphthalenes. Wolfson et al. [8] pyrolysed scrap tyres in a retort heated to 500C and
separated a light and heavy oil fraction. The heavy oil fraction was shown to

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

143

Table 5
Concentration of polycyclic aromatic hydrocarbons (PAH) in tyre derived pyrolysis oils in relation to
pyrolysis temperature (ppm)
PAH

Tyre pyrolysis temperature (C)


450

475

500

525

560

600

Naphthalene
2-Methylnaphthalene
1-Methylnaphthalene
Biphenyl
1-Ethylnaphthalene
2,6-Dimethylnaphthalene
1,7-Dimethylnaphthalene
1,6-Dimethylnaphthalene
1,5-Dimethylnaphthalene
1,2-Dimethylnaphthalene
Acenaphthene
Trimethylnaphthalene
Trimethylnaphthalene
Trimethylnaphthalene
Fluorene
Methylfluorene
2-Methylfluorene
1-Methylfluorene
Methylfluorene
Phenanthrene
Anthracene
Dimethylfluorene
2-Methylphenanthrene
2-Methylanthracene
4-Methylphenanthrene
1-Methylphenanthrene
Dimethlyphenanthrene
2,7-Dimethlyphenanthrene
Dimethlyphenanthrene
Fluoranthene
Dimethlyphenanthrene
Dimethlyphenanthrene
Pyrene
Trimethylphenanthrene
Tetramethylphenanthrene
Chrysene

465
650
460
1030
430
565
550
275
190
770
560
765
665
155
280
65
115
260
135
95
85
215
595
455
355
595
745
1255
105
120
730
440
530
530
30
30

420
570
490
1040
510
565
440
600
305
405
580
470
515
175
210
180
245
340
200
230
160
425
495
640
200
890
1200
1300
120
325
B5
385
105
400
35
15

725
730
625
1630
690
855
365
595
375
600
635
670
425
220
325
135
220
370
170
200
125
425
315
500
140
470
520
1740
455
325
490
305
120
600
65
35

1115
770
645
885
830
755
1020
715
1440
450
700
1050
330
695
290
90
175
310
320
125
135
165
470
1010
275
585
650
525
350
355
445
615
425
470
35
B5

665
1005
895
1320
745
995
700
485
635
460
620
605
830
430
295
240
335
450
195
195
225
320
815
720
605
600
760
1060
1260
790
1490
5195
225
520
30
30

1630
2365
1570
3000
1335
1990
560
1085
880
1385
1070
825
1570
710
605
585
745
555
280
315
295
465
1240
1140
730
555
1290
1075
540
1100
1210
370
115
940
150
60

Total (%)

1.53

1.52

1.72

1.92

2.67

3.43

contain the PAH, biphenyl, acenaphthene and alkyl naphthalenes. Williams and
Taylor [28] using a batch pyrolysis unit with post-pyrolysis heating of the derived
vapours, found a range of PAH including, benzofluoranthene and benzo[a]pyrene,
strongly carcinogenic compounds in tyre derived pyrolysis oils. Sulphur PAH
(PASH), such as dibenzothiophene and benzothiozole, and nitrogen PAH (PANH)
and alkylated quinolines, have been detected in tyre oils in previous studies [34,35].

144

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

The increase in PAH concentration with increasing temperature of pyrolysis has


also been reported by other researchers [5,13,28]. For example, Willliams and
Taylor [28] for the pyrolysis of tyres in a batch reactor with post-pyrolysis heating
of the tyre vapours. They found that the total PAH content of tyre derived oils
increased from 1.4 wt.% at 500C post-pyrolysis temperature to over 10 wt.% total
oil at 720C post-pyrolysis temperature. Cypres and Bettens [13], also using a
post-pyrolysis secondary reactor, found a similar trend of increasing PAH in
derived tyre pyrolysis oils with secondary reactions of the vapours. A DielsAlder
reaction mechanism has been suggested as being responsible for the increase in
aromatic content with temperature [13,28]. A general reaction scheme for the
aromatisation of alkene compounds to aromatic compounds, with a consequent
increase in the methane and hydrogen of the gas phase has been suggested [13].
Pyrolysis of tyres leads to the production of ethene, propene and 1,3 butadiene,
which react to form cyclic alkenes. Dehydrogenation of the cyclic alkene compounds with six carbon atoms occurs to produce single ring aromatic compounds
and as a result of subsequent associative reactions leads to the formation of PAH,
such as naphthalene and phenanthrene. Work by Williams and Bottrill [34] has
shown that a similar increase with temperature can be seen in the content of PASH
compounds in tyre derived oils.
Further evidence that secondary reactions of hydrocarbons at moderate to high
temperatures can produce PAH has been shown in studies on pure hydrocarbon
compounds. For example, Cypres [36] pyrolysed model aliphatic compounds and
confirmed the Diels Alder route to PAH formation in post-pyrolysis cracking
reactions. Pyrolysis of n-decane produced alkenes by thermal degradation, post-pyrolysis cracking of the alkenes between 600 and 900C with a 2 s residence time
showed a decrease in light olefins and the formation of single ring aromatic
compounds, such as benzene, toluene and alkyl-aromatics by DielsAlder reactions. In addition, naphthalene and alkyl-naphthalenes are formed and condensation reactions may continue to produce higher PAH. Fairburn et al. [37] examined
the flash pyrolysis of n-hexadecane in a micro-reactor at temperatures between 576
and 842C. They showed that hexadecane initially pyrolyses to form light alkanes
in the range C1 C3 and alkenes in the range C2 C15. Further decomposition at
higher temperatures (or increased reaction time) produces a decomposition of the
heavier alkenes to lighter alkenes (C4 C6) and subsequently formation of di-alkenes
such as butadiene via dehydrogenation of the lighter alkenes. The butadiene then
immediately combines with available light alkenes such as ethene and propene to
form simple aromatic compounds such as benzene, toluene, ethylbenzene and
styrene via the Diels Alder reaction. Higher molecular weight PAH are then
formed by further reaction between aromatic compounds and olefins. Dupeyre et
al. [38] examined the steam cracking of n-hexadecane at 750850C in a quartz
flow reactor, with the aim of maximising alkene production. They found that
aromatics are increasingly formed at the expense of liquid olefin compounds with
increasing temperature and reaction time. The formation of PAH was attributed to
surface condensation reactions and was noted to increase with temperature, for
example 0.8 wt.% naphthalene was recovered at 750C with a H2O inlet flowrate of

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

145

23.2 g h 1, and 4.6 wt.% at 850C. As the temperature increased, heavier PAH, for
instance anthracene and pyrene, were formed. The formation of PAH via the
DielsAlder reaction scheme is illustrated in Fig. 3 [13].
In addition, it has been shown that increased residence time of pyrolysis vapours
in the hot zone of the reactor will lead to increases in secondary reactions such as
the formation of PAH [28,37]. The long residence time (maximum 120 s) of the
pyrolysis vapours in the hot zone of the reactor in this work suggests that increased
formation of PAH is likely. Where commercial scale tyre pyrolysis reactors have
significant residence times of pyrolysis vapours in the hot zone of the reactor, PAH
are also likely to be a significant proportion of the derived condensed oils.

3.2.4. Volatile hydrocarbons in tyre oils


Table 6 shows the volatile or light hydrocarbons, benzene, toluene, xylenes
(dimethylbenzenes), styrene, limonene and indene and their alkylated homologues.
Limonene exits in the form D,L-limonene (dipentene) and D-and L-limonene, the

Fig. 3. DielsAlder reaction for the formation of polycyclic aromatic hydrocarbons in scrap tyre
pyrolysis [13].

146

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

Table 6
Concentration of light aromatic hydrocarbons in tyre derived pyrolysis oils (ppm)
Volatile hydrocarbon

Benzene
Toluene
Ethylbenzene
1,2-Dimethylbenzene
1,4-Dimethylbenzene
Styrene
1,3-Dimethylbenzene
Trimethylbenzene
Trimethylbenzene
Trimethylbenzene
Methylstyrene
Trimethylbenzene
4-Methylstyrene
Trimethylbenzene
Methylstyrene
Limonene
Indene

Tyre pryolysis temperature (C)


450

475

500

525

560

600

B5
2250
250
2780
2750
1205
920
840
1050
1550
730
1075
730
370
6020
31 320
2190

55
3200
235
3190
2665
1705
1020
825
1265
1350
570
1070
570
440
6025
30 330
2630

770
6095
120
3345
3620
1950
1325
1255
1670
2370
1090
1325
1090
490
7630
29 010
3175

2950
17 740
405
5710
6880
3545
2450
1085
1240
2320
1145
1295
1145
675
8865
28 965
3090

70
7770
370
5875
8350
3635
2570
1285
1530
3210
1590
1395
1590
320
9030
24 590
3105

605
5070
190
3530
3120
1915
1040
820
1200
1450
715
1095
715
330
6950
25 130
1560

dipentene racemic form being found in tyre pyrolysis oils [23]. However, except for
the optical isomerism, dipentene has the same physical properties as D-and Llimonene.
The concentrations of volatile aromatic hydrocarbons increase with increasing
temperature of pyrolysis up to temperatures of 525560C. However at higher
pyrolysis temperature of 600C, there was a fall in concentration. The concentrations of some volatile aromatic hydrocarbons reach significant concentrations, e.g.
toluene (up to 1.77 wt.%), total xylenes (up to 1.68 wt.%), styrene (up to 0.36 wt.%),
limonene (up to 3.13 wt.%) and indene (0.32 wt.%).
Other work on tyre pyrolysis has also shown high concentrations of individual
chemicals in the derived pyrolysis oils. For example, Williams et al [39] reported
0.47 wt.% toluene, 1.78 wt.% xylene/styrene and limonene up to 7.43 wt.% of
pyrolysis oil in a small scale gas purged static batch reactor at 420C. Increasing the
temperature of pyrolysis caused a decrease in the concentration of limonene and an
increase in the concentration of benzene, xylene and styrene. Pakdel et al. [23]
pyrolysed tyres in a 3.5 kg h 1 vacuum pyrolysis unit and found benzene
concentrations in the naphtha fraction of the derived oil of 2.54 wt.% benzene, 6.95
wt.% toluene, 6.12 wt.% xylenes and 14.92 wt.% limonene which represented, 0.68
wt.%, 1.86 wt.%, 1.6 wt.%, 4.00 wt.% in the derived pyrolysis oil respectively. They
also found that increasing temperature of the hearths in the vacuum pyrolysis unit
from 226 to 510C produced an increase in the concentration of benzene, toluene
and xylenes but reduced the concentration of limonene. Cypres and Bettens [13]
have shown much higher concentrations of volatile compounds from the pyrolysis

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

147

of tyres with post-pyrolysis cracking of the pyrolysis vapours. For example, they
reported maximum concentrations of benzene 36.4 wt.%, toluene 16.8 wt.%, xylenes
6.7 wt.% and styrene 5.83 wt.%. Wolfson et al. [8] have reported significant
concentrations of benzene, toluene, xylenes and styrene in the oil derived from the
fixed bed batch pyrolysis of tyres and that their concentration was influenced by
temperature. Kaminsky and Sinn [5] used a fluidised bed pyrolysis unit to pyrolyse
tyres at temperatures from 640 to 840C. Concentrations of volatile aromatic
compounds were high. For example, pyrolysis of tyre pieces at the fluidised bed
temperature of 750C produced an oil with a benzene concentration of approximately, 16 wt.%, toluene 12 wt.% and styrene 0.5 wt.%, xylenes were detected only
in trace concentrations.
The high concentrations of volatile hydrocarbons present in the tyre oils have
been suggested as a potential high value products which could offset the costs of
tyre disposal. For example, xylenes are regarded as major industrial chemicals and
have applications in the plastics industry. o-Xylene is used to produce phthalic
anhydride which is used to produce plasticisers, dyes and pigments, m-xylene
derivatives have applications in the polyester resin and fibre industries and p-xylene
derivatives are used in the production of polyester fibres [40]. Toluene has a wide
range of applications as a chemical feedstock and is used for example, in the
production of pesticides, dyestuffs, surfactants and solvents. Styrene is one of the
most important building blocks used in the production of plastic materials and is
also used to make synthetic rubber and other polymers [40]. Also, indene is
regarded as being of particular technical importance since it is used to produce
indene/coumarone resins which find extensive application, especially in the production of adhesives, as reinforcers and tackifiers in the production of commercial
rubber products, and in paint manufacture [40].
However, it is the high concentration and potential high value of limonene in tyre
pyrolysis oils which has attracted most interest [23]. Limonene is used in the
formulation of industrial solvents, resins and adhesives and as a dispersing agent
for pigments. It is also used as a feed stock for the production of fragrances and
flavourings. It has been used in a wide range of applications including, waterbased
de-greasers, natural lemon scented all-purpose cleaners, hand cleaners and replacements for chlorofluorocarbon solvents to clean electronic circuit boards [23].
Tyres are composed of vulcanized rubber in addition to rubberized fabric with
reinforcing textile cords, steel or fabric belts and steel-wire reinforcing beads. The
rubbers used include, styrene butadienecopolymer (SBR), natural rubber (polyisoprene), nitrile rubber, chloroprene rubber and polybutadiene rubber. Limonene
has been detected as a major product of the pyrolysis of rubbers and the
mechanism of formation discussed. For example, Groves et al. [41] analysed the oil
derived from the pyrolysis of natural rubber in a pyrolysis-gas chromatograph at
500C. They showed that the major products were the monomer, isoprene and the
dimer, dipentene (d,L-limonene) with other oligomers up to the hexamer in
significant concentrations. They suggested that the isoprene monomer is formed via
a depropagating mechanism of the polymer chain and that the dipentene dimer is
formed either as a product resulting from intramolecular cyclisation followed by

148

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

scission, or by monomer recombination via a DielsAlder type reaction. Tamura et


al. [42] have also shown that isoprene and D,L-limonene are formed in high
concentration in natural rubber pyrolysis and they suggest that they are produced
by depolymerisation from polymer radicals, by b-scission of the double bond. The
polymer radicals are liable to form six-membered rings, especially under mild
pyrolysis conditions and therefore the D,L-limonene is formed predominantly at
lower temperature. Bhowmick et al. [43] also examined the pyrolysis of natural
rubber using a TGA. They showed that degradation starts at about 330C in
nitrogen with a peak weight loss at about 400C. They suggest decomposition
follows from radical generation via polymer chain scission and the formation of
isoprene, D,L-limonene and other smaller compounds. Chien and Kiang [44]
pyrolysed natural rubber in helium at 384C and identified isoprene and D,Limonene as the main products of pyrolysis. They also identified a wide range of
other products including, alkane and alkene gases, toluene, xylene, octene and
hydrocarbons up to C16H26. They also suggest a similar mechanism to Bhowmick
et al. [43], in that isoprene and D,L-limonene are formed by polymer chain scission
and minor compounds are formed via chain propagation with or without intramolecular hydrogen transfers. Bhowmick et al. [43] and Chien and Kiang [44]
have suggested that the other products of pyrolysis can be accounted for by the
thermal decomposition of isoprene and dipentene.
The pyrolysis of polybutadiene rubber has been examined by Brazier and
Schwartz [45] in a nitrogen atmosphere up to 550C using TGA. They showed a
two stage thermal decomposition of polybutadiene rubber, with maximum rates of
weight loss at 370C and 470C, which was dependent on heating rate and sample
size. Analysis of the products from the first stage of weight loss showed they were
mainly butadiene and D,L-limonene formed as a result of depolymerisation. Material not undergoing depolymerisation cyclizes and crosslinks to form a residue
which degrades in the second stage. The products of pyrolysis from the second
stage were a complex and varied mixture of hydrocarbons. Madorsky et al. [46]
have also examined the pyrolysis of polybutadiene rubber and similarly showed that
butadiene vinylcyclohexene and D,L-limonene were formed in high concentrations.
Several researchers have reported the degradation of limonene, formed from
scrap tyre or rubber pyrolysis at higher temperature [23,39,42]. Table 6 also shows
that limonene is formed in high concentration at lower temperatures and degrades
as the pyrolysis temperature is increased from 450 to 600C. In addition, there is a
consequent increase in other volatile aromatic hydrocarbons such as toluene,
xylenes, trimethylbenzenes and styrene. Above temperatures of between 525 and
560C there is an overall decrease in volatile hydrocarbon concentration. Early
work by Pines and Ryer [47] on the thermal degradation of limonene at 450C over
copper pellets at atmospheric pressure showed that the main products were,
p-cymene, alkylated benzenes and m-xylene. They proposed a reaction scheme
involving the formation of allo-ocimene via a biallylic diradical and the triene. The
allo-ocimene can then form aromatic hydrocarbons such as m-xylene, m-ethyltoluene, trimethylbenzenes etc. In addition, Traynor et al. [48] pyrolysed limonene
over pyrex chips at 465C and found that it formed biallylic diradicals, which form

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

149

allo-ocimene. They also found that the diradical can also recyclise into partiallyracemic limonene and optically active structural isomers, such as vinylcyclohexene.
The allo-ocimene subsequently formed a variety of conjugated dienes plus m-xylene
as the major aromatic product and other aromatics. The pyrolytic reactions as
described by these studies and are presented diagramatically in Fig. 4 [47,48]. In the
case of tyre pyrolysis products, Pakdel et al. [23] and Williams et al. [39] suggest
that limonene decomposition to a range of products including benzene, xylene,
toluene, trimethylbenzene, styrene and methylstyrene, occurs above 500C.
Above the pyrolysis temperature of 525560C, the volatile hydrocarbons which
are suggested as degradation products of limonene decrease in concentration in the
oils. It is at this temperature range that the formation of PAH increases. Consequently the single ring aromatic species may be decreasing in concentration as a
consequence of Diels Alder aromatisation to form PAH.

Fig. 4. Thermal degradation of limonene [47,48].

150

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

4. Conclusions
1. Pyrolysis of scrap tyres produces an oil similar in properties to a light fuel oil,
with similar calorific value, sulphur and nitrogen contents.
2. The oils contain significant concentrations of polycyclic aromatic hydrocarbons
some of which have been shown to be either carcinogenic and/or mutagenic.
The concentration of PAH increased from 1.5 to 3.4 wt.% of oil as the pyrolysis
temperature increased from 450 to 600C.
3. The formation of PAH was attributed to a DielsAlder type mechanism
involving cyclisation of alkenes and dehydrogenation to form aromatic hydrocarbons.
4. A range of potentially high value volatile hydrocarbons were identified in
significant concentrations in the oils. The concentration of limonene in particular which may represent a potentially high value chemical, reached a maximum
concentration of 3.1 wt.% in the oils.
5. A reaction scheme for the formation of limonene via depolymerisation of the
rubber components of tyres was discussed as the route to limonene formation.
At higher temperatures, limonene is thermally degraded to other volatile aromatic hydrocarbons. The reaction scheme involving an allo-ocimene intermediate was discussed.

Acknowledgements
This work was funded by an award from the University of Leeds Research Fund.
We would also like to acknowledge the support of the John Henry Garner
Scholarship for the support of AMC. The authors wish to thank Peter Thompson,
Patrick Horne and Richard Bottrill for their invaluable help in analyses. We wish
to thank Ed Woodhouse for his work in constructing the pyrolysis unit and his help
in making it operable.

References
[1] G. Bressi (Ed.), Recovery of Materials and Energy from Waste Tyres, Institute of Solid Waste
Associates International, Copenhagen, 1995.
[2] C. Roy, J. Unsworth, Pilot plant demonstration of used tyres vacuum pyrolysis, in: G.L. Ferrero,
K. Maniatis, A. Buekens, A.V. Bridgwater (Eds.), Pyrolysis and Gasification, Elsevier Applied
Science, London, UK, 1989.
[3] S.M. Ogilvie, Opportunities and Barriers to Scrap Tyre Recycling, AEA Technology Report
AEA/CS R1026/C, AEA Technology, Harwell, UK, 1995.
[4] P.T. Williams, S. Besler, D.T. Taylor, Fuel 69 (1990) 1474 1482.
[5] W. Kaminsky, H. Sinn, Pyrolysis of plastic waste and scrap tyres using a fluidised bed process, in:
J.L. Jones, S.B. Radding (Eds.), Thermal Conversion of Solid Wastes and Biomass, ACS Symposium Series 130, American Chemical Society Publishers, Washington DC, 1980.
[6] S. Kawakami, K. Inoue, H. Tanaka, T. Sakai, Pyrolysis process for scrap tyres, in: J.L. Jones, S.B.
Radding (Eds.), Thermal Conversion of Solid Wastes and Biomass, ACS Symposium Series 130,
American Chemical Society Publishers, Washington DC, 1980.

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

151

[7] C. Roy, B. Labrecque, B. de Caumia, Resour. Conserv. Recycl. 4 (1990) 203 213.
[8] D.E. Wolfson, J.A. Beckman, J.G. Walters, D.J. Bennett, Destructive distillation of scrap tyres, US
Dept. of Interior, Bureau of Mines Report of Investigations 7302 (1969).
[9] J. Dodds, W. F. Domenico, D.R. Evans, L.W. Fish, P.L. Lassahn, W.J. Toth, Scrap tyres: a
resource and technology evaluation of tyre pyrolysis and other selected alternative technologies, US
Dept. of Energy Report EGG-2241, 1983.
[10] B. Bilitewski, G. Hardtle, K. Marek, Usage of carbon black and activated carbon in relation to
input and technical aspects of the pyrolysis process, in: G.L Ferrero, K. Maniatis, A. Buekens, A.V.
Bridgwater (Eds.), Pyrolysis and Gasification, Elsevier Applied Science, London, UK, 1989.
[11] A. Lucchesi, G. Maschio, Conserv. Recycl. 6 (3) (1983) 85 90.
[12] G. Collin, Pyrolytic recovery of raw materials from special wastes, in: J.L. Jones, S.B. Radding
(Eds.), Thermal Conversion of Solid Wastes and Biomass, ACS Symposium Series 130, American
Chemical Society Publishers, Washington DC, 1980.
[13] R. Cypres, B. Bettens, Production of benzoles and active carbon from waste rubber and plastic
materials by means of pyrolysis with simultaneous post-cracking, in: G.L. Ferrero, K. Maniatis, A.
Buekens, A.V. Bridgwater (Eds.), Pyrolysis and Gasification, Elsevier Applied Science, London,
UK, 1989.
[14] B. Sahouli, S. Blacher, F. Brouers, H. Durmstadt, C. Roy, S. Kaliaguine, Fuel 75 (1996)
1244 1250.
[15] D.B. Olsen, J.C. Pickens, Combust. Flame 57 (1984) 199.
[16] T.R. Henderson, J.D. Sun, A.P. Li, W.E. Bechtold, T.M. Harvey, J. Shabanowitz, D.F. Hunt,
Environ. Sci. Technol. 18 (1984) 428.
[17] P.T. Williams, K.D. Bartle, G.E. Andrews, Fuel 65 (1986) 1150.
[18] P.T. Williams, M.K. Abbass, G.E. Andrews, K.D. Bartle, Combust. Flame 75 (1989) 1.
[19] A. Herlan, Combust. Flame 31 (1978) 297.
[20] P.T. Williams, S. Besler, D.T. Taylor, Proc. Inst. Mech. Eng. 207 (1993) 55 63.
[21] M.-Y. Wey, B.-H. Liou, S.-Y. Wu, C.-H. Zhang, J. Air Waste Manag. Assoc. 45 (1995) 855.
[22] B. Benallal, H. Pakdel, S. Chabot, C. Roy, Fuel 74 (11) (1995) 1589.
[23] H. Pakdel, C. Roy, H. Aubin, G. Jean, S. Coulombe, Environ. Sci. Technol. 25 (9) (1992) 1646.
[24] D.L. Vassilaros, R.C. Kong, D.W. Later, M.L. Lee, J. Chromatogr. 252 (1982) 1.
[25] C.E. Rostad, W.E. Pereira, J. High Resolut. Chromatogr. Chromatogr. Commun. 9 (1986) 328.
[26] M.L. Lee, D.L. Vassilaros, C.M. White, M. Novotny, Anal. Chem. 51 (1979) 768.
[27] R.K. Sharma, N.N. Bakhshi, Bioresour. Technol. 35 (1991) 57.
[28] P.T. Williams, D.T. Taylor, Fuel 72 (1993) 1469 1474.
[29] P.T. Williams, D.T. Taylor, J. Anal. Appl. Pyrolysis 29 (1994) 111 128.
[30] A.V. Bridgwater, S.A. Bridge, A review of biomass pyrolysis and pyrolysis technologies, in: A.V.
Bridgwater, G. Grassi (Eds.), Biomass Pyrolysis Liquids Upgrading and Utilisation, Elsevier
Applied Science, Essex, UK, 1991
[31] M.L. Lee, M. Novotny, K.D. Bartle, Analytical Chemistry of Polycyclic Aromatic Compounds,
Academic Press, New York, 1981.
[32] J.P. Longwell, in: J. Lahaye, G. Prado (Eds.), Soot in Combustion Systems and its Toxic Properties,
Plenum, New York, 1983.
[33] T.R. Barfnecht, B.M. Andon, W.G. Thilly, R.A. Hites, in: M. Cooke, A.J. Dennis (Eds.), Chemical
Analysis and Biological Fate: Polycyclic Aromatics Hydrocarbons, Battelle Press, Columbus OH,
1980.
[34] P.T. Williams, R.P. Bottrill, Fuel 74 (1995) 736.
[35] S. Mirmiran, H. Pakdel, C. Roy, J. Anal. Appl. Pyrolysis 22 (1992) 205.
[36] R. Cypres, Fuel Process. Technol. 15 (1987) 1.
[37] J.A. Fairburn, L.A. Behie, Y. Svrcek, Fuel 69 (1990) 1537.
[38] D. Depeyre, C. Flicoteaux, C. Chardaire, Ind. Eng. Chem. Process Des. Dev. 24 (1985) 1251.
[39] P.T. Williams, S. Besler, D.T. Taylor, R.P. Bottrill, J. Inst. Energy 68 (1995) 11 21.
[40] H.G. Franck, J.W. Stadelhofer, Industrial aromatic chemistry, Springer-Verlag, London, 1988.
[41] S.A. Groves, R.S. Lehrle, M. Blazso, T. Szekely, J. Anal. Appl. Pyrolysis 19 (1991) 301.
[42] S. Tamura, K. Murakami, H. Kuwazoe, J. Appl. Polym. Sci. 33 (1987) 1122.

152

A.M. Cunliffe, P.T. Williams / J. Anal. Appl. Pyrolysis 44 (1998) 131152

[43] A.K. Bhowmick, S. Rampalli, K. Gallagher, R. Seeger, D. McIntyre, J. Appl. Polym. Sci. 32 (1987)
1125.
[44] J.C.W. Chien, J.K.Y. Kiang, Eur. Polym. J. 15 (1979) 1059.
[45] D.W. Brazier, N.V. Schwartz, J. Appl. Polym. Sci. 22 (1978) 113.
[46] S.L. Madorsky, S. Straus, D. Thompson, L. Williamson, J. Res. Natl. Bur. Stand. 42 (1949) 499.
[47] H. Pines, J. Ryer, J. Am. Chem. Soc. 77 (1955) 4370.
[48] S.G. Traynor, K.J. Crowley, W. Cocker, J. Chem. Res. (S) (1981) 175.

You might also like