Sie sind auf Seite 1von 10

Thermal degradation of polyethylene and polystyrene

from the packaging industry over different catalysts


into fuel-like feed stocks
N. Miskolczi
*
, L. Bartha, Gy. Deak
University of Veszprem, Department of Hydrocarbon and Coal Processing, Egyetem Street 10, Veszprem H-8200, Hungary
Received 14 October 2004; received in revised form 5 January 2005; accepted 6 January 2005
Available online 6 October 2005
Abstract
Thermal degradation of waste polymers was carried out as a suitable technique for converting plastic polymers into liquid hydrocarbons,
which could be used as feed stock materials. The catalytic degradation of waste plastics (polyethylene and polystyrene) was investigated in
a batch reactor over different catalysts (FCC, ZSM-5 and clinoptillolite). The effects of catalysts and their average grain size on the properties
of main degradation products (gases, gasoline, diesel oil) are discussed. The temperature range of 410e450

C was used in the process. Both
equilibrium FCC catalyst and natural clinoptilolite zeolite catalyst had good catalytic activity to produce light hydrocarbon liquids, and ZSM-5
catalyst produced the highest amount of gaseous products. Gases and liquids formed in cracking reactions were analyzed by gas chromatography.
The liquid products consisted of a wide spectrum of hydrocarbons distributed within the C
5
eC
28
carbon number range depending on the crack-
ing parameters. The composition of hydrocarbons had linear non-branched structure in case of polyethylene, while from polystyrene more ar-
omatics (ethyl-benzene, styrene, toluene, and benzene) were produced. The yields of volatile products increased with increasing degradation
temperature. The olen content of liquids was measured with an infrared technique and an olen concentration of 50e60% was observed.
The concentration of unsaturated compounds increased with decreasing temperature, and in the presence of catalysts. The activation energies
were calculated on the basis of the composition of volatile products. The apparent activation energies were decreased by catalysts and catalyst
caused both carbon-chain and double bond isomerisation.
2005 Elsevier Ltd. All rights reserved.
Keywords: Waste polyethylene and polystyrene; Catalysts; Further utilization; Feed stocks; Fuel-like fractions; Octane number
1. Introduction
Our modern society is unimaginable without plastics. Now-
adays both the consumption and production of polymers are
increasing, but the increasing amount of polymer wastes
from them generates further environmental problems. On the
other hand the capacity of landll sites and incinerators is lim-
ited. Due to the limitation, chemical (e.g. thermal and thermo-
catalytic cracking of plastic wastes) recycling is growing in
importance and it is in the focus of the solution of the problem.
In this way waste plastics are converted into fuels or valuable
feed stocks for the petrochemical industry.
Many recent results of investigations of waste polymer deg-
radation of polyethylene and polypropylene with or without
catalysts are published in the literature [1e12]. Mainly polyo-
lens (polyethylene and polypropylene) and polystyrene were
the target polymers, because their cracking resulted in products
with favourable properties for further application [13e17].
The second reason of the high interest of the investigation of
cracking of the above mentioned polymers is that their ratio
in the produced plastics is more than 65%, and this value is
considerably higher in mixture of plastic wastes. The cracking
of polyethylene, polypropylene and even polystyrene in their
individual form is well described [2,3,8,9,12,14,16]. However,
* Corresponding author. Fax: C36 88 423 225.
E-mail address: mnorbert@almos.vein.hu (N. Miskolczi).
0141-3910/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2005.01.056
Polymer Degradation and Stability 91 (2006) 517e526
www.elsevier.com/locate/polydegstab
there are considerably less experimental results on the degra-
dation of mixtures of these wastes, but polyethylene or poly-
styrene are widely used in high amounts for packaging and
they constitute the main components of plastic wastes from
domestic use.
Degradation of polymer wastes produces gases, oils and
a residue fraction. The yield and characteristics of these frac-
tions can be affected with the cracking parameters (e.g. tem-
perature or catalysts etc.). For the further application it is
important that both thermal and thermo-catalytic degradations
produced oils of a relatively wide distribution of hydrocar-
bons. Comparing the degradations with or without catalysts,
catalytic degradations have the advantage of a lower tempera-
ture of degradation. Otherwise as it is well known catalysts
have another positive effect on the degradation, because the
yields of valuable lighter hydrocarbons are increasing and
better quality of products are obtained in the presence of
catalysts. Several reports have been published on the degrada-
tion of waste polymers in the presence of catalysts and give an
excellent summary on the chemical recycling of polymers. It
was described that the most commonly used catalysts in the
catalytic degradation of polymers are acidic zeolite catalysts
[18e23]. Generally silicaealumina, b-zeolite, Y-zeolite,
mordenite, HZSM-5, MCM-41, clinoptilolite, etc. are used
as catalysts, which are of similar pore size and structure or
acidity. Among zeolite type catalysts HZSM-5 signicantly in-
crease the yield of gases and olens due to their great micro-
porous surface area. The ratio of Si/Al is another important
parameter in case of cracking reactions, because the acidic cat-
alysts (HZSM-5, mordenite, Y-zeolite, etc.) have greater ef-
ciency in cracking reactions than the less acidic catalysts,
e.g. amorphous silicaealumina. The structure of catalysts
and their pore size fundamentally determine the cracking
property of the catalyst; e.g. it was found that the difference
in activity of mordenite and zeolite catalysts (e.g. HZSM-5,
Y-zeolite) is due to differences in their structure [22]. Morden-
ite contains pores of relatively large size (about 7 !8 A

),
while the pores of HZSM-5 are smaller (about 5 !5 A

) and
it is known that the length of products is principally deter-
mined by the pore size. The greater the pore sizes the longer
the formed hydrocarbons. Also the geometrics of catalysts
are important. The acidity of catalysts mainly determinates
the cracking behaviour of CeC bonds and the probability of
b-scission. However, the amorphous part of catalysts fosters
the isomerisation and aromatisation reactions. Therefore the
HZSM-5 catalyst shows quite high selectivity in the formation
of parafns and olens and branched hydrocarbons and the
yield of gases is also high. As a matter of fact high yields of
both gases and lighter liquids are due to the great microporous
surface area and the high yield of gases observed with the use
of amorphous silicaealumina without 3D structural pores is
explained with its great acidity. Lots of experiments have
been performed on the investigation of clinoptilolite zeolite
or uid catalytic cracking (FCC) catalysts (fresh or
used) as a catalyst for this process. Clinoptilolite catalysts
usually have pore structure of monoclinic framework consist-
ing of a 10-membered ring (7.6 !3.0 A

) and an 8-membered
ring (3.3 !4.6 A

). This is a silica-rich member of the heu-


landite family. ZSM-5 catalyst contains pores of relatively
small size (about 5 !5 A

) [24]. FCC catalysts usually have


faujasite structure with a ratio of Si/Al greater than 1.5. The
synthetic X- and Y-zeolites have the same crystallographic
structure as faujasite. Y-zeolites have pores of the size of
common molecules and are very useful in rening reactions, be-
cause the structure of Y-zeolite is one of the most open of all
zeolites. FCC catalyst is formed by the arrangement of truncated
octahedra forming cages of 13 A

diameter, and the pore opening


to these cages are 12-membered rings approximately 7.4 A

.
In the present paper, we report the thermal and thermo-
catalytic cracking behaviour of a waste high-density polyethyl-
ene (LDPE) and polystyrene (PS) over different acidic catalysts
(ZSM-5, equilibrium FCC, clinoptilolite) in a batch reactor un-
der isothermal conditions. The goal of our work was to inves-
tigate the cracking of an HDPE and PS mixture over different
catalysts and the effect of catalyst grain size on the properties
of products. Both gas and liquid products formed in cracking
reactions were analyzed by standardized and non-standardized
methods. The effect of the cracking parameters (temperature,
catalysts and their average grain size) on the yields and
structure of products were investigated. The apparent kinetic
parameters of reaction for the overall degradation are obtained
in dependence on temperature. Furthermore the possibility of
further utilization of fuel-like products was investigated.
2. Experimental
2.1. Raw materials
In our experiments commercial high-density polyethylene
(HDPE) and polystyrene (PS) from the packaging industry
were crushed to pieces of 4e7 mm and used as raw material.
The melt ow index (MFI) (at 190

C with 2160 g load) was
0.538 g/10 min, and 0.519 g/10 min in case of HDPE and PS
waste, respectively. The used raw materials had sulphur and
nitrogen content from additives (e.g. antioxidant, re-retardant
and antifuming additives). Waste polymers were used not in
their individual form, because 10% of polystyrene and 90%
of high-density polyethylene were mixed and cracked together.
2.2. Catalysts and their characterization
Three different cracking catalysts were used in thermo-
catalytic cracking: an equilibrium uid catalytic cracking catalyst
(FCC), a commercial ZSM-5 catalyst and a clinoptilolite
containing rhyolite tuff (Clin.). The Si/Al atomic ratio of the
samples was determined by X-ray spectrometry. Nitrogen
adsorptionedesorption isotherms at 77 K of the calcined
catalysts were obtained with a Micromeritics ASAP 2000 in-
strument. Firstly catalysts were degassed under vacuum at
160

C for 5 h, then the surface area of catalysts was calculated
by the application of BET procedure. The pore size distributions
were determined by an Analysette 22 type grain size analyzer
using laser ray according to the Fraunhofer angle deviation
theory. Table 1a and b shows the main properties of the
518 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
uncrushed catalysts. Macro- and microporous surface areas, Si/
Al ratios which are proportional to acidities and cationic compo-
sitions are signicantly different. The characteristic grain
sizes of uncrushed catalysts decrease in sequence of FCC O
Clin. OZSM-5. In order to compare the effect of the average
grain size of catalysts on the yields and characteristics of end
products, catalysts were crushed, and used also in their smaller
grain size. Apart from decreasing the average grain size of cata-
lysts, their other properties were not changed signicantly.
Fig. 1 represents the distribution of average grain size of
crushed and uncrushed catalysts. Uncrushed catalysts had an
average grain size of 12e60 mm, and signicant differences
can be observed between the distribution curves. By crushing
catalysts the grain sizes can be decreased to 1.5e1.8 mm, fur-
thermore the grain size distribution curves become similar.
2.3. Cracking apparatus and the analyses of products
Pyrolysis of HDPE and PS was performed in a Pyrex glass
reactor (volume: 500 cm
3
) under atmospheric pressure by
batch operation at isothermal temperatures (410, 430 and
450

C). The temperature of melted polymer was measured
with a thermometer. In each case 200 g of waste polymer
was fed into the reactor and in case of thermo-catalytic crack-
ing 2% catalyst by weight was added to the mixture of waste
polymers. The applied cracking apparatus is shown in Fig. 2.
For ensuring an inert atmosphere nitrogen was introduced into
the polymer with a ow rate of 7 dm
3
/h. The liquid products
obtained during the degradation were condensed using a water
condenser then collected and analyzed. The gaseous products,
which are not condensed as liquid products, were also mea-
sured and analyzed.
2.4. Analysis of products
Each cracking fraction was analyzed using the following
methods. The following standardized methods were used:
liquid density measurement (MSZ EN ISO 12185), determina-
tion of distillation curve (ASTM-D 1078), determination of
sulphur content of naphtha and diesel oil-like fractions
(ASTMD 6428 99), determination of nitrogen content of naph-
tha and diesel oil-like fractions (ASTMD 6366 99), determina-
tion of ash point (ISO 2719:2002 and MSZ 15967:1979),
determination of CFPP (MSZ EN 116:1999), corrosion test
(MSZ EN ISO2160:2000), determination of kinematic viscos-
ity (MSZ ISO 3105:1998).
Gases were analyzed using a Carlo-Erba Vega Series GC
6000 gas chromatograph with ame ionization detector pro-
vided with a 50 m!0.32 mm fused silica column with
Al
2
O
3
/KCl coating, at 40

C.
Liquid products were analyzed by gas chromatography
with a ame ionization detector using a TRACE GC gas chro-
matograph. It was provided with a 30 m x 0.32 mm Rtx

-1
(Crossbond

100% Dimethyl-polysiloxane) column.


The olen content of liquids was determined by an infrared
technique with a SHIMADZU IR-470 type spectrometer (res-
olution: 2.7 cm
1
, illumination: Globar light, monochromator:
Littrow prism, detector: KRS 5 type detector with a pyrometer
in crystal window vacuum cell) in the 400e4000 cm
1
wave-
number range.
The distribution of olenic double bonds was also obtained
by a method based on infrared technique in the 400e4000
cm
1
wavenumber range. The vinyl double bonds have two
signicant IR absorption bands at 910 and 990 cm
1
while
vinylidene and internal double bonds absorb at 890 and
956 cm
1
. The numbers of eCH
2
e and eCH
3
groups were
determined also by IR technique in the range of 2800e
3100 cm
1
, i.e., in the CeH stretching vibration range of hy-
drocarbons. The number of the eCH
2
e groups is proportional
to the intensity of the asymmetric stretching vibration band at
2927 cm
1
and the number of the eCH
3
groups to that of
asymmetric stretching vibration band at 2958 cm
1
.
Research octane number, motor octane number and cetane
number were determined according to the infrared spectra
with a ZX-101c type instrument.
3. Results
3.1. Yields
Fig. 3a and b shows the yields of the cracking of a mixture
of waste HDPE and PS. It was found that both gas and liquid
yields increased with increasing temperature of degradation,
and nearly complete cracking could be attained at 450

C.
Both the equilibrium FCC catalyst and the natural
Table 1
The main properties of uncrushed and crushed catalysts
Properties Clin. FCC ZSM-5
a. Uncrushed catalysts
Average grain size, mm 32.3 59.7 12.4
Si/Al 5.67 16.4 28.10
BET surface area, m
2
/g 72.6 148.1 300.1
Micropores, m
2
/g 1.3 78.9 225.3
b. Crushed catalysts
Average grain size, mm 1.7 1.5 1.8
Si/Al 5.67 16.4 28.10
BET surface area, m
2
/g 77.4 146.9 297.4
Micropores, m
2
/g 2.1 75.5 217.3
0
25
50
75
100
0,1 1 10 100 1000
Pore size, m
C
u
m
u
l
a
t
i
v
e

d
i
s
t
r
i
b
u
t
i
o
n
,

%
Clin.
FCC
ZSM-5
Clin.*
FCC*
ZSM-5*
Fig. 1. Cumulative distribution of average grain size of catalysts.
519 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
clinoptilolite zeolite catalyst had good catalytic activity for
producing light hydrocarbon liquids, and ZSM-5 catalyst pro-
duced the highest amount of gaseous products. It was ex-
plained with the greater microporous surface area of ZSM-5.
The greater microporous surface area of ZSM-5 catalyst
caused much higher catalyst activity than the other catalysts,
which have lower microporous surface areas; primary crack-
ing reactions of the polymer chain may proceed on the macro-
porous surface of the catalyst, while the smaller fragments are
cracked on their microporous surface.
The yields of both gas and liquid products are affected by
the grain size of catalysts. Smaller average catalyst grain
size resulted in signicantly higher yields of volatile fractions
(gases, liquids) because of greater surface of catalysts grains.
It is well known that in the cracking of CeC bonds the amor-
phous sites of catalysts play an important role. Not only the
presence of catalysts, but also the temperature affects the
yields. The growing yield of volatile components as function
of temperature could be caused by the differences in the ther-
mal stability of polymer chains, because hydrocarbons have
decreasing thermal stability with increasing temperature.
Therefore the CeC bonds cracked more likely at 450

C
than at lower temperature, e.g. at 410

C, and it resulted in
higher yields of volatile products.
3.2. Gases
The composition of gases formed in cracking reactions is
given in Table 2 aec. It can be stated that C
2
, and C
4
fractions
are dominating mainly in gases on the cracking of polyethyl-
ene. The temperature also affected the composition of gases,
because it was found that the concentration of products of
HDPE cracking (e.g. C
2
, and C
4
hydrocarbons) increased,
while that of methane decreased with increasing temperature.
This phenomenon is explained with the difference of activation
energies of polymers. The degradation of polystyrene has
lower activation energy than that of high-density polyethylene,
therefore at lower temperature the ratio of cracking of
polystyrene is greater than that of the other polymer in the
mixture. By increasing the temperature the ratio of cracking
0
25
50
75
100
Y
i
e
l
d
s

o
f

p
r
o
d
u
c
t
s
,

%
Non-cat. Clin. FCC ZSM-5
Gases
Liquids
Residue
Gases
Liquids
Residue
0
25
50
75
100
Y
i
e
l
d
s

o
f

p
r
o
d
u
c
t
s
,

%
Non-cat. Clin. FCC ZSM-5
(a)
(b)
Fig. 3. Yields of products (a) over uncrushedand (b) crushed catalysts at 430

C.
Residue
Liquids
Gases
Water
N
I
T
R
O
G
E
N
Naphtha-like
fraction
Diesel oil-like
fraction
1
2
3
4
5
6
7
Furher separation of liquids
into fuel-like fractions
Fig. 2. Apparatus for cracking of mixtures of plastic wastes. 1, Rotameter; 2, reactor; 3, thermometer; 4, condenser; 5, separator; 6, gas-ow meter; and 7,
distillation.
520 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
of polyethylene is increased, while that of polystyrene de-
creased. Therefore the cracking products of polystyrene were
less dominant at higher temperature. In accordance with liter-
ature [12] the concentration of alkanes was higher than that of
alkenes. The concentration of i-C4 was higher in cracking of
polymer at lower temperature (410 and 430

C), than at higher
temperature (450

C), because perhaps the isomerisation func-
tion of catalysts decreased due to their coking. The cracking of
waste polymers over ZSM-5 resulted the highest concentration
of gaseous branched hydrocarbons. This effect was more sig-
nicant in case of cracking over uncrushed catalysts.
3.3. Liquids
The composition of liquid products was determined by gas
chromatography. Results are shown in Table 3aec. The liquids
contained parafns, olens and aromatic hydrocarbons. Some
differences were observed in the composition of liquids ob-
tained by thermo-catalytic cracking over catalysts with differ-
ent grain size, but these differences were not signicant. On
the other hand the temperature had a signicant effect on
the composition of liquids. When comparing the data of Table
3aec it was observed that the yields of aromatics signicantly
decreased, while the yields of both olens and parafns in-
creased as function of temperature due to higher ratio of crack-
ing of polyethylene. Cracking of polystyrene results aromatics,
while that of polyethylene aliphatic compounds at mild tem-
perature (!460

C).
Table 2
Composition of gases
Non-catalytic FCC FCC
a
ZSM-5 ZSM-5
a
Clin. Clin.
a
a. At 410

C
Methane 23.95 20.46 19.91 19.76 20.32 20.84 20.24
Ethene 18.87 19.62 19.91 19.51 18.76 19.73 20.56
Ethane 25.29 27.02 27.09 27.67 26.18 24.79 25.50
Propene 4.79 5.32 5.73 6.31 6.64 5.92 5.98
Propane 8.05 6.65 6.37 6.98 9.69 8.15 6.38
Butene 9.00 9.98 9.56 9.42 7.81 8.06 9.56
Butane 10.06 9.40 8.58 8.51 7.61 9.60 8.03
i-Butane 0.00 1.55 2.85 1.83 2.99 2.91 3.76
b. At 430

C
Methane 17.34 15.32 14.46 14.97 13.64 14.77 13.59
Ethene 20.51 22.98 20.96 21.71 20.38 22.53 21.53
Ethane 27.48 29.86 29.82 29.70 27.79 26.76 27.82
Propene 5.21 6.15 5.97 5.57 7.49 6.89 6.46
Propane 8.75 8.06 6.82 7.42 10.11 8.74 7.44
Butene 9.78 8.14 9.37 10.03 8.84 8.00 9.87
Butane 10.93 8.43 9.69 8.67 8.38 9.66 8.37
i-Butane 0.00 1.07 2.92 1.92 3.36 2.64 4.91
c. At 450

C
Methane 11.64 10.94 10.49 10.66 0.71 11.37 11.08
Ethene 21.92 24.62 22.60 23.48 20.64 24.15 22.79
Ethane 29.38 31.98 31.84 31.65 28.99 29.99 30.22
Propene 5.56 6.20 6.68 6.53 7.91 6.44 6.74
Propane 9.35 8.43 7.19 8.17 10.11 8.75 6.94
Butene 10.46 7.75 10.28 9.19 9.67 8.05 9.90
Butane 11.69 8.72 8.68 8.17 8.79 9.05 8.92
i-Butane 0.00 1.35 2.24 2.14 3.19 2.20 3.41
a
Crashed catalysts.
Table 3
Composition of liquids
Non-
catalytic
FCC FCC
a
ZSM-5 ZSM-5
a
Clin. Clin.
a
a. At 410

C
Olen content, % 4.27 10.92 10.72 9.15 12.10 10.74 13.27
Vinyl 3.81 5.62 5.38 4.63 5.67 5.45 5.93
Vinylidene 0.11 0.44 0.50 0.35 0.62 0.23 0.40
Internal 0.35 4.88 4.84 4.17 5.81 5.06 6.94
Parafn content, % 4.73 9.84 10.99 8.14 11.63 8.44 13.54
n-Parafns 4.73 7.11 7.84 4.56 7.46 6.40 10.73
i-Parafns 0.00 2.73 3.15 3.58 4.17 2.04 2.81
Aromatics
content, %
91.00 79.24 78.29 82.71 76.27 80.82 73.19
Benzene 1.00 1.88 0.74 1.07 0.81 0.99 0.77
Toluene 3.69 4.96 4.00 4.86 4.03 6.30 3.86
Ethyl-benzene 19.03 15.94 17.54 17.77 15.10 15.68 14.49
Styrene 53.39 45.48 45.13 46.72 45.41 46.96 43.57
Xylenes 0.67 0.06 0.09 0.60 0.09 0.10 0.09
Isopropyl-benzene 3.57 4.58 4.50 3.41 4.53 3.69 4.35
a-Methylstyrene 1.98 2.59 2.59 2.70 2.61 2.41 2.50
C
10
eC
15
0.83 1.57 1.50 0.74 1.51 1.03 1.45
C
16C
6.84 2.18 2.18 4.84 2.19 3.66 2.11
b. At 430

C
Olen content, % 35.86 41.44 40.72 37.86 40.94 41.37 40.89
Vinyl 27.39 14.55 15.85 13.11 16.20 10.60 13.08
Vinylidene 4.49 12.23 12.91 10.24 12.02 4.48 5.70
Internal 3.98 37.08 34.45 32.21 35.16 28.32 30.51
Parafn content, % 37.77 39.96 40.83 36.22 40.96 39.49 40.13
n-Parafns 37.77 32.77 32.28 27.28 32.13 31.03 31.39
i-Parafns 0.00 7.19 8.55 8.94 8.83 8.46 8.74
Aromatics
content, %
26.37 18.6 18.45 25.92 18.1 19.14 18.98
Benzene 0.23 0.37 0.49 0.44 0.53 0.30 0.53
Toluene 2.19 1.28 0.70 1.78 0.75 1.63 0.76
Ethyl-benzene 1.27 3.98 4.43 5.80 3.01 3.81 4.58
Styrene 18.64 10.56 10.62 14.72 11.43 11.03 10.68
Xylenes 0.05 0.02 0.06 0.23 0.07 0.02 0.07
Isopropyl-benzene 1.37 0.91 0.56 1.27 0.60 0.93 0.61
a-Methylstyrene 0.98 0.59 0.41 0.84 0.44 0.42 0.45
C
10
eC
15
0.67 0.20 0.70 0.17 0.75 0.30 0.76
C
16C
0.97 0.69 0.48 0.67 0.52 0.72 0.52
c. At 450

C
Olen content, % 42.02 43.70 45.75 41.67 47.69 47.22 45.45
Vinyl 32.33 18.49 16.73 17.69 14.93 12.35 16.17
Vinylidene 5.11 14.93 14.28 13.88 11.97 4.96 6.30
Internal 4.58 42.77 36.92 40.18 33.55 29.93 33.52
Parafn content, % 46.08 46.03 45.80 43.62 43.15 42.90 43.37
n-Parafns 46.08 41.88 41.81 38.83 37.95 38.37 38.63
i-Parafns 0.00 4.15 3.99 4.79 5.20 4.53 4.74
Aromatics
content, %
11.9 10.27 8.45 14.71 9.16 9.88 11.18
Benzene 0.21 0.17 0.22 0.27 0.21 0.12 0.21
Toluene 0.67 0.74 0.72 1.01 0.71 0.95 0.70
Ethyl-benzene 0.99 2.33 2.02 3.44 1.84 2.12 2.81
Styrene 8.19 5.85 4.34 8.43 5.26 5.52 6.33
Xylenes 0.07 0.01 0.02 0.14 0.02 0.01 0.02
Isopropyl-benzene 0.34 0.44 0.43 0.63 0.43 0.33 0.42
a-Methylstyrene 0.61 0.37 0.27 0.53 0.26 0.25 0.26
C
10
eC
15
0.18 0.11 0.26 0.14 0.26 0.24 0.25
C
16C
0.64 0.25 0.18 0.12 0.18 0.34 0.17
a
Crashed catalysts.
521 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
Table 3aec also shows the concentration of branched
hydrocarbons at different temperatures. The formation of
branched hydrocarbons was not observed without catalysts,
only in thermo-catalytic cases. It was the consequence of the
well known phenomenon that the carbon-chain isomerisation
takes place only in presence of catalysts through carbenium
ions. As data in table show, the yields of branched hydrocar-
bons decreased with increasing temperature. Carbon-chain
isomerisation activity of ZSM-5 catalyst was more signicant,
at lower temperature (e.g. 410

C) than that of the other cata-
lysts (FCC or Clin.). This difference can be attributed to the
higher micro- and macroporous surface area of the ZSM-5 cat-
alyst but the difference diminished with increasing tempera-
ture. The concentration of branched compounds did not
change signicantly with smaller average grain size of cata-
lysts, but it was a little higher when crushed catalysts were
used. According to experimental results branched hydrocar-
bons ranged between C
8
and C
13
, and the maximum amount
of them was between C
10
and C
12
. It is advantageous for fur-
ther utilization of products obtained by cracking of polymer
wastes, because branched hydrocarbons in naphtha-like frac-
tions have high octane numbers.
The olen content of liquids is also an important property
for further utilization. The olen content and its distribution
in liquids were investigated by infrared technique in the range
of 1200e400 cm
1
wavenumber. In addition to the isomerisa-
tion of carbon-chain signicant double bond isomerisation was
observed. It was found that a-olens were the dominating type
of olens in thermal degradation, while the internal ones were
in thermo-catalytic cases due to the shifting of terminal double
bonds to internal position on catalysts.
The distribution of carbon-chain of liquids is shown in
Fig. 4aec. The carbon-chain distribution was considerably
affected by catalysts, because the range of the carbon atom
number of liquid fraction became narrower by using catalysts.
Not only the presence of catalysts affected the yields of hydro-
carbons containing different length of carbon-chain, but also
the temperature. The distribution of carbon-chain became wid-
er with increasing temperature. As Fig. 2 indicates that crack-
ing reactions were carried out in a batch reactor and the whole
process was similar to a reactive distillation. It means that in
case of higher temperature the driving force of distillation
was also higher due to greater difference of temperature be-
tween the melted polymer and the withdrawal point of prod-
ucts. In these circumstances the heavier fraction can be
distilled from the molten polymer. On the other hand the
between thermal and thermo-catalytic cases were less notable
at higher temperature owing to the decreasing cracking activ-
ity of the catalysts. It could be the consequence of the coking
of the catalyst surfaces at higher temperature, which might
block the active sites of catalysts.
There was no signicant difference observed with respect
to the average grain diameter of catalysts. Although ZSM-5
catalysts had greater cracking activity than other catalysts.
There is considerably content of C
8
hydrocarbons in case of
each Table. This is a consequence of the formation of styrene
and a little ethyl-benzene mainly from the cracking of
polystyrene. However, the concentration of C
8
hydrocarbons
decreased with increasing temperature. By increasing the tem-
perature of cracking the probability of scission of CeC bonds
of polymers got higher. Nevertheless the concentration of
polystyrene in the feed polymers was only 10%, and as it
was above demonstrated the waste polystyrene can be con-
verted into styrene and ethyl-benzene with high conversion.
It means that the concentration of C
8
hydrocarbons is about
10% in liquids in case of total conversion of cracking. This
is the reason that the C
8
hydrocarbon content of liquid prod-
ucts converged to the percentage of polystyrene in the mixture
of raw polymers.
In the case of catalytic cracking over catalysts the concen-
tration of heavy aliphatic compounds in the liquids (C
20C
) was
0
3
6
9
12
15
Carbon chain
C
o
m
p
o
s
i
t
i
o
n
,

%
5 10 15 20 25 30
Non-cat.
FCC
ZSM-5
Clin.
FCC*
ZSM-5*
Clin.*
Non-cat.
FCC
ZSM-5
Clin.
FCC*
ZSM-5*
Clin.*
Non-cat.
FCC
ZSM-5
Clin.
FCC*
ZSM-5*
Clin.*
0
3
6
9
12
15
Carbon chain
C
o
m
p
o
s
i
t
i
o
n
,

%
5 10 15 20 25 30
0
3
6
9
12
15
Carbon chain
C
o
m
p
o
s
i
t
i
o
n
,

%
5 10 15 20 25 30
(a)
(b)
(c)
Fig. 4. Distribution of liquid fractions at (a) 410

C, (b) 430

C, and (c)
450

C.
522 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
signicantly lower than in case of pure thermal cracking. It
was observed that the concentration of lighter (C
13
) hydro-
carbons increased in all cases by using catalysts, while the
concentration of heavier (C
13C
) alkene and alkane compounds
decreased at 430

C. However, the maximum carbon number
of distribution curve of aliphatic hydrocarbons increased
with increasing temperature (e.g. at 450

C the maximum of
the curve was at C
18
).
3.4. Reaction kinetics
The kinetics of cracking reactions of waste polymers was
investigated by many researchers. It was found that the degra-
dation of polyolens could be modelled by using a rst-order
kinetic equation [25,26].
The degradation of polystyrene consisted of two steps:
cracking of polymer chain into mainly monomer, ethyl-
benzene and oligomers, and splitting-off substituents from
the aromatic rings. The rst step of degradation is 1.65th or-
der, but the second is 0.76th order [22]. According to the lit-
erature the rate-controlling step is the cracking of long
chains. In the present work 90% of high-density polyethylene
and 10% polystyrene were used, therefore the apparent crack-
ing kinetics was modelled as rst order. The calculated appar-
ent activation energies are in Table 4. By the milling of
catalysts their cracking efciency became better and the ap-
parent activation energy of degradation could be decreased.
As it has been mentioned earlier the degradation of polyo-
lens could be modelled by using rst-order kinetic equation.
Accordingly the activation energies of cracking reactions were
calculated by Eqs. (1) and (2). The reaction rate coefcients
were derived from Eq. (1) with n Z1.

dW
dt
ZkW
n
1
The activation energies were calculated by using the Arrhe-
nius equation:
kZA
0
exp

E
RT

2
Fig. 5 represents the weight of volatile products as function
of cracking time. Integrating Eq. (1), the conversion (c) as
function of time (t) can be expressed by ln(1c) Zkt (Fig. 6).
The reaction rate coefcients were obtained from Fig. 6. It
was found that they increased with the cracking temperatures,
and were higher in the presence of catalysts. From the temper-
ature dependence of the reaction rate constants phenomeno-
logical activation energies were calculated (Table 4). In the
presence of catalysts the average activation energies were
lower than the activation energy of thermal reactions. The
activation energies of thermo-catalytic degradation could be
decreased also by decreasing the average grain size of the
catalysts. The lowest value was found with FCC catalyst.
3.5. Further application of products
One prospective possibility of the further utilization of
liquids formed in the cracking reactions is a fuel-like utiliza-
tion. Therefore liquids obtained in the cracking reactions
were measured and analyzed also as fuel-like products. For
the sake of investigation of this possibility, after thermal and
thermo-catalytic degradation the liquid products from the mix-
ture of waste polymers were separated with distillation into
naphtha-like (b.p. 35e210

C) and diesel oil-like (b.p. 85e
305

C) fractions. Fig. 7a and b shows the gas chromatograms
of naphtha- and diesel oil-like fractions in case of non-catalytic
cracking at 450

C. There are lots of peaks of alkenes and
Table 4
The apparent activation energies
Catalysts E
a
, kJ/mol
None 255
FCC 218
FCC
a
196
ZSM-5 217
ZSM-5
a
209
Clin. 224
Clin.
a
203
a
Crashed catalysts.
0,0
20,0
40,0
60,0
80,0
0 10 20 30 40 50 60 70
Time of degradation, min.
Y
i
e
l
d
s

o
f

v
o
l
a
t
i
l
e

p
r
o
d
u
c
t
s
,

%
Non-cat.
Clin.
FCC
ZSM-5
Fig. 5. The weight of volatile products as function of cracking time using
uncrushed catalysts (at 430

C).
Cracking time, min.
-1,50
-1,00
-0,50
0,00
0 10 20 30 40 50 60
l
n
(
m
/
m
0
)
Clin.
FCC
ZSM-5
Clin.*
FCC*
ZSM-5*
Fig. 6. The ln(1 c) as function of cracking time in case of thermo-catalytic
cracking (430

C).
523 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
alkanes, and aromatics in Fig. 7a, because the cracking reac-
tions of polyethylene resulted great number of hydrocarbons e
alkenes and alkanes e while polystyrene different light
aromatics at moderate conditions.
Table 5a and b presents the main properties of each distil-
late fraction. According to data in Table 5a the naphtha-
like fractions contained C
5
eC
17
hydrocarbons, but mainly
C
8
eC
10
compounds were in the highest concentration in
liquids. Each naphtha-like fraction contained about one third
of aromatics owing to cracking of polystyrene. Due to aro-
matics in naphtha-like fractions their octane numbers were
high. Liquid properties are considerably changed in the
presence of catalysts. Naphtha-like liquids obtained by
thermo-catalytic cracking of polymer wastes had more favour-
able properties for further application, than those obtained by
thermal cracking e.g. higher octane numbers, which can be
due to the greater branched hydrocarbon content. Branched
compounds have moderately high octane numbers and quite
small sensibility. However, neither the type of catalysts nor
their grain size signicantly affected the properties of
products.
Each raw polymer contained sulphur and nitrogen atoms
from different additives (e.g. ame retardant, antioxidants or
antifuming, etc.). The lighter liquid fractions also had sulphur
and nitrogen content, but in very low concentration. The con-
centration of the sulphur and the nitrogen were not more than
15 ppm. It is an important property for their further utilization,
because there is a great ambition to reduce the heteroatom
content (mainly sulphur) in reneries. Furthermore these
processes are quite expensive.
In contrast to the experimental results of naphtha-like frac-
tions, the heavier liquid fractions had considerably lower con-
centration of aromatics. As data show aromatics stayed in
naphtha-like fractions and their concentration was not more
Retention time, min.
Retention time, min.
5 10
C
5
H
10 C
6
H
12
CH
3
CH
3
CH
2
HC
CH
2
C
7
H
14
C
7
H
16
C
8
H
16
C
9
H
18
C
9
H
20
C
10
H
20
C
10
H
22
C
11
H
22
C
12
H
26
C
12
H
24
C
13
H
28
C
13
H
26
C
11
H
24
H
3
C CH
3
CH
2
CH
3
HC
HC
C
8
H
18
C
6
H
14
C
5
H
12
C
11
H
24
C
11
H
22
C
12
H
26
C
12
H
24
C
13
H
28
C
13
H
26
C
14
H
30
C
14
H
28
C
15
H
32
C
15
H
30
C
16
H
34
C
16
H
32
C
17
H
36
C
17
H
34
C
18
H
38
C
18
H
36
C
19
H
40
C
19
H
38
C
20
H
42
C
21
H
42
C
21
H
44
C
20
H
40
C
22
H
46
C
22
H
44
C
23
H
48
C
23
H
46
C
24
H
48
C
24
H
50
C
25
H
50
C
25
H
52
C
26
H
56
C
27
H
54
C
26
H
54
C
26
H
52
xylenes
U,mV
5 10 15 20
Trimer U, mV
(a)
(b)
Fig. 7. The gas chromatogram of (a) naphtha-like and (b) diesel oil-like fractions in case of non-catalytic cracking at 450

C.
524 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
Table 5
Composition and main properties
Non-catalytic FCC FCC
a
ZSM-5 ZSM-5
a
Clin. Clin.
a
a. Of naphtha-like fractions
Yield, % 48.8 59.4 61.6 62.4 63.1 59.7 60.5
Aliphatic content, % 54.39 61.93 64.37 57.44 58.23 62.50 62.82
C
5
eC
7
11.51 20.21 20.68 15.72 14.28 19.97 18.28
C
8
eC
10
26.34 29.60 30.59 28.09 28.24 29.41 29.37
C
11
eC
13
15.09 12.02 13.00 13.38 15.60 13.02 15.06
C
14
eC
16
0.89 0.11 0.11 0.25 0.12 0.11 0.11
C
16C
0.57 0.00 0.00 0.00 0.00 0.00 0.00
Aromatics content, % 45.61 38.07 35.63 42.56 41.77 37.50 37.18
Benzene 0.38 0.28 0.25 0.24 0.35 0.26 0.33
Toluene 3.90 2.34 3.05 4.11 3.60 3.15 3.34
Ethyl-benzene 3.25 2.62 2.70 3.82 3.05 3.84 2.10
Styrene 28.82 28.55 26.50 29.88 29.39 26.03 26.8
Xylenes 0.56 0.35 0.26 0.43 0.30 0.37 0.2
Isopropyl-benzene 0.74 0.60 0.42 0.37 1.15 0.59 056
a-Methylstyrene 2.01 1.64 1.36 1.71 1.70 1.63 0.02
C
10
eC
15
1.42 1.68 1.08 2.00 2.23 1.63 1.94
C
16C
4.53 0.00 0.00 0.00 0.00 0.00 0.00
Density, g/cm
3
0.752 0.730 0.729 0.731 0.734 0.728 0.730
ROM 78 76 75 77 78 76 76
MON 85 84 84 85 85 84 84
(ROMCMON)/2 82 80 80 81 82 80 80
Corrosion test Group 1 Group 1 Group 1 Group 1 Group 1 Group 1 Group 1
Distillation data,

C
IBP 38 35 34 37 33 34 35
20 63 64 67 64 65 63 66
50 135 133 132 137 134 133 134
70 145 149 147 144 146 143 142
FBP 209 207 206 205 201 204 208
Sulphur content, ppm 14 11 13 10 14 9 12
Nitrogen content, ppm 8 6 9 5 9 10 8
M, g/mol 119 102 98 105 96 104 103
C/H 6.0 5.8 5.9 5.9 5.8 5.9 5.9
Other impurities None None None None None None None
b. Of diesel oil-like fractions
Yield. % 51.2 40.6 38.4 37.6 36.9 40.3 39.5
Aliphatic content, % 98.48 99.02 99.97 99.48 99.71 99.47 99.82
C
11
eC
13
11.19 15.95 15.53 16.22 15.48 14.70 14.91
C
14
eC
16
35.59 39.06 40.50 41.24 40.95 39.88 40.01
C
17
eC
19
27.51 28.49 29.69 29.40 29.76 28.93 29.14
C
20
eC
22
13.52 11.00 10.85 10.69 10.06 10.17 10.09
C
23
eC
25
7.34 3.26 2.20 0.58 2.00 3.94 4.51
C
26C
3.33 1.26 1.20 1.35 1.46 1.85 1.16
Aromatics content, % 1.52 0.98 0.03 0.52 0.29 0.53 0.18
Density, g/cm
3
0.823 0.815 0.808 0.801 0.796 0.816 0.811
Cetane number 64 63 63 63 63 64 63
Diesel index 69 69 69 69 69 68 68
Viscosity, mm
2
/s 4.4 4.5 4.4 4.4 4.3 4.3 4.4
CFPP,

C 1 4 5 7 8 5 4
Flash point,

C 92 85 84 83 83 84 85
Corrosion test Group 1 Group 1 Group 1 Group 1 Group 1 Group 1 Group 1
Distillation data,

C
IBP 85 88 84 83 80 81 82
20 163 160 158 162 163 159 164
50 195 194 195 198 193 195 197
70 214 219 218 218 216 215 214
FBP 305 302 298 304 301 298 301
Sulphur content, ppm 15 14 12 11 14 13 10
Nitrogen content, ppm 10 9 11 8 9 10 12
M, g/mol 255 247 240 240 238 247 247
C/H 6.2 6.2 6.2 6.2 6.2 6.2 6.2
Other impurities None None None None None None None
a
Crashed catalysts.
525 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526
than 1% in diesel oil due to the low boiling point of aromatics.
Similarly as it was mentioned in case of lighter liquid frac-
tions, only the presence of catalysts affected the properties
of diesel oil-like liquids and neither the type of catalysts nor
their grain size affected signicantly the properties of products
in case of thermo-catalytic cracking. Both cetane numbers and
diesel indices of products were high enough, while the CFPP
was rather moderately low.
4. Conclusions
Thermal and thermo-catalytic degradation of a mixture of
90% of high-density polyethylene (HDPE) and 10% of poly-
styrene (PS) wastes was investigated in a temperature range
of 410e450

C in Pyrex batch reactor under nitrogen atmo-
sphere. It was found that:
-the mixture of wastes could be converted into lighter frac-
tions with yields of 16e89% depending on temperature
and catalysts,
-the cracking temperature and catalyst grain size signi-
cantly affected the yields, but gas yields depended mainly
on the type of catalyst,
-the activation energy of the overall conversion could be
decreased by about 40 kJ/mol by using catalyst,
-both olen double bond and carbon-chain isomerisation
were observed in the presence of zeolite catalysts,
-liquid hydrocarbon products are of low sulphur content
therefore they may be favourable for fuel production.
Acknowledgement
The authors gratefully acknowledge the nancial support
for this work of the Chemical Engineering Institutes Cooper-
ative Research Centre (especially Ministry of Education of
Hungary and the Hungarian Oil and Gas (MOL) PLC).
References
[1] Kim JR, Yoon JH, Park DW. Catalytic recycling of the mixture of poly-
propylene and polystyrene. Polymer Degradation and Stability 2002;
76:61e7.
[2] Luo K, Suto T, Yasu S, Kato K. Catalytic degradation of high density
polyethylene and polypropylene into liquid fuel in a powder-particle
uidized bed. Polymer Degradation and Stability 2000;70:97e102.
[3] Serrano DP, Aguado J, Escola JM, Garagorri E. Conversion of low
density polyethylene into petrochemical feedstocks using a continuous
screw kiln reactor. Journal of Analytical and Applied Pyrolysis 2001;
58e59:789e801.
[4] Hwang EY, Kim YR, Choi JK, Woo HY, Park DW. Performance of acid
treated natural zeolites in catalytic degradation of polypropylene. Journal
of Analytical and Applied Pyrolysis 2002;62:351e64.
[5] Walendziewski J, Steininger M. Thermal and catalytic conversion of
waste polyolens. Catalysis Today 2001;65:323e30.
[6] Uddin A, Koizumi K, Murata K, Sakata Y. Thermal and catalytic degra-
dation of structurally different types of polyethylene into fuel oil.
Polymer Degradation and Stability 1997;56:37e44.
[7] Breen C, Last PM, Taylor S, Komadel P. Synergic chemical analysis e
the coupling of TG with FTIR, MS and GCeMS 2. Catalytic transfor-
mation of the gases evolved during the thermal decomposition of
HDPE using acid-activated clays. Thermochimica Acta 2000;363:
93e104.
[8] Miranda R, Yang J, Roy C, Vasile C. Vacuum pyrolysis of commingled
plastics containing PVC I. Kinetic study. Polymer Degradation and
Stability 2001;72:469e91.
[9] Ukei H, Hirose T, Horikawa T, Takai Y, Taka M, Azuma N, et al.
Catalytic degradation of polystyrene into styrene and a design of
recyclable polystyrene, with dispersed catalysts. Catalysis Today
2000;62:67e75.
[10] Murata K, Hirano Y, Sakata Y, Uddin A. Basic study on a continous ow
reactor for thermal degradation of polymers. Journal of Analytical and
Applied Pyrolysis 2002;65:71e90.
[11] Ali S, Garforth AA, Harris DH, Rawlence DY, Uemichi Y. Polymer waste
recycling over used catalysts. Catalysis Today 2002;75:247e55.
[12] Grieken R, Serrano DP, Aguado J, Garcia R, Rojo C. Thermal and catalytic
cracking of polyethylene under mild conditions. Journal of Analytical and
Applied Pyrolysis 2001;58e59:127e42.
[13] Karaduman A, Sxims xek EH, C icek B, Bilgesu AY. Flash pyrolysis of
polystyrene wastes in a free-fall reactor under vacuum. Journal of
Analytical and Applied Pyrolysis 2001;60:179e86.
[14] Karaduman A, Sxims xek EH, C icek B, Bilgesu AY. Thermal degradation of
polystyrene wastes in various solvents. Journal of Analytical and Applied
Pyrolysis 2002;62:273e80.
[15] Gao Z, Amasaki I, Kaneko T, Nakada M. Calculation of activation energy
from fraction of bonds broken for thermal degradation of polyethylene.
Polymer Degradation and Stability 2003;81:125e30.
[16] Faravelli T, Bozzano G, Colombo M, Ranzi E, Dente M. Kinetic model-
ing of the thermal degradation of polyethylene and polystyrene mixtures.
Journal of Analytical and Applied Pyrolysis 2003;70:761e77.
[17] Faravelli T, Pinciroli M, Pisano F, Bozzano G, Dente M, Ranzi E.
Thermal degradation of polystyrene. Journal of Analytical and Applied
Pyrolysis 2001;60:103e21.
[18] Seddegi ZS, Budrthumal U, Al-Arfaj AA, Al-Amer AM, Barri SAI.
Catalytic cracking of polyethylene over all-silica MCM-41 molecular
sieve. Applied Catalysis A: General 2002;225:167e76.
[19] Takuma K, Uemichi Y, Ayame A. Product distribution from catalytic
degradation of polyethylene over H-gallosilicate. Applied Catalysis A:
General 2000;(192):273e80.
[20] Serrano DP, Aguado J, Escola JM. Catalytic conversion of polystyrene
over HMCM-41, HZSM-5 and amorphous SiO
2
eAl
2
O
3
: comparison
with thermal cracking. Applied Catalysis B: Environmental 2000;25:
181e9.
[21] Lee KH, Noha NS, Shina DH, Seo Y. Comparison of plastic types for
catalytic degradation of waste plastics into liquid product with spent
FCC catalyst. Polymer Degradation and Stability 2002;78:539e44.
[22] Pinto F, Costa P, Gulyurtulu I, Cabrita I. Pyrolysis of plastic waste 2.
Effect of catalyst on product yield. Journal of Analytical and Applied
Pyrolysis 1999;51:57e71.
[23] Songip AR, Masuda T, Kuwahara H, Hashimoto K. Kinetic studies for
catalytic cracking of heavy oil from waste plastics over REY zeolite.
Energy Fuels 1994;8:131e40.
[24] Woo HC, Lee KH, Lee JS. Catalytic skeletal isomerization of n-butenes
to isobutene over natural clinoptilolite zeolite. Applied Catalysis A:
General 1996;134:147e58.
[25] Manchado LM. Kinetic analysis of the thermal degradation of PP-EPDM
blends. Rubber Chemistry and Technology 2000;73:694e705.
[26] Yang J, Miranda R, Roy C. Using the DTG curve tting
method to determine the apparent kinetic parameters of thermal decom-
position of polymers. Polymer Degradation and Stability 2001;73:
455e61.
526 N. Miskolczi et al. / Polymer Degradation and Stability 91 (2006) 517e526

Das könnte Ihnen auch gefallen