Next Article in Journal
A Novel Control Algorithm of the Air Supply Subsystem: Based on Dynamic Modeling of Proton Exchange Membrane Fuel Cell
Previous Article in Journal
Growth Performance and Meat Quality of Growing Pigs Fed with Black Soldier Fly (Hermetia illucens) Larvae as Alternative Protein Source
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Catalytic Pyrolysis of Municipal Plastic Waste for the Production of Hydrocarbon Fuels

1
Department of Mechanical, School of Engineering Studies, University of Petroleum and Energy Studies, Dehradun 248007, Uttarakhand, India
2
Department of Chemistry and Centre of Alternate Energy, School of Engineering, University of Petroleum and Energy Studies, Dehradun 248007, Uttarakhand, India
3
Department of Chemical Engineering, National Institute of Technology Calicut, Kozhikode 673601, Kerala, India
4
Yogoda Satsanga Mahavidyalaya, Ranchi 834001, Jharkhand, India
5
Biochemical Process Engineering, Division of Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, SE-971 87 Luleå, Sweden
*
Author to whom correspondence should be addressed.
Processes 2022, 10(8), 1497; https://doi.org/10.3390/pr10081497
Submission received: 29 April 2022 / Revised: 13 July 2022 / Accepted: 15 July 2022 / Published: 29 July 2022
(This article belongs to the Section Environmental and Green Processes)

Abstract

:
Currently, the resources of fossil fuels, such as crude oil, natural gas, and coal, are depleting day by day due to increasing energy demands. Nowadays, plastic items have witnessed a substantial surge in manufacturing due to their wide range of applications and low cost. Therefore, the amount of plastic waste is increasing rapidly. Hence, the proper management of plastic wastes for sustainable technologies is the need of the hour. Chemical recycling technologies based on pyrolysis are emerging as the best waste management approaches due to their robustness and better economics. However, research on converting plastic waste into fuels and other value-added goods has yet to be undertaken, and more R&D is required to make waste-plastic-based fuels economically viable. In this review article, the current status of the plastic waste pyrolysis process is discussed in detail. Process-controlling parameters such as temperature, pressure, residence time, reactor type, and catalyst dose are also investigated in this review paper. In addition, the application of reaction products is also described in brief. For example, plasto-oil obtained by catalytic pyrolysis may be utilized in various sectors, e.g., transportation, industrial boilers, and power generation. On the other hand, byproducts, such as solid residue (plasto-char), could be used as a road construction material or to make activated carbon or graphenes, while the non-condensable gases have a good potential to be utilized as heating/energy source.

1. Introduction

The demand of energy is continuously increasing due to increasing population and industrialization. Higher energy demand has led to rising demand for petroleum exploration as well as more environmental pollution [1]. As per the petroleum planning and analysis cell, India consumed almost 133.05 billion litres of fuels in the 2020 financial year. The trend of the fuel market from 2011 to 2019 in India shows that the fuel market increased by 152.7 billion from 2015 to 2019, which was 3.1% higher than that reported for the 2018 financial year [2]. As a result, it is reasonable to expect that it will rise in the next few years as the population’s need for energy rises. Therefore, there will be a need to place more emphasis on alternative fuels.
In recent times, there has been a strong tendency to move away from traditional fossil fuels in favor of novel and renewable energy sources that are cleaner, safer, and inexhaustible. With the supply-demand gap expanding, it is more important than ever to diversify energy sources and look into alternative ways to meet the country’s energy demands while preserving economic growth. Growing environmental concerns present a huge issue for energy companies, emphasizing the need to move to cleaner, more sustainable energy sources, such as compressed natural gas (CNG), liquefied petroleum gas (LPG), hydrogen, biodiesel, and plastic-derived oils. Several state-of-the-art studies show that these alternative fuels are a valuable source of energy, with fewer emissions, and are being used in the energy sector [3,4,5,6,7], e.g., CNG is used in cars, but there is a limit on its power output [8]. Hydrogen is a clean energy source with low toxicity, but it may not be considered the safest fuel in commercial vehicles [9,10]. Similarly, the Indian government has been focusing on biofuels as they have the potential to be used in the transportation sector. However, as compared to conventional diesel, the higher oxygen concentration in biofuels enhances nitrogen oxides (NOx) emissions in the vehicle, but some controversial studies have suggested that it depends on the composition of biodiesel [11]. It can be 10% higher than petroleum-based compounds, which may dissolve in moisture in the atmosphere and generate acid rain. However, fuel produced from municipal plastic waste attracts the researcher due to its higher potential and availability [12].
In the current scenario, plastic is one of the essential products utilized in a huge amount by the growing population due to its durability, lightweight, and ease of production, making it more viable in utilization. Figure 1 depicts the plastic production in 2020 and the plastic production forecasted from 2025 to 2050. Plastics end up as waste after being used once. The Central Pollution Control Board (CPCB) reported that 3.3 million tons of waste plastic were generated in India in 2018–19 [13] that were dumped into the nearby municipal dumping yards.
Collection and segregation of the municipal mixed plastic waste from the dumping yard is again a challenging step, and it should be cost-effective to implement the technology for waste to energy conversion [14]. Technically, it is possible to separate municipal mixed waste into distinct streams; however, this may be more expensive due to the infrastructure requirements for efficient collection and separation. It is noted that existing technology such as tribocharging and electrostatic separation technology can separate the different compositions of waste plastic: high-density polyethylene (HDPE)/polypropylene (PP), low-density polyethylene (LDPE)/PP, and polyethylene terephthalate (PET)/ polyvinyl chloride (PVC) [15]. Other technologies, such as biological treatment to separate nanoparticles from plastic waste and a hierarchical classification strategy to separate LDPE and HDPE, are similar [16]. Thus, different technologies are available to separate municipal mixed plastic waste (MPW), but the employment of the technology to handle the waste plastic may not be economically viable. Sharuddin et al. discussed, in a review study, that different kinds of plastic (PET, PVC, HDPE, LDPE, and MPW) have the potential to be converted into energy as valuable liquid products. However, they concluded that PET and PVC both have a lower tendency to convert into valuable liquid products than other plastics. For instance, thermal degradation of the polyvinyl chloride (PVC) releases hydrochloric acid, adversely affecting liquid product quality and yield [17]. Moreover, contamination of PVC in polyethylene terephthalate (PET) will decompose the resin of PET by becoming brittle and yellowish, which further requires reprocessing. As a result, MPW treatment via the thermal degradation process is not worthwhile, with impurities necessitating some additional operation for waste plastic segregation based on colors, transparency, and resin [18].
A study conducted on the pyrolysis of mixed plastic waste contained 46 wt.% LDPE, 30 wt.% HDPE, and 24 wt.% PP at 650 °C temperature via catalytic pyrolysis (1% of Z–N catalyst), and the ratios of reaction products (gas/liquid/solid) were 6.5/89.0/4.5 wt.%, respectively [19]. On the other hand, some studies have shown that the pyrolysis of municipal plastic waste produces a lower liquid yield of less than 50%, while single plastic feedstocks yielded liquid yields of 97.0 wt.%, 93.1 wt.%, 82.12 wt.%, 80.88 wt.%, 23.1 wt.%, and 12.79 wt.%, respectively, for PS, LDPE, PP, HDPE, PET, and PVC [20,21,22,23,24,25,26]. Similarly, Cepeliogullar et al. explored the pyrolysis of PET waste by using the fixed-bed reactor at a temperature of 500 °C and a heating rate of 10 °C/min, where they observed 23.1 wt.% liquid yields and 76.9 wt.% gaseous with no solid product [24]. However, lower liquid yield indicates low availability of volatile content in PET at around 86.83% compared to the other plastics (>90%) [24]. Salem et al. performed thermal pyrolysis of the reclaimed plastic wastes from unsanitary landfill sites in an auger pyrolysis system at a temperature of 500 °C. The process mass balance was developed on a dry basis, and the yields of pyro-oil, light wax, heavy wax, and gases were 5.5, 23.8, 69.4, and 1.3 wt.%, respectively [27]. The result proves that when increasing the temperature, the liquid yield will increase up to a certain limit, but too-high temperature will reduce the liquid yield and increase the gaseous product. This may be due to the reason that lower temperature leads to volatile formation at a slow rate, which allows sufficient time for volatiles to crack into lower hydrocarbons inside the reactor before leaving to the condensation section, whereas at higher temperature, the volatile formation is rapid due to higher temperature, which does not provide the same residence time for the volatiles to stay inside the reactor to crack down into more smaller hydrocarbons [28]. Furthermore, at higher temperatures, the reaction mechanism is dominated by the inter- and intramolecular hydrogen transfer, followed by b-scission, producing more lighter hydrocarbons. Table 1 shows the volatiles present in different types of plastics.
In the present study, the author focused on thermal pyrolysis and factors affecting the process. Recent development in the pyrolysis process was also investigated, including a comparative study with catalytic thermal pyrolysis with different catalysts and their effect on the solid, liquid, and gaseous hydrocarbons.
Figure 1. Worldwide production of plastic [48,49].
Figure 1. Worldwide production of plastic [48,49].
Processes 10 01497 g001

2. Thermal Decomposition of Plastic Waste

Conventional Pyrolysis Process

Conventional pyrolysis of plastic refers to the pyrolysis process that uses an electrical heater or burner as the source of heat. Many studies have been performed on conventional pyrolysis using different types of plastics, e.g., PS, PP, PS, HDPE, LDPE, and mixed waste plastic [23,50,51,52]. The operating modes for plastic pyrolysis are batch, semi-batch, or continuous reactors, and each has pros and cons of its own. High conversion can be achieved in a batch reactor, but the inconsistent nature of the end product and the high labor costs make it unsuitable for industrial production. Although fixed-bed reactors are noted for their simplicity in design, they have a smaller surface area where reactions can take place. Fluidized-bed reactors, as opposed to fixed-bed reactors, guarantee that the catalyst and fluid are thoroughly mixed, increasing the catalyst’s usable surface area and enhancing heat transmission. Large-sized particles can be handled by conical spouted-bed reactors (CSBR), although product collection and catalyst feeding provide technological hurdles. Ghodke et al. carried out the conventional pyrolysis of virgin mixed plastic and municipal mixed plastic waste (MMPW) at 500 °C and obtained a maximum yield of 62.5 wt.% with MMPW. Heating is the main cost and bottleneck in pyrolysis. The huge cost of the conventional pyrolysis method is due to inefficient heating and heat loss. Thus, researchers have recently attempted to utilize other heating sources, for example, solar [53], microwave [54], and other renewable energy sources [55,56].

3. Recent Development in the Pyrolysis Process

The pyrolysis process decomposes organic matter by heating it at an elevated temperature with no atmospheric oxygen. Heating is required to break down the hydrocarbon molecules at suitable temperatures in the pyrolysis reactor and produce solid, liquid, and gaseous fuels. The quantity of these pyrolysis products depends on feedstock, temperature, heating rate, and the type of reactor being used for the process. On the basis of reaction temperature and heating rate, the pyrolysis process is categorized in slow, fast, and flash pyrolysis. The pros and cons of different pyrolysis processes are presented in Table 2 and Figure 2.

3.1. Slow Pyrolysis Process

Slow pyrolysis is pyrolysis that takes place at a lower operating temperature, with a slower heating rate, and allows more production of coke, tar, and char [59]. The feedstock is pyrolyzed with a low heating rate of around 0.1 to 0.8 °C s−1. Slow pyrolysis heating rates combined with a long feedstock residence time result in more char production. The synthesis of H2, CO, CH4, and C5H12 is aided by a significantly lower non-isothermal heating rate, while fast pyrolysis yields predominantly gases [60].
Das et al. demonstrated in their study that thermally degraded PP and PE wastes produce light hydrocarbon liquid fractions (C6–C20) at low temperatures (400 °C). Slow pyrolysis, on the other hand, produces higher char yields [61].
Thus, the carbonization method was introduced as a novel method for producing charcoal as a valuable product for underserved areas. The process requires a proximate feedstock analysis to analyze the moisture content, ash, fixed carbon, and volatile matters. Although process temperature plays a crucial role in product distribution particle size, heating rate and moisture content also have a big impact on pyrolysis products [56].

3.2. Fast Pyrolysis

Fast pyrolysis is a cutting-edge and promising technology that contemplates rapid thermal degradation with increased heating rates and/or high process temperatures for rapid product quenching and liquid product formation. This process is allowed to produce good-quality liquid fuel with a higher calorific value [62].
Fast pyrolysis results in high yields of gaseous components with low yields of char and oil because the decomposition of the plastic waste is rapid, allowing the molecules to break into small carbon chain compounds such as methane, ethane, propane, butane, etc. During fast heating, the dissociation reaction of the waste plastics allows the β-scission, followed by inter- and intramolecular hydrogen transfer, leading a larger quantity of short-chain hydrocarbons [62]. The process requires a higher heating rate of around 10 to 100 °C s−1, and the feedstock particle size must be less than 3 mm. Mass transfer rate, heat, and chemical kinetics are the essential factors in deciding the pyrolysis product chemistry. Various researchers conducted studies on fast pyrolysis and produced higher liquid yield at a temperature range of 500–550 °C [57,62,63,64,65]. The process produces byproducts such as residue char and pyrolytic gases. The state-of-the-art shows that pyrolytic gases increase by decreasing the liquid oil and solid-product-char [45,63,65].

3.3. Superfast/Ultra-Fast Pyrolysis

Superfast pyrolysis is the most advanced pyrolysis technique. The process produces higher amount of operable liquid oil and pyrolytic gases with lower water content at a high-temperature range of 700 to 1200 °C or above. The process requires a higher heating rate of about 1000 °C s−1 or above to provide a higher temperature to the feedstock. The feedstock particle size should be less than 0.2 mm because of the faster heating rate [66]. The pyrolysis process converts 80% of the energy contained in the feedstock distributed into various energy products, producing higher energy-dense products than the raw feedstock. However, the liquid oil produced contains much oxygen content, indicating that the product is unstable for further use. It also contains impurities such as heavy metals and nitrogen, and removing these pollutants necessitates a considerable amount of hydrogen, which makes the process somehow costly [67].

3.4. Catalytic Thermal Degradation (CTD)

The catalytic thermal degradation (CTD) process is more advanced than the conventional pyrolysis process because the reaction is completed in a catalyst environment, and offers improved liquid fuel quality compared to the conventional thermal process. Thus, selecting a suitable catalyst plays a vital role throughout the process. The CTD process involves a catalyst to increase the reaction rate of thermal degradation. The catalysts lower the activation energy and the process temperature required to improve the overall process and conversion efficiency. The thermal degradation process is energy-intensive as it requires a higher temperature to degrade the waste plastic and transform it into a quality end product, which might not require any upgrading process [63]. Thus, the CTD process is economically viable to produce valuable products such as liquid oil yields of low molecular weight and upgraded quality at reduced thermal and energy demand. This liquid product is obtained with 38–46 MJ kg−1 higher heating value (HHV) and is almost close to conventional diesel, with 45.5 MJ kg−1 [68]. The plastic waste decomposition can be completed in two stages in the presence of a catalyst. The first stage is related to the disintegration of short-length branches, whereas the second stage is associated with the degradation of the molecular backbone of the main chain.
Moreover, increasing the wt.% of the catalyst leads to increased residue char because of the non-decomposable content available in the catalyst and carbon deposition over the catalyst surface. Therefore, it can be concluded that catalyst concentration directly affects the thermal degradation process. The significance of a catalyst is that it starts the decomposition at a lower temperature, which is proportional to the concentration of the catalyst.

4. Catalyst Effect on Thermal Degradation Process

The catalyst plays a vital role in cracking the hydrocarbon chain during the thermal process, and affects the yield of the optimized product and residence time [56,69]. The yield of the product may vary which may depend on the type of plastic waste and the percentage of the catalyst introduced in the process. Muhammad, et al. [70] studied the thermal pyrolysis of post-consumer mixed plastic waste and a simulated plastic mixture using zeolite HZSM-5 catalyst in a two-stage fixed bed reactor. The highest liquid oil yield was 81–97 wt.% in the thermal degradation of plastic. The addition of the catalyst reduced the liquid yield in the range of 44–51 wt.% and produced a higher gas yield due to the cracking of oil volatiles. Researchers studied the impact of zeolite ZSM-5 on the thermal degradation of plastic. It is economically viable and has excellent thermal stability as well as selectivity [71,72,73]. In the catalytic thermal process, zeolite increases the cracking reaction and deoxygenation to produce liquid oil. Onwudili et al. investigated the effect of temperature and zeolite catalyst type on waste plastic pyrolysis [74]. The introduction of a different catalyst at temperatures ranging from 500 to 600 °C had no apparent effect on fuel range liquid products. However, higher temperatures led to an increase in the gaseous product during the process. Muneer et al. investigated the pyrolysis process on feedstock maize stalk and polypropylene at 500 °C with a 1:4 ratio of ZSM-5 catalyst to feedstock; the liquid hydrocarbon yield increased to 66.4% due to the high surface area and selectivity of ZSM-5 catalyst to form the hydrocarbon products [75]. Herein, catalysts with acidic sites on the surface and the ability to contribute hydrogen ions accelerate the rate of isomerization products, increasing the yield of hydrocarbon isomers and the quality of the fuel generated. Catalysts with more acid sites and high density are more effective in cracking polyolefin. However, the catalyst deactivates quickly due to the high acidity and large pore size. According to the literature, it is better to perform polyolefin pyrolysis in the presence of a catalyst with low acidity [76]. The efficiency of hydrocarbon cracking was improved by dissolving the catalyst in a polymer melted state. The catalysts modified the charge distribution in the carbon chain, allowing them to extract hydrocarbon hydride ions and produce carbonium ions. It improved the catalytic impact, allowing for a lower pyrolysis temperature and increased ion production for olefin and aromatic chemicals [19]. As a result, zeolite catalysts influence the degradation process, boost the generation of gaseous hydrocarbons, and result in high liquid hydrocarbons.

4.1. Factors Affecting the Pyrolysis Process

Pyrolysis of municipal plastic waste has been affected by the critical process parameters that play an essential role in optimizing the product yield and quality in the process. In the MPW thermal degradation process, critical parameters influence the final hydrocarbon products. The parameters include temperature and pressure.

4.2. Effect of Temperature

Temperature is the essential parameter for carrying out the pyrolysis of plastic waste because it regulates the cracking response of the polymer chain and influences product distribution. Temperature controls the breakdown process of the hydrocarbon molecules and alters the amount of gaseous and liquid hydrocarbons. Moreover, it also affects the char produced. The van der Waals force attracts molecules together, preventing them from collapsing [77]. The vibration of molecules inside the system increases as the temperature of the system rises, and molecules tend to evaporate away from the feedstock surface. The energy produced by the van der Waals force along polymer chains exceeds the enthalpy of the C–C bond in the chain, causing the carbon chain to break. In polymers, carbon bonds are nonpolar, stable, and highly tight, requiring a lot of energy to break. As a result, it needs a higher temperature range of 300–800 °C. Various studies demonstrate the effect of temperature on liquid yield with various feedstocks [78,79,80]. For instance, Adnan et al. performed a study on polystyrene waste material at 500 °C where they obtained more than 80 wt.% liquid hydrocarbons [81]. The lower temperature range resulted in the formation of a long hydrocarbon chain. However, higher temperature range resulted in the synthesis of a short carbon chain due to the breaking of C–C bonds and the production of aromatic hydrocarbons.

4.3. Effect of Pressure

The pyrolysis of plastic waste is usually performed at atmospheric pressure. As a result, the impact of pressure on pyrolysis is rarely explored in the literature and needs to be addressed in detail, leaving a substantial research gap in the field. However, some researchers have performed pyrolysis under vacuum (lower than atmospheric pressure) or higher pressure conditions (higher than 1 atm). Low pressures (under vacuum or in the presence of an inert diluent) favor the synthesis of primary products, such as monomers, whereas high pressures favor complex liquid fractions. The boiling point of pyrolytic compounds increases when pressure is increased. This raises the pressure in the atmosphere, causing heavy hydrocarbons to be pyrolyzed rather than vaporized at a given working temperature [82].
Lopez et al. studied the waste tire pyrolysis process under vacuum at 25 kPa and 50 kPa under atmospheric pressure at a temperature 500 °C, where they obtained a higher diesel fraction similar to liquid hydrocarbons above the atmospheric pressure [83]. The principal impact of a vacuum pressure over an atmospheric pressure is an increase in the liquid product. Furthermore, a reduction in pore blockage improved the surface areas of the carbon blacks formed, which had a good influence on the porous structural features of the residual carbon black. The vacuum impact on feedstock was introduced to devolatilization and diffused the volatiles inside the particle. Increasing pressure resulted in more gaseous products and a higher yield of lower molecular liquid hydrocarbons [84].

4.4. Residence Time

The residence time of the feedstock material plays a crucial role in the pyrolysis process. It is defined as the amount of time a particle spends in a pyrolysis reactor as long as the products continue to produce, and has a high impact on product quality as well as quantity [85,86]. Higher residence times favor higher liquid production due to the secondary reaction in the pyrolysis reactor. Al-Salem et al. performed a study on residence time, and the results showed that a higher liquid yield was obtained with increasing the residence time. Studies have shown that residence time influences liquid hydrocarbon yield at a given temperature and that increasing residence time may reduce liquid hydrocarbon yield [87]. Similarly, Mastral et al. studied the thermal cracking of HDPE in a fluidized-bed reactor to obtain the effect of residence time. The experiments were performed at temperatures ranging from 650 °C to 850 °C, with residence time ranging from 0.64 s to 2.6 s. At 650 °C, it was observed that the liquid hydrocarbon yields 79.7% at 0.8 s and 68.5% at 1.5 s [79].
Moreover, the increase in the gas yields were 11.4 wt.% and 31.5 wt.% at a residence time of 1 s and 1.5 s, respectively. The lower liquid yield of 9.6 wt.% was obtained at a temperature of 780 °C and 86.4 wt.% of gaseous hydrocarbons by increasing the yield of methane, ethane, and hydrogen [87].
Therefore, increasing residence times resulted in higher gaseous products. However, along with residence time, other factors, such as temperature, heating rate, catalyst, and reactor design, also simultaneously highly influence the hydrocarbon fuel quality [52].

4.5. Type of Reactor

Fluidized-bed (bubbling and circulation), rotary kiln, ablative, auger, and screw reactors are among the most typical pyrolysis reactors used to produce liquid hydrocarbons from MPW. To achieve successful pyrolysis, two essential design characteristics must be addressed, where the reactor should blend the MPW and catalyst effectively, deliver a large amount of heat to the feedstock, and have a short residence period [88]. Large volumes of solid and gaseous hydrocarbons can be generated if the heat transfer rate is low and the residence time is prolonged.

4.5.1. Fixed-Bed Reactor

A fixed-bed reactor is used in the pyrolysis process and is often operated in batch mode. The reactor’s design is simple to construct and suitable for the uniform size of the feedstock, as shown in Figure 3a. However, larger size of the particles or polymers leads to feeding issue into reactors. Heat is provided to this reactor for the thermal breakdown of MPW either externally or by minimal combustion. The products can flow out of the pyrolizer reactor due to expanding the volume of gaseous hydrocarbons. At the same time, the char remains in the reactor. However, a sweep gas is often used to remove the volatiles from the reactor effectively [89].
In a few cases, the solid catalyst was used as a packed bed where the volatiles pass over the bed and improve the efficiency. On the other hand, fixed-bed reactors are often linked to the problems of feeding irregularly shaped polymers with poor heat conductivity. Moreover, researchers have used the fixed-bed reactor for secondary pyrolysis so that the primary product may be easily transferred to the secondary reactor [90]. The catalytic process of polyethylene with Y-zeolite produced 85 wt.% liquid hydrocarbons. However, in the absence of a catalyst, 95 wt.% liquid hydrocarbons were produced at 500 °C. Furthermore, they concluded that increasing the temperature decreased the liquid hydrocarbons with higher gaseous products [91].
When large volumes of material are processed in the traditional fixed-bed reactor, it shows some limitations regarding a slow heating rate, a long residence period, and a non-uniform temperature of the sample within the reactor. Furthermore, the fixed-bed reactor is uneconomic due to its batch-scale nature. As a consequence, the reactor might be used for research as well as small-scale heating and power generation.

4.5.2. Fluidized-Bed Reactor

A fluidized-bed reactor is the most viable reactor, as presented in Figure 3b, and is being used on the industrial scale to process crude oil. The reactor has comprehensive utilization in the industries due to the continuous provision of the feedstock to produce the liquid hydrocarbons. The upward motion of the co-reactant feed in the reactor suspends tiny catalyst particles in suspension. A gas is often used as the fluid in a fluidized-bed reactor, which facilitates particle mixing by maintaining a high flow rate. The particle size is much lower than in a fixed-bed reactor, ranging from 10 to 300 µm with a constant temperature. Thermal and catalytic thermal processes can be carried out at 290–850 °C in the reactor [91]. Luo et al. conducted a comparative study on LDPE on a fluidized-bed reactor and produced a higher liquid yield [92]. Furthermore, the fluidized-bed reactor provides considerably more significant mass enhancement and heat transmission, reducing temperature gradients in the reactor [93,94]. It shows that a reactor with a fluidized bed would be better suited for a large-scale operation in terms of economic sustainability, even though batch reactors are equally efficient at the laboratory scale of the experiment.

4.5.3. Conical Spouted-Bed Reactor (CSBR)

In the pyrolysis of plastic waste and gasification, a conical spouted-bed reactor is employed. As illustrated in Figure 3c, the CSBR comprises a hopper, feeding system, gas mixture, gas preheater, reactor, and condenser. The conical spouted reactor is conical from the end and has a cylindrical part at the top [56]. The CSBR might be used to process the large particle size distribution with adequate mixing. The principal advantages of the CSBR over the fluidized-bed reactor include fast heat transfer rates and negligible defluidization issues when handling highly sticky inputs such as PE and PP at a particular temperature. The reactor works on a slow pyrolysis process at low temperatures and results in higher wax yield. The results of a study using HZSM-5, HY, and Hβ zeolite-based catalysts in the pyrolysis of high-density polyethylene (HDPE) continuously fed into a conical spouted-bed reactor (CSBR) at 500 °C and atmospheric pressure revealed that the HZSM-5 zeolite-based catalyst is very selective to light olefins, yielding 58 wt.%, whereas Hβ and HY result in high yields of non-aromatic C5–C11 products (around 45 wt.%). Due to catalyst deactivation by coke generation, wax yield increases as reactions proceed, notably with HY and Hβ zeolite-based catalysts [94]. Fernandez et al. conducted a study on the biomass steam pyrolysis in a CSBR and achieved 75.4 wt.% bio-oil at a 500–800 °C temperature range [95]. Arabiourrutia et al., on the other hand, employed CSBR to obtain wax products from polyolefins plastics. They discovered that the spouted bed’s design was mainly suited for extracting wax by pyrolysis at low temperatures since the wax yield decreased with rising temperature [96]. This might be because as temperature rises, more waxes are broken into liquid or gaseous products.
Hence, it may be concluded that CSBR provides efficient heat transmission between phases and proper mixing, and the ability to handle large particles of varying densities. However, feeding and entrapment of the catalyst are challenging.

4.5.4. Rotary Kiln Reactor

Figure 3d shows the schematic diagram of the rotary kiln reactor. A rotary kiln reactor is more efficient due to its proper mixing. The reactor efficiently mixes municipal solid waste (MSW) very well compared to other reactors. However, a slow pyrolysis process is performed in a reactor, suggesting that the heating rates are modest. The heating rate is limited to 100 °C/min, and the residence duration is one hour [97]. Rotary drums, furnaces, and kilns are composed of steel drums walled internally with high-resistance and anti-corrosive refractory material. Reactors are intended for prolonged residence durations of 20 min or more, with inclined low-speed rotational movements. The product in rotary kilns varies significantly due to the huge temperature differential in the system. The rotation of the kiln aids in the appropriate mixing of the MPW. Then, it is progressively heated as pyrolysis gases are produced as it moves down the cylinder and carbonizes. The rotary kiln reactor is the most often described for MPW pyrolysis due to distinct advantages [73]. Some of the benefits include proper MPW mixing, flexibility in residence time modification, the ability to feed heterogeneous materials, the absence of the need for MPW pre-treatment, and ease of maintenance.

4.5.5. Auger Reactor

The auger type reactor is also known as the screw type reactor because of having a screw, as shown in Figure 3e. The reactor consists of a hopper, tubular shape, and having a screw that carries the continuous feedstock in operational mode. The screw’s spinning aids in transferring feedstock into the reactor, and the heat required for pyrolysis is carried through the reactor’s tubular wall. The heat from the outside has been observed to be sufficient to heat reactor tubes with lower diameters [98]. If higher production quantities are envisaged, a hot solid carrier filled with feedstock particles is required. This allows feedstock particles to interact more closely as they move through the reactor tubes. Steel and ceramic pellets may be used to make the solid carrier. The liquid hydrocarbons are produced from the vapors obtained by the pyrolysis process in a condenser and compressing. An auger reactor can be constructed to be extremely small and even portable in certain situations. It may also be used near the dumping yard of MPW.

4.6. Type of Catalyst

Catalysts play a crucial role in enhancing product quality while lowering process temperature and retention time, resulting in overall process optimization. Catalytic cracking is more enticing than thermal cracking alone since it is quicker and needs lower temperatures, lowering the energy requirement substantially. Furthermore, catalytic cracking using catalysts such as zeolites produces high-quality products in the spectrum of motor engine fuels, decreasing the need for additional upgrading. Because its products need additional upgrading, thermal pyrolysis is limited to regions of existing oil refineries [68]. Catalysts decrease the desired pyrolysis temperature, shorten reaction time, create diesel components in the appropriate boiling point range of 390–425 °C, improve gasoline selectivity, and accelerate the occurrence of isomerization [99].
Catalysts are widely employed in refineries and in the pyrolysis of plastics to recover hydrocarbons from waste, both of which follow a similar process. The global refinery catalysts market has been significantly rising despite strict rules and regulations and increased demand for petroleum and its derivatives. The industry’s rapid rise is fueled by the growing demand for transportation fuel. Furthermore, strict emission control regulations throughout the world have demonstrated the necessity for efficient, conservative, and high-yield techniques, which is one of the key drivers behind market advancement. As a consequence of this environmental concern, demand for refinery catalysts has increased, enhancing the efficiency of the process. The type of catalyst impacts the process outcome since a large surface area of the catalyst may change the nature of the pyrolysis and, consequently, have a substantial impact on the pyrolysis products [100,101].
Generally, two types of catalysts such as heterogeneous and homogeneous, have been employed in the thermal degradation of plastic waste. The former uses a single phase (reactant and catalyst in the same phase), while heterogenous catalyst refers to the reaction where the catalyst and reactant are in different phases. Catalysts made of Fe/Al2O3 ratio are the most prevalent form of homogeneous catalyst used in plastic waste pyrolysis [101]. Nonetheless, heterogeneous catalysts are the most often utilized in plastic waste pyrolysis because the fluid product can be readily separated from the solid catalyst, easily regenerated, and reused. Nanocrystalline zeolites, the most widely used [102], ordinary acid solids, mesostructured catalysts, for example, metal-supported carbon, and basic oxides. Heterogeneous catalysts have also been reported to endure extreme reaction conditions of up to 1300 °C and 35 MPa and are readily separated from the gas and liquid reactants and products. Heterogeneous catalysts include acid solids, mesostructured, and nanocrystalline zeolites catalysts [103]. Various researchers have shown interest in nanocrystalline zeolites due to their significant role in product distribution [104,105,106]. In addition, various catalysts, such as fly ash, bentonite, silicate, silica-alumina, ceramic, activated char, and MVM-41, have recently attracted researchers’ attention. Table 3 summarizes the catalytic effect on the solid, liquid, and gaseous hydrocarbons with different heating rate and reactors.

4.6.1. Zeolite Catalyst

In the thermocatalytic process, zeolite has been extensively documented as an efficient and selective material catalyst for creating biofuels. ZSM-5 is considered a relatively inexpensive catalyst for transforming plastic waste into liquid hydrocarbons, in addition to its catalyst base. The catalyst also has higher thermal stability, selectivity, activity, and coke deactivation [119]. Zeolites, also known as aluminosilicate crystalline sieves, are solid crystalline structures made up of a three-dimensional framework in which oxygen atoms join tetrahedral sides with open pores which have ion exchange capabilities. The reactivity and efficacy of zeolite are influenced by the SiO2/Al2O3 ratio, which influences the type and quality of the pyrolysis process ultimate yield product. The proportion of silica–alumina in various zeolites varies, and the particular ratio has a distinct reactivity. Ates et al. proved this with varied SiO2/Al2O3 ratios of 30, 80, and 280, utilizing the catalyst HZSM-5 zeolite. SiO2/Al2O3 (30) has the highest acid strength and the largest BET surface area and average pore diameter. Higher acidity of the catalyst is obtained with a SiO2/Al2O3 ratio of 30, sufficient to crack the wax and lead to a higher yield of the light olefins and lower the heavy fraction (C12–C20). However, decreasing the SiO2/Al2O3 from 280 to 30 results in an increase in light olefins output from 35.5 to 58 wt.% and a reduction in C12–C20 and C5–C11 fraction yields from 28 to 5.3 wt.% and 28.8 to 15.2 wt.%, respectively [120]. The pore size and acidity of various zeolite base catalysts are varied, affecting the final product. For example, a larger pore size makes it simpler to convert plastic into polyalkylaromatics, but a smaller pore size converts only to aromatic compounds with small dynamic diameters. According to their findings, choosing the right zeolite to use as a catalyst in the pyrolysis reaction is crucial in deciding the end product formed. The pore size and acidity of various kinds of catalyst base zeolite varies, affecting the end product [121]. Wong et al. investigated LDPE pyrolysis over ZSM-5 zeolite and found that catalytic cracking produced the C1–C8 hydrocarbons due to the efficient heat exchange between catalyst and polymer chain [122].
Sivagami et al. found that LDPE pyrolysis with synthesized ZSM 5 catalyst at 500 °C temperature produced 70% liquid yield with 16 wt.% gas and 14 wt.% solid char. The synthesized ZSM-5 catalyst’s strong acidic characteristics and microporous crystalline structure allow greater cracking and isomerization, resulting in higher breakdown of larger molecules into smaller molecules and more oil production in pyrolysis studies [123].
Susastriawan et al. investigated the influence of zeolite size on the low-temperature pyrolysis of LDPE [80]. The scientists observed that reducing zeolite size might improve reaction rate, pyrolysis temperature, heat transfer rate, and oil products during the pyrolysis process due to larger surface-active area with a smaller size. The oil yields were maximum when the zeolite particle size was 1 mm; however, the results of oil produced were not significant when the particle size was 1–3 mm. According to prior studies published by Kim et al., the presence of the phenolic functional group on the lignin may increase the 39% aromatic hydrocarbon in the final products over zeolite catalysts [124]. Condensation is caused by the presence of a hydroxyl group on the surface of a zeolite-type catalyst that speeds up the formation of aromatic hydrocarbon molecules by dihydroxylation, aromatization, isomerization, and oligomerization pathways [125].
However, researchers have also proven that heat activation and acid activation may improve the catalytic capabilities of natural zeolite. Rashid et al. used natural zeolite that had been modified by thermal and acid activation in the pyrolysis of plastic waste (PS, PE, PP, and PET) and found that acid-activated zeolite has better catalytic activity than thermally activated zeolite [126]. Another advantage of this catalyst is that the spent zeolite catalyst may be recovered and utilized in a similar reaction with a similar goal with more or less efficacy, as well as reduce impurities of sulfur in the liquid product generated [127].

4.6.2. Fluid Catalytic Cracking (FCC) Catalyst

The fluid catalytic cracking catalyst plays a vital role in cracking the polymer during pyrolysis reaction for petroleum-based products. It has been used in the petrochemical industries to convert the high boiling and molecular weight products to the valuable fuel products in gasoline and LPG fractions [128]. Some progress in FCC catalysts has been investigated in order to improve the effectiveness, cracking activity, coke selectivity, and stability of this catalyst, which focused primarily on modifying zeolite using extra aluminum framework where ion exchange takes place [129]. Aisien et al. evaluated that pyrolysis of polyethylene in the presence of an FCC catalyst reduced the liquid yield from 83.3 wt.% to 77.6 wt.%; however, it increased the gaseous hydrocarbons from 13.2 wt.% to 19.7 wt.% and decreased the solid char formation from 3.0 wt.% to 2.7 wt.% [130].
Similarly, Abbas et al. examined the effect of the FCC catalyst in the range of 10 wt.% to 60 wt.% at a temperature range of 450 °C in a stirrer semi-batch reactor and found no major impact on the non-condensable gaseous hydrocarbons, but it seems that there was an increase in the gaseous products [131]. However, it increases the solid hydrocarbons by increasing the catalyst-to-polymer ratio concentration. This might be attributed to greater dehydrogenation and aromatization on the catalyst surfaces, resulting in an increase in solid hydrocarbons on the catalyst surface. Coke causes the catalyst to deactivate, making it ineffective for breaking waxes. Catalyst activity monitoring revealed that an equilibrated catalyst is an excellent choice for plastic waste pyrolysis. Abadi et al. suggested that the FCC catalyst is excellent in the PP pyrolysis to boost the output of liquid hydrocarbons beyond 90 wt.% [132]. Sharuddin et al. [17] examined the spent FCC catalyst over the waste plastics and obtained the 80 wt.% liquid hydrocarbons. With the increasing development in pyrolysis to recycle plastic waste, FCC is a widely used catalyst in the pyrolysis of plastic waste. The fluid catalytic cracking unit (FCCU) is used to supply high-octane gasoline to the refinery gasoline pool, which indicates FCC’s potential in the pyrolysis of plastic waste in the near future [133].

4.6.3. Bimetallic Catalyst

Bimetallic catalysts have been used in the pyrolysis reaction to produce liquid hydrocarbons from the plastic waste. Wen et al. conducted a study on the Ni-loaded CNT for polyolefin and found an appropriate result because of the higher concentration of the carbon available in the catalyst [134]. Bimetallic catalysts provide a synergic effect in the pyrolysis process due to high surface area and stability. Chen et al. conducted a study on the Fe–Ni bimetallic modified MCM-41 with 10 wt.% each, resulting in the 49.9 wt.% liquid hydrocarbon and 65.9% single styrene hydrocarbons [135]. A wide surface area helps the plastic to enter the pore structure directly for cracking. Fe metal acts as a foundation for transforming raw materials to styrene, whereas nickel–metal oxides raise the acidity of the catalyst, allowing multi-ring compounds to be processed into a single hydrocarbon structure. Chemisorption is determined by the external surface character and pore size, which may be designed using the synthesis process. In this scenario, the larger the pore width of the FeNi catalyst, the greater the hydrogen desorption. The large pore size of Co/SBA-15 follows the trend of long-chain products [136].
Furthermore, Mo and Ni have various benefits as pyrolysis catalysts, including cheap cost absorbent, excellent stability and performance, large surface area, and simplicity of manufacture [137]. However, the sulfurization procedure influences the bimetallic catalyst’s reactivity. The improved sulfurization process may consistently improve the proportion of crude oil conversion to fuel oil, which now stands at 86.9%. Zhou et al. conducted the study on the Ni–Fe/ZrO2 catalyst, resulting in the excellent degradation of the PS at a temperature of 500 °C [138]. Hence, the bimetallic catalyst has good selective conversion in the pyrolysis process and may improve the catalytic activity. The combined characteristics of Ni–Fe might lower the water gas shift reaction and other reforming reaction activation energies [139].

5. Physicochemical Properties of the Plastic Oil with and without Catalyst

Table 4 presents the physicochemical properties of the liquid hydrocarbons produced from municipal plastic waste with and without a catalyst. The physicochemical properties of the fuel play a vital role in its utilization in the transport sector. In recent years, conventional fuels such as gasoline and diesel have been used in engine vehicles. However, continuous work is also being carried out on alternative energy. The conventional fuel density for diesel and gasoline is 0.807 gcm−3 and 0.780 gcm−3,, respectively, which is in line with fuels produced with zeolite and kaolin catalysts of 0.871 and 0.800 gcm−3, respectively [140]. Herein, in a combustion chamber, the kinematic viscosity of the fuel appears as a spray pattern and atomization behavior. Too-high viscous oil result in poor fuel atomization and leakage at the fuel injector, respectively, resulting in poor engine performance [141,142]. Therefore, the kinematic viscosity of the fuel should be lower for proper atomization and complete combustion of the fuel in a cylinder engine. Table 4 presents that pyrolysis with a zeolite catalyst obtained a 1.99 cSt lower than other catalysts and close to the diesel viscosity.
Calorific value is the essential characteristic of the obtained liquid hydrocarbons. The calorific value of the liquid hydrocarbon obtained with zeolite and kaolin is 46.67 MJ/kg and 46.47 MJ kg−1, close to the conventional fuel shown in Table 4. Moreover, studies have shown that higher calorific values consume less fuel than lower calorific values, leading to complete combustion in the cylinder and improving the power efficiency of the engine [143,144]. However, the higher pour point of the plastic liquid hydrocarbons lowers the flow characteristic because it may cause the wax to form in the cylinder and make it troublesome in the engine. However, plastic pyrolysis with kaolin clay obtained the pour point at a very low temperature of 18 °C. The flashpoint of the liquid hydrocarbon is very low compared to conventional fuel, indicating the safety hazards of liquid hydrocarbons in terms of explosion and fire during fuel transportation [145,146].
Hence, it is concluded from the table that liquid hydrocarbons have the same characteristics as conventional fuel so liquid hydrocarbons may be used with the blending of diesel and some additives for use in transportation.
Table 4. Physicochemical properties of the liquid hydrocarbons.
Table 4. Physicochemical properties of the liquid hydrocarbons.
Physical Properties No Catalyst
[13]
FCC [118]Natural Zeolite [147]Kaolin
[44]
Silica–Alumina
[148]
Fly Ash [149]Calcium Bentonite
[130]
Gasoline
[23]
Diesel
[23]
Density @ 15 °C (g/cm3)0.8600.7520.8680.8000.7700.800na0.7800.807
Viscosity (cSt)2.482.2732.1912.722.210.1452.321.171.94.1
Calorific Value (MJ/kg)40.42 43.2845.5846.479na41.9343.2842.543.5
Octane Number nananananana 81–85na
Pour Point (°C)18−1124−18nana na6
Flash Point (°C)3529<1040nana 4252
Aniline Point (°C)60nanananana 7177.5
API Gravity@ 60 °F38.1nanananana 5538
Diesel Index nanananana nana
Moisture % vol2.4nanananana 0.4–0.50.1–0.3

6. Byproducts of the Catalytic Pyrolysis Process

The pyrolysis process of municipal solid plastic waste contains byproducts such as solid and gaseous hydrocarbons. Effects of catalyst on byproducts are described in the following.

6.1. Catalyst Effect on Solid Residue and Its Potential

Solid residue as char is unburnt plastic leftover in the pyrolysis reactor. The char formation depends on the temperature, heating rate, and residence time: the lower the value, the more significant the proportion of char in the process [150]. The state-of-the-art has reported that char is comparatively lower than the liquid and gaseous hydrocarbons produced in the process [112,113,114]. The presence of a catalyst increases the hydrocarbon ratio in the char residue. However, char deposition on the catalyst reduces the effectivity of the reaction, which may cause a lower pyrolysis reaction and higher production of ash content as byproducts [151,152,153]. Jamradloedluk et al. [154] performed the pyrolysis of HDPE 400–450 °C, and char (solid residues) obtained were collected and analyzed. Proximate and ultimate analyses showed that pyrolysis char had a large amount of volatile matter (51.40%) and fixed carbon (46.03%), a small amount of moisture (2.41%), and little amount of ash (61.0%). Char-derived briquette showed good potential as a solid fuel for heating application. Some studies showed that it can also be used as a heavy metal absorbent in industrial wastewater and toxic gases or graphene production [56,154].

6.2. Catalytic Effect on Gaseous Hydrocarbons and Its Potential

Gaseous hydrocarbons are the byproducts of the pyrolysis process (Figure 4). Production of the gaseous hydrocarbon depends on the type of plastic and temperature. Polyolefin and polystyrene yielded gaseous hydrocarbons in the 5–20 wt.% range. Pyrolysis of the mixed plastic waste at temperature of 350 °C produced more gaseous products than the individual waste plastics and increased by 8.5 wt.% upon increasing the temperature. It was found that 1 kg of feedstock produced 13–26 wt.% of gaseous products. In addition, the catalyst has a higher impact on the production of gaseous hydrocarbons than liquid hydrocarbons. It increases the gaseous product due to the higher cracking process but lowers the yield of the liquid hydrocarbons. It improves the quality of the liquid hydrocarbons [55]. In conventional pyrolysis of municipal plastic waste, CO, CO2, hydrogen, methane, ethane, propane, butane, and other gaseous hydrocarbons are a source of gaseous products from the pyrolysis process [13,155]. The presence of catalyst leads to the decarboxylation route, which may reduce the concentration of carbon monoxide and increase the carbon dioxide in the gaseous products. However, the state-of-the-art shows that the most significant impact of kaolin and zeolite on the produced gases raises the hydrogen concentration and carbon frame isomerization, which increases i-butane concentration [74,93].
The gases produced by plastic garbage have a significant calorific value. Gases generated by the pyrolysis of agricultural plastic waste include 45.9 and 46.6 MJ kg−1 HHV, respectively [108]. Moreover, the HHV of gases generated by PP and PE pyrolysis is 50 and 42 MJ kg−1, respectively [156]. In addition, there are comparable findings for HHV of waste tires pyrolysis (45 MJ kg−1) [157]. As a result, the generated gases can be used in boilers for heating or in gas turbines for power generation. Moreover, due to their composition, 1-butene and isoprene may be recovered by condensation and employed in the manufacturing of tires. Following separation from other gases, propane and ethane may be used as a chemical feedstock to make polyolefin.

7. Future Challenges

Thermal pyrolysis of municipal mixed plastic waste is a common technology to convert it into liquid, solid, and gaseous hydrocarbons having good potential applications in various fuel industries. However, to achieve more energy and viable products from the pyrolysis process, there is a need for additional research and development for further commercialization.
These are some points that need to be focused on for development:
  • Design and development of the cost-effective and highly efficient pyrolysis reactor.
  • Acquaint the main waste plastic pyrolysis reactor and its process.
  • Understand the constraints and opportunities for improving product quality and yield via plastic pyrolysis.
  • Development in the synthesis of the low-cost catalyst to enhance the plastic oil yield with upgradation.
  • Development of both rapid pyrolysis and plastic-oil upgrading as long as both are aimed at producing useful and valuable products.
  • Post-pyrolysis processing to increase the plastic-oil characteristics of the product.
At present, plastic waste technology is being adopted or is in the initial stage of implementation in various countries such as India, Taiwan, Japan, Thailand, and Malaysia. However, advancements in pyrolysis research are needed for scaling up and commercialization. Therefore, private companies can come forward to take initiatives and provide proficient services of pyrolysis to their customers. Hence, there is a possible chance that the plastic pyrolysis technology program may come under the national-level waste management program in developed and developing countries in the forthcoming years due to the evident benefits.

8. Conclusions

Pyrolysis can convert municipal mixed plastic waste into liquid, solid, and gaseous hydrocarbons. However, produced products have significant calorific value and are likely to be used by recycling and energy industries. The quality of the liquid fuel is affected greatly by the catalyst utilized. To produce liquid products suitable for replacing fossil fuels, particularly gasoline and diesel, low-acidity catalysts with high conversion rates are required. As a result, a bimetallic catalyst such as Ni, Fe, or a Fe–Ni combination may be a supported catalyst to improve the yield and quality of liquid hydrocarbons. Variables including plastic-type, contamination, reactors, pyrolysis, and residence time can also affect liquid hydrocarbon output. They may influence the catalytic pyrolysis mechanism, product yield, and distribution. These influences must be properly identified and handled to ensure process sustainability. Although fast pyrolysis and auger reactors may address product distribution issues, auger pyrolyzers have a simple design and alleviate some of the challenges associated with transferring heat for pyrolyzing plastic waste. The mechanical forces associated with auger reactors improve particle mixing and heat transport, both of which are essential for pyrolysis performance, although they require more investigation to determine their practicality in municipal plastic waste recycling. This review focuses on the pyrolysis of municipal mixed plastic waste in several reactors. However, there are other factors to consider when studying MPW pyrolysis. We should consider using inexpensive catalysts such as clay, red mud, cement, sand, and pyrolysis reactors for fast pyrolysis. Studying MPW pyrolysis kinetics is also important for process design. More study is required to fully understand polymer pyrolysis. More research is needed to fully comprehend MPW pyrolysis.

Author Contributions

Conceptualization, S.P. (Shashank Pal) and A.P.; methodology, S.P. (Shashank Pal), A.K.S. and P.K.G.; software, A.K.; investigation, S.P. (Shashank Pal) and A.K.; writing—original draft preparation, S.P. (Shyam Pandey), A.K. and A.K.S.; writing—review and editing, A.P. and A.K.S.; supervision, A.P. All authors have read and agreed to the published version of the manuscript.

Funding

Not funded by any agencies.

Acknowledgments

Authors are very thankful to the R&D department UPES for utilizing different facilities and software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaewbuddee, C.; Sukjit, E.; Srisertpol, J.; Maithomklang, S.; Wathakit, K.; Klinkaew, N.; Liplap, P.; Arjharn, W. Evaluation of Waste Plastic Oil-Biodiesel Blends as Alternative Fuels for Diesel Engines. Energies 2020, 13, 2823. [Google Scholar] [CrossRef]
  2. Jaganmohan, M. Fuel Market Consumption Volume in India 2011–2020; Arfica Rice Center: Cotonou, Benin, 2011. [Google Scholar]
  3. Pathak, S.K.; Nayyar, A.; Goel, V. Optimization of EGR Effects on Performance and Emission Parameters of a Dual Fuel (Diesel + CNG) CI Engine: An Experimental Investigation. Fuel 2021, 291, 120183. [Google Scholar] [CrossRef]
  4. Simsek, S.; Uslu, S.; Simsek, H.; Uslu, G. Improving the Combustion Process by Determining the Optimum Percentage of Liquefied Petroleum Gas (LPG) via Response Surface Methodology (RSM) in a Spark Ignition (SI) Engine Running on Gasoline-LPG Blends. Fuel Process. Technol. 2021, 221, 106947. [Google Scholar] [CrossRef]
  5. Atilhan, S.; Park, S.; El-Halwagi, M.M.; Atilhan, M.; Moore, M.; Nielsen, R.B. Green Hydrogen as an Alternative Fuel for the Shipping Industry. Curr. Opin. Chem. Eng. 2021, 31, 100668. [Google Scholar] [CrossRef]
  6. Saravanan, S.; Krishnamoorthy, N. Investigation on Reduction in Consequences of Adding Antioxidants into the Algae Biodiesel Blend as a CI Engine Fuel. Fuel 2020, 276, 117993. [Google Scholar] [CrossRef]
  7. Jha, K.K.; Kannan, T.T.M. Recycling of Plastic Waste into Fuel by Pyrolysis—A Review. Mater. Today Proc. 2020, 37, 3718–3720. [Google Scholar] [CrossRef]
  8. Khan, M.I.; Yasmin, T.; Shakoor, A. Technical Overview of Compressed Natural Gas (CNG) as a Transportation Fuel. Renew. Sustain. Energy Rev. 2015, 51, 785–797. [Google Scholar] [CrossRef]
  9. Chintala, V.; Subramanian, K.A. A Comprehensive Review on Utilization of Hydrogen in a Compression Ignition Engine under Dual Fuel Mode. Renew. Sustain. Energy Rev. 2017, 70, 472–491. [Google Scholar] [CrossRef]
  10. Kumar Sharma, A.; Kumar Ghodke, P.; Manna, S.; Chen, W.-H. Emerging Technologies for Sustainable Production of Biohydrogen Production from Microalgae: A State-of-the-Art Review of Upstream and Downstream Processes. Bioresour. Technol. 2021, 342, 126057. [Google Scholar] [CrossRef]
  11. Chen, H.; Xie, B.; Ma, J.; Chen, Y. NOx Emission of Biodiesel Compared to Diesel: Higher or Lower? Appl. Therm. Eng. 2018, 137, 584–593. [Google Scholar] [CrossRef]
  12. Armenise, S.; SyieLuing, W.; Ramírez-Velásquez, J.M.; Launay, F.; Wuebben, D.; Ngadi, N.; Rams, J.; Muñoz, M. Plastic Waste Recycling via Pyrolysis: A Bibliometric Survey and Literature Review. J. Anal. Appl. Pyrolysis 2021, 158, 105265. [Google Scholar] [CrossRef]
  13. Ghodke, P.K. High-Quality Hydrocarbon Fuel Production from Municipal Mixed Plastic Waste Using a Locally Available Low-Cost Catalyst. Fuel Commun. 2021, 8, 100022. [Google Scholar] [CrossRef]
  14. Sarkar, A.; Praveen, G. Utilization of Waste Biomass into Useful Forms of Energy. In Biofuels and Bioenergy; Springer: Cham, Switzerland, 2017; pp. 117–132. [Google Scholar]
  15. Silveira, A.V.M.; Cella, M.; Tanabe, E.H.; Bertuol, D.A. Application of Tribo-Electrostatic Separation in the Recycling of Plastic Wastes. Process Saf. Environ. Prot. 2018, 114, 219–228. [Google Scholar] [CrossRef]
  16. Bonifazi, G.; Capobianco, G.; Serranti, S. A Hierarchical Classification Approach for Recognition of Low-Density (LDPE) and High-Density Polyethylene (HDPE) in Mixed Plastic Waste Based on Short-Wave Infrared (SWIR) Hyperspectral Imaging. Spectrochim. Acta—Part A Mol. Biomol. Spectrosc. 2018, 198, 115–122. [Google Scholar] [CrossRef]
  17. Anuar Sharuddin, S.D.; Abnisa, F.; Wan Daud, W.M.A.; Aroua, M.K. A Review on Pyrolysis of Plastic Wastes. Energy Convers. Manag. 2016, 115, 308–326. [Google Scholar] [CrossRef]
  18. Ghodke, P.K.; Sharma, A.K.; Pandey, J.K.; Chen, W.-H.; Patel, A.; Ashokkumar, V. Pyrolysis of Sewage Sludge for Sustainable Biofuels and Value-Added Biochar Production. J. Environ. Manag. 2021, 298, 113450. [Google Scholar] [CrossRef]
  19. Donaj, P.J.; Kaminsky, W.; Buzeto, F.; Yang, W. Pyrolysis of Polyolefins for Increasing the Yield of Monomers’ Recovery. Waste Manag. 2012, 32, 840–846. [Google Scholar] [CrossRef]
  20. Onwudili, J.A.; Insura, N.; Williams, P.T. Composition of Products from the Pyrolysis of Polyethylene and Polystyrene in a Closed Batch Reactor: Effects of Temperature and Residence Time. J. Anal. Appl. Pyrolysis 2009, 86, 293–303. [Google Scholar] [CrossRef]
  21. Marcilla, A.; Beltrán, M.I.; Navarro, R. Thermal and Catalytic Pyrolysis of Polyethylene over HZSM5 and HUSY Zeolites in a Batch Reactor under Dynamic Conditions. Appl. Catal. B Environ. 2009, 86, 78–86. [Google Scholar] [CrossRef]
  22. Fakhrhoseini, S.M.; Dastanian, M. Predicting Pyrolysis Products of PE, PP, and PET Using NRTL Activity Coefficient Model. J. Chem. 2013, 2013, 7–9. [Google Scholar] [CrossRef]
  23. Ahmad, I.; Ismail Khan, M.; Khan, H.; Ishaq, M.; Tariq, R.; Gul, K.; Ahmad, W. Pyrolysis Study of Polypropylene and Polyethylene Into Premium Oil Products. Int. J. Green Energy 2015, 12, 663–671. [Google Scholar] [CrossRef]
  24. Cepeliogullar, O.; Putun, A.E. Utilization of Two Different Types of Plastic Wastes from Daily and Industrial Life. J. Selcuk Univ. Nat. Appl. Sci. 2013, 2, 694–706. [Google Scholar]
  25. Miranda, R.; Yang, J.; Roy, C.; Vasile, C. Vacuum Pyrolysis of PVC I. Kinetic Study. Polym. Degrad. Stab. 1999, 64, 127–144. [Google Scholar] [CrossRef]
  26. Chintala, V.; Godkhe, P.; Phadtare, S.; Tadpatrikar, M.; Pandey, J.K.; Kumar, S. A Comparative Assessment of Single Cylinder Diesel Engine Characteristics with Plasto-Oils Derived from Municipal Mixed Plastic Waste. Energy Convers. Manag. 2018, 166, 579–589. [Google Scholar] [CrossRef]
  27. Al-Salem, S.M.; Yang, Y.; Wang, J.; Leeke, G.A. Pyro-Oil and Wax Recovery from Reclaimed Plastic Waste in a Continuous Auger Pyrolysis Reactor. Energies 2020, 13, 2040. [Google Scholar] [CrossRef] [Green Version]
  28. Singh, R.K.; Ruj, B. Time and Temperature Depended Fuel Gas Generation from Pyrolysis of Real World Municipal Plastic Waste. Fuel 2016, 174, 164–171. [Google Scholar] [CrossRef]
  29. Suriapparao, D.V.; Kumar, D.A.; Vinu, R. Microwave Co-Pyrolysis of PET Bottle Waste and Rice Husk: Effect of Plastic Waste Loading on Product Formation. Sustain. Energy Technol. Assess. 2022, 49, 101781. [Google Scholar] [CrossRef]
  30. Chattopadhyay, J.; Pathak, T.S.; Srivastava, R.; Singh, A.C. Catalytic Co-Pyrolysis of Paper Biomass and Plastic Mixtures (HDPE (High Density Polyethylene), PP (Polypropylene) and PET (Polyethylene Terephthalate)) and Product Analysis. Energy 2016, 103, 513–521. [Google Scholar] [CrossRef]
  31. Bhoi, P.R.; Rahman, M.H. Hydrocarbons Recovery through Catalytic Pyrolysis of Compostable and Recyclable Waste Plastics Using a Novel Desk-Top Staged Reactor. Environ. Technol. Innov. 2022, 27, 102453. [Google Scholar] [CrossRef]
  32. Rahman, M.H.; Bhoi, P.R.; Saha, A.; Patil, V.; Adhikari, S. Thermo-Catalytic Co-Pyrolysis of Biomass and High-Density Polyethylene for Improving the Yield and Quality of Pyrolysis Liquid. Energy 2021, 225, 120231. [Google Scholar] [CrossRef]
  33. Wijayanti, H.; Irawan, C.; Aulia, N. Copyrolysis of Rice Husk and Plastic Bags Waste from Low-Density Polyethylene (LDPE) for Improving Pyrolysis Liquid Product. IOP Conf. Ser. Earth Environ. Sci. 2022, 963, 012012. [Google Scholar] [CrossRef]
  34. Janajreh, I.; Adeyemi, I.; Elagroudy, S. Gasification Feasibility of Polyethylene, Polypropylene, Polystyrene Waste and Their Mixture: Experimental Studies and Modeling. Sustain. Energy Technol. Assess. 2020, 39, 100684. [Google Scholar] [CrossRef]
  35. Hakeem, I.G.; Aberuagba, F.; Musa, U. Catalytic Pyrolysis of Waste Polypropylene Using Ahoko Kaolin from Nigeria. Appl. Petrochem. Res. 2018, 8, 203–210. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, X.; Burra, K.G.; Wang, Z.; Li, J.; Che, D.; Gupta, A.K. On Deconvolution for Understanding Synergistic Effects in Co-Pyrolysis of Pinewood and Polypropylene. Appl. Energy 2020, 279, 115811. [Google Scholar] [CrossRef]
  37. Bai, B.; Wang, W.; Jin, H. Experimental Study on Gasification Performance of Polypropylene (PP) Plastics in Supercritical Water. Energy 2020, 191, 116527. [Google Scholar] [CrossRef]
  38. Zhao, S.; Wang, C.; Bai, B.; Jin, H.; Wei, W. Study on the Polystyrene Plastic Degradation in Supercritical Water/CO2 Mixed Environment and Carbon Fixation of Polystyrene Plastic in CO2 Environment. J. Hazard. Mater. 2022, 421, 126763. [Google Scholar] [CrossRef] [PubMed]
  39. Ahmad, N.; Ahmad, N.; Maafa, I.M.; Ahmed, U.; Akhter, P.; Shehzad, N.; Amjad, U.E.S.; Hussain, M. Thermal Conversion of Polystyrene Plastic Waste to Liquid Fuel via Ethanolysis. Fuel 2020, 279, 118498. [Google Scholar] [CrossRef]
  40. Wu, Y.; Wang, G.; Zhu, J.; Wang, Y.; Yang, H.; Jin, L.; Hu, H. Insight into Synergistic Effect of Co-Pyrolysis of Low-Rank Coal and Waste Polyethylene with or without Additives Using Rapid Infrared Heating. J. Energy Inst. 2022, 102, 384–394. [Google Scholar] [CrossRef]
  41. Duque, J.V.F.; Martins, M.F.; Debenest, G.; Orlando, M.T.D.A. The Influence of the Recycling Stress History on LDPE Waste Pyrolysis. Polym. Test. 2020, 86, 106460. [Google Scholar] [CrossRef]
  42. Wei, Y.; Fakudze, S.; Zhang, Y.; Ma, R.; Shang, Q.; Chen, J.; Liu, C.; Chu, Q. Co-Hydrothermal Carbonization of Pomelo Peel and PVC for Production of Hydrochar Pellets with Enhanced Fuel Properties and Dechlorination. Energy 2022, 239, 122350. [Google Scholar] [CrossRef]
  43. Wang, Q.; Zhang, J.; Wang, G.; Wang, H.; Sun, M. Thermal and Kinetic Analysis of Coal with Different Waste Plastics (PVC) in Cocombustion. Energy Fuels 2018, 32, 2145–2155. [Google Scholar] [CrossRef]
  44. Mansor, W.N.W.; Razali, N.A.; Abdullah, S.; Jarkoni, M.N.K.; Sharin, A.B.E.; Abd Kadir, N.H.; Ramli, A.; Chao, H.R.; Lin, S.L.; Jalaludin, J. A Review of Plastic-Derived Diesel Fuel as a Renewable Fuel for Internal Combustion Engines: Application, Challanges and Global Potential 2021. IOP Conf. Ser. Earth Environ. Sci. 2022, 1013, 012014. [Google Scholar] [CrossRef]
  45. Kasar, P.; Sharma, D.K.; Ahmaruzzaman, M. Thermal and Catalytic Decomposition of Waste Plastics and Its Co-Processing with Petroleum Residue through Pyrolysis Process. J. Clean. Prod. 2020, 265, 121639. [Google Scholar] [CrossRef]
  46. Mortezaeikia, V.; Tavakoli, O. Thermogravimetric Kinetics of Catalytic and Non-Catalytic Microwave-Assisted Pyrolysis of Waste Abs Plastic. SSRN Electron. J. 2022. [Google Scholar] [CrossRef]
  47. Lazzarotto, I.P.; Ferreira, S.D.; Junges, J.; Bassanesi, G.R.; Manera, C.; Perondi, D.; Godinho, M. The Role of CaO in the Steam Gasification of Plastic Wastes Recovered from the Municipal Solid Waste in a Fluidized Bed Reactor. Process Saf. Environ. Prot. 2020, 140, 60–67. [Google Scholar] [CrossRef]
  48. PCB Annual Report for the Year 2018–19 on Implementation of Plastic Waste Management Rules. Cent. Pollut. Control. Board 2019, 17, 1–18.
  49. Global Plastics Production Forecast 2025–2050; Statista: Singapore, 2022.
  50. Kaminsky, W.; Schmidt, H.; Simon, C.M. Recycling of Mixed Plastics by Pyrolysis in a Fluidised Bed. Macromol. Symp. 2000, 152, 191–199. [Google Scholar] [CrossRef]
  51. Obeid, F.; Zeaiter, J.; Al-Muhtaseb, A.H.; Bouhadir, K. Thermo-Catalytic Pyrolysis of Waste Polyethylene Bottles in a Packed Bed Reactor with Different Bed Materials and Catalysts. Energy Convers. Manag. 2014, 85, 1–6. [Google Scholar] [CrossRef]
  52. Singh, R.K.; Ruj, B.; Sadhukhan, A.K.; Gupta, P. Conventional Pyrolysis of Plastic Waste for Product Recovery and Utilization of Pyrolytic Gases for Carbon Nanotubes Production. Environ. Sci. Pollut. Res. 2022, 29, 20007–20016. [Google Scholar] [CrossRef]
  53. Chintala, V.; Kumar, S.S.; Pandey, J.K.J.K.; Sharma, A.K.A.K.; Kumar, S.S. Solar Thermal Pyrolysis of Non-Edible Seeds to Biofuels and Their Feasibility Assessment. Energy Convers. Manag. 2017, 153, 482–492. [Google Scholar] [CrossRef]
  54. Suriapparao, D.V.; Boruah, B.; Raja, D.; Vinu, R. Microwave Assisted Co-Pyrolysis of Biomasses with Polypropylene and Polystyrene for High Quality Bio-Oil Production. Fuel Process. Technol. 2018, 175, 64–75. [Google Scholar] [CrossRef]
  55. Sivagami, K.; Divyapriya, G.; Selvaraj, R. Catalytic Pyrolysis of Polyolefin and Multilayer Packaging Based Waste Plastics: A Pilot Scale Study. Process Saf. Environ. Prot. 2021, 149, 497–506. [Google Scholar] [CrossRef]
  56. Rajendran, K.M.; Chintala, V.; Sharma, A.; Pal, S.; Pandey, J.K.J.K.; Ghodke, P.; Moorthy Rajendran, K.; Chintala, V.; Sharma, A.; Pal, S.; et al. Review of Catalyst Materials in Achieving the Liquid Hydrocarbon Fuels from Municipal Mixed Plastic Waste (MMPW). Mater. Today Commun. 2020, 24, 100982. [Google Scholar] [CrossRef]
  57. Sipra, A.T.; Gao, N.; Sarwar, H. Municipal Solid Waste (MSW) Pyrolysis for Bio-Fuel Production: A Review of Effects of MSW Components and Catalysts. Fuel Process. Technol. 2018, 175, 131–147. [Google Scholar] [CrossRef]
  58. Lu, J.S.; Chang, Y.; Poon, C.S.; Lee, D.J. Slow Pyrolysis of Municipal Solid Waste (MSW): A Review. Bioresour. Technol. 2020, 312, 123615. [Google Scholar] [CrossRef]
  59. Sogancioglu, M.; Yel, E.; Ahmetli, G. Pyrolysis of Waste High Density Polyethylene (HDPE) and Low Density Polyethylene (LDPE) Plastics and Production of Epoxy Composites with Their Pyrolysis Chars. J. Clean. Prod. 2017, 165, 369–381. [Google Scholar] [CrossRef]
  60. Harussani, M.M.; Sapuan, S.M.; Rashid, U.; Khalina, A.; Ilyas, R.A. Pyrolysis of Polypropylene Plastic Waste into Carbonaceous Char: Priority of Plastic Waste Management amidst COVID-19 Pandemic. Sci. Total Environ. 2022, 803, 149911. [Google Scholar] [CrossRef]
  61. Das, P.; Tiwari, P. The Effect of Slow Pyrolysis on the Conversion of Packaging Waste Plastics (PE and PP) into Fuel. Waste Manag. 2018, 79, 615–624. [Google Scholar] [CrossRef] [PubMed]
  62. Singh, R.K.; Ruj, B.; Sadhukhan, A.K.; Gupta, P. Impact of Fast and Slow Pyrolysis on the Degradation of Mixed Plastic Waste: Product Yield Analysis and Their Characterization. J. Energy Inst. 2019, 92, 1647–1657. [Google Scholar] [CrossRef]
  63. Hussein, Z.A.; Shakor, Z.M.; Alzuhairi, M.; Al-Sheikh, F. Thermal and Catalytic Cracking of Plastic Waste: A Review. Int. J. Environ. Anal. Chem. 2021, 1–18. [Google Scholar] [CrossRef]
  64. Lubongo, C.; Congdon, T.; McWhinnie, J.; Alexandridis, P. Economic Feasibility of Plastic Waste Conversion to Fuel Using Pyrolysis. Sustain. Chem. Pharm. 2022, 27, 100683. [Google Scholar] [CrossRef]
  65. Sharuddin, S.D.A.; Abnisa, F.; Daud, W.M.A.W.; Aroua, M.K. Pyrolysis of Plastic Waste for Liquid Fuel Production as Prospective Energy Resource. IOP Conf. Ser. Mater. Sci. Eng. 2018, 334, 012001. [Google Scholar] [CrossRef]
  66. Kataki, R.; Bordoloi, N.J.; Saikia, R.; Sut, D.; Narzari, R.; Gogoi, L.; Bhuyan, N. Waste Valorization to Fuel and Chemicals Through Pyrolysis: Technology, Feedstock, Products, and Economic Analysis. In Waste to Wealth; Springer: Singapore, 2018; pp. 477–514. [Google Scholar] [CrossRef]
  67. Fang, S.; Gu, W.; Chen, L.; Yu, Z.; Dai, M.; Lin, Y.; Liao, Y.; Ma, X. Ultrasonic Pretreatment Effects on the Co-Pyrolysis of Municipal Solid Waste and Paper Sludge through Orthogonal Test. Bioresour. Technol. 2018, 258, 5–11. [Google Scholar] [CrossRef] [PubMed]
  68. Hafeez, S.; Pallari, E.; Manos, G.; Constantinou, A. Catalytic Conversion and Chemical Recovery; Elsevier Inc.: Amsterdam, The Netherlands, 2018; ISBN 9780128131404. [Google Scholar]
  69. Chen, W.H.; Cheng, C.L.; Lee, K.T.; Lam, S.S.; Ong, H.C.; Ok, Y.S.; Saeidi, S.; Sharma, A.K.; Hsieh, T.H. Catalytic Level Identification of ZSM-5 on Biomass Pyrolysis and Aromatic Hydrocarbon Formation. Chemosphere 2021, 271, 129510. [Google Scholar] [CrossRef] [PubMed]
  70. Muhammad, C.; Onwudili, J.A.; Williams, P.T. Thermal Degradation of Real-World Waste Plastics and Simulated Mixed Plastics in a Two-Stage Pyrolysis–Catalysis Reactor for Fuel Production. Energy Fuels 2015, 29, 2601–2609. [Google Scholar] [CrossRef]
  71. Zhou, N.; Dai, L.; Lyu, Y.; Wang, Y.; Li, H.; Cobb, K.; Chen, P.; Lei, H.; Ruan, R. A Structured Catalyst of ZSM-5/SiC Foam for Chemical Recycling of Waste Plastics via Catalytic Pyrolysis. Chem. Eng. J. 2022, 440, 135836. [Google Scholar] [CrossRef]
  72. Mangesh, V.L.; Padmanabhan, S.; Tamizhdurai, P.; Ramesh, A. Experimental Investigation to Identify the Type of Waste Plastic Pyrolysis Oil Suitable for Conversion to Diesel Engine Fuel. J. Clean. Prod. 2020, 246, 119066. [Google Scholar] [CrossRef]
  73. Chen, D.; Yin, L.; Wang, H.; He, P. Pyrolysis Technologies for Municipal Solid Waste: A Review. Waste Manag. 2014, 34, 2466–2486. [Google Scholar] [CrossRef] [PubMed]
  74. Onwudili, J.A.; Muhammad, C.; Williams, P.T. 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. 2019, 92, 1337–1347. [Google Scholar] [CrossRef]
  75. Muneer, B.; Zeeshan, M.; Qaisar, S.; Razzaq, M.; Iftikhar, H. Influence of In-Situ and Ex-Situ HZSM-5 Catalyst on Co-Pyrolysis of Corn Stalk and Polystyrene with a Focus on Liquid Yield and Quality. J. Clean. Prod. 2019, 237, 117762. [Google Scholar] [CrossRef]
  76. Scheirs, J. Overview of Commercial Pyrolysis Processes for Waste Plastics. In Feedstock Recycling and Pyrolysis of Waste Plastics: Converting Waste Plastics into Diesel and Other Fuels; John Wiley: Hoboken, NJ, USA, 2006; pp. 381–433. [Google Scholar] [CrossRef]
  77. Mahdavi, H.; Eden, N.T.; Doherty, C.M.; Acharya, D.; Smith, S.J.D.; Mulet, X.; Hill, M.R. Underlying Polar and Nonpolar Modification MOF-Based Factors That Influence Permanent Porosity in Porous Liquids. ACS Appl. Mater. Interfaces 2022, 14, 23392–23399. [Google Scholar] [CrossRef] [PubMed]
  78. Kalargaris, I.; Tian, G.; Gu, S. The Utilisation of Oils Produced from Plastic Waste at Different Pyrolysis Temperatures in a DI Diesel Engine. Energy 2017, 131, 179–185. [Google Scholar] [CrossRef]
  79. Mastral, F.J.; Esperanza, E.; Garciía, P.; Juste, M. Pyrolysis of High-Density Polyethylene in a Fluidised Bed Reactor. Influence of the Temperature and Residence Time. J. Anal. Appl. Pyrolysis 2002, 63, 1–15. [Google Scholar] [CrossRef]
  80. Susastriawan, A.A.P.; Sandria, A. Experimental Study the Influence of Zeolite Size on Low-Temperature Pyrolysis of Low-Density Polyethylene Plastic Waste. Therm. Sci. Eng. Prog. 2020, 17, 100497. [Google Scholar] [CrossRef]
  81. Adnan; Shah, J.; Jan, M.R. Effect of Polyethylene Terephthalate on the Catalytic Pyrolysis of Polystyrene: Investigation of the Liquid Products. J. Taiwan Inst. Chem. Eng. 2015, 51, 96–102. [Google Scholar] [CrossRef]
  82. Osman, Y.; Yansaneh, S.H.Z. Recent Advances on Waste Plastic Thermal Pyrolysis: A Critical Overview. Processes 2022, 10, 332. [Google Scholar]
  83. Lopez, G.; Amutio, M.; Elordi, G.; Artetxe, M.; Altzibar, H.; Olazar, M. A Conical Spouted Bed Reactor for the Valorisation of Waste Tires. In Proceedings of the 13th International Conference on Fluidization—New Paradigm in Fluidization Engineering, Gyeong-Ju, Korea, 16–21 May 2010. [Google Scholar]
  84. Krishnamoorthy, V.; Dhanasekaran, R.; Rana, D.; Saravanan, S.; Rajesh Kumar, B. A Comparative Assessment of Ternary Blends of Three Bio-Alcohols with Waste Cooking Oil and Diesel for Optimum Emissions and Performance in a CI Engine Using Response Surface Methodology. Energy Convers. Manag. 2018, 156, 337–357. [Google Scholar] [CrossRef]
  85. Ghodke, P.; Krishna, S.V. Effect of Heterogeneous Catalyst on Esterification of Pyrolysis Oil. In Springer Proceeding in Energy; Springer: Cham, Switzerland, 2018; pp. 219–229. [Google Scholar]
  86. Kalargaris, I.; Tian, G.; Gu, S. Experimental Evaluation of a Diesel Engine Fuelled by Pyrolysis Oils Produced from Low-Density Polyethylene and Ethylene–Vinyl Acetate Plastics. Fuel Process. Technol. 2017, 161, 125–131. [Google Scholar] [CrossRef]
  87. Al-Salem, S.M.; Antelava, A.; Constantinou, A.; Manos, G.; Dutta, A. A Review on Thermal and Catalytic Pyrolysis of Plastic Solid Waste (PSW). J. Environ. Manag. 2017, 197, 177–198. [Google Scholar] [CrossRef]
  88. Musale, H.K.; Bhattacharyulu, Y.C.; Bhoyar, R.K.; Student, P.G.; Professor, A. Design Consideration Of Pyrolysis Reactor For Production Of Bio-Oil. Int. J. Eng. Trends Technol. 2013, 5, 83–85. [Google Scholar]
  89. Hasan, M.M.; Rasul, M.G.; Khan, M.M.K.; Ashwath, N.; Jahirul, M.I. Energy Recovery from Municipal Solid Waste Using Pyrolysis Technology: A Review on Current Status and Developments. Renew. Sustain. Energy Rev. 2021, 145, 111073. [Google Scholar] [CrossRef]
  90. Fogler, H.S. Elements of Chemical Reaction Engineering; Printice-Hall International Editions: Hoboken, NJ, USA, 2016; ISBN 9780133887518. [Google Scholar]
  91. Palomar-Torres, A.; Torres-Jimenez, E.; Kegl, B.; Bombek, G.; Volmajer-Valh, J.; Lešnik, L. Catalytic Pyrolysis of Plastic Wastes for Liquid Oils’ Production Using ZAP USY Zeolite as a Catalyst. Int. J. Environ. Sci. Technol. 2022, 1–14. [Google Scholar] [CrossRef]
  92. Luo, W.; Fan, Z.; Wan, J.; Hu, Q.; Dong, H.; Zhang, X. Study on the Reusability of Kaolin as Catalysts for Catalytic Pyrolysis of Low-Density Polyethylene. Fuel 2021, 302, 121164. [Google Scholar] [CrossRef]
  93. Mohabeer, C.; Guilhaume, N.; Laurenti, D.; Schuurman, Y. Microwave-Assisted Pyrolysis of Biomass with and without Use of Catalyst in a Fluidised Bed Reactor: A Review. Energies 2022, 15, 3258. [Google Scholar] [CrossRef]
  94. Elordi, G.; Olazar, M.; Lopez, G.; Amutio, M.; Artetxe, M.; Aguado, R.; Bilbao, J. Catalytic Pyrolysis of HDPE in Continuous Mode over Zeolite Catalysts in a Conical Spouted Bed Reactor. J. Anal. Appl. Pyrolysis 2009, 85, 345–351. [Google Scholar] [CrossRef]
  95. Fernandez, E.; Santamaria, L.; Amutio, M.; Artetxe, M.; Arregi, A.; Lopez, G.; Bilbao, J.; Olazar, M. Role of Temperature in the Biomass Steam Pyrolysis in a Conical Spouted Bed Reactor. Energy 2022, 238, 122053. [Google Scholar] [CrossRef]
  96. Arabiourrutia, M.; Elordi, G.; Lopez, G.; Borsella, E.; Bilbao, J.; Olazar, M. Characterization of the Waxes Obtained by the Pyrolysis of Polyolefin Plastics in a Conical Spouted Bed Reactor. J. Anal. Appl. Pyrolysis 2012, 94, 230–237. [Google Scholar] [CrossRef]
  97. Fantozzi, F.; Colantoni, S.; Bartocci, P.; Desideri, U. Rotary Kiln Slow Pyrolysis for Syngas and Char Production From Biomass and Waste. J. Eng. Gas Turbines Power. 2007, 129, 901–907. [Google Scholar] [CrossRef]
  98. Park, K.B.; Choi, M.J.; Chae, D.Y.; Jung, J.; Kim, J.S. Separate Two-Step and Continuous Two-Stage Pyrolysis of a Waste Plastic Mixture to Produce a Chlorine-Depleted Oil. Energy 2022, 244, 122583. [Google Scholar] [CrossRef]
  99. Butler, E.; Devlin, G.; Meier, D.; McDonnell, K. A Review of Recent Laboratory Research and Commercial Developments in Fast Pyrolysis and Upgrading. Renew. Sustain. Energy Rev. 2011, 15, 4171–4186. [Google Scholar] [CrossRef] [Green Version]
  100. Lin, H.T.; Huang, M.S.; Luo, J.W.; Lin, L.H.; Lee, C.M.; Ou, K.L. Hydrocarbon Fuels Produced by Catalytic Pyrolysis of Hospital Plastic Wastes in a Fluidizing Cracking Process. Fuel Process. Technol. 2010, 91, 1355–1363. [Google Scholar] [CrossRef]
  101. Yang, R.-X.; Jan, K.; Chen, C.-T.; Chen, W.-T.; Wu, K.C.-W. Thermochemical Conversion of Plastic Waste into Fuels, Chemicals, and Value-Added Materials: A Critical Review and Outlooks. ChemSusChem 2022, e202200171. [Google Scholar] [CrossRef]
  102. Rasul Jan, M.; Shah, J.; Gulab, H. Catalytic Conversion of Waste High-Density Polyethylene into Useful Hydrocarbons. Fuel 2013, 105, 595–602. [Google Scholar] [CrossRef]
  103. Ruhul, A.M.; Kalam, M.A.; Masjuki, H.H.; Fattah, I.M.R.; Reham, S.S.; Rashed, M.M. State of the Art of Biodiesel Production Processes: A Review of the Heterogeneous Catalyst. RSC Adv. 2015, 5, 101023–101044. [Google Scholar] [CrossRef]
  104. Yansaneh, O.Y.; Zein, S.H. Latest Advances in Waste Plastic Pyrolytic Catalysis. Processes 2022, 10, 683. [Google Scholar] [CrossRef]
  105. Figueiredo, A.L.; Araujo, A.S.; Linares, M.; Peral, Á.; García, R.A.; Serrano, D.P.; Fernandes, V.J. Catalytic Cracking of LDPE over Nanocrystalline HZSM-5 Zeolite Prepared by Seed-Assisted Synthesis from an Organic-Template-Free System. J. Anal. Appl. Pyrolysis 2016, 117, 132–140. [Google Scholar] [CrossRef]
  106. Ubando, A.T.; Chueh, C.-C.; Jin, L.; Belbessai, S.; Azara, A.; Abatzoglou, N. Recent Advances in the Decontamination and Upgrading of Waste Plastic Pyrolysis Products: An Overview. Processes 2022, 10, 733. [Google Scholar] [CrossRef]
  107. Inayat, A.; Klemencova, K.; Grycova, B.; Sokolova, B.; Lestinsky, P. Thermo-Catalytic Pyrolysis of Polystyrene in Batch and Semi-Batch Reactors: A Comparative Study. Waste Manag. Res. 2021, 39, 260–269. [Google Scholar] [CrossRef]
  108. Chen, Z.; Wang, Y.; Sun, Z. Application of CoNi Intercalated Vermiculite Catalyst in Pyrolysis of Plastic. J. Phys. Conf. Ser. 2021, 1885, 032030. [Google Scholar]
  109. Panda, A.K. Thermo-Catalytic Degradation of Different Plastics to Drop in Liquid Fuel Using Calcium Bentonite Catalyst. Int. J. Ind. Chem. 2018, 9, 167–176. [Google Scholar] [CrossRef] [Green Version]
  110. Kumagai, S.; Hasegawa, I.; Grause, G.; Kameda, T.; Yoshioka, T. Thermal Decomposition of Individual and Mixed Plastics in the Presence of CaO or Ca(OH)2. J. Anal. Appl. Pyrolysis 2015, 113, 584–590. [Google Scholar] [CrossRef]
  111. Lopez-Urionabarrenechea, A.; de Marco, I.; Caballero, B.M.; Laresgoiti, M.F.; Adrados, A. Catalytic Stepwise Pyrolysis of Packaging Plastic Waste. J. Anal. Appl. Pyrolysis 2012, 96, 54–62. [Google Scholar] [CrossRef]
  112. Senthil Kumar, P.; Bharathikumar, M.; Prabhakaran, C.; Vijayan, S.; Ramakrishnan, K. Conversion of Waste Plastics into Low-Emissive Hydrocarbon Fuels through Catalytic Depolymerization in a New Laboratory Scale Batch Reactor. Int. J. Energy Environ. Eng. 2017, 8, 167–173. [Google Scholar] [CrossRef] [Green Version]
  113. Kassargy, C.; Awad, S.; Burnens, G.; Kahine, K.; Tazerout, M. Gasoline and Diesel-like Fuel Production by Continuous Catalytic Pyrolysis of Waste Polyethylene and Polypropylene Mixtures over USY Zeolite. Fuel 2018, 224, 764–773. [Google Scholar] [CrossRef]
  114. Sangpatch, T.; Supakata, N.; Kanokkantapong, V.; Jongsomjit, B. Fuel Oil Generated from the Cogon Grass-Derived Al–Si (Imperata cylindrica (L.) Beauv) Catalysed Pyrolysis of Waste Plastics. Heliyon 2019, 5, e02324. [Google Scholar] [CrossRef] [Green Version]
  115. Lei, J.; Yuan, G.; Weerachanchai, P.; Lee, S.W.; Li, K.; Wang, J.Y.; Yang, Y. Investigation on Thermal Dechlorination and Catalytic Pyrolysis in a Continuous Process for Liquid Fuel Recovery from Mixed Plastic Wastes. J. Mater. Cycles Waste Manag. 2018, 20, 137–146. [Google Scholar] [CrossRef]
  116. Costa, P.; Pinto, F.; Mata, R.; Marques, P.; Paradela, F.; Costa, L. Validation of the Application of the Pyrolysis Process for the Treatment and Transformation of Municipal Plastic Wastes. Chem. Eng. Trans. 2021, 86, 859–864. [Google Scholar]
  117. Fekhar, B.; Gombor, L.; Miskolczi, N. Pyrolysis of Chlorine Contaminated Municipal Plastic Waste: In-Situ Upgrading of Pyrolysis Oils by Ni/ZSM-5, Ni/SAPO-11, Red Mud and Ca(OH)2 Containing Catalysts. J. Energy Inst. 2019, 92, 1270–1283. [Google Scholar] [CrossRef]
  118. Zhou, N.; Dai, L.; Lyu, Y.; Li, H.; Deng, W.; Guo, F.; Chen, P.; Lei, H.; Ruan, R. Catalytic Pyrolysis of Plastic Wastes in a Continuous Microwave Assisted Pyrolysis System for Fuel Production. Chem. Eng. J. 2021, 418, 129412. [Google Scholar] [CrossRef]
  119. Hedlund, J.; Zhou, M.; Faisal, A.; Öhrman, O.G.W.; Finelli, V.; Signorile, M.; Crocellà, V.; Grahn, M. Controlling Diffusion Resistance, Selectivity and Deactivation of ZSM-5 Catalysts by Crystal Thickness and Defects. J. Catal. 2022, 410, 320–332. [Google Scholar] [CrossRef]
  120. Ateş, F.; Miskolczi, N.; Borsodi, N. Comparision of Real Waste (MSW and MPW) Pyrolysis in Batch Reactor over Different Catalysts. Part I: Product Yields, Gas and Pyrolysis Oil Properties. Bioresour. Technol. 2013, 133, 443–454. [Google Scholar] [CrossRef]
  121. Fadillah, G.; Fatimah, I.; Sahroni, I.; Musawwa, M.M.; Mahlia, T.M.I.; Muraza, O. Recent Progress in Low-Cost Catalysts for Pyrolysis of Plastic Waste to Fuels. Catalysts 2021, 11, 837. [Google Scholar] [CrossRef]
  122. Wong, S.L.; Ngadi, N.; Abdullah, T.A.T.; Inuwa, I.M. Conversion of Low Density Polyethylene (LDPE) over ZSM-5 Zeolite to Liquid Fuel. Fuel 2017, 192, 71–82. [Google Scholar] [CrossRef]
  123. Sivagami, K.; Kumar, K.V.; Tamizhdurai, P.; Govindarajan, D.; Kumar, M.; Nambi, I. Conversion of Plastic Waste into Fuel Oil Using Zeolite Catalysts in a Bench-Scale Pyrolysis Reactor. RSC Adv. 2022, 12, 7612–7620. [Google Scholar] [CrossRef]
  124. Kim, J.Y.; Heo, S.; Choi, J.W. Effects of Phenolic Hydroxyl Functionality on Lignin Pyrolysis over Zeolite Catalyst. Fuel 2018, 232, 81–89. [Google Scholar] [CrossRef]
  125. Miandad, R.; Barakat, M.A.; Aburiazaiza, A.S.; Rehan, M.; Nizami, A.S. Catalytic Pyrolysis of Plastic Waste: A Review. Process Saf. Environ. Prot. 2016, 102, 822–838. [Google Scholar] [CrossRef]
  126. 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. Catalytic Pyrolysis of Plastic Waste: Moving toward Pyrolysis Based Biorefineries. Front. Energy Res. 2019, 7, 27. [Google Scholar] [CrossRef] [Green Version]
  127. Subhaschandra, T.; Nath, T.; Neeranjan, H. A Lab Scale Waste to Energy Conversion Study for Pyrolysis of Plastic with and without Catalyst: Engine Emissions Testing Study. Fuel 2020, 277, 118176. [Google Scholar] [CrossRef]
  128. Saeaung, K.; Phusunti, N.; Phetwarotai, W.; Assabumrungrat, S.; Cheirsilp, B. Catalytic Pyrolysis of Petroleum-Based and Biodegradable Plastic Waste to Obtain High-Value Chemicals. Waste Manag. 2021, 127, 101–111. [Google Scholar] [CrossRef] [PubMed]
  129. Palos, R.; Rodríguez, E.; Gutiérrez, A.; Bilbao, J.; Arandes, J.M. Cracking of Plastic Pyrolysis Oil over FCC Equilibrium Catalysts to Produce Fuels: Kinetic Modeling. Fuel 2022, 316, 123341. [Google Scholar] [CrossRef]
  130. Aisien, E.T.; Otuya, I.C.; Aisien, F.A. Thermal and Catalytic Pyrolysis of Waste Polypropylene Plastic Using Spent FCC Catalyst. Environ. Technol. Innov. 2021, 22, 101455. [Google Scholar] [CrossRef]
  131. Abbas-Abadi, M.S.; Haghighi, M.N.; Yeganeh, H. Evaluation of Pyrolysis Product of Virgin High Density Polyethylene Degradation Using Different Process Parameters in a Stirred Reactor. Fuel Process. Technol. 2013, 109, 90–95. [Google Scholar] [CrossRef]
  132. Abbas-Abadi, M.S.; Haghighi, M.N.; Yeganeh, H.; McDonald, A.G. Evaluation of Pyrolysis Process Parameters on Polypropylene Degradation Products. J. Anal. Appl. Pyrolysis 2014, 109, 272–277. [Google Scholar] [CrossRef]
  133. Orozco, S.; Artetxe, M.; Lopez, G.; Suarez, M.; Bilbao, J.; Olazar, M. Conversion of HDPE into Value Products by Fast Pyrolysis Using FCC Spent Catalysts in a Fountain Confined Conical Spouted Bed Reactor. ChemSusChem 2021, 14, 4291–4300. [Google Scholar] [CrossRef] [PubMed]
  134. Wen, X.; Chen, X.; Tian, N.; Gong, J.; Liu, J.; Rümmeli, M.H.; Chu, P.K.; Mijiwska, E.; Tang, T. Nanosized Carbon Black Combined with Ni2O3 as “Universal” Catalysts for Synergistically Catalyzing Carbonization of Polyolefin Wastes to Synthesize Carbon Nanotubes and Application for Supercapacitors. Environ. Sci. Technol. 2014, 48, 4048–4055. [Google Scholar] [CrossRef] [PubMed]
  135. Chen, T.; Yu, J.; Ma, C.; Bikane, K.; Sun, L. Catalytic Performance and Debromination of Fe–Ni Bimetallic MCM-41 Catalyst for the Two-Stage Pyrolysis of Waste Computer Casing Plastic. Chemosphere 2020, 248, 125964. [Google Scholar] [CrossRef]
  136. Han, J.; Xiong, Z.; Zhang, Z.; Zhang, H.; Zhou, P.; Yu, F. The Influence of Texture on Co/SBA–15 Catalyst Performance for Fischer–Tropsch Synthesis. Catalysts 2019, 8, 661. [Google Scholar] [CrossRef] [Green Version]
  137. Sridhar, A.; Rahman, M.; Infantes-Molina, A.; Wylie, B.J.; Borcik, C.G.; Khatib, S.J. Bimetallic Mo-Co/ZSM-5 and Mo-Ni/ZSM-5 Catalysts for Methane Dehydroaromatization: A Study of the Effect of Pretreatment and Metal Loadings on the Catalytic Behavior. Appl. Catal. A Gen. 2020, 589, 117247. [Google Scholar] [CrossRef]
  138. Zhou, H.; Saad, J.M.; Li, Q.; Xu, Y. Steam Reforming of Polystyrene at a Low Temperature for High H2/CO Gas with Bimetallic Ni-Fe/ZrO2 Catalyst. Waste Manag. 2020, 104, 42–50. [Google Scholar] [CrossRef] [PubMed]
  139. Wu, P.; Sun, J.; Abbas, M.; Wang, P.; Chen, Y.; Chen, J. Hydrophobic SiO2 Supported Fe-Ni Bimetallic Catalyst for the Production of High-Calorie Synthetic Natural Gas. Appl. Catal. A Gen. 2020, 590, 117302. [Google Scholar] [CrossRef]
  140. Damodharan, D.; Rajesh Kumar, B.; Gopal, K.; De Poures, M.V.; Sethuramasamyraja, B. Utilization of Waste Plastic Oil in Diesel Engines: A Review. Rev. Environ. Sci. Biotechnol. 2019, 18, 681–697. [Google Scholar] [CrossRef]
  141. Sharma, A.K.; Sahoo, P.K.; Singhal, S.; Joshi, G. Exploration of Upstream and Downstream Process for Microwave Assisted Sustainable Biodiesel Production from Microalgae Chlorella Vulgaris. Bioresour. Technol. 2016, 216, 793–800. [Google Scholar] [CrossRef] [PubMed]
  142. Khan, K.; Kumar, G.; Sharma, A.K.; Kumar, P.S.; Mandal, C.; Chintala, V. Performance and Emission Characteristics of a Diesel Engine Using Complementary Blending of Castor and Karanja Biodiesel. Biofuels 2018, 9, 53–60. [Google Scholar] [CrossRef]
  143. Chintala, V.; Sharma, A.K.; Karn, A.; Vardhan, H.; Pandey, J.K. Utilization of Biomass-Derived Pyro-Oils in Compression Ignition (CI) Engines–Recent Developments. Energy Sources Part A Recover. Util. Environ. Eff. 2019, 27, 1–5. [Google Scholar] [CrossRef]
  144. Sharma, A.K.; Sharma, P.K.; Chintala, V.; Khatri, N.; Patel, A. Environment-Friendly Biodiesel/Diesel Blends for Improving the Exhaust Emission and Engine Performance to Reduce the Pollutants Emitted from Transportation Fleets. Int. J. Environ. Res. Public Health 2020, 17, 3896. [Google Scholar] [CrossRef] [PubMed]
  145. Sharma, A.K.; Sahoo, P.K.; Singhal, S. Comparative Evolution of Biomass Production and Lipid Accumulation Potential of Chlorella Species Grown in a Bubble Column Photobioreactor. Biofuels 2016, 7, 389–399. [Google Scholar] [CrossRef]
  146. Sharma, A.K.; Sharma, A.; Singh, Y.; Chen, W.H. Production of a Sustainable Fuel from Microalgae Chlorella Minutissima Grown in a 1500 L Open Raceway Ponds. Biomass Bioenergy 2021, 149, 106073. [Google Scholar] [CrossRef]
  147. Syamsiro, M.; Saptoadi, H.; Norsujianto, T.; Noviasri, P.; Cheng, S.; Alimuddin, Z.; Yoshikawa, K. Fuel Oil Production from Municipal Plastic Wastes in Sequential Pyrolysis and Catalytic Reforming Reactors. Energy Procedia 2014, 47, 180–188. [Google Scholar] [CrossRef] [Green Version]
  148. Choi, S.J.; Park, Y.K.; Jeong, K.E.; Kim, T.W.; Chae, H.J.; Park, S.H.; Jeon, J.K.; Kim, S.S. Catalytic Degradation of Polyethylene over SBA-16. Korean J. Chem. Eng. 2010, 27, 1446–1451. [Google Scholar] [CrossRef]
  149. Nalluri, P.; Prem Kumar, P.; Ch Sastry, M.R. Experimental Study on Catalytic Pyrolysis of Plastic Waste Using Low Cost Catalyst. Mater. Today Proc. 2020, 45, 7216–7221. [Google Scholar] [CrossRef]
  150. Thahir, R.; Altway, A.; Juliastuti, S.R. Susianto Production of Liquid Fuel from Plastic Waste Using Integrated Pyrolysis Method with Refinery Distillation Bubble Cap Plate Column. Energy Rep. 2019, 5, 70–77. [Google Scholar] [CrossRef]
  151. Ratnasari, D.K.; Nahil, M.A.; Williams, P.T. Catalytic Pyrolysis of Waste Plastics Using Staged Catalysis for Production of Gasoline Range Hydrocarbon Oils. J. Anal. Appl. Pyrolysis 2017, 124, 631–637. [Google Scholar] [CrossRef]
  152. Sembiring, F.; Purnomo, C.W.; Purwono, S. Catalytic Pyrolysis of Waste Plastic Mixture Catalytic Pyrolysis of Waste Plastic Mixture. IOP Conf. Ser. Mater. Sci. Eng. 2018, 316, 012020. [Google Scholar] [CrossRef]
  153. Qiu, B.; Deng, N.; Zhang, Y.; Wan, H. Application of Industrial Solid Wastes in Catalytic Pyrolysis. Asia-Pacific J. Chem. Eng. 2018, 13, e2150. [Google Scholar] [CrossRef]
  154. Jamradloedluk, J.; Lertsatitthanakorn, C. Characterization and Utilization of Char Derived from Fast Pyrolysis of Plastic Wastes. Procedia Eng. 2014, 69, 1437–1442. [Google Scholar] [CrossRef] [Green Version]
  155. López, A.; de Marco, I.; Caballero, B.M.; Adrados, A.; Laresgoiti, M.F. Deactivation and Regeneration of ZSM-5 Zeolite in Catalytic Pyrolysis of Plastic Wastes. Waste Manag. 2011, 31, 1852–1858. [Google Scholar] [CrossRef] [PubMed]
  156. Jung, S.H.; Cho, M.H.; Kang, B.S.; Kim, J.S. Pyrolysis of a Fraction of Waste Polypropylene and Polyethylene for the Recovery of BTX Aromatics Using a Fluidized Bed Reactor. Fuel Process. Technol. 2010, 91, 277–284. [Google Scholar] [CrossRef]
  157. Leng, S.; Wang, X.; Wang, J.; Liu, Y.; Ma, F.; Zhong, X. Waste Tire Pyrolysis for the Production of Light Hydrocarbons over Layered Catalysts. Energy Technol. 2015, 3, 851–855. [Google Scholar] [CrossRef]
Figure 2. Development in pyrolysis technique.
Figure 2. Development in pyrolysis technique.
Processes 10 01497 g002
Figure 3. Type of reactor: (a) fixed-bed reactor, (b) fluidized-bed reactor, (c) conical spouted-bed reactor, (d) rotary kiln reactor, (e) auger reactor.
Figure 3. Type of reactor: (a) fixed-bed reactor, (b) fluidized-bed reactor, (c) conical spouted-bed reactor, (d) rotary kiln reactor, (e) auger reactor.
Processes 10 01497 g003
Figure 4. Application of the pyrolysis oil, char, and gaseous hydrocarbons [58].
Figure 4. Application of the pyrolysis oil, char, and gaseous hydrocarbons [58].
Processes 10 01497 g004
Table 1. Proximate analysis of different type of plastic wastes.
Table 1. Proximate analysis of different type of plastic wastes.
Type of PlasticFixed Carbon (%)Moisture Content (%) Ash (%) Volatile Matter (%) References
PET14.5NA0.784.8[29]
PET13.9NANA84.1[30]
HDPENANA0.897.15[30]
HDPE16.85NANA83.15[31]
HDPE0.000.000.0199.99[32]
LDPE0.680.303.3795.61[33]
LDPE0.0510.110.02399.816[34]
PP1.620.164.4593.77[35]
PP1.0NANA96.9[30]
PP0.5NA0.0099.5[36]
PP0.430.290.0099.28[37]
PS1.050.320.0998.54[38]
PS0.220.000.0099.78[39]
PS0.0710.090.02599.814[34]
PE0.07NANA99.93[40]
PE0.000.20.499.4[41]
PVC1.970.650.1197.92[42]
PVC4.100.000.0195.89[43]
PA0.690.000.0099.7[44]
ABS0.040.100.9998.87[45]
ABS9.60.0013.676.8[46]
PBT2.880.160.0097.12[45]
MPW5.340.861.2193.45[13]
MPW3.50.003.393.2[47]
Table 2. Recent developments in the pyrolysis process [57,58].
Table 2. Recent developments in the pyrolysis process [57,58].
Name of
Pyrolysis
Temperature RangeResidence TimeHeating Rate
(°C/s)
Feed Stock Size
(mm)
Liquid
(wt.%)
Solid
(wt.%)
Gases
(wt.%)
Slow
pyrolysis
300–700 °C10–100 min0.1–15–50303535
Fast
pyrolysis
400–800 °C0.5–2 s10–100<340–7015–2510–20
Flash pyrolysis700–1200 °C or above<0.5 s1000<0.210–2010–1560–80
Table 3. Catalytic effect on the liquid, solid, and gaseous hydrocarbons.
Table 3. Catalytic effect on the liquid, solid, and gaseous hydrocarbons.
FeedstockReactor TypeHeating Rate
(°C min−1)
Reaction Temperature (°C)Concentration of CatalystType of CatalystReaction ProductsReferences
Liquid
Hydrocarbons
Solid ResidueGaseous Hydrocarbons
PSBatch Reactor10500PelletizedMgO88102[107]
Plastic waste (PS+PE)TL- 200 Tube furnace1048010%Organic vermiculite80.60.119.4[108]
PS+PETL-200 Tube furnace1048010%Co/Vermiculites73.20.126.7[108]
PS+PE
(Plastic waste)
TL-200 Tube furnace1048010%Ni/Vermiculites70.71.328.0[108]
PS+PE
(Plastic waste)
TL-200 Tube furnace1048010%Co-Ni/Vermiculites73.90.126.0[108]
PP+LDPE+ HDPEBatch reactor2050033.3%Calcium bentonite81.312.76[109]
PP+PE+PS+PETVertical tube reactor 600100%Ca(OH)252.29.027.7[110]
PP+PE+PS+PETVertical tube reactor 600100%Cao46.731.817.6[110]
Mixed plastic wasteSemi-batch reactor20440
20 °C/min
10%ZSM-556.93.240.4[111]
Mixed plasticBatch reactor-240-Activated carbon82.4315.222.35[112]
Mixed plasticBatch reactor-240-Charcoal95.542.332.13[112]
Mixed plasticVertical reactor-50010%USY-Zeolite70–80 [113]
Mixed plastic wasteBatch glass reactor535010%Al–Si93.119.270.36[114]
Mixed plastic wasteStirred tank reactor104501.65 kg/h 60 g catalyst (36.36%)Fe-restructured clay83.73 [115]
Mixed plastic wasteVertical tube reactor20500-CaCO3685.726.3[52]
Mixed plastic wasteFixed-batch reactor5.5400--8686[116]
Mixed plastic wasteFixed-bed reactor1060010%FCC69283[74]
Mixed plastic wasteBatch reactor155105%
1:1:2
Red mud: Ca(OH): Ni/SAPO64.2–71.9--[117]
Mixed plastic wasteContinuous microwave-assisted
pyrolysis system
-62010% pelletized formZSM-548.9-49.0[118]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pal, S.; Kumar, A.; Sharma, A.K.; Ghodke, P.K.; Pandey, S.; Patel, A. Recent Advances in Catalytic Pyrolysis of Municipal Plastic Waste for the Production of Hydrocarbon Fuels. Processes 2022, 10, 1497. https://doi.org/10.3390/pr10081497

AMA Style

Pal S, Kumar A, Sharma AK, Ghodke PK, Pandey S, Patel A. Recent Advances in Catalytic Pyrolysis of Municipal Plastic Waste for the Production of Hydrocarbon Fuels. Processes. 2022; 10(8):1497. https://doi.org/10.3390/pr10081497

Chicago/Turabian Style

Pal, Shashank, Anil Kumar, Amit Kumar Sharma, Praveen Kumar Ghodke, Shyam Pandey, and Alok Patel. 2022. "Recent Advances in Catalytic Pyrolysis of Municipal Plastic Waste for the Production of Hydrocarbon Fuels" Processes 10, no. 8: 1497. https://doi.org/10.3390/pr10081497

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop