Next Article in Journal
In Situ Electrochemical Characterization of a Microbial Fuel Cell Biocathode Running on Wastewater
Next Article in Special Issue
Effect of Temperature, Syngas Space Velocity and Catalyst Stability of Co-Mn/CNT Bimetallic Catalyst on Fischer Tropsch Synthesis Performance
Previous Article in Journal
Modified Red Mud Catalyst for Volatile Organic Compounds Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Low-Cost Catalysts for Pyrolysis of Plastic Waste to Fuels

1
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Islam Indonesia, Jl. Kaliurang Km 14, Sleman, Yogyakarta 55584, Indonesia
2
School of Information, Systems and Modelling, Faculty of Engineering and Information Technology, University of Technology Sydney, Sydney, NSW 2007, Australia
3
Research & Technology Innovation, Pertamina, Sopo Del Building, 51st Fl. Jl. Mega Kuningan Barat III, Jakarta Pusat 12950, Indonesia
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(7), 837; https://doi.org/10.3390/catal11070837
Submission received: 9 June 2021 / Revised: 28 June 2021 / Accepted: 7 July 2021 / Published: 10 July 2021
(This article belongs to the Special Issue Current Advanced Technologies in Catalysts/Catalyzed Reactions)

Abstract

:
The catalytic and thermal decomposition of plastic waste to fuels over low-cost catalysts like zeolite, clay, and bimetallic material is highlighted. In this paper, several relevant studies are examined, specifically the effects of each type of catalyst used on the characteristics and product distribution of the produced products. The type of catalyst plays an important role in the decomposition of plastic waste and the characteristics of the oil yields and quality. In addition, the quality and yield of the oil products depend on several factors such as (i) the operating temperature, (ii) the ratio of plastic waste and catalyst, and (iii) the type of reactor. The development of low-cost catalysts is revisited for designing better and effective materials for plastic solid waste (PSW) conversion to oil/bio-oil products.

1. Introduction

Pollution of plastics in the environment has become a serious issue in recent years, producing more than 300 million tons per year [1]. One of the problems is the microplastics issue which is related to the non-degradable properties of the plastic polymer. Many efforts have been reported to reduce and overcome the presence of plastics and microplastic wastes, and one of the promising alternatives is the conversion of plastic waste into renewable energy [2,3]. More than just overcoming the environmental problems, with a designed plastic waste management system, the conversion of plastic waste into renewable energy will also contribute to the vital issue of energy conservation [4,5]. As hydrocarbon is the backbone of plastics, catalytic processes of plastic structure within the pyrolysis technique can produce hydrogen and liquid fuel. The conversion mechanisms include cracking, hydrocracking, and hydrogenation can restore the energy contained in plastic, as a sustainable process for sustainable energy [6,7,8]. For those mechanisms, catalysts play important roles in determining the effectiveness and efficient conversion process. Although in the perspective of kinetics and thermodynamic, the catalyst’s role is to accelerate the reaction, it in fact determines the dominant product of the reaction which is further called selectivity, the optimum condition for the reaction, and also the energy required for the process [9]. The use of the solid catalyst for the catalytic pyrolysis is established and favorable, as it is easily handled, efficient in mass transfer conditions to obtain a high yield. Pyrolysis can be performed by thermal or catalytic processes. However, compared to the thermal decomposition method, the catalytic method has some advantages that could be carried out at a lower temperature and reduce the solid residues such as carbonized char and volatile fraction, short time process, high product selectivity, high octane number, etc. [10]. Moreover, in pyrolysis by cracking catalytic process, the used catalyst can be easy to reuse and reproduce, which can be classified into sustainable approaches [11]. The main factor for the mechanism lays in the domination of radical propagation steps in thermal pyrolysis instead of cationic propagation occurs in the catalytic process, leading to an uncontrolled decomposition. These advantages are feasible from the economic point of view.
Moreover, the combination of plastic or plastic waste with biomass as feed for pyrolysis, also called co-pyrolysis, has also gained much attention. The combination in the catalytic pyrolysis mechanism has been reported to provide synergistic effects such as minimizing coke formation and increasing yield and selectivity towards gasoline and aromatic fraction [12,13]. Wang et al. (2021) reported that the addition of low-density polyethylene (LDPE) to biomass waste could improve the selectivity and aromatics product, including xylene, benzene, and toluene [14]. However, the selectivity conversion also depends on the different types of catalysts used. Dai et al. (2021) studied the pyrolysis process using different types and tandem catalysts [15]. Their study revealed that tandem catalysts could improve the selectivity of naphtha. Many solid catalysts are reported with specific results, and some of them are the catalysts based on zeolite, bimetallic and clay with the comparable popularity expressed in Figure 1. Among these kinds of catalysts, zeolite-based catalysts are the most popular for both pyrolysis and co-pyrolysis for the combined plastic waste/biomass. An intensive catalytic mechanism is provided by reactant migration and surface reaction over the microporous structure of zeolite. Even though a similar mechanism also occurs on clay, clay-based catalyst receives less attention for the processes [16]. By considerations of many possible modifications towards clay, clay-based catalysts are good candidates as low-cost catalysts for plastic waste pyrolysis. With many modifications for zeolite and clay framework with metals or metal oxides zeolite, the catalytic activity enhancements were attempted by increasing the effectiveness of the reaction pathways such as hydrogenation, hydro-deoxygenation, cracking, etc. Catalysts’ thermal and chemical stability refers to the use of high-temperature conditions and very complex reactions involved within the mechanisms as important characters, besides the solid acidity and capability to provide efficient mass transport in the catalytic steps. Referring to their abundant sources in nature, both clay and zeolite materials are found as cheap minerals [17,18,19].
However, there is no clear conclusion or generalization obtained regarding reference [16]’s effect of catalyst performance. Other factors such as temperature, pressure and plastic/biomass ratio in co-pyrolysis allow for the discernment of the impact on the result within these perspectives, and bimetal catalysts are also developed, especially for improvement on gaseous products and hydrogen. In this review, we discuss some essential properties of low-cost catalysts for plastic waste pyrolysis.

2. Clay-Based Catalysts for Plastic Pyrolysis

Catalytic reactions of plastic pyrolysis depend on solid acid mechanisms on the surface which include cracking, isomerization, oligomerization, cyclisation and aromatization reactions. These various mechanisms are governed by acidity, density, porous size, and porous structure of the catalyst surface. Both Lewis and Brønsted acid sites of a clay catalyst play roles in the cracking mechanism which is initiated by the abstraction of the hydride ion from the polymer structure by Lewis acid sites of the catalyst, or the addition of a proton to the C–C bonds by Bronsted acid sites of the catalyst [20]. The higher amount of Brønsted acid sites on the surface of the catalyst provides more hydrogen ions for double bond cleavage and further propagation steps. Meanwhile, with a different role, the Lewis acid sites influence the surface interaction of catalysts with polyolefin, which is an important part of the whole surface reaction in heterogeneous catalysis.
The surface acidity and high specific surface area of the catalyst play an important role in producing liquid products instead of gas products. The availability of micropores in the clay structure has the potency to act as a heat sink and allows a greater residence time for feed molecules to absorb the heat and have interactions that result in hydrogen transfer [21]. The main role of solid catalyst in the liquid product is enhancing the ability to crack the polymer structure to form an intermediate in the mechanism. This influences by increasing the liquid product along with decreasing the wax content. Less wax from the use of bentonite clay refers to the presence of surface acidity but the surface acidity of clay is lower than zeolite as the impact of Si/Al ratio [22]. Less surface acidity led to the lower Brønsted acidity compared with zeolite which minimizes the potential of secondary reaction such as an over-cracking mechanism, so more liquid product distribution is achieved [23]. Increased conversion and selectivity in producing liquid were exhibited by modifying clay structure via pillarization using aluminum. The increasing conversion not only came from increasing the specific surface area and Lewis’s surface acidity distribution, but also came from the thermal stability of the integrity of the clay structure for the sustained pores and surface area for the cracking mechanism [24,25].
Moreover, the stability of pillared clay led to the renewability properties whereas regenerated and reused catalysts showed practically identical conversion and yield values compared with the fresh catalyst. According to the identification of the composition of oil yield from pyrolysis reaction, the presence of metal oxide as a pillar minimizes the side-reaction mechanism. Two main mechanisms: (i) cyclization/aromatization of pyrolytic intermediate and (ii) cracking dominantly occur and play roles in producing diesel fraction in the liquid product together with the high yield of hydrogen gas. Another important aspect from the study on varied metal oxide impregnated onto acid-washed bentonite clay for polypropylene (PP) and high-density polyethylene (HDPE) pyrolysis is the importance of Lewis acid sites from metal ions for facilitating reactions via the formation of the hydration of a proton or the hydride ion due to its surface acidity of catalyst materials [26]. The β-scission of chain-end carbonium ion is the following reaction from acid interaction with polymer chain for the further production of gas and liquid fractions. The β-scission mechanism is presented in Figure 2.
The more effective acid–polymer interaction in the mechanism also prevents residue such as coke formation on the surface as shown by the comparison on acid-washed bentonite clay (AWBC) and metal oxide impregnated AWBC [20]. Table 1 presents the pyrolysis reaction of plastic waster over the different types of clay catalysts.
Referring to the product distribution of pyrolysis reaction, it can be summarized that catalyst surface acidity and pore characteristics are mainly responsible for catalytic performance. The mesoporous structure with a high surface area is closely related to the Si/Al ratio of the catalyst. The surface acidity facilitated the mechanism of the reaction by the formation of the hydride ion or the addition of a proton due to the inherent acidity, which is simultaneously incorporated with the impregnated metals. This mechanism increased the liquid yield as a substantial improvement over thermal cracking which has a tendency to produce gas as the result of the radical mechanism, also called the random scission mechanism [28,29]. The study on HDPE pyrolysis over HZSM-12 revealed that the solid acidity linearly decreased the activation energy (Ea), as proof of the important role of the acid mechanism in accelerating the reaction [33]. However, the extremely high acidity leads to increased yield by the over-cracking that leads to the formation of much smaller molecules. For this reason, aluminum-pillared clay, which has mild acidity while producing a higher liquid yield (~70%) [34], is similar to the use of ultra-stable zeolite (USY (71%) [35], which is higher compared to the yield by ZSM-5 (61%) [36]. The plot in Figure 3 represents the relationship between the solid acidity and the liquid yield from several papers.

3. Plastic Waste-to-Fuel over Zeolite Catalysts

Zeolite (ZSM-5) has been widely reported as an effective and selective material catalyst for producing biofuel through the thermocatalytic reaction [38]. Besides its catalyst base, ZSM-5 is a low-cost catalyst for the conversion of plastic waste to biofuel. The ZSM-5 catalyst also presents excellent thermal stability, good selectivity, activity, and deactivation by coke [39,40]. In thermocatalytic reactions, the ZSM-5 effectively enhances the deoxygenation and cracking reaction to produce stable oil [41,42]. Onwudili et al. (2019) have studied the influence of temperature and type of zeolite as a catalyst for pyrolysis reaction to convert the plastic waste to biofuel liquid products [43]. Using different catalysts at temperatures of 500 and 600 °C showed no significant results of fuel-range liquid products. However, the increasing temperature resulted in the increasing gas composition of the products. Besides that, the high acidic catalyst can lead to a faster production of gases. More acidity promotes the formation of the hydrogenation steps, leading to the synthesis of other free radicals which resulted in β-scission for the gas production [16,44,45]. The related study by Kassargy et al. (2019) also reported that the yields of the liquids fractions are linearly dependent on the proportion and the type of plastic waste in the mixture [44].
Miscolczi et al. (2019) studied the effect of the loading metal to zeolite structure for catalysis plastic to fuel [45]. The authors modified the zeolite catalyst with several metal ions like Fe2+/3+, Cu2+, Ce2+, H+, Mg2+, Ni2+, Sn2+, and Zn2+. Their studies revealed that the presence of metal loading (8–10%) to zeolite structure could affect the pore diameter and the macropore surface area of the materials. The surface area is an essential factor in the pyrolysis reaction process besides the acidity of the catalyst [46,47]. Therefore, modification with several metal ions to zeolite structure can easily control the surface area of the catalyst materials which affects the decreasing temperature decomposition of the plastics during the pyrolysis process. Gorbannezhad et al. (2020) revealed that the co-pyrolysis also depended on the composition of the zeolite catalyst [48]. The authors studied the co-pyrolysis process by combining zeolite (HZSM-5) and sodium carbonate/gamma-alumina for improving the hydrocarbon products of the reaction process. Their study showed that the combination of catalysts could improve the hydrocarbon product to 8.7% at the temperature process of 700 °C. The presence of sodium salt in the composition of the catalyst can improve the deoxygenation reaction so that the breakdown process of macromolecules becomes faster and more effective to produce low molecular compounds such as hydrocarbons. Besides that, the other studies also reported that the catalytic pyrolysis process is affected by the type of zeolite catalyst [49,50]. Several types of catalyst-based zeolite have a difference in the pore size and acidity which affected the produced product [51]. For example, the larger pore size facilitates the conversion of the plastics as source materials to polyalkylaromatics while the smaller pore size only converts to aromatic compounds with small dynamic diameters. Based on their study, the use of suitable zeolite as a catalyst in the pyrolysis reaction is a crucial step to determine the produced product during the reaction process.
Susastriawan et al. (2020) studied the effect of zeolite size on pyrolysis of LDPE plastic waste at low temperatures [52]. The authors revealed that reducing zeolite size could enhance the reaction rate, pyrolysis temperature, heat transfer rate, and the oil products because the smaller size has a high surface-active area to contact with the plastics during a pyrolysis process. The zeolite size of 1 mm showed the highest value of oil yields; however, the particle size of 1–3 mm did not indicate the results of oil produced significantly. In the other studies also reported by Kim et al. (2018), the authors showed that the presence of the phenolic functional group on lignin could enhance 39% of the aromatic hydrocarbon in the totally resulted product [53]. The presence of a hydroxyl group on the chemical structure of lignin gives good selectivity on the decomposition of reactant to form an aromatic hydrocarbon as a major product [54,55]. In addition, the presence of a hydroxyl group on the surface of a zeolite-type catalyst can contribute to the condensation process during pyrolysis. Besides that, hydroxyl can accelerate the formation of aromatic hydrocarbon products through dihydroxylation, aromatization, isomerization, and oligomerization mechanism. Table 2 shows the comparison of zeolite types for pyrolysis reaction of plastic waste. All researchers confirmed that the zeolite catalyst can improve the acid-activation and thermal activation in the pyrolysis reaction of plastic waste (PE, PP, PVC, PET, and PS) [56,57].

4. Effect of Co-Feeding with Biomass Feedstock

The biomass feedstock consists of extractive (0–14%), lignin (16%), hemicellulose (20%), and cellulose (>40%) [61,62,63]. Generally, the yields of the produced biofuels are dependent on several factors such as the composition of the nature of the feedstock, the moisture content of biomass feedstocks, reactor design, and operating temperature conditions [64,65,66,67]. In the pyrolysis process, the biomass feedstock can be mixed with other materials to improve the quantity and the quality of the product as shown in Table 3. There are some pathways for the pyrolysis of biomass like the chemical, biological, and thermochemical conversion as shown in Figure 4. However, the thermochemical reaction is commonly used for the pyrolysis reaction of biomass feedstock because the process has high energy to split and stretch the rigid structure that has biomass [68]. The synergistic effect is remarkable, coming from the presence of produced free radicals from biomass decomposition that contribute to enhancing the scission of chain hydrocarbons from plastics. Decreasing activation energy and the pyrolysis index, representing the easiness of the pyrolytic reaction for producing volatile products (methane, aliphatic hydrocarbon (paraffin), carbon dioxide, aromatic hydrocarbon), are the quantitative parameters revealed to be advantageous of co-pyrolysis [69]. The characteristics of biomass, especially the ratio of hydrogen to carbon effective (H/Ceff), heavily influences the product distribution. Higher H/Ceff significantly improves olefin and aromatic yield and reduces coke formation [70,71,72]. In addition, Bhoi et al. (2019) reported operating conditions such as heating rate, type and particle size of biomass, temperature, carrier gas, type of catalyst, and vapor residence time [73]. The authors and the other related studies reported that the type of catalyst and temperature have major impacts on quality and biofuel yields [74,75,76]. However, the choice of temperature during the pyrolysis reaction depends on the main composition of the biomass sources, for example, hemicellulose and cellulose degrade at approximately 200–350 °C and 330–370 °C, respectively, while lignin occurs at 400 °C [77,78].
Xu et al. (2020) studied the pyrolysis reaction with the mixing of microalgae (Enteromorpha prolifera) and HDPE plastics using the HZSM-5 catalyst [79]. The authors reported that the presence of microalgae in the plastic waste could increase the concentration of aliphatic hydrocarbons and decrease the nitrogen/oxygen-containing compounds and the acidity of the products. Algae as natural resources have received attention because they can produce a high amount of bio-oil than the other resources of biomass [80]. The related study also reported that using microalgae as a pyrolysis feedstock has many advantages like faster rate of growth and high lipid content, lower energy consumption during a process with the percent of energy recovery of 76%, increase in the carbon content to 89%, and decrease in the oxygen content to 0.3% [81,82,83]. Furthermore, Qari et al. (2017) reported that the characteristics of microalgae significantly affected the types and the yields of biofuel products [84]. In the produced bio-oil product, the mixing between biomass and plastic waste can reduce the moisture content because the hydroxyl group in the biomass can directly bind with the plastics which contain the high hydrogen atom to form the hydrogen bonding [85]. Figure 4 shows the process of pyrolysis reaction with the mixing feedstock between natural biomass of microalgae and plastic waste.
Table 3. Effect of co-feeding with other feeds.
Table 3. Effect of co-feeding with other feeds.
Type of Biomass/PolymerCatalystReaction ConditionConversionSelectivity to Remark (Catalyst/Plastic Ratio)Ref
Laminaria japonica/polypropylenePt/mesoporous MFI,
Mesoporous MFI,
HZSM-5
Al-SBA-16
500 °C, 1 atm60.50
59.11
58.15
52.44
20–40% of monoaromatic hydrocarbon, 5–20% of polyaromatic hydrocarbonBiomass/plastic/catalyst ratio = 10:10:1[86]
Corn stalk/HDPEZSM-5700 °C, 1 atm90% of hydrocarbon20–30% of aromatic hydrocarbonBiomass/plastic/catalyst = 1:4:1[87]
pine sawdust/LDPENi-CaO-C-90% of gas product86.74% to H2 gas Biomass/plastic/catalyst: 1:1:2[88]
Corn stalk/PolystyreneZSM-5600 °C90% of liquid product78.89% of monoaromatic hydrocarbon4:1:0.1[89]
L. japonica/polypropyleneAl-SBA 1550030% of liquid product 35% of liquid product is oxygenateSeaweed/polypropylene/catalyst = 1: 1:1[90]
Rice husk/PENi/γ-Al2O3Pyrolysis at 600 °C followed by catalytic reforming at 800 °C80% of H2 and CO 45% of gaseous product is H250~75% PE proportion[91]
Muneer et al. (2019) studied the effect of catalyst ratio to feedstock for the pyrolysis of corn stalk (CS) and polypropylene in a bed reactor at 500 °C [89]. The increase in the catalyst could enhance the liquid oil yield to 66.5% at ZSM-5 catalyst to feedstock ratio of 1:4. ZSM-5 catalyst was found as an effective material for polymer cracking and dehydration of biomass because of its high surface area and high selectivity to produce the hydrocarbon [92]. The production of bio-oil from biomass feedstocks depends on the acidic site and the pore structure of the catalyst because the acidic site increases the rate of polymer cracking. Similar results were also reported by Balasundram et al. (2018); their study revealed that the increasing catalyst amount four times could improve the coke decomposition by 17.1% [93]. It can be summarized that the addition of biomass in the co-pyrolysis enhances the efficiency to produce liquid products and reduce activation energy. By using ZSM-5 catalyst and corn stalk in the co-pyrolysis of HPDE, it was found that increased hydrocarbon yield and H/C eff were obtained on increasing biomass/HDPE ratio, along with decreasing coke [87]. Hydrogen atoms for the co-pyrolysis process were provided by HDPE, leading to an improvement in the rate of hydrocarbon production, meanwhile, oxygenated compounds in the biomass play a role to promote the cracking of HPDE and the chain scission. However, at a certain level, the increasing catalyst dosage may affect the increased charring, which leads to reduce liquid product. For example, the co-pyrolysis of cellulose/polyethylene over montmorillonite K10. A similar trend is also identified on cellulose pyrolysis using montmorillonite on cellulose. The availability of more surface acidity in Al-SBA enhances the co-pyrolysis to produce more C1–C4 hydrocarbon compounds, CO, CO2, and deoxygenation reactions [86].

5. Recent Reports Bimetallic Catalysts for Pyrolysis of Plastics

Bimetallic catalysts have been widely used for the pyrolysis reaction of plastic waste to biofuel. Previously, the monometallic type has been widely studied and reported for the catalytic cracking process. For example, Wen et al. (2014) prepared Ni-loaded to CNTs for polyolefin; although the material showed good performance as the catalyst, the concentration of carbon was relatively very high [94]. Therefore, bimetallic with integrating the different types of material catalyst provides some advantages in the pyrolysis reaction such as large surface area due to its smaller size, good stability, and synergy effect between combined two metals [95]. Yao et al. (2017) studied the Ni-Fe bimetallic catalyst at a ratio of 1:3 for the pyrolysis of waste plastics [96]. The authors reported that the presence of a bimetallic catalyst could enhance H2 production to five times higher compared to the process without a catalyst. Chen et al. (2020) also reported that the Fe-Ni bimetallic modified MCM-41 could improve the produced oil to 49.9% with a percentage of single styrene hydrocarbon of 65.93% at 10%Fe-10%Ni/MCM-41 [97]. Besides that, the presence of bimetallic/MCM-41 can reduce the bromine content from 10% to 2.3% (wt.). The developed catalyst provides a large surface area; thus, the plastics can directly enter the pore structure for the cracking process. The iron metals act as the base site which converts the raw materials to styrene, while the combination with nickel–metal oxides increases the acidity of the catalyst, thus the multi-ring compounds can also be converted to a single hydrocarbon structure. The character of the external surface and pore size determines the chemisorption, and these are designable by the synthesis method. In this case, the higher pore diameter of Fe-Ni catalyst tends to give higher H2 desorption [98], which is in line with the trend of long-chain products by the high pore size of Co/SBA-15 [51].
Li et al. (2016) reported the bi-functional Mo-Ni/SiO2-Al2O3 catalyst for the thermal pyrolysis of crude oil [99]. The presence of the bimetallic of Mo-Ni on the composite catalyst could improve the catalytic reactivity and the amount of yield of fuel oil produced to 57.9%. The metallic of Mo and Ni has several advantages for catalyst in pyrolysis like low-cost adsorbent, excellent stability performance, high surface area, and also ease of regeneration [100]. However, their study reported that the reactivity of the bimetallic catalyst is dependent on the sulfurization process. The increased sulfurization process can regularly improve the percentage of conversion of crude oil to fuel oil until it reaches 86.9%. In addition, some related studies have reported that the bimetallic catalyst type has good selectivity conversion [101,102]. A more specific capability of bimetallic catalyst is shown by Fe-Ni/MCM-41 for not only the decomposition of polymer structure but also to conduct debromination mechanism for plastic waste containing brominated flame-retardant (BFR) pyrolysis. The results indicated that iron showed a satisfactory capability for debromination, in combination with Ni’s ability to produce gaseous products via the hydrogenation mechanism. One of the proposed pathways of the reaction mechanism that occurs is the interaction between organobromine and some metal oxides to produce non-brominated organic compounds, elimination of β-H by Lewis acid sites and dissociative adsorption two-stage reaction [97].
Cai et al. (2020) reported the carbon-based Fe-Ni bimetallic catalyst for fast pyrolysis of plastic waste [103]. The pyrolysis reaction was carried out in a fixed-bed reactor with the ratio of catalyst to plastic waste at 1:2 (wt. ratio) at a temperature of 500 °C. The fixed-bed reactor has some advantages for pyrolysis reaction such as a simple design with the catalyst loaded into the bed column, irregularity in plastic shape, and low thermal experiments [104,105]. The authors explained that Fe-Ni as a bimetallic catalyst played a crucial role in the oxygen reduction reaction (ORR). Moreover, the developed catalyst has good methanol tolerance and stability, avoiding the aggregation and corrosion process on the surface of the catalyst due to its oxygen-containing functional groups on the surface [106]. Zhou et al. (2020) also reported that the bimetallic Ni-Fe/ZrO2 catalyst showed excellent decomposition of polystyrene at a low-temperature process (500 °C) [107]. Their studies revealed that the presence of bimetallic catalysts could improve the catalytic activity in the decomposition of waste. The combined properties between Ni-Fe could decrease the water–gas shift reaction and the activation energy of the reforming reaction [108,109,110]. Table 4 presents the different types of bimetallic catalysts for the pyrolysis process.

6. Factor Affecting in Pyrolysis Process

In the pyrolysis process, many reaction parameters strongly affect the quantity and quality of the resulting products, such as temperature, heating rate, composition blending ratio, and the type of reactor design. However, the reactor designs play a crucial factor in obtaining a high yield quantity and the quality of the product. Previous studies have already reported several types of reactor design for the pyrolysis process, such as fixed-bed, transported bed, rotating cone, plasma pyrolysis, vortex centrifuge, circulating fluidized, entrained flow, etc. [120]. The fixed-bed reactor is commonly used for the pyrolysis process due to the simple process, a large sample quantity and a high product yield. However, the reactor experiences several drawbacks, such as the catalyst being difficult to replace during a pyrolysis process, side reaction and product, and the required high temperature and pressures. Therefore, recently, many types of reactor-based fast pyrolysis reactors have been developed. The principle of the reactors is to optimize the percentage of products in low temperatures and pressure during the pyrolysis process.
Xue et al. (2015) reported the fast pyrolysis of HDPE waste using a fluidized bed reactor [121]. Their study found that HDPE waste could increase the formation of acid and furans in the products and decrease the formation of phenol and vanillin compounds at a relatively lower temperature process. Orozco et al. (2021) studied the pyrolysis of plastic waste using a spouted bed reactor in continuous mode [122]. Their study revealed that the synergy effect between the plastic-type and the total amount of catalyst could decrease the temperature. The reaction design of fast pyrolysis reaction is based on optimizing the product yields by decreasing the size of plastic waste to less than 1 mm and flowing the carrier gas as shown in Figure 5. However, despite the lack of the reactor, there is still a high amount of oxygen content in products. A related study has been reported for removing the oxygen content in pyrolysis products by hydrotreating formate-assisted pyrolysis [123]. Reducing oxygen in the resulting product can improve the percentage of the product by up to 92%.

7. Conclusions and Future Prospect

In summary, we reported and focused on the catalytic decomposition of plastic waste to produce liquid fuel using a low-cost catalyst. Several types of low-cost catalysts were summarized, such as zeolite, clay, and bimetallic. Generally, some studies reported that the low-cost catalyst could be used for catalytic cracking of plastic waste to liquid fuel. However, several factors are still required for improving the quantity and quality of the products, such as decreasing the particle size of waste in feedstocks, temperature, pressure, ratio composition between catalyst to feedstocks, type of catalyst, and type of reactor design. Unbeatably, pyrolysis and co-pyrolysis of plastic waste to produce fuel is a promising sustainable technology. Several conditions of the operation process influence the yield of the products such as (i) heating rate, (ii) the ratio of blending materials, and (iii) the pyrolysis temperature [104]. Additionally, the heterogeneous catalyst plays an important role in the conversion, and from the techno-economic point of view, exploration of the low-cost catalyst with high efficiency and lifetime is still required. Designing an appropriate catalyst is a major challenge in developing the technique, and referring to this study, the selectivity of a catalyst toward a specific product becomes important in the design of the reaction set up, as well as the lifetime of the catalyst. Catalyst solid acidity along with the specific surface area governs the optimum condition of the conversion, and these are related to the structure of solid support and their combination with active metal/bimetals on the surface. Compared to commercial solid acid catalysts, natural zeolite and clays are better options since they are cost-effective and more selective towards liquid products [1]. Referring to decoking capability, tunable performance of bimetallic catalysts and the potencies of zeolite and clay utilization, bimetallic-modified zeolite and clay catalysts are interesting for exploration. The designed Fe-Ni with zeolite or clay supports, for example, is a good candidate for a low-cost pyrolysis catalyst for bromide-containing plastic wastes which are abundantly produced from flame-retarded plastic production. In addition, some strategies to create integrative plastic waste management including co-pyrolysis of biomass/plastic wastes with these catalyst candidates need to be investigated for furthermore well-implemented.

Funding

Funding was provided by World Class Professor Program, Ministry of Research and Higher Education, Republic of Indonesia, 2020.

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miandad, R.; Barakat, M.A.; Rehan, M.; Aburiazaiza, A.S.; Ismail, I.M.I.; Nizami, A.S. Plastic waste to liquid oil through catalytic pyrolysis using natural and synthetic zeolite catalysts. Waste Manag. 2017, 69, 66–78. [Google Scholar] [CrossRef] [PubMed]
  2. Rehan, M.; Nizami, A.-S.; Asam, Z.-U.-Z.; Ouda, O.K.M.; Gardy, J.; Raza, G.; Naqvi, M.; Mohammad Ismail, I. Waste to Energy: A Case Study of Madinah City. Energy Procedia 2017, 142, 688–693. [Google Scholar] [CrossRef]
  3. Samun, I.; Saeed, R.; Abbas, M.; Rehan, M.; Nizami, A.-S.; Asam, Z.-U.-Z. Assessment of Bioenergy Production from Solid Waste. Energy Procedia 2017, 142, 655–660. [Google Scholar] [CrossRef]
  4. Banu, J.R.; Sharmila, V.G.; Ushani, U.; Amudha, V.; Kumar, G. Impervious and influence in the liquid fuel production from municipal plastic waste through thermo-chemical biomass conversion technologies—A review. Sci. Total Environ. 2020, 718, 137287. [Google Scholar] [CrossRef]
  5. Zakir Hossain, H.M.; Hasna Hossain, Q.; Uddin Monir, M.M.; Ahmed, M.T. Municipal solid waste (MSW) as a source of renewable energy in Bangladesh: Revisited. Renew. Sustain. Energy Rev. 2014, 39, 35–41. [Google Scholar] [CrossRef]
  6. Munir, D.; Irfan, M.F.; Usman, M.R. Hydrocracking of virgin and waste plastics: A detailed review. Renew. Sustain. Energy Rev. 2018, 90, 490–515. [Google Scholar] [CrossRef]
  7. Walendziewski, J. Continuous flow cracking of waste plastics. Fuel Process. Technol. 2005, 86, 1265–1278. [Google Scholar] [CrossRef]
  8. Kunwar, B.; Cheng, H.N.; Chandrashekaran, S.R.; Sharma, B.K. Plastics to fuel: A review. Renew. Sustain. Energy Rev. 2016, 54, 421–428. [Google Scholar] [CrossRef]
  9. Zhan, H.; Zhuang, X.; Song, Y.; Liu, J.; Li, S.; Chang, G.; Yin, X.; Wu, C.; Wang, X. A review on evolution of nitrogen-containing species during selective pyrolysis of waste wood-based panels. Fuel 2019, 253, 1214–1228. [Google Scholar] [CrossRef]
  10. 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] [PubMed]
  11. Olaremu, A.G.; Adedoyin, W.R.; Ore, O.T.; Adeola, A.O. Sustainable development and enhancement of cracking processes using metallic composites. Appl. Petrochem. Res. 2021, 11, 1–18. [Google Scholar] [CrossRef]
  12. Campuzano, F.; Brown, R.C.; Martínez, J.D. Auger reactors for pyrolysis of biomass and wastes. Renew. Sustain. Energy Rev. 2019, 102, 372–409. [Google Scholar] [CrossRef]
  13. Al-Hamamre, Z.; Saidan, M.; Hararah, M.; Rawajfeh, K.; Alkhasawneh, H.E.; Al-Shannag, M. Wastes and biomass materials as sustainable-renewable energy resources for Jordan. Renew. Sustain. Energy Rev. 2017, 67, 295–314. [Google Scholar] [CrossRef]
  14. Wang, Z.; Burra, K.G.; Lei, T.; Gupta, A.K. Co-pyrolysis of waste plastic and solid biomass for synergistic production of biofuels and chemicals-A review. Prog. Energy Combust. Sci. 2021, 84, 100899. [Google Scholar] [CrossRef]
  15. Dai, L.; Zhou, N.; Li, H.; Wang, Y.; Liu, Y.; Cobb, K.; Cheng, Y.; Lei, H.; Chen, P.; Ruan, R. Catalytic fast pyrolysis of low density polyethylene into naphtha with high selectivity by dual-catalyst tandem catalysis. Sci. Total Environ. 2021, 771, 144995. [Google Scholar] [CrossRef]
  16. Budsaereechai, S.; Hunt, A.J.; Ngernyen, Y. Catalytic pyrolysis of plastic waste for the production of liquid fuels for engines. RSC Adv. 2019, 9, 5844–5857. [Google Scholar] [CrossRef] [Green Version]
  17. Kumar, B.S.; Dhakshinamoorthy, A.; Pitchumani, K. K10 montmorillonite clays as environmentally benign catalysts for organic reactions. Catal. Sci. Technol. 2014, 4, 2378–2396. [Google Scholar] [CrossRef]
  18. Pienkoß, F.; Ochoa-Hernández, C.; Theyssen, N.; Leitner, W. Kaolin: A Natural Low-Cost Material as Catalyst for Isomerization of Glucose to Fructose. Acs Sustain. Chem. Eng. 2018, 6, 8782–8789. [Google Scholar] [CrossRef]
  19. Król, M. Natural vs. Synthetic Zeolites. Crystals 2020, 10, 622. [Google Scholar] [CrossRef]
  20. Ahmad, I.; Khan, M.; Khan, H.; Ishaq, M.; Tariq, R.; Gul, K.; Ahmad, W. Influence of Metal-Oxide-Supported Bentonites on the Pyrolysis Behavior of Polypropylene and High-Density Polyethylene. J. Appl. Polym. Sci. 2014, 132, 1–19. [Google Scholar] [CrossRef]
  21. Patil, V.; Adhikari, S.; Cross, P. Co-pyrolysis of lignin and plastics using red clay as catalyst in a micro-pyrolyzer. Bioresour. Technol. 2018, 270, 311–319. [Google Scholar] [CrossRef]
  22. Fatimah, I.; Rubiyanto, D.; Prakoso, N.I.; Yahya, A.; Sim, Y.-L. Green conversion of citral and citronellal using tris(bipyridine)ruthenium(II)-supported saponite catalyst under microwave irradiation. Sustain. Chem. Pharm. 2019, 11, 61–70. [Google Scholar] [CrossRef]
  23. Tarach, K.A.; Góra-Marek, K.; Martinez-Triguero, J.; Melián-Cabrera, I. Acidity and accessibility studies of desilicated ZSM-5 zeolites in terms of their effectiveness as catalysts in acid-catalyzed cracking processes. Catal. Sci. Technol. 2017, 7, 858–873. [Google Scholar] [CrossRef] [Green Version]
  24. Vogt, E.T.C.; Weckhuysen, B.M. Fluid catalytic cracking: Recent developments on the grand old lady of zeolite catalysis. Chem. Soc. Rev. 2015, 44, 7342–7370. [Google Scholar] [CrossRef] [Green Version]
  25. Geng, J.; Sun, Q. Effects of high temperature treatment on physical-thermal properties of clay. Thermochim. Acta 2018, 666, 148–155. [Google Scholar] [CrossRef]
  26. Ferrini, P.; Dijkmans, J.; De Clercq, R.; Van de Vyver, S.; Dusselier, M.; Jacobs, P.A.; Sels, B.F. Lewis acid catalysis on single site Sn centers incorporated into silica hosts. Coord. Chem. Rev. 2017, 343, 220–255. [Google Scholar] [CrossRef]
  27. 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]
  28. 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]
  29. De Stefanis, A.; Cafarelli, P.; Gallese, F.; Borsella, E.; Nana, A.; Perez, G. Catalytic pyrolysis of polyethylene: A comparison between pillared and restructured clays. J. Anal. Appl. Pyrolysis 2013, 104, 479–484. [Google Scholar] [CrossRef]
  30. Li, K.; Lei, J.; Yuan, G.; Weerachanchai, P.; Wang, J.-Y.; Zhao, J.; Yang, Y. Fe-, Ti-, Zr- and Al-pillared clays for efficient catalytic pyrolysis of mixed plastics. Chem. Eng. J. 2017, 317, 800–809. [Google Scholar] [CrossRef]
  31. Faillace, J.G.; de Melo, C.F.; de Souza, S.P.L.; da Costa Marques, M.R. Production of light hydrocarbons from pyrolysis of heavy gas oil and high density polyethylene using pillared clays as catalysts. J. Anal. Appl. Pyrolysis 2017, 126, 70–76. [Google Scholar] [CrossRef]
  32. Attique, S.; Batool, M.; Yaqub, M.; Görke, O.; Gregory, D.; Shah, A. Highly efficient catalytic pyrolysis of polyethylene waste to derive fuel products by novel polyoxometalate/kaolin composites. Waste Manag. Res. 2020, 0734242X1989971. [Google Scholar] [CrossRef]
  33. Silva, A.O.S.; Souza, M.J.B.; Pedrosa, A.M.G.; Coriolano, A.C.F.; Fernandes, V.J.; Araujo, A.S. Development of HZSM-12 zeolite for catalytic degradation of high-density polyethylene. Microporous Mesoporous Mater. 2017, 244, 1–6. [Google Scholar] [CrossRef]
  34. Manos, G.; Yusof, I.Y.; Gangas, N.H.; Papayannakos, N. Tertiary Recycling of Polyethylene to Hydrocarbon Fuel by Catalytic Cracking over Aluminum Pillared Clays. Energy Fuels 2002, 16, 485–489. [Google Scholar] [CrossRef]
  35. Kassargy, C.; Awad, S.; Burnens, G.; Kahine, K.; Tazerout, M. Experimental study of catalytic pyrolysis of polyethylene and polypropylene over USY zeolite and separation to gasoline and diesel-like fuels. J. Anal. Appl. Pyrolysis 2017, 127, 31–37. [Google Scholar] [CrossRef]
  36. Ghaffar, N.; Johari, A.; Tuan Abdullah, T.A.; Ripin, A. Catalytic Cracking of High Density Polyethylene Pyrolysis Vapor over Zeolite ZSM-5 towards Production of Diesel. IOP Conf. Ser. Mater. Sci. Eng. 2020, 808, 012025. [Google Scholar] [CrossRef]
  37. Seddegi, Z.S.; Budrthumal, U.; Al-Arfaj, A.A.; Al-Amer, A.M.; Barri, S.A.I. Catalytic cracking of polyethylene over all-silica MCM-41 molecular sieve. Appl. Catal. A Gen. 2002, 225, 167–176. [Google Scholar] [CrossRef]
  38. Lok, C.M.; Van Doorn, J.; Aranda Almansa, G. Promoted ZSM-5 catalysts for the production of bio-aromatics, a review. Renew. Sustain. Energy Rev. 2019, 113, 109248. [Google Scholar] [CrossRef]
  39. Gayubo, A.G.; Aguayo, A.T.; Atutxa, A.; Prieto, R.; Bilbao, J. Deactivation of a HZSM-5 Zeolite Catalyst in the Transformation of the Aqueous Fraction of Biomass Pyrolysis Oil into Hydrocarbons. Energy Fuels 2004, 18, 1640–1647. [Google Scholar] [CrossRef]
  40. Kang, Y.-H.; Wei, X.-Y.; Liu, G.-H.; Ma, X.-R.; Gao, Y.; Li, X.; Li, Y.-J.; Ma, Y.-J.; Yan, L.; Zong, Z.-M. Catalytic Hydroconversion of Ethanol-Soluble Portion from the Ethanolysis of Hecaogou Subbituminous Coal Extraction Residue to Clean Liquid Fuel over a Zeolite Y/ZSM-5 Composite Zeolite-Supported Nickel Catalyst. Energy Fuels 2020, 34, 4799–4807. [Google Scholar] [CrossRef]
  41. Wang, C.; Liu, Q.; Song, J.; Li, W.; Li, P.; Xu, R.; Ma, H.; Tian, Z. High quality diesel-range alkanes production via a single-step hydrotreatment of vegetable oil over Ni/zeolite catalyst. Catal. Today 2014, 234, 153–160. [Google Scholar] [CrossRef]
  42. Hou, X.; Qiu, Y.; Zhang, X.; Liu, G. Analysis of reaction pathways for n-pentane cracking over zeolites to produce light olefins. Chem. Eng. J. 2017, 307, 372–381. [Google Scholar] [CrossRef]
  43. 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 Ins. 2019, 92, 1337–1347. [Google Scholar] [CrossRef]
  44. 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]
  45. Miskolczi, N.; Juzsakova, T.; Sója, J. Preparation and application of metal loaded ZSM-5 and y-zeolite catalysts for thermo-catalytic pyrolysis of real end of life vehicle plastics waste. J. Energy Ins. 2019, 92, 118–127. [Google Scholar] [CrossRef]
  46. Ahmed, M.H.M.; Masuda, T.; Muraza, O. The role of acidity, side pocket, and steam on maximizing propylene yield from light naphtha cracking over one-dimensional zeolites: Case studies of EU–1 and disordered ZSM–48. Fuel 2019, 258, 116034. [Google Scholar] [CrossRef]
  47. Galadima, A.; Muraza, O. Zeolite catalyst design for the conversion of glucose to furans and other renewable fuels. Fuel 2019, 258, 115851. [Google Scholar] [CrossRef]
  48. Ghorbannezhad, P.; Park, S.; Onwudili, J.A. Co-pyrolysis of biomass and plastic waste over zeolite- and sodium-based catalysts for enhanced yields of hydrocarbon products. Waste Manag. 2020, 102, 909–918. [Google Scholar] [CrossRef]
  49. Lin, X.; Zhang, Z.; Wang, Q. Evaluation of zeolite catalysts on product distribution and synergy during wood-plastic composite catalytic pyrolysis. Energy 2019, 189, 116174. [Google Scholar] [CrossRef]
  50. Kianfar, E.; Hajimirzaee, S.; Mousavian, S.; Mehr, A.S. Zeolite-based catalysts for methanol to gasoline process: A review. Microchem. J. 2020, 156, 104822. [Google Scholar] [CrossRef]
  51. 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 2018, 8, 661. [Google Scholar] [CrossRef] [Green Version]
  52. Susastriawan, A.A.P.; Purnomo; 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]
  53. 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]
  54. Li, T.; Ma, H.; Wu, S.; Yin, Y. Effect of highly selective oxypropylation of phenolic hydroxyl groups on subsequent lignin pyrolysis: Toward the lignin valorization. Energy Convers. Manag. 2020, 207, 112551. [Google Scholar] [CrossRef]
  55. Ma, H.; Li, T.; Wu, S.; Zhang, X. Effect of the interaction of phenolic hydroxyl with the benzene rings on lignin pyrolysis. Bioresour. Technol. 2020, 309, 123351. [Google Scholar] [CrossRef]
  56. 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, 1–17. [Google Scholar] [CrossRef] [Green Version]
  57. Yao, D.; Yang, H.; Chen, H.; Williams, P.T. Investigation of nickel-impregnated zeolite catalysts for hydrogen/syngas production from the catalytic reforming of waste polyethylene. Appl. Catal. B Environ. 2018, 227, 477–487. [Google Scholar] [CrossRef]
  58. Akubo, K.; Nahil, M.A.; Williams, P.T. Aromatic fuel oils produced from the pyrolysis-catalysis of polyethylene plastic with metal-impregnated zeolite catalysts. J. Energy Inst. 2019, 92, 195–202. [Google Scholar] [CrossRef]
  59. 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]
  60. López, A.; de Marco, I.; Caballero, B.M.; Laresgoiti, M.F.; Adrados, A.; Aranzabal, A. Catalytic pyrolysis of plastic wastes with two different types of catalysts: ZSM-5 zeolite and Red Mud. Appl. Catal. B Environ. 2011, 104, 211–219. [Google Scholar] [CrossRef]
  61. Sharma, A.; Pareek, V.; Zhang, D. Biomass pyrolysis—A review of modelling, process parameters and catalytic studies. Renew. Sustain. Energy Rev. 2015, 50, 1081–1096. [Google Scholar] [CrossRef]
  62. Kan, T.; Strezov, V.; Evans, T.J. Lignocellulosic biomass pyrolysis: A review of product properties and effects of pyrolysis parameters. Renew. Sustain. Energy Rev. 2016, 57, 1126–1140. [Google Scholar] [CrossRef]
  63. Ha, J.-M.; Hwang, K.-R.; Kim, Y.-M.; Jae, J.; Kim, K.H.; Lee, H.W.; Kim, J.-Y.; Park, Y.-K. Recent progress in the thermal and catalytic conversion of lignin. Renew. Sustain. Energy Rev. 2019, 111, 422–441. [Google Scholar] [CrossRef]
  64. Hossain, Z.; Johnson, E.N.; Wang, L.; Blackshaw, R.E.; Cutforth, H.; Gan, Y. Plant establishment, yield and yield components of Brassicaceae oilseeds as potential biofuel feedstock. Ind. Crop. Prod. 2019, 141, 111800. [Google Scholar] [CrossRef]
  65. Umrigar, V.R.; Chakraborty, M.; Parikh, P. Catalytic activity of zeolite Hβ for the preparation of fuels’ additives: Its product distribution and scale up calculation for the biofuel formation in a microwave assisted batch reactor. J. Environ. Chem. Eng. 2018, 6, 6816–6827. [Google Scholar] [CrossRef]
  66. Ibn Ferjani, A.; Jeguirim, M.; Jellali, S.; Limousy, L.; Courson, C.; Akrout, H.; Thevenin, N.; Ruidavets, L.; Muller, A.; Bennici, S. The use of exhausted grape marc to produce biofuels and biofertilizers: Effect of pyrolysis temperatures on biochars properties. Renew. Sustain. Energy Rev. 2019, 107, 425–433. [Google Scholar] [CrossRef]
  67. Kumar, M.; Olajire Oyedun, A.; Kumar, A. A review on the current status of various hydrothermal technologies on biomass feedstock. Renew. Sustain. Energy Rev. 2018, 81, 1742–1770. [Google Scholar] [CrossRef]
  68. Brigagão, G.V.; de Queiroz Fernandes Araújo, O.; de Medeiros, J.L.; Mikulcic, H.; Duic, N. A techno-economic analysis of thermochemical pathways for corncob-to-energy: Fast pyrolysis to bio-oil, gasification to methanol and combustion to electricity. Fuel Process. Technol. 2019, 193, 102–113. [Google Scholar] [CrossRef]
  69. Akyurek, Z. Sustainable Valorization of Animal Manure and Recycled Polyester: Co-pyrolysis Synergy. Sustainability 2019, 11, 2280. [Google Scholar] [CrossRef] [Green Version]
  70. Zhang, H.; Cheng, Y.-T.; Vispute, T.; Xiao, R.; Huber, G. Catalytic Conversion of Biomass-derived Feedstocks into Olefins and Aromatics with ZSM-5: The Hydrogen to Carbon Effective Ratio. Energy Environ. Sci. 2011, 4, 2297–2307. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, Y.; Bi, P.; Wang, J.; Jiang, P.; Wu, X.; Xue, H.; Liu, J.; Zhou, X.; Li, Q. Production of jet and diesel biofuels from renewable lignocellulosic biomass. Appl. Energy 2015, 150, 128–137. [Google Scholar] [CrossRef]
  72. Zhang, L.; Bao, Z.; Xia, S.; Lu, Q.; Walters, K.B. Catalytic Pyrolysis of Biomass and Polymer Wastes. Catalysts 2018, 8, 659. [Google Scholar] [CrossRef] [Green Version]
  73. Bhoi, P.R.; Ouedraogo, A.S.; Soloiu, V.; Quirino, R. Recent advances on catalysts for improving hydrocarbon compounds in bio-oil of biomass catalytic pyrolysis. Renew. Sustain. Energy Rev. 2020, 121, 109676. [Google Scholar] [CrossRef]
  74. Guedes, R.E.; Luna, A.S.; Torres, A.R. Operating parameters for bio-oil production in biomass pyrolysis: A review. J. Anal. Appl. Pyrolysis 2018, 129, 134–149. [Google Scholar] [CrossRef]
  75. Zhu, Y.; Xu, G.; Song, W.; Zhao, Y.; Miao, Z.; Yao, R.; Gao, J. Catalytic microwave pyrolysis of orange peel: Effects of acid and base catalysts mixture on products distribution. J. Energy Inst. 2021. [Google Scholar] [CrossRef]
  76. Sogancioglu, M.; Yel, E.; Ahmetli, G. Investigation of the Effect of Polystyrene (PS) Waste Washing Process and Pyrolysis Temperature on (PS) Pyrolysis Product Quality. Energy Procedia 2017, 118, 189–194. [Google Scholar] [CrossRef]
  77. Chen, Y.; Fang, Y.; Yang, H.; Xin, S.; Zhang, X.; Wang, X.; Chen, H. Effect of volatiles interaction during pyrolysis of cellulose, hemicellulose, and lignin at different temperatures. Fuel 2019, 248, 1–7. [Google Scholar] [CrossRef]
  78. Collard, F.-X.; Blin, J. A review on pyrolysis of biomass constituents: Mechanisms and composition of the products obtained from the conversion of cellulose, hemicelluloses and lignin. Renew. Sustain. Energy Rev. 2014, 38, 594–608. [Google Scholar] [CrossRef]
  79. Xu, S.; Cao, B.; Uzoejinwa, B.B.; Odey, E.A.; Wang, S.; Shang, H.; Li, C.; Hu, Y.; Wang, Q.; Nwakaire, J.N. Synergistic effects of catalytic co-pyrolysis of macroalgae with waste plastics. Process Saf. Environ. Prot. 2020, 137, 34–48. [Google Scholar] [CrossRef]
  80. Adnan, M.A.; Xiong, Q.; Muraza, O.; Hossain, M.M. Gasification of wet microalgae to produce H2-rich syngas and electricity: A thermodynamic study considering exergy analysis. Renew. Energy 2020, 147, 2195–2205. [Google Scholar] [CrossRef]
  81. Kositkanawuth, K.; Sattler, M.L.; Dennis, B. Pyrolysis of Macroalgae and Polysytrene: A Review. Curr. Sustain. Renew. Energy 2014, 1, 121–128. [Google Scholar] [CrossRef] [Green Version]
  82. Milledge, J.; Benjamin, S.; Dyer, P.; Harvey, P. Macroalgae-Derived Biofuel: A Review of Methods of Energy Extraction from Seaweed Biomass. Energies 2014, 7, 7194–7222. [Google Scholar] [CrossRef]
  83. Li, F.; Srivatsa, S.C.; Bhattacharya, S. A review on catalytic pyrolysis of microalgae to high-quality bio-oil with low oxygeneous and nitrogenous compounds. Renew. Sustain. Energy Rev. 2019, 108, 41–497. [Google Scholar] [CrossRef]
  84. Qari, H.; Rehan, M.; Nizami, A.-S. Key Issues in Microalgae Biofuels: A Short Review. Energy Procedia 2017, 142, 898–903. [Google Scholar] [CrossRef]
  85. Chen, W.; Lu, J.; Zhang, C.; Xie, Y.; Wang, Y.; Wang, J.; Zhang, R. Aromatic hydrocarbons production and synergistic effect of plastics and biomass via one-pot catalytic co-hydropyrolysis on HZSM-5. J. Anal. Appl. Pyrolysis 2020, 147, 104800. [Google Scholar] [CrossRef]
  86. Kim, Y.-M.; Lee, H.W.; Choi, S.J.; Jeon, J.-K.; Park, S.H.; Jung, S.-C.; Kim, S.C.; Park, Y.-K. Catalytic co-pyrolysis of polypropylene and Laminaria japonica over zeolitic materials. Int. J. Hydrogen Energy 2017, 42, 18434–18441. [Google Scholar] [CrossRef]
  87. Zhang, B.; Zhong, Z.; Ding, K.; Song, Z. Production of aromatic hydrocarbons from catalytic co-pyrolysis of biomass and high density polyethylene: Analytical Py–GC/MS study. Fuel 2015, 139, 622–628. [Google Scholar] [CrossRef]
  88. Chai, Y.; Gao, N.; Wang, M.; Wu, C. H2 production from co-pyrolysis/gasification of waste plastics and biomass under novel catalyst Ni-CaO-C. Chem. Eng. J. 2020, 382, 122947. [Google Scholar] [CrossRef]
  89. 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]
  90. Lee, H.; Choi, S.; Park, S.; Jeon, J.-K.; Jung, S.-C.; Kim, S.; Park, Y.-K. Pyrolysis and co-pyrolysis of Laminaria japonica and polypropylene over mesoporous Al-SBA-15 catalyst. Nanoscale Res. Lett. 2014, 9, 376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Xu, D.; Xiong, Y.; Ye, J.; Su, Y.; Dong, Q.; Zhang, S. Performances of syngas production and deposited coke regulation during co-gasification of biomass and plastic wastes over Ni/γ-Al2O3 catalyst: Role of biomass to plastic ratio in feedstock. Chem. Eng. J. 2020, 392, 123728. [Google Scholar] [CrossRef]
  92. Tan, S.; Zhang, Z.; Sun, J.; Wang, Q. Recent progress of catalytic pyrolysis of biomass by HZSM-5. Chin. J. Catal. 2013, 34, 641–650. [Google Scholar] [CrossRef]
  93. Balasundram, V.; Ibrahim, N.; Kasmani, R.M.; Isha, R.; Hamid, M.K.A.; Hasbullah, H.; Ali, R.R. Catalytic upgrading of sugarcane bagasse pyrolysis vapours over rare earth metal (Ce) loaded HZSM-5: Effect of catalyst to biomass ratio on the organic compounds in pyrolysis oil. Appl. Energy 2018, 220, 787–799. [Google Scholar] [CrossRef]
  94. 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]
  95. Kaya, B.; Irmak, S.; Hasanoğlu, A.; Erbatur, O. Developing Pt based bimetallic and trimetallic carbon supported catalysts for aqueous-phase reforming of biomass-derived compounds. Int. J. Hydrog. Energy 2015, 40, 3849–3858. [Google Scholar] [CrossRef]
  96. Yao, D.; Wu, C.; Yang, H.; Zhang, Y.; Nahil, M.A.; Chen, Y.; Williams, P.T.; Chen, H. Co-production of hydrogen and carbon nanotubes from catalytic pyrolysis of waste plastics on Ni-Fe bimetallic catalyst. Energy Convers. Manag. 2017, 148, 692–700. [Google Scholar] [CrossRef]
  97. 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]
  98. Yao, D.; Wang, C.-H. Pyrolysis and in-line catalytic decomposition of polypropylene to carbon nanomaterials and hydrogen over Fe- and Ni-based catalysts. Appl. Energy 2020, 265, 114819. [Google Scholar] [CrossRef]
  99. Li, J.; Xia, H.; Wu, Q.; Hu, Z.; Hao, Z.; Zhu, Z. Hydrocracking of the crude oil from thermal pyrolysis of municipal wastes over bi-functional Mo–Ni catalyst. Catal. Today 2016, 271, 172–178. [Google Scholar] [CrossRef]
  100. 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]
  101. Upare, D.P.; Park, S.; Kim, M.S.; Jeon, Y.P.; Kim, J.; Lee, D.; Lee, J.; Chang, H.; Choi, S.; Choi, W.; et al. Selective hydrocracking of pyrolysis fuel oil into benzene, toluene and xylene over CoMo/beta zeolite catalyst. J. Ind. Eng. Chem. 2017, 46, 356–363. [Google Scholar] [CrossRef]
  102. Hamid, S.; Niaz, Y.; Bae, S.; Lee, W. Support induced influence on the reactivity and selectivity of nitrate reduction by Sn-Pd bimetallic catalysts. J. Environ. Chem. Eng. 2020, 8, 103754. [Google Scholar] [CrossRef]
  103. Cai, N.; Yang, H.; Zhang, X.; Xia, S.; Yao, D.; Bartocci, P.; Fantozzi, F.; Chen, Y.; Chen, H.; Williams, P.T. Bimetallic carbon nanotube encapsulated Fe-Ni catalysts from fast pyrolysis of waste plastics and their oxygen reduction properties. Waste Manag. 2020, 109, 119–126. [Google Scholar] [CrossRef]
  104. 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]
  105. Gao, Z.; Li, N.; Wang, Y.; Niu, W.; Yi, W. Pyrolysis behavior of xylan-based hemicellulose in a fixed bed reactor. J. Anal. Appl. Pyrolysis 2020, 146, 104772. [Google Scholar] [CrossRef]
  106. Lai, C.; Wang, J.; Lei, W.; Xuan, C.; Xiao, W.; Zhao, T.; Huang, T.; Chen, L.; Zhu, Y.; Wang, D. Restricting Growth of Ni3Fe Nanoparticles on Heteroatom-Doped Carbon Nanotube/Graphene Nanosheets as Air-Electrode Electrocatalyst for Zn–Air Battery. Acs Appl. Mater. Interfaces 2018, 10, 38093–38100. [Google Scholar] [CrossRef]
  107. 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]
  108. Baamran, K.S.; Tahir, M. Ni-embedded TiO2-ZnTiO3 reducible perovskite composite with synergistic effect of metal/support towards enhanced H2 production via phenol steam reforming. Energy Convers. Manag. 2019, 200, 112064. [Google Scholar] [CrossRef]
  109. Kharaji, A.G.; Shariati, A.; Takassi, M.A. A Novel γ-Alumina Supported Fe-Mo Bimetallic Catalyst for Reverse Water Gas Shift Reaction. Chin. J. Chem. Eng. 2013, 21, 1007–1014. [Google Scholar] [CrossRef]
  110. 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]
  111. Aboul-Enein, A.A.; Awadallah, A.E. Production of nanostructured carbon materials using Fe–Mo/MgO catalysts via mild catalytic pyrolysis of polyethylene waste. Chem. Eng. J. 2018, 354, 802–816. [Google Scholar] [CrossRef]
  112. Aboul-Enein, A.A.; Awadallah, A.E. Production of nanostructure carbon materials via non-oxidative thermal degradation of real polypropylene waste plastic using La2O3 supported Ni and Ni-Cu catalysts. Polym. Degrad. Stab. 2019, 167, 157–169. [Google Scholar] [CrossRef]
  113. Bajad, G.S.; Tiwari, S.K.; Vijayakumar, R.P. Synthesis and characterization of CNTs using polypropylene waste as precursor. Mater. Sci. Eng. B 2015, 194, 68–77. [Google Scholar] [CrossRef]
  114. Shen, Y.; Gong, W.; Zheng, B.; Gao, L. Ni–Al bimetallic catalysts for preparation of multiwalled carbon nanotubes from polypropylene: Influence of the ratio of Ni/Al. Appl. Catal. B Environ. 2016, 181, 769–778. [Google Scholar] [CrossRef]
  115. Nahil, M.A.; Wu, C.; Williams, P.T. Influence of metal addition to Ni-based catalysts for the co-production of carbon nanotubes and hydrogen from the thermal processing of waste polypropylene. Fuel Process. Technol. 2015, 130, 46–53. [Google Scholar] [CrossRef]
  116. Wu, S.-L.; Kuo, J.-H.; Wey, M.-Y. Thermal degradation of waste plastics in a two-stage pyrolysis-catalysis reactor over core-shell type catalyst. J. Anal. Appl. Pyrolysis 2019, 104641. [Google Scholar] [CrossRef]
  117. Aboul-Enein, A.A.; Awadallah, A.E. Impact of Co/Mo ratio on the activity of CoMo/MgO catalyst for production of high-quality multi-walled carbon nanotubes from polyethylene waste. Mater. Chem. Phys. 2019, 238, 121879. [Google Scholar] [CrossRef]
  118. Wang, J.; Shen, B.; Lan, M.; Kang, D.; Wu, C. Carbon nanotubes (CNTs) production from catalytic pyrolysis of waste plastics: The influence of catalyst and reaction pressure. Catalysis Today 2019, 351, 50–57. [Google Scholar] [CrossRef]
  119. Wu, C.; Nahil, M.A.; Miskolczi, N.; Huang, J.; Williams, P.T. Processing Real-World Waste Plastics by Pyrolysis-Reforming for Hydrogen and High-Value Carbon Nanotubes. Environ. Sci. Technol. 2014, 48, 819–826. [Google Scholar] [CrossRef]
  120. Saravanan, A.; Hemavathy, R.V.; Sundararaman, T.R.; Jeevanantham, S.; Kumar, P.S.; Yaashikaa, P.R. 1—Solid waste biorefineries. In Refining Biomass Residues for Sustainable Energy and Bioproducts; Kumar, R.P., Gnansounou, E., Raman, J.K., Baskar, G., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 3–17. [Google Scholar]
  121. Xue, Y.; Zhou, S.; Brown, R.C.; Kelkar, A.; Bai, X. Fast pyrolysis of biomass and waste plastic in a fluidized bed reactor. Fuel 2015, 156, 40–46. [Google Scholar] [CrossRef]
  122. Orozco, S.; Alvarez, J.; Lopez, G.; Artetxe, M.; Bilbao, J.; Olazar, M. Pyrolysis of plastic wastes in a fountain confined conical spouted bed reactor: Determination of stable operating conditions. Energy Convers. Manag. 2021, 229, 113768. [Google Scholar] [CrossRef]
  123. Khlewee, M.; Gunukula, S.; Wheeler, M.C.; DeSisto, W.J. Hydrotreating of reduced oxygen content bio-oils produced by formate-assisted pyrolysis. Fuel 2019, 254, 115570. [Google Scholar] [CrossRef]
Figure 1. Popularity of different types of catalytic materials for pyrolysis of plastic waste from 2015 to 2020. (Source: Web of Knowledge, https://www.webofknowledge.com, accessed on 1 April 2020).
Figure 1. Popularity of different types of catalytic materials for pyrolysis of plastic waste from 2015 to 2020. (Source: Web of Knowledge, https://www.webofknowledge.com, accessed on 1 April 2020).
Catalysts 11 00837 g001
Figure 2. The mid-chain radical mechanism for obtaining low molecular product.
Figure 2. The mid-chain radical mechanism for obtaining low molecular product.
Catalysts 11 00837 g002
Figure 3. Relationship between solid acidity and liquid yield [27,33,34,35,36,37].
Figure 3. Relationship between solid acidity and liquid yield [27,33,34,35,36,37].
Catalysts 11 00837 g003
Figure 4. The illustration of pyrolysis process with the natural biomass and plastic waste as a feedstock.
Figure 4. The illustration of pyrolysis process with the natural biomass and plastic waste as a feedstock.
Catalysts 11 00837 g004
Figure 5. The illustration of combined pyrolysis reactor designs for improving the yields products.
Figure 5. The illustration of combined pyrolysis reactor designs for improving the yields products.
Catalysts 11 00837 g005
Table 1. Recent works on pyrolysis of plastic waste over clay catalysts.
Table 1. Recent works on pyrolysis of plastic waste over clay catalysts.
NoCatalystPlastic TypeResultRef
1Calcium BentonitePolypropylene (PP), low-density polyethylene (LDPE), and high-density polyethylene (HDPE)The yield was influenced by temperature and ratio of effect to catalyst. The major product as condensable fraction was in the temperature range 400–550 °C and the optimum condition was at 500 °C at the catalyst to plastics ratio of 1:3[27]
2KaolinPPAhoko kaolin exhibited as effective as a low-cost catalyst for producing gasoline/diesel grade fuel with the PP as waste sources. The yield was influenced by catalyst to plastic ratio[28]
3Restructured and pillared claypolyolefinRestructured and pillared clay showed good selectivity towards aliphatic, produced more liquid[29]
5Fe, Ti, Zr- pillared clayHDPE, polystyrene [PS], PPFe-pillared clay showed excellent yield of diesel fraction in liquid product and H2[30]
6Fe-pillared clayHeavy gas oil(HGO)/HDPEThe presence of HGO improved the oil yield from both thermal and catalytic pyrolysis of HDPE[31]
7Tungstophosphoric acid (TPA)/kaolinLow-density polyethylene (LDPE)TPA loaded kaolin (5-TPA-K) produced higher percentages of gasoline-like hydrocarbons (C11–C14)[32]
8Co, Fe, Mn, Zn impregnated acid-washed bentonite clay (AWBC)PP and HDPEMetal oxide impregnation on acid-washed bentonite clay not only improves conversion but also yield reduce coke formation[20]
Manos
Table 2. Pyrolysis of plastic waste over zeolite-based catalysts.
Table 2. Pyrolysis of plastic waste over zeolite-based catalysts.
Type of PlasticCatalystReaction ConditionConversionSelectivity toRemark (Catalyst/Plastic Ratio)Ref
Polystyrene and polyolefeins (PS/PO)Y-zeolite600 °C for 30 min under N2 gasHigh yield valuable aromatics such as benzene and toluene90% of the aromatic content2 g of catalyst and 2 g of plastic[43]
Polyethylene and polypropylene (PE/PP)USY-zeolite500 °CLiquid fractions are dominated by hydrocarbon (C5–C7), C3 and C4 for gaseous products. 80% of liquid productionCatalyst/plastics ratio of 1:10[44]
Polystyrene (PS)Natural/Synthetic zeolite450 °C for 75 min60.8% conversion to ethylbenzene and 38.3% convertsion to alpha-methylstyrene for natural and synthetic zeolite respectively54% and 50% of liquids products for natural and synthetic zeolite, respectivelyCatalys/PS ratio of 0.1 kg:1 kg[1]
High density Polyethylene (HDPE))Co-Y–zeolite600 °C for 30 min40% of gas yield68% of hydrogen productionCatalys/HDPE ratio of 2:1[58]
Plastic mixtures (HDPE/PP/PS/PET/PVC)Regenerated ZSM-5440 °C for 30 minAlmost 60% of plastic waste conversion to liquids phase97.4% of aromatics with 23% of styrene as major composition Catalyst/plastic waste ratio of 1:10[59]
Plastic mixtures (PE/PP/PS/PET/PVC)ZSM-5500 °C for 30 min58.4% conversion to gases phase50.7% of C3–C4 types and 27.9% of styreneCatalyst/plastic waste ratio of 1:10[60]
Table 4. Recent reports bimetallic catalysts for pyrolysis of plastics.
Table 4. Recent reports bimetallic catalysts for pyrolysis of plastics.
Type of Plastic1st Metal2nd MetalCondition, Pressure (atm), Temperature (°C)ConversionSelectivity to Remark (Catalyst/Plastic Ratio)Ref
Low density polyethylene (LDPE)Mo-MgOFeAtmospheric pressure, 750 and 400 °CLDPE waste plastic to carbon nanotubesHigh quality carbon nanostructures materials0.5 g:15 g plastics[111]
Polypropylene (PP) La2O3Ni-Cu500 °C, 700 °C for 2.5 hPP to Carbon nanotubes and carbon nanofibersCarbon yields of 1458% produced0.5 g:15 g plastics[112]
Polypropylene (PP)MgONi/Mo800 °C, 10 minPP to CNT394% of carbon product0.15 g:5 g polymer[113]
Polypropylene (PP)Ni-Al800 °CPP to MWCNTs85%Dependence on the ratio Ni/Al and the amount of Ni-Al catalyst[114]
Polypropylene (PP)Ni-AlZn, Mg, Ca, Ce, Mn500 °CPP to CNTsThe highest carbon deposition 62% and hydrogen 86.4% to Ni-Mn-Al1 g:2 g waste polypropylene[115]
Polyehtylene (PE)NiCe (Ni-Ce core by silica)800 °CPE to hydrogen Hydrogen concentration 60%Weight ratio Catalyst:plastic 1.0[116]
LDPENiFe800 °CCarbon nanotubes (CNTs)Maximum hydrogen concentration and hydrogen yield 73.93% and 84.72 mg.g−10.5 g:1 g waste plastic[96]
Low-density polyethylene (LDPE) wasteMgOCo/Mo400 °CHigh quality multi-walled Carbon Nanotubes and hydrogenOptimum CNTs 1040% wtCoMo(6.5) MgO0.75 g:15 g plastics[117]
Polypropylene (PP)NiFe500 °CCNTs93% filamentous carbon nanotubes0.5 catalyst:1 g PP[118]
HDPENiMn-Al800 °CHydrogen and carbon nanotubes48% total carbon (with no steam), hydrogen yield 94,4% (with steam)0.5 g catalyst:1 g waste plastic (HDPE)[119]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/catal11070837

AMA Style

Fadillah G, Fatimah I, Sahroni I, Musawwa MM, Mahlia TMI, Muraza O. Recent Progress in Low-Cost Catalysts for Pyrolysis of Plastic Waste to Fuels. Catalysts. 2021; 11(7):837. https://doi.org/10.3390/catal11070837

Chicago/Turabian Style

Fadillah, Ganjar, Is Fatimah, Imam Sahroni, Muhammad Miqdam Musawwa, Teuku Meurah Indra Mahlia, and Oki Muraza. 2021. "Recent Progress in Low-Cost Catalysts for Pyrolysis of Plastic Waste to Fuels" Catalysts 11, no. 7: 837. https://doi.org/10.3390/catal11070837

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