You are on page 1of 12

Environmental Technology & Innovation 22 (2021) 101455

Contents lists available at ScienceDirect

Environmental Technology & Innovation


journal homepage: www.elsevier.com/locate/eti

Thermal and catalytic pyrolysis of waste polypropylene plastic


using spent FCC catalyst

Eki Tina Aisien a , Ifechukwude Christopher Otuya b , Felix Aibuedefe Aisien b ,
a
Department of Environmental Management and Toxicology, University of Benin, Benin City, Nigeria
b
Department of Chemical Engineering, University of Benin, Benin City, Nigeria

article info a b s t r a c t

Article history: Waste polypropylene plastic (WPP) is an enormous volume of plastics in the landfill in
Received 26 October 2020 Nigeria. It causes serious environmental problems, such as reduced landfill space and
Received in revised form 22 February 2021 pollution. We made WPP plastic undergo pyrolysis using a batch reactor subjected to
Accepted 24 February 2021
temperatures variation of 300 ◦ C, 350 ◦ C, 375 ◦ C, 400 ◦ C, and the spent fluid catalytic
Available online 13 March 2021
cracking (FCC) catalyst used were 5, 7.5, and 10 wt% catalyst. We heated the reactor at
Keywords: a rate of 15 ◦ C/min. until it reaches the pyrolysis temperature of 400 ◦ C at atmospheric
Polypropylene plastic pressure. We investigated the influence of the FCC catalyst, reaction temperatures, and
Pyrolysis catalyst to plastic ratio. We characterized the pyrolysis liquid oil using density, pour
Fluid catalytic cracking point, API gravity, flash point, viscosity, calorific value, carbon residue, ATSM distillation,
Liquid oil
and GC–MS.
Temperature
The thermal pyrolysis produced maximum liquid oil (83.3 wt%) with gases (13.2 wt%),
and char (3.0 wt%), while the catalytic pyrolysis using 0.1 catalyst to plastic ratio
decreased the liquid oil yield (77.6 wt%), and char (2.7 wt%), with an increase in gases
(19.7 wt%). The GC–MS results of the catalytic pyrolysis of liquid oil showed that the
liquid fractions comprised a wide range of hydrocarbon, mainly distributed within C4
to >C17. The paraffin, olefins, naphthalene, and aromatics yield were 30.83%, 44.6%.
19.44%, and 5.13%, respectively. The liquid oil’s fuel properties were like that of gasoline
and diesel.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction

Plastics of different forms have become an indispensable part of our modern life today. The increasing population and
change in lifestyles of Nigerians have resulted in increasing plastic consumption. Plastics’ production and consumption
rates are high because plastics are lightweight, durable, energy-efficient, inexpensive, and found applications for industrial
and domestic purposes (Miteva et al., 2016). Plastic wastes management is a major environmental problem in Nigeria and
other countries in the world. About 2.5 million tons of plastic wastes are generated per year in Nigeria. The recycling of
these plastic wastes is less than 12%, and about 80% (2 million tons) of these wastes end up in landfills and dumpsites
(Babayemi et al., 2018; Duru et al., 2019; Kehinde et al., 2020).
We divide plastics into two main types, which are thermoplastics and thermosets. The thermoplastics include low-
density polyethylene (LDPE), high-density polyethylene (HDPE), polystyrene (PS), polypropylene (PP), polyvinyl chloride
(PVC), polyethylene terephthalate (PET), polycarbonate (PC), and polymethyl methacrylate (PMM). The thermosets are

∗ Correspondence to: P.M.B. 1154, Ugbowo, Benin City, Nigeria.


E-mail address: aibue.aisien@uniben.edu (F.A. Aisien).

https://doi.org/10.1016/j.eti.2021.101455
2352-1864/© 2021 Elsevier B.V. All rights reserved.
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

polyurethane, polyester, vinyl ester, silicon, and melamine resin (Kyaw and Hmwe, 2015). The consumption of thermo-
plastics is higher than thermosets because we recycle the latter. The average world consumption of plastics is about 35%
HDPE, 23% PP, 10% PS, 13%. PVC, 7% PET, and 12% various polymers. The polyolefins (PE, PP) and PS are the most prevailing
plastics in the waste polymers streams (Kyaw and Hmwe, 2015).
Gaurh and Pramanik (2018) reported that about 24.3% of polypropylene was found in municipal solid waste (MSW).
Polypropylene is a versatile polymer because of its good chemical resistance, mechanical properties, processing, and low
density. Polypropylene is used in food packaging, medical bags, computer components, automotive industry, pipes, fur-
niture, office folders, storage boxes, and general containers (Abbas-Abadi et al., 2014). Plastics are made of petrochemical
hydrocarbons with additives such as flame-retardants, stabilizers, and oxidants. Hence, plastics are non-biodegradable
and will remain in the environment for several thousands of years (Ma et al., 2017). As a result, waste plastics are present
in tremendous amounts on landfills, terrestrial and marine environments with no noticeable degradation (Kyrikou and
Brissoulis, 2007). Hence, the current challenge is that this plastic solid waste (PSW) is now a significant component of
municipal solid waste.
There are several waste management issues and environmental problems from improper waste plastic disposal. Aisien
et al. (2013) reported that this includes water pollution, blockage of side-drain, reduced landfill space, and increased
breeding grounds for insects, rodents, and dangerous reptiles. These problems can be eliminated or minimized through
efficient plastic waste management such as recycling and energy recovery techniques. However, the recycling technique
has some disadvantages, including labor intensive related to the sorting process and water contamination. Also, process
handling is less comfortable and more complex than pyrolysis (Thahir et al., 2019). These drawbacks and energy deficits
in Nigeria today have caused us to research energy recovery techniques.
Plastic waste is one of the most promising resources for fuel production. This is because of its high calorific value,
comparable with gasoline and diesel, and its increasing availability in the environment. We can convert plastic waste
into valuable energy resources. This is mainly through various thermal treatment processes such as the plasma process,
pyrolysis, gasification, and incineration (Moustakas and Loizidou, 2010). However, pyrolysis is one of the environmentally
friendly methods and is the most promising process. The pyrolysis process results in a significant reduction of waste
volume, lower decomposition temperature, low capital cost, and more recoverable energy by producing varieties of
products (Patni et al., 2013). Therefore, pyrolysis is a sustainable management process. Plastic waste has liquid oil as
an energy source, solid char, and gases as value-added products when it undergoes pyrolysis (Fivga and Dimitriou, 2018).
Pyrolysis involves thermal degradation of complex molecules or large chain hydrocarbons into smaller molecules or
shorter chain hydrocarbons at high temperatures (300–800 ◦ C) in the absence of oxygen (Aguado et al., 2007). The process
is divided into thermal pyrolysis and catalytic pyrolysis. The catalytic pyrolysis makes use of catalysts for the thermal
degradation process. Catalysts are widely used in the pyrolysis process to optimize product distribution and increase
product selectivity. Catalysts have also been used in the upgrading of pyrolysis products and to improve the hydrocarbon
distribution. Also, they produce similar liquid oil properties to that of conventional fuels such as diesel and gasoline
(Sharuddin et al., 2016). Many catalysts have been used for the pyrolysis of waste plastics in the past decade. The catalyst
includes red mud, kaolin, SiO2 -Al2 O3 , CuCO3 , ZnO, SiO2 , MgO, Al2 O3 , ZSM-5 Zeolite, CaCO3 , and FCC (Shah et al., 2010;
Lopez et al., 2011; Kumar et al., 2013; Abbas-Abadi et al., 2014; Singh et al., 2018).
We can classify the catalysts used in plastics’ pyrolysis into three types: silica-alumina, zeolite, and FCC catalysts.
Several studies have been carried out on the pyrolysis of waste plastics. Miandad et al. (2016a,b) and Hernandez et al.
(2007) studied the catalytic pyrolysis of plastic waste using ZSM-5, red mud, zeolites Y, FCC others. They reported that
the produced pyrolysis oils were like that of conventional diesel fuel. Moorthy Rajendran et al. (2020) studied the use
of catalysts such as HUSY, HZSM-5, HMOR, Zeolite Y, silica, and FCC for catalytic pyrolysis of municipal plastic waste
into quality fuels. Their results showed that catalysts improve the selectivity; besides that, mild acid catalysts produce
more liquid hydrocarbons. Susastriawan et al. (2020) studied the catalytic pyrolysis of polyethylene using zeolites. They
reported that the smaller the zeolite size and the higher the temperature, the higher the liquid fraction yield. Similarly,
Onwudili et al. (2019) studied catalytic pyrolysis of a mixture of plastics using catalysts such as FCC, ZSM-5, and zeolites
Y. They stated that the yield of the liquid fraction decreased with the addition of the catalyst. The liquid fraction had
properties suitable for fuel, although the number of aromatic compounds present increased.
Santos et al. (2018) studied the catalytic pyrolysis of polyethylene and polypropylene waste plastic using catalysts
such as HZSM-5, USY, NH4ZSM5. They reported that zeolite USY gave a higher amount of liquid fraction, and the major
components were alkylbenzenes, naphthalenes, and olefins. It was also reported that plastic waste catalytic pyrolysis
using catalysts such as ZSM-5, Zeolites Y, Zeolites β , Ca(OH)2 , and others gave yields of the liquid fraction varying from
15 to 93% (Kunwar et al., 2016). They stated that it would be necessary to use high acidity and porosity catalyst and
hydrogenation. López et al. (2017) studied the pyrolysis of polyolefins using different operating conditions and catalysts.
They reported that the HZSM-5 catalyst was suitable for the production of valuable light olefins. Besides, larger pore size
zeolites such as HY, HUSY, or spent FCC catalysts are better alternatives for liquid hydrocarbons production. Similarly,
MCM-41, or the less acidic mesoporous SiO2 -Al2 O3 , are exciting options to produce liquid fuels.
Some researchers have reported studies on the thermal and catalytic pyrolysis of PP. Ahmad et al. (2014) studied the
effect of temperature on the pyrolysis of PP. They said that the highest liquid oil yield was 69.82 wt% at a temperature
of 300 ◦ C. Also, as the temperature increased to 400 ◦ C, the solid residue increased from 1.34 to 5.7 wt%. Sakata et al.
(2013) reported a higher liquid oil yield of 80.1 wt%, gaseous 6.6 wt%, and solid residue 13.3 wt% in the pyrolysis of PP at
2
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

380 ◦ C. Fakhrhoseini and Dastanian (2013) and Abbas-Abadi et al. (2014) increased the pyrolysis temperature to 500 ◦ C,
and they reported 82.12 wt% liquid oil yield. However, Abbas-Abadi et al. (2014) studied the pyrolysis of PP between 420
and 510 ◦ C. They observed that the highest liquid oil yield was 92.3 wt% at 450 ◦ C. Besides, Demirbas (2004) studied the
pyrolysis of PP at 760 ◦ C. He reported a lower liquid oil yield of 48.8 wt% and solid residue 1.6 wt% with a very high
gaseous yield of 49.6 wt%. Jung et al. (2010) reported that PP degradation started at a very low temperature (240 ◦ C)
compared to other feedstocks. Kyong et al. (2003) compared the liquid product yield from PP pyrolysis and PE (LDPE and
HDPE) and PS. They reported that the liquid product yield was in the order of PS > PP > PE.
The FCC catalyst is a combination of zeolite crystals and a non-zeolite acid matrix known as silica-alumina with the
binder (Degnan Jr, 2000). The FCC catalyst that has been used in various plastics pyrolysis is the waste or spent FCC
catalyst. We usually get this from the commercial FCC process in petroleum refineries. The petroleum refinery process
generates a large quantity of spent FCC catalyst yearly. The untreated spent FCC catalyst is disposed of indiscriminately
on land. It occupies valuable land resources, causes severe pollution, and poses an enormous threat to human health (Liu
et al., 2015).
The previous studies on spent FCC catalysts have proven that they could be a suitable catalyst for the pyrolysis of
waste plastics. These studies investigated high temperature and high catalyst to plastic ratio or low temperature and
high catalyst to plastic ratio. Gaurh and Pramanik (2018) reported that PP in the waste streams is high. The liquid oil
yield from PP is more increased than PE, and the primary yield from PS pyrolysis is a waxy product (Kyong et al., 2003).
Besides, Jung et al. (2010) reported that PP degrades at low temperatures than other plastics. Therefore, this research
has investigated waste PP plastic pyrolysis under low temperature and low catalyst to plastic ratio conditions. Hence, we
focused our research on applying spent FCC catalysts on the pyrolysis of waste PP plastic. This will address the serious
environmental problems caused by the indiscriminate disposal of spent FCC catalysts and waste PP plastics. We aim to
investigate thermal and catalytic pyrolysis of waste PP plastics using spent FCC catalysts. The objectives compare thermal
and catalytic pyrolysis of PP at low temperature and low catalyst to plastic ratio. Also, comparing the properties of the
produced liquid oil with local commercial fuels and determining the liquid oil product’s composition.

2. Materials and methods

2.1. Materials collection

We collected the waste PP plastic from Uyi Egharevba landfill at Evbuobanosa, Iduadolor-Benin, Nigeria. We got the
spent FCC catalyst from Warri refinery and petrochemical Company, Warri-Nigeria. We got the insulated jacketed 5 kg
capacity batch pyrolysis reactor, which was earlier designed and fabricated in the Engineering Workshop, University of
Benin, Benin City, from the workshop. We collected analytical grade chemicals from the Chemical Engineering Laboratory,
University of Benin, Benin City.

2.2. Preparation of plastic materials

We washed the waste PP plastics with detergent and use tap water to clean them. We dried the cleaned waste plastics
with sunlight. We then use scissors to cut them into small pieces of 2 to 3 inches and ground them with a grinding
machine. We stored the ground waste PP plastics in a covered plastic container.

2.3. Experimental set-up

The pyrolysis reactor system was made up of an insulated jacketed 5 kg capacity batch reactor. There was a stainless-
steel ampule vessel housed in a vertical furnace made of ceramic clay lined with heating coils. We heated the reactor
with a 3-phase system connected to a K-type thermocouple and an external PID controller for temperature regulation.
We sealed the top of the reactor with the aid of grafoil-gasket flange joints. We connected the reactor outlet tube to the
ice-cold water glass condenser for condensing the pyrolysis condensable gaseous products.

2.4. The pyrolysis studies

We used an insulated 5 kg capacity batch reactor for both the catalytic and thermal pyrolysis processes. We then
charged a mixture of 1 kg of prepared waste PP plastics and 50 g spent FCC catalysts into the reactor’s bottom vessel.
Hence, the catalyst to plastic mass ratio was 0.05. We first evacuated the stain steel ampule with a vacuum pump’s
aid and inserted it inside the furnace heating chamber. We heated the reactor system from 30 ◦ C to the desired pyrolysis
temperature of 300 ◦ C at a heating rate of 15 ◦ C/min. We regulated the reaction temperature with K-thermocouple and PID
controller. The resulting gaseous products pass through the outlet tube connected to the ice-cold water glass condenser.
This ensured that we keep the condensation temperature below 10 ◦ C to ensure maximum vapor condensation to liquid
oil.
We collected the condensable liquid products with a container and weighed them. Also, we weighed the solid residue
left inside the reactor at the end of the pyrolysis process. We then performed four additional trials for this study. We
3
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

tested the physical properties of the produced pyrolysis liquid oil using the standard methods of analysis. We carried out
the product yield evaluation using Eqs. (1) to (3). We used a similar procedure for subsequent catalytic pyrolysis studies.
However, we varied the temperatures (350 ◦ C, 375 ◦ C, and 400 ◦ C) and catalyst to plastic mass ratio (0.05, 0.075, and
0.10) in other studies. Also, we carried out the control study, which is thermal pyrolysis with a similar procedure without
a catalyst.

2.5. Analytical methods

We analyzed the physical properties of the produced pyrolysis liquid oil using standard methods of analysis. We used
the bomb calorimeter for calorific value determination. Also, we determined the distillation profile by ASTM D86, density
by ASTM D1298 viscosity at 40 ◦ C, and flash point by ASTM D93. Besides, the pour point by ASTM D79, ash content
by ASTM D1500, and carbon residue by ASTM D4530. We then analyzed the pyrolysis liquid oil fraction’s chemical
composition and properties using gas chromatography with mass spectrometry (GC–MS) and ASTM distillation. We
calculated the pyrolysis products based on the weight/mass after we weighed each product to complete each study.

2.6. Statistical analysis

We used the mean results of each study’s parameters to test our data, hence n = 5. The Kruskai–Walli H-test or One-
Way Analysis of Variance by ranks using Statistical Package for Social Sciences (SPSS) version 20 with a significance level
of p = 0.05 was used to compare the difference in product yield. We used descriptive analysis of data.
We used the Mabood et al. (2010) formula given in Eqs. (1)–(3) below to calculate the products’ yield as a mass
percentage of the products.
Liquid Yield:
Mass of oil
Oil (wt.%) = x 100% (1)
Mass of polypropylene (PP)
Residue (Char) Yield:
Mass of residue
Char (wt.%) = x 100% (2)
Mass of polypropylene (PP)
Gas Yield:
Gas(wt.%) = 100% − (Oil + Residue) (3)

3. Results and discussion

3.1. The effect of temperature on the pyrolysis of PP plastic

We studied the effect of temperature on the pyrolysis products (liquid oil, gases, and char) distribution at different
catalyst to PP plastic ratios. Figs. 1 and 2 show the effect of temperature on the pyrolysis products at 0.05 and 0.1 catalyst
to plastic ratios, respectively. In Figs. 1 and 2, T represents thermal pyrolysis, and no T represents catalytic pyrolysis. It
can be observed that with the increase in temperature from 300 to 400 ◦ C at 0.05 catalyst to plastic ratio, the liquid oil
yield increased from 72.4 to 83.2 wt%. The char yield increased from 2.6 to 3.8 wt%, and the gases yield decreased from 25
to 13.9 wt%. The results showed a significant difference (p<0.05) in the liquid oil and gas yields between 300 and 400 ◦ C.
This is because the computed value is higher than the tabulated value at a 0.05 level of significance. However, there is a
not significant difference (p<0.05) in the char yield between 300 and 400 ◦ C. This is because the computed value is lower
than the tabulated value at a 0.05 level of significance. Also, we observed a similar trend at a 0.1 catalyst to plastic ratio.
Abbas-Abadi et al. (2012a) observed similar tread. However, some researchers reported a decrease in overall liquid oil
yield with increased gasses and char production when PP plastics pyrolysis were carried out between 450 and 746 ◦ C.
(Jung et al., 2010; Demirbas, 2004; Lopez et al., 2012; Syamsiro et al., 2014a; Abbas-Abadi et al., 2014). Ribeiro et al. (2013)
and Marcilla et al. (2009) reported that the FCC catalyst has strong acidity and could effectively catalyze the pyrolysis of
PP. This is because of its sufficient cracking ability and excellent contact between PP polymer and catalyst particles.
It can be seen from Figs. 1 and 2 that the maximum temperatures for the optimization of liquid oil, gases, and char
yields were 400 ◦ C, 300 ◦ C, and 400 ◦ C, respectively. These optimum temperatures were independent of the catalyst to
plastic ratio. However, as the catalyst to plastic ratio increase, the yield of the liquid oil and char increased, while that of
the gases decreased. Hence, the optimum conditions for the highest yield of liquid oil (83.2 wt%) and char (3.8 wt%) are
400 ◦ C and 0.05 catalyst to plastic ratio. As for the highest gases yield, 300 ◦ C and 0.1 catalyst to plastic ratio were the
optimum conditions.
Figs. 1 and 2 show that thermal pyrolysis gave higher liquid oil, gases, and char yields than catalytic pyrolysis. The
results showed a significant difference (p<0.05) in the liquid oil and gases yields between thermal and catalytic pyrolysis
of waste PP at a 0.1 catalyst to PP ratio. However, there is a no significant difference (p<0.05) in the liquid oil and gases
yields between thermal and catalytic pyrolysis of waste PP at 0.05 catalyst to PP ratio. Also, there is a no significant
4
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Fig. 1. Comparing the variation of pyrolysis products yield with temperature for thermal and catalytic pyrolysis at 0.05 FCC to PP ratio.

Fig. 2. Comparing the variation of pyrolysis products yield with temperature for thermal and catalytic pyrolysis at 0.1 FCC to PP ratio.

difference (p<0.05) in the char yield between/within thermal and catalytic pyrolysis. This is because PP’s branching
structure makes it easily degradable (Hakki Metecan et al., 2005). Also, the higher proportion of tertiary carbons present
in the polypropylene chains promotes C–C bonds’ thermal cleavage (Aguado et al., 2000). Besides, Jung et al. (2010) stated
that the tertiary carbon in PP encourages carbocation formation during its thermal degradation process. They explained
that this might probably be the reason for achieving maximum PP degradation at even lower temperatures without a
catalyst.
It can be observed from Table 1 that the average residence time was 65 min and 33 min for thermal and catalytic
pyrolysis, respectively. This shows that the spent FCC catalyst is very effective in accelerating the PP pyrolysis reaction
rate, and there is a significant difference (p<0.05) in the reduction in residence time between thermal and catalytic
pyrolysis of waste PP. These results agree with that from other studies (Abbas-Abadi et al., 2012b; Jung et al., 2010;
Abbas-Abadi et al., 2014; Ahmad et al., 2014; Sakata et al., 1999; Costa et al., 2010; Fakhrhoseini and Dastanian, 2013;
Demirbas, 2004).

3.2. The effect of catalyst to plastic ratio on the pyrolysis of PP plastic

Fig. 3 shows the effect of catalyst to plastic ratio on the pyrolysis products distribution waste PP at different
temperatures. In Fig. 3, A, B, C and D represent 300 ◦ C, 350 ◦ C, 375 ◦ C and 400 ◦ C respectively. It can be observed from
Fig. 3 that as the catalyst to plastic ratio increased from 0.05 to 0.1 at 300 ◦ C (A), the liquid oil yield decreases from 72.4
to 64.7 wt%. The char yield decreased from 2.6 to 2.2 wt%, and the gases yield increased from 25.5 to 33.1 wt%. The results
showed no significant difference (p<0.05) in the product yields between 300 and 400 ◦ C.
Miandad et al. (2016a,b) and Kunwar et al. (2016) reported that catalyst usually causes an increase in the gaseous
fraction yield. Besides, a decrease in liquid fraction and an increase in the amount of char. According to Akpanudoh et al.
(2005) and Aboulkas et al. (2010), catalysts’ highly acidic behavior can lead to further cracking of the liquid hydrocarbons
in the reaction chamber. This, they said, could cause more gases and coke yields and reduce liquid yield. Also, Kim et al.
5
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Fig. 3. The variation of pyrolysis products yield with catalyst to plastic ratio at different temperature.

(2002) reported that catalysts with low/high acidity and BET surface areas with microporous structures favor PP’s initial
degradation. This, they said, might lead to the maximum production of gases. Obali et al. (2012) carried out PP’s pyrolysis
with an alumina-loaded catalyst and reported maximum production of gases. The formation of carbocation during PP
degradation is because of tertiary carbon in its carbon chain, favoring gas production (Jung et al., 2010). Syamsiro et al.
(2014a) studied PP and PS’s catalytic pyrolysis with an acid (HCl) activated natural zeolite catalyst. They reported the
production of more gases than liquid using thermal activated natural zeolite catalyst. This, they said, was because of its
high acidity and BET surface area.
We observed a similar trend in other pyrolysis temperatures at 350 ◦ C (B), 375 ◦ C (C), and 400 ◦ C (D), as shown in Fig. 3.
Also, it can be seen that there is a no significant difference (p<0.05) in the product yields between 0.05 and 0.1 catalyst
to plastic ratio. Besides, we observed that the effect of the spent FCC on PP plastic pyrolysis in terms of residence time is
a no significant difference (p<0.05), as shown in Table 1. However, the results showed a significant difference (p<0.05)
between thermal and catalytic pyrolysis in residence time. Hence, the low residence time. Budsaereechai et al. (2019)
stated that an increase in the amount of catalyst did not significantly affect the yield and fuel properties of produced
liquid oil. We got similar results in our study. Hence, it seems unnecessary to use high catalyst loading. However, we
observed that these results agree with other results reported by Abbas-Abadi et al. (2014), Kyong et al. (2002, 2003),
Sharuddin et al. (2016). They said a similar effect of spent FCC catalyst on the yield of the pyrolysis products. Therefore,
we can state that the FCC catalyst’s use enhanced the ability to decompose or crack the PP plastics. Also, it increases the
non-condensable gas (gases) yield and ultimately produced more gases. Luo et al. (2000) and Abbas-Abadi et al. (2014)
stated that the FCC catalyst is one of the best catalysts for optimizing liquid oil production from plastic pyrolysis. They
reported a high liquid yield above 90 wt% for PP pyrolysis. Also, Sharuddin et al. (2016) reported that the spent FCC
catalyst still has high catalytic performance with liquid oil yield above 80 wt% for all plastics.

3.3. Effect of FCC catalyst on the products yield and residence time

Table 1 shows the product yield distribution and the residence time for thermal and catalytic pyrolysis of PP plastic. We
observed that the liquid oil yield from thermal pyrolysis was higher than catalytic pyrolysis. Hakki Metecan et al. (2005)
explained that it was because of PP’s branching structure, which makes it easily degradable. Also, the higher proportion
of tertiary carbons present in the polypropylene chains promotes C–C bonds’ thermal cleavage (Aguado et al., 2000; Jung
et al., 2010). In the catalytic pyrolysis, the liquid oil and char yields decrease with an increased catalyst to plastic ratio.
There is a significant difference (p<0.05) between thermal and catalytic pyrolysis in the residence time. The decrease
in residence time for catalytic pyrolysis is almost half of that for thermal pyrolysis. However, there is a no significant
difference (p<0.05) in residence time between 0.05 and 0.1 catalyst to plastic ratio. This shows that the FCC catalyst only
reduces the residence time for the pyrolysis of PP. These results were in agreement with that reported by Altohami and
Mustafa (2018). However, Marcilla et al. (2005) and Abbas-Abadi et al. (2012b) reported that inadequate heat transfer
could decrease the catalyst efficiency. This they said to increase the undesirable products (char and gaseous products).

3.4. The characteristics of PP plastic pyrolysis liquid oil

3.4.1. Physical properties


We characterized the liquid oil produced from the catalytic pyrolysis of waste PP plastic. Hence, we determined the
viscosity, flash point, API gravity, density, pour point, ash content, and the calorific value. We then compared the values of
6
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Table 1
Products yield distribution and retention time for thermal and catalytic pyrolysis of PP plastic at 300 ◦ C using a different
catalyst to plastic ratio.
Catalyst: plastic ratio Liquid oil (wt.%) Gases (wt.%) Char (wt.%) Residence time (min.)
0 74.0 23.7 2.3 65
0.05 72.4 25.0 2.6 35
0.075 69.3 28.3 2.4 33
0.1 64.7 33.1 2.2 3.2

Table 2
The comparison of the physical parameters of produced liquid oil at different temperatures with standard commercial fuel.
Physical properties Temperature (o C) @ 0.1 catalyst : PP ratio Commercial fuel standard (Ahmad et al., 2014)
300 ◦ C 350 ◦ C 375 ◦ C 400 ◦ C Gasoline Diesel Kerosene
Viscosity @ 40 ◦ C (cSt) 1.899 1.964 2.2052 2.2732 0.775–1.03 2.0–5.3 0.9–2.2
Density (g/cm3 ) 0.761 0.75 0.78 0.752 0.72–0.736 0.83–0.85 0.78–0.82
Flash point (o C) 29 28.4 28 29 20.8–42 55–60 50–55
API gravity @ 40 ◦ F 34 35 36 35 62.34–65.03 34.97–38.98 41.06–49.91
Pour point (o C) −10 −10 −9 −11 – 6 -
Ash (wt%) 0.039 0.037 0.042 0.044 – 0.01 –
Carbon residue (wt.%) 0.56 0.54 0.6 0.6
Calorific residue (MJ/kg) 43.435 43.561 44.318 43.286 46.86 42.51 45.46

the pyrolysis liquid oil’s physical properties with that of commercial fuel (gasoline, diesel, and kerosene). We characterized
liquid oil to find out its improved quality and suitability for energy generation and heating purposes.
Table 2 shows the physical properties of pyrolysis liquid oil at different temperatures in comparison with commercial
fuel. It can be observed that there is a no significant difference (p<0.05) in the physical properties between 300 and
400 ◦ C. It was also observed that the values of all the physical properties determined could be compared either with
kerosene, gasoline, or diesel.
We found that the density and API gravity to be 0.7617 g/cm 3 and 35, respectively. This is comparable to the range
of 0.72 to 0.85 g/cm 3 and 34.97 to 65.03 for gasoline, diesel, and kerosene. The viscosity was 2.085 cSt, which is close to
that for diesel and kerosene.
We found it to be 28.6 ◦ C, which is comparable to that of gasoline. Shah and Jan (2014) explained that the reason is
that liquid oil comprises some aromatic hydrocarbons.
We found the calorific value to be 43.65 MJ/kg. This result is within the range reported for conventional gasoline,
diesel, and kerosene (Shakirullah et al., 2010; Ahmad et al., 2014). Saptoadi and Pratama (2015) said that a slightly lower
calorific value of pyrolysis oil could be suitably utilized as a fuel or after mixing with kerosene oil. Rehan et al. (2016)
stated that the energy content is one of the essential characteristics of any fuel for its applications. The pour point of the
liquid pyrolysis oil was −10 ◦ C. This is lower than the commercial diesel of 6 ◦ C. This shows aromatic compounds (Shah
and Jan, 2014). It also relates to the lower calorific value of PP plastic compared with commercial diesel and gasoline.
The average ash content was 0.04 wt%, which was close to that of commercial diesel. Santaweesuk and Janyalertadun
(2017) reported that a small ash content value shows the liquid oil is free from metal contamination and high molecular
weight soot.
The carbon residue of the pyrolysis liquid oil was 0.575 wt%. This shows that the liquid oil has a very low aromatic
content, hence reducing carbon to deposition on the diesel engine.
Besides, we observed no significant difference (p<0.05) in the physical properties between 0.05 and 0.1 catalytic
to plastic ratio. Therefore, we concluded that the liquid oil’s physical properties from waste PP were independent of
temperature or catalyst to plastic ratio. These results agree with those of some researchers (Gaurh and Pramanik, 2018;
Bozbas, 2008).

3.4.2. Chemical properties


Fig. 4 shows the variation of ASTM temperature characteristics with volume vaporized pyrolysis liquid oil and standard
fuel. The ASTM temperature characteristics of hydrocarbons are essential in terms of the safety and performance of the
fuel. We compared the produced pyrolysis liquid oil with standard fuel (kerosene, diesel, and gasoline). The results show
that the boiling point range of the pyrolysis liquid oil at 300 ◦ C lies between diesel and gasoline for 0%–30% recovery
of distillate (Fig. 4). This further affirms its suitability for energy production. These results agree with that of Gaurh and
Pramanik (2018), in which they characterized pyrolysis liquid oil from polyethylene.
We investigated the distribution of hydrocarbons with different carbon chain lengths. The optimum pyrolysis condi-
tions were 300 ◦ C and 0.1 catalyst to plastic ratio. We observed that the liquid oil from the catalytic pyrolysis has a low
viscosity. This shows enhanced breakdown and cracking of PP plastic into low molecular weight liquid hydrocarbons.
7
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Fig. 4. Variation ASTM temperature characteristics of pyrolysis and standard fuel with volume percent vaporized.

Fig. 5. GC–MS analysis showing various effects of catalytic pyrolysis on liquid oil from PP plastics.

Miandad et al. (2016a,b) and Yuliansyah et al. (2015) reported that produced liquid oil composition is affected by
types of feedstock, catalyst, and process parameters. Besides, inadequate heat transfer can affect PP’s pyrolysis and the
molecular weight distribution of the products (Marcilla et al., 2005). These agree with GC–MS results in Table 3 and
Figs. 5 and 6, which show more significant proportions of low molecular weight hydrocarbon fractions of C4 - C12 . The
results also showed that the pyrolysis liquid oil is a complex mixture of hydrocarbons with carbon atoms ranging from
C4 to more significant than C17 . Hence, the liquid oil from the catalytic pyrolysis of PP comprises distinct classes such as
paraffin (alkanes), olefins (alkene), naphthenes (cycloalkane), and aromatics. As shown in Table 3 and Fig. 5, the results
are paraffin (30.83%), olefins (44.60%), naphthenes (19.44%), and aromatics (5.13%). This report is in agreement with that
of Jung et al. (2010).
Table 3 and Fig. 5 show that the GC–MS characterization of the produced liquid oil has different hydrocarbon ranges.
These are C4 -C9 , C10 -C13 , C14 -C17 , and > C17 . The results showed that C4 -C9 , C10 -C13 , C14 -C17 , and > C17 were 20.41%,
24.49%, 36.73%, and 18.37%, respectively. Also, we got liquid fraction products range as gasoline range (C4 -C12 ), diesel
range (C12 -C23 ), and kerosene range (C10 -C18 ). The values are 38.78%, 65.31%, and 71.43% for gasoline, diesel, and kerosene,
respectively. The order is kerosene > diesel > gasoline.
The GC–MS results in Table 3 and Fig. 5 show that there are many hydrocarbon compounds. The major compounds
include Benzene, 3-Octadene, 6-Tridecene, Cyclopropane, 1-methyl-1-(2-methyl), Toluene, 8-Heptadecene, 3-Eicosene, 6-
Tridecene and Cyclooctane, 1,4-dimethyl- cis. These findings agreed with other studies on the chemical composition of
produced liquid oil from plastics (Abbas-Abadi et al., 2014; Sharuddin et al., 2016; Kyong et al., 2003, 2002; Sarker and
Rashid, 2013; Marcilla et al., 2004). However, some researchers reported other major compounds such as Xylene, Biphenyl,
3-methyl pentene, Octadecene, Tetracosane, 3-Undecene, 1,3 Pentadiene, 2,4-dimethyl-, 2-Butanone. 3-methyl-1-phenyl-
, Heptane, 2,4-dimethyl-, 2, 4, dimethyl -1- heptane, 2,4 dimethyl cyclooctane, 1-Noandecene (Sharuddin et al., 2016;
Miandad et al., 2019; Panda, 2018; Al-Salem et al., 2017). This difference was explained by Miandad et al. (2016a,b). They
reported that plastic types are one determinant of liquid oil’s chemical composition from the pyrolysis of plastics. Others
include various process conditions such as temperature, reaction time, type of catalyst, and amount of catalyst used.
8
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Table 3
GC–MS composition of pyrolysis oil from polypropylene plastic.
S/N Retention Area (%) Compound Molecular
time (min) formula
1 4.311 13.253 1-Propene, 2-methyl C4 H8
2 4.311 13.253 2-Pentene C5 H10
3 4.311 13.253 Cyclopropane, 1,2-dimethyl- cis C5 H10
4 13.253 5.165 2,4-Hexadene C6 H10
5 13.253 5.165 1-Pentene, 3,4-dimethyl C7 H14
6 13.253 5.165 Ethylbenzene C7 H14
7 6.051 28.803 1-Hexane, 2-methyl C7 H14
8 6.051 28.803 Toluene C7 H8
9 6.051 28.803 Benzene, (1-methyl-ethyl) C7 H8
10 6.379 17.230 2-Hexene, 3,5-dimethyl C8 H16
11 6.379 17.230 Naphthalene C10 H8
12 6.379 17.230 Benzene, 2-propenyl- C9 H10
13 6.541 4.150 Nonane, 2,6-dimethyl- C9 H10
14 6.541 4.150 1,2-diethyl cyclooctane C10 H20
15 6.541 4.150 Decane, 4-methylene- C11 H22
16 8.865 3.680 7-methyl-1-undecene C12 H24
17 8.865 3.680 Cis-6-tridecene C13 H26
18 9.117 4.300 Trans-4-tetradecene C14 H28
19 9.117 4.300 Trans-6-tridecene C13 H26
20 9.117 4.300 5-butyl-4-nonene C13 H26
21 9.245 2.340 Trans-3-tetradecene C14 H28
22 9.245 2.340 1-Tetradecene C14 H28
23 9.245 2.340 Trans-6-tetradecene C14 H28
24 9.374 2.730 Trans-5-tetradecene C14 H28
25 9.374 2.730 Cyclobutane,2-(1,1-dimethyethyl) C8 H14 O
26 9.374 2.730 11-methyldodecanol C13 H28 O
27 11.412 1.770 Cetene C16 H32
28 11.412 1.770 2-methyl-cis-7-hexadecene C16 H32
29 11.412 1.770 8-heptadecene C17 H34
30 11.693 3.190 1-methyl-1-(2- methylpropyl)-2-
31 11.693 3.190 nonylcyclopropane C17 H34
32 11.693 3.190 Benzene, 1,1’-(2-butene-1,4-diyl)bis- C16 H16
33 11.693 3.190 Cis-1,4-dimethyl cyclooctane C10 H20
34 11.793 3.340 Anthracene C14 H10
35 11.793 3.470 Trans-3-octadecene C18 H36
36 11.793 3.470 8-heptadecene C17 H34
37 11.793 3.470 1-octadecene C18 H36
38 11.793 3.470 Diphenylbenzene C18 H14
39 12.078 1.780 11-methyldodecanol C13 H28 O
40 12.078 1.780 9-Eicosene C20 H40
41 13.998 1.220 Trans-3-Eicosene C20 H40
42 13.998 1.220 2-hexyl-1-decanol C16 H34 O
43 13.998 1.220 2-methyl-1-hexadecanol C17 H36 O
44 13.998 1.220 Carbonic acid, but-2-yn-1-yl tetradecyl ester C19 H34 O3

4. Conclusions

In conclusion, we have proven:

1. Temperature and catalyst to plastic ratio significantly impact the pyrolysis of PP plastic in terms of the product
composition of the liquid oil, gases, and char.
2. The operating temperature required depends strongly on product preference.
3. The thermal pyrolysis produced maximum liquid oil (83.3 wt%) with gases (13.2 wt%) and char (3.0 wt%).
4. The catalytic pyrolysis using 0.1 catalyst to plastic ratio decreased the liquid oil yield (77.6 wt%) and char (2.7 wt%),
with an increase in gases (19.7 wt%).
5. The GC–MS results of the catalytic pyrolysis of liquid oil showed that the liquid fractions comprised a wide range
of hydrocarbon mainly distributed within C4 to >C17 . The paraffin, olefins, naphthalene, and aromatics yield were
30.83%, 44.6%. 19.44%, and 5.13%, respectively.
6. The pyrolysis liquid oil has good fuel properties such as calorific value (43.56 MJ/kg), viscosity (2.2cP), density (0.736
g/cm 3 ), and flash point (28.6 ◦ C).
7. The composition and quality of the pyrolysis liquid oil were like that of gasoline and diesel.

9
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Fig. 6. GC–MS of the pyrolysis oil from polypropylene at optimize condition of 300 ◦ C at 0.1 FCC catalyst to PP ratio.

Funding

This research did not receive any specific grant from funding agencies in public, commercial, or not-for-profit sectors.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Acknowledgment

The authors wish to gratefully acknowledge Gabriel Okoduwa for his kind provision to design and fabricate batch
pyrolysis reactor.

Appendix A. Supplementary data

Supplementary material related to this article can be found online at https://doi.org/10.1016/j.eti.2021.101455.

References

Abbas-Abadi, M.S., Haghighi, M.N., Yeganeh, H., 2012a. The effect of temperature, catalyst, different carrier gases, and stirrer on the produced
transportation hydrocarbons of LLDPE degradation in a stirred reactor. J. Anal. Appl. Pyrolysis 95, 198–204.
Abbas-Abadi, M.S., Haghighi, M.N., Yeganeh, H., 2012b. Evaluation of pyrolysis product of virgin high-density polyethylene degradation using different
process parameters in a stirred reactor. Fuel Process. Technol. 109, 90–95.
Abbas-Abadi, M.S., Haghighi, M.N., Yeganeh, H., McDonald, A.G., 2014. Evaluation of pyrolysis process parameters on polypropylene degradation
products. J. Anal. Appl. Pyrolysis 109, 272–277.
Aboulkas, A., El Harfi, K., El Bouadili, A., 2010. Thermal degradation behaviors of polyethylene and polypropylene. Part 1: Pyrolysis, kinetics, and
mechanisms. Energy Convers. Manage. 51, 1363–1369.
Aguado, J., Serrano, D.P., Garagorri, F., Fernández, J.A., 2000. Catalytic conversion of polyolefins into fuels over zeolite beta. Polym. Degrad. Stab. 69,
11–16.
Aguado, J., Serrano, D.P., Miguel, G.S., Castro, M.C., Madrid, S., 2007. Feedstock recycling of polyethylene in a two-step thermo-catalytic reaction
system. J. Anal. Appl. Pyrolysis 79, 415–423.
Ahmad, I., Khan, M.I., Khan, H., Ishaq, M., Tariq, R., Gul, K., et al., 2014. Pyrolysis study of polypropylene and polyethylene into premium oil products.
Int. J. Green Energy 12, 663–671.
Aisien, F.A., Amenaghawon, A.N., Adeboyejo, A.R., 2013. Application of recycled rubber from scrap tire in the removal of phenol from aqueous solution.
Pac. J. Sci. Technol. 14 (2), 330–341.
Akpanudoh, N.S., Gobin, K., Manos, G., 2005. Catalytic degradation of plastic waste to liquid fuel over commercial cracking catalysts: Effect of polymer
to catalyst ratio/acidity content. J. Mol. Catal. A: Chemical 235, 69–73.
Al-Salem, S.M., Antelava, A., Constantinou, A., Manos, G., Dutta, A., 2017. A review on thermal and catalytic pyrolysis of plastic solid waste (PSW). J.
Environ. Manag. 197, 177–198.
Altohami, E.K., Mustafa, M.A., 2018. Effect of Kaolin as a catalyst on the yield of catalytic pyrolysis of waste plastic mixtures UofKEJ. 8, (2), pp. 17–24.
Babayemi, J.O., Ogundiran, M.B., Weber, R., Osibanjo, O., 2018. Initial inventory of plastics imports in Nigeria as a basis for more sustainable
management policies. J. Health Pollut. 8 (18), 36–44.
Bozbas, K., 2008. Biodiesel as an alternative motor fuel: production and policies in the European Union. Renew. Sustain. Energy Rev. 12, 542–552.

10
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Budsaereechai, S., Andrew, J.H., Ngernyen, H., 2019. Catalytic pyrolysis of plastic waste for the reduction of liquid fuels for engines. Royal Soc. Chem.
Adv. 9, 5844–5857.
Costa, P., Pinto, F., Ramos, A.M., Gulyurtlu, I., Cabrita, I., Bernardo, M.S., 2010. Study of the pyrolysis kinetics of a mixture of polyethylene,
polypropylene, and polystyrene. Energy Fuels 24 (12), 6239–6247.
Degnan Jr, T.F., 2000. Applications of zeolites in petroleum refining. Top. Catal. 13, 349–356.
Demirbas, A., 2004. Pyrolysis of municipal plastic wastes for recovery of gasoline-range hydrocarbons. J. Anal. Appl. Pyrolysis 72, 97–102. http:
//dx.doi.org/10.1016/j.jaap.2004.03.001.
Duru, R.U., Ikpeama, E.E., Ibekwe, J.A., 2019. Challenges and prospects of plastic waste management in Nigeria. Waste Dispos. Sustain. Energy 1,
117–126.
Fakhrhoseini, S.M., Dastanian, M., 2013. Predicting Pyrolysis Products of PE, PP, and PET using NRTL Activity Coefficient Model. Hindawi Publishing
Corporation, pp. 1–5.
Fivga, A., Dimitriou, I., 2018. Pyrolysis of plastic waste for the production heavy fuel substitute. Energy 149, 865–874.
Gaurh, P., Pramanik, H., 2018. Production and characterization of pyrolysis oil using waste polyethylene in a semi-batch reactor. Indian J. Chem.
Technol. 25, 336–344.
Hakki Metecan, I., Ozkan, A.R., Isler, R., Yanik, J., Saglam, M., Yuksel, M., 2005. Naphtha derived from polyolefins. Fuel 84 (5), 619–628.
Hernandez, M.R., Gomez, A., García, A.N., Agullo, J., Marcilla, A., 2007. Effect of the temperature in the nature and extension of the primary and
secondary reactions in the thermal and HZSM-5 catalytic pyrolysis of HDPE. Appl. Catal. A: General 317, 183.
Jung, S.H., Cho, M.H., Kang, B.S., Kim, J.S., 2010. Pyrolysis of a fraction of waste polypropylene and polyethylene for the recovery of BTX aromatics
using a fluidized bed reactor. Fuel Process. Technol. 91, 277–284.
Kehinde, O., Ramonu, O.J., Babaremu, K.O., Justin, L.D., 2020. Plastic wastes: Environmental hazard and instrument for wealth creation in Nigeria.
Heliyon. 6 (10), e05131.
Kim, J.R., Yoon, J.H., Park, D.W., 2002. Catalytic recycling of the mixture of polypropylene and polystyrene. Polym. Degrad. Stab. 76, 61–67.
Kumar, S., Prakash, R., Murugan, S., Singh, R.K., 2013. Performance and emission analysis of blends of waste plastic oil obtained by catalytic pyrolysis
of waste HDPE with diesel in a CI engine. Energy Convers. Manage. 74, 323–331.
Kunwar, B., Cheng, H.N., Chandrashekaran, S.R., Sharma, B.K., 2016. Plastics to fuel: A review. Renew. Sustain. Energy Rev. 54, 421–428.
Kyaw, K.T., Hmwe, C.S.S., 2015. Effect of various catalysts on fuel oil pyrolysis process of mixed plastic waste. Int. J. Adv. Eng. Tech. 8 (5), 794–802.
Kyong, H.L., Nam, S.N., Dae, H.S., Seo, Y., 2002. Comparison of plastic types for catalytic degradation of waste plastics into liquid product with spent
FCC catalyst. Polym. Degrad Stab. 78, 539–544.
Kyong, H.L., Sang, G.J., Kwang, H.K., Nam, S.N., Dae, H.S., Park, J., et al., 2003. Thermal and catalytic degradation of waste high-density polyethylene
(HDPE) using spent FCC catalyst. Korean J. Chem. Eng. 20, 693–697.
Kyrikou, I., Brissoulis, D., 2007. Biodegradation of agricultural plastic films: A critical review. J. Polym. Environ. 15, 125–150.
Liu, T., Qiu, Z.F., Yang, J., Cao, L.M., Zhang, W., 2015. Output hazard and treatment methods of spent refinery catalysts in China. Environ. Prot. Chem.
Ind. 35, 159–164.
López, G., Artetxe, M., Amutio, M., Bilbao, J., Olazar, M., 2017. Thermochemical routes for the valorization of waste polyolefinic plastics to produce
fuels and chemicals. A review. Renew. Sustain. Energy Rev. 73, 346–368.
Lopez, A., Marco, I.D., Caballero, B.M., Laresgoiti, M.F., Adrados, A., 2012. Catalytic stepwise pyrolysis of packaging plastic waste. J. Anal. Appl. Pyrolysis
96, 54–62.
Lopez, A., de Marco, I., Caballero, B.M., Laresgoiti, M.F., Adrados, A., Aranzabal, A., 2011. The effect of natural clays catalysts on thermal degradation
of a plastic waste mixture. Appl. Catal. B 104, 211–2119.
Luo, G., Suto, T., Yasu, S., Kato, K., 2000. Catalytic degradation of high-density polyethylene and polypropylene into liquid fuel in a powder-particle
fluidized bed. Polym. Degrad. Stab. 70, 97–102.
Ma, C., Yu, J., Wang, B., Song, Z., Xiang, J., Hu, S., et al., 2017. Catalytic pyrolysis of flame retarded high impact polystyrene over various solid acid
catalysts. Fuel Process. Technol. 155, 32–41.
Mabood, F., Jan, M.R., Shah, J., Jabeen, F., Hussain, Z., 2010. Catalytic conversion of waste low-density polyethylene into fuel oil. J. Iran. Chem. Res.
3, 121–131.
Marcilla, A., Beltrán, M.I., Hernández, F., Navarro, R., 2004. HZSM5 and HUSY deactivation during the catalytic pyrolysis of polyethylene. Appl. Catal.
A Gen. 278, 37–43.
Marcilla, A., Beltrán, M.I., Navarro, R.A., 2009. Thermal and catalytic pyrolysis of polyethylene over HZSM5 and HUSY zeolites in a batch reactor
under dynamic conditions. Appl. Catal. B: Environ 86, 78–86.
Marcilla, A., García-Quesada, J.C., Sánchez, S., Ruiz, R., 2005. Study of the catalytic pyrolysis behavior of polyethylene-polypropylene mixtures. J. Anal.
Appl. Pyrolysis 74, 387–392.
Miandad, R., Barakat, M.A., Aburiazaiza, A.S., Rehan, M., Nizami, A.S., 2016a. Catalytic pyrolysis of plastic waste: A review. Process. Saf. Environ. Prot.
102, 822–838.
Miandad, R., Nizami, A.S., Rehan, M., Barakat, M.A., Khan, H., Mustafa, A., Ismail, I.M.I., Murphy, J.D., 2016b. Influence of temperature and reaction
time on the conversion of polystyrene waste to pyrolysis liquid oil. Waste Manage..
Miandad, R., Rehan, M., Barakat, M.A., Aburiazaiza, A.S., Khan, H., Ismail, I.M.I., Dhavamani, J., Gardy, J., Hassanpour, A., Nizami, A.S., 2019. Catalytic
pyrolysis of plastic waste: Moving toward pyrolysis-based biorefineries. Front. Energy Res. 7, 27.
Miteva, K., Aleksovski, S., Bogoeva-Gaseva, G., 2016. Catalytic pyrolysis of waste plastic into liquid fuel. Zsatita Materijala 57 (4), 600–604.
Moorthy Rajendran, K., Chintala, V., Sharma, A., Pal, S., Pandey, J.K., Ghodke, 2020. Review of catalyst materials in achieving the liquid hydrocarbon
fuels from municipal mixed plastic waste (MMPW). Mater Today Commun. 24, 100982.
Moustakas, K., Loizidou, M., 2010. Solid waste management through the application of thermal methods. In: Kumar, S. (Ed.), Waste Management. In:
Tech, Greece, pp. 89–124.
Obali, Z., Sezgi, N.A., Do?gu, T., 2012. Catalytic degradation of polypropylene over alumina loaded mesoporous catalysts. Chem. Eng. J. 207, 421–425.
Onwudili, J.A., Muhammad, C., Williams, P.T., 2019. Influence of catalyst bed temperature and properties of zeolite catalysts on pyrolysis-catalysis of
a simulated mixed plastics sample for the production of upgraded fuels and chemicals. J. Energy Inst. 92, 1337–1347.
Panda, A.K., 2018. Thermo-catalytic degradation of different plastics to drop in liquid fuel using calcium bentonite catalyst. Int. J. Ind. Chem. 9,
167–176.
Patni, N., Shah, P., Agarwal, S., Singhal, P., 2013. Alternate strategies for conversion of waste plastic to fuels. ISRN Renew. Energy 2013, 1–7.
Rehan, M., Nizami, A.S., Shahzad, K., Ouda, O.K.M., Ismail, I.M.I., Almeelbi, T., Iqbal, T., Demirbas, A., 2016. Pyrolytic liquid fuel: a source of renewable
electricity generation in Makkah. Energy Sources Part A 38, 2598–2603.
Ribeiro, A.M., Machado Júnior, H.F., Costa, D.A., 2013. Kaolin and commercial FCC catalysts in the cracking of loads of polypropylene under refinery
conditions. Br. J. Chem. Eng. 30 (04), 825–834.
Sakata, Y., Uddin, M.A., Muto, A., 1999. Degradation of polyethylene and polypropylene into fuel oil by using solid acid and non-solid acid catalysts.
J. Anal. Appl. Pyrolysis 51, 135–155.

11
E.T. Aisien, I.C. Otuya and F.A. Aisien Environmental Technology & Innovation 22 (2021) 101455

Santaweesuk, C., Janyalertadun, A., 2017. The production of fuel oil by conventional slow pyrolysis using plastic waste from a municipal landfill. Int.
J. Environ. Sci. Dev. 8 (3), 168–173.
Santos, B.P.S., Almeida, D., Marques, M.F.V., Henriques, C.A., 2018. Petrochemical feedstock from pyrolysis of waste polyethylene and polypropylene
using different catalysts. Fuel 215, 515–521.
Saptoadi, H., Pratama, N.N., 2015. Utilization of plastics wastes oil as a partial substitute for kerosene in pressurized cookstoves. Int. J. Environ. Sci.
Dev. 6, 363–371.
Sarker, M., Rashid, M.M., 2013. A waste plastic mixture of polystyrene and polypropylene into light grade fuel using Fe2O3 catalyst. Intern. J. Renew.
Energy Tech. Res. 2 (1), 17–28.
Shah, J., Jan, M.R., 2014. Conversion of waste polystyrene through catalytic degradation into valuable products. Korean J. Chem. Eng. 31, 1389–1398.
Shah, J., Jan, M.R., Mabood, F., Jabeen, F., 2010. Catalytic pyrolysis of LDPE leads to valuable resource recovery and reduction of waste problems.
Energy Convers Manage. 51 (2), 2791–2801.
Shakirullah, M., Ahmad, I., Ahmad, W., Ishaq, M., 2010. Oxidative desulphurization study of gasoline and kerosene: Role of some organic and inorganic
oxidants. Fuel Process. Technol. 91 (11), 1736–1741.
Sharuddin, S.D.A., Abnisa, F., Daud, W.M.A.W., Aroua, M.K., 2016. A review on pyrolysis of plastic wastes. Energy Convers. Manage. 115, 308–326.
Singh, M.V., Kumar, S., Sarkerb, M., 2018. Waste HD-PE plastic, deformation into liquid hydrocarbon fuel using pyrolysis-catalytic cracking with a
CuCO3 catalyst. Sustain. Energy Fuels 2, 1057–1068.
Susastriawan, A.A.P., Purnomo, Sandria, A., 2020. Experimental study of the influence of zeolite size on low-temperature pyrolysis of low-density
polyethylene plastic waste. Therm. Sci. Eng. Prog. 17, 100497.
Syamsiro, M., Cheng, S., Hu, W., Saptoadi, H., Pratama, N.N., Trisunaryanti, W., et al., 2014a. Liquid and gaseous fuel from waste plastics by sequential
pyrolysis and catalytic reforming processes over Indonesian natural zeolite catalysts. Waste Technol. 2, 44–51.
Thahir, R., Altway, A., Juliastuti, S.R., 2019. Production of liquid fuel from plastic waste using integrated pyrolysis method with refinery distillation
bubble cap plate column. Energy Report 5, 70–77.
Yuliansyah, T.A., Agus, P., Ramadhan, A.A.M., Laksono1, R., 2015. Pyrolysis of plastic waste to produce pyrolytic oil as an alternative fuel. Int. J.
Technol. 7, 1076–1083.

12

You might also like