You are on page 1of 10

This article was downloaded by: [Moskow State Univ Bibliote]

On: 09 October 2013, At: 23:51


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Petroleum Science and Technology


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lpet20

Transportation Fuels From Catalytic


Co-pyrolysis of Plastic Wastes With
Petroleum Residues: Evaluation of
Catalysts by Thermogravimetric Analysis
a a
M. F. Ali & M. S. Qureshi
a
H.E.J. Research Institute of Chemistry, International Center for
Chemical and Biological Sciences, University of Karachi , Karachi ,
Pakistan
Published online: 24 Jun 2013.

To cite this article: M. F. Ali & M. S. Qureshi (2013) Transportation Fuels From Catalytic Co-pyrolysis
of Plastic Wastes With Petroleum Residues: Evaluation of Catalysts by Thermogravimetric Analysis,
Petroleum Science and Technology, 31:16, 1665-1673, DOI: 10.1080/10916466.2010.551239

To link to this article: http://dx.doi.org/10.1080/10916466.2010.551239

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Petroleum Science and Technology, 31:1665–1673, 2013
Copyright © Taylor & Francis Group, LLC
ISSN: 1091-6466 print/1532-2459 online
DOI: 10.1080/10916466.2010.551239

Transportation Fuels From Catalytic Co-pyrolysis of


Plastic Wastes With Petroleum Residues: Evaluation of
Catalysts by Thermogravimetric Analysis
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

M. F. Ali1 and M. S. Qureshi1


1
H.E.J. Research Institute of Chemistry, International Center for Chemical and Biological
Sciences, University of Karachi, Karachi, Pakistan

Thermal or catalytic pyrolysis processes have attracted much attention for the conversion of plastic
wastes to a mixture of their basic hydrocarbons, which can be valuable either as fuels or as raw
chemicals. Especially catalytic pyrolysis has the advantage over non-catalytic pyrolysis as the process
can be tailored to produce much useful transportation fuels such as gasoline and diesel oils. The
effect of different catalysts on the thermal degradation of polypropylene (PP) and its copyrolysis
with petroleum vacuum residue (VR) has been studied by thermogravimetric analysis (TGA). TGA
experiments were used to compare the activity of different catalysts towards PP degradation alone
and mixture with VR. All the hydrocracking (HC) catalysts enhanced the copyrolysis of PP/VR
mixtures. The reaction rates were found to increase with catalyst fraction, acidity, and a reduction in
the catalysts particle size. The temperatures of onset, Ton , of maximum-rate, Tmax , and of end, Tend ,
of the degradation also shifted to lower values. Among HC catalysts, it seems that Z-713 performed
better at all three weight fractions; Z-713 reduced both T1% and T99% better than others. On the other
hand, the FCC catalyst (RCD-8) performed poorly for the degradation of PP. Various titania based
catalysts prepared in our laboratory performed well. The number of acid sites and catalyst pore size
along with impregnation with transition metals such as W/Ni and Ni/Mo are found as the key factors
for the energy efficient conversion of polymers and VR to liquid hydrocarbons. The results will help
in our ongoing study on conversion of waste plastics into feed stocks for chemicals and transportation
fuels.

Keywords: TGA, pyrolysis, catalytic cracking, polypropylene, petroleum residues, waste plastics,
catalysts

1. INTRODUCTION

The applications of plastic materials are continually increasing in the modern world. Presently
plastics are manufactured for various uses such as: consumer packaging, wires, pipes, containers,
appliances, electrical/electronics parts, and automotive parts. The increase in plastic materials
consumption has led to a parallel rise in the generation of plastic wastes. As not all waste plastic
may be recycled and most of them are not biodegraded, the disposal of this waste has become

Address correspondence to M. F. Ali, H.E.J. Research Institute of Chemistry, International Center for Chemical and
Biological Sciences, University of Karachi, Karachi, Pakistan. E-mail: mfali2k@yahoo.com

1665
1666 M. F. ALI AND M. S. QURESHI

a major social concern. Therefore, measures have to be implemented to reduce plastic wastes
negative impact on the environment. Land filling of plastic waste is not the right solution due to the
danger of leaching and soil impregnation of plastic additives, such as various dyes and phthalates,
polluting ground water. Moreover, the number of landfill sites is decreasing. Incineration process
is also not acceptable because it produces several pollutants that are dangerous to the environment.
In addition, both processes do not allow the recovery of the organic content of plastic waste, which
should be part of the organic life-cycle (Pinto et al., 1999). Feedstock recycling of plastic wastes
is regarded currently as a promising alternative for the management of plastic wastes, as this
process transforms plastics to their constituent monomers or their basic hydrocarbon feed stock.
Thermal or catalytic cracking of plastic wastes yields a mixture of their basic hydrocarbons,
which can be valuable either as fuels or as raw chemicals (Serrano et al., 2000). The energy
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

consumption during the thermal cracking (pyrolysis) is very high, and the molecular weight
distribution of the products obtained is quite broad, varying with the conditions used (Marcilla
et al., 2001). However, the selectivity of the product obtained can be controlled by the use of
suitable catalysts (catalytic cracking) under appropriate conditions of temperature, pressure and
atmosphere (N2 , H2 , or air). Catalytic cracking of plastics requires lower energy consumption
(temperatures from 350ı C–550ıC), and the chemical distribution of the product is narrower than
in thermal process, leading to the production of more valuable products (Ali and Siddiqui, 2005).
Due to the collection problems of waste plastics, it is often difficult to maintain a steady supply
of material to the processing plants. To overcome such problems the waste plastics can be co-
processed by blending with other hydrocarbon sources such as heavy petroleum oils and residues.
This approach would provide a steady supply as well as upgrade both waste plastics and heavy
residues. The co-processing of waste plastics with heavy petroleum oils, tar mats, and petroleum
residues has been reported (Joo and Curtis, 1996; Luo and Curtis, 1996; Ali and Siddiqui, 2006).
Typical catalysts used for the cracking of heavy petroleum residues are acidic solids like
amorphous silica-alumina, zeolites, mesophorous materials, and activated carbon (Kim et al.,
2000). Similar catalysts were found to be effective for the pyrolysis of polymers. The number of
acid sites and catalyst pore size are described as the two key factors for the cracking of long chain
hydrocarbons present in petroleum residues and polymers. Many researchers have studied the use
of various zeolites, such as ZSM-5, Y, beta, and mordenite (Geraldo and Filho, 2005). Owing to
its strong acidity, ZSM-5 has been investigated most intensively. The application of ZSM-5 to the
pyrolysis of waste plastics, however, has revealed several technical problems, such as the relatively
high quantity of gas products and coke formation near the pore entrance. Mesoporous zeolites, such
as, MCM-41 was found to have a high potential for use as a pyrolysis catalyst for waste plastics
because of its high surface area, tunable uniform mesopores (from 20ı to 100ıA), and moderate
acid strength. The kinetics of degradation of polyolefins by thermogravimetric analysis (TGA)
in the presence of solid catalysts of different acidity was reported (Craniti and Gervasini, 2001).
In the present work, the pyrolytic behavior of plastics alone and mixed with heavy petroleum
residues were investigated under controlled atmosphere using thermogravimetric analysis (TGA).
TGA is a thermal analysis technique which measures the amount and rate of change in the
weight of the material as a function of temperature or time in a controlled atmosphere. TGA
measurements are used primarily to determine the composition of materials and to predict their
thermal stability up to elevated temperatures. However, with proper experimental procedures, TGA
technique allows a comparison of the activity of different catalysts towards cracking of polymers
and petroleum residues. The activity of different hydrocracking catalysts and a composite titania-
alumina and USY zeolite catalyst (prepared by the authors of this study) has been evaluated
on the basis of TGA data. The information thus collected can be used for feedstock recycling
of plastic waste alone or coprocess with other hydrocarbon sources. Among various plastics,
polypropylene (PP) has relatively high melting point and a high resistance to chemicals; and the
demand for polypropylene (PP) in the world is approximately 25 wt% of thermoplastic and is
TRANSPORTATION FUELS FROM CATALYTIC CO-PYROLYSIS 1667

steadily increasing. Consequently PP, as one of the largest parts of both industrial and domestic
wastes, is selected for this study as a representative of waste plastics. The information regarding
the degradation mechanism and reaction parameters, thus obtained will help in designing a pilot
plant reactor to convert plastic wastes into chemicals and fuels.

2. EXPERIMENTAL DESIGN

2.1. Materials
The model plastics that were chosen in this study included low-density polyethylene (LDPE),
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

high-density polyethylene (HDPE), polypropylene (PP) and polystyrene (PS). These plastics were
used in the same form as received from the Saudi Basic Industries Corporation (SABIC), Riyadh.
Vacuum residue (VR) oil samples were obtained from National Refinery Ltd., Karachi, Pakistan.
The properties and characteristics of VR are shown in Table 1.

2.2. Catalysts
The suitability of solids as heterogeneous catalysts for the cracking processes as in the hydrocrack-
ing of petroleum residues and pyrolysis of plastics is based on the four essential characteristics:
activity, selectivity, ease of regeneration and mechanical strength. The four commercially available
catalysts (C-2 to C-5) were used because of their known selectivity towards the hydrocracking
reactions. Catalyst C-1 represents one of various titania based catalysts prepared in our laboratory
using a method of mixing, metals impregnation, drying, and calcinations according to method
described by Ali and Asaoka (2009). The description and characteristics of all catalysts (C-1 to
C-5) are given in Table 2.

2.3. Thermal Analysis (TGA)


The non-isothermal thermo gravimetric analysis was performed on a TA Instruments SDT Q600,
simultaneous TGA-DTA-DSC analyzer. The experiments were conducted under flowing atmo-
sphere of nitrogen at a purge rate of 200 mL/min. The samples were studied in the fine powder
form. A quantity of 3:5 ˙ 0:3 mg was placed in an open ceramic sample pan. The sample was then
equilibrated to 200ıC before being heated to 600ıC at different heating rates (3ıC–10ıC/min).
The actual heating rate was calculated from temperature measurements made during the period
of sample’s decomposition. In this study, this analysis serves primarily as an assessment tool in
the screening of various potential catalysts for the cracking of polymer/VR mix. An assessment
of the catalyst performance prior to use in a reactor reduces cost.

TABLE 1
Properties of Petroleum Vacuum Residue (VR)

Vacuum Residue (Resids)

Gravity ı API 9.03


Sulfur (wt%) 3.85
Carbon residue (wt%) 13.2
Asphaltene (wt%) 10.9
Viscosity, cst @ 98.9ı C 39.75
1668 M. F. ALI AND M. S. QURESHI

TABLE 2
Properties of Catalysts

Pore Wt. Loss Wt. Loss


Description and Surface Volume, Total Prior After
Name Composition Area, m2 /g cc/g Acidity 100ı C 100ıC

C-1 NiMo loaded on [TiO2 C alpha 359.0 0.42 1.48 3.26 7.36
alumina C AP-1 C USY] Extrudates
C-2 KC-2710 (AKZO Nobel) 182.0 0.23 0.85 2.96 8.40
C-3 Z-713 (Zeolyst International) 221 0.34 1.15 1.45 4.57
C-4 HC-100 (UOP) 231 0.25 1.20 3.67 9.73
C-5 RCD-8 (UOP) 210 0.21 1.50 2.19 6.28
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

2.4. TGA Sample Preparation


For the catalytic pyrolysis study by TGA, the samples were prepared by mixing dried proportions
of plastic and catalysts (dry mixing). Plastic was first frozen in liquid nitrogen and crushed into a
very fine powder before mixing with an appropriate amount of very fine catalyst. For co-processing
reactions PP and VR were mixed in a 1:1 ratio.

3. RESULTS AND DISCUSSION

3.1. Catalyst(s) Weight Loss


The thermal stability of catalysts was assessed by recording weight loss during TGA runs on all
five catalysts (Figure 1). Calcium oxalate monohydrate was used as a calibration standard. This
material has become popular for demonstrating thermobalance performance, as its TG gives three
distinct weight losses over a wide temperature range. The measured losses (Figure 1) are with in
the agreeable limits with theoretical losses, according to the usual scheme as shown in Eq. (1):

CaC2 O4 :H2 O ! CaC2 O4 ! CaCO3 ! CaO: (1)

The curve gives a value of 18.50% lost, compared with the theoretical value of 19.2%, for the
decomposition of the anhydrous oxalate to calcium carbonate approximately 500ı C, which is due
to the disproportionation of carbon monoxide to carbon dioxide and carbon [2CO ! C C CO2 ].
The residual calcium oxide was pale gray due to the carbon deposited.
All five catalysts showed weight loss approximately 100ıC which indicates water loss. The
weight change is observed to be more pronounced for the hydrocracking catalysts C-2 and C-4.
The catalyst prepared by these authors (C-1) also showed a relatively high loss.

3.2. Pyrolysis Study of Individual Materials by TGA


The pyrolysis behavior of the individual components (plastics and petroleum residue) of the
copyrolizing system were studied in order to establish a baseline and also to determine the nature
of any interactions between copyrolizing materials. Figure 2 illustrates TGA curves for different
plastics (HDPE, LDPE, PP, PS, PET, and PVC). As is evident from one main distinct weight loss
step in the figures, the degradation behaviors of all plastics except polyvinyl chloride polymer
(PVC) are similar showing a constant degradation behavior at the involved temperature range
TRANSPORTATION FUELS FROM CATALYTIC CO-PYROLYSIS 1669
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

FIGURE 1 Weight loss % for five catalysts and calcium oxalate monohydrate.

FIGURE 2 Thermogravimetric analysis (TGA) thermograms for virgin plastics.


1670 M. F. ALI AND M. S. QURESHI

and are completely volatilized, leaving no char yield. Also, these plastics have a very narrow
temperature range (420ıC and 490ı C) over which they volatilize. However, the TG behavior of
PVC shows a completely different behavior due to loss of chlorine in the form of HCl and other
organo halogen compounds. Similar to other plastics, PVC are completely volatilized, leaving no
char yield. The different macromolecular structure of plastic types and different mechanisms for
their decomposition is responsible for this different TGA behavior. These results clearly indicate
the usefulness of TG analysis for the identification of different plastic types in samples collected
from municipal wastes.
TGA and DTG curves for PP, petroleum residue (VR), and VR/PP blend are illustrated in
Figure 3. It can be seen that VR decomposed at lower temperature than PP and decomposition
process is also longer (300ı C–515ıC). VR shows relatively higher weight loss in low temperature
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

oxidation region (< 300ıC). The bulk of VR sample is volatilized at high temperature, leaving
approximately 5% char yield. This lower decomposition temperature and low ash contents of VR
are useful because it should allow VR to act as a solvent media for waste plastics and to intimately
mix during cocracking of VR with plastic wastes. Both TGA and DTG analysis results for PP/VR
blends show that the higher temperatures of initial weight loss for PP and VR are significantly
decreased on mixing and that there is a significant interaction between VR and PP. The results
indicated that the higher temperature of initial weight loss of PP was decreased significantly by
the addition of VR. Also the cracking of PP, VR, and PP/VR mixture has shown a single peak in
their DTG plots.

3.3. Catalytic Pyrolysis of Polypropylene, Vacuum Residue and PP/VR Mix


Thermogravemetric analysis on the performance of all five catalysts in the catalytic pyrolysis of
PP, VR, and PP/ VR mix was carried out using 3, 5, and 8 wt% of each catalyst. TGA runs

FIGURE 3 Thermograms for vacuum residue C polypropylene.


TRANSPORTATION FUELS FROM CATALYTIC CO-PYROLYSIS 1671

at 8 wt% of all five catalysts were repeated to check for repeatability of the TGA data. It was
concluded that the data is repeatable within experimental limits.
All hydrocracking (HC) catalysts (C-2, C-3, C-4, and C-5) were effective in lowering the onset
of degradation. At 3 wt% catalyst, HC catalysts lowered T1% by approximately 8–9%, whereas
titania/alumina catalyst (C-1) lowered T1% by approximately 5%. Similarly, T99% decreased with
increasing catalyst weight fraction for all the catalysts. Thus, the degradation temperature range
is significantly reduced with the addition of the HC catalysts. Furthermore, it appears that the
Z-713 catalyst (C-3) is most effective and FCC/RFCC catalyst (C-5) is least effective.
The temperature at the maximum rate of conversion (Tmax ) and conversion maximum weight
loss (Cmax ) with the addition of the various catalysts and with increasing catalyst weight fractions
are shown in Table 3. However, no significant correlation between Cmax and the catalyst type and
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

amount was observed, although a distinct increase in conversion percent occurred with increasing
catalyst weight fraction. The results in Table 3 show that all catalysts except the FCC/RFCC
catalysts (C-5) enhanced the pyrolysis process. Zeolyst Z-713 catalyst (C-3) was found to be
most effective in reducing the Tmax for conversion reactions. The FCC/RFCC catalysts (C-5)
had almost no effect. Although AKZO 2710 (C-2) is also a Ni/W catalyst, it is less effective
than Z-713 in its catalyzing the reactions. The number of acid sites on a solid catalyst plays a
key role in the catalytic degradation rate of polyolefins. This number increases with increasing
aluminum incorporation into the zeolite crystal. This may explain AKZO 2710’s reduced catalytic
activity (Al/Si D 1.15) in comparison to the Z-713 catalyst (Al/Si D 6.5). Another influence on
the conversion reactions using microporous materials is the catalyst pore size. Z-713 is relatively
large pore zeolite and high acidity, whereas the base-zeolite of HC catalysts is characterized by
medium pores and relatively low acidity. Given that PP molecules/ VR asphaltenes are much
larger than the pore size of other catalysts, the degradation of the primary decomposition products
(large molecules) occurs preferably over the surface of Z-713 catalysts, forming smaller molecules
that can be permitted into the pores of the other HC catalysts for further cracking. Thus, larger
pore of Z-713 will permit for further degradation of polymers within its pore, unlike medium pore

TABLE 3
Temperature and Conversion at Maximum Rate of Degradation
of Polypropylene (PP)

Sample Catalyst (%) Tmax ı C  Tmax (%) Cmax (%)

PP Nil 470 0 55
VR Nil 310 0 94.5
PP C VR (1:1) Nil 395 0 89.0
PP C VR (1:1) C-1 (3%) 350 11.4 88.0
PP C VR (1:1) C-1 (5%) 340 13.9 90.0
PP C VR (1:1) C-1(8%) 342 14.4 92.0
PP C VR (1:1) C-2 (3%) 365 7.6 92.5
PP C VR (1:1) C-2 (5%) 357 9.6 89.5
PP C VR (1:1) C-2 (8%) 355 10.1 89.5
PP C VR (1:1) C-3 (3%) 340 13.9 93.0
PP C VR (1:1) C-3 (5%) 330 16.4 94.0
PP C VR (1:1) C-3 (8%) 325 17.7 94.5
PP C VR (1:1) C-4 (3%) 355 10.1 90.5
PP C VR (1:1) C-4 (5%) 345 12.6 91.2
PP C VR (1:1) C-4 (8%) 343 12.3 91.0
PP C VR (1:1) C-5 (3%) 390 1.3 88.0
PP C VR (1:1) C-5 (5%) 385 2.5 89.0
PP C VR (1:1) C-5 ( 8%) 380 3.8 88.5
1672 M. F. ALI AND M. S. QURESHI

TABLE 4
Activation Energy (Ea) for Catalytic Degradation
of Polypropylene (PP)

Regression
Catalyst,a Ea , for Line Used
Sample % w/w kJ/mol for Ea

PP 0 306.8 0.993
VR 0 101.5 0.995
PP C VR 0 140.5 0.989
PP C VR C-1 (5%) 125.0 0.987
PP C VR C-2 (5%) 118.5 0.972
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

PP C VR C-3 (5%) 130.8 0.992


a
Measured weight fraction.

catalysts. The total acidity of the other HC catalysts was found to have lower values (0.84 mmol/g)
compared with the Z-713 (1.3 mmol/g). The metal loading on the titania-alumina in Catalyst
C-1 was found to increase the acidity of titania-alumina from 0.84 to 1.1 mmol/g. However,
impregnation with the metals (Ni/Mo) of silica-alumina base (Catalysts C-1) were less effective
than the metals Ni/W (catalysts C-2 and C-3) and Ni/Mo (catalyst C-4). This direct influence of
catalyst acidity and pore size on the catalytic degradation of PP was also observed in literature
(Durmus and Kasgöz, 2005). The results also illustrate the Tmax is lowered with increasing catalyst
weight for all the five catalysts employed. The number of acid sites that are available for polymer
cracking increases with the catalyst amount, thereby, enhancing the degradation reactions. From
the plots it is evident that all the HC catalysts enhanced the cocracking of PP with VR.

3.4. Estimation of Activation Energy (Ea) for PP Degradation


The kinetic parameters, activation energy (Ea) and the pre-exponential factor (A), were estimated
for the thermal and catalytic pyrolysis of PP, VR, and PP/ VR mixture by various catalysts.
A comparison of the activities amongst the catalysts was thus performed. The Freeman-Carroll
differential approach (Westerhout et al., 1997) was used to determine kinetic parameters. The
kinetic parameters were estimated for the samples containing 5 wt% of catalysts (C-1 to C-3) and
Ea and regression for line used for Ea calculations were estimated by the Arrhenius Equation.
The results are shown in Table 4. The pyrolysis of virgin PP, VR gave overall activation energy as
306.8 kj/mole and 101.5 kj/mole respectively. The Ea of PP C VR mixture was found to be 140
kj/mole. The activation energies for catalyzed decomposition of PP C VR showed considerable
lowering (values between 112.5 and 120.8 kj/mole).

4. CONCLUSION

Thermogravimetric analysis of the copyrolysis of polypropylene with petroleum vacuum residue


had clearly indicated a synergistic effect in their decompositions. The Tmax for PP was lowered and
Tmax of VR rose from that of individual component. Ea for the overall PP C VR copyrolysis had
also shown a lowering of activation energy than PP alone. The catalyzed copyrolysis reactions of
PP C VR have shown further lowering of activation energy. TG data allowed estimating activation
energy (Ea) and the pre-exponential factor (A) for the catalytic degradation of PP. The activities
of five different catalysts for the decomposition of polypropylene were evaluated by TGA, and
TRANSPORTATION FUELS FROM CATALYTIC CO-PYROLYSIS 1673

DTA analysis. The maximum activity was observed for the more acidic catalyst loaded with
Ni/W metals (zeolyst Z-713). The temperature of PP degradation decreased and the conversion
increased when the Z-713 concentration was increased from 3 to 8 wt%. The RFCC catalyst
(RDC-8) performed poorly for the degradation of PP. The performance of the catalyst prepared
by us (C-1) was also satisfactory.

ACKNOWLEDGEMENT

The authors would like to thank the Higher Education Commission, Govt. of Pakistan for financial
support. The facility support provided by H. E. J. Research Institute of Chemistry, ICCBS, and
Downloaded by [Moskow State Univ Bibliote] at 23:51 09 October 2013

University of Karachi is gratefully acknowledged.

REFERENCES

Ali, M. A., and Asaoka, S. (2009). Ni-Mo-titania-alumina catalysts with USY zeolite for low pressure hydrodesulfurization
and hydrocracking. Pet. Sci. Technol. 27:984–997.
Ali, M. F., and Siddiqui, M. N. (2005). Thermal and Catalytic decomposition behavior of PVC mixed plastic waste with
petroleum residue. J. Anal. Appl. Pyrolysis. 74:282–289.
Ali, M. F., and Siddiqui, M. N. (2006). In Feedstock Recycling and Pyrolysis of Waste plastics. J. Scheirs and W.
Kaminsky, eds. John Wiley & Sons Ltd. 363–380.
Carniti, P., and Gervasini, A. (2001). Thermogravimetric study of the kinetics of degradation of polypropylene with solid
catalysts. Thermochim. Acta 379:51–58.
Durmus, A., and Kasgöz, A. (2005). Thermal-catalytic degradation kinetics of polypropylene over BEA, ZSM-5 and MOR
zeolites. Appl. Catal. B. 61:316–322.
Geraldo, J., and Filho, A. P. (2005). Thermo gravimetric kinetics of polypropylene degradation on ZSM-12 and ZSM-5
catalysts. Catal. Today. 107–108:507–512.
Joo, H. K., and Curtis, C. W. (1996). Catalytic coprocessing of plastics with coal and petroleum resid using NiMo/Al2 O3 .
Energ. & Fuels 10:603–611.
Kim, J. R., Yoon, J. H., and Park, D. W. (2000). Catalytic recycling of the mixture of polypropylene and polystyrene.
Polym. Degrad. Stab. 76:61–67.
Luo, M., and Curtis, C. W. (1996). Thermal and catalytic processing of Illinois No. 6 coal with model and comingled
plastics. Fuel Process. Tech. 49:91–117.
Marcilla, A., Beltran, M., and Conesa, J. A. (2001). Catalyst addition in polyethylene pyrolysis: Thermogravimetric study.
J. Anal. Appl. Pyrolysis 58-59:117–126.
Pinto, F., Costa, P., Gulyurtlu, I., and Cabrita, I. (1999). Pyrolysis of plastic wastes: 2. Effect of catalyst on product yield.
J. Anal. Appl. Pyrolysis 51:57–71.
Serrano, D. P., Aguado, J., and Escola, J. M. (2000). Catalytic conversion of polystyrene over HMCM-41, HZSM-5 and
amorphous SiO2 -Al2 O3 : Comparison with thermal cracking. Appl. Catal. B. 25:181–189.
Westerhout, R., Wanders, J., Kuipers J., and Van Swaaij, W. (1997). Kinetics of the low-temperature pyrolysis of polyethene,
polypropene, and polystyrene modeling, experimental determination and comparison with literature models and data.
Ind. Eng. Chem. Res. 36:1955–1964.

You might also like