Você está na página 1de 19

Resources, Conservation and Recycling 23 (1998) 163–181

Catalytic plastics cracking for recovery of


gasoline-range hydrocarbons from municipal
plastic wastes

A.G. Buekens *, H. Huang


Department of Chemical Engineering and Industrial Chemistry - CHIS 2, Free Uni6ersity of Brussels,
Pleinlaan 2, Brussels, 1050, Belgium

Accepted 5 May 1998

Abstract

This paper reviews recent developments in plastics cracking, a process developed to recycle
plastic wastes into useful petrochemical materials. Under thermal cracking conditions, plastic
wastes can be decomposed into three fractions: gas, liquid and solid residue. The liquid
products are usually composed of higher boiling point hydrocarbons. By adopting customary
fluid cracking catalysts and reforming catalysts, more aromatics and naphthenes in the
C6 –C8 range can be produced, which are valuable gasoline-range hydrocarbons. More tests
are, however, needed to verify the pyrolysis process in a pilot scale particularly for treatment
of mixtures of bulk plastics. Plastics cracking is only an elementary conversion technology;
its application has to be combined with other technologies such as municipal solid waste
collection, classification and pretreatment at the front end, as well as hydrocarbon distillation
and purification at the back end. Social, environmental and economic factors are also
important in industrial implementation of the technology. © 1998 Elsevier Science B.V. All
rights reserved.

Keywords: Plastic waste; Plastics cracking; Hydrocarbons; Municipal

* Corresponding author. Tel.: +32 2 6293244; fax: + 32 2 6293333; e-mail: vnet3.vub.ac.be

0921-3449/98/$19.00 © 1998 Elsevier Science B.V. All rights reserved.


PII S0921-3449(98)00025-1
164 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

1. Introduction

In recent years the production and consumption of plastics have increased


drastically; as a consequence the responsible disposal of plastic wastes has created
serious social and environmental arguments. At present both landfilling and
incineration of plastic wastes are widely practised. In Japan, the percentage of
municipal plastic wastes, as a fraction of municipal solid waste (MSW), that was
landfilled in the early 1980s was estimated to be 45%, incineration was 50%, and
the other 5% was subjected to separation and recycling [1]. In the USA, more
than 15% of the total MSW was incinerated in 1990; only about 1% of post-con-
sumer plastics were recycled [2 – 4].
Landfilling of plastic wastes is expected to decrease in the future as landfill
space is depleted, and plastic wastes are resistant to environmental degradation.
Co-incineration of plastic wastes with other municipal solid wastes may be in-
creasingly practised, because the high caloric value of plastics can enhance the
heating value of MSW and facilitate an efficient incineration, while their energy
content can also be recovered. But the potential relationship between plastics fed
into an incinerator and the formation of some highly toxic pollutants such as
dioxins and furans is still unclear. It has been suggested that the chlorine content
in PVC and other plastics is related to the formation of dioxins and furans, which
are chlorinated polynuclear aromatic compounds. And although there is consider-
able evidence that these pollutants would still be generated in the absence of
plastics, environmental pressures against incineration have never completely disap-
peared.
Plastics pyrolysis, on the other hand, may provide an alternative means for
disposal of plastic wastes with recovery of valuable gasoline-range hydrocarbons.
In pyrolysis or thermal cracking, the polymeric materials are heated to high
temperatures, so their macromolecular structures are broken down into smaller
molecules and a wide spectrum of hydrocarbons are formed. These pyrolytic
products can be divided into a gas fraction, a liquid fraction consisting of
paraffins, olefins, naphthenes and aromatics, and solid residues. In catalytic crack-
ing, more aromatics and naphthenes are selectively formed in the presence of
commercial fluid cracking catalysts (FCC) or reforming catalysts, so that the
productivity and economics of pyrolysis processes are improved. This paper re-
views recent laboratory experiments of catalytic plastics pyrolysis for converting
plastic wastes into gasoline-range hydrocarbons and the industrial aspects of
implementing this technology.

2. Sources and properties of plastic wastes

Plastic wastes can be classified as industrial and municipal plastic wastes ac-
cording to their origins; these groups have different qualities and properties and
are subjected to different management strategies.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 165

2.1. Industrial plastic wastes

Industrial plastic wastes are those arising from the plastics manufacturing and
processing industry. Usually they are homogeneous or heterogeneous plastic resins,
relatively free of contamination and available in fairly large quantities. Recycling
technologies for industrial plastic wastes are currently based on pelletization and
molding into low grade plastic products; the recycled products have poor mechan-
ical and color qualities and a lower market value [5]. The reclaimed product outputs
of Japan in the early 1980s already amounted to some 15% of total industrial
plastic wastes [1]. Thus for industrial plastic wastes, repelletization and remolding
seem to be a simple and effective means of recycling. But when plastic wastes are
heterogeneous or consist of mixed resins, they are unsuitable for reclamation. In
this case thermal cracking into hydrocarbons may provide a suitable means of
recycling, which is termed chemical recycling.

2.2. Municipal plastic wastes

Municipal plastic wastes normally remain a part of municipal solid wastes as they
are discarded and collected as household wastes. Plastics usually account for about
7% of the total MSW by weight and much more by volume. In order to recycle
municipal plastic wastes, separation of plastics from other household wastes is
required. Although MSW separation technologies have been studied extensively, it
is still not possible to classify MSW mechanically and obtain marketable fractions.
So waste separation at the household is required with regard to recycling of
municipal plastic wastes.
If household wastes are separately disposed into three parts: (1) combustibles
such as paper, kitchen waste, textiles, and wood, (2) incombustibles such as metals,
glass, ceramics, and (3) plastics, then the collected plastics will be mixed plastic
wastes with major components of PE, PP, PS, PVC, etc. For mixed plastics some
mechanical separation equipment is currently available [1,6]. For example, using a
wet separation process mixed plastics can be separated into two groups: those with
a density greater than water such as PS and PVC, and those with a density lower
than that of water such as PE, PP and expanded PS. The latter group is much
larger than the first group. Consequently, recycling of municipal plastic wastes
should deal with plastic mixtures of PE, PP and PS, provided that the above
separation procedures are practised. Typical composition of such plastic mixture
will be three parts PE, one part PP and one part PS. More investigations are needed
to identify the sources and properties of plastic wastes, and their suitability for
various recycling methods such as repelletization, remolding and pyrolysis.

3. Gasoline-range hydrocarbons

The pyrolysis of plastic wastes produces a whole spectrum of hydrocarbons


including paraffins, olefins and aromatics. But not all these hydrocarbons are
166 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

suitable for gasoline usage. Gasoline qualities are usually measured in terms of
volatility and octane number of the hydrocarbons.
(a) Volatility. Adequate volatility is required for smooth operation of petrol
engines. Light hydrocarbons have higher volatilities; they may cause vapor lock
when the engine is hot. Heavy hydrocarbons have lower volatilities, but they may
not be volatile enough to start the engine when the engine is cold. Suitable
hydrocarbons are in the C5 – C8 range; some C9 and above may be added accord-
ing to climate and season.
(b) Octane number. This is a measure of the gasoline quality for prevention of
early ignition which leads to cylinder knock. Higher octane numbers are pre-
ferred. The octane numbers of selected hydrocarbons are listed in Table 1.
Generally speaking, for gasoline production, aromatics, naphthenes and iso-
alkanes are highly desirable, whereas olefins and n-paraffins are less desired.
Accordingly, it is desirable that the pyrolytic conversion of plastic wastes pro-
duces more aromatics, naphthenes and iso-alkanes with carbon number in the
range C5 – C8.

Table 1
Octane numbers and boiling points of some hydrocarbons [7,8]

Compounds Octane number (blending, research method) Boiling point (°C)

Paraffins
n-Butane 113 0
n-Pentane 62 36
n-Hexane 19 69
n-Heptane 0 98
Isooctane 100
2-Methylhexane 41 90
2,2-Dimethylpentane 89
2,2,3-Trimethylbutane 113
Naphthenes
Cyclopentane 141 49
Methylcyclopentane 107 72
Cyclohexane 110 81
Olefins
2-Methyl-1-hexene 91a 91
3-Methyl-2-pentene 97a 68
Aromatics
Benzene 99 80
Toluene 124 111
m-Xylene 145 139
1,3-Dimethylbenzene 145
1,3,5-Trimethylbenzene 171
Isopropylbenzene 132

a
Actual octane number.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 167

Table 2
Decomposition mechanisms and monomeric yields of some polymers [14,15]

Polymers Decomposition mechanisms Monomeric yield (weight %)

Polymethylmetacrylate End-chain scission 91 – 98


Polytetrafluorethylene End-chain scission 95
Polymethacrylonitrile End-chain scission 90
Polyethylstyrene End-chain scission 82 – 94
Polystyrene End-chain and random-chain scission 42 – 45
Polyisobutene End-chain and random-chain scission 18 – 25
Polyethylene Random-chain scission 0.03
Polypropylene Random-chain scission 0 – 17
Polybutadiene Random-chain scission 1
Polyvinylchloride Chain-stripping 0 – 0.07
Polyvinyl fluoride Chain-stripping 0
Polyacrylonitrile Chain-stripping 5

4. Thermal cracking

Pyrolysis and gasification of plastics and other carbonaceous fuels have been
studied extensively in the past [9–11]. Recent progress in converting plastic wastes
into petrochemicals by means of pyrolysis in the absence of a catalyst has been
reviewed by Kaminsky [12,13]. A comprehensive treatment of the mechanism of
plastics pyrolysis has been presented by Cullis and Hirschler [14].
Four types of mechanisms of plastics pyrolysis have been proposed:
(a) End-chain scission or depolymerization. The polymer is broken up from the
end groups successively yielding the corresponding monomers.
(b) Random-chain scission. The polymer chain is broken up randomly into
fragments of uneven length.
(c) Chain-stripping. Elimination of reactive substitutes or side groups on the
polymer chain, leading to the evolution of a cracking product on one hand, and
a charring polymer chain on the other.
(d) Cross-linking. Formation of a chain network, which often occurs for ther-
mosetting polymers when heated.
These different mechanisms and product distributions are to some extent related
to bond dissociation energies, the chain defects of the polymers, and the aromaticity
degrees, as well as the presence of halogen and other hetero-atoms in the polymer
chains. For common plastics the decomposition mechanisms and associated
monomer yield are listed in Table 2. The pyrolysis of PMMA occurs following the
end-chain scission mechanism and a monomer recovery of more than 90% can
easily be achieved [15]; the pyrolysis of PS occurs by both end-chain and random-
chain scission and the monomer recovery is only some 45% [16]. For PE and PP,
the main components of municipal plastic wastes, the pyrolysis occurs through the
random-chain scission mechanism and a whole spectrum of hydrocarbon products
is obtained [10]. The gas and oil yields from polyolefin pyrolysis are about 50 and
40% wt. of the feed at 750°C, respectively, as shown in Table 3; and the oil fraction
168 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

consists mainly of higher boiling point hydrocarbons (tar). Usually the gas can only
be burned to provide the process heat, and the overall yield of valuable gasoline-
range hydrocarbons is poor, so that pyrolysis process as a means for chemical
recycling of municipal plastic wastes is rarely practised on an industrial scale at
present.

5. Catalytic cracking

A number of experimental studies have been carried out by various researchers


with the objective of improving gasoline yield from plastics pyrolysis by introducing
suitable catalysts, as listed in Table 4. Common plastics such as PE and PP have
already been tested extensively; the catalysts tested are mainly those used in the
petrochemical refinery industry. The laboratory experimental set-up in these studies
is mostly flow reactors; it may be useful to distinguish between two modes of
catalyst usage: ‘liquid phase contact’ and ‘vapor phase contact’ [35]. In ‘liquid phase
contact’, the catalyst is contacted with melted plastics and acts mainly on the
partially degraded oligomers from the polymer chains; in ‘vapor phase contact’, the
polymer is thermally degraded into hydrocarbon vapors which are then contacted
with the catalyst. In this section these laboratory studies are discussed following the
type of catalyst adopted, i.e. FCC catalyst, reforming catalyst, active carbon, etc.

5.1. FCC catalyst

Fluid cracking catalysts (FCC) have been developed for the cracking of large
hydrocarbon molecules from crude petroleum or its denser fractions into gasoline-
range hydrocarbons with suitable volatilities. Major commercial FCC catalysts are
listed in Table 5. Zeolite-based catalysts have much higher activities and selectivities
than the former alumino-silicate catalysts.
FCC catalysts have been found to have significant effects in plastics pyrolysis.
Some representative data are given in Tables 6–10. It is of interest to note that in
these data the determination of liquid, tar and residue fractions is dependent on the
experimental set-up and analytical methodology. Generally speaking, liquid frac-
tion has average molecular weight 100–250, tar has 300–800, and the residue may
include coke and undegraded polymer. However, in some experiments tar has been

Table 3
Product distribution in plastics pyrolysis (% wt. of feed) [13]

Feed Pyrolysis temperature (°C) Gas Oil Residue Other

PE 760 55.8 42.4 1.8 —


PP 740 49.6 48.8 1.6 —
PS 581 9.9 24.6 0.6 64.9 Styrene
Mixture PE/PP/PS 750 52.0 46.6 1.4 —
Table 4
Recent laboratory experiments on catalytic plastics cracking

Investigator Experimental setup Plastics tested Catalysts tested

Uemichi et al. Flow reactor (plastics are melted in a PE, PP Silica alumina, CaX zeolite, activated
[17–20] melter and fed to a reactor packed with carbon, metal supported on
catalyst in N2 flow) silica-alumina or activated carbon
Ishihara et al. Flow reactor (plastics are melted in a PE, PP, PS Active charcoal, silica-alumina, NaY
[21–25] melter and fed to a reactor packed with zeolite
catalyst in N2 flow); batch reactor
(plastics and catalyst are mixed and
heated in N2 flow)
Mordi et al. [26,27] Sealed reactor (plastics and catalyst are LDPE, PP H-ZSM-5, H-mordenite, H-theta-1
mixed end heated) zeolites
Beltrame et al. Flow reactor (a mixture of plastics and PE, PS Silica gel, alumina, silica-alumina, rare
[28,29] catalyst is heated); TGA (a mixture of earths-Y and H-Y zeolites
plastics and catalyst in a N2 stream)
Vasile et al. [30–32] Flow reactor (plastics are melted in an LDPE, HDPE, PP Silica-alumina, ZSM-5 zeolite,
extruder and fed to a reactor packed with dealkylation catalyst
catalyst)
Ueno et al. [33,34] Flow reactor (thermally degraded PE, PS Silica-alumina, HZSM-5 zeolite, active
products of plastics are passed through a carbon, metal oxides
fixed bed of catalyst in N2 flow)
Sakata et al. [35–37] Flow reactor (a mixture of plastics and LDPE, HDPE, linear LDPE, cross-linked Silica-alumina
catalyst is heated); flow reactor (thermally LDPE, PP, PET, PVC
degraded products of plastics are passed
through a fixed bed of catalyst)
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181
169
170 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

Table 5
FCC catalysts

Catalyst Major cations exchanged Composition (wt. %)

SiO2 Al2O3 Na Ca Re

Silica-alumina — 88 – 74 12 – 26 — — —
Zeolite A Ca2+ 42.5 37.4 7.85 13.0 —
Zeolite X and Y Ca2+ 47.8 31.5 7.7 12.3 —
RE, NH+
4 40.0 33.0 0.22 — 26.5
Zeolite ZK5 H+ 26.8 23.1 0.47 — —
+
Mordenite H 80.1 13.4 0.3 1.54 —

RE, rare earth metals.

combined with liquid and grossly reported as oil fraction, or combined with
undegraded polymer and taken as residue fraction.
The main effects of catalyst addition in plastics pyrolysis are as follows. (1) The
pyrolysis temperature for achieving a certain conversion is reduced drastically.
Comparing the data in Tables 3, 6 and 7 it can be seen that under non-catalytic
conditions a temperature above 700°C is usually required to reach a reasonable
conversion; on the other hand under catalytic conditions, the pyrolysis temperature
can be lowered to 300 – 400°C; and as the catalyst/plastics ratio is increased, the
pyrolysis temperature can be further lowered. For example, in polypropylene
pyrolysis when the catalyst/plastics ratio was 6, a liquid yield of 46% was obtained
within 1 h of reaction time even at temperature as low as 220°C [25]. (2) More
iso-alkanes and aromatics in the C5 –C10 range can be produced which are highly
desirable gasoline-range hydrocarbons (see Tables 8–10). (3) The reaction rate is
increased significantly; e.g. the initial rate of degradation of polypropylene was
reported to be approximately four times faster than that of non-catalytic thermal
degradation [35].
The catalytic effects of FCC catalysts in oil cracking are due to their acidic
properties and have been explained by a carbonium ion theory [7,8]. Similar
mechanisms have been postulated for plastics pyrolysis in the presence of FCC or
other acidic catalysts [18,19,23 – 25,30,38]. The catalytic mechanisms may be briefly
described using polyethylene as an example.
(1) Initiation. Initiation may occur on some defect sites of the polymer chains.
For instance, an olefinic linkage could be converted into an on-chain carbonium ion
by proton addition:
– CH2CH2CHCHCH2CH2 – + HX“ –CH2CH2+ CHCHCH2CH2 + X −
(1)
Then the polymer chain may be broken up through b-scission:
+ +
– CH2CH2 C HCH2 – CH2CH2 – “ –CH2CH2CHCH2 + C H2CH2 – (2)
Table 6
Product distribution of polyethylene pyrolysis in batch reactors

Temperature (°) Catalyst Catalyst/plastics ratio (w/w) Reaction time (h) Product distribution (wt. %) Data source

Gas Liquid Tar Residue

260 None 0 1 0 3 15 82 [23]


260 NaY zeolite 1 1 10 38 36 16 [23]
350 None 0 2 34 15 46 [27]
350 H-mordenite 0.2 2 43 32 25 [27]
400 None 0 1 10 0 90 [28]
400 Silica-alumina 0.1 1 7 93 0 [28]
400 H-Y zeolite 0.1 1 9 91 0 [28]
400 None 0 \7 20 44 13 17 [33]
400 Silica-aluminaa 0.1 4 37 52 4 Trace [33]
400 HZSM-5 zeolitea 0.1 B3 50 45 1 Trace [33]
600 None 0 1 6 29 65 [28]

a
Vapor phase contact.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181
171
172

Table 7
Product distribution of polypropylene pyrolysis in batch reactors

Temperature (°) Catalyst Catalyst/plastics ratio (w/w) Reaction time (h) Product distribution (wt. %) Data source

Gas Liquid Tar Residue

220 Silica-alumina 1 1 4 18 32 47 [25]


220 Silica-alumina 6 1 12 46 36 5 [25]
260 Silica-alumina 1 1 6 32 29 33 [25]
350 H-mordenite 0.2 2 43 24 15 [26]
380 None 0 13.3 24.7 64.9 10.4 [35]
380 Silica-aluminaa 0.1 13.3 35 54.5 10.5 [35]
380 Silica-alumina 0.1 5.8 24.8 68.8 6.4 [35]

a
Vapor phase contact.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 173

Table 8
Product composition in thermal cracking of polypropylene at 526°C (wt. %) [18]

Carbon number Normal alkanes Iso-alkanes Alkenes Total

H2 Trace
1 1.0 1.0
2 2.1 1.0 3.1
3 0.4 8.3 8.7
4 3.5 3.5
5 4.0 1.5 5.5
6 0.3 4.8 5.1
7 2.5 2.5
8 0.8 1.3 2.1
9 0.5 11.4 11.9
10 1.0 1.0
11 0.4 0.5 0.9
12 1.7 1.7
13 0.8 0.8
14 0.2 0.2 0.4
15 3.0 3.0
16 1.1 1.1
17–24 5.1 5.1
HBP 42.9
Total 7.5 2.2 47.7 100.3

HBP, higher boiling point hydrocarbons.

Table 9
Product composition of polypropylene cracking over silica-alimina at 477 °C (wt. %) [18]

Carbon number Normal alkanes Iso-alkanes Alkenes Aromatics Total

H2 0.1
1 0.8 0.8
2 0.4 0.5 0.9
3 1.6 9.6 11.2
4 1.8 13.6 13.1 28.5
5 1.0 10.0 10.7 21.7
6 0.4 6.7 6.5 0.3 13.9
7 0.1 2.3 2.7 2.9 8.0
8 Trace 1.3 0.7 3.5 5.5
9 0.6 0.2 2.0 2.8
10 0.2 1.4 1.6
11 0.1 0.2 0.3
12 Trace 0.3 0.3
13 –16 0.5
HBP 3.8
Total 6.1 34.8 44.0 10.6 99.9
174 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

Table 10
Product composition of polypropylene cracking over CaX type zeolite at 477 °C (wt. %) [18]

Carbon number Normal alkanes Iso-alkanes Alkenes Aromatics Total

H2 Trace
1 0.9 0.9
2 0.9 0.5 1.4
3 2.6 8.3 10.9
4 3.6 19.1 8.0 30.7
5 2.1 17.1 3.9 23.1
6 1.1 11.4 1.3 0.2 14.0
7 0.4 6.0 0.2 1.1 7.7
8 0.2 2.5 2.3 5.0
9 0.1 0.9 1.8 2.8
10 Trace 0.2 0.6 0.8
11 0.1 0.1 0.2
12 Trace Trace Trace
13 –16
HBP 2.6
Total 11.9 57.3 22.2 6.1 100.1

Initiation may also take place through random hydride-ion abstraction by


low-molecular-weight carbonium ions (R + ):
– CH2CH2CH2CH2CH2CH2 – + R + “ –CH2CH2+ CHCH2CH2CH2 – + RH
(3)
The newly formed on-chain carbonium ion then undergoes b-scission as in Eq.
(2).
(2) Depropagation. The molecular weight of the main polymer chains may be
reduced through successive attacks by acidic sites or other carbonium ions and
chain cleavage, yielding an oligomer fraction (approximately C30 –C80). Further
cleavage of the oligomer fraction probably by direct b-scission of chain-end
carbonium ions leads to gas formation on the one hand, and a liquid fraction
(approximately C10 – C25) on the other.
(3) Isomerization. The carbonium ion intermediates can undergo rearrangement
by hydrogen- or carbon-atom shifts leading to, e.g. a double-bond isomerization of
an olefin:
H+ + −H+
CH2CH – CH2 – CH2 – CH3 “ CH3 – C H–CH2 –CH2 –CH3 “ CH3 –CHCH
– CH2 – CH3
Other important isomerization reactions are methyl-group shift and isomeriza-
tion of saturated hydrocarbons.
(4) Aromatization. Some carbonium ion intermediates can undergo cyclization
reactions. An example is when hydride-ion abstraction first takes place on an olefin
at a position several carbons removed from the double bond, the result being the
formation of an olefinic carbonium ion:
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 175

R1+ +R2CH —-CH – CH2CH2CH2CH2CH3UR1H+ R2CH—-CH


+
– CH2CH2CH2 C HCH3
This carbonium ion could undergo intramolecular attack on the double bond:

Such reactions provide a route to cyclization and formation of aromatics.


The experimental findings that in plastics cracking the FCC catalysts lower the
pyrolysis temperature, enhance the production of iso-alkanes and aromatics, and
increase the reaction rate are well explained by the above carbonium ion mecha-
nism of acidic catalysts.

5.2. Reforming catalysts

Reforming catalysts are based on transition metals supported on silica-alumina.


The most commonly used reforming catalyst is Pt/SiO2-Al2O3 with about 0.5 wt. %
Pt only. These catalysts are said to be bifunctional, in view of the two kinds of
active sites playing different roles. The metallic sites catalyze hydrogenation/dehy-
drogenation reactions, while the acidic sites on the support catalyze isomerization
reactions. A proper combination of these two functions can promote a variety of
reforming reactions such as isomerization where straight-chain paraffins are isomer-
ized to branched-chain molecules, dehydrocyclization where straight-chain paraffins
are converted to cycloparaffins, and dehydrogenation in which naphthenes are
dehydrogenated to aromatics; the octane numbers of gasoline-range hydrocarbons
are improved without changing their carbon numbers. The following overall
reaction scheme for n-pentane isomerization illustrates the reforming reactions:
176 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

Table 11
Product distribution of polyethylene cracking over reforming and activated carbon catalysts at 526°C
[20]

Catalysts Product distribution (wt. %)

Hydrogen Normal alkanes Iso-alkanes Alkenes Aromatics HBP

A12O3 1 21 2 40 4 28
Pt/Al2O3 5 17 2 22 12 42
SiO2-Al2O3 0.5 18 26 42 13.5 0
Pt/SiO2-Al2O3 3 10 18 42 22 5
C 1 60 5 10 24 0
Pt/C 5 28 2 12 53 0

Similarly, normal heptane could be reformed to iso-heptane, and normal hex-


ane and iso-hexane could be reformed to benzene [8].
When reforming catalysts are applied to plastics cracking, the acidic sites in
the alumina support are expected to function as an FCC catalyst, and the
metal sites expected to catalyze the isomerization and dehydrocyclization reac-
tions, so that a higher selectivity to aromatic products can be achieved. This
has been demonstrated by Uemichi et al. [20]; some of their experimental data
are reproduced in Table 11. When only Al2O3 was used, the higher-boiling-
point hydrocarbon fraction was 28% and the aromatic fraction was only 4%
probably due to insufficient acidic activity; when Pt was impregnated on Al2O3,
the aromatic yield was increased to 12%. Similarly, when SiO2-Al2O3 was used,
the yields of iso-alkanes and alkenes were increased significantly; when Pt im-
pregnated on SiO2-Al2O3 was tested, the aromatic yield was increased at the
expenses of normal and iso-alkanes. This data set seems to support that re-
forming catalysts enhance aromatic formation in plastics cracking.

5.3. Acti6ated carbon

Activated carbon has also been found to catalyze plastics cracking. The
product distribution in the presence of activated carbon catalyst is, however,
very different from that of FCC or reforming catalysts. As shown in Table 11,
normal alkanes are produced in a large quantity, whereas the amount of iso-
alkanes is very small. The catalytic mechanism of activated carbon is therefore
unlikely to be the same as the carbonium ion mechanism of acidic catalysts
because of the absence of branched alkanes, as pointed out by Uemichi et al.
[19,20]. When Pt impregnated on activated carbon was tested, the aromatic
yield was reported to reach as high as 50% of the plastic sample; it was sug-
gested that this is due to a good combination of cracking and dehydrocycliza-
tion activities of the catalyst [20]. More experimental work on this type of
catalyst seems desirable.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 177

5.4. Other catalysts

Other types of catalysts such as metal oxides [33,34], metal complexes of the type
MCln -AlCl3 or M(AlCl4)n (M =Li, Na, K, Mg, Ca, Ba; n= 1–2) [38], and alkali
metal carbonates or alkaline metal carbonates [39] have also been tested for
polymer degradation. These studies appear to be directed mainly for enhancement
of monomer recovery.

6. Process design

All the experiments discussed above were carried out in laboratory-scale reactors
with fixed beds of catalysts. The reported experimental data varied from laboratory
to laboratory, owing to different reactor set-ups, experimental procedures and
sample analysis techniques. Few data are available concerning plastics cracking in
fluidized beds of catalysts, which could have a different cracking product distribu-
tion because of the different catalyst/plastics contact and temperature profiles.
Several problems remain in industrial process design based on laboratory studies as
follows.
(1) Reactor type. To achieve continuous operation on an industrial scale,
fluidized bed reactors seem to be suitable with their higher heating rate to the
feedstocks due to good heat and mass transfer. A higher heating rate is believed to
result in a lower tar and a higher liquid production. The nominal residence time in
the high temperature zone has to be controlled carefully. Sometimes the products
may be over-cracked into too small molecules, resulting in more gas and less liquid
production. Fast fluidized bed reactors such as that reported by Scott [40] could
achieve a high productivity and throughput.
(2) Catalyst regeneration. Similar to the FCC process, catalysts used in plastics
cracking are deactivated rapidly by coke deposition on the catalyst surfaces. For
example, Uemichi et al. [18] reported that for polypropylene cracking at 477°C in
the presence of CaX zeolite, the higher-boiling-point hydrocarbon fraction was
2.6% at the initial stage of the reaction, but was increased to 46.2% at 130 min after
starting reaction due to catalyst deactivation. Thus for continuous operation the
cokes need to be burned off regularly to regenerate the catalytic activity. The FCC
configuration of a fluidized bed cracker, a fluidized bed regenerator and two
circulating loops is required also for catalytic plastics cracking.
(3) Processing of pyrolytic oil. In order to obtain useful gasoline-range hydrocar-
bons from the pyrolytic oil, fractional distillation is preferred for product separa-
tion. More valuable chemical raw materials including benzene, toluene and other
condensed aromatic hydrocarbons may be obtained by refining the pyrolytic oil.
(4) Effects of plastic mixtures. Most laboratory experiments of catalytic plastics
cracking were performed on pure polymer components; only a few data have been
reported for mixtures of different types of plastics [41]. Because of the interactions
of their degradation mechanisms, the pyrolytic behavior and product distribution of
a plastic mixture may be different from that of the individual component [36].
178 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

Further testing using a mixture of PE, PP, PS and PVC should provide more useful
data.
(5) Effects of plastic additives. Commercial plastic products have many additives
serving as plastic stabilisers, flame retardants, pigments, fillers, etc. Sometimes the
chlorine content of commercial polyethylene products can reach as high as 10% wt.
The gasoline product from plastics cracking may have an unacceptable high
chlorine content, so dechlorination and neutralization processing, such as that
adopted in petrochemical refining, may be necessary on some occasions.

7. Industrial applications

Some attempts to commercialize the pyrolysis technology for recovery of gaso-


line-range hydrocarbons from plastic wastes have been reported [42–47]. So far
these attempts appear to be unsuccessful due to logistic and economic factors. For
instance, thermal plastics pyrolysis has been tested in a 10-kg/h laboratory unit [42]
consisting of a fluidized bed reactor (45 mm o.d.), operated at temperatures of
400 – 600°C, having a quartz sand bed without the presence of a catalyst; nitrogen
was used as fluidizing gas and heat was provided by a tubular furnace. After the
completion of laboratory testing, a pilot plant with a capacity of 100 kg/h was built
[43,44]. However, for an industrial scale plant having a capacity of 25000 t/year, the
investment is about $20 million if the plant processes delivered and cleaned plastic
wastes, so that the treatment costs of plastic wastes are about $250/t. The product
from the plant is similar to typical naphtha from crude oil, which is a feedstock to
crackers. But naphtha is worth only $180/t, leaving a ‘gate fee’ (charged to the
waste supplier) of about $100/t of waste for the plant to break even, so that a
subsidy of $100/t of waste is required for the pyrolysis process to be economically
viable. It is thus apparent that the industrial implementation of pyrolysis technol-
ogy strongly depends on economic factors, especially oil prices. At the current low
oil prices, pyrolysis technology is not competitive. In the future, as the pyrolysis
process is further improved by such means as the addition of catalysts, and as the
oil price rises, pyrolysis technology may have a wider industrial application in
converting plastic wastes into gasoline-range hydrocarbons.
On the other hand social and environmental issues are also important in
implementing plastics pyrolysis technology. For example, municipal solid wastes are
extremely heterogeneous and separate disposal is required in the household in order
to collect plastic wastes in large quantities. A social waste disposal separation
scheme has to be practised if plastic wastes are to be recycled into valuable
petrochemicals by pyrolysis processes. Environmental issues include for example
dioxin formation in municipal solid waste incineration. The role of plastics in
incineration and its relation to dioxin formation is still under investigation. If
plastics are found to be one of the main factors of high dioxin formation, then the
current practise of co-combustion of plastics with other wastes in municipal
incinerators should be avoided, and waste plastic treatment may be shifted to
pyrolysis or gasification processes.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 179

8. Conclusions

Plastic wastes may be converted into gasoline-range hydrocarbons by pyrolysis


processes. The products of thermal cracking are mainly paraffins, olefins as well as
higher-boiling-point hydrocarbon products (carbon number \ 24). The products of
catalytic cracking are composed of more iso-alkanes and aromatics, which are
highly desirable gasoline-range hydrocarbons. Reforming catalysts have the highest
selectivity to aromatics, probably due to their combination of cracking and
reforming activities. In the best laboratory testing conditions, aromatic production
has been shown to reach as high as 50% wt. of the plastic sample at the initial stage
of reaction using a platinum on activated carbon catalyst.
Most of the tests on catalytic cracking so far were performed in fixed-bed flow
reactors. Few catalytic cracking data in fluidized bed units are available. The effects
of different plastic additives presented in the commercial plastic products on the
gasoline quality from pyrolysis process is unclear. Further tests are necessary for
plastic mixtures (e.g. three parts of PE, one part of PP and one part of PS) in the
presence of reforming catalysts (platinum impregnated on a silica-alumina support
or activated carbon support) in fluidized bed reactors.
The industrial implementation of pyrolysis technology on a wider scale depends
strongly on economic, environmental and political issues. Currently, due to the
moderate oil prices, pyrolysis technology is not likely to be economically competi-
tive. However, environmental considerations may hasten the industrial implementa-
tion of pyrolysis technology in the near future.

References

[1] Plastic Wastes: Resource Recovery and Recycling in Japan. Tokyo: Plastic Waste Management
Institute, 1985.
[2] Andrews GD, Subramanian PM, editors. Emerging Technologies in Plastics Recycling. ACS Symp
Ser 513. Washington, DC: 1992.
[3] Curlee TR, Das S. Identifying and assessing of opportunity for plastics recycling. Resourc Conserv
Recycl 1991;5:343–63.
[4] Yakowitz H. Incineration of MSW: scientific and technical evaluation of the state-of the-art by an
expert panel. Resourc Conserv Recycl 1990;4:241 – 51.
[5] Bennett RA. Recycled plastics: product applications and potential. In: Andrews GD, Subramanian
PM, editors. Emerging Technologies in Plastics Recycling. ACS Symp Ser 513. Washington, DC:
1992:26–38.
[6] Bahr A. The sorting of plastic wastes. In: Thomé-Kozmiensky KJ, editor. International Recycling
Congress, Berlin, 1979. Berlin: Freitag Verlag, 1979:1202 – 1210.
[7] Satterfield CN. Heterogeneous Catalysis in Practice. New York: McGraw-Hill, 1980.
[8] Gates BC, Katzer JR, Schuit GCA. Chemistry of Catalytic Processes. New York: McGraw-Hill,
1979.
[9] Ferrero GL, Maniatis K, Buekens A, Bridgwater AV, editors. Pyrolysis and Gasification. London:
Elsevier, 1989.
[10] Schoeters JG, Buekens A. Pyrolysis of plastics in a steam fluidised bed. In: Thomé-Kozmiensky KJ,
editor. International Recycling Congress, Berlin, 1979. Berlin: Freitag Verlag, 1979:674 – 680.
180 A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181

[11] Kaminsky W, Sinn H, Doring J, Janning J, Timman H. Raw material recovery from scrap tires and
plastic wastes. In: Thomé-Kozmiensky KJ, editor. International Recycling Congress, Berlin, 1979.
Berlin: Freitag Verlag, 1979:681–685.
[12] Kaminsky W. Recycling of polymeric materials by pyrolysis. Makromol Chem Macromol Symp
1991;48/49:381–93.
[13] Kaminsky W, Pyrolysis of polymers. In: Andrews GD, Subramanian PM, editors. Emerging
Technologies in Plastics Recycling. ACS Symp Ser 513. Washington, DC: 1992:60 – 70.
[14] Cullis CF, Hirschler MM. The Combustion of Organic Polymers. Oxford: Clarendon Press, 1981.
[15] Buekens A, De Wolf F, Schoeters J. PMMA pyrolysis: fundamental and experimental investigation.
In: Ferrero GL, Maniatis K, Buekens A, Bridgwater AV, editors. Pyrolysis and Gasification.
London: Elsevier, 1989:580–585.
[16] Mertens JJR, Jacobs E, Callaerts A, Buekens A. Kinetic study of the thermal decomposition of
polystyrene by means of pyrolysis gas chromatography. Makromol Chem Rapid Commun
1982;3:349–56.
[17] Uemichi Y, Ayame A, Kashiwaya Y, Kanoh H. Gas chromatographic separation of the products
of degradation of polyethylene over a silica-alumina catalyst. J Chromatogr 1983;259:69 – 77.
[18] Uemichi Y, Kashiwaya Y, Tsukidate M, Ayame A, Kanoh H. Product distribution in degradation
of polypropylene over silica-alumina and CaX zeolite catalysts. Bull Chem Soc Jpn 1983;50:2768 –
73.
[19] Uemichi Y, Kashiwaya Y, Ayame A, Kanoh H. Formation of aromatic hydrocarbons in degrada-
tion of polyethylene over activated carbon catalyst. Chem Lett 1984:41 – 44.
[20] Uemichi Y, Makino Y, Kanazuka T. Degradation of polyethylene to aromatic hydrocarbons over
metal-supported activated carbon catalysts. J Anal Appl Pyrolysis 1989;14:331 – 44.
[21] Ide S, Nanbu H, Kuroki T, Ikemura T. Catalytic degradation of polystyrene in the presence of
active charcoal. J Anal Appl Pyrolysis 1984;6:69 – 80.
[22] Ishihara Y, Nanbu H, Ikemura T, Takesue T. Catalytic decomposition of polyethylene using a
tubular flow reactor system. Fuel 1990;69:978 – 84.
[23] Ishihara Y, Nanbu H, Saido K, Ikemura T, Takesue T. Back biting reactions during the catalytic
decomposition of polyethylene. Bull Chem Soc Jpn 1991;64:3585 – 92.
[24] Ishihara Y, Nanbu H, Saido K, Ikemura T, Takesue T. Mechanism of gas formation in
polyethylene catalytic decomposition. Polymer 1992;33:3482 – 6.
[25] Ishihara Y, Nanbu H, Saido K, Ikemura T, Takesue T, Kuroki T. Mechanism of gas formation in
catalytic decomposition of polypropylene. Fuel 1993;72:1115 – 9.
[26] Morid RC, Fields R, Owyer J. Gasoline-range chemicals from zeolite-catalyzed thermal degradation
of polypropylene. J Chem Soc Chem Commun 1992:374 – 375.
[27] Morid RC, Fields R, Owyer J. Thermolysis of low density polyethylene catalysed by zeolites. J Anal
Appl Pyrolysis 1994;29:45–55.
[28] Beltrame PL, Carniti P, Audisio G, Bertini P. Catalytic degradation of polymers: Part II -
Degradation of polyethylene. Polym Degrad Stab 1989;26:209 – 20.
[29] Audisio G, Bertini F, Beltrame PL, Carniti P. Catalytic degradation of polymers: Part III -
Degradation of polystyrene. Polym Degrad Stab 1990;29:191 – 200.
[30] Vasile C, Onu P, Barboiu V, Sabliovschi M, Moroi G. Catalytic decomposition of polyolefins. II.
Considerations about the composition and structure of reaction products and the reaction mecha-
nism on silica-alumina cracking catalyst. Acta Polym 1985;36:543 – 50.
[31] Vasile C, Onn P, Barboiu V, Sabliovschi M, Moroi G. Catalytic decomposition of polyolefins. III.
Decomposition over the ZSM-5 catalyst. Acta Polym 1988;39:306 – 10.
[32] Vasile C, Sabliovschi M, Barboiu V. Catalytic decomposition of polyolefins over various catalysts.
XI. Rev Roum Chim 1995;40:679–91.
[33] Ohkita H, et al. Acid properties of silica-alumina catalysts and catalytic degradation of
polyethylene. Ind Eng Chem Res 1993;32:3112 – 6.
[34] Zhang Z, et al. Chemical recycling of waste polystyrene into styrene over solid acids and bases. Ind
Eng Chem Res 1995;34:4514–9.
[35] Sakata Y, Uddin MA, Koizumi K, Murata K. Catalytic degradation of polypropylene into liquid
hydrocarbons using silica-alumina catalyst. Chem Lett 1996:245 – 246.
A.G. Buekens, H. Huang / Resources, Conser6ation and Recycling 23 (1998) 163–181 181

[36] Sakata Y, Uddin MA, Koizumi K, Murata K. Thermal degradation of polyethylene mixed with
poly(vinyl chloride) and poly(ethyleneterephthalate). Polym Degrad Stab 1996;53:111 – 7.
[37] Uddin MA, Koizumi K, Murata K, Sakata Y. Thermal and catalytic degradation of structurally
different types of polyethylene into fuel oil. Polym Degrad Stab 1997;56:37 – 44.
[38] Ivanova SR, Gumeova EF, Minsker KS, Zaikov GE, Berlin AA. Selective catalytic degradation of
polyolefins. Prog Polym Sci 1990;15:193 – 215.
[39] Wingfield RC Jr, Braslaw J, Gealer RL. Pyrolytic conversion of plastic and rubber waste to
hydrocarbons with basic salt catalysts. United States Patent, US 4,515,659, 1985.
[40] Scott DS. Fast pyrolysis of plastic wastes. Energy Fuels 1990;4:407 – 11.
[41] Sakata Y, Uddin MA, Muto A, Koizumi K, Narazaki M, Murata K. Catalytic degradation of
waste plastics into fuel oil by three-phase reactor. 5th Asia Conference on Fluidized-Bed and
Three-Phase Reactors, 16–20 December, 1996, Hsitou, Taiwan. 1996.
[42] Kirkwood KC, Leng SA, Sims DW. Polymer cracking. European Patent, EP 502,618, 1992.
[43] BP commercialize a thermal plastics recycling route. Chem Eng, May 1992:17.
[44] Plastics recycling: struggles for profitability. Chem Eng, August 1993:23.
[45] Miller A. Industry invests in reusing plastics. Environ Sci Technol 1994;28:16A.
[46] Lechert H, Sung Q, Kaminsky W, Sinn H. Verfahren zum Aufarbeiten von Abfallmaterial.
European Patent, EP 321,867, 1989.
[47] Plastic-waste decomposition/liquefaction system. Report. Nippon Steel, March 1992.

Você também pode gostar