Next Article in Journal
Comparative In Vitro Biocompatibility Study of the Two Orthodontic Bonding Materials of Different Types
Next Article in Special Issue
Computational Exploration of Bio-Degradation Patterns of Various Plastic Types
Previous Article in Journal
Progressive Methods in Studying the Charred Layer Parameters Change in Relation to Wood Moisture Content
Previous Article in Special Issue
Polyurethane Foam Waste Upcycling into an Efficient and Low Pollutant Gasification Syngas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Critical Review on the Progress of Plastic Bioupcycling Technology as a Potential Solution for Sustainable Plastic Waste Management

by
Passanun Lomwongsopon
and
Cristiano Varrone
*
Section of Bioscience and Engineering, Department of Chemistry and Bioscience, Aalborg University, Fredrik Bajers Vej 7H, 9220 Aalborg, Denmark
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(22), 4996; https://doi.org/10.3390/polym14224996
Submission received: 31 October 2022 / Revised: 11 November 2022 / Accepted: 14 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Advances in Biodegradation of Plastics)

Abstract

:
Plastic production worldwide has doubled in the last two decades and is expected to reach a four-fold increase by 2050. The durability of plastic makes them a perfect material for many applications, but it is also a key limitation to their end-of-life management. The current plastic lifecycle is far from circular, with only 13% being collected for recycling and 9% being successfully recycled, indicating the failure of current recycling technology. The remaining plastic waste streams are thus incinerated, landfilled, or worse, mismanaged, leading to them leaking into the environment. To promote plastic circularity, keeping material in the loop is a priority and represents a more sustainable solution. This can be achieved through the reuse of plastic items, or by using plastic waste as a resource for new materials, instead of discarding them as waste. As the discovery of plastic-degrading/utilizing microorganisms and enzymes has been extensively reported recently, the possibility of developing biological plastic upcycling processes is opening up. An increasing amount of studies have investigated the use of plastic as a carbon source for biotechnological processes to produce high-value compounds such as bioplastics, biochemicals, and biosurfactants. In the current review, the advancements in fossil-based plastic bio- and thermochemical upcycling technologies are presented and critically discussed. In particular, we highlight the developed (bio)depolymerization coupled with bioconversion/fermentation processes to obtain industrially valuable products. This review is expected to contribute to the future development and scale-up of effective plastic bioupcycling processes that can act as a drive to increase waste removal from the environment and valorize post-consumer plastic streams, thus accelerating the implementation of a circular (plastic) economy.

Graphical Abstract

1. Introduction

Demand for plastic has dramatically increased during the last decades and continues growing, reaching 460 Mt in 2019, thus doubling from the 234 Mt reported in 2000 (Figure 1) [1]. Plastics are used in a wide variety of products, dominating and outperforming other materials, as they are versatile, cheap, lightweight, and resistant; however, they are also very diverse and typically designed for endurance rather than recyclability, which often makes the end-of-life management of these materials rather challenging. As a consequence, we’ve witnessed an increased accumulation of plastic wastes in the environment and the phenomenon has now reached such an extent so as to be recognized as a global problem.
To give some examples, according to the United States Environmental Protection Agency (US EPA), in 2018, 12.2% of municipal solid waste in the United States was made of plastics, the majority of which (75%) were landfilled (Figure 2). The statistics also showed that landfilled plastics have been increasing every year since 1960, while the recycling rates are still relatively low (a little higher than 8%) [2]. Somewhat better statistics are observed in the European Union (EU), where 35% is reported to be collected for recycling, though most plastic still goes to combustion for energy recovery (42%), while the rest (23%) is landfilled [3]. Unfortunately, only a fraction of these collected 35% is then really recycled at the highest level possible, with process losses and downcycling resulting in further reduction of the actual global recycling flow [4]. On a global scale, a disappointing 13% is actually collected for recycling [1].
The mismanagement of end-of-life plastics has thus become a threat to the environment and our health, with estimated 22 Mt of global plastic leakage into the environment in 2019 [1]. By 2050, it is expected that the production and incineration of plastic could release 2.8 gigatons of CO2 per year, equal to the emissions from 615,500-megawatt coal plants [5]. For the overall plastic pollution in the marine biosphere, more than 123 Mt of plastics leaked into the ocean from 1950–2015 (Figure 3), which is the reason for the death of 1 million sea birds and 100,000 sea animals yearly [6]. According to the study of Liebmann et al., PP, PET, PS and PE were detected in fecal samples of tested participants consuming seafood, implying that plastic pollution has already started to affect human health [7]. In fact, microplastics could potentially cause alterations in human chromosomes, leading to infertility, obesity and cancer [8].
Notably, the problems of the current plastic sector are not only related to the management of end-of-life plastic, but also the considerable use of virgin fossil resources in plastic production, which is accelerating the world to petroleum depletion. The linear flow of the plastics value chain (produce–use–discard) is among the main causes of the above-mentioned negative impacts. The European Commission has therefore recently introduced its circular economy action plan, to encourage a more sustainable value chain and ensure that post-consumer waste is kept in the loop (at the highest possible level) for as long as possible [11]. To reach this goal, new technological solutions are needed, with recycling strategies that help improving the techno-economic feasibility of the recycled plastics (still not competitive with virgin fossil ones [12] or that lead to improved value/properties of the new plastics (defined as “upcycling”) [13]. In fact, the currently implemented recycling technology mainly transforms the plastic waste into lower-value products (downcycling) or, at best, into the same level. In order to boost the development of these new technologies for plastic upcycling, the EC has been financing a significant amount of research and innovation projects, addressing both, chemical (i.e., iCAREPLAS, https://www.icareplast.eu (accessed on 7 March 2022); MultiCycle, http://multicycle-project.eu (accessed on 7 Mar 2022)) and biological processes (UPLIFT, https://upliftproject.eu (accessed on 7 March 2022); PRESERVE, https://www.preserve-h2020.eu (accessed on 7 March 2022); UpPE-T, https://uppet.eu (accessed on 7 March 2022); MIX-UP, https://www.mix-up.eu (accessed on 7 March 2022); BioICEP https://www.bioicep.eu (accessed on 7 March 2022); MIPLACE, https://miplacebio.com (accessed on 7 March 2022); Enzycle, https://www.enzycle.eu (accessed on 7 March 2022)).
Biological plastic upcycling is performed by using plastic waste streams as a carbon substrate for biotechnological processes, similarly to the approach used for lignocellulosic feedstocks [14]. Several enzymes have been identified that present hydrolytic properties to depolymerize certain plastics, while microbial fermentation processes can be developed to convert depolymerized plastics to higher-value products, for example, biopolymers [15,16,17]. Nevertheless, the biological route is generally hindered by plastic recalcitrance, typically related to hydrophobicity, and the crystallinity. The biodegradability of plastics can be very different, depending on their chemical structure: polyolefins such as polyethylene (PE) and polypropylene (PP) (which are among the most abundant plastics, Figure 4), for instance, cannot be easily degraded by microorganisms because of their recalcitrant C–C bonds, while polyesters are much more prone to be attacked at the hydrolyzable ester units. Several other factors, e.g., surface conditions, molecular weight, thermal properties, etc., also affect the biodepolymerization of plastic [18].
Most of the reviews dealing with plastic waste management focus on conventional processes, e.g., landfilling, incineration, mechanical, and chemical recycling [19,20,21] that typically downcycle the materials. They do not analyze the advantages of integrated processes that combine thermochemical and biochemical technologies. Up to date, relatively few studies have been focusing on plastic bioupcycling strategies; however, new low-cost biotechnological processes could act as fundamental drives for the valorization of plastic waste streams that currently are not effectively recycled. In fact, by converting such streams into higher-value upcycled polymers, the post-consumer plastic would no longer be regarded as a waste but as a valuable feedstock, thus contributing to keeping the materials in the loop. Therefore, the overall aim of this paper is to provide a comprehensive and critical review on plastic bioupcycling by identifying the potential strengths and limitations of these technologies, and thus speeding up the development of a more sustainable and economically feasible plastic sector. The review is structured based on different plastic types, where the depolymerization and fermentation steps are indicated separately, to highlight the challenges and technological solutions of the different steps.

2. Recycling of Conventional Plastic Wastes

Conventional recycling methods cannot keep up with the increasing amount and variety of plastic wastes. Mechanical recycling mainly involves grinding and pelletizing of relatively clean plastic streams into small particles, which are then reformed into new products without significant changes in the chemical structure. It represents the most mature recycling technology, with relatively low greenhouse gas emissions, and has been extensively reported in the literature [22]. Examples of mechanical recycling include the remolding of polyolefins waste for outdoor furniture, decking, and fencing, or of polyethylene terephthalate (PET) bottles to shoes. However, in the real situation, large streams of post-consumer plastic waste come in the form of (contaminated) mixed plastics, e.g., multilayer films (combination of different polymer types), galvanized polymers (combination of plastic with other materials such as metals, carbon, glass fibers, etc.), and additives (such as flame retardants and plasticizers) [23,24,25]. Mechanical recycling requires a sorting and cleaning process before recycling, which makes the handling of contaminated and/or mixed plastic streams extremely challenging, often leading to downcycling into less-valuable products. The number of cycles is also limited, due to the deterioration of plastic materials [26], and as a consequence, new technologies such as chemical recycling have been developed to reduce such limitation.
Chemical recycling includes thermochemical and catalytic conversions such as pyrolysis, gasification, fluid-catalyzed cracking, hydrocracking, and chemolysis (glycolysis, hydrolysis, methanolysis, aminolysis) [27]. These processes break down the polymer at high temperatures, with or without catalysts, to a mixture of oligomers/monomers and/or gaseous products and are thus applicable for handling heterogenous and contaminated plastic [28]. However, these recycling methods are usually costly, frequently use very large amounts of chemicals, and/or are energy-intensive, leaving behind hazardous gases and toxic residues [29]. Notably, chemical conversion emits more greenhouse gases (per ton of plastic utility treated) than most other treatment types, with the only exception of incineration.
Overall, even though conventional recycling methods reduce the amount of plastic going to landfills, the existing technologies often tend to reduce the properties of plastic waste-derived monomers and accelerate their end-of-lifetime [30,31]. As a consequence, new studies are starting to present biotechnological recycling strategies as a much-needed complementary solution to the end-of-life of those challenging plastic waste streams that currently are not effectively recycled. In fact, applying biological depolymerization and bio-recycling to plastics waste provides the opportunity to produce higher-value products in more sustainable and more mild conditions (lower temperature and energy requirements, absence of toxic chemicals, etc.), without the need of previous sorting. A good example is the new biotechnology developed by Carbios and the University of Toulouse that enables the efficient enzymatic depolymerization of post-consumer PET bottles on an industry-relevant scale and processing time (90% depolymerization into monomers in only 10 h) [32].

3. Bioupcycling

Besides biotechnological recycling, new upcycling processes are now extensively studied and under development. The possibility to convert post-consumer PET to polyhydroxyalkanoate (PHA) by enzymatic depolymerization and subsequent bacteria fermentation [16], for instance, allows researchers to obtain a bio-material with good technical substitution potential, novel properties (depending on the co-polymer) and biodegradability. From this perspective, post-consumer plastic can be upcycled rather than only recycled. Moreover, biological methods have the advantage that they can be applied to contaminated plastic waste (i.e., food or soil) and do not require previous separation of the different fractions. Furthermore, the high selectivity of enzymes could allow for a stepwise removal of specific components of the mixed-plastic waste, facilitating the downstream processing; thus, it can go beyond the limits of mechanical and chemical recycling [33].
Recent research in the biodegradation of plastic waste allowed for the establishment of plastic biodepolymerization processes for some plastic types, thus paving the path for greener plastic recycling processes [34,35,36,37,38,39,40,41].
Both natural and engineered enzymes for plastic depolymerization have been extensively studied [32,34,42,43,44,45,46]. Even though the cost of the enzymes is still of concern, a recent study on enzymatic recycling of PET has determined that it should only contribute to 4% of the overall operating costs [47]. The authors highlighted that an enzyme-based recycling process can be cost-competitive, and the constant development of enzyme performance and optimization of the process remains an opportunity to further improve the economic viability of this process. Once bioprocesses for bulk enzyme production have been established, the enzymatic degradation of plastic is a promising technology that will be implemented in the near future. Depolymerization enzymes can be utilized in multiple ways, including free enzymes, immobilized enzymes, extracellular enzymes of whole-cell biocatalysts, surface enzymes, and/or in the form of enzyme cocktails [48]. Moreover, synthetic biology can be applied to improve the catalytic activity of enzymes through protein engineering, e.g., direct evolution and rational protein design [49]. Plastic-degrading enzymes in microorganisms can evolve from their natural activity on recalcitrant biopolymers such as lignocellulose, chitin, and cutin. Polyester plastics could, for instance, be depolymerized by hydrolytic enzymes produced by bacteria or fungi, such as cutinases, esterases, lipases, ureases, and proteases, as they have hydrolyzable ester bonds in their backbone (Satti and Shah, 2020). An excellent example is the bacterial enzyme polyurethane hydrolases (PUase), capable of hydrolyzing polyurethane (PU) [14], or the well-known PETase and MHETase from Ideonella sakaiensis 201-F6, able to hydrolyze PET [45]. For plastics with a non-hydrolyzable C–C backbone, such as polyolefins (PE and PP), oxidative enzymes play a significant role in introducing active functional groups into the backbone, which consequently can undergo biodegradation [50]. Alkane hydroxylases from isolated Pseudomonas sp. E4 was also reported to play an important role in PE biodegradation [51]. This enzyme acts on the hydroxylation of C–C bonds to release primary or secondary alcohols, which are then further oxidized to ketones or aldehydes, and finally to carboxylic acids [52]. Thus, the identification and optimization of efficient enzymes represents a prerequisite for the development of bioupcycling processes. In fact, the enzymes are not going to use the polymers as a carbon source, as would happen with a microbial cell that degrades the polymers. This allows to fully recover the monomers for the subsequent upcycling step, for instance through the contribution of fermentation processes [13].
So, in conclusion, the established knowledge on biodepolymerization is expected to boost the development of new biocatalytic plastic upcycling processes, to produce value-added products and/or generate new (and more renewable) plastics with better properties, out of conventional plastic wastes. In this sense, the cost of renewable plastic production (which is still too high compared to conventional fossil-based plastics and related to feedstocks’ price and fluctuation) could be alleviated. Hence, renewable plastic will become more economically viable for general commodities, once large-scale production is reached. At present, bioupcycling with endogenous or engineered metabolic pathways has been demonstrated as a proof of concept and is a hot topic for researchers worldwide. The most relevant published studies have been reviewed in the sub-section below.

3.1. Bioupcycling of Polyethylene Terephthalate (PET)

PET is a petrochemical-based plastic produced on multimillion tons worldwide. It is a polyester made of the repeating units of ethylene glycol (EG) and terephthalic acid (TA). It is cheap and has a very low permeability to gas and moisture, making it an ideal material for single-use plastic bottles. The demand of PET worldwide is around 29 Mt in 2022 [53], while in Europe it reached 4.1 Mt in 2020 [3]. However, even if PET has a higher recycling rate than other plastic types (50% in Europe and 23% in the US for PET bottles), 69% of recycled PET was used for lower-grade applications such as trays, film, strapping, or fiber [54]. From these statistics, the PET management system clearly does not show a very high level of circularity; therefore, there is a clear need for new upcycling approaches to valorize PET into higher-value products or, at least, keep the material at the same level in the value chain.
Various upcycling strategies have been reported to valorize PET waste, including biotechnological processes and bioupcycling (Table 1). It typically combines depolymerization and fermentation/bioconversion strategies to produce new valuable products. Chemical processes (e.g., hydrolysis, alcoholysis, glycolysis, aminolysis, ammonolysis, and hydrogenolysis [55]) and thermal processes (e.g., hydrothermal liquefaction, pyrolysis, and microwave irradiation) are generally used for the depolymerization of PET. Interestingly, a considerable achievement through enzymatic or microbial degradation has also been reported, especially on PET, that can lead to the development of more sustainable bioupcycling processes. Since PET monomers are linked through hydrolyzable ester bonds, an increasing number of PET hydrolyzing (and/or surface modifying) enzymes have been reported recently. One of the major findings was, for instance, the discovery of Ideonella sakaiensis [45], a bacterium that can grow on PET as the only carbon source, thanks to the synergy of its enzymes IsPETase and MHETase that break down PET to mono-(2-hydroxyethyl)terephthalic acid (MHET), di-(2-hydroxyethyl)terephthalic acid (BHET), and finally to EG and TA. Subsequently, several studies reported the improvement of the mesophilic IsPETase through protein engineering, for example by increasing the hydrolytic activity [56] or the thermal stability, using rational protein engineering [34,43,57].
In order to boost enzyme production, several studies have been investigating the addition of signal peptides to the IsPETase gene to enhance its secretion during heterologous expression in Escherichia coli [58]. The use of heterogeneous immobilized biocatalysis through the use of different binding modules and linkers has also been extensively investigated [59,60]. Some of the most effective enzymes that act on PET hydrolysis are thermostable cutinases such as Humicola insolens cutinase (HiC) [35], Thermobifida fusca cutinase (TfCut2) [46], and leaf-branch compost cutinase (LCC) [61]. Improving their hydrolysis activity has also been reported using various methods, for example, by enhancing electrostatic interaction between TfCut2 and PET by cationic surfactant additive-based approach, and showed impressive biodegradation of PET by 97% within 30 h [36]. In general, the design of thermostable hydrolases has been intensively investigated during the last years [62] and resulted in the development of promising processes for enzymatic depolymerization. One of the biggest breakthroughs so far is probably represented by the study by Tournier and colleagues that engineered the LCC cutinase and increased its optimal reaction temperature up to the glass transition temperature (Tg) of PET [32]. Such bioprocesses can represent a valuable and sustainable alternative to thermochemical depolymerization and are laying the basis for the further conversion of plastic waste-derived monomers.
Bioupcycling post-consumer PET (or even polyolefins) to PHA has been gaining attention since this approach promotes the valorization of plastic waste into renewable biopolymers. PHA is a general term for microbial polyester of R-3-hydroxyalkanoic acids [63]. It is a promising substitute for several petroleum-based plastics, due to its superior thermal processability, biodegradability, and biocompatibility properties [64,65]. Kenny et al. reported using a two-step chemo-biotechnological process for upcycling PET to PHA [63]. First, the PET waste was hydrolyzed by pyrolysis at 450 °C at a feed rate of 1 kg/h, obtaining 77% of solid fraction (TA, oligomers, benzoic acid and others), 18% gas fraction (CO2, CO, H2, ethene, and others), and 6.3% liquid fraction (EG, acetic aldehyde and others). Second, the TA was dissolved in NaOH to generate sodium terephthalate, which was used as the sole carbon source to grow a PHA-accumulating strains. Pseudomonas putida GO16, P. putida GO19, and P. frederiksbergensis GO23 were found to consume sodium terephthalate and accumulated PHA up to 23–27% of CDW. GO19 was the most efficient at converting TA to PHA, with a productivity of 8.4 mgPHA/L/h. The research group continued to develop the bioprocess to enhance PHA production from TA by co-feeding with waste glycerol (WG) [66]. The fermentation was designed to have two distinct phases: the biomass growth phase and the PHA production phase. They found that when P. putida GO16 was fed with WG only during the growth phase (0–24 h) and WG and TA during the PHA production phase (24–48 h), the highest total PHA production was achieved (5.30 g/L). This strategy promoted a 2.0-fold higher PHA production than feeding with TA alone. This study showed that bioprocess engineering strategies are key to develop highly efficient bioupcycling of plastic waste.
Tiso et al. reported the bioupcycling of PET to PHA and hydroxyalkanoyloxyalkanoate (HAA) by using a combination of enzymatic hydrolysis and whole-cell biocatalyst [16]. PET was hydrolyzed by LCC, a polyester hydrolase capable of efficient PET depolymerization to TA and EG. Pseudomonas umsongensis GO16 KS3 was found to consume both TA and EG (the latter at a 3.5-fold lower rate than TA) within 23 h of cultivation, in a 5 L bioreactor producing PHA. The PHA production only reached 0.15 g/L, or 7% of CDW. Interestingly, HAA was produced from TA only, while EG was exclusively used for growth. A maximum HAA concentration of 35 mg/L was achieved with a production rate and yield of 5 mg/L/h and 0.01 gHAA/gTA, respectively [16]. Despite the low performance, this represents a highly interesting approach with potential industrial applications (once optimized). In fact, HAA can be directly polymerized with 4,4′-methyl diphenyl diisocyanate and butanediol (BDO) to form biopoly(amide urethane) (bio-PU).
Another study reported the bioconversion of TA, a monomer of PET, to muconic acid (MA), which is currently used to produce adipic acid (AA), an important monomer for various plastics [67]. PET was first depolymerized by microwave radiation at 230 °C for 50 min to TA and EG. Then, the E. coli strain CTL-1 (expressing TphAabc, TphB, and AroY, which is responsible for converting TA to catechol, via 1,2-dihydroxy-3,5-cyclohexadiene-1,4-dicarboxylate (DCD) and protocatechuic acid (PCA)) and the strain MA-1 (expressing CatA, which is a catechol 1,2-dioxygenase responsible for converting catechol to MA), were combined for MA synthesis. The MA concentration reported was 2.7 mM, accounting for a 85.4% molar yield (MA/TA). In the same study, gallic acid (GA), pyrogallol, vanillic acid (VA), and glycolic acid (GLA) were also produced from engineered stains, using TA. In addition, GLA was produced by EG-fermenting Gluconobacter oxydans KCCM 40109.
The recent study by Sadler and Wallace (2021) showed the development of a one-pot bioprocess to convert TA from PET waste into a value-added molecule, vanillin. PET from a post-consumer plastic bottle was firstly hydrolyzed to TA by semi-purified LCC at 72 °C. Then, the reaction was cooled down and the engineered strain, E. coli RARE_pVanX, was added to perform the bioconversion. E. coli RARE_pVanX was constructed with two plasmids that were encoded for different enzymes (terephthalate 1,2-dioxygenase, dihydroxy-3,5-cyclohexadiene-1,4-dicarboxylic acid dehydrogenase, carboxylic acid reductase, and catechol O-methyltransferase), which convert TA to vanillin via intermediates (PCA, VA, and dihydroxybenzaldehyde). The process optimization was performed by screening the protein-expression media (M9-glucose supplemented with L-Met and n-butanol (nBuOH)), increasing E. coli cell membrane permeability to TA (addition of 1% v/v n-BuOH), adjusting pH (5.5) and temperature (22 °C), and mitigating the toxicity of vanillin by in situ product removal, using oleyl alcohol. The final production of vanillin reached 789 μM (119 mg/L) or 79% conversion. [68].
The chemo-biological upcycling of PET to the multifunctional coating material, catechol, was also reported. First, PET waste was glycolyzed to a mixture of PET oligomers. Then, an enzymatic hydrolysis of the glycolyzed products was performed, turning the mixture without previous purification into TA, by Bacillus subtilis esterase (Bs2Est). The catechol production from TA was consequently conducted, using a catechol biosynthesis strain (obtained through the combination of the TA degradation module and catechol biosynthesis module in E. coli). The final titer of catechol was 5.97 mM, accounting for 99.5% conversion by mol of TA. Catechol shows great functions as a coating material without the need for an adherence layer, and its antibacterial activity is comparable to chitosan [69].
Table 1. PET bioupcycling.
Table 1. PET bioupcycling.
Depolymerization StrategyDepolymerization
Products Used as a
Feedstock for
Fermentation Step
Fermentation StrategyProducts from FermentationTiterProductivityYieldRef.
Hydrolytic pyrolysis
at 450 °C
Solid product mixture (terephthalic acid (TA), oligomers, benzoic acid, and others)Fermentation in shake flask containing 4.2 g/L of PET-derived sodium terephthalate and 67 mg/L of nitrogen at 30 °C for 48 h by Pseudomonas putida GO16medium chain length PHA (mclPHA)0.25 g/L8.4 mgPHA/L/h0.27 gPHA/gCDW[63]
Hydrolytic pyrolysis
at 450 °C
Solid product mixture (TA, oligomers, benzoic acid, and others)Fermentation in shake flask containing 4.2 g/L of PET-derived sodium terephthalate and 67 mg/L of nitrogen at 30 °C for 48 h by P. putida GO19mclPHA0.25 g/L8.4 mgPHA/L/h0.23 gPHA/gCDW[63]
Hydrolytic pyrolysis
at 450 °C
Solid product mixture (TA, oligomers, benzoic acid, and others)Fermentation in shake flask containing 4.2 g/L of PET-derived sodium terephthalate and 67 mg/L of nitrogen at 30 °C for 48 h by P. putida GO23mclPHA0.27 g/L4.4 mgPHA/L/h0.24 gPHA/gCDW[63]
PyrolysisTAFed-batch fermentation in 19.5 L-stirred tank reactor with controlled pH of 6.9 and dissolved oxygen (DO) level above 40% at 30 °C for 48 h by P. putida GO16 supplied with TA as the sole growth and PHA substratemclPHA2.61 g/L0.05 g/L/h0.30 gPHA/gCDW[66]
PyrolysisTAFed-batch fermentation in 19.5 L-stirred tank reactor with controlled pH of 6.9 and DO level above 40% at 30 °C for 48 h by P. putida GO16 supplied with waste glycerol (WG) as growth substrate and TA as PHA substratemclPHA5.22 g/L0.11 g/L/h0.36 gPHA/gCDW[66]
PyrolysisTAFed-batch fermentation in 19.5 L-stirred tank reactor with controlled pH of 6.9 and DO level above 40% at 30 °C for 48 h by P. putida GO16 supplied with TA as the sole growth and PHA substratemclPHA5.30 g/L0.11 g/L/h0.35 gPHA/gCDW[66]
PyrolysisTAFed-batch fermentation in 19.5 L-stirred tank reactor with controlled pH of 6.9 and DO level above 40% at 30 °C for 48 h by P. putida GO16 supplied with WG as growth and PHA substrate and TA as PHA substrate onlymclPHA4.98 g/L0.10 g/L/h0.35 gPHA/gCDW[66]
PyrolysisTAFed-batch fermentation in 19.5 L-stirred tank reactor with controlled pH of 6.9 and DO level above 40% at 30 °C for 48 h by P. putida GO16 supplied with WG and TA as both growth and PHA substratesmclPHA4.42 g/L0.09 g/L/h0.36 gPHA/gCDW[66]
Enzymatic degradation by recombinant leaf-branch compost cutinase (LCC)TA, ethylene glycol (EG), mono-(2-hydroxyethyl)terephthalic acid (MHET), di-(2-hydroxyethyl)terephthalic acid (BHET)Fermentation in 5 L-stirred tank reactor with controlled pH of 7.0 and DO level above 20% at 30 °C for 28 h Pseudomonas umsongensis GO16 KS3 supplied with hydrolyzed PET at the amount to yield 40 mM of TA and EG and limited inorganic nutrientmclPHA0.15 g/LNA0.014 gPHA/gSubstrate[16]
Enzymatic degradation by recombinant LCCTA, EG, MHET, BHETFermentation in shake flask containing Delf medium with diluted (1:20) hydrolyzed PET (TA and EG concentration of 15–18 mM) at 30 °C for 24 h by P. umsongensis GO16 KS3 pSB01Hydroxyalkanoyloxy-alkanoate (HAA)35 mg/L5 mg/L/h0.01 gHAA/gTA[16]
Microwave radiation for 50 min
at 230 °C
TABioconversion by metabolically engineered E. coli strain PCA-1 and HBH-2 to convert TA to intermediate protocatechuic acid (PCA), and then to gallic acid (GA), at 30 °C and 250 rpm for 24 h in 50 mM Tris buffer (pH 7.0) containing 2% (w/v) glycerolGA2.7 mMNA0.925 MGA/MTA[67]
Microwave radiation for 50 min
at 230 °C
TABioconversion by metabolically engineered E. coli strain PG-1a to convert TA to intermediate PCA, GA, and then pyrogallol, at 30 °C and 250 rpm for 6 h in 50 mM Tris buffer (pH 7.0) containing 2% (w/v) glycerolPyrogallol1.1 mMNA0.327 MPyrogallol/MTA[67]
Microwave radiation for 50 min
at 230 °C
TABioconversion for 6 h by metabolically engineered E. coli strain CTL-1 and MA-1 to convert TA to intermediate catechol, and then to muconic acid (MA)MA2.7 mMNA0.854 MMA/MTA[67]
Microwave radiation for 50 min
at 230 °C
TABioconversion using double-catalyst VA-2a system for 48 h by metabolically engineered E. coli strain PCA-1 and OMT-2His to convert TA to intermediate PCA and then to vanillic acid (VA), in 50 mM Tris buffer (pH 7.0) containing 10% (w/v) glycerol, 10 g/L yeast extract, 20 g/L peptone, and 2.5 mM L-methionineVA1.4 mMNA0.416 MVA/MTA[67]
Microwave radiation for 50 min
at 230 °C
EGBioconversion by Gluconobacter oxydans KCCM 40109 using 10.7 mM of EG from PET hydrolysate as a feedstock, at 30 °C and 220 rpm in shake flask at the working volume of 1 LGlycolic acid (GLA)NANA0.986 MGLA/MEG[67]
-EG (mock substrate to study upcycling of PET-derived monomer)Fermentation in shake flask containing 10% (v/v) EG in 250 mM potassium phosphate buffer (pH 7.0) at 30 °C with gentle stirring and aeration at 1 VVM for 120 h by Pichia naganishii AKU 4267GLA105 g/LNA0.880 MGLA/MEG[70]
-EG (mock substrate to study upcycling of PET-derived monomer)Fermentation in shake flask containing 10% (v/v) EG in 250 mM potassium phosphate buffer (pH 7.0) at 30 °C with gentle stirring and aeration at 1 VVM for 120 h by Rhodotorula sp. 3Pr-126GLA110 g/LNA0.922 MGLA/MEG[70]
-EG (mock substrate to study upcycling of PET-derived monomer)Fermentation in shake flask containing 100 mM of EG in nitrogen limiting M9 medium (0.132 g/L of (NH4)2SO4) at 30 °C for more than 72 h by P. putida MFL185 (engineered strain that has the tac promoter inserted before the native glycolate oxidase operon and harbor overexpression)mclPHANANA0.32 gPHA/gCDW and 0.06 gPHA/gEG[71]
-EG (mock substrate to study upcycling of PET-derived monomer)Anaerobic fermentation of 50 mM EG at 30 °C by acetogenic bacterium Acetobacterium woodiiAcetate10.4 mM3.6 μmol/mg/hNA[72]
-EG (mock substrate to study upcycling of PET-derived monomer)Anaerobic fermentation of 50 mM EG at 30 °C by acetogenic bacterium A. woodiiEthanol12.0 mM4.8 μmol/mg/hNA[72]
Enzymatic degradation by semi-purified LCC (pH 10.0) at 72 °C for 48 hPET hydrolysateBioconversion using metabolically engineered E. coli RARE_pVanX to convert TA to intermediate protocatechuate (PC), and then to vanillin using optimized condition: M9-glucose supplemented with L-Met and nBuOH as a protein expression media, pH 5.5, room temperature for 24 h, in situ product removal (ISPR) by oleyl alcoholVanillin300–400 μMNANA[68]
-TA (mock substrate to study upcycling of PET-derived monomer)Bioconversion using metabolically engineered E. coli RARE_pVanX to convert TA to intermediate PC and then to vanillin using optimized condition: M9-glucose supplemented with L-Met and nBuOH as a protein expression media, pH 5.5, room temperature for 24 h, ISPR by oleyl alcoholVanillin789 μMNA0.79 Mvanillin/MTA[68]
Chemical glycolysis
at 200 °C for 3 h
Mixture of BHET, MHET, and PET oligomers at 84.8, 7.7, and 8.7%, respectivelyEnzymatic hydrolysis of the glycolyzed products (the mixture) into TA by Bacillus subtilis esterase (Bs2Est) (2 U/mL at 30 °C and 1000 rpm), following by producing catechol from PET hydrolysates using a catechol biosynthesis strain that was established using the combination of the TA degradation module and catechol biosynthesis module in E. coli (in 12 h)Catechol5.97 mMNA0.995 MCatechol/MTA[69]
Chemocatalytic glycolysis: PET was glycolyzed with EG as a solvent (1:4 w/w) and catalyzed by 1% (w/w) titanium (IV) butoxide at 220 °C overnightBHETFermentation in 3 L-bioreactor with batch culture in the first 4 h fed with 4-hydroxybenzoic acid to induce the β-ketoadipate pathway, followed by fed-batch culture using BHET as a carbon source (pulse adding at 9.1, 23.3, 32.8, 48.2 and 73.9 h). The sequential metabolic engineered Pseudomonas putida KT2440 (constitutive expression of native genes for EG utilization, expression of gene for TA catabolism, expression of PETase and MHETase for BHET hydrolysis, and gene deletion to enhance β-ketoadipic acid production) was used for bioconversion.β-ketoadipic acid15.1 g/L0.16 g/L/h0.76 Mβ-ketoadipic acid/MBHET[73]
Chemocatalytic glycolysis and enzymatic hydrolysis: PET was glycolyzed with EG as a solvent and catalyzed by betaine at 190 °C for 30–120 min, followed by enzymatic hydrolysis (PETase and MHETase)TAWhole cell bioconversion of TA (4.5 g/L) to protocatechuic acid by metabolically engineered E. coli PCA-1 was performed in shake flask at 30 °C and 250 rpmPCA3.8 g/L-0.904 MPCA/MTA[74]
Chemocatalytic glycolysis and enzymatic hydrolysis: PET was glycolyzed with EG as a solvent and catalyzed by betaine at 190 °C for 30–120 min, followed by enzymatic hydrolysis (PETase and MHETase)EGWhole cell bioconversion of EG (30.6 g/L) to GLA by Gluconobacter oxydan KCCM 40109 was performed in shake flask at 30 °C and 200 rpm.GLA31.4 g/L-0.916 MGLA/MEG[74]

3.2. Bioupcycling of Polyurethanes (PU)

PU is a general term used for a class of polymers typically derived from the polycondensation of (poly)isocyanates (−NCO) and polyols (exothermic reactions) [75]. There are three main types of PU: polyester, polycaprolactone, and polyether urethanes, depending on the type of long-chain diol used in PU production [76]. PUs are used in many fields of application, such as, building insulation, pillows and mattresses, and insulating foams for fridges [3]. In 2020, PU production was 3.81 Mt in Europe, representing 7.8% of total plastic production [3]. Since many PU types have a thermoset nature with covalent cross links, recycling is still extremely challenging. Almost 50% of PU produced in the European Union goes to a landfill, 45% being incinerated with energy recovery, and only 5% is mechanically recovered [77]. The statistics from the US show that the mechanical recycling of PU on a large scale has only been performed on flexible PU foams by shredding and mixing the scrap polymers with binders to produce padding-type products; such processes have the limitation that they can be repeated only 8–10 times, until reaching the end-of-life [78]; this represents a common problem with mechanical recycling. Chemical recycling methods used for PU may include hydrolysis, glycolysis, and aminolysis, which use water, glycol, and ammonia or ammonium hydroxide as a nucleophile, respectively [79]. Europe’s first chemical recycling plant for post-consumer PU foam was constructed very recently in Semoy, France, by the Dow Chemical Company and partners, employing glycolysis for recovering polyols [80]. Another method called acidolysis, based on a combination of two carboxylic acids and an unspecified non-metallic catalyst, has been successfully employed by H&S, a Poland-based company, to break up PU and recover polyols [81]. Even though these processes can recover chemical constituents from PU waste, the negative impact from the toxic isocyanides and the huge amount of chemical catalysts consumption cannot be neglected. Therefore, despite the recalcitrance of these polymers, biodegradation of PU using microorganisms (i.e., fungal species) or enzymes is gaining more interest lately, as a more environmentally friendly and sustainable alternative [82,83]. On the other hand, since PUs represent a large class of plastics, it is difficult to point to a single enzyme that can degrade them all. In the presence of hydrolase-type enzymes, e.g., lipase, esterase, and protease, PUs can be hydrolyzed by a three-step mechanism: (1) chemical dissolution of ester and amide bonds in the polymer chain; (2) decreasing molecular weight and viscosity; (3) cleaving all polymer chains [76]. However, urethane forms stable bonds that can be hydrolyzed at a slower rate than ester bonds [84]. Bacillus subtilis strain MZA-75 [85] and Pseudomonas aeruginosa strain MZA-85 [86] were reported for polyester-based PU degradation. These two strains were also studied in bacterial co-culture and found to utilize polyester PU as a carbon and energy source more efficiently, as compared to the individual strains. The main degradation products were BDO and AA [87]. Another study investigated PU degradation by the enriched mixed-microbial consortia and reported the presence of extracellular enzymes, such as esterases, proteases, and ureases, affecting ester, urethane, ether, aromatic, and aliphatic groups of the plastic [88]. Fungi, such as Aspergillus tubingensis [89], Penicillium sp. [90], Cladosporium cladosporioides complex [91], or fungal communities [92] were also reported for their ability to degrade PU.
Besides BDO and AA, other PU degradation products are EG, 2,4′-toluenediamine (TDA), and 4,4′-methylenedianiline (MDA). They can be used to synthesize new PU or other polyesters, e.g., PHA, polybutylene succinate (PBS), poly(1,3-propylene succinate-ran-1,4-butylene succinate) (PPBS), and poly(1,3-propylene adipate-ran-1,4-butylene adipate) (PPBA) [79].
Last but not least, a recent study reported the upcycling processes of PU waste to rhamnolipid biosurfactants, [75] (Table 2). They successfully showed the growth of a defined mixed culture composed of three Pseudomonas putida KT2440 strains obtained by adaptive laboratory evolution on mock PU hydrolysate. The highest specific growth rate was 0.21 h−1 on the mixture of AA, 1,4-BDO, and EG. After engineering of the mixed culture, 0.10 g/L of rhamnolipids were produced from the AA, BDO, and EG mixture.

3.3. Bioupcycling of Polyolefins

Polyolefins represent the most abundant polymers, accounting for 42% of all plastics globally produced and 50% of plastics produced in Europe (Figure 4). In the packaging sector, around 70% of the total plastic waste generated is made of polyolefins [1]. The highest demand of polyolefins is PE (LLDPE, LDPE, MDPE, and HDPE), followed by PP. Polyolefins are non-polar materials, durable, chemically resistant, and have low permeability; thus, they are used in various applications, especially for flexible food packaging [94]. However, their durability also make them retain in nature for a very long time (the half-life can be up to 5000 years on land) [95]. This is due to the stable carbon to carbon bonds of the polyolefins that make biodegradation much more challenging than in polyesters (e.g., PET) [96]. However, due to its abundance in our waste, it is imperative to look into bioupcycling strategies for this type of plastic. Instead of mineralization or depolymerization to gaseous monomers, polyolefins’ recalcitrancy could be seen as an opportunity for upcycling by recovering and converting oligomers released during biofragmentation.

3.3.1. Polyethylene (PE)

A major type of polyolefins used globally is PE, accounting for around 25% of the market share [1]. It is widely used in various applications such as plastic bags, composite films, food and drink packaging, and bubble wraps. PE is a very high-molecular-weight hydrocarbon plastic with recalcitrant C–C bonds; thus, biodegradation of PE needs to be initiated by a pre-oxidation step, i.e., physicochemical oxidation with the formation of free radicals. The radicals can react with oxygen and form peroxyl groups under aerobic conditions, while they create end terminal unsaturated double bonds under anaerobic conditions [95]. The hydroperoxyl group is highly reactive and can generate several types of oxygen-containing products, including carbonyl groups. The oxidation of polyolefins can be followed by FTIR through the carbonyl peak formation and expressed as a carbonyl index (the ratio of carbonyl absorbance (1715 cm−1) to an invariant absorbance characteristic) [97].
The polymer containing oxygenated groups is much more prone to undergo biodegradation; however, not one but a cascade of enzymes is thought to be involved in polyolefins biodegradation. Studies reported that oxidative enzymes such as peroxidases, oxygenases, and laccases could play a significant role in PE biodegradation [98,99,100,101,102,103,104]. A laccase from Rhodococcus ruber C208, for instance, was found to reduce 20% and 15% of the average molecular weight (Mw) and average molecular number (Mn) of PE and increase the carbonyl group formation in PE structure, as evidenced by gel permeation chromatography (GPC) and attenuated total reflection Fourier-transform infrared spectroscopy (ATR-FTIR) [104]. A recent study by Yao et al. reported different effects of laccases from Botrytis aclada (BaLac) and Bacillus subtilis (BsLac) on high-temperature UV-irradaited PE, depending on their different redox potential. New functional groups, including -OH, -C=O, and C=C, were detected in PE chains after exposure to the laccase-mediator system. The Mw of PE was reduced by 40% with the BaLac degradation. The degradation products were aldehydes, ketones, and alcohols between C4–C8 (BaLac-treated) and C7–C15 (BsLac-treated) [37].
Alkane hydroxylases (AHs) were also reported to be involved in polyolefins degradation. AHs are known to convert terminal CH3 of alkane to 1-alkanol, which then can be oxidized by several other enzymes to fatty acids and then enter the β-oxidation pathway to provide energy for microbial growth [105]. AH systems include alkane monooxygenase (AlkB), electron-transport protein rubredoxin, and rubredoxin reductase, catalyzing hydroxylation of alkanes at the terminal carbon (ω-position) in aerobic bacteria [40,106]. Jeon and Kim showed that AHs from P. aeruginosa E7 are involved in the biodegradation of low-molecular-weight PE, confirmed by the conversion of approximately 19% of PE into CO2 during biodegradation at 37 °C, using an engineered E. coli strain expressing AH [40]. They also investigated the effect of induction conditions: the addition of n-hexadecane induced the transcription level of alkB1 in E. coli, whereas alkB2 was induced by both n-hexadecane and n-dodecane [107]. Microbial oxidation of PE is proposed to be similar with the metabolic pathway of linear n-alkanes, as shown in three pathways below [108,109].
1. 
Terminal oxidation:
RCH3 → RCH2OH → RCHO → RCOOH
2. 
Bi-terminal oxidation:
H3CRCH3 → H3CRCH2OH → H3CRCHO → H3CRCOOH → HOH2CRCOOH → OHCRCOOH → HOOCRCOOH
3. 
Subterminal oxidation:
RCH2CH2CH3 → RCH2CH(OH)CH3 → RCH2C(O)CH3 → RCH2OC(O)CH3 → RCH2OH + CH3COOH
Cytochromes P450, an iron (Fe)-containing heme protein, among the most versatile biocatalysts in nature, are as also foreseen to be part of the PE-degrading machinery [106]. They are monooxygenases that are able to introduce one atom of O2 into a wide variety of organic substrate molecules. P450s can catalyze many reaction types: C-H hydroxylation; C=C double bond epoxidation; heteroatom oxygenation; O-, N-, and S-dealkylation; aromatic coupling; and C-C bond cleavage [106]. In particular, they can hydroxylate linear alkanes, alcohols, and fatty acids of various chain lengths [110], which are intermediates in the proposed degradation pathway of polyolefins. However, further effort is still needed to prove the exact mechanism of these enzymes. In addition to the above-mentioned enzymes, intracellular enzymes including alcohol dehydrogenase, aldehyde dehydrogenase, Baeyer–Villiger monooxygenase, and esterase were also reported to be involved in polyolefins biodegradation to alcohol, aldehydes, ketones, and aliphatic acids, which can then be further metabolized by bacteria via the β-oxidation pathway and subsequently enter the citric acid cycle [52,111].
Recently, a few publications have reported new innovative bioprocesses for PE biodegradation. Peixoto et al. (2022) reported new evidence on microbial oxidation of PE by nitric oxide (NO) produced from nitrifier and denitrifying bacteria. Three genera, including Comamonas sp., Delftia sp., and Stenotrophomonas sp. can oxidize ammonia (NH4) to nitrite (NO2), releasing NO as an intermediate, which oxidizes PE through the introduction of a nitro group into the PE chain (the high peak at 1550 cm−1 detected by FTIR) [112]. Furthermore, a latex clearing protein (LcpK30) from Streptomyces sp. strain K30 was reported to fragment UV-pretreated PE and PP, by adding -OH and -C=O functional groups to the polyolefins’ backbone. The Mw of PE was reduced by 42%, and ketones were detected by GC-MS as degradation products [38].
Several other microorganisms have been identified during PE biodegradation, including, for example, bacteria (Bacillus, Pseudomonas, Ralstonia, Rhodococcus, Staphylococcus, Streptomyces and Stenotrophomonas), fungi (Aspergillus, Cladosporium, Lentinus, Phanerochaete and Penicillium), and gut microbiomes of mealworms and waxworms [113,114,115,116,117,118]. The very recent publication by Sanluis-Verdes and team reported the first discovery of animal enzymes, belonging to the phenol oxidase family, obtained from saliva of waxworms’ larvae (Galleria mellonella) that are capable of oxidizing and depolymerizing PE. The authors reported an increase in the carbonyl index and changes in the average molecular weight of PE, together with the release of oxidized aliphatic ketones as degradation products after saliva contact [39]. A process utilizing different microorganisms and enzymes was recently patented by the Technical Research Centre of Finland VTT, that included (among other things) the use of Bacillus licheniformis, Bacillus flexus, Bacillus subtilis, and Bacillus cereus’ metal-dependent hydrolases [119]. Nevertheless, in contrast with polymers made of ester bond linkages, major advancements in enzyme discovery and optimization are still needed to reach a breakthrough in the case of polyolefins.
Consequently, as biodegradation/depolymerization of PE is still being developed, upcycling of PE has been so far mainly reported as a combination of thermochemical and biological process. Different studies have tried to develop processes for bioupcycling PE waste into biodegradable bioplastic PHA (Table 3). Guzik and colleagues, for instance, examined a two-step chemo-biotechnology conversion. Firstly, PE was pyrolyzed to a mixture of hydrocarbons (C8–C32), called PE hydrolysis wax, which is composed of 90% alkanes and 10% alkenes. The PE pyrolysis wax was then used as a sole carbon source to grow PHA-accumulating strains. Next, the PHA production was enhanced by changing the nitrogen source from NH4Cl to NH4NO3 and adding a biosurfactant (rhamnolipid). The authors discussed that rhamnolipids enhanced the uptake of aliphatic hydrocarbon. The highest PHA was produced from Pseudomonas aeruginosa GL-1 with 18.9% of CDW (0.074 g/L) [120].
Another study performed oxidative degradation of PE at 145 °C to obtain an oxidized polyethylene wax (O-PEW), which was used as a novel carbon source. Ralstonia eutropha H16 was cultured in tryptone soya broth (TSB) with 4 g/L O-PEW to produce PHA. The results showed 1.24 g/L PHA production after 48 h, accounting for 33.8% CDW [121]. Researchers also investigated the use of non-oxidized PE wax (N-PEW), which is cheaper and easier to produce than O-PEW [122]. Not surprisingly, a lower PHA production was found, with 0.46 g/L (32% CDW). The presence of less recalcitrant carbon sources in O-PEW, such as fatty and carboxylic acids, could be the reason that makes it more accessible to bacteria. There are some studies investigating the direct use of untreated PE to produce PHA by different strains, such as Cuprividus necator H16, Pseudomonas putida LS46, and Acinetobacter pittii IRN19; however, the reported PHA yields are very low [109]. To date, depolymerization of PE by thermal processes seems to be one of the most effective strategies to allow for efficient (subsequent) microbial usage. To open the window for efficient polyolefins bioupcycling technologies, the development of cheap pretreatments and more efficient biodegradation processes are still needed.
Table 3. PE bioupcycling.
Table 3. PE bioupcycling.
Depolymerization StrategyDepolymerization Products
Used as a Feedstock for
Fermentation Step
Fermentation StrategyProducts from FermentationTiterProductivityYieldRef.
PyrolysisPE hydrolysis wax (a mixture of hydrocarbons (C8–C32): 90% alkanes and 10% alkenes)Fermentation in shake flask containing 0.05% (w/v) PE pyrolysis wax as a sole carbon source and 0.025% (w/v) of NH4Cl as a nitrogen source at 30 °C for 48 h by Pseudomonas aeruginosa GL-1PHA0.023 g/LNA0.10 gPHA/gCDW[120]
PyrolysisPE hydrolysis wax (a mixture of hydrocarbons (C8–C32): 90% alkanes and 10% alkenes)Fermentation in shake flask containing 2% (w/v) PE pyrolysis wax as a sole carbon source and 0.019% (w/v) of NH4NO3 as a nitrogen source at 30 °C for 48 h by P. aeruginosa GL-1, in the presence of 0.05% (w/v) rhamnolipidsPHA0.074 g/LNA0.19 gPHA/gCDW[120]
PyrolysisPE hydrolysis wax (a mixture of hydrocarbons (C8–C32): 90% alkanes and 10% alkenes)Fermentation in shake flask containing 2% (w/v) PE pyrolysis wax as a sole carbon source and 0.019% (w/v) of NH4NO3 as a nitrogen source at 30 °C for 48 h by P. aeruginosa PAO1, in the presence of 0.05% (w/v) rhamnolipidsPHA0.045 g/LNA0.15 gPHA/gCDW[120]
Oxidative degradation in a two-phase system (gas-liquid phase), after melting at 145 °C and using oxygenOxidized polyethylene wax (O-PEW)Fermentation in shake flask containing 4 g/L melted O-PEW emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by Ralstonia eutropha H16PHA1.25 g/LNA0.34 gPHA/gCDW[121]
--Fermentation in shake flask containing Ramsey’s media with 1% LDPE particles at 30 °C and 150 rpm for 21 d by Cuprividus necator H16short chain length PHA (sclPHA)NANA0.0318 gPHA/gCDW[123]
--Fermentation in shake flask containing Ramsey’s media with 1% LDPE particles at 30 °C and 150 rpm for 21 d by Pseudomonas putida LS46mclPHANANA0.0054 gPHA/gCDW[123]
--Fermentation in shake flask containing Ramsey’s media with 1% LDPE particles at 30 °C and 150 rpm for 21 d by Acinetobacter pittii IRN19mclPHANANA0.0049 gPHA/gCDW[123]
PyrolysisNon-oxidized PE wax (N-PEW)Fermentation in shake flask containing 4 g/L melted N-PEW emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by Cupriavidus necator H16PHA0.46 g/LNA0.32 gPHA/gCDW[122]

3.3.2. Polypropylene (PP)

PP is the second most produced plastic after PE, contributing to 16.6% of global plastic [1]. It is used for various applications, mainly food packaging, automobile, and the textile industry. The large production of PP, its recalcitrance, and the short application lifespan creates a large amount of post-consumption waste, posing a threat to the environment.
PP is categorized as a polyolefin, along with PE, that has hydrophobic backbones composed of long carbon chains and high molecular weight [52]. The enzyme machinery required to degrade PP is similar to PE; however, the presence of a methyl side-chain in PP makes it even more resistant to microbial or enzymatic depolymerization [124]. Similar to other polyolefins, increasing PP hydrophilicity by oxidation should, in principle, enable microbial colonization on its surface and initiate degradation through the extracellular enzymes.
PP biodegradation has been reported on a range of screened microorganisms. Jeon and Kim isolated a PP degrading strain from the municipal solid waste stack field and identified it as Stenotrophomonas panacihumi PA3-2. The strain degraded 20.3% and 12.7% of low- and high-molecular-weight PP, respectively, at 37 °C and a period of 90 days [124]. The single strain Enterobacter sp nov. bt DSCE01, Enterobacter cloacae nov. bt DSCE02, and Pseudomonas aeruginosa nov. bt DSCE-CD03 isolated from cow dung also showed similar degradation of PP and LDPE (less than 20%) in the study of Skariyachan et al. [125]. However, when they were grown together as a consortium, 63% and 64% degradation of PP and LDPE were found, respectively. A similar result was reported by using thermophilic consortia of Aneurinibacillus aneurinilyticus, Brevibacillus agri, Brevibacillus sp. and Brevibacillus brevis, which degraded 56% of PP [126]. Further PP biodegradation advances were reported using endophytic fungi [99] and gut microbiota of invertebrate larvae [127]. So far, a superior biodegradation efficiency was obtained by using mixed culture. Skariyachan et al. reported the formation of aldehyde and methyl groups, as well as of cis-2-chlorovinyl acetate, tri-decanoic acid, and octa-decanoic acid products from polyolefins biodegradation [126]. Such observations are aligned with the proposed metabolism of polyolefins, which is expected to start with an oxidation that leads to the formation of carbonyl groups, followed by a further oxidation to aldehyde or ketone, and finally to the formation of carboxylic acids [128].
As already seen in PE, the developed upcycling strategies for PP rely on high temperature depolymerization steps (Table 4). Johnston et al. employed thermal oxidation in the presence of oxygen-ozone as a first step to oxidize PP. Then, the thermal-oxidized PP was emulsified in the TSB media before being used as a feedstock for PHA production by Cupriavidus necator H16, in shake flasks. After 48 h, 1.36 g/L of PHA was accumulated, accounting for 42% of CDW [129]. Another research group developed a combined process using pyrolysis (540 °C) and yeast fermentation (with Yarrowia lipolytica), to upcycle PP into fatty acids (C18 and C16) [130,131].

3.4. Bioupcycling of Polystyrene (PS)

PS reached 6.1% of European plastic demand in 2020 [3] and 4.8% worldwide in 2019 [1]. Thanks to its structural stability, it has been used for many purposes ranging from food packaging, household gadgets, electrical devices, and building insulation. The main building block of PS is styrene, which can be combined with additives, colorants, and/or other plastics [15]. As plastic waste has become a pollutant, contaminating landfills and the marine environment, it should be mentioned that one-third of landfilling plastic is PS [132]. Styrene monomers leaching from PS are also very dangerous, being considered carcinogenic in humans [133].
Nonetheless, biodegradation approaches have been studied also for PS. Since it is also a thermoplastic that resists hydrolysis, PS needs a preliminary oxidation process to generate hydrolyzable functional groups. Kim et al. confirmed that carbonyl groups were produced during PS degradation by Pseudomonas aeruginosa strain DSM 50071 (isolated from the gut of the superworms Zophobas atratus), associated with the enzymes serine hydrolase and S-formylglutathione hydrolase [134]. Ho et al. gathered further information on microbial degradation of PS, identifying relevant fungi, bacteria, and archaea strains [135], and several other studies have isolated PS degrading strains from various sources, such as marine environments [136,137] and gut microbiota of invertebrate larvae [134,138,139].
A previous attempt to turn PS waste into valuable products was performed by Savoldelli and colleagues [132]. They used the combination of thermal (240 °C) and bacterial (Pseudomonas putida) degradation to break down PS to naphthalene derivatives, benzene derivatives, as well as some styrene; however, this study did not show the full upcycling of PS. Bioupcycling of PS was, however, successfully performed through another combined approach, using pyrolysis followed by bacterial conversion (Table 5). The fermentation of the styrene pyrolysis oil in a 7.5 L stirred tank reactor accumulated 57% PHA of CDW with a yield of around 0.1 gPHA/g styrene oil (10% conversion). The PHA was characterized as mclPHA with the monomers (R)-3-hydroxyhexanoate, (R)-3-hydroxyoctanoate and (R)-3-hydroxydecanoate [140]. An increase in PHA titer and yield was found when styrene oil after pyrolysis was distilled before being used as a substrate for fermentation [141].
Thermal oxidation with a pro-oxidant/pro-degradant additive of PS coupled with microbial fermentation was also reported to upcycle post-consumer PS to PHA. PS was subjected to thermal oxidation for 20 h at 60 °C (with a flow rate of the oxygen-ozone mixture of 7.5 L/h) and then used as the feedstock for PHA fermentation by Cupriavidus necator H16, cultured in tryptone soya broth (TSB). The results showed PHA accumulation at 48% of CDW, accounting for 1.72 g/L. The major subunit was 3-hydroxybutyrate, with up to 12 mol% of 3-hydroxyvalerate and 3- hydroxyhexanoate co-monomeric units [15].

3.5. Bioupcycling of Polyvinyl Chloride (PVC)

PVC has low biodegradability, but possesses the highest proportion of plasticizer among other synthetic plastics, which is susceptible to microbial attack [144]. The study by Giacomucci and colleagues investigated the ability of five microbial strains (Pseudomonas chlororaphis (DSM 50083), Pseudomonas citronellolis (DSM 50332), Bacillus subtilis subsp. spizizenii (DSM 15029), Bacillus flexus (DSM 1320), and Chelatococcus daeguensis (DSM 22069)) to biodegrade plastic, showing that a microbial biofilm was found after 3-month incubation of PVC containing 30% (w/w) of plasticizers, while no biofilm formation was detected on PE, PP, and PS during the same period [145]. Tenebrio molitor larvae were reported to reduce weight-, number- and size-average molecular weights (Mw, Mn and Mz) of PVC by 33.4%, 32.8% and 36.4%, respectively, in 16 days. The PVC depolymerization was found to be a gut microbe-dependent biodegradation associated with four families: Streptococcaceae, Spiroplasmataceae, Enterobacteriaceae, and Clostridiaceae [146]. Another study investigated the anaerobic biodegradation of PVC by marine consortia and reported gravimetric weight losses of 12% over 7 months [147]. PVC biodegradation by fungi was reported by Ali et al., showing that Phanerochaete chrysosporium PV1 caused a significant reduction in the molecular weight of PVC and appearance and shifting of FTIR peaks at the 2370–2350 cm−1 region, corresponding to HC-Cl [148].
Also in this case, thermochemical methods proved to be a valuable approach to upcycle PVC into valuable products. For example, a one-pot process including a sequence of dechlorination by Cl-fixative (ZnO or KOH), carbonization of dechlorinated polyenes and further modifications, was used to upcycle PVC into valuable carbon materials, pipeline-quality pyrolysis gas, and chlorides [149]. However, to our knowledge, bioupcycling of PVC, using only bioprocesses, has not been reported so far. The pyrolysis products from PVC, especially those with removed chlorines [150], could be the interesting substrate for further bioconversion steps, such as in the case of recalcitrant polyolefins.

3.6. Bioupcycling of Mixed Plastic Waste

The ability to upcycle unsorted, mixed plastic waste would represent a key technology to address real post-consumer plastic streams [151]. To the best of our knowledge, bioupcycling of real mixed plastic waste has hardly been addressed so far. The most promising technology going in this direction was recently proposed by Sullivan and colleagues, that valorized mixed plastic waste (obtained by mixing HDPE, polystyrene and PET), through chemical oxidation coupled to biological funneling [152]. Basically, the polymers were treated with metal-promoted autoxidation that deconstructed them into a mixture of oxygenated intermediates. The latter have an enhanced water solubility, which facilitates their use as a feedstock for the following bioupcycling step, by an engineered Pseudomonas putida strain. Besides this study, the selective recovery of monomers in mixed plastic stream has also been proposed, by using enzymes cocktails and microbial mixed consortia processes [151]. The two-enzyme PETase/MHETase system was reported to improve PET and MHET conversion to their monomers, and was highlighted as a model for the future development of multienzyme systems for depolymerization of mixed polymer wastes [60]. Another study by Edwards and team investigated the microbial community-based degradation of diverse carbon sources. The consortium of 2 Bacillus and 3 Pseudomonas strains was able to degrade PET and also utilize putrescine, an alkane substrate, as a carbon source indicating the potential ability to degrade polyolefins. The strains were also capable of degrading common plasticizers, including phthalates, paraben, and other aromatic and phenolic compounds [153]. This result presents a possibility of using defined mixed consortia for bio-recycling and upcycling of mixed plastic stream.

4. Future Perspectives and Conclusions

Plastic is a cheap and ubiquitous material in our society that has been used for multiple functions and applications, to the point to be considered a major characteristic of the current period (the so-called “plastic age”). Even if plastics have countless benefits, they also create a complex waste problem on a global scale. The current linear plastic lifecycle causes environmental pollution, and the negative effects are becoming more and more visible, indicating a non-sustainable value chain. Moreover, the way that plastics are produced, used, and discarded leads to a continuous loss of material from the loop and fails to comply with the circular economy approach. To tackle this problem, the EU has established the waste framework directive, introducing a waste five-step hierarchy that sets the order of preference for managing and disposing of waste (Figure 5) [154]. Preventing unnecessary plastic usage, reducing whenever possible single-use plastics, and reusing plastic materials as much as possible, are the preferred options (higher priority in the hierarchy) to lower amount of plastic waste generated and the carbon footprint. Landfilling is the least desirable, as it causes the loss of material and often leads to the contamination of the environment. The EU commission is thus pushing toward “end-of-waste” or “zero waste” criteria, where new technologies, policies and supply chains will allow us to consider the waste streams as a secondary raw material that should be recovered, reused, or recycled at the highest possible level. Currently, the mechanical recycling is the dominant technology, but, as already mentioned, it does not involve the higher level of the waste hierarchy (reduce and reuse; Figure 5), and it fails to handle contaminated/mixed/multilayered plastics. Moreover, it inevitably leads to downcycling, already after a limited number of cycles. In this sense, bioupcycling can contribute better to the circular economy, by converting post-consumer plastic via (integrated) biotechnological processes to more renewable and carbon-neutral high-value chemicals and new polymers, or at least to recycle the monomers from biodepolymerization to the same type of plastic with preserved quality.
The conversion of fossil-based plastic waste to bioplastics also brings a new opportunity in terms of advanced/add-on properties, such as the improved barrier properties of PEF compared to PET [13]. It also opens the window towards new/innovative materials with different properties from conventional plastics, such as functionalized plastics or the new poly(amide urethane) (bio-PU) obtained from upcycled PET waste (to HAA) [16]. Moreover, the new materials can be developed to be more readily recyclable than the parent polymers, to align with the ambitious goal of Europe’s new plastic strategy that requires 55% of plastic packaging being recyclable by 2030 [4]. A recent study by Roux and Varrone showed that bioupcycling of fossil-derived PET to bio-PTT and bio-PEF can represent an important drive towards plastic waste valorization and increased recycling rates [12]. In fact, the techno-economic analysis showed that recycling PEF and PTT to the required 55% would decrease their production cost by 50%, obtaining a minimum selling price of 1.61 USD/Kg, which is lower than rPET prices [47]. In this consideration, bioupcycling is expected to play an important role towards a smarter, more innovative, and more sustainable plastic sector.
Due to the incredible complexity and diversity of plastic waste, the future success of upcycling approaches will ultimately depend on the proper combination and integration of different complementary technologies and processes. In this sense, cross-disciplinarity will be a key enabler. A good example is the sequential process of chemical depolymerization of PET, followed by biological conversion of PET hydrolysates to higher values building blocks. Chemocatalytic glycolysis produces, for instance, BHET, which can be bioconverted by engineered P. putida to β-ketoadipic acid, a valuable monomer for the production of a performance-advantaged nylon-6,6 analog polymer [73]. Other chemical depolymerization strategies (e.g., methanolysis to produce dimethyl terephthalate, aminolysis to produce terephthalamide, or alcoholysis to produce terephthalate diesters [155,156]) could be combined with biotechnological upcycling as well. The incredible diversity of metabolic pathways and enzymes can be exploited to develop new bioprocesses to valorize the complex depolymerization broths obtained from thermochemical processes. New high-value products can be obtained from specific hydrolysates (without previous purification), by the use of consolidated bioprocesses and synthetic biology. Moreover, while thermochemical processes can boost the (bio)degradation and upcycling steps, biological processes (such as enzyme technology) might remove specific unwanted compounds and impurities to increase overall thermochemical conversion yields (i.e., removing PA before pyrolysis to avoid formation of cyanide gasses and costly downstream gas cleaning), thus overcoming mutual limitations.
According to the plastics microbial biodegradation database (PMBD), 949 microorganism–plastic relationships and 79 genes involved in the biodegradation of plastics have been reported [157]. There is no doubt that these numbers are going to dramatically increase in the coming years thanks to advanced metagenomic and proteomic approaches, new modeling tools, new synthetic biology, and protein engineering techniques [158]. The development of new high-throughput screening methods will be paramount. The constant discovery and engineering of microbes and enzymes that can selectively degrade certain types of polymers will enable more effective recycling of dirty mixed plastic waste fractions that currently are incinerated or go to landfill. An excellent example is the interesting coupled bioleaching/enzyme-based process for the recycling of multilayer packaging that was recently developed by Kremser and colleagues [159]. The study successfully demonstrated an innovative method to recover pure PE and aluminum hydroxide from beverage cartons. The achievement of this study opens for the biological recycling of multilayer materials. All these findings anticipate the possibility of developing novel biocatalysts and bioprocesses that will allow for cost-effective and scalable plastic bioupcycling.
In conclusion, it is important to underline that only an extra effort on assessing economic feasibility and environmental impacts of such coupled biological processes will illustrate their viability and contribution to a circular economy.

Author Contributions

Conceptualization, P.L. and C.V.; writing—original draft preparation, P.L.; writing—review and editing, P.L. and C.V.; visualization, P.L.; supervision, C.V.; project administration, C.V.; funding acquisition, C.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the H2020 UPLIFT project (Grant Agreement no. 953073).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank the European Commission, under H2020 Grant Agreement no. 953073 (acronym: UPLIFT), and the Department of Chemistry and Bioscience, Aalborg University for the financial support of this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. OECD. Global Plastics Outlook: Economic Drivers, Environmental Impacts and Policy Options; OECD Publishing: Paris, France, 2022; ISBN 9789264584068. [Google Scholar]
  2. US EPA Plastics: Material-Specific Data. Available online: https://www.epa.gov/facts-and-figures-about-materials-waste-and-recycling/plastics-material-specific-data (accessed on 6 March 2022).
  3. PlasticsEurope The Facts 2021. Available online: https://plasticseurope.org/knowledge-hub/plastics-the-facts-2021/ (accessed on 16 February 2022).
  4. European Commission. A European Strategy for Plastics in a circular economy. Eur. Com. 2018. [Google Scholar]
  5. Hamilton, L.A.; Feit, S.; Muffett, C.; Kelso, M.; Rubright, S.M.; Bernhardt, C.; Schaeffer, E.; Moon, D.; Morris, J.; Labbé-Bellas, R. Plastic & Climate the Hidden Costs of a Plastic Planet; Center for International Environmental Law (CIEL): Washington, DC, USA, 2019. [Google Scholar]
  6. The Ocean Conference Factsheet: Marine Pollution. Available online: https://sustainabledevelopment.un.org/content/documents/Ocean_Factsheet_Pollution.pdf (accessed on 6 March 2022).
  7. Liebmann, B.; Köppel, S.; Reiberger, T.; Schwab, P. Assessment of Microplastic Concentrations in Human Stool-Preliminary Results of a Prospective Study. In Proceedings of the United European Gastroenterology (UEG) Week, Vienna, Austria, 20–24 October 2018; pp. 1–16. [Google Scholar]
  8. Sharma, S.; Chatterjee, S. Microplastic pollution, a threat to marine ecosystem and human health: A short review. Environ. Sci. Pollut. Res. 2017, 24, 21530–21547. [Google Scholar] [CrossRef] [PubMed]
  9. Geyer, R.; Jambeck, J.R.; Law, K.L. Production, use, and fate of all plastics ever made. Sci. Adv. 2017, 3, e1700782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Lebreton, L.; Egger, M.; Slat, B. A global mass budget for positively buoyant macroplastic debris in the ocean. Sci. Rep. 2019, 9, 12922. [Google Scholar] [CrossRef] [Green Version]
  11. European Commission. Circular Economy Action Plan. Available online: https://ec.europa.eu/environment/strategy/circular-economy-action-plan_en (accessed on 7 March 2022).
  12. Roux, M.; Varrone, C. Assessing the Economic Viability of the Plastic Biorefinery Concept and Its Contribution to a More Circular Plastic Sector. Polymer 2021, 13, 3883. [Google Scholar] [CrossRef]
  13. Lomwongsopon, P.; Varrone, C. Contribution of Fermentation Technology to Building Blocks for Renewable Plastics. Fermentation 2022, 8, 47. [Google Scholar] [CrossRef]
  14. Wierckx, N.; Prieto, M.A.; Pomposiello, P.; de Lorenzo, V.; O’Connor, K.; Blank, L.M. Plastic waste as a novel substrate for industrial biotechnology. Microb. Biotechnol. 2015, 8, 900–903. [Google Scholar] [CrossRef]
  15. Johnston, B.; Radecka, I.; Hill, D.; Chiellini, E.; Ilieva, V.; Sikorska, W.; Musioł, M.; Zięba, M.; Marek, A.; Keddie, D.; et al. The Microbial Production of Polyhydroxyalkanoates from Waste Polystyrene Fragments Attained Using Oxidative Degradation. Polymers 2018, 10, 957. [Google Scholar] [CrossRef] [Green Version]
  16. Tiso, T.; Narancic, T.; Wei, R.; Pollet, E.; Beagan, N.; Schröder, K.; Honak, A.; Jiang, M.; Kenny, S.T.; Wierckx, N.; et al. Towards bio-upcycling of polyethylene terephthalate. Metab. Eng. 2021, 66, 167–178. [Google Scholar] [CrossRef]
  17. Narancic, T.; Salvador, M.; Hughes, G.M.; Beagan, N.; Abdulmutalib, U.; Kenny, S.T.; Wu, H.; Saccomanno, M.; Um, J.; O’Connor, K.E.; et al. Genome analysis of the metabolically versatile Pseudomonas umsongensis GO16: The genetic basis for PET monomer upcycling into polyhydroxyalkanoates. Microb. Biotechnol. 2021, 14, 2463–2480. [Google Scholar] [CrossRef]
  18. Tokiwa, Y.; Calabia, B.P.; Ugwu, C.U.; Aiba, S. Biodegradability of Plastics. Int. J. Mol. Sci. 2009, 10, 3722–3742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Pan, D.; Su, F.; Liu, C.; Guo, Z. Research progress for plastic waste management and manufacture of value-added products. Adv. Compos. Hybrid Mater. 2020, 3, 443–461. [Google Scholar] [CrossRef]
  20. Babaremu, K.O.; Okoya, S.A.; Hughes, E.; Tijani, B.; Teidi, D.; Akpan, A.; Igwe, J.; Karera, S.; Oyinlola, M.; Akinlabi, E.T. Sustainable plastic waste management in a circular economy. Heliyon 2022, 8, e09984. [Google Scholar] [CrossRef] [PubMed]
  21. Idumah, C.I.; Nwuzor, I.C. Novel trends in plastic waste management. SN Appl. Sci. 2019, 1, 1402. [Google Scholar] [CrossRef] [Green Version]
  22. Ragaert, K.; Delva, L.; Geem, K.V. Mechanical and chemical recycling of solid plastic waste. Waste Manag. 2017, 69, 24–58. [Google Scholar] [CrossRef]
  23. Gala, A.; Guerrero, M.; Serra, J.M. Characterization of post-consumer plastic film waste from mixed MSW in Spain: A key point for the successful implementation of sustainable plastic waste management strategies. Waste Manag. 2020, 111, 22–33. [Google Scholar] [CrossRef]
  24. Hahladakis, J.N.; Velis, C.A.; Weber, R.; Iacovidou, E.; Purnell, P. An overview of chemical additives present in plastics: Migration, release, fate and environmental impact during their use, disposal and recycling. J. Hazard. Mater. 2018, 344, 179–199. [Google Scholar] [CrossRef]
  25. Hatti-Kaul, R.; Nilsson, L.J.; Zhang, B.; Rehnberg, N.; Lundmark, S. Designing Biobased Recyclable Polymers for Plastics. Trends Biotechnol. 2020, 38, 50–67. [Google Scholar] [CrossRef]
  26. Abu-Thabit, N.Y.; Pérez-Rivero, C.; Uwaezuoke, O.J.; Ngwuluka, N.C. From waste to wealth: Upcycling of plastic and lignocellulosic wastes to PHAs. J. Chem. Technol. Biotechnol. 2021, 97, 3217–3240. [Google Scholar] [CrossRef]
  27. Jamil, N.; Kumar, P.; Batool, R. Soil Microenvironment for Bioremediation and Polymer Production; John Wiley & Sons: Hoboken, NJ, USA, 2019; ISBN 9781119592129. [Google Scholar]
  28. Soni, V.K.; Singh, G.; Vijayan, B.K.; Chopra, A.; Kapur, G.S.; Ramakumar, S.S.V. Thermochemical Recycling of Waste Plastics by Pyrolysis: A Review. Energy Fuels 2021, 35, 12763–12808. [Google Scholar] [CrossRef]
  29. Thiounn, T.; Smith, R.C. Advances and approaches for chemical recycling of plastic waste. J. Polym. Sci. 2020, 58, 1347–1364. [Google Scholar] [CrossRef]
  30. Enthaler, S. Illustrating Plastic Production and End-of-Life Plastic Treatment with Interlocking Building Blocks. J. Chem. Educ. 2017, 94, 1746–1751. [Google Scholar] [CrossRef]
  31. Fox, J.A.; Stacey, N.T. Process targeting: An energy based comparison of waste plastic processing technologies. Energy 2019, 170, 273–283. [Google Scholar] [CrossRef]
  32. Tournier, V.; Topham, C.M.; Gilles, A.; David, B.; Folgoas, C.; Moya-Leclair, E.; Kamionka, E.; Desrousseaux, M.L.; Texier, H.; Gavalda, S.; et al. An engineered PET depolymerase to break down and recycle plastic bottles. Nature 2020, 580, 216–219. [Google Scholar] [CrossRef]
  33. Vollmer, I.; Jenks, M.J.F.; Roelands, M.C.P.; White, R.J.; van Harmelen, T.; de Wild, P.; van der Laan, G.P.; Meirer, F.; Keurentjes, J.T.F.; Weckhuysen, B.M. Beyond Mechanical Recycling: Giving New Life to Plastic Waste. Angew. Chem. Int. Ed. 2020, 59, 15402–15423. [Google Scholar] [CrossRef] [Green Version]
  34. Son, H.F.; Cho, I.J.; Joo, S.; Seo, H.; Sagong, H.Y.; Choi, S.Y.; Lee, S.Y.; Kim, K.J. Rational Protein Engineering of Thermo-Stable PETase from Ideonella sakaiensis for Highly Efficient PET Degradation. ACS Catal. 2019, 9, 3519–3526. [Google Scholar] [CrossRef]
  35. Carniel, A.; Valoni, É.; Nicomedes, J.; Gomes, A.d.C.; de Castro, A.M. Lipase from Candida antarctica (CALB) and cutinase from Humicola insolens act synergistically for PET hydrolysis to terephthalic acid. Proc. Biochem. 2017, 59, 84–90. [Google Scholar] [CrossRef]
  36. Furukawa, M.; Kawakami, N.; Tomizawa, A.; Miyamoto, K. Efficient Degradation of Poly(ethylene terephthalate) with Thermobifida fusca Cutinase Exhibiting Improved Catalytic Activity Generated using Mutagenesis and Additive-based Approaches. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef] [Green Version]
  37. Yao, C.; Xia, W.; Dou, M.; Du, Y.; Wu, J. Oxidative degradation of UV-irradiated polyethylene by laccase-mediator system. J. Hazard. Mater. 2022, 440, 129709. [Google Scholar] [CrossRef]
  38. Zhang, H.; Kong, D.; Wang, L.; Xia, W.; Yao, C.; Wu, J. Degradation of UV-pretreated polyolefins by latex clearing protein from Streptomyces sp. Strain K30. Sci. Total Environ. 2022, 806, 150779. [Google Scholar] [CrossRef]
  39. Sanluis-Verdes, A.; Colomer-Vidal, P.; Rodríguez-Ventura, F.; Bello-Villarino, M.; Spinola-Amilibia, M.; Ruiz-López, E.; Illanes-Vicioso, R.; Castroviejo, P.; Cigliano, R.A.; Montoya, M.; et al. Wax worm saliva and the enzymes therein are the key to polyethylene degradation by Galleria mellonella. bioRxiv 2022, 13, 5568. [Google Scholar] [CrossRef] [PubMed]
  40. Jeon, H.J.; Kim, M.N. Functional analysis of alkane hydroxylase system derived from Pseudomonas aeruginosa E7 for low molecular weight polyethylene biodegradation. Int. Biodeterior. Biodegrad. 2015, 103, 141–146. [Google Scholar] [CrossRef]
  41. Yang, Y.; Yang, J.; Wu, W.M.; Zhao, J.; Song, Y.; Gao, L.; Yang, R.; Jiang, L. Biodegradation and Mineralization of Polystyrene by Plastic-Eating Mealworms: Part 1. Chemical and Physical Characterization and Isotopic Tests. Environ. Sci. Technol. 2015, 49, 12080–12086. [Google Scholar] [CrossRef] [PubMed]
  42. Shi, L.; Liu, H.; Gao, S.; Weng, Y.; Zhu, L. Enhanced extracellular production of IsPETase in Escherichia coli via engineering of the pelB signal peptide. J. Agric. Food Chem. 2021, 69, 2245–2252. [Google Scholar] [CrossRef]
  43. Rennison, A.; Winther, J.R.; Varrone, C. Rational Protein Engineering to Increase the Activity and Stability of IsPETase Using the PROSS Algorithm. Polymer 2021, 13, 3884. [Google Scholar] [CrossRef]
  44. Nakamura, A.; Kobayashi, N.; Koga, N.; Iino, R. Positive charge introduction on the surface of thermostabilized PET hydrolase facilitates PET binding and degradation. ACS Catal. 2021, 11, 8550–8564. [Google Scholar] [CrossRef]
  45. Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.; Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K. A bacterium that degrades and assimilates poly(ethylene terephthalate). Science 2016, 351, 1196–1199. [Google Scholar] [CrossRef]
  46. Roth, C.; Wei, R.; Oeser, T.; Then, J.; Föllner, C.; Zimmermann, W.; Sträter, N. Structural and functional studies on a thermostable polyethylene terephthalate degrading hydrolase from Thermobifida fusca. Appl. Microbiol. Biotechnol. 2014, 98, 7815–7823. [Google Scholar] [CrossRef]
  47. Singh, A.; Rorrer, N.A.; Nicholson, S.R.; Erickson, E.; DesVeaux, J.S.; Avelino, A.F.T.; Lamers, P.; Bhatt, A.; Zhang, Y.; Avery, G.; et al. Techno-economic, life-cycle, and socioeconomic impact analysis of enzymatic recycling of poly(ethylene terephthalate). Joule 2021, 5, 2479–2503. [Google Scholar] [CrossRef]
  48. Ellis, L.D.; Rorrer, N.A.; Sullivan, K.P.; Otto, M.; McGeehan, J.E.; Román-Leshkov, Y.; Wierckx, N.; Beckham, G.T. Chemical and biological catalysis for plastics recycling and upcycling. Nat. Catal. 2021, 4, 539–556. [Google Scholar] [CrossRef]
  49. Jenkins, S.; Quer, A.M.I.; Fonseca, C.; Varrone, C. Microbial Degradation of Plastics: New Plastic Degraders, Mixed Cultures and Engineering Strategies. In Soil Microenvironment for Bioremediation and Polymer Production; Jamil, N., Kumar, P., Batool, R., Eds.; Wiley Online Book: Hoboken, NJ, USA, 2019; pp. 215–238. ISBN 9781119592129. [Google Scholar]
  50. Arkatkar, A.; Arutchelvi, J.; Bhaduri, S.; Uppara, P.V.; Doble, M. Degradation of unpretreated and thermally pretreated polypropylene by soil consortia. Int. Biodeterior. Biodegrad. 2009, 63, 106–111. [Google Scholar] [CrossRef]
  51. Yoon, M.G.; Jeon, H.J.; Kim, M.N. Biodegradation of Polyethylene by a Soil Bacterium and AlkB Cloned Recombinant Cell. J. Bioremediat. Biodegrad. 2012, 3, 1–8. [Google Scholar] [CrossRef] [Green Version]
  52. Mohanan, N.; Montazer, Z.; Sharma, P.K.; Levin, D.B. Microbial and Enzymatic Degradation of Synthetic Plastics. Front. Microbiol. 2020, 11, 2837. [Google Scholar] [CrossRef] [PubMed]
  53. Tiseo, I. PET Demand Worldwide 2030. Available online: https://www.statista.com/statistics/1128658/polyethylene-terephthalate-demand-worldwide/ (accessed on 6 October 2022).
  54. Zero Waste Europe New Report: PET is Far from Real Circularity. Available online: https://www.recycling-magazine.com/2022/02/16/new-report-pet-is-far-from-real-circularity/ (accessed on 16 February 2022).
  55. Hou, Q.; Zhen, M.; Qian, H.; Nie, Y.; Bai, X.; Xia, T.; Rehman, M.L.U.; Li, Q.; Ju, M. Upcycling and catalytic degradation of plastic wastes. Cell Rep. Phys. Sci. 2021, 2, 100514. [Google Scholar] [CrossRef]
  56. Chen, C.C.; Han, X.; Li, X.; Jiang, P.; Niu, D.; Ma, L.; Liu, W.; Li, S.; Qu, Y.; Hu, H.; et al. General features to enhance enzymatic activity of poly(ethylene terephthalate) hydrolysis. Nat. Catal. 2021, 4, 425–430. [Google Scholar] [CrossRef]
  57. Brott, S.; Pfaff, L.; Schuricht, J.; Schwarz, J.N.; Böttcher, D.; Badenhorst, C.P.S.; Wei, R.; Bornscheuer, U.T. Engineering and evaluation of thermostable IsPETase variants for PET degradation. Eng. Life Sci. 2022, 22, 192–203. [Google Scholar] [CrossRef]
  58. Cui, L.; Qiu, Y.; Liang, Y.; Du, C.; Dong, W.; Cheng, C.; He, B. Excretory expression of IsPETase in E. coli by an enhancer of signal peptides and enhanced PET hydrolysis. Int. J. Biol. Macromol. 2021, 188, 568–575. [Google Scholar] [CrossRef]
  59. Ferrario, V.; Todea, A.; Wolansky, L.; Piovesan, N.; Guarneri, A.; Ribitsch, D.; Guebitz, G.M.; Gardossi, L.; Pellis, A. Effect of Binding Modules Fused to Cutinase on the Enzymatic Synthesis of Polyesters. Catalyst 2022, 12, 303. [Google Scholar] [CrossRef]
  60. Knott, B.C.; Erickson, E.; Allen, M.D.; Gado, J.E.; Graham, R.; Kearns, F.L.; Pardo, I.; Topuzlu, E.; Anderson, J.J.; Austin, H.P.; et al. Characterization and engineering of a two-enzyme system for plastics depolymerization. Proc. Natl. Acad. Sci. USA 2020, 117, 25476–25485. [Google Scholar] [CrossRef]
  61. Sulaiman, S.; Yamato, S.; Kanaya, E.; Kim, J.J.; Koga, Y.; Takano, K.; Kanaya, S. Isolation of a novel cutinase homolog with polyethylene terephthalate-degrading activity from leaf-branch compost by using a metagenomic approach. Appl. Environ. Microbiol. 2012, 78, 1556–1562. [Google Scholar] [CrossRef] [Green Version]
  62. Wei, R.; Haugwitz, G.V.; Pfaff, L.; Mican, J.; Badenhorst, C.P.S.; Liu, W.; Weber, G.; Austin, H.P.; Bednar, D.; Damborsky, J.; et al. Mechanism-Based Design of Efficient PET Hydrolases. ACS Catal. 2022, 12, 3382–3396. [Google Scholar] [CrossRef] [PubMed]
  63. Kenny, S.T.; Runic, J.N.; Kaminsky, W.; Woods, T.; Babu, R.P.; Keely, C.M.; Blau, W.; O’Connor, K.E. Up-cycling of PET (Polyethylene Terephthalate) to the biodegradable plastic PHA (Polyhydroxyalkanoate). Environ. Sci. Technol. 2008, 42, 7696–7701. [Google Scholar] [CrossRef] [PubMed]
  64. Saratale, R.G.; Cho, S.K.; Saratale, G.D.; Kadam, A.A.; Ghodake, G.S.; Kumar, M.; Bharagava, R.N.; Kumar, G.; Kim, D.S.; Mulla, S.I.; et al. A comprehensive overview and recent advances on polyhydroxyalkanoates (PHA) production using various organic waste streams. Bioresour. Technol. 2021, 325, 124685. [Google Scholar] [CrossRef] [PubMed]
  65. Choi, S.Y.; Cho, I.J.; Lee, Y.; Kim, Y.J.; Kim, K.J.; Lee, S.Y. Microbial Polyhydroxyalkanoates and Nonnatural Polyesters. Adv. Mater. 2020, 32, 1907138. [Google Scholar] [CrossRef]
  66. Kenny, S.T.; Runic, J.N.; Kaminsky, W.; Woods, T.; Babu, R.P.; O’Connor, K.E. Development of a bioprocess to convert PET derived terephthalic acid and biodiesel derived glycerol to medium chain length polyhydroxyalkanoate. Appl. Microbiol. Biotechnol. 2012, 95, 623–633. [Google Scholar] [CrossRef]
  67. Kim, H.T.; Kim, J.K.; Cha, H.G.; Kang, M.J.; Lee, H.S.; Khang, T.U.; Yun, E.J.; Lee, D.H.; Song, B.K.; Park, S.J.; et al. Biological Valorization of Poly(ethylene terephthalate) Monomers for Upcycling Waste PET. ACS Sustain. Chem. Eng. 2019, 7, 19396–19406. [Google Scholar] [CrossRef]
  68. Sadler, J.C.; Wallace, S. Microbial synthesis of vanillin from waste poly(ethylene terephthalate). Green Chem. 2021, 23, 4665–4672. [Google Scholar] [CrossRef]
  69. Kim, H.T.; Ryu, M.H.; Jung, Y.J.; Lim, S.; Song, H.M.; Park, J.; Hwang, S.Y.; Lee, H.S.; Yeon, Y.J.; Sung, B.H.; et al. Chemo-Biological Upcycling of Poly(ethylene terephthalate) to Multifunctional Coating Materials. ChemSusChem 2021, 14, 4251–4259. [Google Scholar] [CrossRef]
  70. Kataoka, M.; Sasaki, M.; Hidalgo, A.R.G.D.; Nakano, M.; Shimizu, S. Glycolic acid production using ethylene glycol-oxidizing microorganisms. Biosci. Biotechnol. Biochem. 2001, 65, 2265–2270. [Google Scholar] [CrossRef] [Green Version]
  71. Franden, M.A.; Jayakody, L.N.; Li, W.J.; Wagner, N.J.; Cleveland, N.S.; Michener, W.E.; Hauer, B.; Blank, L.M.; Wierckx, N.; Klebensberger, J.; et al. Engineering Pseudomonas putida KT2440 for efficient ethylene glycol utilization. Metab. Eng. 2018, 48, 197–207. [Google Scholar] [CrossRef]
  72. Trifunović, D.; Schuchmann, K.; Müller, V. Ethylene glycol metabolism in the acetogen Acetobacterium woodii. J. Bacteriol. 2016, 198, 1058–1065. [Google Scholar] [CrossRef] [PubMed]
  73. Werner, A.Z.; Clare, R.; Mand, T.D.; Pardo, I.; Ramirez, K.J.; Haugen, S.J.; Bratti, F.; Dexter, G.N.; Elmore, J.R.; Huenemann, J.D.; et al. Tandem chemical deconstruction and biological upcycling of poly(ethylene terephthalate) to β-ketoadipic acid by Pseudomonas putida KT2440. Metab. Eng. 2021, 67, 250–261. [Google Scholar] [CrossRef] [PubMed]
  74. Kim, D.H.; Han, D.O.; Shim, K.I.; Kim, J.K.; Pelton, J.G.; Ryu, M.H.; Joo, J.C.; Han, J.W.; Kim, H.T.; Kim, K.H. One-Pot Chemo-bioprocess of PET Depolymerization and Recycling Enabled by a Biocompatible Catalyst, Betaine. ACS Catal. 2021, 11, 3996–4008. [Google Scholar] [CrossRef]
  75. Utomo, R.N.C.; Li, W.J.; Tiso, T.; Eberlein, C.; Doeker, M.; Heipieper, H.J.; Jupke, A.; Wierckx, N.; Blank, L.M. Defined Microbial Mixed Culture for Utilization of Polyurethane Monomers. ACS Sustain. Chem. Eng. 2020, 8, 17466–17474. [Google Scholar] [CrossRef]
  76. Mahajan, N.; Gupta, P. New insights into the microbial degradation of polyurethanes. RSC Adv. 2015, 5, 41839–41854. [Google Scholar] [CrossRef]
  77. Bisceglie, F.D.; Quartinello, F.; Vielnascher, R.; Guebitz, G.M.; Pellis, A. Cutinase-Catalyzed Polyester-Polyurethane Degradation: Elucidation of the Hydrolysis Mechanism. Polymer 2022, 14, 411. [Google Scholar] [CrossRef]
  78. Liang, C.; Gracida-Alvarez, U.R.; Gallant, E.T.; Gillis, P.A.; Marques, Y.A.; Abramo, G.P.; Hawkins, T.R.; Dunn, J.B. Material Flows of Polyurethane in the United States. Environ. Sci. Technol. 2021, 55, 14215–14224. [Google Scholar] [CrossRef]
  79. Liu, J.; He, J.; Xue, R.; Xu, B.; Qian, X.; Xin, F.; Blank, L.M.; Zhou, J.; Wei, R.; Dong, W.; et al. Biodegradation and up-cycling of polyurethanes: Progress, challenges, and prospects. Biotechnol. Adv. 2021, 48, 107730. [Google Scholar] [CrossRef]
  80. Europur The Netherlands: A World-Leader in Mattress Recycling. Available online: https://www.europur.org/news-events/item/70-the-netherlands-a-world-leader-in-mattress-recycling (accessed on 18 February 2022).
  81. Robinson, S. Recycling Polyols Back into Flexible Foam. Available online: https://www.utech-polyurethane.com/node/807346/printable/print (accessed on 18 February 2022).
  82. Cregut, M.; Bedas, M.; Durand, M.J.; Thouand, G. New insights into polyurethane biodegradation and realistic prospects for the development of a sustainable waste recycling process. Biotechnol. Adv. 2013, 31, 1634–1647. [Google Scholar] [CrossRef]
  83. Jin, X.; Dong, J.; Guo, X.; Ding, M.; Bao, R.; Luo, Y. Current advances in polyurethane biodegradation. Polym. Int. 2022, 71, 1384–1392. [Google Scholar] [CrossRef]
  84. Gewert, B.; Plassmann, M.M.; Macleod, M. Pathways for degradation of plastic polymers floating in the marine environment. Environ. Sci. Proc. Impacts 2015, 17, 1513–1521. [Google Scholar] [CrossRef] [PubMed]
  85. Shah, Z.; Krumholz, L.; Aktas, D.F.; Hasan, F.; Khattak, M.; Shah, A.A. Degradation of polyester polyurethane by a newly isolated soil bacterium, Bacillus subtilis strain MZA-75. Biodegradation 2013, 24, 865–877. [Google Scholar] [CrossRef]
  86. Shah, Z.; Hasan, F.; Krumholz, L.; Atkas, D.; Shah, A.A. Degradation of polyester polyurethane by newly isolated Pseudomonas aeruginosa strain MZA-85 and analysis of degradation products by GC-MS. Int. Biodeterior. Biodegrad. 2013, 77, 114–122. [Google Scholar] [CrossRef]
  87. Shah, Z.; Gulzar, M.; Hasan, F.; Shah, A.A. Degradation of polyester polyurethane by an indigenously developed consortium of Pseudomonas and Bacillus species isolated from soil. Polym. Degrad. Stab. 2016, 134, 349–356. [Google Scholar] [CrossRef]
  88. Vargas-Suárez, M.; Fernández-Cruz, V.; Loza-Tavera, H. Biodegradation of polyacrylic and polyester polyurethane coatings by enriched microbial communities. Appl. Microbiol. Biotechnol. 2019, 103, 3225–3236. [Google Scholar] [CrossRef] [PubMed]
  89. Khan, S.; Nadir, S.; Shah, Z.U.; Shah, A.A.; Karunarathna, S.C.; Xu, J.; Khan, A.; Munir, S.; Hasan, F. Biodegradation of polyester polyurethane by Aspergillus tubingensis. Environ. Pollut. 2017, 225, 469–480. [Google Scholar] [CrossRef]
  90. Magnin, A.; Hoornaert, L.; Pollet, E.; Laurichesse, S.; Phalip, V.; Avérous, L. Isolation and characterization of different promising fungi for biological waste management of polyurethanes. Microb. Biotechnol. 2019, 12, 544–555. [Google Scholar] [CrossRef]
  91. Álvarez-Barragán, J.; Domínguez-Malfavón, L.; Vargas-Suárez, M.; González-Hernández, R.; Aguilar-Osorio, G.; Loza-Tavera, H. Biodegradative activities of selected environmental fungi on a polyester polyurethane varnish and polyether polyurethane foams. Appl. Environ. Microbiol. 2016, 82, 5225–5235. [Google Scholar] [CrossRef] [Green Version]
  92. Zafar, U.; Nzeram, P.; Langarica-Fuentes, A.; Houlden, A.; Heyworth, A.; Saiani, A.; Robson, G.D. Biodegradation of polyester polyurethane during commercial composting and analysis of associated fungal communities. Bioresour. Technol. 2014, 158, 374–377. [Google Scholar] [CrossRef]
  93. Magnin, A.; Pollet, E.; Perrin, R.; Ullmann, C.; Persillon, C.; Phalip, V.; Avérous, L. Enzymatic recycling of thermoplastic polyurethanes: Synergistic effect of an esterase and an amidase and recovery of building blocks. Waste Manag. 2019, 85, 141–150. [Google Scholar] [CrossRef]
  94. Bauer, A.S.; Tacker, M.; Uysal-Unalan, I.; Cruz, R.M.S.; Varzakas, T.; Krauter, V. Recyclability and Redesign Challenges in Multilayer Flexible Food Packaging—A Review. Foods 2021, 10, 2702. [Google Scholar] [CrossRef] [PubMed]
  95. Chamas, A.; Moon, H.; Zheng, J.; Qiu, Y.; Tabassum, T.; Jang, J.H.; Abu-Omar, M.; Scott, S.L.; Suh, S. Degradation Rates of Plastics in the Environment. ACS Sustain. Chem. Eng. 2020, 8, 3494–3511. [Google Scholar] [CrossRef] [Green Version]
  96. Ammala, A.; Bateman, S.; Dean, K.; Petinakis, E.; Sangwan, P.; Wong, S.; Yuan, Q.; Yu, L.; Patrick, C.; Leong, K.H. An overview of degradable and biodegradable polyolefins. Prog. Polym. Sci. 2011, 36, 1015–1049. [Google Scholar] [CrossRef]
  97. Albertsson, A.C.; Andersson, S.O.; Karlsson, S. The mechanism of biodegradation of polyethylene. Polym. Degrad. Stab. 1987, 18, 73–87. [Google Scholar] [CrossRef]
  98. Mukherjee, S.; Kundu, P.P. Alkaline fungal degradation of oxidized polyethylene in black liquor: Studies on the effect of lignin peroxidases and manganese peroxidases. J. Appl. Polym. Sci. 2014, 131, 8982–8990. [Google Scholar] [CrossRef]
  99. Sheik, S.; Chandrashekar, K.R.; Swaroop, K.; Somashekarappa, H.M. Biodegradation of gamma irradiated low density polyethylene and polypropylene by endophytic fungi. Int. Biodeterior. Biodegrad. 2015, 105, 21–29. [Google Scholar] [CrossRef]
  100. Pometto, A.L.; Lee, B.; Johnson, K.E. Production of an extracellular polyethylene-degrading enzyme (s) by Streptomyces species. Appl. Environ. Microbiol. 1992, 58, 731–733. [Google Scholar] [CrossRef] [Green Version]
  101. Iiyoshi, Y.; Tsutsumi, Y.; Nishida, T. Polyethylene degradation by lignin-degrading fungi and manganese peroxidase. J. Wood Sci. 1998, 44, 222–229. [Google Scholar] [CrossRef]
  102. Zhao, J.; Guo, Z.; Ma, X.; Liang, G.; Wang, J. Novel surface modification of high-density polyethylene films by using enzymatic catalysis. J. Appl. Polym. Sci. 2004, 91, 3673–3678. [Google Scholar] [CrossRef]
  103. Fujisawa, M. Degradation of polyethylene and nylon-66 by the laccase-mediator system. J. Polym. Environ. 2001, 9, 103–108. [Google Scholar] [CrossRef]
  104. Santo, M.; Weitsman, R.; Sivan, A. The role of the copper-binding enzyme—laccase—in the biodegradation of polyethylene by the actinomycete Rhodococcus ruber. Int. Biodeterior. Biodegrad. 2013, 84, 204–210. [Google Scholar] [CrossRef]
  105. Li, H.; Wang, X.L.; Mu, B.Z.; Gu, J.D.; Liu, Y.D.; Lin, K.F.; Lu, S.G.; Lu, Q.; Li, B.Z.; Li, Y.Y.; et al. Molecular detection, quantification and distribution of alkane-degrading bacteria in production water from low temperature oilfields. Int. Biodeterior. Biodegrad. 2013, 76, 49–57. [Google Scholar] [CrossRef]
  106. Yeom, S.J.; Le, T.K.; Yun, C.H. P450-driven plastic-degrading synthetic bacteria. Trends Biotechnol. 2022, 40, 166–179. [Google Scholar] [CrossRef] [PubMed]
  107. Jeon, H.J.; Kim, M.N. Comparison of the functional characterization between alkane monooxygenases for low-molecular-weight polyethylene biodegradation. Int. Biodeterior. Biodegrad. 2016, 114, 202–208. [Google Scholar] [CrossRef]
  108. Watanabe, M.; Kawai, F.; Shibata, M.; Yokoyama, S.; Sudate, Y. Computational method for analysis of polyethylene biodegradation. J. Comput. Appl. Math. 2003, 161, 133–144. [Google Scholar] [CrossRef] [Green Version]
  109. Montazer, Z.; Najafi, M.B.H.; Levin, D.B. Microbial degradation of low-density polyethylene and synthesis of polyhydroxyalkanoate polymers. Can. J. Microbiol. 2019, 65, 224–234. [Google Scholar] [CrossRef]
  110. Hammerer, L.; Winkler, C.K.; Kroutil, W. Regioselective Biocatalytic Hydroxylation of Fatty Acids by Cytochrome P450s. Catal. Lett. 2017, 148, 787–812. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, Y.; Pedersen, J.N.; Eser, B.E.; Guo, Z. Biodegradation of polyethylene and polystyrene: From microbial deterioration to enzyme discovery. Biotechnol. Adv. 2022, 60, 107991. [Google Scholar] [CrossRef]
  112. Peixoto, J.; Vizzotto, C.; Ramos, A.; Alves, G.; Steindorff, A.; Krüger, R. The role of nitrogen metabolism on polyethylene biodegradation. J. Hazard. Mater. 2022, 432, 128682. [Google Scholar] [CrossRef]
  113. Danso, D.; Chow, J.; Streita, W.R. Plastics: Environmental and biotechnological perspectives on microbial degradation. Appl. Environ. Microbiol. 2019, 85, 2. [Google Scholar] [CrossRef] [Green Version]
  114. Raut, S.; Raut, S.; Sharma, M.; Srivastav, C.; Adhikari, B.; Sen, S.K. Enhancing Degradation of Low Density Polyethylene Films by Curvularia lunata SG1 Using Particle Swarm Optimization Strategy. Indian J. Microbiol. 2015, 55, 258–268. [Google Scholar] [CrossRef] [PubMed]
  115. Mierzwa-Hersztek, M.; Gondek, K.; Kopeć, M. Degradation of Polyethylene and Biocomponent-Derived Polymer Materials: An Overview. J. Polym. Environ. 2019, 27, 600–611. [Google Scholar] [CrossRef] [Green Version]
  116. Restrepo-Flórez, J.M.; Bassi, A.; Thompson, M.R. Microbial degradation and deterioration of polyethylene-A review. Int. Biodeterior. Biodegrad. 2014, 88, 83–90. [Google Scholar] [CrossRef]
  117. Sohn, Y.J.; Kim, H.T.; Baritugo, K.A.; Jo, S.Y.; Song, H.M.; Park, S.Y.; Park, S.K.; Pyo, J.; Cha, H.G.; Kim, H.; et al. Recent Advances in Sustainable Plastic Upcycling and Biopolymers. Biotechnol. J. 2020, 15, 1900489. [Google Scholar] [CrossRef]
  118. Taghavi, N.; Singhal, N.; Zhuang, W.Q.; Baroutian, S. Degradation of plastic waste using stimulated and naturally occurring microbial strains. Chemosphere 2021, 263, 127975. [Google Scholar] [CrossRef] [PubMed]
  119. Koivuranta, K.; Andberg, M.B.; Nygren, H.; Castillo, S. Enzymes, Micro-Organisms and Uses Thereof, and a Method of Degrading Hydrocarbon Chains. International Patent No. WO2022/175592, 25 August 2022. [Google Scholar]
  120. Guzik, M.W.; Kenny, S.T.; Duane, G.F.; Casey, E.; Woods, T.; Babu, R.P.; Nikodinovic-Runic, J.; Murray, M.; O’Connor, K.E. Conversion of post consumer polyethylene to the biodegradable polymer polyhydroxyalkanoate. Appl. Microbiol. Biotechnol. 2014, 98, 4223–4232. [Google Scholar] [CrossRef]
  121. Radecka, I.; Irorere, V.; Jiang, G.; Hill, D.; Williams, C.; Adamus, G.; Kwiecień, M.; Marek, A.A.; Zawadiak, J.; Johnston, B.; et al. Oxidized polyethylene wax as a potential carbon source for PHA production. Materials 2016, 9, 367. [Google Scholar] [CrossRef] [Green Version]
  122. Johnston, B.; Jiang, G.; Hill, D.; Adamus, G.; Kwiecień, I.; Zięba, M.; Sikorska, W.; Green, M.; Kowalczuk, M.; Radecka, I. The Molecular Level Characterization of Biodegradable Polymers Originated from Polyethylene Using Non-Oxygenated Polyethylene Wax as a Carbon Source for Polyhydroxyalkanoate Production. Bioengineering 2017, 4, 73. [Google Scholar] [CrossRef] [Green Version]
  123. Montazer, Z.; Habibi-Najafi, M.B.; Mohebbi, M.; Oromiehei, A. Microbial Degradation of UV-Pretreated Low-Density Polyethylene Films by Novel Polyethylene-Degrading Bacteria Isolated from Plastic-Dump Soil. J. Polym. Environ. 2018, 26, 3613–3625. [Google Scholar] [CrossRef]
  124. Jeon, H.J.; Kim, M.N. Isolation of mesophilic bacterium for biodegradation of polypropylene. Int. Biodeterior. Biodegrad. 2016, 115, 244–249. [Google Scholar] [CrossRef]
  125. Skariyachan, S.; Taskeen, N.; Kishore, A.P.; Krishna, B.V.; Naidu, G. Novel consortia of enterobacter and pseudomonas formulated from cow dung exhibited enhanced biodegradation of polyethylene and polypropylene. J. Environ. Manag. 2021, 284, 112030. [Google Scholar] [CrossRef]
  126. Skariyachan, S.; Patil, A.A.; Shankar, A.; Manjunath, M.; Bachappanavar, N.; Kiran, S. Enhanced polymer degradation of polyethylene and polypropylene by novel thermophilic consortia of Brevibacillus sps. and Aneurinibacillus sp. screened from waste management landfills and sewage treatment plants. Polym. Degrad. Stab. 2018, 149, 52–68. [Google Scholar] [CrossRef]
  127. Yang, S.S.; Ding, M.Q.; He, L.; Zhang, C.H.; Li, Q.X.; Xing, D.F.; Cao, G.L.; Zhao, L.; Ding, J.; Ren, N.Q.; et al. Biodegradation of polypropylene by yellow mealworms (Tenebrio molitor) and superworms (Zophobas atratus) via gut-microbe-dependent depolymerization. Sci. Total Environ. 2020, 756, 144087. [Google Scholar] [CrossRef] [PubMed]
  128. Muenmee, S.; Chiemchaisri, W.; Chiemchaisri, C. Enhancement of biodegradation of plastic wastes via methane oxidation in semi-aerobic landfill. Int. Biodeterior. Biodegrad. 2016, 113, 244–255. [Google Scholar] [CrossRef]
  129. Johnston, B.; Radecka, I.; Chiellini, E.; Barsi, D.; Ilieva, V.I.; Sikorska, W.; Musiol, M.; Ziȩba, M.; Chaber, P.; Marek, A.A.; et al. Mass spectrometry reveals molecular structure of polyhydroxyalkanoates attained by bioconversion of oxidized polypropylene waste fragments. Polymers 2019, 11, 1580. [Google Scholar] [CrossRef] [Green Version]
  130. Mihreteab, M.; Stubblefield, B.A.; Gilbert, E.S. Microbial bioconversion of thermally depolymerized polypropylene by Yarrowia lipolytica for fatty acid production. Appl. Microbiol. Biotechnol. 2019, 103, 7729–7740. [Google Scholar] [CrossRef]
  131. Mihreteab, M.; Stubblefield, B.A.; Gilbert, E.S. Enhancing polypropylene bioconversion and lipogenesis by Yarrowia lipolytica using a chemical/biological hybrid process. J. Biotechnol. 2021, 332, 94–102. [Google Scholar] [CrossRef]
  132. Savoldelli, J.; Tomback, D.; Savoldelli, H. Breaking down polystyrene through the application of a two-step thermal degradation and bacterial method to produce usable byproducts. Waste Manag. 2017, 60, 123–126. [Google Scholar] [CrossRef]
  133. Farrelly, T.A.; Shaw, I.C. Polystyrene as Hazardous Household Waste. In Household Hazardous Waste Management; InTechOpen: London, UK, 2017. [Google Scholar] [CrossRef] [Green Version]
  134. Kim, H.R.; Lee, H.M.; Yu, H.C.; Jeon, E.; Lee, S.; Li, J.; Kim, D.H. Biodegradation of Polystyrene by Pseudomonas sp. Isolated from the Gut of Superworms (Larvae of Zophobas atratus). Environ. Sci. Technol. 2020, 54, 6987–6996. [Google Scholar] [CrossRef]
  135. Ho, B.T.; Roberts, T.K.; Lucas, S. An overview on biodegradation of polystyrene and modified polystyrene: The microbial approach. Crit. Rev. Biotechnol. 2018, 38, 308–320. [Google Scholar] [CrossRef]
  136. Auta, H.S.; Emenike, C.U.; Fauziah, S.H. Screening of Bacillus strains isolated from mangrove ecosystems in Peninsular Malaysia for microplastic degradation. Environ. Pollut. 2017, 231, 1552–1559. [Google Scholar] [CrossRef] [PubMed]
  137. Syranidou, E.; Karkanorachaki, K.; Amorotti, F.; Franchini, M.; Repouskou, E.; Kaliva, M.; Vamvakaki, M.; Kolvenbach, B.; Fava, F.; Corvini, P.F.-X.; et al. Biodegradation of weathered polystyrene films in seawater microcosms. Sci. Rep. 2017, 7, 17991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Yang, Y.; Wang, J.; Xia, M. Biodegradation and mineralization of polystyrene by plastic-eating superworms Zophobas atratus. Sci. Total Environ. 2020, 708, 135233. [Google Scholar] [CrossRef] [PubMed]
  139. Wang, Z.; Xin, X.; Shi, X.; Zhang, Y. A polystyrene-degrading Acinetobacter bacterium isolated from the larvae of Tribolium castaneum. Sci. Total Environ. 2020, 726, 138564. [Google Scholar] [CrossRef] [PubMed]
  140. Ward, P.G.; Goff, M.; Donner, M.; Kaminsky, W.; O’Connor, K.E. A two step chemo-biotechnological conversion of polystyrene to a biodegradable thermoplastic. Environ. Sci. Technol. 2006, 40, 2433–2437. [Google Scholar] [CrossRef]
  141. Goff, M.; Ward, P.G.; O’Connor, K.E. Improvement of the conversion of polystyrene to polyhydroxyalkanoate through the manipulation of the microbial aspect of the process: A nitrogen feeding strategy for bacterial cells in a stirred tank reactor. J. Biotechnol. 2007, 132, 283–286. [Google Scholar] [CrossRef]
  142. Ward, P.G.; Roo, G.D.; O’Connor, K.E. Accumulation of polyhydroxyalkanoate from styrene and phenylacetic acid by Pseudomonas putida CA-3. Appl. Environ. Microbiol. 2005, 71, 2046–2052. [Google Scholar] [CrossRef] [Green Version]
  143. Nikodinovic-Runic, J.; Casey, E.; Duane, G.F.; Mitic, D.; Hume, A.R.; Kenny, S.T.; O’Connor, K.E. Process analysis of the conversion of styrene to biomass and medium chain length polyhydroxyalkanoate in a two-phase bioreactor. Biotechnol. Bioeng. 2011, 108, 2447–2455. [Google Scholar] [CrossRef]
  144. Ru, J.; Huo, Y.; Yang, Y. Microbial Degradation and Valorization of Plastic Wastes. Front. Microbiol. 2020, 11, 442. [Google Scholar] [CrossRef] [Green Version]
  145. Giacomucci, L.; Raddadi, N.; Soccio, M.; Lotti, N.; Fava, F. Polyvinyl chloride biodegradation by Pseudomonas citronellolis and Bacillus flexus. N. Biotechnol. 2019, 52, 35–41. [Google Scholar] [CrossRef]
  146. Peng, B.Y.; Chen, Z.; Chen, J.; Yu, H.; Zhou, X.; Criddle, C.S.; Wu, W.M.; Zhang, Y. Biodegradation of Polyvinyl Chloride (PVC) in Tenebrio molitor (Coleoptera: Tenebrionidae) larvae. Environ. Int. 2020, 145, 106106. [Google Scholar] [CrossRef] [PubMed]
  147. Giacomucci, L.; Raddadi, N.; Soccio, M.; Lotti, N.; Fava, F. Biodegradation of polyvinyl chloride plastic films by enriched anaerobic marine consortia. Mar. Environ. Res. 2020, 158, 104949. [Google Scholar] [CrossRef] [PubMed]
  148. Ali, M.I.; Ahmed, S.; Robson, G.; Javed, I.; Ali, N.; Atiq, N.; Hameed, A. Isolation and molecular characterization of polyvinyl chloride (PVC) plastic degrading fungal isolates. J. Basic Microbiol. 2014, 54, 18–27. [Google Scholar] [CrossRef] [PubMed]
  149. Zhou, X.L.; He, P.J.; Peng, W.; Yi, S.X.; Lü, F.; Shao, L.M.; Zhang, H. Upcycling waste polyvinyl chloride: One-pot synthesis of valuable carbon materials and pipeline-quality syngas via pyrolysis in a closed reactor. J. Hazard. Mater. 2022, 427, 128210. [Google Scholar] [CrossRef] [PubMed]
  150. Yuan, G.; Chen, D.; Yin, L.; Wang, Z.; Zhao, L.; Wang, J.Y. High efficiency chlorine removal from polyvinyl chloride (PVC) pyrolysis with a gas–liquid fluidized bed reactor. Waste Manag. 2014, 34, 1045–1050. [Google Scholar] [CrossRef] [PubMed]
  151. Ballerstedt, H.; Tiso, T.; Wierckx, N.; Wei, R.; Averous, L.; Bornscheuer, U.; O’Connor, K.; Floehr, T.; Jupke, A.; Klankermayer, J.; et al. MIXed plastics biodegradation and UPcycling using microbial communities: EU Horizon 2020 project MIX-UP started January 2020. Environ. Sci. Eur. 2021, 33, 99. [Google Scholar] [CrossRef] [PubMed]
  152. Sullivan, K.P.; Werner, Z.A.; Ramirez, K.J.; Ellis, L.E.; Bussard, J.R.; Black, B.A.; Brander, D.G.; Bratti, F.; Buss, B.L.; Dong, X.; et al. Mixed plastic waste valorization through tandem chemical oxidation and biological funneling. Science 2022, 378, 207–211. [Google Scholar] [CrossRef]
  153. Edwards, S.; León-zayas, R.; Ditter, R.; Laster, H.; Sheehan, G.; Anderson, O.; Beattie, T.; Mellies, J.L. Microbial Consortia and Mixed Plastic Waste: Pangenomic Analysis Reveals Potential for Degradation of Multiple Plastic Types via Previously Identified PET Degrading Bacteria. Int. J. Mol. Sci. 2022, 23, 5612. [Google Scholar] [CrossRef]
  154. European Comission. Waste Framework Directive. Available online: https://environment.ec.europa.eu/topics/waste-and-recycling/waste-framework-directive_en (accessed on 28 September 2022).
  155. Sheel, A.; Pant, D. Chemical Depolymerization of PET Bottles via Glycolysis. Recycl. Polyethyl. Terephthalate Bottles 2019, 1, 61–84. [Google Scholar] [CrossRef]
  156. Sinha, V.; Patel, M.R.; Patel, J.V. Pet waste management by chemical recycling: A review. J. Polym. Environ. 2010, 18, 8–25. [Google Scholar] [CrossRef]
  157. Gan, Z.; Zhang, H. PMBD: A Comprehensive Plastics Microbial Biodegradation Database. Database 2019, 2019, baz119. [Google Scholar] [CrossRef] [PubMed]
  158. Zhu, B.; Wang, D.; Wei, N. Enzyme discovery and engineering for sustainable plastic recycling. Trends Biotechnol. 2022, 40, 22–37. [Google Scholar] [CrossRef] [PubMed]
  159. Kremser, K.; Gerl, P.; Borrás, A.B.; Espinosa, D.R.; Martínez, B.M.; Guebitz, G.M.; Pellis, A. Bioleaching/enzyme-based recycling of aluminium and polyethylene from beverage cartons packaging waste. Resour. Conserv. Recycl. 2022, 185, 106444. [Google Scholar] [CrossRef]
Figure 1. Global plastic use by region. The figure is created according to the data from OECD [1].
Figure 1. Global plastic use by region. The figure is created according to the data from OECD [1].
Polymers 14 04996 g001
Figure 2. Post-consumer plastic waste management in US (2018) and EU27+3 (2020). The figure is created according to the data from US EPA [2] and PlasticsEurope [3].
Figure 2. Post-consumer plastic waste management in US (2018) and EU27+3 (2020). The figure is created according to the data from US EPA [2] and PlasticsEurope [3].
Polymers 14 04996 g002
Figure 3. The pathway of plastics from primary production until end-of-life, during the period 1950–2015, based on the study of Geyer et al. (2017) [9] and Lebreton et al. (2019) [10].
Figure 3. The pathway of plastics from primary production until end-of-life, during the period 1950–2015, based on the study of Geyer et al. (2017) [9] and Lebreton et al. (2019) [10].
Polymers 14 04996 g003
Figure 4. Distribution of plastic demand globally (a) and in EU27+3 (b) by polymer types. The figure is created according to the data from OECD [1] and PlasticsEurope [3].
Figure 4. Distribution of plastic demand globally (a) and in EU27+3 (b) by polymer types. The figure is created according to the data from OECD [1] and PlasticsEurope [3].
Polymers 14 04996 g004
Figure 5. The hierarchy of plastic waste management according to EU Waste Framework Directive.
Figure 5. The hierarchy of plastic waste management according to EU Waste Framework Directive.
Polymers 14 04996 g005
Table 2. PU bioupcycling.
Table 2. PU bioupcycling.
Depolymerization StrategyDepolymerization Products Used as a Feedstock for
Fermentation Step
Fermentation StrategyProducts from
Fermentation
TiterProductivityYieldRef.
Enzymatic degradation of polycaprolactone polyol-based PU by esterase (E3576) in 0.1 M phosphate buffer (pH 7.0). The enzyme solution was replaced every 3–4 d to overcome a loss of enzymatic activity.6-hydroxycaproic acid (1 g/L)-----[93]
-Adipic acid (AA) (mock substrate to study upcycling of PU-derived monomer)Bioconversion (at 30 °C and 200 rpm for 135 h) using metabolically engineered P. putida KT2440 A12.1p pPS05 to convert AA into HAA and then to rhamnolipidRhamnolipid0.02 g/LNA0.014 gRhamnolipid/gSubstrate[75]
-1,4-Butanediol (BDO) (mock substrate to study upcycling of PU-derived monomer)Bioconversion (at 30 °C and 200 rpm for 135 h) using metabolically engineered P. putida KT2440 B10.1 pPR05 to convert BDO into HAA and then to rhamnolipidRhamnolipid0.13 g/LNA0.088 gRhamnolipid/gSubstrate[75]
-EG (mock substrate to study upcycling of PU-derived monomer)Bioconversion (at 30 °C and 200 rpm for 135 h) using metabolically engineered P. putida KT2440 ∆gclR ∆PP_2046 ∆PP_2662::14d to convert EG into HAA and then to rhamnolipidRhamnolipid0.07 g/LNA0.038 gRhamnolipid/gSubstrate[75]
-AA + BDO + EG (mock hydrolysate to study upcycling of PU-derived monomers)Bioconversion (at 30 °C and 200 rpm for 210 h) using mixed culture of three metabolically engineered P. putida KT2440 to convert the mock hydrolysate into HAA and then to rhamnolipidRhamnolipid0.1 g/LNA0.008 gRhamnolipid/gSubstrate[75]
AA + BDO + EG + 2,4-toluenediamine (TDA) (mock hydrolysate to study upcycling of PU-derived monomers)Bioconversion (at 30 °C and 200 rpm for 210 h) using mixed culture of three metabolically engineered P. putida KT2440 to convert the mock hydrolysate into HAA and then to rhamnolipid without extraction of TDARhamnolipid0.02 g/LNA0.002 gRhamnolipid/gSubstrate[75]
AA + BDO + EG + TDA (mock hydrolysate to study upcycling of PU-derived monomers)Bioconversion (at 30 °C and 200 rpm for 210 h) using mixed culture of three metabolically engineered P. putida KT2440 to convert the mock hydrolysate into HAA and then to rhamnolipid with extraction of TDA at pH 3.5Rhamnolipid0.07 g/LNA0.005 gRhamnolipid/gSubstrate[75]
Table 4. PP bioupcycling.
Table 4. PP bioupcycling.
Depolymerization StrategyDepolymerization
Products Used as a
Feedstock for
Fermentation Step
Fermentation StrategyProducts from
Fermentation
TiterProductivityYieldRef.
Pro-degradation at 180 °C with 1% (w/w) cobalt stearate as pro-oxidant/pro-degradant additiveOxidatively pro-degraded PPFermentation in shake flask containing 2 g/L oxidatively pro-degraded PP emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H16PHA0.58 g/LNA0.26 gPHA/gCDW[129]
Oxidatively pro-degraded PP was subjected to oxidative degradation in a two-phase system (gas-liquid phase), after melting at 60–80 °C and using oxygen-ozone mixtureThermal oxidized PPFermentation in shake flask containing 2 g/L thermal-oxidized PP emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H16PHA1.36 g/LNA0.42 gPHA/gCDW[129]
Pyrolysis at 540 °CPyrolysis oil contained branched chain fatty alcohols (51%) and alkenes (25%)Fermentation in shake flask containing OP4 medium (15 g/L pyrolysis oil, 5.4 g/L Tween 80, 4.5 g/L oleic acid, 1.25 g/L (NH4)2SO4, 2.5 g/L KH2PO4, and 0.830 g/L MgSO4·7H2O) at 30 °C for 312 h by Yarrowia lipolytica strain 78-003Fatty acids with C18 compounds (oleic acid, linoleic acid, and stearic acid) as dominant products, followed by C16 compounds (palmitic and palmitoleic acids).492 mg/LNANA[130]
Table 5. PS bioupcycling.
Table 5. PS bioupcycling.
Depolymerization StrategyDepolymerization Products
Used as a Feedstock for
Fermentation Step
Fermentation StrategyProducts from
Fermentation
TiterProductivityYieldRef.
-Styrene (mock substrate to study upcycling of PS-derived monomer)Fermentation in shake flask containing 1.85 g/L styrene as a sole carbon source and 67 gN/L NaNH4HPO4·4H2O as a nitrogen source at 30 °C for 48 h by P. putida CA-3PHANANA0.099 gPHA/gStyrene[142]
Pyrolysis at 520 °CStyrene oil (82.8% (w/w) styrene as well as low level of α-methylstyrene, toluene, styrene dimer, and traces of other aromatic compounds)Fermentation in shake flask containing styrene oil as a sole carbon source and 1 g/L NaNH4HPO4·4H2O as a nitrogen source at 30 °C by P. putida CA-3PHA0.14 g/LNA0.0625 gPHA/gStyrene oil (0.25 gPHA/gCDW)[140]
Pyrolysis at 520 °CStyrene oil (82.8% (w/w) styrene as well as low level of α-methylstyrene, toluene, styrene dimer, and traces of other aromatic compounds)Fermentation in 7.5 L stirred tank reactor feeding a sole carbon source through the gaseous phase contained styrene oil at a concentration of 9.5 mg/L (flow rate 0.15 L/min for the first 3 h of growth and increased to 0.25 L/min for the subsequent 3 h, and finally, to 0.65 L/min for the remainder) at 30 °C by P. putida CA-3PHA0.32 g/LNA0.1 gPHA/gStyrene oil (0.57 gPHA/gCDW)[140]
PyrolysisDistilled styrene oil (89.9% styrene, 5.63% α-methylbenzene, 2.63% toluene, 1.05% ehtylbenzene, 0.43% benzene, 0.19% 1-ethyl-2-methy benzene, and 0.17% unknown)Fermentation in stirred tank reactor feeding distilled styrene oil at a feed rate of 75 mg/L/h (equivalent to 69 mgC/L/h) and NaNH4HPO4·4H2O at a feed rate of 1.5 mg/L/h at 30 °C by P. putida CA-3PHA0.82 g/LNA0.28 gPHA/gStyrene oil (0.42 gPHA/gCDW)[141]
-Styrene (mock substrate to study upcycling of PS-derived monomer)Fed-batch fermentation in stirred tank reactor feeding styrene as a carbon source overtime through air sparger and NH4Cl as a nitrogen source at different feed rate during the operation period. The fermentation was conducted at 30 °C and pH 6.9 by P. putida CA-3mclPHA3.36 g/LNA0.32 gPHA/gCDW[143]
Pro-degradationPro-degraded PSFermentation in shake flask containing 3.7 g/L prodegraded PS emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H16PHA0.52 g/LNA0.39 gPHA/gCDW[15]
Pro-degraded PP was subjected to thermal oxidation (60 °C) in a two-phase system (gas-solid phase) using oxygen-ozone mixtureThermal oxidized PS (60 °C)Fermentation in shake flask containing 3.7 g/L thermal oxidized PS emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H17PHA1.72 g/LNA0.48 gPHA/gCDW[15]
Pro-degraded PP was subjected to thermal oxidation (80 °C) in a two-phase system (gas-solid phase) using oxygen-ozone mixtureThermal oxidized PS (80 °C)Fermentation in shake flask containing 3.7 g/L thermal oxidized PS emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H18PHA1.28 g/LNA0.42 gPHA/gCDW[15]
Pro-degraded PP was subjected to thermal oxidation (100 °C) in a two-phase system (gas-solid phase) using oxygen-ozone mixtureThermal oxidized PS (100 °C)Fermentation in shake flask containing 3.7 g/L thermal oxidized PS emulsified in TSB by sonication as a sole carbon source at 30 °C for 48 h by C. necator H19PHA0.96 g/LNA0.36 gPHA/gCDW[15]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lomwongsopon, P.; Varrone, C. Critical Review on the Progress of Plastic Bioupcycling Technology as a Potential Solution for Sustainable Plastic Waste Management. Polymers 2022, 14, 4996. https://doi.org/10.3390/polym14224996

AMA Style

Lomwongsopon P, Varrone C. Critical Review on the Progress of Plastic Bioupcycling Technology as a Potential Solution for Sustainable Plastic Waste Management. Polymers. 2022; 14(22):4996. https://doi.org/10.3390/polym14224996

Chicago/Turabian Style

Lomwongsopon, Passanun, and Cristiano Varrone. 2022. "Critical Review on the Progress of Plastic Bioupcycling Technology as a Potential Solution for Sustainable Plastic Waste Management" Polymers 14, no. 22: 4996. https://doi.org/10.3390/polym14224996

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

Article Metrics

Back to TopTop