Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4256 2023-02-22 06:43:16 |
2 Some of the new methods plastic waste treatment and references has been added + 3524 word(s) 7780 2023-02-22 07:16:14 | |
3 format change -3012 word(s) 4768 2023-02-22 08:01:02 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Rafey, A.; Pal, K.; Bohre, A.; Modak, A.; Pant, K.K. Conventional Techniques for Thermal Treatment of Plastic Waste. Encyclopedia. Available online: https://encyclopedia.pub/entry/41515 (accessed on 05 September 2024).
Rafey A, Pal K, Bohre A, Modak A, Pant KK. Conventional Techniques for Thermal Treatment of Plastic Waste. Encyclopedia. Available at: https://encyclopedia.pub/entry/41515. Accessed September 05, 2024.
Rafey, Abdul, Kunwar Pal, Ashish Bohre, Arindam Modak, Kamal Kishore Pant. "Conventional Techniques for Thermal Treatment of Plastic Waste" Encyclopedia, https://encyclopedia.pub/entry/41515 (accessed September 05, 2024).
Rafey, A., Pal, K., Bohre, A., Modak, A., & Pant, K.K. (2023, February 22). Conventional Techniques for Thermal Treatment of Plastic Waste. In Encyclopedia. https://encyclopedia.pub/entry/41515
Rafey, Abdul, et al. "Conventional Techniques for Thermal Treatment of Plastic Waste." Encyclopedia. Web. 22 February, 2023.
Conventional Techniques for Thermal Treatment of Plastic Waste
Edit

Plastic waste poses a serious threat to the environment and it has been increasing at an alarming rate. In 2022, global plastic waste generation was reported to be around 380 million tonnes as compared to 353 million tonnes in 2019. Production of liquid fuel from plastic waste is regarded as a viable method for disposing of the plastic and utilizing its energy. A wide range of technologies have been explored for turning plastic waste into fuel, including the conventional pyrolysis, incineration, gasification and advanced oxidation.

plastic waste pyrolysis Gasification

1. Introduction

Plastic materials like polyethylene terephthalate (PET), low-density polyethylene (LDPE), high-density polyethylene (HDPE), polyvinyl chloride (PVC), polypropylene (PP) and polystyrene (PS) play a significant role in our daily lives including uses in a wide variety of products, like packaging, electronics, aircrafts, sporting goods and automobiles. Its widespread use results from its unmatchable advantages and characteristics when compared to other available materials [1]. One of the most significant components of solid waste management, and a notable fraction of municipal solid waste (MSW), are plastics with an overall global production of around 390 million metric tonnes in 2021, much higher than in the year 1950 (1.7 million metric tonnes). This is expected to reach 1800 million metric tonnes in 2050 [2][3]. While plastic production has increased at such a rapid rate, environmental issues have recently drawn attention and concern on a global scale.
Today’s greatest challenges are the management of plastic waste and the reduction of environmental plastic pollution, especially keeping in mind the increasing demand for plastic products. Owing to the challenges present in the current scenario, there is a need to urgently switch the strategy of plastic handling from landfilling and incineration towards a more sustainable and environmentally friendly alternative. Plastic waste can be treated via different pathways as illustrated in Figure 1.
Figure 1. Pathways for plastic waste management.
One of the most efficient ways of handling plastic waste is to extract and utilize energy from it. Compared to incineration and landfill disposal, fuel conversion from plastics can reduce hazardous emissions and pathogen contamination. In the process of creating fuel from waste plastic, reaction factors (such as temperature, residence time and rate of increase of temperature) can control the overall conversion process.
Various fractions of hydrocarbons can be produced from plastic waste when subjected to chemical treatment (namely, gasification, pyrolysis and catalytic degradation) [4]. By using these thermochemical processes, plastic waste can be transformed into variety of energy products (electricity, process heat for industrial facilities and automobile fuels) [5]. In this regard, chemical treatment of plastic waste emerges as a more sustainable approach as the conversion efficiency is sufficiently high and energy can be recovered from the plastic waste, leading to fuel production, syngas/hydrogen (H2) production and various other chemical productions [6]. Reviews on the pyrolysis and catalytic cracking of plastic waste into fuels with C8-C22 aromatic and aliphatic hydrocarbon have been published [7][8][9]. The composition and characteristics of product fuels were analysed in relation to the effects of pyrolysis and catalytic cracking parameters [10][11][12].
Notably, many studies on innovative technologies, like hydrothermal liquefaction and advanced oxidation processes, have been undertaken recently describing plastic waste conversion [13][14]. Compared to conventional pyrolysis, hydrothermal liquefaction allows for plastic waste conversion into excellent liquid oils at a significantly lower temperature. Additionally, high purity oil can be produced by advanced oxidation of plastic at normal temperature and pressure. In advanced oxidation systems, polymers break down to H2 in the presence of hydroxyl (·OH) species [15]. During the process, formation of carboxylic acids (R-COOH) takes place, making plastics act as carbon source for chemical sector. Additionally, while using photoreduction or electrocatalysis, the CO2 generated during the advanced oxidation process can be transformed into acetic acid [16].

2. Conventional Techniques for Thermal Treatment of Plastic Waste

2.1. Pyrolysis

Pyrolysis refers to thermal degradation of plastic waste in the absence of oxygen to produce liquid and gaseous fuels [17]. Temperature for the pyrolysis process generally ranges between 300–700 °C. It offers several benefits as compared to other ways of handling plastic waste, such as greater strength and flexibility, elimination of cost associated with segregation, washing and blending before mechanical recycling, the ability to treat both thermoset and thermoplastic and in the handling of polymer composites in engineering applications [18]. Composition of final product and yield is dependent on the composition of feedstock, temperature, reactor configuration, rate of heating and catalyst [19]. The choice of reactor has a significant impact on the efficiency of the reaction leading to the desired end product, residence time, heat transfer, and mixing of the polymers and catalysts. Most of the plastic pyrolysis in laboratories was carried out in batch, semi-batch or continuous-flow reactors like fluidized beds, fixed-bed reactors and conical spouted beds (CSBR).
A batch reactor is essentially a closed system as no reactants or products can enter or leave it while the reaction is taking place. One of the benefits of a batch reactor is that a high conversion can be attained by keeping the reactant in the reactor for a long time. The variability of the product from batch to batch, high labor expenses in each batch and the difficulties of large-scale manufacturing are the drawbacks associated with the batch reactor. A semi-batch reactor, however, enables simultaneous reactant addition and product removal. Regarding reaction selectivity, the semi-batch reactor also has the option of gradually introducing reactants. Semi-batch reactors have similar labor cost issues as batch reactors. As such, they are better suited for small-scale production.
The catalyst is often palletized and packed in a static bed in fixed-bed reactors. Although it is simple to design, there are certain limitations, like the irregular particle size and type of plastics used as feedstock, which present issues during the feeding process. Additionally, the reaction has a finite amount of access to the catalyst’s surface area. In some circumstances, fixed-bed reactors are only employed as the secondary pyrolysis reactor, because the product from primary pyrolysis may be readily fed into the fixed-bed reactor, which typically comprises liquid and gaseous phases.
However, some of the issues that arise within fixed-bed reactors are resolved by fluidized bed reactors. In contrast to a fixed-bed reactor, the catalyst in fluidized bed reactors is carried in a fluid condition by the fluidizing gas as it travels over a distributer plate. Since the catalyst is well-mixed with the fluid, a greater surface area for the reaction to occur is provided and there is improved access to the catalyst [20]. Due to the lower operating cost, a fluidized bed reactor would therefore be the optimal reactor to deploy in the pilot plant on a conventional design size.
The conical spouted bed reactor (CSBR) offers good mixing and can handle larger particles, a wide range of particle densities and broad particle size distribution. Olazar et al. [21] asserted that a CSBR outperformed bubbling fluidized beds in terms of attrition and bed segregation. Additionally, it had a moderate de-fluidization issue while working with sticky solids and a high heat transfer rate between phases. However, several technical difficulties with the reactor’s operation, including catalyst feeding, catalyst entrainment and product (solid and liquid) collection, make it less advantageous.

2.1.1. Thermal Pyrolysis

In this process, depolymerization or cracking of plastics take place at a temperature ranging between 350–900 °C in absence of oxygen, thus producing liquid and gaseous fuel and carbonized char [22]. The condensable component of the volatile product is normally recovered after condensation as liquid fuel, while the remaining portion is a non-condensable gas of high calorific value. Thermal pyrolysis of plastic waste can used to convert the plastic waste to both liquid and gaseous fuel on an economically viable scale. Table 1 summarizes various studies that have been undertaken on thermal pyrolysis.
Table 1. Summary of research studies carried out by various research groups on thermal pyrolysis of plastic waste.

2.1.2. Microwave-Assisted Pyrolysis

In this process, a dielectric material is heated by microwave radiation and thus referred to as microwave pyrolysis. The interaction of microwave radiation with materials depends on their dielectric qualities. The incoming microwaves may be reflected as with conductors or transmitted as in the case of perfect insulators. They may even get absorbed and decay inside some material. Agitation of molecules in the presence of electromagnetic fields results in the generation of heat. With plastic, the microwave energy is absorbed through the absorbent and, then it transfers thermal heat to the plastic by conduction. The uniformity of heating dispersion is influenced by the absorbent’s physical qualities and volume ratio. Additionally, varying microwave power results in radically different product dispersion. Microwave-induced pyrolysis has been proven to be a viable method for producing value-added chemicals and fuels due to its significant advantages over thermal and catalytic pyrolysis [49]. The method could be used to heat plastics quickly, volumetrically and selectively for energy recovery. Plastics, meanwhile, are unable to absorb microwave energy due to their low dielectric loss factor. Therefore, to favour pyrolysis, plastic must be combined with the absorbent. Plastic materials with a high dielectric loss factor, like shredded tyres and silicon carbide, carbon and iron mesh, are ideal choices [32][33][34]. Various studies have been conducted on pyrolysis of plastic waste using microwave dielectric heating as shown in Table 2.
Table 2. Summary of studies on microwave pyrolysis of plastic waste.

2.1.3. Catalytic Pyrolysis

A catalyst speeds up the chemical reactions without affecting the process as it progresses. In industry and research, catalysts are frequently employed to enhance product selectivity and optimize product distribution. In order to produce products with significant economic value like automobile fuel (diesel and gasoline) and C2-C4 olefins, which are in high demand in the petrochemical industry, catalytic degradation is therefore particularly intriguing. When a catalyst is utilized, the process activation energy is reduced, accelerating the pace of reaction. As a result, a catalyst lowers the temperature needed, which is significant as the pyrolysis process consumes a lot of energy (it is extremely endothermic). Heat is one of the most expensive factors in any industry; therefore, using a catalyst may assist in saving energy. Additionally, numerous researchers employed catalysts for product improvement and to enhance the hydrocarbon distribution in order to produce pyrolysis liquid with qualities comparable to those of conventional fuels like gasoline and diesel.
Polymer degradation is brought about by catalytic pyrolysis in which the material is heated in the presence of catalyst and in absence of oxygen. Catalyst usage offers several advantages like reduced consumption of energy, modified composition of product on cracking and reduced process time. In literature, zeolites, fluid catalytic cracking (FCC) and silica-alumina based catalysts are mostly reported for pyrolysis of plastic waste [35][36][37]. Zeolite-based catalysts have a greater acid strength compared to non-zeolite catalysts, and they often accomplish higher conversion and produce more gaseous product.
There are two methods for catalytic upgradation of pyrolysis vapors: in-situ and ex-situ. In the in-situ approach, the catalyst is combined with the feedstock and added to the pyrolysis reactor, while in the ex-situ mode, the catalyst is added to a different reactor situated downstream of the pyrolysis reactor. In-situ catalytic pyrolysis does not require a separate catalytic reactor, thereby lowering the overall capital cost of setup manufacturing and operational costs. However, the catalyst deactivates more quickly if the feedstock has a high mineral and ash content [50].
The fact that the catalytic temperature is the same as the pyrolysis temperature, and that the optimal pyrolysis temperature may not be ideal for catalytic cracking, is another drawback of the in-situ catalyst application approach [51]. An ex-situ catalytic upgrading mode is thought to be more appealing for feedstock with a high ash concentration than an in-situ mode. Additionally, char may be obtained as a useful product and catalyst regeneration is also simple. An additional element that makes this mode more adaptable and hence desirable is the provision of separate temperature controls for pyrolysis and catalysis [52].
Either mode of adding the catalyst will result in the conversion of oxygenated molecules into stable hydrocarbons. However, the yield of important hydrocarbons in oil also varies based on the feedstock, the kind of reactor and the operating circumstances. An in-situ contact method increased aromatics production in the catalytic pyrolysis of polyethylene (PE), polypropylene (PP) and polyethylene terephthalate (PET), but an ex-situ mode was shown to be more successful in increasing aromatic yield in the case of polystyrene (PS) [50]. However, there are very little research examining the impact of the catalyst contact method in the simultaneous pyrolysis of plastic and biomass.

2.2. Gasification

Gasification is a thermo-chemical process involving multiple chemical reactions wherein a carbon containing feedstock like plastic waste is converted into synthetic gas in a partial supply of air, oxygen or steam [53]. The process operates at a sufficiently high temperature (>600–1000 °C) in order to thermally degrade the plastic waste for yielding syngas [54].
Several advantages are associated with gasification, like increased heating value of fuel by rejection of non-combustibles like nitrogen and water, reduction in oxygen content of fuel, exposure to H2 at high pressure or exposure to steam at high temperatures and pressures where H2 is added to the product will raise the products relative hydrogen content (H/C ratio).

2.2.1. Gasifying Medium

One of the important parameters to be considered during gasification of plastic waste is the gasifying medium. Gasification can be done in mediums like steam, air, carbon dioxide or oxygen or a combination of these gases [55]. The gasifying medium plays a significant role in converting solid carbon and heavier hydrocarbons into CO and H2, which are low molecular weight gases. Oxygen is primarily used as a gasifying medium in either pure form or via air, and the products include CO and CO2 when subjected to low and high oxygen, respectively. Gasification proceeds towards combustion when oxygen is supplied over a threshold limit, resulting in the formation of flue gas instead of synthesis/producer gas. The formed combustion product or the flue gases possess no residual heating value.
Similarly, when steam is used as a gasifying medium, the formed product contains more hydrogen per unit of carbon, thus increasing the H/C ratio. On the other hand, if air is directly used instead of oxygen, nitrogen present in air would dilute the product and the heating value of the gas would be reduced when compared to the heating value of the gas produced from oxygen/steam gasification. Thus, it can be concluded that the highest heating value is obtained when oxygen is used as a gasifying medium, followed by steam and air. The gases produced after air gasification is generally called producer gas whereas the gases produce through oxygen/steam gasification are known as synthesis gas.
Air gasification is a simple process and is advantageous as no external energy is required. Additionally, the output gas has less tar when compared to when steam gasification is used [56]. The syngas produced by steam gasification has a higher H2/CO ratio than the syngas produced by direct air gasification, making it more suitable for use in chemical synthesis applications [57]. The biggest difficulty with this method is the amount of heat needed to fuel the endothermic steam reforming reactions inside the reactor. Gasification with pure O2 is an alternative to air and steam as it includes the benefits of both gasifying agents. This method is more complex and expensive, especially for medium-scale applications, due to the high fixed assets and running costs for air separation [58]. More recently, it has been suggested that the pyrolysis and in-line reformation of pyrolysis volatiles is a potential method for valorizing waste plastics for production of H2 [31][59][60][61]. Additionally, this method uses very effective reforming catalysts to produce syngas that is entirely tar-free, addressing the primary difficulty in the traditional gasification of plastic.

2.2.2. Classification of Gasifiers

The primary criteria used to classify gasifiers are the gas-solid contacting mechanism and the gasification medium. Gasifiers have been divided into three main categories based on the manner of gas-solid contact: (1) Fixed or moving bed, (2) Fluidized bed, and (3) Entrained-flow bed. As illustrated in Figure 2, the direction in which the gasifying medium passes through the bed further categorizes each type of gasifier [6].
Figure 2. Classification of gasifiers.
A specific gasifier type might not be appropriate for the entire range of gasifier capacities. For example, moving-bed (updraft and downdraft) types are used for smaller units (under 10 MWth); fluidized-bed types are better suited for intermediate units (between 5 and 100 MWth); and entrained flow reactors are used for high capacity units (above 50 MWth). Major differences between the three gasifier types are summarized in Table 3.
Table 3. Major differences between the gasifier types.
In a fixed-bed gasifier, sometimes referred to as a moving-bed gasifier, the fuel is supported on a grate. Due to the fuel’s ability to slide down the gasifier like a plug, this type is also known as a moving bed. One of the main benefits of fixed-bed gasifiers is that they may be built in small quantities at low cost. Because of this, there are a lot of small-scale moving-bed biomass gasifiers in operation all over the world. It is challenging to produce equal distribution of fuel, temperature and gas composition over the cross-section of the gasifier due to poor mixing and heat transfer in the moving (fixed) bed. As a result, during gasification, fuels that are prone to agglomeration may form agglomerates. This is why large-capacity fixed-bed gasifiers are not very successful for biomass fuels or coal with a high caking index.
Fluidized-bed gasifiers are well known for temperature uniformity and mixing. A fluidized bed is made up of bed materials, which are granular particles that are retained in a semi-suspended form (fluidized state) by the movement of the gasifying medium across them at proper velocities. This type of gasifier is essentially unaffected by the quality of the fuel due to the superior gas solid mixing and the substantial thermal inertia of the bed. Additionally, the temperature homogeneity significantly lowers the possibility of fuel agglomeration. For gasifying biomass, the fluidized-bed concept has proven to be quite beneficial.

2.2.3. Other Gasifier Types

Spouted Bed

For waste valorization operations, spouted beds are an alternative to fluidized beds due to their specific characteristics like rapid heat and mass transfer rates, effective solid mixing and relatively better gas-solid contact. Additionally, their vigorous cyclic solid circulation prevents de-fluidization issues and makes it easier to handle sticky materials, irregular particles and particles with a wide size range. Their use in gasification processes is primarily constrained by the volatile’s short residence times, which impede tar cracking reactions. This technology has been extensively employed in bench scale setups for the pyrolysis of various solid wastes. The biomass pyrolysis process has been successfully scaled up to 25 kg per hour.

Plasma Gasifier

Waste materials are converted into gaseous products by plasma gasification operations in an oxidizing atmosphere. The primary benefit of plasma reactors for plastic gasification is the high temperature attainment, which encourages practically complete breaking of tar compounds and, consequently, high gas yields by improving the elimination of hazardous compounds. According to the methods used for plasma discharge, three kinds of plasma technologies—radio frequency, microwave, and direct current—have been established.

Pyrolysis-Reforming Process

This procedure, which aims to produce hydrogen, is carried out in two reactors that are connected in order to perform the pyrolysis and reforming processes. Most of the bench and laboratory scale units with various reactor configurations have been used to study this unique method. Czernik and French [62] conducted the ground-breaking research in a continuous system that consumed 0.06 kg of plastic every hour. The experimental unit consisted of a fluidized bed for plastic pyrolysis and a stationary bed for the catalytic reforming of the volatiles produced.
This method has several benefits over conventional steam gasification and steam reforming of bio-oil or plastic pyrolysis oil, including: (a) operation under optimum conditions because the two reactors are integrated into one unit (the reforming temperature is lower than that used in the gasification process, reducing potential catalyst sintering issues); (b) prevention of tar formation; and (c) direct contact between the feedstock and the catalyst. However, the effectiveness of the reforming catalyst will determine the development of the combined pyrolysis and in-line reforming process. Thus, to increase H2 production, obtain complete conversion of pyrolysis volatiles and prevent tar formation, highly active and selective catalysts are needed [63].

2.2.4. Gasification Reactions

The reactions that occur inside the gasifier are complicated reactions as shown in Table 4. The reactions are generally classified into five types: (i) carbon reactions, (ii) oxidation reactions, (iii) shift reactions, (iv) methanation reactions and (v) steam reforming reactions.
Table 4. Main gasification reactions [64].
An overview of several studies for the effect of temperature on gas production, as well as composition during gasification of plastic waste, can be found in Table 5.
Table 5. Summary of plastic waste gasification studies carried out by various research groups around the globe.
Syngas is rich in H2 and CH4, which can be used as combustible gases, as opposed to pyrolysis, which produces gas products that primarily consist of C3–C6. However, because of the dilution effect, the increased gas flow rate during gasification operation leads to lower throughputs, more difficult separation and lower calorific values of products, which has a detrimental influence on the overall economic benefits. Additionally, the gasification process harms the environment by producing noxious gases like NOX. Additionally, the produced syngas contains various pollutants such NH3, H2S, NOx, alkali metals and significant amounts of tars, necessitating an extra step in purification and raising the cost. A high-temperature environment is necessary for gasification, which raises the cost and requires the use of expensive machinery.

3. Advanced Oxidation Techniques for Treatment of Plastic Waste

3.1 Photocatalytic Oxidation

Photocatalysis is a technology that imitates the photosynthesis occurring in nature. Usually, there are three separate steps involved: (1) Charge carriers are excited via light absorption by photocatalysts, (2) the photogenerated charge carriers are separated and transported, and (3) the catalyst surface undergoes redox reaction. Photocatalytic oxidation involves oxidative breakdown of polymers into lower molecular weight materials in the presence of ultraviolet (UV) radiation and a photocatalyst which dominates the conversion process.
For photocatalysis to take place, incident photon energies must be greater than or equal to the band gap of the photocatalyst. While the reduction potential of photoelectrons in the conduction band should be greater than that of the reactant to be reduced, the oxidation potential of holes in the valence band of the photocatalyst should be greater than that of the substrate to be oxidised. For the photocatalytic valorisation of plastic waste, the holes in the photocatalysts’ valence band are frequently used to oxidise the plastic to produce organic products or degrade it to CO2, while the electrons in the photocatalyst’s conduction band can be used to reduce the protons in water to H2, CO2 to carbon derived fuels or capturing by oxygen to involve it in the subsequent plastic oxidation.

3.2. Electrocatalytic Oxidation

Conversion of plastics through electrocatalytic oxidation can be brought about using two different methods: direct oxidation and indirect oxidation. Direct oxidation refers to the electrophilic attack on a polymer by ·OH produced by water discharge on the anode surface. When strong oxidising intermediates dominate in the plastic conversion process, it is referred to as indirect oxidation. With an external voltage (0.55 V) applied in H3PO4 solution at 200 °C, polyvinyl alcohol (PVA) was successfully converted to H2 (9.5 mol/min) [77]. Also, for the production of carboxylic acid (75%), electrocatalytic degradation of PVC was carried out on TiO2/C cathode (−0.7 V) at 100 °C [78].

A plastic polymer is reduced when it receives electrons from the cathode (TiO2/C), which undergoes dechlorination at a high temperature. Additionally, the polymer is oxidised with ·OH to produce carbonyl and hydroxyl groups, which subsequently breaks down into tiny molecules (e.g., alcohols, carboxylic acids and esters). Finally, these chemicals partially mineralize to CO2 and H2O. Electrocatalytic breakdown of plastic wastes may result in a single product that could be turned directly into fuels. Electrolysis alone cannot yield as many fuel components as pyrolysis and electrolysis combined. 

3.3. Fenton Oxidation

The Fenton reaction is an advanced technique that is often used to degrade chemical compounds that are resistant to conventional methods. Fenton oxidation is the process of converting polymers into small molecules by generating ·OH from Fe2+ activated H2O2. The Fenton oxidation process for plastic conversion is usually carried out under mild reaction conditions. Due to rapid breakdown of hydrogen peroxide (H2O2), the conversion process is effective, making it advantageous for large-scale applications. However, the procedure uses a lot of H2O2, which increases the cost of operation. Through chemical oxidation reactions, the Fenton reaction can regulate the decomposition of polymer waste to create high-value fine chemicals.

References

  1. Weiland, F.; Lundin, L.; Celebi, M.; van der Vlist, K.; Moradian, F. Aspects of chemical recycling of complex plastic waste via the gasification route. Waste Manag. 2021, 126, 65–77.
  2. Bai, B.; Jin, H.; Fan, C.; Cao, C.; Wei, W.; Cao, W. Experimental investigation on liquefaction of plastic waste to oil in supercritical water. Waste Manag. 2019, 89, 247–253.
  3. Gallo, F.; Fossi, C.; Weber, R.; Santillo, D.; Sousa, J.; Ingram, I.; Nadal, A.; Romano, D. Marine litter plastics and microplastics and their toxic chemicals components: The need for urgent preventive measures. Environ. Sci. Eur. 2018.
  4. Brems, A.; Dewil, R.; Baeyens, J.; Zhang, R. Gasification of Plastic Waste as Waste-to-Energy or Waste-to-Syngas Recovery Route. In Solid Waste as a Renewable Resource; Apple Academic Press: Palm Bay, FL, USA, 2015; pp. 241–263. ISBN 9780429154485.
  5. Arena, U.; Di Gregorio, F.; Amorese, C.; Mastellone, M.L. A techno-economic comparison of fluidized bed gasification of two mixed plastic wastes. Waste Manag. 2011, 31, 1494–1504.
  6. Lopez, G.; Artetxe, M.; Amutio, M.; Alvarez, J.; Bilbao, J.; Olazar, M. Recent advances in the gasification of waste plastics. A critical overview. Renew. Sustain. Energy Rev. 2018, 82, 576–596.
  7. Uzoejinwa, B.B.; He, X.; Wang, S.; El-Fatah Abomohra, A.; Hu, Y.; Wang, Q. Co-pyrolysis of biomass and waste plastics as a thermochemical conversion technology for high-grade biofuel production: Recent progress and future directions elsewhere worldwide. Energy Convers. Manag. 2018, 163, 468–492.
  8. Wong, S.L.; Ngadi, N.; Abdullah, T.A.T.; Inuwa, I.M. Current state and future prospects of plastic waste as source of fuel: A review. Renew. Sustain. Energy Rev. 2015, 50, 1167–1180.
  9. 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. Manage. 2017, 197, 177–198.
  10. Mohanraj, C.; Senthilkumar, T.; Chandrasekar, M. A review on conversion techniques of liquid fuel from waste plastic materials. Int. J. Energy Res. 2017, 41, 1534–1552.
  11. 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.
  12. Kunwar, B.; Cheng, H.N.; Chandrashekaran, S.R.; Sharma, B.K. Plastics to fuel: A review. Renew. Sustain. Energy Rev. 2016, 54, 421–428.
  13. Okolie, J.A.; Nanda, S.; Dalai, A.K.; Berruti, F.; Kozinski, J.A. A review on subcritical and supercritical water gasification of biogenic, polymeric and petroleum wastes to hydrogen-rich synthesis gas. Renew. Sustain. Energy Rev. 2020, 119, 109546.
  14. Pichler, C.M.; Bhattacharjee, S.; Rahaman, M.; Uekert, T.; Reisner, E. Conversion of Polyethylene Waste into Gaseous Hydrocarbons via Integrated Tandem Chemical-Photo/Electrocatalytic Processes. ACS Catal. 2021, 11, 9159–9167.
  15. Weber, R.S.; Ramasamy, K.K. Electrochemical oxidation of lignin and waste plastic. ACS Omega 2020, 5, 27735–27740.
  16. Liu, L.; Zhao, H.; Andino, J.M.; Li, Y. Photocatalytic CO2 reduction with H2O on TiO2 nanocrystals: Comparison of anatase, rutile, and brookite polymorphs and exploration of surface chemistry. ACS Catal. 2012, 2, 1817–1828.
  17. 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.
  18. Eze, W.U.; Madufor, I.C.; Onyeagoro, G.N.; Obasi, H.C. The effect of Kankara zeolite-Y-based catalyst on some physical properties of liquid fuel from mixed waste plastics (MWPs) pyrolysis. Polym. Bull. 2020, 77, 1399–1415.
  19. Qureshi, M.S.; Oasmaa, A.; Pihkola, H.; Deviatkin, I.; Tenhunen, A.; Mannila, J.; Minkkinen, H.; Pohjakallio, M.; Laine-Ylijoki, J. Pyrolysis of plastic waste: Opportunities and challenges. J. Anal. Appl. Pyrolysis 2020, 152, 104804.
  20. Kaminsky, W.; Kim, J.S. Pyrolysis of mixed plastics into aromatics. J. Anal. Appl. Pyrolysis 1999, 51, 127–134.
  21. Olazar, M.; Lopez, G.; Amutio, M.; Elordi, G.; Aguado, R.; Bilbao, J. Influence of FCC catalyst steaming on HDPE pyrolysis product distribution. J. Anal. Appl. Pyrolysis 2009, 85, 359–365.
  22. Eze, W.U.; Umunakwe, R.; Obasi, H.C.; Ugbaja, M.I.; Uche, C.C.; Madufor, I.C. Plastics waste management: A review of pyrolysis technology. Clean Technol. Recycl. 2021, 1, 50–69.
  23. Al-Salem, S.M. Thermal pyrolysis of high density polyethylene (HDPE) in a novel fixed bed reactor system for the production of high value gasoline range hydrocarbons (HC). Process Saf. Environ. Prot. 2019, 127, 171–179.
  24. 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.
  25. Sharma, B.K.; Moser, B.R.; Vermillion, K.E.; Doll, K.M.; Rajagopalan, N. Production, characterization and fuel properties of alternative diesel fuel from pyrolysis of waste plastic grocery bags. Fuel Process. Technol. 2014, 122, 79–90.
  26. 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.
  27. López, A.; de Marco, I.; Caballero, B.M.; Laresgoiti, M.F.; Adrados, A. Influence of time and temperature on pyrolysis of plastic wastes in a semi-batch reactor. Chem. Eng. J. 2011, 173, 62–71.
  28. Fakhrhoseini, S.M.; Dastanian, M. Predicting pyrolysis products of PE, PP, and PET using NRTL activity coefficient model. J. Chem. 2013, 2013, 7–9.
  29. Williams, P.T.; Slaney, E. Analysis of products from the pyrolysis and liquefaction of single plastics and waste plastic mixtures. Resour. Conserv. Recycl. 2007, 51, 754–769.
  30. 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.
  31. 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.
  32. 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.
  33. Lee, K.H.; Noh, N.S.; Shin, D.H.; Seo, Y. Comparison of plastic types for catalytic degradation of waste plastics into liquid product with spent FCC catalyst. Polym. Degrad. Stab. 2002, 78, 539–544.
  34. Luo, G.; Suto, T.; Yasu, S.; Kato, K. Catalytic degradation of high density polyethylene and polypropylene into liquid fuel in a powder-particle fluidized bed. Polym. Degrad. Stab. 2000, 70, 97–102.
  35. Demirbas, A. Pyrolysis of municipal plastic wastes for recovery of gasoline-range hydrocarbons. J. Anal. Appl. Pyrolysis 2004, 72, 97–102.
  36. Sakata, Y.; Uddin, M.A.; Muto, A. Degradation of polyethylene and polypropylene into fuel oil by using solid acid and non-acid catalysts. J. Anal. Appl. Pyrolysis 1999, 51, 135–155.
  37. Adnan; Shah, J.; Jan, M.R. Thermo-catalytic pyrolysis of polystyrene in the presence of zinc bulk catalysts. J. Taiwan Inst. Chem. Eng. 2014, 45, 2494–2500.
  38. Uddin, M.A.; Koizumi, K.; Murata, K.; Sakata, Y. Thermal and catalytic degradation of structurally different types of polyethylene into fuel oil. Polym. Degrad. Stab. 1997, 56, 37–44.
  39. Williams, P.T.; Williams, E.A. Fluidised bed pyrolysis of low density polyethylene to produce petrochemical feedstock. J. Anal. Appl. Pyrolysis 1999, 51, 107–126.
  40. Miskolczi, N.; Bartha, L.; Deák, G.; Jóver, B.; Kalló, D. Thermal and thermo-catalytic degradation of high-density polyethylene waste. J. Anal. Appl. Pyrolysis 2004, 72, 235–242.
  41. 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.
  42. Elordi, G.; Olazar, M.; Lopez, G.; Artetxe, M.; Bilbao, J. Product yields and compositions in the continuous pyrolysis of high-density polyethylene in a conical spouted bed reactor. Ind. Eng. Chem. Res. 2011, 50, 6650–6659.
  43. 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.
  44. 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.
  45. Cho, M.H.; Jung, S.H.; Kim, J.S. Pyrolysis of mixed plastic wastes for the recovery of benzene, toluene, and xylene (BTX) aromatics in a fluidized bed and chlorine removal by applying various additives. Energy Fuels 2010, 24, 1389–1395.
  46. Grause, G.; Matsumoto, S.; Kameda, T.; Yoshioka, T. Pyrolysis of mixed plastics in a fluidized bed of hard burnt lime. Ind. Eng. Chem. Res. 2011, 50, 5459–5466.
  47. Bagri, R.; Williams, P.T. Catalytic pyrolysis of polyethylene. J. Anal. Appl. Pyrolysis 2002, 63, 29–41.
  48. Ali, M.F.; Ahmed, S.; Qureshi, M.S. Catalytic coprocessing of coal and petroleum residues with waste plastics to produce transportation fuels. Fuel Process. Technol. 2011, 92, 1109–1120.
  49. Kaminsky, W. Plastics recycling. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 1992; ISBN 3-52730-385-5.
  50. Xue, Y.; Johnston, P.; Bai, X. Effect of catalyst contact mode and gas atmosphere during catalytic pyrolysis of waste plastics. Energy Convers. Manag. 2017, 142, 441–451.
  51. Wan, S.; Wang, Y. A review on ex situ catalytic fast pyrolysis of biomass. Front. Chem. Sci. Eng. 2014, 8, 280–294.
  52. Luo, G.; Resende, F.L.P. In-situ and ex-situ upgrading of pyrolysis vapors from beetle-killed trees. Fuel 2016, 166, 367–375.
  53. Hu, M.; Guo, D.; Ma, C.; Luo, S.; Chen, X.; Cheng, Q.; Laghari, M.; Xiao, B. A novel pilot-scale production of fuel gas by allothermal biomass gasification using biomass micron fuel (BMF) as external heat source. Clean Technol. Environ. Policy 2016, 18, 743–751.
  54. Nyakuma, B.B.; Ivase, T.J.P. Emerging trends in sustainable treatment and valorisation technologies for plastic wastes in Nigeria: A concise review. Environ. Prog. Sustain. Energy 2021, 40, e13660.
  55. Salaudeen, S.A.; Arku, P.; Dutta, A. Gasification of Plastic Solid Waste and Competitive Technologies; Elsevier Inc.: Amsterdam, The Netherlands, 2018; ISBN 9780128131404.
  56. Gil, J.; Corella, J.; Aznar, M.P.; Caballero, M.A. Biomass gasification in atmospheric and bubbling fluidized bed: Effect of the type of gasifying agent on the product distribution. Biomass Bioenergy 1999, 17, 389–403.
  57. Erkiaga, A.; Lopez, G.; Amutio, M.; Bilbao, J.; Olazar, M. Syngas from steam gasification of polyethylene in a conical spouted bed reactor. Fuel 2013, 109, 461–469.
  58. Xiao, R.; Jin, B.; Zhou, H.; Zhong, Z.; Zhang, M. Air gasification of polypropylene plastic waste in fluidized bed gasifier. Energy Convers. Manag. 2007, 48, 778–786.
  59. Degnan, T.F. Applications of zeolites in petroleum refining. Top. Catal. 2000, 13, 349–356.
  60. Artetxe, M.; Lopez, G.; Amutio, M.; Elordi, G.; Bilbao, J.; Olazar, M. Cracking of high density polyethylene pyrolysis waxes on HZSM-5 catalysts of different acidity. Ind. Eng. Chem. Res. 2013, 52, 10637–10645.
  61. Garforth, A.A.; Lin, Y.H.; Sharratt, P.N.; Dwyer, J. Production of hydrocarbons by catalytic degradation of high density polyethylene in a laboratory fluidised-bed reactor. Appl. Catal. A Gen. 1998, 169, 331–342.
  62. Czernik, S.; French, R.J. Production of hydrogen from plastics by pyrolysis and catalytic steam reform. Energy Fuels 2006, 20, 754–758.
  63. Olazar, M.; Santamaria, L.; Lopez, G.; Fernandez, E.; Cortazar, M.; Arregi, A.; Bilbao, J. Progress on catalyst development for the steam reforming of biomass and waste plastics pyrolysis volatiles: A review. Energy Fuels 2021, 35, 17051–17084.
  64. Kannan, P.; Al Shoaibi, A.; Srinivasakannan, C. Energy recovery from co-gasification of waste polyethylene and polyethylene terephthalate blends. Comput. Fluids 2013, 88, 38–42.
  65. Arena, U.; Zaccariello, L.; Mastellone, M.L. Fluidized bed gasification of waste-derived fuels. Waste Manag. 2010, 30, 1212–1219.
  66. Martínez-Lera, S.; Torrico, J.; Pallarés, J.; Gil, A. Design and first experimental results of a bubbling fluidized bed for air gasification of plastic waste. J. Mater. Cycles Waste Manag. 2013, 15, 370–380.
  67. Sancho, J.A.; Aznar, M.P.; Toledo, J.M. Catalytic Air Gasification of Plastic Waste (Polypropylene) in Fluidized Bed. Part I: Use of in-Gasifier Bed Additives. Ind. Eng. Chem. Res. 2008, 47, 1005–1010.
  68. Arena, U.; Di Gregorio, F. Energy generation by air gasification of two industrial plastic wastes in a pilot scale fluidized bed reactor. Energy 2014, 68, 735–743.
  69. Lee, J.W.; Yu, T.U.; Lee, J.W.; Moon, J.H.; Jeong, H.J.; Park, S.S.; Yang, W.; Lee, U. Do Gasification of mixed plastic wastes in a moving-grate gasifier and application of the producer gas to a power generation engine. Energy Fuels 2013, 27, 2092–2098.
  70. Zaccariello, L.; Mastellone, M.L. Fluidized-bed gasification of plastic waste, wood, and their blends with coal. Energies 2015, 8, 8052–8068.
  71. Wilk, V.; Hofbauer, H. Conversion of mixed plastic wastes in a dual fluidized bed steam gasifier. Fuel 2013, 107, 787–799.
  72. Friengfung, P.; Jamkrajang, E.; Sunphorka, S.; Kuchonthara, P.; Mekasut, L. NiO/dolomite catalyzed steam/O2 gasification of different plastics and their mixtures. Ind. Eng. Chem. Res. 2014, 53, 1909–1915.
  73. Rutberg, P.G.; Kuznetsov, V.A.; Serba, E.O.; Popov, S.D.; Surov, A.V.; Nakonechny, G.V.; Nikonov, A.V. Novel three-phase steam-air plasma torch for gasification of high-caloric waste. Appl. Energy 2013, 108, 505–514.
  74. Park, Y.; Namioka, T.; Sakamoto, S.; Min, T.J.; Roh, S.A.; Yoshikawa, K. Optimum operating conditions for a two-stage gasification process fueled by polypropylene by means of continuous reactor over ruthenium catalyst. Fuel Process. Technol. 2010, 91, 951–957.
  75. Barbarias, I.; Lopez, G.; Artetxe, M.; Arregi, A.; Santamaria, L.; Bilbao, J.; Olazar, M. Pyrolysis and in-line catalytic steam reforming of polystyrene through a two-step reaction system. J. Anal. Appl. Pyrolysis 2016, 122, 502–510.
  76. Wu, C.; Williams, P.T. Pyrolysis-gasification of plastics, mixed plastics and real-world plastic waste with and without Ni-Mg-Al catalyst. Fuel 2010, 89, 3022–3032.
  77. Hori, T.; Kobayashi, K.; Teranishi, S.; Nagao, M.; Hibino, T. Fuel cell and electrolyzer using plastic waste directly as fuel. Waste Manag. 2020, 102, 30–39.
  78. Miao, F.; Liu, Y.; Gao, M.; Yu, X.; Xiao, P.; Wang, M.; Wang, S.; Wang, X. Degradation of polyvinyl chloride microplastics via an electro-Fenton-like system with a TiO2/graphite cathode. J. Hazard. Mater. 2020, 399, 123023.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 1.0K
Revisions: 3 times (View History)
Update Date: 22 Feb 2023
1000/1000
ScholarVision Creations