Feedstock recycling of polymer wastes

7
Feedstock recycling of polymer wastes Arthur A. Garforth a, * , Salmiaton Ali b , Jesu ´s Herna ´ndez-Martı ´nez a , Aaron Akah a a Environmental Technology Centre, Department of Chemical Engineering, The University of Manchester, P.O. Box 88, Manchester M60 1QD, UK b Department of Chemical and Environmental Engineering, Faculty of Engineering, UNIPUTRA, 43400 UPM Serdang, Selangor, Malaysia Abstract Current common polymer waste recycling methods, mechanical recycling and energy recovery, have drawbacks such as labour intensive sorting and atmospheric pollution. Feedstock recycling has emerged as an environmentally successful alternative for poly- mer waste management. Ó 2005 Published by Elsevier Ltd. Keywords: Plastics; Polymer recycling; Feedstock recycling; Tertiary recycling; Catalytic cracking 1. Polymer recycling Polymer waste might be regarded as a potentially cheap source of chemicals and energy, although its recy- cling varies widely across Europe [*1]. Disposing of the waste to landfill is becoming undesirable due to legisla- tive pressures (where waste to landfill must be reduced by 35% over the period from 1995 to 2020) [2], rising costs, the generation of explosive greenhouse gases (such as methane) and the poor biodegradability of commonly used packaging polymers. The two main alternatives for treating municipal and industrial polymer wastes are energy recycling, where wastes are incinerated with some energy recovery and mechanical recycling. The incineration of polymer waste meets with strong societal opposition [3] and, there is the Kyoto Protocol to consider, as the UK moves towards its domestic goal of reducing carbon dioxide emissions by 20% by 2010 [2]. Mechanical recycling (the conver- sion of ‘‘scrap’’ polymer into new products) is a popular recovery path for manufacturers and is carried out on single-polymer waste streams as a market for recycled products can only be found if the quality is close to that of the original. Unfortunately these products are often more expensive than virgin plastic [4,5]. In 2002 in the UK, only 17% of 3.8 million tonnes of polymer waste was recycled by these methods, the remainder was land-filled or incinerated (without energy recovery) [*1]. 2. Feedstock recycling—current state of the art Feedstock recycling, also known as chemical recy- cling or tertiary recycling, aims to convert waste polymer into original monomers or other valuable chemicals. These products are useful as feedstock for a variety of downstream industrial processes or as trans- portation fuels. There are three main approaches: depo- lymerisation, partial oxidation and cracking (thermal, catalytic and hydrocracking) [*6]. 2.1. Depolymerisation Polymers are divided into two groups: (i) condensa- tion polymers and (ii) addition polymers. Condensation polymers which include materials such as polyamides, polyesters, nylons and polyethylene terephthalate (PET), can be depolymerised via reversible synthesis 1359-0286/$ - see front matter Ó 2005 Published by Elsevier Ltd. doi:10.1016/j.cossms.2005.04.003 * Corresponding author. E-mail address: [email protected] (A.A. Gar- forth). Current Opinion in Solid State and Materials Science 8 (2004) 419–425

Transcript of Feedstock recycling of polymer wastes

Current Opinion in Solid State and Materials Science 8 (2004) 419–425

Feedstock recycling of polymer wastes

Arthur A. Garforth a,*, Salmiaton Ali b, Jesus Hernandez-Martınez a, Aaron Akah a

a Environmental Technology Centre, Department of Chemical Engineering, The University of Manchester, P.O. Box 88, Manchester M60 1QD, UKb Department of Chemical and Environmental Engineering, Faculty of Engineering, UNIPUTRA, 43400 UPM Serdang, Selangor, Malaysia

Abstract

Current common polymer waste recycling methods, mechanical recycling and energy recovery, have drawbacks such as labour

intensive sorting and atmospheric pollution. Feedstock recycling has emerged as an environmentally successful alternative for poly-

mer waste management.

� 2005 Published by Elsevier Ltd.

Keywords: Plastics; Polymer recycling; Feedstock recycling; Tertiary recycling; Catalytic cracking

1. Polymer recycling

Polymer waste might be regarded as a potentially

cheap source of chemicals and energy, although its recy-

cling varies widely across Europe [*1]. Disposing of the

waste to landfill is becoming undesirable due to legisla-

tive pressures (where waste to landfill must be reduced

by 35% over the period from 1995 to 2020) [2], risingcosts, the generation of explosive greenhouse gases (such

as methane) and the poor biodegradability of commonly

used packaging polymers.

The two main alternatives for treating municipal and

industrial polymer wastes are energy recycling, where

wastes are incinerated with some energy recovery and

mechanical recycling. The incineration of polymer waste

meets with strong societal opposition [3] and, there is theKyoto Protocol to consider, as the UK moves towards

its domestic goal of reducing carbon dioxide emissions

by 20% by 2010 [2]. Mechanical recycling (the conver-

sion of ‘‘scrap’’ polymer into new products) is a popular

recovery path for manufacturers and is carried out on

single-polymer waste streams as a market for recycled

1359-0286/$ - see front matter � 2005 Published by Elsevier Ltd.

doi:10.1016/j.cossms.2005.04.003

* Corresponding author.

E-mail address: [email protected] (A.A. Gar-

forth).

products can only be found if the quality is close to that

of the original. Unfortunately these products are often

more expensive than virgin plastic [4,5]. In 2002 in the

UK, only 17% of 3.8 million tonnes of polymer waste

was recycled by these methods, the remainder was

land-filled or incinerated (without energy recovery) [*1].

2. Feedstock recycling—current state of the art

Feedstock recycling, also known as chemical recy-

cling or tertiary recycling, aims to convert waste

polymer into original monomers or other valuable

chemicals. These products are useful as feedstock for a

variety of downstream industrial processes or as trans-

portation fuels. There are three main approaches: depo-lymerisation, partial oxidation and cracking (thermal,

catalytic and hydrocracking) [*6].

2.1. Depolymerisation

Polymers are divided into two groups: (i) condensa-

tion polymers and (ii) addition polymers. Condensation

polymers which include materials such as polyamides,polyesters, nylons and polyethylene terephthalate

(PET), can be depolymerised via reversible synthesis

Fig. 1. Typical zeolites used in polymer cracking [57]: (a) H-ZSM-5,

(b) H-Mor (Mordenite), (c) H-Y or HUS-Y and (d) H-Beta.

420 A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425

reactions to initial diacids and diols or diamines. Typical

depolymerisation reactions such as alcoholysis, glycoly-

sis and hydrolysis yield high conversion to their raw

monomers [7].

In contrast, addition polymers which includematerials

such as polyolefins, typically making up 60–70% of muni-cipal solid waste plastics [*1,3,4], cannot be easily depoly-

merised into the original monomers. However, the results

obtained in the thermal depolymerisation of polymethyl-

methacrylate (PMMA) are noteworthy since at 723 K, a

98% yield to the monomer has been reported [*8].

2.2. Partial oxidation

The direct combustion of polymer waste, which has a

good calorific value, may be detrimental to the environ-

ment because of the production of noxious substances

such as light hydrocarbons, NOx, sulfur oxides and

dioxins. Partial oxidation (using oxygen and/or steam),

however, could generate a mixture of hydrocarbons

and synthesis gas (CO and H2), the quantity and quality

being dependent on the type of polymer used. Borgianniet al. [9] showed the possibilities of recovering energy

from waste containing polyvinyl chloride (PVC) by a

gasification process without additional dechlorination

facilities. A new type of waste gasification and smelting

system using iron-making and steel-making technologies

has been described by Yamamoto et al. [10], reportedly

to produce a dioxin-free and high-calorie purified gas.

Hydrogen production efficiency of 60–70% from poly-mer waste has been reported for a two-stage pyrolysis

and partial oxidation process [*11]. Co-gasification of

biomass with polymer waste has also been shown to in-

crease the amount of hydrogen produced while the CO

content reduced [12]. The production of bulk chemicals,

such as acetic acid, from polyolefins via oxidation using

NO and/or O2, is also possible [13,14].

2.3. Cracking: hydro-, thermal- and catalytic

Cracking processes break down polymer chains into

useful lower molecular weight compounds. This can be

achieved by reaction with hydrogen, known as hydro-

cracking or by reaction in an inert atmosphere (pyrolytic

methods), which can be either thermal or catalytic

cracking.Hydrocracking of polymer waste typically involves

reaction with hydrogen over a catalyst in a stirred batch

autoclave at moderate temperatures and pressures (typ-

ically 423–673 K and 3–10 MPa hydrogen). The work

reported, mainly focuses on obtaining a high quality

gasoline starting from a wide range of feeds. Typical

feeds include polyolefins, PET, polystyrene (PS), polyvi-

nyl chloride (PVC) and mixed polymers [15,*16,17–21],polymer waste from municipal solid waste and other

sources [17,18,22–26], co-mixing of polymers with coal

[24,25,27–31], co-mixing of polymers with different refin-

ery oils such as vacuum gas–oil [32–36] and scrap tyres

alone or co-processed with coal [37–41]. To aid mixing

and reaction, solvents such as 1-methyl naphthalene,

tetralin and decalin have been used with some success

[25,28,41]. Several catalysts, classically used in refineryhydrocracking reactions, have been evaluated and in-

clude transition metals (e.g., Pt, Ni, Mo, Fe) supported

on acid solids (such as alumina, amorphous silica–alu-

mina, zeolites and sulphated zirconia). These catalysts

incorporate both cracking and hydrogenation activities

and although gasoline product range streams have been

obtained, little information on metal and catalyst sur-

face areas, Si/Al ratio or sensitivity to deactivation isquoted.

In thermal degradation, the process produces a broad

product range and requires high operating temperatures,

typically more than 773 K and even up to 1173 K

[*8,42–49]. On the other hand, catalytic degradation

might provide a solution to these problems by allowing

control of the product distribution and reducing the

reaction temperature [48,50–53].Catalytic cracking studies have been mainly limited

to pure polymers (predominantly using polyolefins and

PS) and fresh, pure acid catalysts (zeolites predominat-

ing). Zeolites are crystalline, porous aluminosilicates

[54–56] characterised by channel networks and pore

openings of molecular dimensions (see Fig. 1, [57])

leading to increased shape selectivity in petrochemical

Fig. 2. Mixture of HDPE/ZSM-5 at 478 K, 200 times magnification.

Reprinted from Garforth and co-workers [*85].

A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425 421

reactions [54–56,58]. PVC is problematic because HCl

strips from the polymer at relatively low temperatures

[59,60]. In Japan, a low temperature thermal cracking

stage is employed prior to catalytic cracking, however,

co-mingling with other polymer waste is required to

boost the H content of the residual partially crackedpolymer waste. Typically iron-based catalysts have been

employed to dechlorinate the PVC/mixed polymer-de-

rived oil [52,61–63]. The catalytic cracking of PS to ben-

zene, toluene and xylene (BTX), as well as styrene

monomer, has been carried out by a number of research-

ers at operating temperatures from 623–823 K over acid

catalysts such as zeolites (HMOR HZSM-5, HY), amor-

phous SiO2–Al2O3, BaO powder and a sulfur-promotedzirconia [64–67]. Predominantly, catalytic cracking re-

search has focussed on the degradation of polyolefins

to gas, liquid and waxy products using a range of acid

catalysts (typically, amorphous silica alumina and zeo-

lites). For example, in Japan, legislative pressures have

resulted in research targeting a stable liquid product.

Other researchers have targeted an end use, such as, gas-

oline-range hydrocarbons and others the production ofethene and propene [15,48,68–74].

A variety of reactor types has been used including

batch [53,70,75] and fixed bed [52,68,69,71,76–79],

or non-catalytically using thermal degradation in a

fluidised bed reactor or kiln [59,80,*81]. With batch

reactors, secondary cracking reactions predominate,

yielding a broad range of products including heavy aro-

matics, coke and saturated hydrocarbons. Fixed bedreactors are prone to blocking due to the viscous nature

of melted polymer presenting problems when scaling-up

[82]. Non-catalytic thermal cracking using a fluidised

bed reactor with sand as a fluidising agent or kiln

requires a higher operating temperature and produces

products in a very broad range [*81].

On the other hand, the use of a fluidised bed reactor

has advantages in terms of heat and mass transfer, aswell as constant temperatures throughout the reactor

[48,82,83]. Recent work has logically extended studies

to fluid catalytic cracking (FCC) catalysts with compar-

isons to pure zeolites and silica alumina [74,84,*85].

Before design predictions can be made for a pyrolysis

process on an industrial scale, an understanding of the

interface between the polymer and the catalyst must be

developed. The mechanism of interaction is highly com-plex, with three phases (liquid polymer, solid catalyst

and gaseous products), mass transfer by diffusion, con-

vection and bulk flow as well as cracking-type reactions

with a large number of products. Fig. 2 shows a scan-

ning electron micrograph (SEM) of a finely blended

mixture of high density polyethylene (HDPE) and

ZSM-5 after heating from room temperature to 478 K.

The polymer particles have melted and flowed but indi-vidual catalyst and polymer particles were still notice-

able. On increasing the temperature to 573 K, the

melted polymer has completely ‘‘wetted’’ the zeolite par-

ticles (Fig. 3).The catalytic degradation of HDPE has been carried

out in a laboratory fluidised bed reactor using pure zeo-

lites and fresh, steam deactivated and ‘‘equilibrium’’ cat-

alysts (E-Cats) with different rare earth oxides and Ni

and V loadings (listed in Table 1) [48,83,86–88].

At 723 K, the products from polymer cracking were

mostly gases in the range C1–C9 (determined by GC

analysis), and coke and unreacted polymer (determinedby thermogravimetric analysis) [48,89]. As expected,

trends in polymer cracking (Table 3) reflected the differ-

ent nature of catalysts, with fresh commercial FCC cat-

alysts and pure ZSM-5 catalyst converting 85–90% of

their feeds to gaseous, liquid and carbonaceous prod-

ucts. The lower activity of pure US-Y (ex Crosfield

Chemical) was expected due to its rapid deactivation.

On the other hand, the less active steamed and equilib-rium catalysts showed only 60–70% conversion to the

volatile products. The E-Cats showed negligible loss in

overall conversion of HDPE due to metal contamina-

tion, although the products of polymer degradation

were olefin-rich compared with steam deactivated

Cat-1S and -7S.

Figs. 4 and 5 show selected olefin and paraffin prod-

ucts in the carbon range of C3–C6, respectively, for thecatalysts. US-Y is the major active component in com-

mercial FCC catalysts and therefore the product yields

compared favourably. The level of activity of the vari-

ous catalysts was reflected in the amount of primary

(olefin) versus secondary (paraffin) products observed.

With high acidities of both fresh catalysts (Cat-1 and -7,

Table 1

Catalyst details (supplied by Engelhard Corporation, USA) [88]

Catalyst Commercial name REO (wt%) UCS (A) MSA (m2/g) ZSA (m2/g) Ni (ppm) V (ppm)

Cat-1 Fresh commercial FCC catalyst 0.8 24.4 112 264 – –

Cat-7 Fresh commercial FCC catalyst 9.6 24.7 90 331 – –

Cat-1S a Steam deactivated FCC catalyst 0.8 24.3 90 198 – –

Cat-7S a Steam deactivated FCC catalyst 9.6 24.5 72 241 – –

E-Cat 1 Equilibrium FCC catalysts 1.3 24.3 76 99 171 217

E-Cat 2 Equilibrium FCC catalysts 1.6 24.3 32 95 5400 6580

a Steaming conditions: 4 h/1061 K/100% steam.

Fig. 3. Mixture of HDPE/ZSM-5 at 573 K, 200 times magnification. Reprinted from Garforth and co-workers [*85].

Table 2

Weight% of product distributions at T = 723 K; C/P = 6:1 [88]

ZSM-5 US-Y Cat-1 Cat-7 Cat-1S Cat-7S E-Cat1 E-Cat2

Gaseous 83.7 55.9 75.0 71.8 50.4 55.8 64.5 65.8

Liquid 2.0 0.5 9.0 6.8 7.2 7.8 1.4 1.4

Coke 2.4 4.5 6.5 7.2 3.0 4.9 1.5 1.2

Involatile 11.9 39.1 9.5 14.2 39.4 31.5 32.6 31.6

Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0

Gaseous product distribution

C1–C4 68.6 36.6 44.4 47.4 38.4 44.4 35.2 37.1

C5–C9 23.1 60.2 52.2 48.8 60.2 52.8 63.4 62.6

BTX 8.3 3.2 3.4 3.8 1.4 2.8 1.4 0.3

Total 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0

Total gaseous product

Paraffins 27.0 48.8 53.7 60.0 31.4 48.7 23.6 23.0

Olefins 64.7 47.8 42.5 35.7 67.1 48.6 74.9 76.6

Yield (wt%) = (P(g)/Polymer feed (g)) · 100.

422 A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425

Table 2), high reactivity was expected and a high yield of

secondary products, paraffins, was observed.

By contrast, the used catalysts with lower acidities and

poisoned with heavy metals yielded predominantly ole-

finic products mostly in the carbon range of C3–C6. Evi-

dence of high REO stabilisation of steam deactivated

catalyst, Cat-7S, was noted with a yield of balanced pri-

mary and secondary products (Figs. 4 and 5). During the

0

5

10

15

20

25

ZSM5 USY Cat1 Cat7 Cat1S Cat7S ECat1 ECat2

C3= C4=

C5= C6=

Fig. 4. Selected olefin products (wt%) at T = 450 �C; C/P = 6:1.

0

5

10

15

20

25

ZSM5 USY Cat1 Cat7 Cat1S Cat7S ECat1 ECat2

C3 C4

C5 C6

Fig. 5. Selected paraffin products (wt%) at T = 723 K; C/P = 6:1.

Table 3

Total plastic waste generated and recovered in Western Europe (kt) [*1]

1993 1995 1997 1999 2001 2003

Total plastics waste 16,211 16,056 16,975 19,166 19,980 21,150

Total plastics waste recovered 3340 4019 4364 6183 7357 8230

Mechanical recycling 915 1222 1455 1888 2521 3130

Feedstock recycling 0 99 334 346 298 350

Energy recovery 2425 2698 2575 3949 4538 4750

%Total plastics waste recovered 21 26 26 32 37 39

A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425 423

steaming process in the FCC regenerator, catalysts will

lose some of framework aluminium ions, creating defects

in the crystals and leading to decreased catalyst acidities

[87]. Nevertheless, with the presence of RE in the FCC

catalysts, the steam dealumination is hindered. There-

fore, with RE, the catalyst activities are maintained by

reducing the amount of crystal destruction as seen here.

3. Conclusions

Although feedstock recycling has been heralded as

having great potential, polymer waste recycling levels

have remained virtually unchanged at 350 kt since

1997 (see Table 3, [*1]). High costs associated with col-

lection, sorting and transportation to provide a guaran-teed supply of low chlorine-containing polymer waste to

recycling sites remain significant. Schemes such as

Duales System Deutschland [*90] in Germany (‘‘green

dot’’) have addressed this issue but there remains the

high energy and process costs of the feedstock recycling

technology. Thermal and catalytic cracking although

effective require significant operating temperatures and

are strongly endothermic, leading to large adiabatic tem-

perature falls across reactors. However, improving the

economics of the process itself by using exhaustedzero-cost catalysts to produce a tailored product will

help to make the process viable [83].

Oxidation methods, energetically more favourable,

are at high temperature and have associated difficul-

ties such as dangerous emissions, product quality and

expensive materials of construction. Hydrocracking

studies have been limited to date and merit further study

since the process is exothermic and can be carried out atsignificantly lower temperatures.

Another strategy worth considering is the targeting of

large volume guaranteed waste streams, such as, from

paper recycling plants to reduce collections costs. If this

is linked with careful characterisation of this type of

waste stream, the supply of a quality controlled polymer

waste should be possible.

Acknowledgements

This work was performed with the financial support

of the University Putra Malaysia. Thank you to Dr.

D.H. Harris (Engelhard Corporation) for catalysts and

technical advice. Thanks also to Miss S. Maegaard for

her contribution on thermal analysis and microscopyand also Mr. R.J. Plaisted of the Centre of Micropo-

rous Materials. Special thanks to Dr. D.L. Cresswell

for useful discussion during the preparation of this

review.

References

The papers of particular interest have been high-

lighted as:

* of special interest.

[*1] Plastics Europe (Association of Plastic Manufacturers). An

analysis of plastics consumption and recovery in Europe (2002-

2003). Published in 2004.

424 A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425

[2] DEFRA, Department for Environment, Food and Rural Affairs.

Waste Strategy 2000 for England and Wales: part II chapter 5,

<www.defra.gov.uk/environment/waste/strategy/cm4693/index. htm>.

[3] British Plastic Federation. Plastic waste management, <http://

www.bpf.co.uk/bpfissues/Waste/Management.cfm>.

[4] Brandrup J. In: Menges G, editor. Recycling and recovery of

plastics. Munich: Hanser; 1996. p. 393–412.

[5] Lee M. Recycling polymer waste. Chemistry in Britain 1995;7:

515–6.

[*6] Aguado J, Serrano DP. Feedstock recycling of plastic wastes,

RSC clean technology monographs. Cambridge: Royal Society

of Chemistry; 1999. p. 192.

[7] Cornell DD. In: Richard F, editor. Plastics, rubber, and paper

recycling—a pragmatic approach. Washington: American

Chemical Society; 1995. p. 72–9.

[*8] Kaminsky W, Predel M, Sadiki A. Feedstock recycling of

polymers by pyrolysis in a fluidised bed. Polym Degrad Stabil

2004;85(3):1045–50.

[9] Borgianni C, De Filippis P, Pochetti F, Paolucci M. Gasification

process of wastes containing PVC. Fuel 2002;81(14):1827–33.

[10] Yamamoto T, Isaka K, Sato H, Matsukura Y, Ishida H.

Gasification and smelting system using oxygen blowing for

municipal waste. ISIJ Int 2000;40(3):260–5.

[*11] Wallman PH, Thorsness CB, Winter JD. Hydrogen production

from wastes. Energy 1998;23(4):271–8.

[12] Pinto F, Franco C, Andre RN, Miranda M, Gulyurtlu I, Cabrita

I. Co-gasification study of biomass mixed with plastic wastes.

Fuel 2002;81(3):291–7.

[13] Pifer A, Sen A. Chemical recycling of plastics to useful organic

compounds by oxidative degradation. Angew Chem Int Ed

1998;37(23):3306–8.

[14] Shiono T, Niki E, Kamiya Y. Oxidative degradation of

polymers. II. Thermal oxidative degradation of atactic polypro-

pylene in solution under pressure. J Appl Polym Sci 2003;

21(6):1625–46.

[15] Walendziewski J, Steininger M. Thermal and catalytic conver-

sion of waste polyolefines. Catal Today 2001;65:323–30.

[*16] Venkatesh KR, Hu J, Wang W, Holder GD, Tierney JW,

Wender I. Hydrocracking and hydroisomerisation of long-chain

alkanes and polyolefins over metal-promoted anion-modified

zirconium oxides. Energy Fuels 1996;10:1163–70.

[17] Ding WB, Liang J, Anderson LL. Thermal and catalytic

degradation of high density polyethylene and commingled post-

consumer plastic waste. Fuel Process Technol 1997;51:47–62.

[18] Ding WB, Liang J, Anderson LL. Hydrocracking and hydroiso-

merisation of high density polyethylene and waste plastic over

zeolite and silica–alumina-supported Ni and Ni–Mo sulfides.

Energy Fuels 1997;11:1219–24.

[19] Hesse ND, White RL. Polyethylene catalytic hydrocracking by

PtHZSM-5, PtHY, and PtHMCM-41. J Appl Poly Sci 2004;92:

1293–301.

[20] Dufaud V, Basset J-M. Catalytic hydrogenolysis at low temper-

ature and pressure of polyethylene and polypropylene to diesels

or lower alkanes by a zirconium hydride supported on silica–

alumina: A step toward polyolefin degradation by the micro-

scopic reverse of Ziegler–Natta polymerisation. Angew Chem Int

Ed 1998;37(6):806–10.

[21] Murty MVS, Rangarajan P, Grulke EA, Bhattacharyya D.

Thermal degradation/hydrogenation of commodity plastics and

characterisation of their liquefaction products. Fuel Process

Technol 1996;49:75–90.

[22] de la Puente G, Klocker C, Sedran U. Conversion of waste

plastics into fuels: Recycling polyethylene in FCC. Appl Catal B:

Environ 2002;36(4):279–85.

[23] Walendziewski J. Engine fuel derived from waste plastics by

thermal treatment. Fuel 2002;81(4):473–81.

[24] Feng Z, Zhao J, Rockwell J, Bailey D, Huffman G. Direct

liquefaction of waste plastics and co-liquefaction of coal–plastic

mixtures. Fuel Process Technol 1996;49:17–30.

[25] Luo M, Curtis CW. Effect of reaction parameters and catalyst

type on waste plastics liquefaction and coprocessing with coal.

Fuel Process Technol 1996;49:177–96.

[26] Ramdoss PK, Tarrer AR. High-temperature liquefaction of

waste plastics. Fuel 1998;77(4):293–9.

[27] Ding WB, Tuntawiroon W, Liang J, Anderson LL. Depolymer-

isation of waste plastics with coal over metal-loaded silica–

alumina catalysts. Fuel Process Technol 1996;49:49–63.

[28] Rothenberger KS, Cugini AV, Thompson RL, Ciocco MV.

Investigation of first-stage liquefaction of coal with model plastic

waste mixtures. Energy Fuels 1997;11:849–55.

[29] Taghiei MM, Feng Z, Huggins FE, Huffman G. Coliquefaction

of waste plastics with coal. Energy Fuels 1994;8:1228–32.

[30] Wang L, Chen P. Development of first-stage coliquefaction of

Chinese coal with waste plastics. Chem Eng Process 2004;43:

145–8.

[31] Yasuda H, Yamada O, Zhang A, Nakano K, Kaiho M.

Hydrogasification of coal and polyethylene mixture. Fuel 2004;

83:2251–4.

[32] Karagoz S, Yanik J, Ucar S, Song C. Catalytic coprocessing of

low density polyethylene with VGO using metal supported on

activated carbon. Energy Fuels 2002;16:1301–8.

[33] Karagoz S, Yanik J, Ucar S, Saglam M, Song C. Catalytic and

thermal degradation of high density polyethylene in vacuum gas

oil over non-acidic and acidic catalysts. Appl Catal A: Gen 2003;

242:51–62.

[34] Karagoz S, Karayildirim T, Ucar S, Yuksel M, Yanik J.

Liquefaction of municipal waste plastics in VGO over acidic

and non-acidic catalysts. Fuel 2003;82(4):415–23.

[35] Karayilidirim T, Yanik J, Ucar S, Saglam M, Yuksel M.

Conversion of plasticsrHVGO mixtures to fuels by two-step

processing. Fuel Process Technol 2001;73:23–35.

[36] Joo HS, Curtis CW. Catalytic coprocessing of plastics with coal

and petroleum resid using NiMo/Al2O3. Energy Fuels 1996;10:

603–11.

[37] Harrison G, Ross AB. Use of tyre pyrolysis oil for solvent

augmentation in two-stage coal liquefaction. Fuel 1996;75(8):

1009–13.

[38] Anderson LL, Callen M, Ding W, Liang J, Mastral AM,

Mayoral MC, et al. Hydrocoprocessing of scrap automotive

tires and coal. Analysis of oils from autoclave coprocessing. Ind

Eng Chem Res 1997;36:4763–7.

[39] Ibrahim MM, Seehra MS. Effects of hydrogen pressure and

temperature on free radicals produced in coal–tire coprocessing.

Fuel Process Technol 1996;49:197–205.

[40] Liu Z, Zondlo JW, Dadyburjor DB. Tire liquefaction and its

effect on coal liquefaction. Energy Fuels 1994;8:607–12.

[41] Tang Y, Curtis CW. Thermal and catalytic coprocessing of waste

tires with coal. Fuel Process Technol 1996;46:195–215.

[42] Demirbas A. Pyrolysis of municipal plastic wastes for recovery of

gasoline-range hydrocarbons. J Anal Appl Pyrol 2004;72(1):

97–102.

[43] Cunliffe AM, Jones N, Williams PT. Pyrolysis of Composite

Plastic Waste. Environ Technol 2003;24(5):653–63.

[44] Babu BV, Chaurasia AS. Modeling for pyrolysis of solid particle:

Kinetics and heat transfer effects. Energy Convers Manage 2003;

44(14):2251–75.

[45] Faravelli T, Pinciroli M, Pisano F, Bozzano G, Dente M, Ranzi

E. Thermal degradation of polystyrene. J Anal Appl Pyrol 2001;

60(1):103–21.

[46] Fink JK. Pyrolysis and combustion of polymer wastes in

combination with metallurgical processes and the cement indus-

try. J Anal Appl Pyrol 1999;51(1–2):239–52.

A.A. Garforth et al. / Current Opinion in Solid State and Materials Science 8 (2004) 419–425 425

[47] Conesa JA, Font R, Marcilla A, Garcia AN. Pyrolysis of

polyethylene in a fluidised bed reactor. Energy Fuels 1994;8(6):

1238–46.

[48] Garforth AA, Lin Y-H, Sharratt PN, Dwyer J. Production of

hydrocarbons by catalytic degradation of high density polyeth-

ylene in a laboratory fluidised bed reactor. Appl Catal A: Gen

1998; 169(2):331–42.

[49] Buekens AG, Schoeters JG. Technical methods in plastics

pyrolysis. Macromol Symp 1998;135:63–81.

[50] Sakata Y. Catalytic degradation of polyethylene and polypro-

pylene to fuel oil. Macromol Symp 1998;135:7–18.

[51] Sakata Y, Uddin MA, Koizimi K, Murata K. Catalytic

degradation of polypropylene into liquid hydrocarbons using

silica–alumina catalyst. Chem Lett 1996;3:254–6.

[52] Uemichi Y, Takuma K, Ayame A. Chemical recycling of poly

(ethylene) by catalytic degradation into aromatic hydrocarbons

using H–Ga-silicate. Chem Commun 1998;18:1975–6.

[53] Yoshida T, Ayame A, Kano H. Gasification of polyethylene over

solid catalysts (part 1). Study by the use of a micro batchwise

reactor. Bull Jpn Petrol Inst 1975;17(2):218–25.

[54] Barrer RM. Zeolites and clay minerals as sorbents and molecular

sieves, 496. London: Academic Press; 1978.

[55] Martens JA, Jacobs PA. In: Jansen JC, editor. Introduction to

zeolite science and practise. Studies in surface science and

catalysis, vol. 58. Amsterdam: Elsevier; 2001. p. 633–72.

[56] Ward JW. In: Leach BE, editor. Applied industrial catalysis. 3rd

ed. New York: Academic Press; 1984. p. 271–8.

[57] International Zeolite Association. Atlas of zeolite structure

types, <http://www.iza/structure.org/databases>.

[58] Breck DW. Zeolite molecular sieves: Structure, chemistry and

use. New York: Wiley-Interscience; 1974. p. 752.

[59] Brophy BJH. In: Menges G, editor. Recycling and recovery of

plastics. Munich: Hanser; 1996. p. 422–33.

[60] Burnett RH. In: Richard F, editor. Rubber, and paper recycling:

A pragmatic approach. Washington: American Chemical Soci-

ety; 1995. p. 97–103.

[61] Paci M, La Mantia FP. Influence of small amounts of polyvi-

nylchloride on the recycling of polyethyleneterephthalate. Polym

Degrad Stab 1999;63(1):11–4.

[62] Tang C, Wang Y-Z, Zhou Q, Zheng L. Catalytic effect of Al–Zn

composite catalyst on the degradation of PVC-containing

polymer mixtures into pyrolysis oil. Polym Degrad Stab

2003;81(1):89–94.

[63] Lingaiah N, Uddin AM, Muto A, Sakata Y, Imai T, Murata K.

Catalytic dechlorination of chloroorganic compounds from

PVC-containing mixed plastic-derived oil. Appl Catal A: Gen

2001; 207(1–2):79–84.

[64] Nanbu H, Sakuma Y, Ishihara Y, Takesue T, Ikemura T.

Catalytic degradation of polystyrene in the presence of aluminum

chloride catalyst. Polym Degrad Stab 1987;19(1):61–76.

[65] Serrano DP, Aguado J, Escola JM. Catalytic conversion of

polystyrene over HMCM-41, HZSM-5 and amorphous SiO2–

Al2O3: Comparison with thermal cracking. Appl Catal B:

Environ 2000;25(2–3):181–9.

[66] Ukei H, Hirose T, Horikawa S, Takai Y, Taka M, Azuma N,

et al. Catalytic degradation of polystyrene into styrene and a

design of recyclable polystyrene with dispersed catalysts. Catal

Today 2000;62(1):67–75.

[67] de la Puente G, Sedran U. Recycling polystyrene into fuels by

means of FCC: Performance of various acidic catalysts. Appl

Catal B: Environ 1998;19(3–4):305–11.

[68] Uemichi Y, Ayame A, Kashiwaya Y, Kano H. Gas chromato-

graphic determination of the products of degradation of poly-

ethylene over a silica–alumina catalyst. J Chromatogr 1983;

259(1): 69–77.

[69] Uemichi Y, Nakamura J, Itoh T, Sugioka M, Garforth AA,

Dwyer J. Conversion of polyethylene into gasoline-range fuels by

two-stage catalytic degradation using silica–alumina and HZSM-

5 zeolite. Ind Eng Chem Res 1999;38(2):385–90.

[70] Aguado J, Serrano DP, Escola JM, Garagorri E, Fernandez JA.

Catalytic conversion of polyolefins into fuels over zeolite beta.

Polym Degrad Stab 2000;69(1):11–6.

[71] Luo G, Suto T, Yasu S, Kato K. Catalytic degradation of high

density polyethylene and polypropylene into liquid fuel in a

powder-particle fluidised bed. Polym Degrad Stab 2000;70(1):

97–102.

[72] Negelein DL, Lin R, White RL. Effects of catalyst acidity and

HZSM-5 channel volume on polypropylene cracking. J Appl

Poly Sci 1998;67(2):341–8.

[73] Hesse ND, Lin R, Bonnet E, Cooper III J, White RL. In situ

analysis of volatiles obtained from the catalytic cracking of

polyethylene. J Appl Poly Sci 2001;82(12):3118–25.

[74] Cardona SC, Corma A. Tertiary recycling of polypropylene by

catalytic cracking in a semibatch stirred reactor: Use of spent

equilibrium FCC commercial catalyst. Appl Catal B: Environ

2000;25(2–3):151–62.

[75] Audisio G, Bertini F, Beltrame PL, Carniti P. Catalytic

degradation of polyolefins. Macromol Symp 1992;57:191–209.

[76] Uemichi Y, Kashiwaya Y, Tsukidate M, Ayame A, Kanoh H.

Product distribution in degradation of polypropylene over silica–

alumina and CaX zeolite catalysts. Bull Jpn Petrol Inst

1983;56(9): 2768–73.

[77] Songip AR, Masuda T, Kuwahara H, Hashimoto K. Test to

screen catalysts for reforming heavy oil from waste plastics. Appl

Catal B: Environ 1993;2(2–3):153–64.

[78] Ohkita H, Nishiyama R, Tochihara Y, Mizushima T, Kakuta N,

Morioka Y, et al. Acid properties of silica–alumina catalysts and

catalytic degradation of polyethylene. Ind Eng Chem Res 1993;

32(12):3112–6.

[79] Mordi RC, Fields R, Dwyer J. Thermolysis of low density

polyethylene catalysed by zeolites. J Anal Appl Pyrol 1994;29(1):

45–55.

[80] Dent I, Hardman S. IChemE Environ Protect 1996;44:1–8.

[*81] Hardman S, Leng SA, Wilson DC. BP Chemicals Limited.

Polymer cracking. European Patent Application, EP 567292,

1993.

[82] Lin Y-H. Theoretical and experimental studies of catalytic

degradation of polymers. PhD Thesis. University of Manchester

Institute of Science and Technology. Manchester 1998.

[83] Ali S, Garforth AA, Harris DH, Rawlence DJ, Uemichi Y.

Polymer waste recycling over ‘‘used’’ catalysts. Catal Today

2002; 75(1–4):247–55.

[84] Gobin K, Manos G. Thermogravimetric study of polymer

catalytic degradation over microporous materials. Polym Degrad

Stab 2004;86(2):225–31.

[*85] Ali S, Garforth AA, Harris DH, Shigeishi RA. Polymer waste

recycling over ‘‘used’’ catalyst. In: Keane MA, editor. Interfacial

applications in environmental engineering. Surfactant science

series, vol. 108. NewYork: MarcelDekker, Inc.; 2002. p. 295–328.

[86] Dwyer J, Rawlence DJ. Fluid catalytic cracking: Chemistry.

Catal Today 1993;18(4):487–507.

[87] Mitchell MM, Hoffman JF, Moore HF. In: Mitchell MM, editor.

Fluid catalytic cracking: Science and technology. Studies in

surface science and catalysis, vol. 76. Amsterdam: Elsevier;

1993. p. 293–338.

[88] Ali S, Garforth AA. Feedstock recycling of polyolefins by

catalytic cracking. in 2nd International symposium on feedstock

recycling of plastics and other innovative plastics recycling

techniques, Ostend, Belgium, 2002.

[89] Lin Y-H, Sharratt PN, Garforth AA, Dwyer J. Deactivation of

US-Y zeolite by coke formation during the catalytic pyrolysis of

high density polyethylene. Termochim Acta 1997;294(1):45–50.

[*90] Patel M, von Thienen N, Eberhard J, Worrell E. Recycling of

plastics in Germany. Resour Conserv Recycl 2000;29:65–90.