You are on page 1of 9

Polymer Degradation and Stability 92 (2007) 813e821

www.elsevier.com/locate/polydegstab

Chemical catalysed recycling of waste polymers: Catalytic


conversion of polypropylene into fuels and chemicals over
spent FCC catalyst in a fluidised-bed reactor
Y.-H. Lin*, M.-H. Yang
Department of Biochemical Engineering & Graduate Institute of Environmental Polymer Materials, Kao Yuan University, 821 Kaohsiung, Taiwan, ROC
Received 27 October 2006; received in revised form 26 January 2007; accepted 31 January 2007
Available online 11 February 2007

Abstract
Polypropylene (PP) was pyrolysed over spent FCC commercial catalyst (FCC-s1) using a laboratory fluidised-bed reactor operating isothermally at ambient pressure. The influence of reaction conditions including catalyst, temperature, and ratio of polymer to catalyst feed and flow
rates of fluidising gas was examined. The yield of gaseous and liquid hydrocarbon products at 390  C for spent FCC commercial catalyst
(87.8 wt%) gave much higher yield than silicate (only 17.1 wt%). Greater product selectivity was observed with FCC-s1 as a post-use catalyst
with about 61 wt% olefins products in the C3eC7 range. The selectivity could be further influenced by changes in reaction conditions. Valuable
hydrocarbons of olefins and iso-olefins were produced by low temperatures and short contact times used in this study. It is also demonstrated that
a post-use catalyst system under appropriate conditions the resource potential of polymer waste can be economically recovered and also can
address the recycling desire to see an alternative to solve a major environment problem.
2007 Elsevier Ltd. All rights reserved.
Keywords: Polymer; Fluidised-bed reactor; Pyrolysis; Catalyst; Selectivity

1. Introduction
Polymer waste can be regarded as a potential source of
chemicals and energy. Methods for recycling polymer waste
have been developed and new recycling approaches are being
investigated [1]. Chemical recycling, i.e., thermal and/or catalytic conversion of waste polymers into fuels or chemicals, has
been recognised as an ideal approach and could significantly
reduce the net cost of disposal [2]. The most widely used conventional chemical methods for waste polymer treatment are
pyrolysis and catalytic reforming. Since thermal degradation
demands relatively high temperatures and its products require
further processing for their quality to be upgraded, catalytic
degradation of polymer waste offers considerable advantages.

* Corresponding author. Tel.: 886 7 6077777; fax: 886 7 6077788.


E-mail address: lin@cc.kyu.edu.tw (Y.-H. Lin).
0141-3910/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2007.01.028

Suitable catalysts have the ability to control both the product


yield and product distribution from polymer degradation as
well as to reduce significantly the reaction temperature, potentially leading to a cheaper process with more valuable products
[3e5].
Studies of the effects of catalysts on the catalytic degradation of polymer have been performed by (i) contacting melted
polymer with catalyst in fixed bed reactors [6e8], (ii) heating
mixtures of polymer and catalyst powders in reaction vessels
[9e11] and (iii) passing the products of polymer pyrolysis
through fixed bed reactors containing cracking catalysts
[12e14]. However, the configuration of the pyrolysis-reforming reactors poses serious engineering and economics constraints. For the study of these pyrolysis-reforming reactions,
it is difficult to measure the exact mass-flow rate of the
reactant from the pyrolysis to catalytic zones, and consequently it is virtually impossible to identify and quantify the
reactant and to control its quality. In addition, the use of fixed

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

814

beds or adiabatic batch where polymer and catalyst are contacted directly leads to problems of blockage and difficulty
in obtaining intimate contact over the whole reactor. Without
good contact the formation of large amounts of residue are
likely, and scale-up to industrial scale is not feasible. In order
to compare the polymer cracking properties of different catalysts, it is preferable to examine the effects of catalysts without
extensive complications due to reactions of primary cracking
products, e.g. olefins, with unconverted polymer by using techniques that minimise such interactions. For this purpose, a laboratory fluidized-bed reactor has been used to study catalytic
cracking of polymers by limiting the contact between primary
volatile products and the catalyst/polymer mixture [15e18].
The catalyst increases significantly the commercial potential of
a recycling process based on catalytic degradation, as cracking catalysts could cope with the conversion of plastic waste co-fed into
a refinery FCC unit. It is certainly possible to develop commercial
processes based on these approaches. Therefore, a more interesting approach is that of adding polymer waste into the FCC process,
under suitable process conditions with the use of zero value of
spent FCC catalysts, a large number of waste plastics can be economically converted into valuable hydrocarbons. However, much
less is known about performance of fluid catalytic cracking (FCC)
commercial catalysts on the degradation of polymer waste. Although spent FCC catalysts have been investigated in some trials,
the results for those studies are sketchy and were carried out by the
use of batch or fixed bed reactors [19,20]. Potential concepts has
been investigated (Part I) using thermal analysis to estimate the kinetic behaviours and as a potential method to evaluate the effect of
spent FCC catalysts on the catalytic degradation of polypropylene
(PP). It is the objective of this work to investigate the product slate
of the FCC process made by PP recycled into the usual chemicals
and fuels, by means of its conversion over post-use FCC commercial catalysts in a catalytic fluidised-bed reaction system, and specifically for identification of suitable reaction conditions for
enhancing the potential benefits of catalytic polymer recycling.
2. Experimental
2.1. Materials and experimental procedures
The catalysts employed are described in Table 1. All the catalysts were pelleted, crushed and sieved to give particle sizes

ranging from 75 to 180 mm. The catalyst was then dried by


heating in flowing nitrogen (50 ml min1) to 120  C at
60  C h1. After 2 h the temperature was increased to 520  C
at a rate of 120  C h1 to active the catalyst for 5 h. In contrast
with the micrometer size of the crystals in conventional zeolites
[21], ZSM-5 and HUSY used in this work synthesised with
smaller crystallite sizes [22,23] present a high proportion of external surface, which accounts for approximately 20e30% of
the total zeolite surface area. The polymer used in this study
was pure polypropylene (PP; isotactic, r 853.6 kg m3,
MW z 332,000, Aldrich). High purity nitrogen was used as
the fluidising gas and the flow was controlled by a needle valve
and pre-heated in the bottom section of the reactor tube. Flowmeters were used to measure the full range of gas velocities
from the incipient to fast fluidisation. Before catalytic pyrolysis
experiments were started, several fluidisation runs were performed at ambient temperature and pressure to select (i) suitable particle sizes (both catalyst and polymer waste) and (ii)
optimise the fluidising gas flow rates to be used in the reaction.
The particle size of both catalyst (75e180 mm) and polymer
(75e250 mm) were chosen to be large enough to avoid entrainment but not too large as to be inadequately fluidised. High
flow rates of fluidising stream improve catalystepolymer
mixing and external heat transfer between the hot bed and the
cold catalyst. On the other hand, an excessive flow rate could
cause imperfect fluidisation and considerable entrainment of
fines.
2.2. Experimental procedures and product analysis
A process flow diagram of the experimental system is given
elsewhere [15] and shown schematically in Fig. 1. A threezone heating furnace with digital controllers was used and
the temperatures of the furnace in its upper, middle and bottom
zones were measured using three thermocouples. By these
means the temperature of the pre-heated nitrogen below the
distributor and catalyst particles in the reaction volume could
be effectively controlled to within 1  C. The polymer feed
system was designed to avoid plugging the inlet tube with
melted polymer and to eliminate air in the feeder. The feed
system was connected to a nitrogen supply to evacuate polymer into the fluidised catalyst bed. Thus, commingled polymer
particles were purged under nitrogen into the top of the reactor

Table 1
Catalysts used in the catalytic degradation of polypropylene
Catalyst

FCC-s1
ZSM-5
HUSY
HAHA
Silicalite
a
b
c
d
e

Si/Al

2.1
17.5
13.6
3.6
>1000

Surface area (cm2/g)


BETa

Micropore

External

147
426
547
268
362

103
263
429
21
297

44
128
118
247
65

Total surface area (BET).


The catalyst was a mixture of zeolite, a silicaealumina matrix and binder, not determined.
Chinese Petroleum Corp., CPC, Taiwan, ROC.
BP Chemicals, Sunbury-on-Thames, UK.
Single-point BET determined.

Pore size (nm)

Commercial Name

eb
0.55  0.51
0.74
3.28e
0.55  0.51

Equilibrium catalystsc
ZSM-5 zeolited
Ultrastabilised Y zeolitec
Amorphous silica aluminac
Synthesized in-house

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

N2

815

V1

TC1

V3

TC2

V2

10

11

V4

TC3

Fig. 1. Schematic diagram of a catalytic fluidised-bed reactor system: (1) Feeder, (2) furnace, (3) sintered distributor, (4) fluidised catalyst, (5) reactor, (6) condenser, (7) flow meter, (8) 16-loop automated sample system, (9) gas bag, (10) GC, and (11) digital controller for three-zone furnace.

and allowed to drop freely into the fluidised-bed at t 0 min.


At sufficiently low polymer/catalyst ratios (as used in this
study) the outside of the catalyst particles are not wet with
polymer, so the catalyst particles move freely.
Volatile products leaving the reactor were passed through
a glass-fibre filter to capture catalyst fines, followed by an iceacetone condenser to collect any condensable liquid product.
A three-way valve was used after the condenser to route product
either into a sample gas bag or to an automated sample valve
system with 16 loops. The Tedlar bags, 15 L capacity, were
used to collect time-averaged gaseous samples. The bags were
replaced at intervals of 10 min throughout the course of reaction. The multiport sampling valve allowed frequent, rapid sampling of the product stream when required. Spot samples were
collected and analysed at various reaction times (t 1, 2, 3, 5,
8, 12, 15, and 20 min). The rate (Rgp, wt% min1) of hydrocarbon production of gaseous products collected by automated
sample system in each run was defined by the relationship:

after the polymer degradation was deemed residues and


contained involatile products and coke. The amount and nature
of the residues were determined by thermogravimetric analysis
(TGA).

3. Results and discussion


The reactor and various units of the collection system were
weighted before and after the runs to determine the mass balance. Catalytic pyrolysis products (P) are grouped together as
hydrocarbon gaseous (<C5), liquids (gasoline (C5eC9), condensate in condenser and captured in filter and BTX) and residues (coke and products, involatile at reaction temperature,
deposited on catalyst) to enable the overall pyrolysis processes
to be described more easily. A number of runs were repeated
in order to check their reproducibility. The term yield as
used in this paper is defined by the relationship:

Rgp hydrocarbon production rate of gaseous products in each spot rung=min


 100=total hydrocarbon production of gaseous products over the whole spot runsg
Gaseous hydrocarbon products were analysed using a gas
chromatograph equipped with (i) a thermal conductivity detector (TCD) fitted with a 1.5 m  0.2 mm i.d. molecular sieve
13X packed column and (ii) a flame ionisation detector
(FID) fitted with a 50 m  0.32 mm i.d. PLOT Al2O3/KCl capillary column. A calibration cylinder containing 1% C1eC5
hydrocarbons was used to help identify and quantify the gaseous products. The remaining solid deposited on the catalyst

yieldwt% Pg  100=polymer fedg


Mass balances of 90  5% were obtained for all experiments. Due to the high nitrogen flow rates used in this study,
it is difficult to completely recover all the lower molecular
weight material, and this results in some loss in the mass
balance.

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

816

3.1. Degradation of PP over silicalite and spent FCC


commercial catalyst
Product distributions for PP degradation over silicalite (Si/
Al > 1000) in the 330e450  C range are summarised in Table
2. At temperatures below 390  C, a large amount of solid residue, presumably unconverted commingled polymer and high
molecular weight degradation products, remained on the silicalite catalyst. The yield of gaseous and liquid hydrocarbon
products at 390  C was only 17.1 wt% (Table 2) compared
with 87.8 wt% (Table 3) when spent FCC commercial equilibrium catalyst (FCC-s1) was used. Typically thermal degradation productions were observed with silicalite showing
primary cracking products and an even spread of carbon numbers consisting of C3eC6 olefins products with some isomerisation of BTX. At higher temperatures, product streams
containing C1eC9 hydrocarbons were produced with gaseous
yield 21.2 wt% and liquid range yield 18.7 wt% of polymer
converted at 450  C.
Some similar trends in product yields were observed with
spent FCC catalyst (FCC-s1) as the reaction temperature was
increased. Gaseous and coke yields increased and involatile
residues (unreacted or partially reacted PP) and liquids decreased. Product distributions with FCC-s1 catalyst contained
more olefinic materials in the range of C3eC7 (about 61 wt%

Table 2
Summary of products of PP degradation over silicalite catalyst
Degradation results

360

Degradation results

330

390

420

450

c
d

Yield (wt% feed)


P
Gaseous ( C1eC4)
3.5
7.1
Liquid
2.5
4.3
P
Gasoline ( C5eC9)
1.7
2.8
Condensate liquida
0.7
1.4
0.1
0.1
BTXb
Residuec
94.0
88.6
Distribution of C1eC9 hydrocarbon products (wt%
C1
n.d
n.d
n.d
n.d
C2
C]
n.d
0.2
2
C3
n.d
0.1
1.8
2.8
C]
3
C4
n.d
0.4
C]
1.6
3.6
4
C5
n.d
n.d
0.9
1.8
C]
5
C6
n.d
ed
]
C6
0.5
0.7
C7
n.d
n.d
0.1
0.3
C]
7
C8
n.d
n.d
d
]
C
e
ed
8
P
C9
n.d
n.d

10.2
6.9
5.1
1.6
0.2
82.9
feed)
n.d
n.d
0.2
0.3
4.2
0.6
4.9
n.d
3.1
ed
1.4
0.2
0.3
n.d
0.1
ed

15.9
13.0
8.9
2.1
0.2
72.9

21.2
18.7
13.4
2.8
0.4
60.1

ed
0.1
0.3
0.8
5.6
0.9
8.2
0.3
3.8
0.1
2.6
0.3
1.4
ed
0.3
0.1

0.1
0.1
0.5
0.9
6.8
1.3
11.5
0.5
5.2
0.2
4.1
0.5
1.8
0.2
0.6
0.3

Fluidising N2 rate 600 ml min1, catalyst particle size 75e180 mm, polymer to catalyst ratio 30 wt% and total time of collection 60 min.
a
Condensate liquid: condensate in condenser and captured in filter.
b
BTX: benzene, toluene and xylene.
c
Residue: coke and involatile products.
d
e: less than 0.01 (wt%); n.d: not detectable.

Temperature ( C)
360

Yield (wt% feed)


P
Gaseous ( C1eC4)
20.8
22.3
Liquid
64.4
64.5
P
Gasoline ( C5eC9)
57.4
56.9
Condensate liquida
6.6
6.9
0.4
0.7
BTXb
Residuec
14.8
13.2
Involatile residue
13.2
11.3
Coke
1.6
1.9
Mass balance (%)
88.2
89.3
Distribution of C1eC9 hydrocarbon products (wt%
C1
n.d
n.d
C2
n.d
n.d
C]
ed
0.1
2
C3
1.1
1.5
C]
5.6
6.6
3
1.9
1.6
C4
C]
12.2
12.6
4
C5
1.8
2.4
C]
19.8
18.9
5
3.4
3.5
C6
C]
17.5
16.2
6
C7
3.7
4.6
6.6
6.2
C]
7
C8
1.2
1.2
]
C
2.8
3.1
8
P
C9
0.6
0.8
a

Temperature ( C)
330

Table 3
Summary of products of PP degradation over FCC-s1 catalyst (fluidising N2
rate 600 ml min1, catalyst particle size 75e180 mm, polymer to catalyst
ratio 30 wt%, and total time of collection 30 min)

390

420

450

26.9
60.9
54.3
5.4
1.2
12.2
9.9
2.3
91.8
feed)
ed
ed
0.1
1.7
8.8
2.2
14.1
2.7
16.5
3.5
14.2
4.5
7.4
1.5
2.8
1.2

31.2
58.2
52.7
3.9
2.0
10.2
7.6
2.6
92.7

34.2
56.7
50.9
3.3
2.5
9.1
6.0
3.1
93.2

ed
ed
0.3
2.3
10.3
2.9
15.4
3.3
16.2
3.2
11.9
3.8
8.0
1.8
2.8
1.7

ed
0.2
0.4
2.7
11.2
3.2
16.5
3.4
13.7
4.6
10.3
3.3
8.7
2.1
3.2
1.6

Condensate liquid: condensate in condenser and captured in filter.


BTX: benzene, toluene and xylene.
Residue: coke and involatile products.
e: less than 0.01 (wt%); n.d: not detectable.

at 390  C) with minor products, methane and ethane, only


detectable at the higher reaction temperatures. The rate of hydrocarbon production as a function of time for PP degradation
over FCC-s1 catalyst at different reaction temperatures is compared in Fig. 2 and, as expected, faster rates were observed at
higher temperatures. At 450  C, the maximum rate of hydrocarbon production was 36 wt% min1 after only 2 min with
all the polymer degraded after approximately 11 min. As the
temperature of reaction was decreased, the initial rate of
hydrocarbon production dropped and the time for PP to be degraded lengthened. At 330  C the rate of hydrocarbon production was significantly lower with the polymer being degraded
more slowly over 20 min.
3.2. Effect of reaction conditions on PP degradation
over spent FCC catalyst
The effect of reaction conditions including flow rates of fluidising gas (270e900 ml min1), ratios of polymer to catalyst
feed (1:1e1:6) and catalyst type (FCC-s1) has been investigated. The results shown in Fig. 3 illustrate that for efficient
PP degradation good mixing is required, with a dramatic
drop-off in the rate of degradation observed only at the lowest

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821
40
360C

Degradation results

390C

30

Rgp (wt%/min)

Table 4
Product distributions shown from FCC-s1 catalysed degradation of PP at different fluidising N2 rates

330C

35

450C

20
15
10
5
0

10

12

14

16

18

20

Time (min)
Fig. 2. Comparison of hydrocarbon yields as a function of time at different
reaction temperatures for the catalytic degradation of PP over spent FCC
commercial catalyst (FCC-s1) (rate of fluidisation gas 600 ml min1,
catalyst particle size 75e180 mm, polymer to catalyst ratio 30 wt%).

fluidising flow used (300 ml min1). Furthermore, changing


the fluidising flow rate influences the product distribution. At
low flow rates (high contact times for primary products), secondary products are observed with increased amounts of coke
precursors (BTX) although the overall degradation rate is
slower as shown by increasing amounts of partially depolymerised products (Table 4).
The amount of FCC-s1 used in the degradation of PP polymer remained constant and, therefore, as more waste polymers
were added to the reactor then fewer catalytic sites per unit
weight of catalyst were available for cracking. The overall effect of increasing the polymer to catalyst ratio from 0.1:1 to
0.6:1 on the rate of hydrocarbon generation was small but predictable (Fig. 4). As the polymer to catalyst ratio increases, the
possibility of PP polymer adhesion to the reactor wall

Yield (wt% feed)


P
Gaseous ( C1eC4)
Liquid
P
Gasoline ( C5eC9)
Condensate liquida
BTXb
Residuec
Involatile residue
Coke
Mass balance (%)

450

300

29.6
60.5
55.1
4.9
0.5
9.9
7.7
2.2
89.2

28.8
60.4
54.7
4.9
0.8
10.8
8.5
2.3
89.6

26.9
60.9
54.3
5.4
1.2
12.2
9.9
2.3
91.8

26.3
60.7
54.6
4.5
1.6
13.0
10.5
2.5
90.3

26.1
60.1
53.7
4.3
2.1
13.8
11.2
2.6
94.1

10% wt/wt
30

600 ml/min
450 ml/min

20% wt/wt
30% wt/wt

25

300 ml/min

Rgp (wt%/min)

Rgp (wt%/min)

600

35

750 ml/min

25

750

increases as the amount of unreacted polymer waste in the reactor rises. The total product yield after 20 min showed only
a slight downward trend even after a 6-fold increase in added
polymer waste. This can be attributed to the sufficient cracking
ability of FCC-s1 and excellent contact between PP polymer
and catalyst particles. As more PP was added, lower C5eC9
gasoline yields but higher liquid yields and involatile products
were observed (Table 5). More BTX (coke precursor) was produced but increasing the polymer to catalyst ratio had only virtually no effect on C1eC4 hydrocarbon gases production.
Both the carbon number distribution of the products of PP
polymer cracking at 390  C over FCC-s1 catalyst, zeolitic catalysts (ZSM-5 and HUSY) and non-zeolitic amorphous silica
alumina (SAHA) used in this study and the nature of the product distribution were found to vary with the catalyst used. As
shown in Table 6, the yield of volatile hydrocarbons for

900 ml/min
30

900

Reaction temperature 390  C, catalyst particle size 75e180 mm, polymer


to catalyst ratio 30 wt%, and total time of collection 30 min.
a
Condensate liquid: condensate in condenser and captured in filter.
b
BTX: benzene, toluene and xylene.
c
Residue: coke and involatile products.

35

20
15
10
5
0

fluidising N2 rates (mL/min)

420C

25

817

40% wt/wt
60% wt/wt

20
15
10
5

10

12

14

16

18

20

Time (min)
Fig. 3. Comparison of hydrocarbon yields as a function of time at different
fluidisation gas for the degradation of PP over FCC-s1 catalyst (reaction
temperature 390  C, catalyst particle size 75e180 mm, polymer to catalyst
ratio 30 wt%).

10

12

14

16

18

20

Time (min)
Fig. 4. Comparison of hydrocarbon yields as a function of time at different
ratios of polymer to catalyst for the degradation of PP over FCC-s1 catalyst
(reaction temperature 390  C, catalyst particle size 75e180 mm, rate of
fluidisation gas 600 ml min1).

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

818

35

Table 5
Product distributions shown from FCC-s1 catalysed degradation of PP at different ratios of polymer to catalyst (reaction temperature 390  C, catalyst
particle size 75e180 mm, fluidising N2 rate 600 ml min1, and total time
of collection 30 min)

Yield (wt% feed)


P
Gaseous ( C1eC4)
Liquid
P
Gasoline ( C5eC9)
Condensate liquida
BTXb
Residuec
Involatile residue
Coke
Mass balance (%)
a
b
c

Ratio of polymer to catalyst (% wt)


10

20

30

40

60

26.7
63.2
57.4
4.9
0.9
10.1
7.6
2.5
89.5

27.8
61.7
55.4
5.3
1.0
10.5
8.3
2.2
90.6

26.9
60.9
54.3
5.4
1.2
12.2
9.9
2.3
91.8

27.6
60.3
52.9
5.8
1.6
12.1
10.4
1.7
92.5

28.2
59.2
51.4
6.0
1.8
12.6
10.7
1.9
90.4

Condensate liquid: condensate in condenser and captured in filter.


BTX: benzene, toluene and xylene.
Residue: coke and involatile products.

zeolitic catalysts (ZSM-5 z HUSY) gave higher yield than


spent FCC commercial catalyst (FCC-s1) and non-zeolitic catalysts (FCC-s1 z SAHA) and the highest was obtained for
ZSM-5 (nearly 90 wt%). Overall, the bulk of the products observed with these acidic cracking catalysts (FCC-s1, ZSM-5,
HUSY and SAHA) were in the gaseous and liquid phase
with less than 13 wt% solid residues of involatile residue
and coke collected. The differences in the product distributions between those catalysts can be seen with ZSM-5 producing a much more C1eC4 hydrocarbon gases (56 wt%) than
FCC-s1, HUSY and SAHA catalysts. Some similarities
were observed between FCC-s1 and SAHA with C1eC4 and
C5eC9 yields, which were approximately 27e30 wt%
and 53e54 wt%, respectively. The highest level of unconverted polymer was observed with silicalite, while the highest
coke yields were observed with HUSY. The rate of gaseous
hydrocarbon evolution further highlights the slower rate of
degradation over silicalite catalyst as shown in Fig. 5 when
comparing all catalysts under identical conditions. The results
Table 6
Summary of products of PP degradation over various commercial catalysts
Degradation results

Yield (wt% feed)


P
Gaseous ( C1eC4)
Liquid
P
Gasoline ( C5eC9)
Condensate liquida
BTXb
Residuec
Involatile residue
Coke
Mass balance (%)

Catalyst type
FCC-s1

ZSM-5

HUSY

SAHA

Silicalite

26.9
60.9
54.3
5.4
1.2
12.2
9.9
2.3
91.8

56.2
38.8
33.5
3.5
1.8
5.0
3.1
1.9
94.5

33.5
58.7
54.8
3.4
0.5
7.8
2.9
4.9
92.4

30.2
56.9
52.9
3.7
0.3
12.9
10.5
2.4
90.5

10.2
6.9
5.1
1.6
0.2
82.9
80.3
2.6
94.1

Reaction temperature 390  C, fluidising N2 rate 600 ml min1, polymer to


catalyst ratio 30 wt%, and total time of collection 30 min.
a
Condensate liquid: condensate in condenser and captured in filter.
b
BTX: benzene, toluene and xylene.
c
Residue: coke and involatile products.

HUSY
ZSM-5

Rgp (wt%/min)

Degradation results

FCC-s1
30
25

SAHA
Silicalite

20
15
10
5
0

10

12

14

16

18

20

Time (min)
Fig. 5. Comparison of hydrocarbon yields as a function of time for the catalytic degradation of PP at 390  C over different catalysts (waste polymer to
catalyst ratio 30 wt%, rate of fluidisation gas 600 ml min1).

of the products of PP degradation reflect the differing cracking


effect of FCC-s1 catalyst compared with the zeolite and nonzeolitic materials. The maximum rate of generation was observed after 2 min with the zeolite catalysts whereas the maximum was observed after 3 min with FCC-s1 and SAHA.

3.3. Variation of product stream with catalytic selectivity


and catalyst deactivation
P
P ]
Equilibrium ratios of i-butene/ butenes (i-C]
C4 ) and
4 /
i-butane/n-butane (i-C4/n-C4) were predicted using Gibbs free
energy minimisation on the PRO/II package for the temperatures used experimentally and are presented alongside the corresponding experimental results in Table 6. The i-butene/
P
butenes ratio is very close to the predicted equilibrium
values and thus the reactions involved in the production and
interconversion of butenes are very fast over FCC-s1, and their
ratio is primarily equilibrium controlled. The i-butane/n-butane ratio reflects the involvement of tertiary C4 carbenium
ions in bimolecular hydrogen transfer reactions and since tertiary carbenium ions are more stable than primary ions,
a higher yield of iso-butane would be expected. As can be
seen in Table 7, the observed i-butane/n-butane ratios at
390  C are well above calculated equilibrium values consistent
with the cracking of long chain hydrocarbon molecules to
yield iso-butylcarbenium ions which provide a source for i-butane, via hydrogen transfer, or i-butene. The selectivity could
be varied by changes in different operating conditions used
in this study. Further evidence of the increase in secondary reactions, for example, bimolecular
hydrogen
transfer, was seen
P
P
in the lowering of the
olefin/ paraffin (o/p 4.52 at
330  C versus o/p 2.86 at 450  C) and i-butane/n-butane
(i-C4/n-C4 3.23 at 330  C versus i-C4/n-C4 1.83 at
450  C) ratios as temperature increases, in the experimental
range. At fast flow rates, primary cracking products are favoured
as evidenced
by the increasing ratios of i-butene/
P ]
P
1
/
C
N2 fluidising
butenes (i-C]
4
4 2.12 in 300 ml min

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

819

Table 7
Influence of reaction conditions on product selectivity for the catalysed degradation of PP over FCC-s1 catalyst: experimental and predicted equilibrium results
Reaction conditiona

Ratio

Temperatureb ( C)
P
i-Butene/ butenes
P
i-Butene/ butenese
i-Butane/n-butane
i-Butane/n-butanee
P
SOlefins/ paraffinsf
a
b
c
d
e
f

N2 rated (ml min1)

P/C ratioc (wt%)

330

390

450

10

30

60

300

600

900

0.55
0.54
3.23
1.04
4.52

0.49
0.50
2.47
0.89
3.91

0.43
0.45
1.83
0.77
2.86

0.57

0.52

0.54

0.46

0.51

0.59

3.46

2.47

2.29

2.12

2.47

3.53

3.61

3.91

3.43

3.65

3.91

4.28

Represents a series of runs with different reaction temperature, polymer to catalyst ratio and N2 fluidising rate.
Polymer to catalyst ratio 30 wt% and 600 ml min1 N2 fluidising rate.
Reaction temperature 390  C and fluidising N2 600 ml min1.
Polymer to catalyst ratio 30 wt% and reaction temperature 390  C.
Predicted equilibrium data.
Denotes the ratio of the sum all olefinic to paraffinic products.

P ]
rate versusPi-C]
C4 3.53 in 900 ml min1 N2 fluidising
4 / P
rate) and olefin/ paraffin (o/p 3.65 in 300 ml min1 N2
fluidising rate versus o/p 4.28 in 900 ml min1 N2 fluidising
rate).

FCC-s1

Wt %

30
20
10
0

Wt %

20
10
0

ZSM-5

30

Wt %

HUSY

30

20
10
0

SAHA

30

Wt %

The relation in catalytic activity to catalyst deactivation


was examined by the transient change in the amount of gaseous compounds produced. Rapid variation in the product
stream of FCC-s1 and HUSY catalysts was observed (Fig. 6)

20
10
0

Time (min)
i-C4

Fig. 6. Some of hydrocarbon products of isobutene (i-C4),


catalysts.

tot C4=

tot C5=

P
butanes (tot C4]) and pentanes (tot C5]) as a function of time for PP degradation over different

820

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821

when the spot samples, taken during the course of the reaction,
were analysed. The deactivation is reflected in the decrease of
the amount of iso-butane (i-C4) produced (product of bimolecular reaction) and the relative increase in olefins (product of
]
monomolecular reaction), exemplified by, C]
4 and C5 . The
spent FCC-s1 catalyst with bimodal pore structures, which is
composed of both the micropore of zeolite and the mesopore
of silicaealumina, may allow bulky reactions to occur, ultimately leading to the generation of coke and subsequently deactivation of the catalyst. The deactivation was more
exaggerated in the case of HUSY with its large pore openings
and internal supercages. In contrast, ZSM-5 is resistant to coking when coke builds up on outer surface and the product
stream remains essentially unchanged, whereas the weakness
and lower density of the acid sites in SAHA along with the increased tolerance to coke in the amorphous structure is most
likely the reason for the lack of variation in the product stream
over this catalyst.

4. Conclusions
A laboratory catalytic fluidised-bed reactor has been used to
obtain a range of volatile hydrocarbons by catalytic degradation
of polypropylene in the temperature range of 330e450  C. The
catalytic degradation of PP over spent FCC catalyst performed
in fluidised-bed reactor was shown to be a useful method for the
production of potentially valuable hydrocarbons. In the presence of the spent FCC commercial catalyst at 390  C used in
this work, conversion of PP polymer to volatile hydrocarbons
in the catalytic fluidised-bed reactor was more than 87 wt% of
feed in 20 min, while silicalite yielded less than 18 wt% of
feed after 60 min. The use of fluidised-bed reaction system
coped with a spent FCC equilibrium catalyst can be a better
option from economical point of view since it can gives
a good conversion with comparable short reaction time, and
even its activity is lower than that of the zeolites (ZSM-5 and
HUSY) and silicaealuminas (SAHA), this can be compensated
by increasing the catalyst to PP ratio.
Product distributions with FCC-s1 catalyst contained more
olefinic materials in the range of C3eC7 (about 56 wt% at
390  C). Experiments carried out with FCC-s1 catalyst gave
good yields of volatile hydrocarbons with differing selectivity
in the final products dependent on reaction conditions. However, silicalite give very low conversions of polymer waste
to volatile hydrocarbons compared with spent FCC catalyst
(FCC-s1) under the same reaction conditions. The selectivity
could be further influenced by changes in operating conditions; in particular, olefins and iso-olefins were produced by
low temperatures and short contact times. From a practical
point of view, the use spent equilibrium catalyst from FCC
units can be the most adequate solution. It is concluded that
the use of spent FCC commercial catalyst and under appropriate reaction conditions can have the ability to control both the
product yield and product distribution from polymer degradation, potentially leading to a cheaper process with more valuable products.

Acknowledgements
The authors would like to thank the National Science Council (NSC) of the Republic of China (ROC) for financial support (NSC 95-2211-E-244-013). Thanks also are due to
Professor M.D. Ger and Dr. C.-M. Chiu for samples of
ASA, spent FCC commercial catalyst and surface area/pore
size measurements.

References
[1] Scott G. Polymers and the environment. London: Royal Society of
Chemistry; 1999.
[2] Brandrup J, Bittner M, Michaeli W, Menges G. Recycling and recovery
of plastics. Munich, New York: Carl Hanser Verlag; 1996.
[3] Lin YH, Sharratt PN, Garforth AA, Dwyer J. Catalytic conversion of
polyolefins to chemicals and fuels over various cracking catalysts. Energy Fuels 1998;12:767.
[4] Puente G, Sedran U. Recycling polystyrene into fuels by means of FCC:
performance of various acidic catalysts. Appl Catal B Environ 1998;
19(3):305.
[5] Aguado J, Serrano DP, Escola JM, Garagorri E, Fernandez JA. Catalytic
conversion polyolefins into fuels over zeolite beta. Polym Degrad Stab
2000;70(1):97.
[6] Lee SY, Yoon JH, Kim JR, Park DW. Catalytic degradation of polystyrene over natural clinoptilolite zeolite. Polym Degrad Stab 2001;
74(2):297.
[7] Dawood A, Miura K. Catalytic pyrolysis of g-irradiated polypropylene
over HY-zeolite for enhancing the reactivity and the product selectivity.
Polym Degrad Stab 2002;76(1):45.
[8] Monos G, Yusof IY, Gangas NH, Papayannakos NK. Tertiary recycling of
polyethylene to hydrocarbon fuel by catalytic cracking over aluminum
pillared clays. Energy Fuels 2002;16:485.
[9] Marcilla A, Gomez A, Reyes-Labrta A, Giner A. Catalytic pyrolysis
of polypropylene using MCM-41: kinetic model. Polym Degrad Stab
2003;80:233.
[10] Gao Z, Ksuyoshi K, Amasaki I, Nakada M. A kinetic study of thermal
degradation of polypropylene. Polym Degrad Stab 2003;80:269.
[11] Durmus A, Koc SN, Pozan GS, Kasgoz A. Thermal-catalytic degradation
kinetics of polyethylene over BEA, ZSM-5 and MOR zeolites. Appl
Catal B Environ 2005;61:316.
[12] Ohkita H, Nishiyama R, Tochihara Y, Mizushima T, Kakuta N,
Morioka Y, et al. Acid properties of silicaealumina catalysts and catalytic degradation of polyethylene. Ind Eng Chem Res 1993;32:3112.
[13] Uemichi Y, Nakamura J, Itoh T, sugioka Garforth AA, Dwyer J. Conversion of polyethylene into gasoline-range fuels by twoestage catalytic
degradation using silicaealumina and HZSM-5 zeolite. Ind Eng Chem
Res 1999;38:385.
[14] Puente G, Klocker C, Sedran U. Conversion of waste plastics into fuels:
Recycling polyethylene in FCC. Appl Catal B Environ 2002;36(4):279.
[15] Lin YH, Hwu WH, Ger MD, Yeh TF, Dwyer J. A combined kinetic and
mechanistic modeling of the catalytic of polymers. J Mol Catal A Chem
2001;171:143.
[16] Lin YH, Yang MH, Yeh TF, Ger MD. Catalytic degradation of high density polyethylene over mesoporous and microporous catalysts in a fluidised-bed reactor. Polym Degrad Stab 2004;86:121.
[17] Lin YH, Yen HY. Fluidised bed pyrolysis of polypropylene over cracking
catalysts for producing hydrocarbons. Polym Degrad Stab 2005;89:101.
[18] Lin YH, Yang MH. Catalytic reactions of post-consumer polymer waste
over fluidized cracking catalysts for producing hydrocarbons. J Mol Catal
A Chem 2005;231:113.
[19] Salvador CC, Avelino C. Tertiary recycling of polypropylene by catalytic
cracking in a semibatch stirred reactor: use of spent equilibrium FCC
commercial catalyst. Appl Catal B Environ 2000;25:151.

Y.-H. Lin, M.-H. Yang / Polymer Degradation and Stability 92 (2007) 813e821
[20] Lee KH, Noh NS, Shin DH, Seo YH. Comparison of plastic types for catalytic degradation of waste plastics into product with spent FCC catalyst.
Polym Degrad Stab 2002;78:539.
[21] Barrer RM. Hydrothermal chemistry of zeolites. London: Academic
Press; 1982.

821

[22] Lin YH. Experimental and theoretical studies on the catalytic degradation of polymers, PhD Thesis, UMIST; 1998.
[23] Camblor MA, Corma A, Martinez A, Mocholi FA. Benefits in activity
and selectivity of small Y zeolite crystallites stabilized by a higher silicon to aluminium ratio by synthesis. Appl Catal 1989;55:65.

You might also like