Anda di halaman 1dari 13

JOURNALOI

ANALYTICAL and
APPLIED PYROLYSIS
Journal of Analytical and Applied Pyrolysis
ELSEVIER 38 (1996) 75-87

Pyrolysis of polyolefins with steam to yield olefins


C.M. Simon*, W. Kaminsky, B. Schlesselmann
L’nit.rr.sitj~of Hamburg. Institute jbr Technical and Macromolwular Chemistry, Bunde.wtr& 45.
D-20146 Hamburg, Germanic

Received I April 1996: accepted 5 July 1996

Abstract

The polyolefinic fraction of post consumer plastics waste was pyrolyzed in a fluidized bed
lab scale reactor. In order to improve olefin yields, steam was used as the fluidizing agent.
The influence of the reaction temperature on olefin yields was investigated in the range
600-800°C. Furthermore, the residence time was varied at a given temperature of 700°C
between 1.8 and 3.2 s.
The product distribution shows a characteristic influence of the cracking temperature on
the yields of the different components. High amounts of olefins are obtained at temperatures
around 7OO”C, with 20-31 wt.% ethene, 14-18 wt.‘% propene. 3-6 wt.‘% butenes and.
additionally, 19-23 wt.‘% pyrolysis gasoline as the main products. These yields are similar to
the product distribution of naphtha cracked in a steam cracker, with 29 wt.‘%) ethene. 16
wt.‘% propene. 10 wt.% butenejbutadiene, and 24 wt.% pyrolysis gasoline.
After separation the olefins can be reused as monomers for the production of polyolefins.

Key\caor&ls: Fluidized bed reactor: Olefins: Polyolefins: Pyrolysis; Steam

1. Introduction

During the last few years the recycling of plastics has been an extremely
controversial issue which has been studied and examined in many industrial
countries, especially Germany. Whereas for glass or metals recycling possibilities
already existed, plastics were only recycled in the case of clean production wastes.
Under the influence of the German packaging decree, the plastics industry was
forced to come up with possibilities for recycling post-consumer plastic wastes.
During the last few years, different approaches have been followed by industry and

* Corresponding author.

0165-2370;96/$15.00 Copyright 0 1996 Elsevier Science B.V. All rights reserved


PII SOI65-2370(96)00950-3
16 C.M. Simon et al. /J. Anal. Appl. Pyrolysis 38 (1996) 75-87

scientific institutes. Good overviews can be found in Refs. [l-4]. Considering the
different petrochemical processes (pyrolysis, cat-cracking, visbreaking, coking, ga-
sification and hydrogenation) under evaluation for the recycling of hydrocarbon
backbone plastics, all processes involve a thermal cracking step. This article
describes the possibilities of thermal cracking in a fluidized bed reactor. The special
advantages of fluidized beds are summarized in detail by Kunii and Levenspiel [5].
The original aim of the Hamburg pyrolysis process was the production of a
petrochemical oil mainly consisting of benzene, toluene and xylene (BTX-aromat-
its) and a gas of high calorific value from plastic wastes [6,7].
Since 1991 other process variants have been investigated, such as
production of aliphatic oils/waxes at low pyrolysis temperatures for use as a feed
for petrochemical plants, e.g. a steam cracker [8,9],
direct conversion to olefins starting from polyolefins or mixed plastics [lo],
selective depolymerisation of the specific polymers poly(methylmethacrylate) [l l]
or polystyrene [12] to their monomers at low temperatures of 450-550°C.
For two reasons we concentrated our work on the conditions required to
maximize yields of ethene, propene and butenes out of polyolefinic waste. First,
polyolefins dominate the production and use of plastics. Due to their short life
application, polyolefins contribute most to the plastic waste stream. For Western
Europe, the portion of plastics in municipal solid waste can be divided into
different types (Table 1) [13].
The second reason is the relatively high value of these products. One concept of
integrating the plastic waste stream into an available petrochemical infrastructure
by the use of a pyrolysis process is shown in Fig. 1. Mixed plastic waste (MPW)
(polyethylene (PE), polypropylene (PP) and polystyrene (PS)) with a low
polyvinylchlorine (PVC) content can be pyrolyzed directly at temperatures of
700-800°C to produce large quantities of aromatics (40-50 wt.%) (BTX-process)
[14]. Even higher yields of more valuable products can be obtained by pyrolysis of
previously separated polyolefin and polystyrene fractions. Polystyrene can be
depolymerized at low temperatures to give styrene in yields of around 75 wt.% [15].
For the polyolefin fraction, there are two possibilities: the production of aliphatic
oils and waxes at low temperatures of 450-550°C [8] for further use as a feedstock
for olefin production, or one-step pyrolysis in an inert gas, such as nitrogen or
steam, to directly yield olefins in the temperature range of 650&75O”C.

Table 1
Composition of the plastics portion of municipal solid waste by type in Western Europe [13]

Component Amount (wt.%)

Polyolefins (PE, PP) 66.9


Polystyrene + extended polystyrene 13.3
Polyvinylchloride 10.3
Polyethyleneterephthate 5.3
Others 4.2

PE, polyethylene.
PP. polypropylene.
C.M. Simon et al. //)J. Anal. Appt. P?wlJsis 38 (1996) 75-87 71

Fig. 1. Integrating plastic waste into petrochemical:refinery infrastructures

The reported experiments were carried out in cooperation with DSM (former
Dutch State Mines, The Netherlands).

2. Experimental

For the experiments, a l-3 kg h ’ laboratory scale system at the University of


Hamburg was used (Fig. 2). The fluidized bed reactor, with a free diameter of 154
mm, was heated indirectly by burning propane in a separate concentric combustion
chamber. Steam was generated in three consecutive electrically heated vessels and
preheated in a coil up to temperatures close to 400°C. Before reaching the fluidized
bed, the steam was further heated to a temperature of 500°C. The plastic material
was feed to the reactor using a screw system, rapidly heated and converted to

Fig. 2. Flow scheme of the l-3 kg h ’ laboratory scale pyrolysis system


78 C.M. Simon et al. ! J. Anal. Appl. Pyrolysis 38 (1996) 75-87

Table 2
Parameters of the pyrolysis experiments

Experiment LF I LF 2 LF 3 LF 4 LF 5 LF 6 LF 7

Temperature fluidized bed (“C) 605 655 690 700 700 750 805
Temperature freeboard (“C) 605 655 685 695 700 755 805
Residence time reactor (s) 3.5 2.7 1.8 2.3 3.2 2.9 3.1

Input (kg) 3.9 5.0 5.2 5.6 4.7 4.9 4.6


Duration of input (min) 220 160 150 210 260 185 215
Feed rate (kg hh’) 1.1 1.9 2.1 1.6 1.1 1.6 1.3

Water throughput (kg/h) 2.8 3.2 4.1 3.0 2.5 2.4 2.1

Steam stream (m’ hh’) 10.4 11.7 18.4 13.5 10.2 10.2 9.6
Auxiliary nitrogen stream (m’ hh’) 1.2 1.0 0.8 0.8 1.0 1.0 1.0
Sum (m’ hh’) 11.6 12.7 19.2 14.3 11.2 11.2 10.6

Ratio water/plastic kg kg-’ 2.7 1.7 2.0 1.9 2.3 1.5 1.7

volatiles. The resulting pyrolysis gas/steam mixture passed a cooling and separation
unit. Dust, fine sand and soot were separated in a cyclone. Afterwards, tar and high
boiling substances were condensed and collected in the first cooler and a heated
(160°C) electrostatic precipitator. In subsequent coolers the water phase and the
low boiling pyrolysis oil were condensed. In the last coolers the gas was even cooled
to temperatures below - 10°C. A second electrostatic precipitator removed hydro-
carbon oil droplets from the pyrolysis gas.
The pressure was controlled manually and excess gas was pumped out of the
system by a compressor, then measured by a gas meter and burned in a flame.
Some parameters of the experiments are listed in Table 2.
Experimental tests (LF 1,2,5-7) were conducted in order to investigate the
influence of the reaction temperature. In combination with experiment LF 5, two
further experiments (LF 3 + 4) were carried out in order to test the influence of the
residence time in the hot reaction zone of the fluidized bed and freeboard zone. The
residence time in the reactor is the residence time of gaseous products in the free
volume of the fluidized bed and the hot freeboard. It was calculated assuming a
plug flow without back mixing of the gas, as is assumed in the two-phase model of
Werther [16].
The MPW was collected from households in Belgium. It was washed and reduced
in size to 5-10 mm before it was separated by density separation with water
(density: 1 .OO g cm - 3). The resulting light fraction, mainly composed of PE and PP,
was used as the input feed material.
The results of the averaged elementary analysis for the light fraction were, in %
by weight: C, 84.7; H, 14.0; Cl, 0.075; ash, 1.3; and C/H-ratio, 0.51. This indicates
that the mixture was mainly composed of PE and PP with only a small portion of
PS (3 wt.%) and PVC (theoretically 0.13 wt.%).
Prior to pyrolysis the material was extruded at temperatures of 210-220°C to
give a granulate product which was easier to handle. This recyclate had melting
C.M. Simon et al. )))J. Anal. Appl. Pyrol~~sis 38 (1996) 75- 87 19

flow indices (MFI) equal to 4.0 g 10 min- ’ (190°C. 5 kg) and 1.O g 10 min ’
(190°C 2 kg).
In all experiments, gases, water/oil-mixtures and tars, as well as solid residues.
were obtained as product fractions. Samples of the liquid pyrolysis products were
first separated using a separating funnel into a water and an oil phase fraction. A
sample of the liquid oil phase was distilled at reduced pressure (1.33 x lo4 Pa) up
to a temperature of 215”C, yielding an oil fraction and a distillation residue. These
values correspond to the boiling point (bp) of fluorene (bp = 295°C at normal
pressure). The products were analysed by a variety of methods, including gas
chromatography (GC), gas chromatography/mass spectroscopy (GCi’MS), elemen-
tary analysis, the Wickbold-method for chlorine analysis, the Karl-Fischer method,
and neutron-activation energy analysis (NAA). Pyrolysis gases were separated on
different packed columns and measured with TCD- and FID-detectors. For GC
analysis of the oils, a capillary column SE 52. 50 m (Macherey and Nagel) was
used.

3. Results and discussion

Conducted in each experiment, mass balances were close to 98 wt.‘%. Losses can
be attributed mainly to uncertainties in the calculation for the gas fraction from the
discontinuous analysis results and the gas volumes. These losses were mainly
distributed in the gas fraction. The product distribution of all fractions and some
components are given in Table 3. Product yields are reported as wt.% of the feed
on a moisture-free and ash-free basis. Apart from experiment LF 1, the gas fraction
is the main product (58-75 wt.%), having a maximum yield at a temperature of
around 700°C. The influence of the residence time on the product distribution is
only minor for the range studied.
Only in experiment LF 1 was the liquid fraction the main product (51 wt.%). In
all the other experiments, oils were produced with an l&24 wt.% yield. The
distillation residue shows an interesting dependency on the pyrolysis temperature.
At low and high reaction temperatures, these residues have high values ( 13- 19
wt.‘%,), whereas at moderate temperatures of 700- 750°C the residue is clearly
decreased. Short residence times seem to reduce the residues (LF 3-5). The soot
fraction shows no distinct changes with the temperature. It seems as if soot
formation is reduced by short residence times.

3. I. GUS ,fiwc.tion

In all experiments, the pyrolysis gas is rich in olefins. The highest gas yields are
produced between 650 and 750°C whereas lower and higher temperatures result in
less gas production. At temperatures between 650 and 700°C the gas fraction
resembled, in its composition (ethene 24431 wt.‘%,, propene 14-18 wt.% and
butenes 337 wt.%), the gas fraction produced by a steam cracker. Valuable
products could be separated and purified by these common processes. The remain-
ing methane could be used as an energy supply.
80 C.M. Simon et al. /J. Anal. Appl. Pyrolysis 38 (1996) 75-87

If, as in steam crackers, the obtained ethane is recirculated to the cracking zone,
the ethene yields could be increased by l-3 wt.%.
Carbon monoxide and carbon dioxide are formed in low quantities with a
positive temperature effect, as reported previously [lo], possibly because of in-
creased rates of soot gasification.
For the most important substances (ethene, propene, butadiene and n-/iso-
butene), the influence of temperature on their yields is plotted in Fig. 3.
Shorter residence times of the order investigated have no significant effects on the
yields of these products as can be seen from the data presented in Table 3.

Table 3
Product distribution of the pyrolysis experiments (refering to total organic input)

Experiment LF 1 LF 2 LF 3 LF 4 LF 5 LF 6 LF 7

Temperature (“C) 605 655 690 700 700 750 805


Residence time reactor (s) 3.5 2.7 1.8 2.3 3.2 2.9 3.1

Products (wt.%)
Gases 30 68 75 12 72 69 58
Hydrogen 0.03 0.4 0.6 0.6 0.6 0.9 1.3
Carbon monoxide 1.2 1.3 1.3 1.0 1.4 0.8 1.4
Carbon dioxide 0.6 0.6 0.9 1.4 1.3 1.6 2.2
Methane 1.8 5.4 8.9 11 8.8 13 I3
Ethene 8.1 24 29 31 30 36 30
Ethyne 0.01 0.08 0.3 0.3 0.4 0.8 I.3
Ethane 1.5 2.9 3.3 3.4 2.8 0.8 I.7
Propene 7.7 18 18 14 15 8.4 3.2
Propane 0.6 0.6 0.7 0.5 0.5 0.2 0.07
niiso-Butene 4.4 6.7 4.7 2.2 3.0 0.6 0.2
cisltrans-2-Butene 0.6 0.7 0.9 0.6 0.7 0.3 0.08
Butadiene 3.1 7.1 6.9 5.4 6.4 4.0 2.2

Aliphatics 31 9.8 7.2 3.7 4.5 1.5 0.7


C,-Hydrocarbons 6.1 6.9 4.6 2.8 3.4 1.5 0.7
C,-Hydrocarbons 6.4 2.4 1.2 0.5 0.6 0.03 0.01

BTX-aromatics 1.6 5.9 9.8 13 9.9 14 I6


Benzene 0.8 3.9 6.8 9.4 7.4 11 13
Toluene 0.7 1.9 2.1 3.2 2.4 2.8 2.4

Other aromatics 1.7 2.6 4.6 6.0 4.6 7.7 7.7


Styrene B I.0 1.6 1.8 1.4 2.1 2.1
Indene B 0.2 0.5 0.7 0.6 1.0 1.1
Naphthalene d 0.2 0.7 I .4 0.8 2.7 2.9

Sum oil 51 18 22 23 19 24 24
Distillation residue 19 13 3.4 4.7 8.5 6.4 I8
soot b 0.6 0.4 0.9
0.6 0.8 0.7

“Not detected
‘Not found.
C.M. Simon rt 01. )))J. And. Appl. Pyrol~~.vi.s38 (1996) 75 87 Xl

40

35

30

E 25

%
F 20

5
F l5

10

0
655 705 755

Temperature [“Cl

Fig. 3. Influence of temperature on the product yield.

3.2. Oil ,fbxtion

The oils show a typical change from aliphatic to aromatic character with
increasing temperature. BTX-aromatics increase from a low of 1.6 wt.‘%, at 605°C
up to 16 wt.‘% at 805°C. The aliphatic oil/wax consists of alkenes, alkadienes and
alkanes. Lower hydrocarbons are more common and the yields decrease with a
rising carbon number of hydrocarbons. The main components of the aromatic oils
are benzene, toluene, styrene, naphthalene and indene. All compounds show a
positive temperature effect as expected for pyrolysis reactions.
The main results of the oil analysis are given in Table 4.
After distillation, the oils have only a low water content and show no signs of ash
after combustion. The chlorine content varies between 22 and 70 mg kg ’ and did
not seem to be influenced by the temperature or steam/plastics ratio. If the oil was
used for production purposes the chlorine content would have to be lowered to
values below 10 mg kg ~ ’

Table 4
Analysis of the distilled pyrolysis oils

Experiment LF I LF 2 LF 3 LF 4 LF 5 LF 6 L 1: 7

Temperature (“C) 605 655 690 700 700 750 X05


Residence time reactor (s) 3.5 2.7 1.8 7.3 3.2 2.9 i I
Water content (wt.%) 0.02 0.03 0.02 0.03 0.06 0.06 0.3.3
Chlorine content (ppm) 60 56 22 70 45 31 38

Elementary analysis (wt.‘%)


Carbon 85.2 88.9 90.9 91.7 91.6 92.0 92.6
Hydrogen 14.4 10.8 9.1 8.3 x.3 7.7 7.4

C’H ratio (atomic ratio) 0.53 0.69 0.84 0.93 0.92 I .oo I .05
82 CM. Simon et al. / J. And. Appl. Pyrolysis 38 (1996) 75-87

Table 5
Analysis of the distillation residues

Experiment LF 1 LF 2 LF 3 LF 4 LF 5 LF 6 LF 7

Temperature (“C) 605 655 690 700 700 750 805


Residence time reactor (s) 3.5 2.7 1.8 2.3 3.2 2.9 3.1

Ash (wt.‘%) 0.25 0.13 2.45 0.64 0.06 r’ ”


Chlorine (mg Cl g-‘) 2.4 0.87 0.54 0.39 0.58 0.54 0.52

Elementary analysis (wt.‘%)


Carbon 85.0 88.4 88.2 90.2 92.5 93.1 93.7
Hydrogen 14.0 11.1 7.5 6.7 7.0 6.2 5.9

C/H ratio (atomic ratio) 0.54 0.67 0.98 1.13 1.11 I .26 1.33

“Below 0.1 wt.‘%

3.3. Distillation residue/tar

The distillation residue contains high boiling substances with boiling points
above 295°C (fluorene and higher boiling substances) and tar.
As mentioned earlier, residues are formed in higher proportions at low and high
reaction temperatures. This can be explained by the fact that at lower temperatures
(LF 1 + 2) more heavy hydrocarbons ( > C,,,) are produced, whereas higher temper-
atures (LF 7) lead to the formation of more condensed aromatic, high boiling
compounds. These two products are left in the distillation residue.
Analysis results of the residues are given in Table 5.
As shown in Table 5, reaction temperatures above 690°C give a higher C/H ratio
( > 1 .O) indicative of a higher aromatic fraction. Chlorine is enriched in this fraction
and reaches values that range from 390 to 2400 mg Cl kg- ‘, which increases
further at low reaction temperatures. At higher gas velocities and shorter residence
times, a greater part of fine solids is transported along into the liquid fraction and,
after phase separation and distillation, is found as ash in the distillation residue (LF
3-5).

3.4. soot

This fraction contains the solid organic parts, which were found in the cyclone
and the reactor residue. It consists of fine soot and soot deposited on the reactor
sand.

3.5. Water fraction

The water fractions were obtained by phase separation of the liquid products in
a separation funnel. Typical analysis results are shown in Table 6.
The organic material in the waterphase is due to the same type of hydrocarbon
compounds found in the oil phase. The major components detected are benzene
and toluene, in addition to some phenol detected by GC/MS.
C.M. Simon et al. )/IJ. Ancll. A{)pl. P?~rol~~.vi.~
3X (1996) 75 87 83

Table 6
Analysis of the water fractions

Experiment LF I LF 2 LF 3 LF 4 LF 5 LF 6 LF 7

Temperature (“C) 605 655 690 700 700 750 805


Residence time reactor (s) 3.5 2.7 1.8 2.3 3.2 2.9 3. I

pH value 3.1 2.8 3.3 2.‘) -.-


7, 2.7 ) Y
_.
Chlorine content (mg I-‘) 130 245 I63 I75 705 225 ‘00

TOC (mg I ~‘) 40 60 ~’ I’ 39 25 I7


COD (mg I -‘) IO00 1470 ‘I .a 950 512 300

.‘Not measured

The total organic carbon (TOC) analyses of the water samples were determined
on filtered samples. The TOC values are typical of those found for refinery waste
waters [17]. The high values of the chemical oxygen demand (COD) can be
explained by the presence of hydrocarbons in the water fraction and particulate
carbon associated with the very fine soot particles not separated by the cyclone and
filtration processes.
The pH values are influenced by the steam/plastics ratio and are therefore a little
higher in the experiments LF 3 and LF4. The pH values correlate well with the
chloride content.
The water fractions may be treated in the existing waste water facilities found in
a refinery.

The series of experiments shows the expected influence of temperature on the


distribution of hydrogen amongst the product fractions. Table 7 contains the CH
ratio of the gases, oils and distillation residues.
At lower temperatures, carbon and hydrogen ratios are similar in all product
fractions, which is also similar to the C/H ratio of the feed polyolefin waste. Higher
temperatures lead to changes due to different cracking conditions. The gas fraction

Table 7
C’H ratios (atomic ratios) of the product fractions

Experiment LF I LF 2 LF 3 LF 4 LF 5 LF 6 LF 7

Temperature (“C) 605 655 690 700 700 750 805


Residence time reactor (s) 3.5 2.7 1.8 2.3 3.2 2.9 3.1

Product fraction
Gas 0.53 0.47 0.43 0.42 0.46 0.43 0.40
Oil 0.53 0.69 0.84 0.93 0.92 I .oo I .05
Residue 0.54 0.67 0.98 I I .3 I.11 I.76 I.33
84 C.M. Simon et al. /J. Anal. Appi. Pyrolysis 38 (1996) 75-87

Table 8
Distribution of chlorine over the product fractions

Experiment LF 2 LF 7

Chlorine input (g)” 3.17 3.43

Chlorine in products (g)


Gas 0.015 0.014
Oil 0.03 0.03
Residue 0.08 0.07
Deposit (tar) 0.83 1.24
soot 0.25 0.23
Water 2.22 1.71

Sum products (g) 3.43 3.29


Recovery rate (“/;t) 91 96

‘Considering an average chlorine content of 0.075 wt.%.

becomes richer in hydrogen, while the other products lose hydrogen and become
more and more aromatic as the reaction temperature is increased.

3.7. Chlorine distribution

The distribution of chlorine between the various pyrolysis fractions is shown in


Table 8 for the experiments LF 2 and LF 7.
Most of the chlorine is found in the solids or in the water. The gas and oil
fractions are less contaminated with chlorine.

Table 9
Heavy metal concentrations of selected product fractions (mg kgg’)

Product fraction Input material Experiment LF 2 Experiment LF 7

Element LF Oil Residue Tar Oil Residue Tar soot

Na 42 0.5 5 170 1 32 18 1700


K 35 to.09 2.8 355 <0.08 4 33 3400
Cr 40 1.7 6.3 180 1.4 5 38 2300
Mn 6 0.025 0.53 28 0.04 0.33 4.1 840
Zn 225 20 18.1 165 16 10 120 350000
Cd 40 to.03 4.7 100 <0.03 8 60 4000
As 0.12 0.075 2.1 2 0.028 1.7 0.71 13

Sb 8 0.002 3 50 0.065 3.7 10 490


Se 5 2 4.6 15 4 2.9 3.1 350
C. M. Simon et al. ! J. Anal. Appi. Pyroljxis 38 (1996) 75 87 85

Table IO
Product distribution of the pyrolysis of polyoletins under different parameters (refering to total organic
input)

Experiment LF 4 LF 5 PE I MPW PE 2 PPI BI B7

Temperature (“c) 700 700 700 700 740 740 680 740
Reactor LWS 3 LWS I FB
Input material LF LF PE MPW PE PP PE PE
Fluidizing gas Steam Nitrogen Steam

Products (wt.‘%,)
Methane II 8.8 10 6.4 6.7 7.8 2.6 6.9
Ethene 31 30 36 I9 34 I3 I4 33
Ethane 3.4 2.8 2.7 2.0 3.5 3.9 1.7 3.2
Propene I4 I5 I5 II 22 36 7.8 19
2: Butenes 2.8 3.1 2.4 3.1 I2 I6 8.9 I8
Butadiene 5.4 6.4 7.0 4.7 IO 5.3 ,I (8
Styrene 1.8 I.4 I.1 I9 0.05 0. I ,’ ’
E Ethenejpropene 45 45 51 30 56 49 21 52

FB, Fluidized bed; LF, light fraction (PE/PP) postconsumer waste: PE. polyethylene; PP. polypropylene;
and MPW. mixed plastics waste (PEjPP/PS).
“Value not reported.

3.8. Hracy metal distribution

The heavy metal contents of some of the product fractions compared to those
found in the input feeds are presented in Table 9. The majority of these metals
accumulated in the solids and the residue although some are found in the oils, for
example the elements chromium (l-2 mg kg -I), selenium (2-4 mg kg ‘) and zinc
(16630 mg kg- ‘) which reached concentrations above 1 mg kg- ‘.

4. Comparison

For discussion of the results, the experiments are compared to other data
produced earlier in our research group and to two experiments conducted by
Buekens et al. [15].
The experiments PE 1 and MPW were carried out on the same pyrolysis system
only with different input materials [lo]. The experiments PE 2 and PP 1 with brand
new PE and PP, respectively, as input materials both used a smaller, lab scale
apparatus (20-40 g hh ‘, diameter 55 mm) and employed a shorter residence time
with nitrogen as the fluidizing gas [6]. Experiments B 1 and B 2 reported in [ 151 also
represent experiments with steam as the fluidizing gas and new PE as the input
material.
Table 10 shows that the highest values of ethene and propene were produced on
the lab scale pyrolysis systems with virgin plastic as feed and with very low
residence times and a temperature above 700°C (experiments PE 2 and PP 1). In
86 C.M. Simon et al. /J. Anal. Appl. Pyrolysis 38 (1996) 75-87

larger pyrolysis systems, the yields of ethene and propene drop a little (experiment
PE 1) due to longer residence times, resulting in increased secondary reactions of
the primary formed olefins. For the light fraction (LF 4 and LF 5) a sum of 45
wt.% of ethene and propene was reached. This lower value could be expected from
the results of the other experiments, but is actually higher than the yield from
MPW. On the other hand, MWP gives a higher yield of styrene due to the higher
PS content. All results are in accordance with the values reported from the
experiments B 1 and B 2. For further reduction of secondary reactions and
optimization of the olefin yield, shorter residence times should be used. This
observation is in keeping with the results documented in a recent US patent [18],
where the conversion of mixed plastics in a circulating fluidized bed is reported to
yield 42 wt.% ethene and 7 wt.% propene.
The results of experiments LF 4 and LF 5 (Table 11) show that a fluidized bed
process using a light fraction as the input material can yield a similar product
distribution as naphtha cracked in a steam cracker [19].

5. Conclusions

The results presented in this study suggest that pyrolysis may be a promising way
for feedstock recycling of polyolefinic plastic wastes. It is possible to obtain yields
of the main products, which are comparable to those of common technical
processes producing the same components.
By further variation of the parameters, it should be possible to increase the yields
of the desired olefins, ethene and propene, to around 50 wt.%. Higher yields of
specific olefins will be difficult to obtain using a thermal process with random bond
breaking resulting in a variety of hydrocarbons [20].

Table 11
Comparison of product yields from steamcrackers for different feed materials [18]

Experiment LF 4 LF 5
Temperature (“C) 700 700
Reactor FB FB Steam cracker
Input material LF LF LPG Naphtha AGO
Fluidizing gas Steam Steam
Products (wt.%)
Ethene 31 30 31 29 25
Propene 14 15 22 16 14
Butene, butadiene 8 10 13 10 9
Pyrolysis benzine 23 19 8 24 19
Gas, others 24 26 26 21 33
Z Ethene/propene 45 45 53 45 39

FB, Fluidized bed; LF, light fraction (PE/PP) post consumer waste; LPG, liquid petrol gasoline; and
AGO, atmospheric gasoline.
C.M. Simon et ul. lIJ. Anal. Appl. Pyro!,k 38 (1996) 75-87 87

Once the optimum conditions have been established, more work will be required
to ensure a satisfactory chlorine balance and a heavy metal distribution. The
addition of ammonia to the steam could be one way of reducing the chlorine
concentration of the products (gas and oil) and represents new experiments planned
for the future. Further studies concerning the energy balance are also important
prior to a detailed economics evaluation of the process.

Acknowledgements

The author wishes to thank Volkswagen-Stiftung, Germany for financial support.

References

[I] H. Sutter. Erfassung und Verwertung von Kunststoff. EF-Verlag, Berlin, 1993.
[2] J. Brandrup, Die Wiederverwertung von Kunststoffen. Hanser, Munich, 1995.
[3] R.D. Leaversuch, Mod. Plast. Int., July (1991) 26.
[4] M. Gebauer, Kunststoffe, 85(2) (1995) 214.
[5] D. Kunii and 0. Levenspiel, Fluidization Engineering. Butterworth-Heinemann. Boston. 1991. p.
IO.
[6] H. Sinn, W. Kaminsky and J. Janning, Angewandte Chemie, 88 (1976) 737.
[7] W. Kaminsky, Makromol. Chem. Makromol. Symp., 4849 (1991) 381.
[8] H. Kastner and W. Kaminsky, Hydrocarbon Process., May (1995) 109.
[9] J.H. Brophy and S. Hardman, in H. Sutter (Ed.), Erfassung und Verwertung von Kunststoff.
EF-Verlag, Berlin, 1993, p. 426.
[IO] W. Kaminsky. B. Schlesselmann and C. Simon, J. Anal. Appl. Pyrolysis, 32 (1995) 19.
[I I] W. Kaminsky and J. Franck, J. Anal. Appl. Pyrolysis. 19 (1991) 311.
[12] R. Rahnenfiihrer, Ph.D. Thesis, University of Hamburg, 1993.
[I 31 APME, Plastics Recovery in Perspective, Brussels. 1994.
[I41 B. Schlesselmann, W. Kaminsky and C.M. Simon, Polymer Degradation, in press.
[15] J.G. Schoeters, A.G. Buekens, Recycling Berlin ‘79. Verlag fiir Umwelttechnik, Berlin, 1979.
[16] J. Werther, Int. Chem. Eng., 20 (1980) 529.
[17] ATV. Grundsatze fur Bemessung und Betrieb von Abwasserreinigungsanlagen in Erdiilrafhnerien.
Regelwerk Abwasser-Abfall, St. Augustin, 1986, p, 201.
[18] US Patent. US 005326919, 1994.
[19] B.P. Deutsche, Das Buch vom Erdiil, Reuter und Klockner Verlag, Hamburg, 1986.
[20] G. Menges, in J. Brandrup (Ed.), Die Wiederverwertung von Kunststoffen. Hanser, Munich. 1995.
D. 3.

Anda mungkin juga menyukai