Вы находитесь на странице: 1из 16

I NTERNATIONAL J OURNAL OF C HEMICAL

R EACTOR E NGINEERING
Volume 5 2007 Article A85

A Comparison of Gasification with


Pyrolysis for the Recycling of Plastic
Containing Wastes

Ruby Ray∗ RB Thorpe†


Fluids & Systems, School of Engineering, FEPS (J2), University of Surrey,
ruby che@yahoo.com

Fluids & Systems, School of Engineering, FEPS (J2), University of Surrey,
rex.thorpe@surrey.ac.uk
ISSN 1542-6580
Copyright 2007
c The Berkeley Electronic Press. All rights reserved.
A Comparison of Gasification with Pyrolysis for the
Recycling of Plastic Containing Wastes
Ruby Ray and RB Thorpe

Abstract
In the present investigation, pyrolysis and gasification, two widely used ther-
mochemical processes, are compared as potential chemical recycling methods for
MWP and plastic rich MSW in terms of products of high value and their end uses.
High temperature pyrolysis results in a wide spectrum of products which also con-
tain monomers of C2-C4 range such as ethylene and 1,3-butadiene. Recovery of
monomers from their isomers and other products is difficult and energy-intensive.
Gasification breaks solid wastes into simple molecules (mainly CO & H2 ) which
subsequently can be converted to value added liquid chemicals (namely alcohols)
by a catalytic synthesis processes. Synthetic alcohol then can be converted to
the desired petrochemical precursors. After reviewing different aspects of both
pyrolysis and gasification, recycling through gasification is chosen as the pre-
ferred route for project SPORT as syngas product can be converted into several
key petrochemical products in high yield.

KEYWORDS: MPW, recycling, pyrolysis, gasification


Ray and Thorpe: A Comparison of Gasification with Pyrolysis 1

1. INTRODUCTION

The amounts of mixed plastic wastes (MPW) (including separated auto-shredder waste (ASW) and waste
agricultural plastics (WAP)) and plastic-containing municipal solid waste (MSW) have been increasing steadily in
recent years and disposal of these wastes has become a global issue that has attracted the attention of researchers,
businessmen, politicians, regulators and environmental activists. The world's annual consumption of plastic
materials has increased from around 5 million tonnes in the 1950s to nearly 100 million tonnes today (Waste online,
2007). The amount of plastic waste generated annually in the UK is estimated to be nearly 3 million tonnes (Waste
online, 2007). It is estimated that more than 50% of all plastics that end up in this waste is packaging, three-quarters
of which is from households. Only 7% by weight of total plastic waste is currently being recycled in the UK (Waste
online, 2007). Due to their typically low density, the polymeric wastes occupy a large volume of the waste stream
leading to the problem of disposal.

The disposal options for plastic wastes are either landfilling or incineration. The current recovery options
are energy recovery or recycling. Recycling includes mechanical processes as well as chemical and various
emerging technologies. As almost all plastics are not biodegradable, composting is not an option. The recovery
options for plastics are discussed below.

Mechanical recycling is often regarded as the ideal recovery option for plastics, and this is certainly true for
items such as blow moulded containers and pallet wrap film, which are easily identified by polymer type and
separated from the waste stream. However, this is not the case for the majority of mixed plastic waste. Mixed waste
plastics encompass a wide range of materials with wide ranging properties, developed specifically to suit each
individual application. Mixed Plastic Wastes are mainly composed of low density polyethylene (LDPE), high
density polyethylene (HDPE), polypropylene (PP), polystyrene (PS), polyvinyl chloride (PVC) and polyethylene
terephthalate (PET). Due to the complex nature and the ever increasing use of composite materials and flexible
plastic films, separating individual plastic components, that are suitable for mechanical recycling, is becoming more
difficult, costly and energy intensive. As a consequence, mechanical recycling is not the most sustainable recovery
option for the majority of mixed plastic waste.

Energy recovery, i.e. combustion of mixed waste plastics with the recovery of the energy content is another
proven option, which is widely used for plastic recycling in many parts of the world. As all polymers have high
chemical heat content, energy recovery offers a way of winning back some value from polymer wastes. Carefully
controlled energy recovery processes are above landfilling in the Waste Hierarchy (DEFRA, UK), but there is
persistent public perception issues associated with energy recovery, and there continues to be opponents to this form
of recovery. The concerns appear to be centred on the potential loss of resource and the potential formation of
dioxins and furans from energy recovery processes. There is a clear need for a more sustainable recycling
technology (in terms of economics, environmental performance, and social acceptance) for MWP and plastics-rich
MSW to produce fuels or valuable monomer molecules.

Chemical recycling implies a change of the chemical structure of the material, but in such a way that some
of the resulting chemicals can be used to produce the original polymer(s) again or other valuable chemicals.
Chemical recycling is often used to refer to the various processes for recycling pure waste PET. These include
glycolysis, methanolysis, hydrolysis, saponification and pyrolysis. Costs are often high with chemical recycling
processes and large scale operations are required for them to be economic.

Pyrolysis can also be considered to be chemical recycling process as it cracks the polymeric chains into
smaller molecules without changing their basic chemical structure. Gasification is also a chemically based process
and showing increasing potential as a sustainable recycling option for plastics and other waste materials because it
enables a broader range of carbon materials to be broken down into basic molecules, to be used as building blocks
for a range of new products.

The absolute value of recycling mixed waste plastics through pyrolysis or gasification depends on the
particular products targeted. Different products ranging from monomers to synthetic transport fuel can be obtained
by chemical recycling of MPW and plastic rich MSW. The market prices of these products can vary greatly.

Published by The Berkeley Electronic Press, 2007


2 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

However, the value relative to energy recovery is more consistent and can be expressed relative to oil price.
Representative prices at $40 oil are given in Table 1 below.

Table 1. Margin of chemical recycling over thermal recycling

Product Representative Price @ $40 oil Spot prices, 2005 (Brown, 2005)
Diesel $340/t $447/t
Gasoline $365/t $598/t
Fuel Oil $240/t $356/t
Ethylene $840/t $816/t
Propylene $630/t $760/t
1, 3 Butadiene $530/t $1014/t
Benzene, Toluene, Xylene $540/t $861/t
Styrene $1020/t $1432/t

Pyrolysis and gasification are related technologies as both processes decompose organic waste by exposing
it to high temperatures and limit the amount of oxygen present during decomposition; gasification allows a small
amount of oxygen whereas pyrolysis allows none. With pyrolysis processes, a substantial amount of the intrinsic
chemical energy of plastic waste can be recovered through products for instance monomers and transport fuel. But
pyrolysis, which is an endothermic process, requires significant energy input to drive the process. The heat transfer
issue often restricts pyrolysis to small scale plants. Gasification breaks solid wastes into simple molecules (mainly
carbon monoxide & hydrogen) which subsequently can be converted to value-added petrochemical building blocks
such as monomers. But during gasification, a fraction of the intrinsic chemical energy of the feed plastic is lost due
to CO2 formation. In both processes, the product distribution and consequently the economics of the process are
strongly influenced by the operating conditions used.

The present study is under project SPORT (‘Sustainable Plastics to Olefins Recycling Technology’) which
is funded by the UK DTI. The aim of Project SPORT is to identify and adapt a sustainable recycling process for
mixed plastic wastes that crack the waste plastics back to basic valuable monomer molecules mainly C2-C4
monomers for petrochemical recycling and to established its economical viability. It is estimated that 4% of the
world's annual oil production is used as a feedstock for plastics production and an additional 3-4% during
manufacture of those plastics (Waste online, 2007). The majority of this plastic is not recycled after use but
landfilled. Therefore, efficient waste plastic recycling technology to produce valuable monomer molecules for
petrochemical recycling can play a key role to reduce fossil fuel consumption. The main objective of this paper is a
comparison of pyrolysis and gasification of MWP and plastic rich MSW in terms of products of high value and their
end uses and to suggest the preferred route of chemical recycling of plastic wastes for project SPORT. The review of
this suggestion using more modern decision making tools such as LCA (Life Cycle Analysis) is beyond the scope of
this paper.

2. CHEMICAL RECYCLING PROCESSES

2.1 Pyrolysis
Pyrolysis refers to the thermal decomposition of organic material at elevated temperatures in the absence of oxygen.
Pyrolysis differs from other thermal conversion processes, i.e. gasification and combustion, in that it is an
endothermic process. Yang and Roy (1996) reported that the overall endothermic heat of pyrolysis of NR (natural
rubber) and SBR (styrene-butadiene rubber) are 870 and 550 kJ/kg respectively at 510 °C. Pyrolysis process
produces a mixture of combustible gases (primarily methane, other hydrocarbons, hydrogen and carbon monoxide),
hydrocarbon liquids (contain several constituents over a wide range of molecular weight) and carbon rich solid
residues. The relative quantities of gas, liquid and solid depend on the mode of pyrolysis and operating conditions
like temperature, rate of heating, pressure and residence time. More liquid is produced at lower temperatures
whereas higher temperatures and longer residence times produce more gas. Lower process temperature and longer
vapour residence time favour the production of charcoal (Bridgwater, 2003).

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 3

Pyrolysis offers an attractive technique of chemical recycling of mixed plastics waste. It can decompose
long polymeric chains of the plastic to reusable monomers, hydrocarbon fuels having carbon number distribution in
the range of C1 – C50 and valuable aromatic solvents like benzene, toluene etc. The product distribution of pyrolytic
reaction of polymeric material highly depends on the nature of the feed plastic waste and operating conditions
mainly temperature. Changing these variables, it is possible to obtain a predominantly liquid hydrocarbon product
with potential for use as a fuel or a mixture of gaseous and liquid products that can be used as a petrochemical
building block (Mader and Mennicken, 1995).

Thermal degradation of plastic materials follows both consecutive and competing reaction paths to give a
wide array of products and intermediates. The sequences of reactions are complex in nature and involve both
endothermic and exothermic processes. Several investigators (Bockhorn et al., 1999; Zhou et al., 2006; Saha et al.,
2007) have reported the overall kinetic parameters for the thermal decomposition of individual plastics namely
waste low-density polyethylene (LDPE), polypropylene (PP), waste polyethylene terephthalate (PET), polyvinyl
chloride (PVC) etc, but work on pyrolysis kinetics of mixed plastic waste has yet to be done. The mechanism of
thermal decomposition of plastics has been expressed by the investigators (Bockhorn et al., 1999; Zhou et al., 2006;
Saha et al., 2007) using formal degressive rate expressions


= k (T )(. 1 − α )
a
(1)
dt

Where α is the degree of conversion which is defined as α = (m 0 − m ) / (m 0 − mα ) with m0 being the initial
mass, m the actual mass and mα the final mass; k is the reaction rate constant and a the apparent order of the overall
decomposition reaction. Introducing the Arrhenius type temperature dependency for the rate constant,
k = k 0 e − E / RT and the heating rate, β = dT / dt , Equation (1) is transformed into

dα  k 0  − E / ( RT )
=  .e .(1 − α )a (2)
dt  β 

The overall kinetic parameters k0, E and a have been evaluated using Equation (2). Though most of the
researchers reported the kinetic parameters for the decomposition of plastics, Bockhorn et al. (1999) also studied
kinetics of formation of products like HCl, benzene from the isothermal degradation of PVC.

During the present study, we have examined a number of commercially available polymeric pyrolysis
process options which could produce valuable monomer molecules mainly C2-C4 range for project SPORT. It is
observed that most of the commercial organisations like Cynar Plc, Alphakat Technologies, Gossler Envitec,
offering pyrolysis technology to the investor, predominantly target low temperature pyrolysis of waste plastic to
produce high quality synthetic fuels namely gasoline, diesel and fuel oil. A pyrolysis temperature within the range of
350-450 oC is the best operating temperature for plastic cracking to valuable liquid petroleum fuels. It is understood
(Ballice, 2002; Lee and Shin, 2007) that at low temperature, polymeric feedstock primarily devolatilize to produce
high molecular weight hydrocarbons with carbon number ranging from C5–C50. The plastic cracking process of
Cynar Plc, known as the ThermoFuel process, is a process whereby scrap and waste plastics are gently 'cracked' at
relatively low temperatures (370-420 ºC) to give synthetic fuel with an average carbon range similar to diesel oil.
The Alphakat Technologies claims to produce synthetic diesel in one step by catalytic depolymerisation of
hydrocarbon containing waste material such as mixed waste plastic with a reactor temperature of 350 ºC
(Krzesinski, 2006). Gossler Envitec has developed thermal-catalytic cracking processes with temperature range of
350-400 oC for the recycling of polyolefin’s like polypropylene and polyethylene. Their polyolefin recycling
technology mainly produces synthetic fuel (gasoline, diesel and fuel oil). Noting the product range, low temperature
pyrolysis is not a desirable option for project SPORT but because of the high current value of transport fuel, it does
seem commercially interesting.

In high temperature pyrolysis, where the temperature varies within the range of 650-850 oC, secondary gas
phase cracking occurs and the primary devolatilization products are cracked to result a wide spectrum of smaller
hydrocarbons (ranging from C1–C9) like alkenes, alkanes, aromatics and gaseous by-products. Several investigators

Published by The Berkeley Electronic Press, 2007


4 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

(Arena and Mastellone, 2000; Herna´ndez et al., 2005; Mastral et al., 2002; Sodero et al., 1996) have studied various
aspects of high temperature pyrolysis of polymers mainly polyolefin’s which is the major fraction of mixed plastic
waste. High temperature pyrolysis of plastic rich in polyolefin’s typically produces 65-85% gases, 18-22% liquid
and the rest soot (Sodero et al., 1996). A typical composition of the gas is 4-10% methane, 22-31 wt% ethylene, 1.5-
2.7 wt% ethane, 10-14 wt% propylene, 0.5-1 wt% propane, 4.5–5 wt% n/iso-butene, 0.3-0.6 wt% cis/trans-2-butene,
6-7 wt% butadiene (mostly 1,3-butadiene) (Sodero et al., 1996). The temperature and residence time have a great
influence on the product distribution in high temperature pyrolysis. Mastral et al. (2002) obtained their highest gas
yield of 86.4 wt% by pyrolysing HDPE at 780°C. Above 850°C, the gas yield decreased due to cyclation reactions
that formed aromatic hydrocarbons. They also reported that yield of methane increases with increasing residence
time at higher temperature whereas fractions of C3 and C4 compounds decrease with increasing residence time.
However, C2H2 and C2H6 yields were not affected by the residence time. So, the gaseous products contain valuable
monomers of C2-C4 range like ethylene, propylene and 1,3-butadiene for petrochemical recycling, but the recovery
of the valuable monomers from their isomers and other products is difficult, costly and energy-intensive. For
instance, recovery of propylene from propane and 1,3-butadiene from n/iso butane would be very difficult by
distillation due to their close boiling point (Table 2). In fact this problem is well understood which is why olefin
units have dedicated distillation and solvent extraction units for separating these monomers from the streams of
close boiling compounds. These separations are expensive and need to be performed at scale to be economic. Where
it is possible to integrate a pyrolysis plant with a refinery or olefins unit, the additional capital cost of the light gas
separations could be minimised. However, transporting plastic waste over long distances to refineries is not an ideal
option because it implies high transport costs and environmental impact. Due to high calorific value, gaseous
products can be used as fuel, but this is also not the most desired end use for project SPORT. Liquid product mainly
contains 32-35% aliphatics (mainly C5-C9 hydrocarbons) and 65-68% aromatics namely benzene, toluene, styrene,
naphthalene etc (Day et al., 1999; Kaminsky and Kim, 1999). Day et al. (1999) reported that the yield of the
aliphatic compounds falls off rapidly as the pyrolysis temperature increases while that of the aromatic compounds
remain relatively constant. As C5-C9 hydrocarbons are gasoline range products, they can be mixed after treatment
into the gasoline pool. Some valuable chemicals can be recovered from this mixture if their percentage is high
enough. As the actual product distribution of C5-C9 range liquid obtained form pyrolysis of mixed plastic is not
available in literature or from any commercial organisation, it is very difficult to identify the quantity of any
valuable compounds for petrochemical recycling.

Aromatic fraction of liquid product from plastic pyrolysis is of much interest from recycling point of view.
Day et al. (1999) reported that the mono-aromatics are the most dominant organic compounds identified in the
pyrolysis liquid at high temperature, with benzene, toluene, naphthalene, indene and styrene being the major
components. Other aromatics such as ethylbenzene, indene, p-xylene are also detected in the aromatic mixture.
Among all these mono-aromatics, styrene is the most valuable monomer. But the close boiling point of styrene to
ethylbenzene, and other components preclude the use of classical distillation as a separation method. A solvent-
based extractive distillation system, which is complicated and costly, can extract and purify styrene. Benzene and
toluene are also valuable products. But from boiling point table (Table2), it is clear that benzene, when recovered by
distillation, would be associated with other C6 compounds e.g. cyclohexene, cyclohexane. Toluene would also be
associated with other close boiling compounds e.g. 2,5-dimethylhexane, 2,4-dimethylhexane, cyclopentane etc if
they present in the liquid mixture. To avoid such contamination, general practices of recovering BTX aromatics are
by either liquid-liquid extraction or solvent-based extractive distillation.

Moreover, in project SPORT we are interested in recycling ASR as well. Automobile shredder residue
(ASR), highly heterogeneous polymeric waste stream produced by shredding operations during the recycling of
automobiles, contains not only different plastics but rubber, foam, textiles and metals. Product distribution of
pyrolysis of ASR is more complicated. Pyrolysis of ASR produces more solid residue than other polymeric waste.
Day et al. (1996) reported that solid residue is the major product of ASR pyrolysis process, ranging from 80% at
500°C to 70% at 750 °C whereas gas and liquid product in pyrolysate are of 20% and 10% respectively. The solid
residues can be considered to be composed of three components. An organic fraction composed of organic material
(not fully degraded polymeric species and high molecular weight products), a carbonaceous char and inorganic ash
(including metals residues). Therefore, a major part of the intrinsic chemical energy in ASR is lost in solid residue.
Disposal of this solid residue is a major environmental concern. Another major concern with ASR pyrolysis
processes is the possible formation of carcinogenic compounds of environmental concern such as polycyclic
aromatic hydrocarbons (PAH’s), polychlorinated biphenyls, dioxins and furans. Rausa and Pollesel (1997)
pyrolyzed ASR at 650, 800 and 850 °C and detected several PAH’s as well as phenol, benzaldehyde, and

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 5

chlorobenzene as pyrolysis products. At 650°C, the PAH’s accounted for 1.3% of the pyrolyzate while at 800 and
850 °C they increased to 32.3 and 55.1%, respectively. Hence, pyrolysis of ASR does not appear to be an attractive
option.

Table 2. Boiling points of monomers, BTX and close boiling compounds thereto

Components Normal boiling point, Components Normal boiling point,


(oC) (oC)
Ethylene -103 2,4-dimethylpentane 80.5
Ethane -88 Cyclohexane 80.7
Propylene -47.8 Cyclohexene 82.9
Propane -42.1 Ethylcyclopentane 103.4
n/iso Butene -6.3 2,5-dimethylhexane 109.1
1,3-Butadiene -4.5 2,4-dimethylhexane 109.4
Trans-2-butene 0.8 Toluene 110.6
Cis-2-butene 3.7 3,3-dimethylhexane 111.9
1,2-Butadiene 10.8 cyclopentane 118.7
Methyl-cyclopentane 71.8 Ethylbenzene 136.1
2,3-dimethyl-2-butene 73.2 p-Xylene 138.3
Benzene 80.1 Styrene 145.1

2.2 Gasification
In thermal gasification, combustible material is partially oxidized into a gaseous energy carrier in a high-temperature
reducing atmosphere, using air, steam or oxygen as the gasification agent. Gasification occurs in a number of
sequential steps:
• Drying to evaporate moisture
• Pyrolysis to produce volatiles, vaporised tars or oils and a solid char residue
• Gasification or partial oxidation of the solid char, pyrolysis tars and pyrolysis gases

Typically, the exothermic reaction between carbon and oxygen provides the heat energy required to drive
the pyrolysis and endothermic char gasification reactions. The main prevailing reactions along with their kinetic
information occurring in the gasifier are given in the Table 3. This kinetic information could be useful in the sizing
of a gasifier. The key components of product gas, which is commonly known as synthesis gas, include H2, CO, CO2,
CH4 and N2 (if air is used as the gasification agent).

Attractive features of gasification include: the ability to produce a consistent synthesis gas product and the
ability to accommodate a wide variety of feedstock. Gasification of waste has many potential benefits when
compared with conventional options such as incineration. One of them is that highly toxic dioxin and furan
compounds are not expected to be present in the synthesis gas (Orr and Maxwell, 2000). This is for the two
following reasons. First, the reducing gas environment precludes the formation of the free chlorine from HCl, thus
preventing chlorination of any dioxin precursor in the syngas. Secondly, the high temperature in the gasification
process effectively destroys any dioxin/furan compounds or their precursors in the feed. The reducing atmosphere in
gasification system also prevents the oxidation of sulphur and nitrogen compounds to SOx and NOx (Orr and
Maxwell, 2000).

The gasification process has also few downsides. It produces CO2 which is a detrimental greenhouse gas
and a fraction of the intrinsic chemical energy of the feed waste is also lost through this CO2. Moreover, synthesis
gas from a waste gasifier contains several impurities like particulates (entrained ash, char, fluid bed material etc.),
tars (refractory aromatics), hydrogen halides, volatile heavy metals and heavy metal compounds (e.g. Hg, Pb, Cd)
and alkaline compounds, depending on the fuel composition, mode of gasification and the type of gasifier used.
Many of these are toxic substances that pose ecological and human health risks if released into the environment.
Because of these impurities, the syngas needs to pass through extensive gas cleaning systems before being used by
subsequent downstream processes.

Published by The Berkeley Electronic Press, 2007


6 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

Table 3. Kinetic data of main prevailing reactions occurring in the gasifier

Reactions Rate constant Heat of reactions


(kJ/mol)
Oxidation 407.4
 160 x10 3  -1
C + O 2 → CO2 k = 5.67 × 10 9 × exp − s (Levenspiel, 1982)
 RT 
(Di Blasi, 2000)
Gasification
Boudouard reaction:  22409.64  -1 -173.8
C + CO2 → 2CO k = 1.6 × 10 12 × exp − s
 T  (Caram and
Amundson, 1977)
(Freund, 1986)
Steam carbon reaction:  17500  -1 -131.0
k = 1.33 × 10 3 T × exp − s
C + H 2 O → CO + H 2  T  (Arai and Hasatani,
(Arai and Hasatani, 1987) 1987)

k = 5.14 × 10 4 × exp 3.995 × 10 − 4 T 


Water gas shift reaction:
43.47
CO + H 2 O → CO2 + H 2   (Arai and Hasatani,
(kmol/m3)-1s-1 (Arai and Hasatani, 1987) 1987)

 2762 
CO2 + H 2 → CO + H 2 O k = 2.63 × 10 6 × exp 3.995 × 10 −4 T −  -43.47
 T  (Arai and Hasatani,
(kmol/m3)-1s-1 (Arai and Hasatani, 1987) 1987)

Air gasification produces a poor-quality gas in terms of calorific value (4-7 MJ/m3 higher heating value)
which is suitable as a fuel for boilers, engines and turbines, but is not suitable for use as synthesis gas for conversion
to liquid products due to its high N2 content. Gasification with pure oxygen results in a medium heating value gas
with a higher quality mixture of carbon monoxide and hydrogen and virtually no nitrogen but has drawbacks; high
production cost of oxygen and hazards associated with handling of oxygen during gasification. Gasification with
steam is more commonly called “reforming” and results in a hydrogen and carbon dioxide rich synthesis gas. The
main disadvantage of steam gasification is the quantity of heat that needs to be supplied to the reaction and
consequently, the complexity of the reactor (Bridgwater, 1995).

The syngas gas from oxygen or steam gasification, which is rich of CO and H2 can be used as a chemical
building block from which a wide assortment of commercial chemicals may be manufactured. A major use of
syngas is for methanol manufacture. Syngas conversion to methanol is an well known process where the conversion
takes place in a high pressure (80 – 100 barg) fixed bed reactor using a copper based catalyst at around 240 oC.
Methanol is a commodity chemical used as a feed to many manufacturing processes including Methanol to Olefins
and Methanol to Propylene. Syngas can also be converted to mixed alcohols by gas-to-liquid catalytic conversion
processes e.g. processes using ultrafine Mo–Co–K catalysts (Zhang et al., 2001) or molybdenum sulphide
(K/Co/MoS2) as catalyst (Nirula, 1996). The general stoichiometry for alcohol formation may be represented as
follows:

nCO + 2nH2 → CnH2n+lOH + (n-l)H2O (3)

Alcohol synthesis reactions are highly exothermic (∆H = -101.74 kJ/mol CO converted). The exothermic
heat of reaction can be used for steam generation which steam can subsequently be utilized to produce electricity.
Alcohol synthesis reactions are always associated with the equilibrium water gas shift reaction:

CO + H2O ↔ CO2 + H2 (4)

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 7

By choosing an appropriate catalyst, greater flexibility in terms of methanol/higher alcohols ratios, lower
water content in the reaction product and higher selectivities to the desired liquid alcohol products in the C2-C4
range can be achieved.

Dow/Union Carbide (Nirula, 1996) developed a modified Fischer-Tropsch catalyst system based on
molybdenum sulphide for their process of mixed alcohols from syngas. The following table (Table 4) reveals the
product distribution based on 40% CO conversion obtained in Dow/Union Carbide patent.

Table 4. Product selectivities (mole%) based on 40% CO conversion (Nirula, 1996)

Product Mole%
Methane 9.45
Ethane 0.75
Methanol 13.46
Ethanol 28.08
Propanols 9.32
Butanols 3.13
C5 alcohols 1.63
Methyl acetate 1.16
Ethyl acetate 1.02
Carbon dioxide 32.00

The above selectivity corresponds to 85% liquid product (alcohols and esters) and 15% gas product on a
CO2-free basis. The liquid product contains 41.3 % ethanol and 13.7 % propanol. Methanol, ethanol and propanol
are of particular interest as routes are available to convert these alcohols into important petrochemical monomers
such as ethylene and propylene.

Plastics are good feedstock candidate for gasification process because many plastics have high calorific
values and simple chemical structure composed primarily of carbon and hydrogen. But because of the size and the
shape of the plastic products, the size reduction is necessary to produce a feed of high bulk density to facilitate ease
of feeding.

In the present study, we have reviewed a number of commercially available gasification processes that can
handle mixed plastic waste and plastic rich municipal solid waste. Different methods (Thermoselect process, Texaco
gasification process, BASF conversion process, the Akzo Nobel steam gasification process, the Linde gasification
process etc) have been developed in this field which claim to process mixed plastic waste and plastic rich municipal
solid waste. For instance, the Themoselect process (Drost et al., 2004) claims to be an innovative high temperature
waste recycling technology that can efficiently gasify mixed waste plastics mainly automotive shredder residues
(ASR) mixed with MSW. Solid wastes (45%ASR and 55%MSW) can be continuously processed in fixed bed
oxygen blown gasifier and residue melting reactor to achieve a maximum recovery of recyclable raw materials, with
simultaneous utilization of the chemical energy contained within the waste material.

3. COMPARISON

After reviewing thoroughly both pyrolysis and gasification process for chemical recycling of mixed waste plastic, it
appears that both have positive and negative aspects.

Low temperature pyrolysis of mixed plastic waste mainly produces synthetic transport fuel, which is not
the desired product for project SPORT, but does currently have a reasonably high value. High temperature pyrolysis
of polyolefin rich mixed plastic produces valuable monomers like ethylene, propylene, butadiene and styrene. By
choosing proper operating conditions, namely temperature and residence time, it may be possible to produce
reasonable fractions of valuable monomers for petrochemical recycling. However, the main problem associated with
pyrolysis is its complex product distribution and product purification requirements. The gaseous products are well
identified, but separation of valuable monomers like propylene and butadiene from their isomers and other

Published by The Berkeley Electronic Press, 2007


8 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

compounds is cost and energy intensive. An additional material issue is the need to remove acidic gases eg. HCl
from the product stream. The liquid product (C5+) consists of a heterogeneous mixture of large number of
compounds and proper analysis of this steam to identify all compounds has not been published by any organisation.
A few of the aromatic compounds are named in literature, like styrene, benzene, toluene, but separation processes,
which would involve solvent extraction and extractive distillation due to the presence of close boiling compounds in
the mixture, is costly and complicated and cannot be designed without a proper product analysis.

As the maximum practical size for a plastic processing plant is determined by the volume of suitable wastes
that can be economically collected and transported to the plant and since the bulk density of plastic wastes is low
and the transport costs high, even wastes generated comparatively close to the plant may be uneconomic to recover.
That is why project SPORT is not limited to polyolefin rich mixed plastics feed. We are considering ASR and
plastic rich MSW as well. But, ASR pyrolysis produces mainly solid residue which contains substantial amount of
carbonaceous compounds. Efficiency of ASR pyrolysis process decreases as substantial amount intrinsic chemical
energy of ASR is lost as solid residue. Moreover, it produces carcinogenic compounds of environmental concern
such as polycyclic aromatic hydrocarbons (PAH’s), polychlorinated biphenyls, dioxins and furans etc.

Another difficulty with pyrolysis is that it is a less proven technology for mixed waste plastic on large
scale. We reviewed several commercially-available pyrolysis process during this study but none of them are proven
at anywhere near the desired scale (>100 kt p.a.) of neither project SPORT nor do they have a track record of
reliable near continuous operation. For instance, ThermoFuel process of Cynar Plc can process only 10-20 tonnes of
waste plastic per day i.e. 3.65-7.30 kt p.a.

On the other hand, gasification is a well defined technology and is already commercially available on the
large scale. Information on the analysis of the gas product stream is available in the literature and in more and
sufficient detail on a commercially confidential basis. Gasification can handle different types of feedstock whether
it is mixed waste plastic including ASR or plastic rich MSW. Gasification converts the carbonaceous part of the
solid wastes into synthesis gas which consists of simple molecules (mainly CO & H2). It also produces solid residue,
which contains mainly the inorganic part of the feedstock (a slag) and can be used as building material. CO & H2
can be converted to value added liquid chemicals (namely alcohols) by a catalytic synthesis processes as explained
above. The product obtained by this catalytic synthesis process is not complicated, it is well defined. By choosing
appropriate catalyst, higher selectivities to alcohol products in the C1-C4 range can be achieved which then can be
utilised in a number of petrochemical processes.

The gasification process is self sustained as no external heat is required to drive the process. As gasifier
operates at high temperature, synthesis gas is generally available at 1100-1200 oC. The heat energy of synthesis gas
can be utilized to produce hot water or steam for electricity generation. This energy exaction increases the overall
efficiency of the gasification process. During the present study, a flowsheet has been developed of the conversion of
syngas to mixed alcohol using the Aspen Plus simulation package (Aspen flowsheet and discussion are in Section
4). The syngas composition and temperature is based on a process suitable for the gasification of the mixed plastic
waste, the Thermoselect process (Drost et al., 2004). From the flowsheet, we have estimated the electricity generated
utilising heat energy of hot synthesis gas and other hot process stream. The estimated electricity generated by the
process is 713 kWh/t of Thermoselect waste feed.

However, chemical recycling through gasification has several downsides as well. The best form of
gasification with respect to downstream processes is oxygen blown gasification, but this requires oxygen, which is
expensive and has to be separated from air with a fairly high energy cost reported as 250 kWh per tonne of oxygen
produced.

Petrochemical monomers (ethylene & propylene) obtained via a gasification route involves several
significant and separate processing steps, namely
• MSW feed preparation
• Gasification
• Syngas cleanup
• Alcohol synthesis reaction
• Separation of desired alcohols

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 9

• Further conversion of alcohol to valuable monomers

An extensive gas cleanup system is necessary as impurities in synthesis gas may deactivate the catalyst
used in alcohol synthesis process. The combined effect of all the necessary processes may increase the monomer
production cost. Another issue is that gasification produces significant amount of CO2 which has detrimental
greenhouse effect. But if we consider the CO2 produced in supplying endothermic heat of reaction associated with
pyrolysis process, the value appears to be comparable or sometimes more than the CO2 emitted from gasifier
depending upon the feed composition. We have calculated the typical CO2 emission from power plant to generate
electricity required to supply the endothermic heat (~ 1390 kWh/t of waste) of pyrolysis of polyolefin (50%
polyethylene and 50% polypropylene) feed pyrolysing at 690 0C. Calculated CO2 emission associated with
polyolefin pyrolysis is ~1.25 kg/kg of polyolefin feed. Whereas, Thermoselect process reported that CO2 emission
from gasifier processing solid waste (45% ASR and 55% MSW) feed is ~0.5 kg/kg of feed (Drost et al., 2004).

Finally we present a comparison in terms of energy and product of pyrolysis and gasification. We have
information for the pyrolysis of polyolefins, a very favourable but unrealistic feed for project SPORT. Whereas we
do have information for the gasification of a more general, more plentiful, mixed waste stream that contains plastic
which is one of the proposed feed for project SPORT. Based on all the information outlined in the paper, we have
constructed Table 5. Though we have tried to present a reasonable comparison in terms of energy and product of
pyrolysis and gasification process in Table 5, we have to keep in mind the different feed for the processes. We
cannot perform a fair economic comparison of pyrolysis with gasification as detailed product analysis of pyrolysis
of mixed plastic waste or plastic rich MSW is not available either in literature or from any commercial organisation.
Unless we know the exact product distribution from pyrolysis process, it is also not possible to design proper
downstream processes such as gas cleanup system, product separation processes etc.

Table 5. Comparison in terms of energy and product


Pyrolysis of polyolefins Gasification of a mixed waste
Energy Comparison

Energy required to produce O2 - 250 kWh/t O2 produced or 114.75


kWh/t of waste (based on
Thermoselect Process)

Heat energy required to drive ~1390 kWh/t of waste -


the process

Electricity generated utilizing - 713 kWh/t of waste (based on


heat energy of product Thermoselect process)

Product Comparison High Temperature Pyrolysis Product Alcohol Products (approximate


(approximate wt%) wt%) based on 40% CO
conversion

Ethylene (26 wt%) Methanol (14.3 wt%)


Propylene (12 wt%) Ethanol (43 wt%)
Butadiene (6.5 wt%) Propanol (18.6 wt%)
Styrene (1.6 wt%) Butanol (7.7 wt%)
Benzene/Toluene/Xylene (9 wt%) C5-alcohols (4.8 wt%)

Low Temperature Pyrolysis Product

Gasoline
Diesel
Fuel Oil

Published by The Berkeley Electronic Press, 2007


10 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

4. SIMULATION OF MIXED WASTE PLASTIC TO OLEFINS PROCESS BASED


ON GASIFICATION

In the present study under project SPORT, recycling of mixed waste plastic via gasification is considered as the
suitable route. The Thermoselect gasifier (Drost et al., 2004; Drost and Kaiser, 2004) which claims to be suitable for
processing mixed waste plastic including ASR or plastic rich MSW, has been chosen for base case study. The
Thermoselect process is a proven technology available in a large scale demonstration facility and commercialised.
Commercial plants are in operation in Karlsruhe, Germany, in Tokyo-Chiba, Japan and in Mutsu, Japan.

The whole process is explained in great detail by Drost et al. (2004) and Drost and Kaiser (2004). The
gasifier used in the process is fixed bed oxygen blown gasifier with a residue melting reactor at the bottom to
achieve a maximum recovery of recyclable raw materials. Prior to its gasification, the waste is processed in a
degassing channel into dense plugs or briquettes. Radiated heat from the gasification reactor initiates the waste
drying and decomposition processes in the degassing channel. The dried and charred briquettes emerge from the
degassing channel into the gasifier and are exposed to steam (from water in the waste) and controlled injection of
pure oxygen as the gasification medium. All organic materials in the waste are transformed into a synthesis gas with
a composition that reflects the thermodynamic equilibrium at the top of the reactor (approximately 1200 °C, 1.2
bar). The metals and minerals continue to move to the bottom of the reactor which is a residue melting reactor
where they are melted at 2000 °C by further injection of natural gas and oxygen. The liquid metals in the effluent
from the residue melting reactor are recovered as two separate phases.

The composition of synthesis gas obtained from the Thermoselect gasifier at 1200 °C is 35 vol% H2, 35
vol% CO, 25 vol% CO2 and rest nitrogen. Synthesis gas also contains several impurities such as K, P2O5, H2S, HCl,
Hg, Zn etc. Because of these impurities, the syngas needs to pass through extensive gas cleaning systems before
being used by subsequent downstream processes. Though Thermoselect process uses wet gas cleaning system, in
project SPORT dry gas cleaning process involving intimately mixing the flue gas stream with a solid/liquid
sorbent/reactant that fixes the pollutant gases in a solid particle which is then collected, along with any process
particulates/fly ash, on a fabric/ceramic filter (bag house filter) has been proposed.

During the present investigation, a flowsheet of conversion of syngas to mixed alcohol using the Aspen
Plus simulation software has been developed (Figure1). As alcohol synthesis reactor operates at high temperature
and high pressure, clean syngas coming from gas cleaning system are compressed using multiple compressors with
intercoolers. CO2 is removed from the syngas which is nothing but a diluent and also increasing the volume of the
alcohol synthesis reactor. Mixed alcohol products have been separated by means of distillation. During the
development of the base case flowsheet of mixed alcohol process, purge gas from the reactor recycle has been
evolved which has been utilized for power generation using a combination of gas turbine and steam turbines.
Sensible heat energy of the hot synthesis gas coming out from the gasifier is also used to generate the saturated high
pressure steam (62 bar) which is subsequently utilized in high pressure steam turbine for electricity generation.
Saturated HP steam at 62 bar has also been raised using heat energy released by alcohol synthesis reactor. From the
Aspen simulation, we have estimated the electricity generated (Table 6) using the combination of gas turbine and
low pressure and high pressure steam turbines. The estimated electricity generated by the process is 713 kWh/t of
Thermoselect waste feed with a feed rate of 45.66 t/hr. Due to the commercial confidentiality, we can not reveal all
the process details in this paper.

Table 6. Electricity generated in the syngas conversion to mixed alcohol process

Equipments Electricity generated


Gas turbine 9.05 MW
High pressure steam turbine 7.64 MW
Low pressure steam turbine 4.75 MW

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 11

5. CONCLUSION

Comparing pyrolysis and gasification and reviewing different aspects of them, recycling through gasification is
chosen as the preferred route for project SPORT. The main driving forces towards gasification are its simple and
well defined product (syngas) and its availability in large scale in commercial operation. Syngas can be converted to
mixed alcohol products by a catalytic synthesis processes. Mixed alcohol products obtained from alcohol synthesis
reactor is also well defined and choosing proper catalyst, it is also possible to obtain higher selectivity towards C1-
C4 alcohols which subsequently can be converted into several key petrochemical products in high yield. As detail
product distribution is also available for mixed alcohol products, proper design of downstream separation processes
is possible.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the financial support of the UK DTI (Department of Trade and Industry) for
project SPORT.

NOTATION

E activation energy, J/mol


k reaction rate constant, s-1
R universal gas constant, J mol-1 K-1
t time, s
T temperature, K

Subscripts
o initial

Published by The Berkeley Electronic Press, 2007


12 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

Combusted gas
to s tack
Heat recovery from hot syngas & HP Steam raising
44
Syngas to gas cleaning system Saturate d s team at 3.6 bar
BFW at 3.6 bar 42 43
Superhe ate d s team
Boiler1 at 3.6 bar
50 35 Mixer3 Splitter2
36 37
Boiler2 HE4
BFW at 62 bar 49 Gas Turbine
32 38

Saturated steam at 62 bar


Air 28 Superhe ate d s team
30

LP steam to alcohol
at 62 bar LP steam
34

seperation units
29 33 turbine
51 Compressor Turbine
Combustor HE3
Syngas from 48 31 40
gasifie r at 1200 oC HP steam
27 turbine
Mixer2 41 Water
47 (BFW re turn)
Cooler2

Valve2
18
46
Off gas recycle Splitter1
Steam raising in Alcohol Synthesis Reactor 19 17

BFW at 62 bar 45 Cooler1

Off gas
HE5 Saturate d s team at 62 bar 14 15
Compressor1
Mixed alcohol products 39
HE2
20

Flash1
9
Multiple compressor Multiple compressor 21
CO2 absorber with intercooler Water separator
with intercooler
4 6 8 10
1 3 11 16
Mixer1 HE1 Seperated
Cle an Syngas 5 Alcohol Synthesis Mixed alcohol products
from Gas cleaning s ystem 2 alcohol products
7 Reactor MULTIFRAC
13
water Alcohol s eperation
water CO2 water units
12
Mixed alcohol products

Figure1: Aspen flowsheet of syngas conversion to mixed alcohol products along with heat integration

http://www.bepress.com/ijcre/vol5/A85
Ray and Thorpe: A Comparison of Gasification with Pyrolysis 13

REFERENCES

Arai, N., Hasantani, M., “A mathematical model of Simultaneous formation of volatile - NO and char – NO during
the packed bed combustion of coal char particles under an NH3 - O2 -Ar gas stream”, Fuel, Vol. 66, 1418-1426
(1987).

Arena, U., Mastellone, M. L., “Defluidization phenomena during the pyrolysis of two plastic wastes”, Chem. Eng.
Sc., Vol. 55, 2849-2860 (2000).

Ballice, L., “Classification of volatile products evolved during temperature programmed co-pyrolysis of low density
polyethylene (LDPE) with polypropylene”, Fuel, Vol. 81, 1233-1240 (2002).

Bockhorn, H., Hornung, A., Hornung, U., “Mechanisms and kinetics of thermal decomposition of plastics from
isothermal and dynamic measurements”, J. Anal. Appl. Pyrolysis, Vol. 50, 77–101(1999).

Bridgwater, A.V., “The Technical and Economical Feasibility of Biomass Gasification for Power Generation”, Fuel,
Vol. 74, No. 5, 631-653 (1995).

Bridgwater, A.V., “Renewable fuels and chemicals by thermal processing of biomass”, Chem. Eng. J., Vol. 91, 87-
102 (2003).

Brown, R., “Chemical Commodities Monthly”, Chemical Market Reporter, September, 2005.

Caram, H. S., Amundson, N., “Diffusion and reaction in a stagnant boundary layer about a carbon particle”, Ind. &
Eng. Chem. (Fundamentals), Vol. 16, (2), 171 (1977).

Cynar PLC UK - Recycling waste plastic to synthetic diesel fuel. URL: http://www.cynarplc.com/thermo-fuel.asp

Day, M., Cooney, J.D., Shen, Z., “Pyrolysis of automobile shredder residue: an analysis of the products of a
commercial screw kiln process”, J. Anal. Appl. Pyrolysis, Vol. 37, 49-61 (1996).

Day, M., Shen, Z., Cooney, J.D., “Pyrolysis of auto shredder residue: experiments with a laboratory screw kiln
reactor”, J. Anal. Appl. Pyrolysis, Vol. 51,181–200 (1999).

DEFRA, UK – Environmental Protection - Recycling and waste.


URL: http://www.defra.gov.uk/environment/waste/topics/waste-hierarchy.pdf

Di Blasi, C., “Dynamic Behavior of Stratified Downdraft Gasifiers”, Chem. Eng. Sc., Vol. 55, 2931 (2000).

Drost, U., Eisenlihr, F., Kaiser, B., Kaiser, W., Stahlberg, R., “Report on the operating trial with automotive
shredder residue (ASR)”, 4th International Automobile Recycling Congress, Geneva, Switzerland, 1-20 (2004).

Freund, H., “Gasification of carbon by CO2: a transient kinetics experiment”, Fuel, Vol. 65, 63-66 (1986).
Gossler Envitec. http://www.gossler-envitec.de

Herna´ndez, M.R., Garcıa, A.N., Marcilla, A., “Study of the gases obtained in thermal and catalytic flash pyrolysis
of HDPE in a fluidized bed reactor”, J. Anal. Appl. Pyrolysis, Vol. 73, 314–322 (2005).

Kaminsky, W., Kim, J.S., “Pyrolysis of mixed plastics into aromatics”, J. Anal. Appl. Pyrolysis, Vol. 51, 127–134
(1999).

Krzesinski, E., “Transforming biomass to Diesel fuel by catalytic depolimerization. Low pressure, low temperature,
low costs. The KDV-method from Alphakat, Germany”, World Bioenergy, Sweden, 1-4 (2006).

Published by The Berkeley Electronic Press, 2007


14 International Journal of Chemical Reactor Engineering Vol. 5 [2007], Article A85

Lee, K.H., Shin, D.H., “Characteristics of liquid product from the pyrolysis of waste plastic mixture at low and high
temperatures: Influence of lapse time of reaction”, Waste Management, Vol. 27, 168–176 (2007).

Levenspiel, O., Chemical Reaction Engineering, 2nd U.S. Edn. Wiley eastern Ltd. (1982).

Mader, F., Mennicken, T., “The Feedstock Option”, Shell Chemicals Europe Magazine, 20-23 (1995).

Mastral, F.J., Esperanza, E., Garcı´a, P., Juste, M., “Pyrolysis of high-density polyethylene in a fluidised bed reactor.
Influence of the temperature and residence time”, J. Anal. Appl. Pyrolysis, Vol. 63, 1–15 (2002).

Nirula, S.C., “Dow/Union carbide process for Mixed alcohols from Syngas”, Process Economics Review, PEP ’85-
1, March, 1-27 (1996).

Orr, D., Maxwell, D., “A Comparison of Gasification and Incineration of Hazardous Waste”, final report DCN
99.803931.02, NETL, US Department of Energy (2000).

Range Fuels – Biomass to Energy. URL: http://www.rangefuels.com/

Rausa, R., Pollesel, P., “Pyrolysis of automotive shredder residue (ASR) - Influence of temperature on the
distribution of products”, J. Anal. Appl. Pyrolysis, Vol. 40–41, 383–401(1997).

Saha, B., Reddy, P.K., Ghoshal, A.K., “Hybrid genetic algorithm to find the best model and the globally optimized
overall kinetics parameters for thermal decomposition of plastics”, Chem. Eng. J., article in press, (2007).

Sodero, S.F., Berruti, F., Behie, L.A., “Ultrapyrolytic Cracking of Polyethylene - A high yield recycling method”,
Chem. Eng. Sc., Vol. 51, No 1, 2805-2810 (1996).

Syntech Biofuel Process - Ethanol from Biomass. URL: http://www.syntecbiofuel.com/

Waste online, Plastic recycling in UK (2007). URL: http://www.wasteonline.org.uk/index.aspx

Yang, J., Roy, C., “A new method for DTA measurement of enthalpy change during the pyrolysis of rubbers”,
Thermochimica Acta 288, 155 168 (1996).

Zhang, Y., Sun, Y., Zhong, B., “Synthesis of Higher Alcohols from Syngas over Ultrafine Mo—Co—K Catalysts”,
Catalysis Letters, Vol. 76, No. 3-4/October (2001).

Zhou, L., Wang, Y., Huang, Q., Cai, J., “Thermogravimetric characteristics and kinetic of plastic and biomass
blends co-pyrolysis”, Fuel Processing Technology, Vol. 87, 963–969 (2006).

http://www.bepress.com/ijcre/vol5/A85

Вам также может понравиться