You are on page 1of 12

Journal of Hazardous Materials 424 (2022) 127460

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Review

Conversion of plastic waste into fuels: A critical review


Ning Li a, Hengxin Liu a, Zhanjun Cheng a, Beibei Yan a, Guanyi Chen a, b, c, 1, *, Shaobin Wang d
a
School of Environmental Science and Engineering/Tianjin Engineering Research Center of Bio Gas/Oil Technology, Tianjin University, Tianjin 300072, China
b
School of Mechanical Engineering, Tianjin University of Commerce, Tianjin 300134, China
c
School of Science, Tibet University, Lhasa 850012, Tibet Autonomous Region, China
d
School of Chemical Engineering, The University of Adelaide, Adelaide, SA 5005, Australia

A R T I C L E I N F O A B S T R A C T

Editor: Dr. R Teresa Plastic wastes have posed serious threats to the environment, including decrease of soil nutrient effectiveness and
agricultural production as well as emerge of ecological instability. Fuel conversion from plastic waste is regarded
Keywords: as a promising strategy for its disposal and energy utilization. Plastic wastes can be converted into target fuels by
Plastic adjusting cracking of chemical bonds. Currently, numerous technologies regarding fuel conversion from plastic
Fuel
wastes have been reported, including conventional pyrolysis, novel heat treatment and advanced oxidation.
Pyrolysis
However, systematic summary and comparative analysis of different technologies are still scarcely reported. In
Catalytic cracking
Hydrothermal liquefaction this review, fuel conversion from plastic wastes was summarized comprehensively, highlighting novel heat
Advanced oxidation treatment and advanced oxidation technologies reported in recent years. Furthermore, the superiority and
drawbacks of each technology were analyzed, and future prospects of technology application were proposed.
With lower reaction temperature and higher-value fuel, novel heat treatment of plastics is more popular than
traditional one. Advanced oxidation can be controlled to convert plastics into fuels under room temperature and
pressure, guiding the new normal in energy utilization of plastic wastes. This review aims to provide inspiration
for energy utilization of solid waste, addressing the issues of white pollution and energy shortage.

1. Introduction microplastics in the soil changes the isoelectric cohesive Coulomb and
van der Waals forces between soil particles, affecting the formation of
In recent years, the global demand of plastics has increased with the soil agglomerates (Machado et al., 2018). With plastic wastes mixed in
continuous development of industry, agriculture and fishery. As a result, the soil, the pores of the root cell wall are blocked. As a result, the ab­
more and more plastic wastes generate year by year. According to the sorption of water and nutrients via root system is affected, leading to
United Nations Environment Programme, the global production of reduction of output (Mbachu et al., 2021); (2) threaten the survival of
plastic wastes was about 360 million tons in 2018. As we all know, animals. Waste plastics on land surface or in water are easily taken by
plastic is a high molecular polymer with stable structure, which is hard animals, causing animal death; (3) produce microplastics and induce
to be destroyed by microorganisms in nature. As a kind of solid ecological security. The toxic effect of plastic wastes refers to organism
pollutant, plastic waste has posed serious threats to the environment in adverse reactions generated by the invasion from plastic to organisms.
the following aspects: (1) affect soil function and agricultural produc­ Microplastics can inhibit the expression of genes involved in oxidative
tion. The concentrations of dissolved organic nitrogen and phosphorus stress and transmembrane transport in the plant body, hence affecting its
decrease owing to the interaction between plastic wastes and humus, growth (Zhou et al., 2020). More seriously, microplastics hinder the
thus affecting soil nutrient effectiveness (Pathan et al., 2020). Besides, extracellular transmission of cellular transport carriers and cause human

Abbreviations: PP, Polypropylene; PS, Polystyrene; PE, Polyethylene; LDPE, Low density polyethylene; PET, Polyethylene terephthalate; HDPE, High density
polyethylene; RH, Rice husk; WS, Wheat straw; HZSM-5, Hydrogen zeolite socony mobile-five; FCC, Fluid catalytic cracking; SZ, Sulfated zirconia; Ecat-1, Com­
mercial catalyst-one; PVC, Polyvinyl chloride; PAM, Polyacrylamide; PLA, Polylactic acid; PUR, Polyurethane; CNX|Ni2P, Carbon nitride/nickel phosphide; PVA,
Polyvinyl alcohol; PCBs, polychlorodiphenyls.
* Corresponding author at: School of Environmental Science and Engineering/Tianjin Engineering Research Center of Bio Gas/Oil Technology, Tianjin University,
Tianjin 300072, China.
E-mail address: chen@tju.edu.cn (G. Chen).
1
Postal address: No.135, Yaguan Road, Jinnan District, Tianjin City, P.R. China.

https://doi.org/10.1016/j.jhazmat.2021.127460
Received 12 August 2021; Received in revised form 3 October 2021; Accepted 6 October 2021
Available online 8 October 2021
0304-3894/© 2021 Elsevier B.V. All rights reserved.
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

metabolic disorders (Nor et al., 2021). Also, microplastics ingested by et al., 2010). Combustible gases and liquid fuel can be used directly as
animals and accumulated in the gut, attacking the intestinal mucosa and secondary energy to realize energy utilization of plastic waste (Nanda
causing intestinal damage (Varo et al., 2021). On one hand, micro­ and Berruti, 2020). Hydrothermal liquefaction and catalytic cracking
plastics refer to plastic particles less than 5 mm in size. Owing to the are regarded as novel heat technologies. In this way, plastic waste is
great specific surface area of microplastics, it is prone to adsorb other converted to gaseous hydrocarbons and >C13 non-aromatic liquid fuel at
pollutants to trigger combined toxic effects. On the other hand, micro­ 250–450 ◦ C (Shen, 2020; Budsaereechai et al., 2019). Furthermore,
plastic belongs to polymer and requires more than 100 years to be plastic polymers are decomposed to H2 in the presence of ⋅OH species in
degraded in nature. The toxic effect of microplastics accumulates advanced oxidation systems (Weber and Ramasamy, 2020). At the same
gradually in food chain, affecting the stability of ecosystem (Zhang et al., time, organic products such as carboxylic acid and sulfonic acid are
2017); (4) cause white pollution. The plastic wastes in the environment formed, making plastic waste a carbon resource for the chemical in­
can affect the cityscape and pose a visual shock to people; (5) trigger dustry. In addition, the CO2 produced in the process of advanced
diseases. Plastic wastes in the ocean show higher charge and hydro­ oxidation, can be further converted to acetic acid with photoreduction
philicity than that on land, which can be converted into microplastics in or electrocatalysis (Liu et al., 2012; Zhao et al., 2017).
the presence of microorganisms. Bacteria could accumulate and adhere Reviews on energy conversion of plastic waste into C5-C20 aromatic
to the surface of plastic wastes easily, forming micro colonies and and aliphatic hydrocarbon fuels have been reported by pyrolysis and
resulting in damage to organisms in the ocean. For instance, with catalytic cracking (Al-Salem et al., 2017; Uzoejinwa et al., 2018; Wong
physical injury from plastic wastes to coral, genus Vibrio on plastic et al., 2015). The effects of pyrolysis and catalytic cracking parameters
wastes can invade coral tissues and induce white syndrome (Lamb et al., on the composition and properties of target fuels were summarized
2018; Rasool et al., 2021). Also, accidental ingestion of microplastics by (Mohanraj et al., 2017; Anuar Sharuddin et al., 2016; Kunwar et al.,
fish induces a false sense of satiety, thus reducing their food intake. 2016). Noticeably, many new technologies have been reported
Decrease of fish population size and individual quality occurs due to regarding the conversion of plastics into fuels in recent years, such as
unavailability of growth energy (Mallik et al., 2021). Furthermore, hydrothermal liquefaction and advanced oxidation (Okolie et al., 2020;
microplastics in the soil can be absorbed by plants, leading to growth Pichler et al., 2021). In hydrothermal liquefaction, plastic wastes can be
retardation and plant height inhibition (Xu et al., 2020). People feed on converted into superior oils at relatively lower temperature compared to
plants, resulting in microplastics transfer to human body, which can conventional pyrolysis. Besides, high purity oil can generate by plastic
cause intestinal inflammation and affect human growth and develop­ conversion via advanced oxidation at room temperature and pressure.
ment readily (Yee et al., 2021). Therefore, it is particularly important for However, systematic summary of new technologies and comparison of
the proper treatment and disposal of plastic wastes. different processes have not yet been published. This review provides
Large amounts of dioxins and polychlorodiphenyls (PCBs) can readers with novel technologies for fuel conversion from plastic wastes
generate via incineration of plastic wastes. Dioxins absorbed by human and clarifies the superiority and drawbacks of different systems. In
respiratory tract trigger reproductive decline and blurred vision. The addition, more efficient processes can be designed based on different
dynamic balance of reactive oxygen species in plants is disrupted by types of plastic wastes, promoting target fuels generation by further
PCBs readily, causing free radical accumulation and plant damage. PCBs parameter regulation in the system. Herein, this review focused on novel
ingestion by animals show acute hepatic necrosis and bone deformation heat treatment and advanced oxidation technologies for energy utili­
owing to PCBs accumulation in fatty tissues (Bursian et al., 2018; Sub­ zation of plastics. The advantages and shortcomings of different tech­
ramanian et al., 2017; Yao et al., 2017). Also, landfill is one of the nologies were also analyzed by comparison. Furthermore, the outlook of
common ways for plastic wastes treatment currently. Heavy metals (e.g., future applicable technologies was conducted. Also, the unresolved is­
cadmium, copper and mercury) and pathogens carried by plastic wastes sues and development directions were proposed. This review aims to
enter the groundwater with rainwater washing. The contaminated improve summaries of technologies for converting waste plastics into
groundwater can induce heart failure and bone softening after humans fuels, promoting the harmless disposal and energy utilization of organic
and animals consumption (Sall et al., 2020). solid waste. This review could provide inspiration and reference for
The energy utilization of plastic wastes is one of the effective ways researchers in the field of solid waste utilization.
for its proper disposal. Fuel conversion from plastics can reduce harmful
emissions and pathogen contaminants compared to incineration and 2. Conversion technologies of plastic wastes into fuels
landfill disposal. The species and density of oxidants can be regulated by
reaction parameters (e.g., temperature, reaction time and temperature 2.1. Traditional heat treatment
increasing rate) in fuel conversion from plastic wastes (Yin et al., 2021).
Furthermore, chemical bonds are broken selectively, and generate The pyrolysis process can realize the conversion of plastic wastes into
C4-C18 gasoline hydrocarbon and diesel hydrocarbon fractions (Dogu fuel by anaerobic heating at 350–600 ◦ C. By providing thermal energy to
et al., 2021). In this way, most of the carbon and hydrogen elements are meet the activation energy required for polymer cracking, the plastic
stored in the liquid fuel, reducing the formation of polycyclic aromatic waste is converted into combustible gas and liquid oil. Liquid oil could
hydrocarbons effectively (Gin et al., 2021; Liu et al., 2021). Also, fuel be obtained from the pyrolysis of polyvinyl chloride (PVC), of which the
conversion from plastic wastes via heat treatment is conducted with calorific value was as high as 40 MJ/kg (Miranda et al., 1999). By
anoxic condition at 250–600 ◦ C (Banu et al., 2020). Most of aerobic comparison, the calorific value of non-aromatic fuel produced by py­
pathogens intolerant to high temperature are eliminated during the heat rolysis of polystyrene (PS) at 350 ◦ C was slightly lower (40.6 MJ/kg)
treatment of plastic wastes. (Bajad et al., 2017). Obviously, the calorific value of fuel obtained from
Currently, plastic wastes can be converted into fuels, fine chemical polypropylene (PP) pyrolysis is closer to that of gasoline (44 MJ/kg),
feedstocks and functional materials via appropriate technologies (Chen appearing as an alternative energy after treatment. PS pyrolysis at
et al., 2021; Lin et al., 2010; Singh et al., 2020). Traditional heat 500 ◦ C could produce 96.0% of aromatic hydrocarbon, of which C5-C12
treatment, novel heat treatment and advanced oxidation technologies liquid gasoline hydrocarbon accounted for 82.5% (Zhang et al., 2020).
have been reported for fuel conversion from plastic wastes (Fadillah The product composition via polyethylene (PE) pyrolysis at 400 ◦ C was
et al., 2021; Hibino et al., 2021; Dwivedi et al., 2021). Initially, gasifi­ studied (Sarker et al., 2012b). Results showed the percentage of C3-C27
cation as a conventional pyrolysis technology, could convert plastic volatile olefins reached 97.0%. It can be seen that the pyrolysis oil from
wastes into liquid fuel of C5-C20 aromatic and aliphatic hydrocarbon at plastic wastes contains a large amount of gasoline and diesel fractions.
350–600 ◦ C under oxygen deprivation (Papari et al., 2021). Moreover, However, it is unsuitable for direct use as fuel due to wide carbon
combustible gases such as H2, CO and CH4 could also generate (Panda number distribution, low octane value of gasoline and large number of

2
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

olefins. Hence, further upgrading of the pyrolysis oil is needed. As shown


in Table 1, mixed plastic wastes pyrolyzed to obtain chlorinated oil,
which was further refined to light liquid fraction including alkanes and
aromatics. The chlorine content of the light fraction reduced and the oil
composition was closer to gasoline and diesel products after refining
(Lopez-Urionabarrenechea et al., 2015). Overall, fuels with high calo­
rific value (>40 MJ/kg) can generate by plastic pyrolysis at scale, alle­
viating energy crisis potentially.
The efficiency of fuel conversion from plastic wastes is mainly
related to heating rate, pyrolysis temperature and feed composition. The
effect of heating rate on fuel conversion process of low density poly­
ethylene (LDPE) was investigated. It was found that the yield of non-
aromatic liquid oil decreased from 80.4% to 76.5% with ramp rate
increasing from 6 ◦ C/min to 10 ◦ C/min. The yield of the above products
further decreased to 71.1% with heating rate increasing to 14 ◦ C/min
(FakhrHoseini and Dastanian, 2013; Singh et al., 2019a) pyrolyzed
PE/PP/PS mixed plastic wastes slowly (10 ◦ C/min) at 500 ◦ C and found
the major products were C5-C14 aromatics and aliphatic hydrocarbons.
Comparatively, the mixed waste plastics were mainly converted to
gaseous products such as methane and ethane under fast pyrolysis
(20 ◦ C/min). In addition, pyrolysis of PE and polyethylene terephthalate
(PET) was conducted at heating rate of 15 ◦ C/min and 20 ◦ C/min,
respectively. PE was converted into >C13 non-aromatic high-value fuels
while PET was turned into aromatic fuels (Cheng et al., 2020; Singh
et al., 2019b). Therefore, slow pyrolysis can promote the conversion of
plastic into high-value liquid fuels and reduce the formation of
by-products.
As shown in Fig. 1a, the effect of pyrolysis temperature on fuel
production from mixed plastics (PP and LDPE) was investigated. The
yield of C5-C12 gasoline fraction increased from 18.3% to 48.6% with
temperature rising from 220 ◦ C to 275 ◦ C (Shah et al., 2010). Besides,
the yield of diesel hydrocarbon in PP pyrolysis rose from 57.3% to 69.8%
with the increase of pyrolysis temperature from 250 ◦ C to 300 ◦ C. The
production of naphthenic hydrocarbons increased while that of diesel
hydrocarbon decreased to 63.2% under higher temperature (400 ◦ C) Fig. 1. (a) Effect of pyrolysis temperature on oil yield from PP and LDPE; (b)
Pyrolysis mechanism of mixed plastic wastes. The Fig. 1(a) is taken from
(Ahmad et al., 2014). The impact of temperature on PP pyrolysis in a
Ref. (Prurapark et al., 2020) with copyright permission.
vertical cracking reactor was investigated. The fuel yield declined from
74.4% to 50.3% when the temperature increased from 550 ◦ C to 625 ◦ C
(Jin et al., 2018). After being understood the influence of related pa­ the diesel percentage by pyrolysis of PP alone was 46.3%, which
rameters, high density polyethylene (HDPE) and PET have been con­ improved by 9.3% and 5.2% after mixing with RH and WS (Suriapparao
verted to kerosene, gasoline and diesel fractions successfully (Prurapark et al., 2020). It was reported that fuel could be obtained by co-pyrolysis
et al., 2020). Consequently, high temperature pyrolysis (>550 ◦ C) of of red oak with HDPE at 575 ◦ C, of which the calorific value reached
plastic wastes is detrimental to the formation of liquid hydrocarbons. It 35.2 MJ/kg, much higher than that with HDPE pyrolysis alone (Xue
is crucial to select pyrolysis temperature rationally according to the et al., 2015). Co-pyrolysis of biomass and plastics required less micro­
target products. wave energy or thermal energy. Also, the loss of thermal conductivity
The effect of feedstock on the pyrolysis of PS and PP was studied. It was low, which is conducive to generating high-value fuels and
was observed that the yield of diesel fraction obtained by pyrolysis of improving the pyrolysis efficiency. In the future, co-pyrolysis of plastic
single PS was 20.8%, which increased slightly to 27.4% and 26.7% after wastes with other biomass can be tried to enhance the target fuel for­
addition of rice husk (RH) and wheat straw (WS) respectively. Besides, mation to the desired level.

Table 1
Fuel conversion from plastics via traditional heat treatment.
Type of plastics Conditions Products Yield (%) Refs.

PET T = 400/425/450 ◦ C C10-C14 kerosene / (Prurapark et al., 2020)


HDPE T = 575 ◦ C, Ratio of plastics to biomass = 1:4 Liquid oil 57.6 (Xu et al., 2020)
PE/PP/PS (Mixture) T = 500 ◦ C, Heating rate = 20 ◦ C/min C5-C19 liquid oil 50.9 (Singh et al., 2019a)
PS T = 500 ◦ C, Filling ratio of heat carrier = 15% C5-C12 liquid oil 82.5 (Zhang et al., 2017)
PE T = 400 ◦ C Liquid oil 92.0 (Sarker et al., 2012a)
LDPE/PP (Mixture) T = 275 ◦ C, Time = 80 min C5-C12 gasoline fraction 48.6 (Shah et al., 2010)
PP T = 550 ◦ C Liquid oil 74.4 (Jin et al., 2018)
LDPE T = 500 ◦ C, Heating rate= 6 ◦ C/min Liquid oil 80.4 (FakhrHoseini and Dastanian, 2013)
PP T = 300 ◦ C C6-C16 naphtha 50.0 (Ahmad et al., 2014)
PP T = 450 ◦ C, Ratio of plastics to RH = 1:1 C6-C14 diesel 55.6 (Suriapparao et al., 2020)
PS T = 350 ◦ C, Time = 40 min Liquid oil 88.0 (Bajad et al., 2017)
PE/PP/PS/PET/PVC (Mixture) T = 500 ◦ C, Time = 30 min C5-C16 liquid oil 87.0 (Lopez-Urionabarrenechea et al., 2015)

PET: Polyethylene terephthalate; HDPE: High density polyethylene; PE: Polyethylene; PP: Polypropylene; PS: Polystyrene; LDPE: Low density polyethylene; PVC:
Polyvinyl chloride; T: Temperature; RH: Rice husk.

3
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

The conversion mechanism of mixed plastic wastes to fuel by py­ In catalytic cracking process, the plastic conversion is regulated by
rolysis is shown in Fig. 1b. Heat is absorbed by plastics, inducing temperature, flow rate of scrubbing gas, ratio of catalyst to plastic and
thermal-initiation reactions at 350–600 ◦ C (Eq. (1)). Long-chain free catalyst properties. The effect of temperature on the conversion of HDPE
radicals generated by random breakage of PE and PP, or end fracture of to fuels with HZSM-5 catalyst was studied. As shown in Table 2. The
PS, which can be converted to paraffin and diolefin via hydrogen yield of C12-C18 diesel fraction raised from 14.3% to 41.7% as temper­
transfer (Eq. 2). Besides, Vinyl can be formed through beta-scission of ature increased from 400 ◦ C to 450 ◦ C (Miskolczi et al., 2004). Also,
long chain radicals. Moreover, the reorganization and aromatization HDPE could be converted to C7-C16 alkanes by HZSM-5 catalysis at
occur in vinyl, promoting the formation of alkanes and olefins (Eqs. 400 ◦ C (Fig. 2a). The main products were aromatics and alkanes after
(3–8)). When free radicals, cations and anions are presented in the ter­ the temperature further increased by 40 ◦ C (Pan et al., 2018). In addi­
minal of plastic molecules, the bond energy at this position is distinct tion, by optimizing the pyrolysis temperature, the conversion of PP to
from that of the neighboring groups. Hence, the plastic is susceptible to C5-C9 gasoline was achieved at 360 ◦ C (Sarker et al., 2012a). Therefore,
suffer end-break degradation. In this model, pyrolysis starts at the end of in contrast to conventional pyrolysis, the decomposition of plastics can
plastic molecular chain. The molecular weight of the plastic diminishes be accelerated by catalytic cracking at relatively low temperatures. The
slowly during the reaction. Besides, a large amount of monomeric matter effect of temperature on HDPE conversion by FCC catalyst was investi­
releases. In general, most of the vinyl polymers in the α-substituted gated. The proportion of C5-C12 aliphatic hydrocarbons decreased from
position are degraded via this reaction pattern. Random degradation 86.2% to 81.1% with temperature increasing from 420 ◦ C to 510 ◦ C
occurs anywhere in the polymer chain, and the plastic is decomposed (Abbas-Abadi et al., 2013). Besides, the HDPE catalytic cracking was
into small molecular fragments. However, no monomeric substance is conducted with silica alumina. The yield of C6-C8 cycloalkanes
produced during the degradation process and the active site is not decreased from 75.3% to 72.9% as the temperature increased from
required in the polymer chain to drive the reaction. Typically, vinyl 400 ◦ C to 500 ◦ C (Luo et al., 2000). After optimising the related pa­
polymers such as polystyrene and polyacrylonitrile are fragmented by rameters, PE was converted into C6-C9 aliphatic hydrocarbons success­
random chain breaking mode. The main reaction equations of mixed fully (Fushimi and Gaffney, 2014). Hence, relatively high temperature
plastic wastes in pyrolysis are listed as follows: (>400 ◦ C) can inhibit the formation of gasoline hydrocarbons.
The influence of N2 flow rate on catalytic cracking of plastic waste
(1) Thermal-initiating reaction: was summarized. Mixed plastic wastes (PE/PP/PS) was converted into
fuels with HZSM-5 catalyst at 360 ◦ C. The yield of C1-C9 aliphatic and
(1)

− R − (R)n − R − → − R − (R)n − R aromatic hydrocarbons increased from 90.3% to 91.2% with N2 flow
rate rising from 270 mL/min to 420 mL/min. When the N2 flow rate
reached 900 mL/min, the production of the above fuels decreased to
(2) Chain-breaking reaction:
88.1% (Lin and Yang, 2005). Also, a mixture of PE, PP and PS plastic
− R − (R)n − R∙ → − R − (R)n − R∙ + R (2) wastes was decomposed via fluid catalytic cracking (FCC). With growth
in the flow rate of N2 from 300 mL/min to 600 mL/min, the output of
combustible gases (e.g., benzene, styrene, toluene and xylene) increased
(3) Chain-terminating reactions: by 1.1%. Further enlarging the N2 flow rate to 900 mL/min, a 2.2%
increase of combustible gases yield was realized (Lin, 2009). Therefore,
R∙ + R∙ → R− R (3)
the cost of plastic wastes conversion would decrease via selecting
R∙ + RO∙→ OR (4) appropriate flow rate of gas.
The catalyst composition plays an important role in LDPE conver­
R∙ + ROO∙→ ROOR (5) sion. The catalytic decomposition of LDPE by Pt loading sulfated zir­
RO∙ + RO∙ → ROOR (6) conia (SZ) at 250 ◦ C was investigated. With the growth of Pt
concentration in SZ from 0.4% to 1.2%, the output of C5-C12 alkanes
RO∙ + ROO∙→ ROR + O2 (7) increased by 16.2% (Utami et al., 2018). The catalytic cracking of mixed
plastics (PP/PE/PS) by different Ni/Si-Al catalysts was conducted at
ROO∙ + ROO∙→ ROOR + O2 (8)
410 ◦ C. The content of C5-C12 aliphatic hydrocarbons rose from 70.2%
to 78.5% with the increase of molar ratio of Si and Al from 0:1–1:1
(Wang and Wang, 2011). Obviously, the excellent performance of metal
catalysts was shown in the catalytic cracking of plastic wastes, which
2.2. Novel heat treatment promoted the formation of C5-C12 aliphatic hydrocarbons.
The catalyst mass presents significant influences on catalytic
2.2.1. Catalytic cracking cracking of plastics. The conversion of mixed plastics (PE/PP/PS) was
Catalytic cracking refers to the heating process of mixed catalysts investigated with different mass ratios of catalysts at 400 ◦ C. As the mass
with plastic wastes, aiming to obtain liquid oil. By adding catalysts, the ratio of commercial catalyst (ECat-1) to plastic waste increased from
activation energy can be reduced. As a result, the cracking of plastic 1:10–3:5, the yield of C5-C9 aromatics and C1-C4 combustible gases
wastes and the production of liquid oil are facilitated. With high calorific decreased by 5.3% and 1.6%, respectively (Lin and Yang, 2007). The
value and excellent performance, cracked oil has the potential to sub­ catalytic cracking of PS by SiO2/Al2O3 at 300 ◦ C was reported. A 9.1%
stitute fossil energy such as oil and diesel. The composition and prop­ decrease of styrene output was found owing to an increase of the mass
erties of PS cracked oil under natural zeolites catalysis was examined. ratio of catalyst to PS from 1:20–1:10 (Owusu et al., 2018). Moreover,
The cracked oil was composed of aromatic hydrocarbons, including with the mass ratios of catalysts (BaTiO3、HZSM-5 and CuCO3) to
styrene, ethylbenzene, benzene, asphaltene, and toluene. The calorific polymer reaching 1:19, 1:10 and 1:9, LDPE and HDPE were converted to
value of the obtained fuel was as high as 45 MJ/kg, which is extremely C5-C9 aromatic and >C13 non-aromatic fuels respectively (Ahmad et al.,
similar to the quality of conventional diesel fuel (Miandad et al., 2017). 2013; Marcilla et al., 2009; Singh et al., 2018). Once the mass ratio of
Besides, PP cracking was performed under azoisobutylnitrile catalysis, catalyst to plastics is exceeded the optimal value, agglomeration of
obtaining C8-C12 aliphatic hydrocarbons and cycloalkanes (Siddiqui, catalysts occurs, leading to reduced surface area, lower catalytic effi­
2009). In comparison with direct pyrolysis, catalytic cracking of plastics ciency and easier generation of carbon. Therefore, the catalyst dosage in
could generate liquid oil with narrower carbon number distribution. the catalytic cracking of plastics should not be too large. Catalytic
Moreover, the yield of gasoline and kerosene fractions was enhanced. cracking of PP could produce diesel fraction with Ni-Mo/Laponite

4
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Table 2
Fuel conversion from plastics via novel heat treatment.
Type of plastics Catalyst Conditions Products Yield Refs.
(%)

PE/PP/PS Commercial equilibrium T = 400 ◦ C, Time = 30 min, N2 flow rate = 600 mL/min, C5-C9 gasoline 56.5 (Liang et al., 2013)
(Mixture) catalyst RCP = 1: 10
HDPE BaTiO3 T = 350 ◦ C, Time = 30 min, RCP = 1: 19 C6-C16 liquid oil 68.6 (Ahmad et al., 2013)
PE/PP/PS Ni/Si-Al T = 410 ◦ C, Time = 60 min, Ratio of Si to Al= 1: 1 C5-C12 gasoline 78.5 (Wang et al., 2014)
(Mixture)
PE/PP/PS HZSM-5 T = 360 ◦ C, Time = 20 min, N2 flow rate = 420 mL/min, C1-C4 gases 91.2 (Lin and Yang, 2005)
(Mixture) RCP = 2: 5
HDPE Ni/Al2O3 T = 400 ◦ C, Time = 120 min, RCP = 1: 50, P = 1 Mpa C7-C27 liquid oil 78.1 (Pan et al., 2018)
PS Natural zeolite T = 450 ◦ C, Time = 75 min, RCP = 1: 10, Heating rate Ethylbenzene 60.8 (Miandad et al., 2017)
= 10 ◦ C/min
PE/PP/PS Commercial equilibrium T = 390 ◦ C, N2 flow rate = 900 mL/min, RCP = 3: 7 Combustible 84.6 (Lin, 2009)
(Mixture) catalyst gases
LDPE Pt-promoted SZ T = 250 ◦ C, Pt ratio = 1.5% Olefins 55.6 (Utami et al., 2018)
PS SiO2/Al2O3 T = 300 ◦ C, Time = 120 min, RCP = 1: 20 Liquid oil 59.6 (Owusu et al., 2018)
HDPE HZSM-5 T = 550 ◦ C, RCP = 1: 10, Heating rate = 5 ◦ C/min Combustible 72.6 (Marcilla et al., 2009)
gases
HDPE HZSM-5 T = 450 ◦ C, Time = 60 min, RCP = 3: 97 C12-C18 diesel 41.7 (Miskolczi et al., 2004)
HDPE CuCO3 T = 500 ◦ C, RCP = 1: 19 Liquid oil 94.0 (Singh et al., 2018)
PP HZSM-5 T = 360 ◦ C, RCP = 1: 19 C3− C25 liquid oil 92.0 (Sarker et al., 2012a)
LDPE Cu@TiO2 T = 300 ◦ C, Time = 30 min, RCP = 1: 19 C13-C19 liquid oil 86.4 (Ukarde and Pawar, 2021)
PE sugarcane bagasse T = 280 ◦ C, Time = 60 min, RCP = 17: 3 Bio-oil 31.5 (Baloch et al., 2020b)
HDPE / T = 425 ◦ C, Time = 60 min, P = 0.62 Mpa, S = Deionized Diesel 45.0 (Jin et al., 2020)
water
PS / T = 490 ◦ C, Time = 10 min, P = 30 Mpa, S = Deionized Ethylbenzene 50.0 (Bai et al., 2019)
water
PS / T = 350 ◦ C, Time = 30 min, P = 25 Mpa, S = Pure water Liquid oil 85.4 (Sarker et al., 2012b)
LDPE sugarcane bagasse T = 280 ◦ C, Time = 75 min, S = Ethanol Bio-oil 32.3 (Baloch et al., 2020b)
PP/PS/PET / T = 300 ◦ C, Time = 30 min, P = 25 Mpa, S = Deionized Liquid oil 32.0 (Seshasayee and Savage,
(Mixture) water 2021)
PS / T = 350 ◦ C, Time = 60 min, S = Ethanol Liquid oil 84.7 (Ahmad et al., 2020)
PP / T = 425 ◦ C, Time = 120 min, P = 23 Mpa, S = Deionized C6-C13 liquid oil 64.7 (Chen et al., 2019)
water

PE: Polyethylene; PP: Polypropylene; PS: Polystyrene; HDPE: High density polyethylene; LDPE: Low density polyethylene; PET: Polyethylene terephthalate; HZSM-5:
Hydrogen zeolite socony mobile-five; T: Temperature; RCP: Ratio of catalyst to polymer; P: Pressure; S: Solvent; SZ: sulfated zirconia.

catalyst (mass ratio of Ni-Mo/Laponite to PP = 1:10), of which C10-C25 decreased to 76.3% with temperature rising to 325 ◦ C (Ukarde and
aliphatic hydrocarbons accounted for 64.1% (Mangesh et al., 2020). Pawar, 2021). As shown in Fig. 3a, hydrothermal liquefaction of mixed
Also, PP cracking with FCC catalyst was studied at 1:10 mass ratio of plastic wastes (PP, PS and PET) was conducted at 25 MPa. The liquid oil
catalyst to PP. Results showed the percentage of C12-C22 aliphatic hy­ yield decreased from 32.0% to 30.1% with temperature increasing from
drocarbons reached 85.8% (Lee et al., 2002). Therefore, FCC catalyst is 300 ◦ C to 350 ◦ C (Seshasayee and Savage, 2021). Therefore, it is easier
more conducive to diesel conversion from PP than Ni-Mo/laponite. for hydrothermal liquefaction of plastic wastes to fuels at relatively low
The catalytic cracking process of PE, PP, PS, PET and PVC mixture is temperature (300 ◦ C).
illustrated in Fig. 2b. The mechanism of catalytic cracking is similar to Reaction time could influence oil production via hydrothermal
that of conventional pyrolysis. In catalytic cracking process, thermal- liquefaction of plastics. The effect of reaction time on fuel generation
initiating, chain-breaking and chain-terminating reactions are during hydrothermal liquefaction of HDPE was investigated. The con­
included. Notably, thermal-initiating reaction is initiated at 250–450 ◦ C centration of n-alkanes and cycloalkanes products increased within
and the paraffins obtained from β-scission can be converted to short- 40–60 min and reached a peak (43.0%) in 90–120 min. In the reaction
chain olefins via catalyst. In comparison with conventional pyrolysis, time range of 120–240 min, the yield of n-alkane was almost constant
catalytic cracking can generate C5-C16 aliphatic and C8-C18 aromatic while that of aromatic hydrocarbons increased (Jin et al., 2020). Be­
hydrocarbons with higher quality. sides, hydrothermal liquefaction of PS was performed at 400 ◦ C and
30 MPa. Styrene and propane were main products within 20 min.
2.2.2. Hydrothermal liquefaction Extending the reaction to 60 min, the main products were toluene and
Hydrothermal liquefaction refers to the conversion of plastics into ethylbenzene, accompanied with the formation of polycyclic aromatic
liquid oil in subcritical or supercritical water at 280–450 ◦ C and hydrocarbons (Bai et al., 2019). Therefore, plastic wastes are converted
7–30 MPa. Hydrothermal liquefaction is influenced by temperature, into fuels by hydrothermal liquefaction, which can be applied in prac­
reaction time and solvent. The effect of temperature on PP liquefied oil tical situations after refining. To avoid coking of products, the time of
was studied at 23 MPa. With temperature increasing from 380 ◦ C to hydrothermal liquefaction should not be too long. Hydrothermal
450 ◦ C, the content of C6-C13 aliphatic hydrocarbons decreased while liquefaction of LDPE was performed at 425 ◦ C. The yield of C5-C12
that of aromatic hydrocarbons increased (Chen et al., 2019). In the aliphatic hydrocarbons increased from 62.3% to 87.2% as reaction time
process of PET conversion, benzoic acid changed from minor to major rising from 30 min to 60 min. However, the product yield decreased to
product due to increase of temperature from 400 ◦ C to 450 ◦ C (Sesha­ 77.7% with time further prolonging to 120 min (Čolnik et al., 2021).
sayee and Savage, 2020). Apparently, high temperature promotes the Also, the production of C5-C12 liquid oil from PE could be realized at
hydrothermal liquefaction of plastic wastes to aromatic compounds. 400 ◦ C. The yield of C5-C12 liquid oil was improved by 5.2% in
LDPE was transformed into liquid oil in an autoclave. The yield of 30–60 min and then declined slightly to 24.5% at 120 min (Song et al.,
C13-C19 cycloalkanes and aromatic hydrocarbons increased from 17.2% 2021). Hence, long-term hydrothermal liquefaction is not conducive to
to 86.1% with temperature increasing from 250 ◦ C to 300 ◦ C, which the formation of gasoline fractions. The reaction time needs to be limited

5
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Fig. 2. (a) Oil composition of HDPE conversion in catalytic pyrolysis; (b) Fig. 3. (a) Effect of temperature on oil yield obtained by mixture of PP, PS and
Mechanism of mixed plastic wastes in catalytic pyrolysis. The Fig. 2(a) is taken PET at 25 MPa; (b) Mechanism of hydrothermal liquefaction. The Fig. 3(a) is
from Ref. (Pan et al., 2018) with copyright permission. taken from Ref. (Seshasayee and Savage, 2021) with copyright permission.

to 1 h approximately. occurred in unsaturated aliphatics, forming C8-C18 cycloalkanes and C6-


Currently, water is widely utilized as a solvent in hydrothermal C9 aromatic hydrocarbons. The process of hydrothermal liquefaction of
liquefaction studies. Comparatively, ethanol with low critical point plastic wastes can be summarized as follows: (1) depolymerization, (2)
achieves the mild liquefaction more easily. The effect of solvent type on monomers degradation via fracture, dehydration, decarboxylation and
the conversion of PS to fuels was studied. The yield of liquid oil reached deamidation, (3) fragments repolymerization.
84.7% in the presence of ethanol while that obtained with water was
74.2% (Ahmad et al., 2020). Moreover, taking ethanol as a solvent, 2.3. Advanced oxidation process
LDPE was converted to C5-C10 aromatic and >C13 non-aromatic fuels
with high heat value (29.6 MJ/kg) (Baloch et al., 2020b). Therefore, 2.3.1. Photocatalytic oxidation
ethanol is conducive to improving the output and quality of oil product Photocatalytic oxidation refers to oxidative decomposition of plastics
derived from hydrothermal liquefaction of plastics. In addition, into lower molecular substances via oxidants such as cavities and⋅OH
co-liquefaction of plastics with biomass has been reported. The yield of under photocatalysis. The photocatalyst dominates in the conversion
liquid oil achieved 9.1% in the co-liquefaction mode with mass ratio of process. LDPE could be converted to C2 fuel by polyacrylamide (PAM)/
Jingou lignite (JG), WS and PET (5:4:1). Compared with the hydro­ TiO2 and PAM/ZnO under UV light respectively (Kamalian et al., 2020;
thermal liquefaction of single PET, the yield of liquid oil via Liang et al., 2013). Evidently, oxidative species are formed via PAM/­
co-liquefaction was 4.8% higher under same conditions (Wang et al., TiO2 or PAM/ZnO photocatalysis, producing low molecular fuels. Con­
2014). The co-liquefaction of PE with bagasse at 280 ◦ C was conducted. version of polylactic acid (PLA), PET and polyurethane (PUR) to fuels by
The percentage of bio-oil was 31.5%, of which the calorific value CdS/CdOx photocatalysis under solar light was examined. The H2
reached 34.9 MJ/kg (Baloch et al., 2020a). In conclusion, product accounted for 38.8%, 16.6% and 22.5% respectively (Uekert
co-liquefaction of plastic wastes and biomass can produce high-value et al., 2018). Additionally, a two-step conversion route of C–C bond
fuels, but the yield of liquid oil remains to be enhanced. Plastic con­ cleavage and coupling was designed via Nb2O5 photocatalysis. PE, PP
version via novel heat treatment (250–450 ◦ C) can decrease carbon and PVC were degraded to CO2, which was further photo-reduced to
number distribution of products, forming gasoline and diesel fractions CH3COOH (<40.0%) (Jiao et al., 2020). The current photocatalysts
(C5-C16 aliphatic hydrocarbons) in the scale-up application. possess poor capability of CO2 reduction, leading to low yield of C2 fuels.
The mechanism of hydrothermal liquefaction is shown in Fig. 3b, Therefore, it is critical for rational design of photocatalyst and optimi­
which is similar to that of conventional pyrolysis. Taking PE and PP zation of C–C bond breaking and coupling to achieve efficient con­
mixture as an example, depolymerization is triggered by dissociation of version of plastic wastes into multi-carbon fuels at room temperature
free radicals at 280–450 ◦ C and 7–30 MPa. Noticeably, paraffin gener­ and pressure. Furthermore, Uekert et al. (2019) conducted PET and PLA
ated by saturated aliphatics could be transformed into C2-C4 combus­ conversion via carbon nitride/nickel phosphide (CNX=Ni2P) photo­
tible gases via further cracking. Besides, cyclization and aromatization catalysis under alkaline conditions. As a result, acetate and formate

6
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

⋅COOH +⋅COOH → HOOC− COOH (13)


generated, which could be applied as fuels after updating in Table 3. In
comparison with two-step conversion, direct generation of low carbon HOOC− COOH + 6e− + 6 H+ → CH3COOH + 2 H2O (14)
organics improves the sustainability and value of photocatalytic system.
The photocatalytic process for plastic conversion is controllable with where h+ and e- denote photo-induced holes and photo-generated
no extra oxidant, avoiding secondary contamination. In scale-up appli­ electrons respectively.
cations, plastics could induce depolymerization by conventional heat
treatment. Provided by thermal energy, the plastics could be trans­ 2.3.2. Fenton oxidation
formed into gas and liquid fuels. According to Eq. (9), the pyrolysis Fenton oxidation refers to polymer oxidation to small molecular
process requires 314.4–658.1 RMB per ton plastic for obtaining thermal substances via⋅OH generated by Fe2+ activated H2O2. Fenton oxidation
energy (Fivga and Dimitriou, 2018). However, the primary energy of sulfonated LDPE and HDPE was conducted at 65 ◦ C. LDPE and HDPE
source for plastic photocatalysis is solar light, which triggers plastic were converted to carboxylic acids with yields of 87.5% and 88.8%
decomposition and sustains photocatalytic conversion of plastic wastes respectively (Chow et al., 2016). Also, the sulfonated PVC was decom­
into fuels. Therefore, plastic conversion via photocatalysis can save posed to carboxylic acids (75.5%) via Fenton oxidation at 65 ◦ C (Chow
operating cost due to no thermal energy consumption compared to et al., 2017). Obviously, PE can be converted into carboxylic acids more
pyrolysis. efficiently than PVC. Generally, PE and PVC can be fragmented by
Fenton oxidation, while PS requires UV light to promote the trans­
S=(1000×q)/3600×K (9) formation process. PS could be converted into fuels after sulfonation by
Fenton or light-assisted Fenton process at 75 ◦ C. However, no reaction
Where S is pyrolysis cost per ton of dry plastics, q is the energy
was found by direct Fenton oxidation of PS within 5 h. Noticeably, the
requirement for dry plastic pyrolysis (1316.1 kJ/kg), K is the electricity
conversion of PS to CO2 by photo-Fenton oxidation exceeded 89.0% at
price for industrial consumption in China (0.86–1.80 RMB/kWh).
300 min under UV light (Feng et al., 2011). Therefore, the rapid
However, solar energy utilization for plastic conversion is insuffi­
oxidation of PS cannot be achieved with⋅OH in Fenton process. Several
cient due to the low proportion of UV light. More visible-light-driven
oxidants react together in light-assisted Fenton oxidation system.
photocatalyts needs to be developed for fuel conversion from plastics.
However, the exact oxidation mechanism is not yet clarified and re­
The mechanism of plastics conversion via photocatalysis is shown in
quires further investigation. Moreover, photo-Fenton oxidation of sul­
Fig. 4a. Semiconductor photocatalysts usually have a band structure and
fonated PE, PP and PVC was reported. The main product was CO2, which
narrow width of band gap. The photocatalytic conversion process of PE
could be converted to acetic acid after photo-reduction (Chow et al.,
is described in Eqs. (10–14). When the photon energy acquired by
2018).
semiconductor is larger than its band gap, the electrons can be stimu­
Plastic conversion via Fenton oxidation usually occurs at mild re­
lated from valence band to conduction band, leaving photo-induced
action conditions. The conversion process is efficient owing to rapid
holes on the valence band. Also, holes can oxidize water to ⋅OH,
decomposition of hydrogen peroxide, benefiting for large-scale appli­
which further oxidizes PE to CO2. Moreover, CO2 is converted to acetic
cations. Nevertheless, the process requires large amounts of hydrogen
acid via the reduction of electrons on conduction band. The equations of
peroxide, resulting in high operating costs.
photocatalytic conversion of PE are listed below:
Fig. 4b shows the Fenton oxidation process for plastic (PE as an
H2O + h+ →⋅OH + H+ (10) example) conversion to fuels. The addition of Fenton reagent to PE
generates a large amount of ⋅OH, which could break the C–C bond. PE is
(11) decomposed into adipic acid, glutaric acid and succinic acid by α-break,
∙OHOH
− (CH2 − CH2)n ̅→2nCO2
− ̅̅̅
β-break and γ-break, respectively. As described in Eqs. (15–18), Fe2+
CO2 + e− + H+ →⋅COOH (12) reacts with H2O2 to form Fe3+ and ⋅OH. Besides, H2O2 is converted to
⋅HO2 with Fe3+ catalysis, which can react with H2O2 to form more ⋅OH.

Table 3
Fuel conversion from plastics via advanced oxidation process.
Type of Technology System Conditions Products Yield Refs.
plastics

LDPE Photocatalysis polyacrylamide grafted UV light CO2 / (Liang et al.,


TiO2 2013)
PET Photocatalysis Carbon Nitride/Nickel Solar light, T = 25 ◦ C H2 77.2 μmol/g (Uekert et al.,
Phosphide 2019)
PET Photocatalysis CdS/CdOx Visible light H2 278.2 mmol/ (Uekert et al.,
g 2018)
LDPE Photocatalysis ZnO UV light CO2 / (Kamalian et al.,
2020)
PVC Fenton FeCl3, H2O2 T = 50 ◦ C, C/S Ratio = 8.0 C1-C2 acids 82.5% (Chow et al.,
2017)
LDPE Fenton FeCl3, H2O2 Chlorosulfuric acid = 2.5%, No light C1-C4 acids 62.4% (Chow et al.,
2016)
PE Fenton FeCl3, H2O2 UV-Vis light, pH = 2.5, Time = 2 h CO2 > 99% (Chow et al.,
2018)
PS Photo-Fenton FeCl3, H2O2 UV light, Time = 300 min CO2 89.0% (Feng et al.,
2011)
PVA Electrocatalysis H3PO4 T = 100 ◦ C, External voltage = 0.55 V H2 9.5 µmol/min (Hori et al.,
2020)
PVC Electrocatalysis TiO2/graphite cathode, T = 100 ◦ C, pH = 3.0, Oxygen flow = 40 mL/min, Carboxylic 75.0% (Miao et al.,
Time = 6 h, Applied potentia = − 0.7 V acid 2020)
PP Solar Thermo-coupled NaOH, KOH T = 350 ◦ C, Time = 60 min, External voltage Gas 33.1% (Jiang et al.,
electrocatalysis = 1.5 V 2020)

LDPE: Low density polyethylene; PET: Polyethylene terephthalate; PVC: Polyvinyl chloride; PE: Polyethylene; PS: Polystyrene; PVA: Polyvinyl alcohol. PP: Poly­
propylene; T: Temperature

7
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Fig. 4. Mechanism of plastic conversion to fuels (a) Photocatalytic oxidation; (b) Fenton oxidation; (c) Electro-catalytic oxidation; (d) advanced oxidation. The Fig. 4
(a) is taken from Ref. (Jiao et al., 2020) with copyright permission. The Fig. 4(b) is taken from Ref. (Chow et al., 2016) with copyright permission. The Fig. 4(c) is
taken from Ref. (Miao et al., 2020) with copyright permission.

8
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Besides, the interaction between ⋅OH and RH generates R⋅(Eq. 19). Overall, plastic conversion via Fenton, photocatalysis and electro­
Eventually, the polymer is transformed into CO2 and H2O with breakage catalysis can generate⋅OH, converting plastic wastes into long-chain free
of C–C bond (Eq. 20). The equations about Fenton oxidation of plastics radicals (Fig. 4d). Furthermore, C2-C4 carboxylic acid (e.g., HCOOH,
to fuels are listed below: CH3COOH and C₂H₂O₄) and H2 are formed with further oxidation.

Fe2+ + H2O2 →Fe3+ +⋅OH + OH− (15)


3. Conclusion and perspective
Fe 2+ 3+
+⋅OH →Fe + OH −
(16)
In recent decades, fuel conversion from plastic wastes has received
Fe3+ + H2O2 →Fe2+ +⋅HO2 + H+ (17) extensive attentions in order to alleviate environmental pollution and
energy crisis. This review summarizes the available technologies for
⋅HO2 + H2O2 → O2 + H2O +⋅OH (18)
plastic wastes conversion into fuels comprehensively. Some conclusions
RH +⋅OH → R⋅+ Fe2+ (19) are put forward below:
+ +
R + O2 → ROO →⋅⋅⋅→ CO2 + H2O (20) (1) The fuels, obtained from plastic wastes conversion via traditional
thermal treatment, possess high calorific value and superior
performance approaching gasoline and diesel. However, the
2.3.3. Electrocatalytic oxidation plastic pyrolysis process needs high temperature and long reac­
Electrocatalytic conversion of plastics contains two routes of direct tion time, producing liquid oil with low yield, octane value and
and indirect oxidation. Direct oxidation route refers to the electrophilic wide range of carbon number distribution. Besides, the pipeline is
attack of polymer by ⋅OH generated via water discharge on the surface of clogged easily due to the formation of olefins and paraffins during
anode. Indirect oxidation means strong oxidizing intermediates domi­ reactions. What’s worse, volatile compounds generate in the
nate in the plastic conversion process. Polyvinyl alcohol (PVA) was pyrolysis process, resulting in environmental pollution.
converted to H2 (9.5 µmol/min) successfully with an external voltage (2) Efficient conversion of plastic wastes can be realized at relatively
(0.55 V) applied in H3PO4 solution at 200 ◦ C (Hori et al., 2020). Also, low temperature via novel heat treatment. The obtained fuel is
electrocatalytic degradation of PVC was conducted for carboxylic acid highly selective, among which isomerization and aromatization
(75%) formation on TiO2/C cathode (− 0.7 V) at 100 ◦ C (Miao et al., products account for a large proportion. Nevertheless, catalysts
2020). Obviously, electrocatalytic degradation of plastic wastes could with high demand in plastics conversion are difficult to recover.
generate single product, which can be converted into fuels without The catalyst deactivation occurs owing to impurity deposition in
post-treatment. In contrast to single electrolysis, the combination of cokes and plastic wastes. Besides, the reaction equipment re­
pyrolysis and electrolysis can produce abundant fuel components. The quires strict resistance to high pressure and corrosion.
PP conversion to H2 and C1-C5 fuels was realized through solar (3) Advanced oxidation can achieve target fuel conversion from
thermo-coupled electrolysis. Furthermore, the PP depolymerization was plastic wastes under room temperature and pressure, reducing
enhanced with rising temperature (350–390 ◦ C) and electrolysis current energy consumption considerably compared to heat treatment.
(0–400 mA) (Jiang et al., 2020). Therefore, by considering fuel yield Furthermore, the advanced oxidation process is controllable with
with temperature and electrolysis current, it is prospective to design an efficient plastic conversion and high product purity. However,
electrochemical system with high-efficiency under low temperature and the reaction process requires highly-efficient catalysts, and the
current. In this way, the industrial utilization of plastic wastes can be liquid products have relatively low calorific value. Interestingly,
achieved with electrocatalysis. active sites on catalyst surface can be adjusted by appropriate
The electrocatalytic oxidation is controllable for plastic conversion design, realizing oxidative species regulation and high-quality-
at room temperature and pressure. The voltage and electrode could fuels production. Overall, fuel conversion from plastic wastes
affact plastic oxidation directly. High purity fuels will be formed at via advanced oxidation is promising for future applications.
appropriate conditions (e.g., voltage). However, the electrode life is
short and the electrocatalytic power consumption is high. Although certain progress has been obtained in the fuel conversion
Electrocatalytic conversion route of plastics (e.g., PVC) is described from plastic wastes, the following challenges still remain to be overcome
in Fig. 4c. After receiving electrons from the cathode (TiO2/C), PVC is in future research:
reduced, triggering dechlorination at high temperature. Also, PVC oxi­
dization occurs with⋅OH to form carbonyl and hydroxyl groups, which (1) The addition of biomass in conventional thermal treatment of
are further converted into small molecules (e.g., alcohols, carboxylic plastic wastes decreases microwave energy, heat energy and
acids and esters). At last, these substances are mineralized to CO2 and thermal conductivity losses, improving fuel yield by synergistic
H2O partially. The mechanism of electrocatalytic oxidation of plastic effect. However, few studies have been focused on the influential
wastes is proposed. Initially, H2O in the acidic solution is adsorbed on parameters (e.g., temperature, feed ratio of plastics to biomass
the surface of metal oxide anode. As shown in Eq. (21), H2O loses and reaction time) and related mechanism of plastics/biomass co-
electrons under the electric field, forming MOx(⋅OH) (MOx represents pyrolysis. The optimization of reaction parameters based on
anode). mechanism deserves further investigations to achieve higher fuel
yield. The maximization of economic benefit can be realized by
MOx + H2O → MOx(⋅OH) + H+ + e− (21) considering process parameters and product quality
simultaneously.
Furthermore, metal peroxide (MOx+1) generated due to the invasion
(2) Catalyst plays a crucial role in the novel heat treatment of plastic
of oxygen atoms to MOx lattice (Eq. (22)).
wastes. Many kinds of plastic wastes exist, but a majority of
MOx(⋅OH) → MOx+1 + H+ + e− (22) catalysts are efficient for single polymer decomposition merely.
The catalyst presents the drawbacks of easy deactivation and high
Selective oxidation of polymer was conducted via MOx(⋅OH)z and cost. Therefore, it is essential to develop catalysts with excellent
MOx+1 in solution (Eqs. (23) and (24)). adaptability, high activity and recyclability. In order to avoid
clogging, the reactor design should be optimized to separate the
R + MOx(⋅OH)z → CO2 + MOx + ZH+ + Ze− (23)
concentrated brine in time and deter salt deposition. Moreover,
R + MOx+1 → RO + MOx (24)

9
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

the equipment performance can be enhanced via the develop­ the upper Hudson River to predict dietary and hepatic concentrations and health
effects. Environ. Toxicol. Chem. 37 (2), 563–575. https://doi.org/10.1002/etc.3997.
ment of novel pressure and corrosion resistant materials.
Chen, H., Wan, K., Zhang, Y., Wang, Y., 2021. Waste to wealth: chemical recycling and
(3) The liquid products, obtained by advanced oxidation of plastics chemical upcycling of waste plastics for a great future. ChemSusChem 14, 1–5.
have low calorific value. The species and density of oxidants in https://doi.org/10.1002/cssc.202100652.
the system can be regulated via appropriate modulation of active Chen, W.-T., Jin, K., Linda Wang, N.-H., 2019. Use of supercritical water for the
liquefaction of polypropylene into oil. ACS Sustain. Chem. Eng. 7 (4), 3749–3758.
sites on catalyst surface. Accordingly, selective rupture of https://doi.org/10.1021/acssuschemeng.8b03841.
chemical bonds in polymers and directional conversion of target Cheng, L., Gu, J., Wang, Y., Zhang, J., Yuan, H., Chen, Y., 2020. Polyethylene high-
fuels can be realized. Furthermore, the photocatalysts should be pressure pyrolysis: better product distribution and process mechanism analysis.
Chem. Eng. J. 385, 123866 https://doi.org/10.1016/j.cej.2019.123866.
designed rationally via energy band modulation, heterojunctions Chow, C.F., Wong, W.L., Ho, K.Y., Chan, C.S., Gong, C.B., 2016. Combined chemical
construction and co-catalysts loading. The separation of photo­ activation and Fenton degradation to convert waste polyethylene into high-value
generated carriers is desired to be accelerated, prolonging the fine chemicals. Chem. Eur. J. 22 (28), 9513–9518. https://doi.org/10.1002/
chem.201600856.
service time of photocatalysts. Besides, optimization of C–C Chow, C.F., Wong, W.L., Chan, C.W., Chan, C.S., 2018. Converting inert plastic waste
bond breakage and coupling process of intermediates may realize into energetic materials: a study on the light-accelerated decomposition of plastic
efficient transformation of multi-carbon fuel from plastic wastes waste with the Fenton reaction. Waste Manag. 75, 174–180. https://doi.org/
10.1016/j.wasman.2018.01.034.
under natural illumination. Chow, C.-F., Wong, W.-L., Chan, C.-S., Li, Y., Tang, Q., Gong, C.-B., 2017. Breakdown of
plastic waste into economically valuable carbon resources: Rapid and effective
chemical treatment of polyvinylchloride with the Fenton catalyst. Polym. Degrad.
Declaration of Competing Interest Stab. 146, 34–41. https://doi.org/10.1016/j.polymdegradstab.2017.09.016.
Čolnik, M., Kotnik, P., Knez, Ž., Škerget, M., 2021. Hydrothermal decomposition of
polyethylene waste to hydrocarbons rich oil. J. Supercrit. Fluids 169, 105136.
The authors declare that they have no known competing financial
https://doi.org/10.1016/j.supflu.2020.105136.
interests or personal relationships that could have appeared to influence Dogu, O., Pelucchi, M., Van de Vijver, R., Van Steenberge, P.H.M., D’Hooge, D.R.,
the work reported in this paper. Cuoci, A., Mehl, M., Frassoldati, A., Faravelli, T., Van Geem, K.M., 2021. The
chemistry of chemical recycling of solid plastic waste via pyrolysis and gasification:
state-of-the-art, challenges, and future directions. Prog. Energy Combust. Sci. 84,
Acknowledgements 100901 https://doi.org/10.1016/j.pecs.2020.100901.
Dwivedi, U., Naik, S.N., Pant, K.K., 2021. High quality liquid fuel production from waste
plastics via two-step cracking route in a bottom-up approach using bi-functional Fe/
This research was financially supported by the National Natural HZSM-5 catalyst. Waste Manag. 132, 151–161. https://doi.org/10.1016/j.
Science Foundation of China (52100156 and 521176197) and Shenzhen wasman.2021.07.024.
Science and Technology Program (GJHZ20200731095801005 and Fadillah, G., Fatimah, I., Sahroni, I., Musawwa, M.M., Mahlia, T.M.I., Muraza, O., 2021.
Recent progress in low-cost catalysts for pyrolysis of plastic waste to fuels. catalysts
JCYJ20200109150210400).
11 (7), 837. https://doi.org/10.3390/catal11070837.
FakhrHoseini, S.M., Dastanian, M., 2013. Predicting pyrolysis products of PE, PP, and
References PET using NRTL activity coefficient model. J. Chem. 2013, 1–5. https://doi.org/
10.1155/2013/487676.
Feng, H.M., Zheng, J.C., Lei, N.Y., Lei, Y.U., Kong, H.K., Han-Qing, Y.U., Lau, T.C.,
Abbas-Abadi, M.S., Haghighi, M.N., Yeganeh, H., 2013. Evaluation of pyrolysis product
Lam, M.H.W.J.E.S., 2011. Photoassisted Fenton degradation of polystyrene. Environ.
of virgin high density polyethylene degradation using different process parameters
Sci. Technol. 45 (2), 744. https://doi.org/10.1021/es102182g10.1021/es102182g.
in a stirred reactor. Fuel Process. Technol. 109, 90–95. https://doi.org/10.1016/j.
Fivga, A., Dimitriou, I., 2018. Pyrolysis of plastic waste for production of heavy fuel
fuproc.2012.09.042.
substitute: a techno-economic assessment. Energy 149, 865–874. https://doi.org/
Ahmad, I., Ismail Khan, M., Ishaq, M., Khan, H., Gul, K., Ahmad, W., 2013. Catalytic
10.1016/j.energy.2018.02.094.
efficiency of some novel nanostructured heterogeneous solid catalysts in pyrolysis of
Fushimi, R., Gaffney, A.M., 2014. The new chempren process for the conversion of waste
HDPE. Polym. Degrad. Stab. 98 (12), 2512–2519. https://doi.org/10.1016/j.
plastic to chemicals and fuel. Top. Catal. 57 (17–20), 1412–1418. https://doi.org/
polymdegradstab.2013.09.009.
10.1007/s11244-014-0312-6.
Ahmad, I., Khan, M.I., Khan, H., Ishaq, M., Tariq, R., Gul, K., Ahmad, W., 2014. Pyrolysis
Gin, A.W., Hassan, H., Ahmad, M.A., Hameed, B.H., Din, A.T.M., 2021. Recent progress
study of polypropylene and polyethylene into premium oil products. Int. J. Green
on catalytic co-pyrolysis of plastic waste and lignocellulosic biomass to liquid fuel:
Energy 12 (7), 663–671. https://doi.org/10.1080/15435075.2014.880146.
the influence of technical and reaction kinetic parameters. Arab. J. Chem. 14 (4),
Ahmad, N., Ahmad, N., Maafa, I.M., Ahmed, U., Akhter, P., Shehzad, N., Amjad, U.-e-s,
103035 https://doi.org/10.1016/j.arabjc.2021.103035.
Hussain, M., 2020. Thermal conversion of polystyrene plastic waste to liquid fuel via
Hibino, T., Kobayashi, K., Teranishi, S., Hitomi, T., 2021. Solid oxide fuel cell using
ethanolysis. Fuel 279, 118498. https://doi.org/10.1016/j.fuel.2020.118498.
municipal solid waste directly as fuel: biomass, resin, plastic, and food waste. ACS
Al-Salem, S.M., Antelava, A., Constantinou, A., Manos, G., Dutta, A., 2017. A review on
Sustain. Chem. Eng. 9 (8), 3124–3136. https://doi.org/10.1021/
thermal and catalytic pyrolysis of plastic solid waste (PSW). J. Environ. Manag. 197,
acssuschemeng.0c07657.
177–198. https://doi.org/10.1016/j.jenvman.2017.03.084.
Hori, T., Kobayashi, K., Teranishi, S., Nagao, M., Hibino, T., 2020. Fuel cell and
Anuar Sharuddin, S.D., Abnisa, F., Wan Daud, W.M.A., Aroua, M.K., 2016. A review on
electrolyzer using plastic waste directly as fuel. Waste Manag. 102, 30–39. https://
pyrolysis of plastic wastes. Energ. Convers. Manag. 115, 308–326. https://doi.org/
doi.org/10.1016/j.wasman.2019.10.019.
10.1016/j.enconman.2016.02.037.
Jiang, T., Zhao, X., Gu, D., Yan, C., Jiang, H., Wu, H., Wang, B., Wang, X., 2020. STEP
Bai, B., Jin, H., Fan, C., Cao, C., Wei, W., Cao, W., 2019. Experimental investigation on
polymer degradation: Solar thermo-coupled electrochemical depolymerization of
liquefaction of plastic waste to oil in supercritical water. Waste Manag. 89, 247–253.
plastics to generate useful fuel plus abundant hydrogen. Sol. Energy Mater. Sol. Cells
https://doi.org/10.1016/j.wasman.2019.04.017.
204, 110208. https://doi.org/10.1016/j.solmat.2019.110208.
Bajad, G., Jain, R., Harhare, W., Vijayakumar, R.P., Bose, S.J.A.M.P., 2017. Synthesis of
Jiao, X., Zheng, K., Chen, Q., Li, X., Li, Y., Shao, W., Xu, J., Zhu, J., Pan, Y., Sun, Y.,
fuel oil and carbon nanotubes in an autoclave using plastic waste as precursor.
Xie, Y., 2020. Photocatalytic conversion of waste plastics into C2 fuels under
Mater. Manuf. Process. 32 (5), 495–500. https://doi.org/10.1080/
simulated natural environment conditions. Angew. Chem. 59 (36), 15497–15501.
10426914.2016.1244830.
Jin, K., Vozka, P., Kilaz, G., Chen, W.-T., Wang, N.-H.L., 2020. Conversion of
Baloch, H.A., Nizamuddin, S., Siddiqui, M.T.H., Mubarak, N.M., Mazari, S., Griffin, G.J.,
polyethylene waste into clean fuels and waxes via hydrothermal processing (HTP).
Srinivasan, M.P., 2020a. Co-liquefaction of synthetic polyethylene and polyethylene
Fuel 273, 117726. https://doi.org/10.1016/j.fuel.2020.117726.
bags with sugarcane bagasse under supercritical conditions: a comparative study.
Jin, Z., Chen, D., Yin, L., Hu, Y., Zhu, H., Hong, L., 2018. Molten waste plastic pyrolysis
Renew. Energ. 162, 2397–2407. https://doi.org/10.1016/j.renene.2020.10.008.
in a vertical falling film reactor and the influence of temperature on the pyrolysis
Baloch, H.A., Siddiqui, M.T.H., Nizamuddin, S., Mubarak, N.M., Khalid, M.,
products. Chin. J. Chem. Eng. 26 (2), 400–406. https://doi.org/10.1016/j.
Srinivasan, M.P., Griffin, G.J., 2020b. Solvothermal co-liquefaction of sugarcane
cjche.2017.08.001.
bagasse and polyethylene under sub-supercritical conditions: optimization of process
Kamalian, P., Khorasani, S.N., Abdolmaleki, A., Karevan, M., Khalili, S., Shirani, M.,
parameters. Process Saf. Environ. Prot. 137, 300–311. https://doi.org/10.1016/j.
Neisiany, R.E., 2020. Toward the development of polyethylene photocatalytic
psep.2020.01.018.
degradation. J. Polym. Eng. 40 (2), 181–191. https://doi.org/10.1515/polyeng-
Banu, J.R., Sharmila, V.G., Ushani, U., Amudha, V., Kumar, G., 2020. Impervious and
2019-0230.
influence in the liquid fuel production from municipal plastic waste through thermo-
Kunwar, B., Cheng, H.N., Chandrashekaran, S.R., Sharma, B.K., 2016. Plastics to fuel: a
chemical biomass conversion technologies - a review. Sci. Total Environ. 718,
review. Renew. Sust. Energ. Rev. 54, 421–428. https://doi.org/10.1016/j.
137287 https://doi.org/10.1016/j.scitotenv.2020.137287.
rser.2015.10.015.
Budsaereechai, S., Hunt, A.J., Ngernyen, Y., 2019. Catalytic pyrolysis of plastic waste for
Lamb, J.B., Willis, B.L., Fiorenza, E.A., Couch, C.S., Howard, R., Rader, D.N., True, J.D.,
the production of liquid fuels for engines. RSC Adv. 9 (10), 5844–5857. https://doi.
Kelly, L.A., Ahmad, A., Jompa, J., Harvell, C.D., 2018. Plastic waste associated with
org/10.1039/c8ra10058f.
disease on coral reefs. Science 359 (6374), 460–462. https://doi.org/10.1126/
Bursian, S.J., Kern, J., Remington, R.E., Link, J.E., 2018. Use of polychlorinated biphenyl
science.aar3320.
and toxic equivalent concentrations in scat from mink (Neovison vison) fed fish from

10
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Lee, K.H., Noh, N.S., Shin, D.H., Seo, Y.J.P.D., 2002. Comparison of plastic types for Panda, A.K., Singh, R.K., Mishra, D.K., 2010. Thermolysis of waste plastics to liquid fuelA
catalytic degradation of waste plastics into liquid product with spent FCC catalyst. suitable method for plastic waste management and manufacture of value added
Polym. Degrad. Stab. 78 (3), 539–544. https://doi.org/10.1016/S0141-3910(02) products—a world prospective. Renew. Sust. Energ. Rev. 14 (1), 233–248. https://
00227-6. doi.org/10.1016/j.rser.2009.07.005.
Liang, W., Luo, Y., Song, S., Dong, X., Yu, X., 2013. High photocatalytic degradation Papari, S., Bamdad, H., Berruti, F., 2021. Pyrolytic conversion of plastic waste to value-
activity of polyethylene containing polyacrylamide grafted TiO2. Polym. Degrad. added products and fuels: a review. Materials 14 (10), 2586. https://doi.org/
Stab. 98 (9), 1754–1761. https://doi.org/10.1016/j.polymdegradstab.2013.05.027. 10.3390/ma14102586.
Lin, Y., Yang, M., 2007. Catalytic conversion of commingled polymer waste into Pathan, S.I., Arfaioli, P., Bardelli, T., Ceccherini, M.T., Nannipieri, P., Pietramellara, G.,
chemicals and fuels over spent FCC commercial catalyst in a fluidised-bed reactor. 2020. Soil pollution from micro- and nanoplastic debris: a hidden and unknown
Appl. Catal. B 69 (3–4), 145–153. https://doi.org/10.1016/j.apcatb.2006.07.005. biohazard. Sustainability 12 (18), 1–31. https://doi.org/10.3390/su12187255.
Lin, Y.H., Yang, M.H., 2005. Catalytic reactions of post-consumer polymer waste over Pichler, C.M., Bhattacharjee, S., Rahaman, M., Uekert, T., Reisner, E., 2021. Conversion
fluidised cracking catalysts for producing hydrocarbons. Mol. Catal. A Chem. 231 of polyethylene waste into gaseous hydrocarbons via integrated tandem chemical-
(1–2), 113–122. https://doi.org/10.1016/j.molcata.2005.01.003. photo/electrocatalytic processes. ACS Catal. 11 (15), 9159–9167. https://doi.org/
Lin, Y.H., Yang, M.H., Wei, T.T., Hsu, C.T., Wu, K.J., Lee, S.L., 2010. Acid-catalyzed 10.1021/acscatal.1c02133.
conversion of chlorinated plastic waste into valuable hydrocarbons over post-use Prurapark, R., Owjaraen, K., Saengphrom, B., Limthongtip, I., Tongam, N., 2020. Effect
commercial FCC catalysts. J. Anal. Appl. Pyrolysis 87 (1), 154–162. https://doi.org/ of temperature on pyrolysis oil using high-density polyethylene and polyethylene
10.1016/j.jaap.2009.11.006. terephthalate sources from mobile pyrolysis plant. Front. Energy Res. 8, 541535
Lin, Y.-H., 2009. Production of valuable hydrocarbons by catalytic degradation of a https://doi.org/10.3389/fenrg.2020.541535.
mixture of post-consumer plastic waste in a fluidized-bed reactor. Polym. Degrad. Rasool, F.N., Saavedra, M.A., Pamba, S., Perold, V., Mmochi, A.J., Maalim, M.,
Stab. 94 (11), 1924–1931. https://doi.org/10.1016/j. Simonsen, L., Buur, L., Pedersen, R.H., Syberg, K., Jelsbak, L., 2021. Isolation and
polymdegradstab.2009.08.004. characterization of human pathogenic multidrug resistant bacteria associated with
Liu, L., Zhao, H., Andino, J.M., Li, Y., 2012. Photocatalytic CO2 reduction with H2O on plastic litter collected in Zanzibar. J. Hazard. Mater. 405, 124591 https://doi.org/
TiO2 nanocrystals: comparison of anatase, rutile, and brookite polymorphs and 10.1016/j.jhazmat.2020.124591.
exploration of surface chemistry. ACS Catal. 2 (8), 1817–1828. https://doi.org/ Sall, M.L., Diaw, A.K.D., Gningue-Sall, D., Aaron, S.E., Aaron, J.-J., 2020. Toxic heavy
10.1021/cs300273q. metals: impact on the environment and human health, and treatment with
Liu, S., Kots, P.A., Vance, B.C., Danielson, A., Vlachos, D.G., 2021. Plastic waste to fuels conducting organic polymers, a review. Environ. Sci. Pollut. Res. 27 (24),
by hydrocracking at mild conditions. Sci. Adv. 7 (17), eabf8283 https://doi.org/ 29927–29942. https://doi.org/10.1007/s11356-020-09354-3.
10.1126/sciadv.abf8283. Sarker, M., Rashid, M.M., Molla, M., 2012a. Waste polypropylene plastic conversion into
Lopez-Urionabarrenechea, A., de Marco, I., Caballero, B.M., Laresgoiti, M.F., liquid hydrocarbon fuel for producing electricity and energies. Environ. Technol. 33
Adrados, A., 2015. Upgrading of chlorinated oils coming from pyrolysis of plastic (24), 2709–2721. https://doi.org/10.1080/09593330.2012.676075.
waste. Fuel Process. Technol. 137, 229–239. https://doi.org/10.1016/j. Sarker, M., Rashid, M.M., Rahman, M.S., Molla, M.J.C.O., 2012b. Low and high density
fuproc.2015.04.015. polyethylene waste plastics conversion into liquid hydrocarbon fuel. Chim. Oggi 30
Luo, G., Suto, T., Yasu, S., Kato, K., 2000. Catalytic degradation of high density (6), 42–43. https://www.researchgate.net/publication/286672978.
polyethylene and polypropylene into liquid fuel in a powder-particle fluidized bed. Seshasayee, M.S., Savage, P.E., 2020. Oil from plastic via hydrothermal liquefaction:
Polym. Degrad. Stab. 70, 97–102. https://doi.org/10.1016/S0141-3910(00)00095- production and characterization. Appl. Energy 278, 115673. https://doi.org/
1. 10.1016/j.apenergy.2020.115673.
Machado, A.Ad.S., Lau, C.W., Till, J., Kloas, W., Lehmann, A., Becker, R., Rillig, M.C., Seshasayee, M.S., Savage, P.E., 2021. Synergistic interactions during hydrothermal
2018. Impacts of microplastics on the soil biophysical environment. Environ. Sci. liquefaction of plastics and biomolecules. Chem. Eng. J. 417, 129268 https://doi.
Technol. 52 (17), 9656–9665. https://doi.org/10.1021/acs.est.8b02212. org/10.1016/j.cej.2021.129268.
Mallik, A., Xavier, K.A.M., Naidu, B.C., Nayak, B.B., 2021. Ecotoxicological and Shah, S.H., Khan, Z.M., Raja, I.A., Mahmood, Q., Bhatti, Z.A., Khan, J., Farooq, A.,
physiological risks of microplastics on fish and their possible mitigation measures. Rashid, N., Wu, D., 2010. Low temperature conversion of plastic waste into light
Sci. Total Environ. 779, 146433 https://doi.org/10.1016/j.scitotenv.2021.146433. hydrocarbons. J. Hazard. Mater. 179 (1–3), 15–20. https://doi.org/10.1016/j.
Mangesh, V.L., Perumal, T., Subramanian, S., Padmanabhan, S., 2020. Clean energy from jhazmat.2010.01.134.
plastic: production of hydroprocessed waste polypropylene pyrolysis oil utilizing Shen, Y., 2020. A review on hydrothermal carbonization of biomass and plastic wastes to
Ni–Mo/Laponite catalyst. Energy Fuels 34 (7), 8824–8836. https://doi.org/ energy products. Biomass Bioenergy 134, 105479. https://doi.org/10.1016/j.
10.1021/acs.energyfuels.0c01051. biombioe.2020.105479.
Marcilla, A., Beltrán, M.I., Navarro, R., 2009. Thermal and catalytic pyrolysis of Siddiqui, M.N., 2009. Conversion of hazardous plastic wastes into useful chemical
polyethylene over HZSM5 and HUSY zeolites in a batch reactor under dynamic products. J. Hazard. Mater. 167 (1–3), 728–735. https://doi.org/10.1016/j.
conditions. Appl. Catal. B 86 (1–2), 78–86. https://doi.org/10.1016/j. jhazmat.2009.01.042.
apcatb.2008.07.026. Singh, M.V., Kumar, S., Sarker, M., 2018. Waste HD-PE plastic, deformation into liquid
Mbachu, O., Jenkins, G., Kaparaju, P., Pratt, C., 2021. The rise of artificial soil carbon hydrocarbon fuel using pyrolysis-catalytic cracking with a CuCO3 catalyst. Sustain.
inputs: reviewing microplastic pollution effects in the soil environment. Sci. Total Energy Fuels 2 (5), 1057–1068. https://doi.org/10.1039/c8se00040a.
Environ. 780, 146569 https://doi.org/10.1016/j.scitotenv.2021.146569. Singh, R.K., Ruj, B., Sadhukhan, A.K., Gupta, P., 2019a. Impact of fast and slow pyrolysis
Miandad, R., Barakat, M.A., Rehan, M., Aburiazaiza, A.S., Ismail, I.M.I., Nizami, A.S., on the degradation of mixed plastic waste: product yield analysis and their
2017. Plastic waste to liquid oil through catalytic pyrolysis using natural and characterization. J. Energy Inst. 92 (6), 1647–1657. https://doi.org/10.1016/j.
synthetic zeolite catalysts. Waste Manag. 69, 66–78. https://doi.org/10.1016/j. joei.2019.01.009.
wasman.2017.08.032. Singh, R.K., Ruj, B., Sadhukhan, A.K., Gupta, P., 2019b. Thermal degradation of waste
Miao, F., Liu, Y., Gao, M., Yu, X., Xiao, P., Wang, M., Wang, S., Wang, X., 2020. plastics under non-sweeping atmosphere: part 1: effect of temperature, product
Degradation of polyvinyl chloride microplastics via an electro-Fenton-like system optimization, and degradation mechanism. J. Environ. Manag. 239, 395–406.
with a TiO2/graphite cathode. J. Hazard. Mater. 399, 123023 https://doi.org/ https://doi.org/10.1016/j.jenvman.2019.03.067.
10.1016/j.jhazmat.2020.123023. Singh, R.K., Ruj, B., Sadhukhan, A.K., Gupta, P., 2020. Conventional pyrolysis of plastic
Miranda, R., Yang, J., Roy, C.J.P.D., 1999. Vacuum pyrolysis of PVC I. Kinetic study. waste for product recovery and utilization of pyrolytic gases for carbon nanotubes
Polym. Degrad. Stab. 64 (1), 127–144. https://doi.org/10.1016/S0141-3910(98) production. Environ. Sci. Pollut. Res. Int. https://doi.org/10.1007/s11356-020-
00186-4. 11204-1.
Miskolczi, N., Bartha, L., Deák, G., Jóver, B., Kalló, D., 2004. Thermal and thermo- Song, Z., Xiu, F.R., Qi, Y., 2021. Degradation and partial oxidation of waste plastic
catalytic degradation of high-density polyethylene waste. J. Anal. Appl. Pyrolysis 72 express packaging bags in supercritical water: resources transformation and
(2), 235–242. https://doi.org/10.1016/j.jaap.2004.07.002. pollutants removal. J. Hazard. Mater. 423, 127018 https://doi.org/10.1016/j.
Mohanraj, C., Senthilkumar, T., Chandrasekar, M., 2017. A review on conversion jhazmat.2021.127018.
techniques of liquid fuel from waste plastic materials. Int. J. Energy Res. 41 (11), Subramanian, S., Schnoor, J.L., Van Aken, B., 2017. Effects of polychlorinated biphenyls
1534–1552. https://doi.org/10.1002/er.3720. (PCBs) and their hydroxylated metabolites (OH-PCBs) on Arabidopsis thaliana.
Nanda, S., Berruti, F., 2020. Thermochemical conversion of plastic waste to fuels: a Environ. Sci. Technol. 51 (12), 7263–7270. https://doi.org/10.1021/acs.
review. Environ. Chem. Lett. 19 (1), 123–148. https://doi.org/10.1007/s10311-020- est.7b01538.
01094-7. Suriapparao, D.V., Yerrayya, A., Nagababu, G., Guduru, R.K., Kumar, T.H., 2020.
Nor, N.H.M., Kooi, M., Diepens, N.J., Koelmans, A.A., 2021. Lifetime accumulation of Recovery of renewable aromatic and aliphatic hydrocarbon resources from
microplastic in children and adults. Environ. Sci. Technol. 55 (8), 5084–5096. microwave pyrolysis/co-pyrolysis of agro-residues and plastics wastes. Bioresour.
https://doi.org/10.1021/acs.est.0c07384. Technol. 318, 124277 https://doi.org/10.1016/j.biortech.2020.124277.
Okolie, J.A., Nanda, S., Dalai, A.K., Berruti, F., Kozinski, J.A., 2020. A review on Uekert, T., Kuehnel, M.F., Wakerley, D.W., Reisner, E., 2018. Plastic waste as a feedstock
subcritical and supercritical water gasification of biogenic, polymeric and petroleum for solar-driven H2 generation. Energy Environ. Sci. 11 (10), 2853–2857. https://
wastes to hydrogen-rich synthesis gas. Renew. Sust. Energ. Rev. 119, 109546 doi.org/10.1039/c8ee01408f.
https://doi.org/10.1016/j.rser.2019.109546. Uekert, T., Kasap, H., Reisner, E., 2019. Photoreforming of nonrecyclable plastic waste
Owusu, P.A., Banadda, N., Zziwa, A., Seay, J., Kiggundu, N., 2018. Reverse engineering over a carbon nitride/nickel phosphide catalyst. J. Am. Chem. Soc. 141 (38),
of plastic waste into useful fuel products. J. Anal. Appl. Pyrolysis 130, 285–293. 15201–15210. https://doi.org/10.1021/jacs.9b06872.
https://doi.org/10.1016/j.jaap.2017.12.020. Ukarde, T.M., Pawar, H.S., 2021. A Cu doped TiO2 catalyst mediated Catalytic Thermo
Pan, Z., Xue, X., Zhang, C., Wang, D., Xie, Y., Zhang, R., 2018. Evaluation of process Liquefaction (CTL) of polyolefinic plastic waste into hydrocarbon oil. Fuel 285,
parameters on high-density polyethylene hydro-liquefaction products. J. Anal. Appl. 119155. https://doi.org/10.1016/j.fuel.2020.119155.
Pyrolysis 136, 146–152. https://doi.org/10.1016/j.jaap.2018.10.011.

11
N. Li et al. Journal of Hazardous Materials 424 (2022) 127460

Utami, M., Wijaya, K., Trisunaryanti, W., 2018. Pt-promoted sulfated zirconia as catalyst risks, interactions and fate - a review. Crit. Rev. Environ. Sci. Technol. 50 (21),
for hydrocracking of LDPE plastic waste into liquid fuels. Mater. Chem. Phys. 213, 2175–2222. https://doi.org/10.1080/10643389.2019.1694822.
548–555. https://doi.org/10.1016/j.matchemphys.2018.03.055. Xue, Y., Zhou, S., Brown, R.C., Kelkar, A., Bai, X., 2015. Fast pyrolysis of biomass and
Uzoejinwa, B.B., He, X., Wang, S., El-Fatah Abomohra, A., Hu, Y., Wang, Q., 2018. Co- waste plastic in a fluidized bed reactor. Fuel 156, 40–46. https://doi.org/10.1016/j.
pyrolysis of biomass and waste plastics as a thermochemical conversion technology fuel.2015.04.033.
for high-grade biofuel production: recent progress and future directions elsewhere Yao, M., Hu, T., Wang, Y., Du, Y., Hu, C., Wu, R., 2017. Polychlorinated biphenyls and its
worldwide. Energ. Convers. Manag. 163, 468–492. https://doi.org/10.1016/j. potential role in endometriosis. Environ. Pollut. 229, 837–845. https://doi.org/
enconman.2018.02.004. 10.1016/j.envpol.2017.06.088.
Varo, I., Osorio, K., Estensoro, I., Naya-Catala, F., Sitja-Bobadilla, A., Navarro, J.C., Yee, M.S.-L., Hii, L.-W., Looi, C.K., Lim, W.-M., Wong, S.-F., Kok, Y.-Y., Tan, B.-K.,
Perez-Sanchez, J., Torreblanca, A., Piazzon, M.C., 2021. Effect of virgin low density Wong, C.-Y., Leong, C.-O., 2021. Impact of microplastics and nanoplastics on human
polyethylene microplastic ingestion on intestinal histopathology and microbiota of health. Nanomaterials 11 (2), 496. https://doi.org/10.3390/nano11020496.
gilthead sea bream. Aquaculture 545, 737245. https://doi.org/10.1016/j. Yin, F., Zhuang, Q., Chang, T., Zhang, C., Sun, H., Sun, Q., Wang, C., Li, L., 2021. Study
aquaculture.2021.737245. on pyrolysis characteristics and kinetics of mixed plastic waste. J. Mater. Cycles
Wang, B., Huang, Y., Zhang, J., 2014. Hydrothermal liquefaction of lignite, wheat straw Waste Manag. 23 (5), 1984–1994. https://doi.org/10.1007/s10163-021-01271-y.
and plastic waste in sub-critical water for oil: Product distribution. J. Anal. Appl. Zhang, C., Chen, X., Wang, J., Tan, L., 2017. Toxic effects of microplastic on marine
Pyrolysis 110, 382–389. https://doi.org/10.1016/j.jaap.2014.10.004. microalgae Skeletonema costatum: interactions between microplastic and algae.
Wang, J.L., Wang, L.L., 2011. Catalytic pyrolysis of municipal plastic waste to fuel with Environ. Pollut. 220, 1282–1288. https://doi.org/10.1016/j.envpol.2016.11.005.
nickel-loaded silica-alumina catalysts. Energ. Source Part A 33 (21), 1940–1948. Zhang, Y., Ji, G., Chen, C., Wang, Y., Wang, W., Li, A., 2020. Liquid oils produced from
https://doi.org/10.1080/15567030903436814. pyrolysis of plastic wastes with heat carrier in rotary kiln. Fuel Process. Technol.
Weber, R.S., Ramasamy, K.K., 2020. Electrochemical oxidation of lignin and waste 206, 106455 https://doi.org/10.1016/j.fuproc.2020.106455.
plastic. ACS Omega 5 (43), 27735–27740. https://doi.org/10.1021/ Zhao, G., Huang, X., Wang, X., Wang, X., 2017. Progress in catalyst exploration for
acsomega.0c03989. heterogeneous CO2 reduction and utilization: a critical review. J. Mater. Chem. A 5
Wong, S.L., Ngadi, N., Abdullah, T.A.T., Inuwa, I.M., 2015. Current state and future (41), 21625–21649. https://doi.org/10.1039/c7ta07290b.
prospects of plastic waste as source of fuel: a review. Renew. Sust. Energ. Rev. 50, Zhou, J., Cao, Y., Liu, X., Jiang, H., Li, W., 2020. Bladder entrance of microplastic likely
1167–1180. https://doi.org/10.1016/j.rser.2015.04.063. induces toxic effects in carnivorous macrophyte Utricularia aurea Lour. Environ. Sci.
Xu, B., Liu, F., Cryder, Z., Huang, D., Lu, Z., He, Y., Wang, H., Lu, Z., Brookes, P.C., Pollut. Res. 27 (25), 32124–32131. https://doi.org/10.1007/s11356-020-09529-y.
Tang, C., Gan, J., Xu, J., 2020. Microplastics in the soil environment: occurrence,

12

You might also like