Next Article in Journal
Integrating Soil Diagnostics and Life Cycle Assessment to Enhance Vineyard Sustainability on a Volcanic Island (Tenerife, Spain)
Previous Article in Journal
Circular Economy Framework for Plastic Waste Management: A Case Study from Coastal Hotels in Zanzibar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biocatalytic Recycling of Polyethylene Terephthalate: From Conventional to Innovative Routes for Transforming Plastic and Textile Waste into Renewable Resources

by
Damayanti Damayanti
1,2,*,
David Septian Sumanto Marpaung
3,
Abdul Rozak Kodarif
1,
Andri Sanjaya
1,
Desi Riana Saputri
1,
Yunita Fahni
1,
Lutfia Rahmiyati
1,
Putri Zulva Silvia
1,
Dewi Qurrota A’yuni
1,
Calaelma Logys Imalia
1,
Dikri Uzlifah Janah
1 and
Ho Shing Wu
4,*
1
Department of Chemical Engineering, Institut Teknologi Sumatera, Jl. Terusan Ryacudu, Way Huwi, Kec. Jati Agung, Lampung Selatan 35365, Lampung, Indonesia
2
Department of Energy System Engineering, Institut Teknologi Sumatera, Jl. Terusan Ryacudu, Way Huwi, Kec. Jati Agung, Lampung Selatan 35365, Lampung, Indonesia
3
Department of Biosystems Engineering, Institut Teknologi Sumatera, Jl. Terusan Ryacudu, Way Huwi, Kec. Jati Agung, Lampung Selatan 35365, Lampung, Indonesia
4
Department of Chemical Engineering and Materials Science, Yuan Ze University, 135 Yuan-Tung Road, Chung-Li, Taoyuan 32003, Taiwan
*
Authors to whom correspondence should be addressed.
Resources 2025, 14(11), 176; https://doi.org/10.3390/resources14110176
Submission received: 21 October 2025 / Revised: 13 November 2025 / Accepted: 14 November 2025 / Published: 20 November 2025

Abstract

The rapid accumulation of plastic and textile waste, particularly polyethylene terephthalate (PET), has emerged as a global challenge for sustainable resource management. Conventional recycling methods, including mechanical and chemical routes, recover limited value and often degrade material quality while consuming substantial energy. Biocatalytic recycling, by contrast, offers a resource-efficient alternative that transforms post-consumer PET into high-purity monomers under mild and environmentally benign conditions. This review examines advances in enzymatic PET depolymerization, focusing on hydrolases such as cutinases, PETases, MHETases, and lipases. The discussion highlights enzyme engineering, reactor design, and process integration that improve kinetics, thermostability, and yield. From a resource perspective, biocatalytic recycling redefines PET waste as a renewable carbon feedstock capable of re-entering industrial cycles, thereby reducing reliance on virgin petrochemicals and mitigating greenhouse gas emissions. Ultimately, this review positions biocatalytic PET recycling as a cornerstone technology for achieving circularity and advancing global resource sustainability.

1. Introduction

Plastics are usually man-made polymers chemically synthesized, with a backbone composed entirely of carbon–carbon bonds. Most of their raw materials come from fossil fuels. They have become essential in modern life due to their versatility, durability, and low cost, making them vital across industries such as packaging, textiles, electronics, and healthcare [1,2,3]. Global plastic production is estimated to reach around 1.1 billion tons by 2050 [4]. Its widespread use is driven by advantages like ease of processing, lightweight nature, and exceptional barrier properties, especially in food and beverage packaging [5]. However, the rapid growth of plastic consumption has led to severe environmental challenges, including plastic waste accumulation, greenhouse gas emissions, and threats to marine and terrestrial ecosystems. The majority of plastic waste originates from single-use packaging, which accounts for approximately 50% of global plastic waste [6]. In 2015, it was estimated that emissions from oil extraction to refining reached 68 million tons of CO2 equivalents (CO2e) [7]. This calculation was based on the weighted average carbon intensity of energy production from 8966 active oil fields across 90 countries, with approximately 4% of crude oil allocated for plastic manufacturing. In addition to the production and manufacturing stages, greenhouse gas emissions are also generated during the extraction and transportation of raw materials, as well as from plastic waste management and the release of plastics into the environment [8]. Emissions from production facilities are typically influenced by factors such as the efficiency, design, and lifespan of the equipment used. Even after plastics are discarded, their impact on the global climate continues, with most environmental effects occurring after their useful life has ended [9]. Implementing effective plastic waste management practices is essential for addressing sustainability challenges and mitigating environmental problems.
Apart from plastic waste, textiles that enhance our daily lives with beauty and comfort also pose a significant challenge for solid waste management. Garments typically have an average lifespan of around 5 years, after which most post-consumer textile waste is discarded, amounting to an estimated 65–92 million metric tons annually worldwide [10]. In 2018, global textile production reached 105 million metric tons [11]. The amount of textile waste ending up in landfills or incinerators continues to rise significantly [12], with less than 1% of the materials used in clothing production being recycled into new garments [13]. Currently, recycling technologies face five main challenges: (i) the absence of commercially viable methods for recycling low-grade textile fractions, (ii) limited large-scale processes and expertise for separating fiber types in mixed blends and composite structures, (iii) high costs associated with recovery processes, (iv) a recycling market dominated mainly by low-quality materials and blends, and (v) expensive logistics combined with a lack of sufficient textile recycling facilities at both local and regional levels [14,15]. To address these challenges, it is essential to deconstruct and separate the different fiber components into purified streams, enabling more efficient recycling and material recovery.
Both textile and plastic waste contain PET as a raw material. PET is widely used because of its durability, lightweight nature, and versatility, making it a popular choice in industries like packaging, especially for beverage bottles, and textiles, notably for synthetic fibers such as polyester. Compared to polyvinyl chloride (PVC), PET is considered a safer material since PVC in plastic waste can emit harmful chlorinated compounds, such as gaseous hydrogen chloride and chlorine-containing dioxins, during recycling or thermal energy recovery processes [16]. With the rapid growth of the plastic bottle market, PET consumption has been rising significantly to meet increasing demand [17]. PET also makes up a significant portion of consumer textile fibers, accounting for 88% of global textile fiber production when combined with other dominant fibers [18]. The primary approach to ensuring the sustainability of PET production is through recycling. Recycled PET is mainly processed into fibers (72%), followed by bottles (10%), sheets (10%), tapes (5%), and other products (3%), using chemical and mechanical recycling methods [19]. However, if the proportion of recycled PET in the production system is not optimal, the recycling process can still have negative environmental impacts [20]. To address this issue, efforts such as process optimization and enhancing recycling systems are crucial for further reducing greenhouse gas emissions and improving overall environmental performance.
The recycling of PET, derived from either plastic or textile waste, can be carried out through mechanical, chemical, or biological processes, with the primary goal of transforming PET into economically reusable forms. Mechanical recycling is typically categorized into primary and secondary recycling, while chemical and biological recycling are generally classified as tertiary and quaternary recycling. Each method offers distinct advantages and limitations, and the choice of approach is dependent on specific user requirements and application goals. Mechanical recycling transforms post-consumer plastics and textiles into raw materials for manufacturing new plastic or textile products [21]. Mechanical recycling represents the most widely used approach for PET products. The process involves collecting and shredding post-consumer PET into small fragments, washing and removing contaminants, and then melting the material into PET resin [22,23]. The purified PET flakes are then remelted and molded into new products, such as bottles, fibers, and films. Mechanical recycling offers a cost-effective, energy-efficient approach that helps reduce greenhouse gas emissions and conserve natural resources [24]. It also reduces the amount of PET products sent to landfills or oceans, aiding the shift toward a circular economy. However, this approach faces challenges such as inconsistent quality and availability of collected PET, contamination issues, and material property degradation. In such situations, alternative methods such as chemical recycling might offer better solutions.
Chemical recycling is an emerging technology that depolymerizes PET products into their constituent monomers, which can then be repolymerized to produce new PET materials [25,26]. Chemical recycling addresses the limitations of mechanical recycling by converting PET waste into high-quality raw materials, including terephthalic acid (TPA) and monoethylene glycol (MEG). This approach can also lower greenhouse gas emissions and reduce energy consumption by up to 50% compared to the production of virgin PET [27]. However, chemical recycling demands specialized equipment and technical expertise, making the process both costly and complex. In addition, the process is more expensive and intricate than mechanical recycling, and significant challenges remain regarding its scalability and commercial application [28]. Therefore, biological recycling could serve as an alternative to mechanical and chemical approaches.
Global plastic and textile waste has raised concerns about resource depletion and environmental pollution. Among synthetic polymers, PET plays a key role in both sectors due to its use in packaging and fibers. This dual purpose makes PET not just a pollutant but also a valuable, recoverable carbon resource. Traditional mechanical and chemical recycling methods, though common, often yield lower-quality materials and consume substantial energy. Therefore, developing sustainable recycling strategies that preserve monomeric value and conserve resources is crucial. Biocatalytic recycling—using enzymes to break down PET under mild conditions—offers an environmentally friendly and resource-efficient solution.
Biorecycling, or enzymatic recycling, is a growing technique that uses enzymes to break down PET products into their basic monomers. Usually, biological recycling relies on either microorganisms or synthetic enzymatic reactions [29]. The biorecycling of PET employs enzymes as biological catalysts to depolymerize the polymer into its constituent monomers [30]. The process generally consists of several key steps: collection and sorting, size reduction, enzymatic depolymerization, purification, and repolymerization. During collection and sorting, PET waste is separated according to color, type, and quality. Size reduction then involves shredding the waste into smaller fragments to increase surface area and enhance enzymatic degradation. Enzymatic depolymerization yields a mixture of monomers and oligomers, which are subsequently separated and purified. Finally, the purified monomers undergo polymerization to produce new PET, which can be used to manufacture bottles, films, fibers, and other products [31]. This process offers several benefits over traditional recycling methods and has the potential to help build a more sustainable future.
From a resource management perspective, biocatalytic PET recycling is more than just a technological innovation—it is a strategic approach for conserving resources and promoting a circular economy. Global PET production exceeds 30 million metric tons annually, using large amounts of petroleum-based feedstocks and creating significant post-consumer waste. By enzymatically transforming discarded plastics and textiles back into high-purity monomers like TPA and MEG, biocatalytic processes directly close the resource loop, decreasing reliance on virgin fossil materials and reducing waste buildup. Compared to traditional mechanical and chemical methods, enzymatic recycling operates under milder conditions, reducing energy use and greenhouse gas emissions. This shift turns PET waste from an environmental problem into a secondary raw material source, a renewable carbon supply that can be reused in production. Understanding biocatalytic PET recycling in this light supports broader goals of sustainable resource use, waste valorization, and low-carbon material management.
This review discusses current technologies for biological recycling of PET, focusing on its conversion and reuse to reduce CO2 emissions. Among the various PET recycling approaches, biological recycling offers an environmentally friendly solution, relying on biocatalytic reactions. Several strategies have been developed for biocatalytic PET recycling, including the use of microorganisms and enzymes. The review also covers the process of obtaining new biologically recycled materials from PET, as well as the potential implementation of biocatalytic recycling on a large industrial scale. A deeper understanding of these biocatalytic methods is essential for policymakers to identify the most effective strategies to address the significant challenge of carbon emissions and promote sustainable recycling practices.

2. Resources: Plastic and Textile Waste to PET

The environmental impact of the textile industry is rising as global demand for textiles grows. Interestingly, 99% of polyester, a common synthetic fiber made from petroleum, is produced from post-consumer PET bottles, which make up 54% of the global fiber market. PET, the third most common polymer, is widely used in packaging and beverage bottles because of its durability, transparency, and lightweight qualities [32,33]. Global manufacturing of PET is predicted to reach 34 billion metric tons by 2050. PET is produced via a condensation process that utilizes TPA and EG, both sourced from petroleum feedstocks [34]. Furthermore, PET fiber was used worldwide to produce filament yarns and staple fibers. PET staple fibers are primarily used in the production of cotton-PET blended textiles; however, recycling these blended fabrics poses challenges due to the complex segregation of the mixed ingredients. Consequently, these textiles are predominantly discarded in landfills, resulting in environmental issues [35].
Converting plastic waste into PET usually starts with collecting and separating post-consumer plastics. Since waste streams often contain mixed polymers, proper sorting is essential to remove contaminants such as PVC, which can greatly reduce product quality during processing [36]. After sorting, size reduction methods such as shredding or cutting are used to increase the surface area, making subsequent cleaning and processing easier. Washing then removes labels, dirt, and organic residues, resulting in clean PET flakes ready for reprocessing. These initial steps are crucial for producing high-quality recycled PET, as impurities can harm the mechanical and thermal properties of the final product.
Once prepared, PET flakes can be melted down again and shaped into pellets or granules, which serve as raw materials for making new products. In mechanical recycling, these pellets are directly used to produce bottles, fibers, or films. In contrast, during chemical recycling, PET can be broken down into its monomers, such as TPA and MEG. These monomers can then be rebuilt into virgin-quality PET [37]. Such conversion routes not only extend the material life cycle but also reduce reliance on fossil-based raw materials, reduce greenhouse gas emissions, and promote circular-economy goals by reintegrating plastic waste into the production chain.
In textiles, converting waste into PET involves both mechanical and chemical methods. Mechanical recycling typically involves processes such as shredding and carding, where textile waste, primarily polyester-rich fabrics, is broken down into fibers that can be respun into yarn [33]. The mechanical approach has limitations, especially for post-consumer waste with blended or low-quality fibers, which often produces weaker fibers that need to be mixed with virgin materials to regain strength and durability [9]. Despite these issues, mechanical recycling remains a cost-efficient method and is commonly used to create insulation, padding, or nonwoven products from textile waste.
On the other hand, chemical recycling provides a more efficient pathway to recover PET from textile waste. This method depolymerizes polyester fibers into monomers such as PTA and MEG, which can then be purified and repolymerized into high-quality PET. Techniques like glycolysis, hydrolysis, and enzymatic hydrolysis are frequently applied, and advanced approaches such as using ionic liquids or subcritical water treatment are being developed to handle mixed fibers (e.g., polyester–cotton blends) [38,39]. Compared to mechanical methods, chemical recycling can regenerate PET with properties comparable to virgin plastic, creating new fibers, packaging materials, or even textiles [11]. This makes chemical recycling a key strategy for closing the loop in the textile-to-PET circular economy.
After plastic and textile waste are converted into PET, the material becomes suitable for biocatalytic degradation through enzymatic processes. In this approach, specific enzymes break down the PET polymer chains into their monomeric building blocks, primarily TPA and MEG [40]. This enzymatic degradation is a sustainable alternative to conventional recycling methods, as it operates under mild conditions, reduces energy consumption, and minimizes secondary pollution. The resulting monomers can be purified and repolymerized to produce new PET with properties comparable to virgin material, thereby closing the loop in a circular economy framework. In some cases, biodegradation can even occur directly on plastic or textile waste without prior conversion [10,41].

3. PET Characteristics and Recycling Context

Understanding the physical and chemical properties of PET is fundamental to optimizing recycling performance. The polymer’s high crystallinity and chemical resistance confer durability during use but hinder degradability post-consumption. These characteristics define PET as both an engineering success and a recycling challenge. In resource terms, the same molecular stability that ensures material longevity also creates kinetic barriers to depolymerization. Recognizing this paradox underscores the necessity for advanced recycling methods capable of reclaiming PET’s intrinsic carbon value.
PET’s lightweight, strong, and flexible characteristics make it a crucial material in today’s society [40,42]. Post-consumer and post-industrial PET waste differ from virgin PET due to their sources and contamination levels. Post-industrial waste from production is clean and readily recyclable, but post-consumer waste from consumer items might be polluted and difficult to recycle. Post-consumer PET waste contains soil, polymers, glue, labels, and small-molecule pollutants from bottles, packaging, and other uses [43]. PET is a semicrystalline polymer. The degree of crystallinity is determined by numerous criteria, including the molecular weight distribution, molecular weight, and crystallization temperature [44]. The crystallization rate of PET within the temperature range of 150–180 °C is based on the degree of chain orientation, molecular weight, and the characteristics of the polymerization catalyst. It is utilized in the process of producing PET [26,45]. Crystallinity is produced by heating until the glass transition temperature (Tg) is attained, which is followed by molecular orientation [46]. In addition to that, the crystalline microstructures demonstrate considerable resistance to the process diffusion of enzyme and water, which serves as a key component as a catalyst and reactant to biocatalytic depolymerization, respectively [47,48].
The melting temperature (Tm) of commercial PET is between 255 and 265 °C, with more crystalline PET reaching 265 °C. Virgin PET’s Tg ranges from 67 to 140 °C [26]. Furthermore, the proximate and ultimate analyses of PET for volatile matter, moisture, and ash were 88.2%, 0.5%, and 11.2%, respectively. On the other hand, the percentages of the ultimate analysis, such as carbon, hydrogen, nitrogen, and oxygen, are 64.2, 4.4, 0, and 31.4 wt%, respectively [49]. Furthermore, rheological studies are limited by crystallization at low temperatures and polymer breakdown at higher temperatures. PET rheological tests are typically performed at temperatures ranging from 260 to 290–300 °C. Above the latter temperature, rapid degradation occurs, limiting the practicality of the experiments [44]. In addition, the surface morphology of crystalline PET was studied by Daggubati. It compares SEM images of virgin and recycled PET to demonstrate the similarities in surface morphology between the two. Surface roughness in recycled PET is shown as holes, flakes, cracks, and pitting [50].
Kinetic studies of solid-state PET processes provide essential knowledge that can be used. Kinetic analysis aims to precisely predict the thermal behavior of a process under multiple circumstances that differ from those utilized in the PET studies [51]. In addition, Dubdud investigated various methods to determine the activation energy of PET. For instance, the activation energies of PET using Starnik, FWO, KAS, and Friedman were reported as 139.2, 142.8, 138.8, and 167.5 kJ/mol, respectively [49]. Furthermore, the physical and chemical properties were reached by thermogravimetric analysis, and derivative thermogravimetric analysis demonstrates that the thermal decomposition started at 370 °C, with the temperature maximum decomposition up to 441 °C, and the end of the degradation process of PET is 502 °C, with a heating rate of 10 °C/min [52].
Furthermore, the comparison with mechanical, chemical, and biological recycling approaches to the recycling of PET is listed in Table 1. The mechanical process is conducted by several methods, such as separation process, crushing, heat, and extraction. The waste from PET bottles/containers is melted to produce recycled PET pellets. On the other hand, the chemical process of PET recycling uses chemical and/or thermochemical methods to degrade PET waste into its constituent monomers, oligomers, and a mixture of solid, liquid, and gaseous hydrocarbons [53]. Furthermore, Biocatalysts and microorganisms capable of degrading plastics have received significant attention from the scientific community and the media, with news outlets presenting “plastic-eating” microorganisms and enzymes as potential solutions to the worldwide plastic crisis [54]. Enzymes facilitate rapid and selective chemical transformations in biological catalysis; however, they often exhibit limited tolerance to extreme processing conditions and additives, as their activity depends heavily on maintaining protein structure stability [55].
Table 1. Comparison of various technologies for the recycling of PET.
Table 1. Comparison of various technologies for the recycling of PET.
ApproachMechanical RecyclingChemical RecyclingBiological RecyclingRef
Technology recycling of PET
  • Injection molding
  • Extrusion molding
  • Inflation molding
  • Blow molding
  • Vacuum molding
-
Hydrolysis
-
Methanolysis
-
Ammonolysis
-
Glycolysis
-
Aminolysis
-
Methanolysis–Hydrolysis
-
Glycolysis–Methanolysis
-
Steam hydrolysis
Various microorganism is used to recycle PET[26]
ProductNew PET bottleDimethyl terephthalate, Bis(2-hydroxyethyl) terephthalate, TPA, MEG, and oligomersDimethyl terephthalate, Bis(2-hydroxyethyl) terephthalate, TPA, MEG[26]
Rate of degradation High speedHigh speedModerate[56]
Separation processApplicableComplicated to separate the product Applicable[56]
Temperature/PressureHighHigh Moderate[26]
Greenhouse gasLowHighModerate
Advantages
-
Cost-effective
-
High yield
-
Simple method
-
Raw materials flexibility
-
Environmentally friendly
-
High yield
[26,57]
Drawbacks
-
Cleaning PET waste is a necessary initial step to eliminate contaminants.
-
Poor mechanical stress of PET
-
The life cycle of recycled PET will affect its quality
-
High production cost
-
High corrosiveness for some chemical recycling of PET
-
Less environmentally friendly
-
Hazardous volatile organic substances for by-products
-
High consumption of energy
-
Long reaction time
-
Enzyme cost
[57,58]
PET recycling routes can be categorized into mechanical, chemical, and biocatalytic pathways. Mechanical recycling remains the most accessible yet offers limited resource efficiency, as material properties deteriorate after repeated processing. Chemical recycling restores monomer quality but consumes high energy and often requires toxic reagents. Biocatalytic recycling, in contrast, converts PET waste into reusable monomers under mild conditions, enabling circular resource recovery at lower environmental cost. This progression from mechanical to biocatalytic routes reflects an evolving strategy—from waste disposal to resource regeneration.

4. Classes of PET-Hydrolyzing Enzymes

In the beginning, polyester hydrolases were discovered using traditional microbiological methods, such as isolating, cultivating, and enriching microbial strains capable of hydrolyzing natural polymers [48]. The following research resulted in the identification of bacterial and fungal species, which included those from Thermobifida and Fusarium species. It is found to degrade the synthetic polyesters, such as PET [59]. Furthermore, biocatalytic chemical recycling has the potential to revolutionize the commercial recycling industry by preserving the valuable properties of waste materials while remaining energy-efficient [60]. Enzyme-based PET depolymerization is becoming more popular as a possible alternative to other chemical recycling methods. This is mainly because (a) it requires lower temperatures, which use less energy; (b) it avoids the need for potentially hazardous substances like highly polar solvents and strong acids; and (c) enzymes have an excellent selectivity for polymer ester linkages, which helps recover monomers and allows the technique to be used with mixed materials, thus eliminating the need for difficult plastic sorting [61].
PET hydrolase (PETase) hydrolyzes PET via a classical serine-hydrolase mechanism, as shown in Figure 1 [47]. PETase uses a Ser–His–Asp catalytic triad, in which the activated serine attacks the ester bond in PET to form a tetrahedral intermediate and then an acyl–enzyme intermediate, releasing one fragment of the polymer. Next, a water molecule is activated by the His–Asp pair and attacks the acyl–enzyme, generating a second tetrahedral intermediate that collapses to release the carboxylate product and regenerate free serine. By repeating this two-step acylation–deacylation cycle along the polymer chain, PETase progressively depolymerizes PET into soluble oligomers and monomers such as MHET, TPA, and EG.
Currently, two types of enzymes have been identified that act on PET: (1) enzymes that attack the polymer surface, increasing its surface hydrophilicity, but having little or no effect on the polymer shape. The next category may include enzymes like nonspecific esterases, serine proteases, lipases, and certain genuine PETases. (2) This category includes genuine PET hydrolases, which are often derived from thermostable cutinases, as well as thermostable enzymes from compost metagenome libraries and variations of genuine PETase. Some of these enzymes could be useful for bio-recycling because they break down the inner blocks of PET and alter its structure [62]. Furthermore, PET can be degraded with several enzymes, mostly hydrolases, including PET hydrolase (PETase—EC 3.1.1.101, carboxyl ester hydrolases, and EC 3.1.1.2, arylesterase), cutinases (EC 3.1.1.74), carboxylesterases (EC 3.1.1.1), and lipases (triacylglycerol lipase, EC 3.1.1.3) [63,64]. Cutinases have a shallow, open active site on their surface, allowing them to hydrolyze water-insoluble, hydrophobic polyesters. In contrast, lipases and esterases feature an active site within a tunnel-like structure, enabling vertical hydrolysis. Research has progressed from screening natural strains to engineering highly active enzymes with improved thermostability and substrate affinity. This shift marks a move from environmental observation to intentional design for resource recovery.

4.1. Cutinase

Cutin is the insoluble, waxy polymer that is an essential part of the cuticle, which serves as a structural element in plants. It contains 12-monounsaturated 9,10,18-trihydroxy C18 acids, dihydroxy palmitic acids, fatty acids, and 12-monounsaturated 18-hydroxy 9,10-epoxy C18 acids. It shows a lesser quantity of phenolic groups compared to its similar, suberin, which exists in potatoes [65]. The initial enzyme identified for the degradation of PET fibers was a cutinase derived from Thermobifida fusca, which also possesses significant industrial potential for PET enzymatic degradation [66]. Furthermore, Cutinases (EC 3.1.1.74) are the only enzymes known to break down the inner block of PET films; nevertheless, lipases, esterases, and cutinases can also modify their surface [64,67].
Nevertheless, Cutinases are serine hydrolases found in numerous species, for instance Fusarium oxysporum, Saccharomonospora viridis, Thermobifida alba, Thermobifida cellulosilytica, Fusarium solani pisi, Thermobifida fusca, Humicola insolens, Kineococcus radiotolerans, Moniliophthora roreri [68,69]. These enzymes can digest cutin, a fatty acid polyester that forms the main part of the plant cuticle. Enzymes like these have become essential biocatalysts in breaking down synthetic polymers such as PET [70]. Compared to other enzymes that hydrolyze PET, cutinases can effectively break the ester bonds in PET because of the structural similarities between cutin and PET. Table 2 shows various comparisons of the degradation of PET by cutinase. In addition, cutinases have a greater attraction for esters of short-chain to medium-chain fatty acids [71]. The study by Won et al. focused on a cutinase enzyme found in Rhodococcus strains from the Ross Sea. It demonstrated outstanding activity under alkaline conditions. Additionally, it is possible to slowly break down PET with MEG-terephthalate ester bonds [72].

4.2. Carboxylesterases

Carboxylesterases (EC 3.1.1. x) have become common enzymes that exist in several species, including some viruses [73]. Carboxylesterases are a broad distribution of hydrolases that catalyze the breaking of ester bonds in a wide range of substrates. They successfully hydrolyze a wide range of molecules with particular functional groups, such as carboxylic acid esters, amides, and thioesters [74,75]. It follows that the taxonomy and categorization of bacterial carboxylesterases have been continuously revised and updated. Hitch and Clavel created a new classification system for bacterial lipolytic enzymes, dividing them into 35 families and 11 authentic lipase subfamilies based on their sequences and conserved features. The updated classification method, which relates to sequences for each category, was used for the study and categorization of new target esterases [76].
Previous research described the depolymerization of PET by carboxylesterases derived from Bacillota, namely the p-nitrobenzyl esterase (BsEstB) from Bacillus subtilis and the esterase (Cbotu_EstA) from Clostridium botulinum [60]. These types of enzymes, which are greater than cutinases, have a temperature optimum of 40 °C and demonstrate a minimal degree of PET hydrolysis. Consequently, they are classified as PET surface-modifying enzymes as compared to PET hydrolases [31]. In addition, carboxylesterases from I. sakaiensis and Thermobifida fusca KW3, identified as MHETase and TFCa, respectively, were demonstrated to assist in PET depolymerization by IsPETase and T. fusca cutinase. When degradation is conducted at elevated temperatures, inhibition mediated by MHET is alleviated [77]. Furthermore, terephthalate esters are generally deconstructed using two hydrolytic processes by carboxylesterases Figure 2; (1) PET is initially converted to mono(2-hydroxyethyl) terephthalic acid (MHET) by an aromatic polyester breaking down hydrolase, for instance PET digesting enzyme—also known as PETase, (2) MHETase, a second enzyme, breaks down MHET into TPA and MEG [78]. The combination of microorganisms and enzymes can enhance the depolymerization process of PET microplastics under elevated temperature [79].
Table 2. Comparison of the depolymerization of PET through Cutinase.
Table 2. Comparison of the depolymerization of PET through Cutinase.
SubstrateType of EnzymeSourceCondition OperationsPET Degradation, %Ref
T, °Ct, hpH
PET packageLC-cutinaseMetagenome from leaf branch compost70248≤25[80]
Amorphous PET filmVariant of
TfCut2
T. f usca KW360248≤25[81]
Amorphized and micronized PETLC-cutinase
variant
Metagenome from leaf branch compost7210990[82]
PET filmCut190**SSEscherichia coli70488.610.1[83]
Amorphous PET filmThermobifida fusca cutinase TfCut2B. subtilis strain RH 114967096850[84]
Textile PET fibresCutinase ICCGDAQIE. coli BL21 (DE3)7014997[85]
PET-GF Variant of
Cut190
Saccharomonospora viridis
AHK 90
63NA±833.6 ± 3.0[86]
PET-GF filmHiC (Novo)Humicola insolens70966.5–9.597 ± 3[87]
PETThcCut1-G63A/F210I/D205C
/E254C/Q93G (ThcCut1-AICCG)
NA7096NA96.2[68]
Post-industrial PET fibres Cutinase ICCGDAQIE. coli7024997.8[88]
Post-consumer PET bottlesCutinase Est1_5MThermobifida alba AHK1196524–72890.8[89]
PET fabricsTfu_0883Thermobififida fusca60487.550[90]
** thermostable mutant of cutinase.

4.3. Lipases

Lipases are triacylglycerol acylhydrolases (E.C.3.1.1.3) that are classified within the hydrolases category. In addition to that, Extracellular triacylglycerol acyl hydrolases are known as lipases, which could be a trigger as a catalyst for the hydrolysis of long-chain triglycerides into glycerol and fatty acids [91]. Lipases demonstrate unique properties that make them especially interesting for synthetic applications, such as versatility, stability, recyclability, extensive tolerance to hydrophobic substrates, activity in non-aqueous environments, and the capacity to operate independently of cofactors [92].
In addition to that, Lipases are water-soluble enzymes due to a lid domain that blocks the hydrophobic active site. Numerous fungal species produce lipases and can depolymerize PET, such as Humicola sp., Thermomyces lanuginosus, Candida sp., Pseudomonas sp., and Thermobidifida fusca [91,93]. At the same time, lipase (TfH) was mainly purified and isolated from Thermobifida fusca. It is possible to hydrolyze PET films by 40–50% within three weeks. Meanwhile, lipase has been purified from Thermomyces lanuginosus, demonstrating its ability to hydrolyze PET [94,95,96]. Furthermore, Lipase B (CALB) derived from yeast Candida antarctica, previously classified as Pseudozyma antarctica, is recognized for its exceptional selectivity and catalytic efficacy. The function of CALB resembles the activities of IsPETase and IsMHETase [97]. The kinetic parameters of PET degradation through several enzymes are shown in Table 3. Swiderek et al. studied the process of hydrolysis of PET oligomers catalyzed by the promiscuous lipase B from Candida antarctica using molecular dynamics simulations validated by experimental Michaelis-Menten kinetics. The selectivity of the CALB substrate in relation to the pH value. The kcat values for BHET hydrolysis remain similar regardless of the pH of the reaction. However, the kcat values for MHET hydrolysis decrease as the pH increases. The kM of CALB toward both substrates, BHET and MHET, follows the same trajectory with regard to substrate affinity: kM (pH 9); (61.0 mM ± 25.7 mM, and 100 mM) > (pH 5) (22.5 mM ± 9.6 mM, and 23.8 mM ± 8.8 mM) > (pH 7) (13.3 mM ± 4.3 mM, and 15.1 mM ± 3.2 mM), respectively [98]. Furthermore, the kinetic study of various cutinases is shown in Table 3. Acero et al. studied the kinetic properties of the cutinases from Thermobifida. The Thermobifida cellulosilytica DSM44535 (Thc_Cut1 and Thc_Cut2) and Thermobifida fusca DSM44342 (Thf42_Cut1) enzymatically hydrolyze PET. The three different Thermobifida cutinases exhibit significant similarity, with a maximum of 18 amino acid discrepancies; nonetheless, they have distinct kinetic characteristics on soluble substrates. Their kcat and Km values for pNP–acetate ranged from 2.4 to 211.9 s−1 and 127 to 200 μM, respectively, whereas for pNP–butyrate, kcat and Km values ranged from 5.3 to 195.1 s−1 and 1483 to 2133 μM, respectively. Thc_Cut1 generated the greatest amount of TPA and MHET from PET [59].
Furthermore, the PET macromolecular polymers cannot directly penetrate microbial cells; hence, the first breakdown of PET into monomeric products predominantly depends on extracellular enzymes released by the host microorganisms. As seen in Figure 2, lipases can utilize a variety of enzymatic systems to biodegrade PET into smaller, water-soluble monomeric compounds, including MHET to become MEG, and TPA [99]. Moreover, using CALB demonstrated highly efficient hydrolysis processes and polymer depolymerization, resulting in the buildup of terephthalic acid. Additionally, CALB and Humicola insolens cutinase achieved full PET depolymerization, with a mole fraction reaching 0.88 and a 7.7-fold enhancement in PET yield, generating TPA [93,100]. Furthermore, the hydrolytic efficacy of lipases on PET can be determined of their substrate-binding pocket. This is due mainly to the lid structure that blocks the active site, as a result inhibiting enzymes from directly interacting with the macromolecular substrates [95,101]. Safdar et al. studied the biodepolymerization of PET by the extracellular lipase of Aspergillus niger—the synthesis of lipase from Aspergillus niger MG654699.1 using agro-industrial byproduct (wheat bran) by solid-state fermentation. The synthesized lipase exhibited an activity of 176.55 U/mL, with optimum conditions of 37 °C and pH 7.0. The 3.6% weight loss of PET was as a result of the biocatalytic activity of 30 KDa lipase [102].
Nevertheless, low temperatures will affect the lower specific volume of polymeric materials, raising the thermodynamic and kinetic barriers to geometrical change, resulting in limited substrate deformability [103,104]. Directed evolution and genome mining research investigations have discovered mutant PETases that demonstrate improved thermostabilities and catalytic efficiencies, such as PHL7, HOT-PETase, LCCICCG, LCCWCCG, and Fast-PETase [105]. Interestingly, in less than three days, one such designed PETase can degrade 90% of 20 kg of PET waste that has been processed using extrusion and micronization [106]. Furthermore, PETase serves as the enzyme responsible for the complex process of continually breaking down the PET polymeric chain into MHET, with small amounts of TPA and BHET as byproducts [107].
The kinetic data summarized in Table 3 reveal clear performance differences among the principal classes of PET-hydrolyzing enzymes. Cutinases and PETases exhibit the highest overall catalytic turnover and substrate affinity under typical experimental conditions (40–70 °C, pH 7–9), followed by MHETase, which complements the degradation pathway by converting the intermediate mono(2-hydroxyethyl) terephthalate (MHET) to terephthalic acid (TPA) and ethylene glycol (EG). Lipases and carboxylesterases, while capable of hydrolyzing short-chain model esters, show limited activity toward crystalline PET due to steric hindrance and restricted active-site access. These distinctions underscore that effective PET depolymerization relies on both enzyme selection and synergistic catalytic cascades.
The apparent catalytic constants (kcat) in Table 3 vary by two orders of magnitude across enzyme classes. Cutinases such as TfCut2 and HiC exhibit kcat values around 10–40 s−1, while wild-type PETase typically shows 2–10 s−1. Substrate crystallinity significantly affects apparent kinetic performance. Enzymes achieve nearly complete hydrolysis (>90%) for amorphous or pretreated PET films, whereas only 10–25% conversion is observed for highly crystalline materials under identical conditions. This discrepancy supports the hypothesis that chain mobility, not just surface adsorption, governs reaction kinetics. Pretreatment methods such as thermal amorphization, glycol-assisted swelling, or surface etching can reduce activation energy barriers and enhance kcat/km ratios across enzyme classes.
Table 3. Comparison of the Kinetic Depolymerization of PET by Lipase and Cutinase.
Table 3. Comparison of the Kinetic Depolymerization of PET by Lipase and Cutinase.
SubstrateCrystallinity, %Type of
Enzyme
SourceCondition
Operations
Kinetic Parameter *ProductRef
T, °Ct, hpHkcatkmR2
Post-consumer PET41.1Humicola insolens cutinaseNovozymes70967224.2 ± 34.0
(h−1)
0.0041 ± 0.0010
(L g−1)
0.981
TPA + MHET + BHET[108]
Bis(benzoyloxyethyl) terephthalate (3PET) with (p-nitrophenyl acetate (PNPA))NATha_Cut1Escherichia coli BL21-Gold(DE3)50NA72.72 ± 0.2
(h−1)
213 19 (μmol L−1)NAMHET, TPA, Benzoic acid, 2-hydroxyethyl benzoate[109]
Bis(benzoyloxyethyl) terephthalate (3PET) with (p-nitrophenyl butyrate (PNPB))NATha_Cut1E.coli BL21-Gold(DE3)50NA76.03 ± 0.59 (s−1)1933 ± 306 (μmol L−1)NAMHET, TPA, Benzoic acid, 2-hydroxyethyl benzoate[109]
PET nanoparticles9.8TfCut2Thermobifida fusca6018.5147.673 ± 4.928
(min−1)
0.010 ± 7.40 × 10−4
(mLcm−2)
0.998
BHET, MHET[110]
Amorphous PETNACut190**SSE. coli3724724.9
(s−1)
0.082 (mM) NANA[111]
PET with (p-nitrophenyl acetate (PNPA))37Thf42_Cut1Thermobifida fusca DSM4434250120739.5 ± 3.0
(s−1)
167 ± 29 (μmol L−1)NABHET, MHET, TPA, Benzoic acid, 2-hydroxyethyl benzoate[59]
PET with (p-nitrophenyl butyrate (PNPB))37Thf42_Cut1Thermobifida fusca DSM4434250120730.9 ± 8.6 (s−1)2100 ± 361 (μmol L−1)NABHET, MHET, TPA, Benzoic acid, 2-hydroxyethyl benzoate[59]
* kcat: specific activity, km: Michaelis constant. ** thermostable mutant of cutinase.

4.4. PETase

PETase (EC 3.1.1.101), is an enzyme known as PET hydrolase (PETase), is capable of converting PET into its monomers [112]. PET hydrolases can degrade PET, thereby offering an environmentally favorable solution to the buildup of PET in the environment [113]. A unique structural characteristic of PETase is its atypically shallow and solvent-exposed active-site cleft, with more hydrophobic pockets found in thermophilic homologs, such as leaf-branch compost cutinase (LCC) and Thermobifida fusca cutinase (TfCut2) [114]. In addition, PETase’s structural flexibility, activity at ambient temperature, and unique active-site structure make it an effective biocatalyst for the degradation of PET. Nonetheless, its drawbacks, especially regarding thermostability and efficiency in high-crystallinity PET, have prompted significant technological efforts to realize its full potential for environmental and industrial applications [114,115]. Meanwhile, PETase is heat-sensitive and will be inactivated at temperatures greater than 40 °C. At mild temperatures of 25−30 °C, PETase may lose its degrading activity after more than 24 h of incubation. PETase has lost efficiency for industrial applications due to the need for more extended incubation periods and elevated reaction temperatures of about 50−65 °C to increase the degradation rate [116,117]. The mesophilic PET depolymerase moves the degradation polymer of the system into the physiological temperature range, demonstrating a feasible strategy for the free biological transformation of PET into any product that an organism can generate [118].
Two α/β-hydrolase fold enzymes, PETase and MHETase, collaborate to degrade PET in a two-step process through MHET, resulting in TPA and MEG, which are the essential precursors for a new cycle of PET production [119], as seen in Figure 3. Moreover, determining the kinetic parameters of PETase is inherently difficult due to the insolubility and heterogeneity of PET [114], as listed in Table 4. Burgin et al. studied the kinetic reaction mechanism of the PETase enzyme. The reaction rate constant is predicted to be k = 0.82 ± 0.10 s−1. As well as the reaction rate constant, Erickson et al. utilized the Michaelis-Menten equation to determine that the kcat values for amorphous PET film and crystalline PET powder were 1.5 ± 0.5 and 0.8 ± 0.0 s−1, respectively, at 30 °C [120,121].
Table 4. Comparison of kinetic degradation of PET through PETase, IsPETase, and IsMHETase.
Table 4. Comparison of kinetic degradation of PET through PETase, IsPETase, and IsMHETase.
Raw MaterialsType of EnzymesOrganismCondition OperationsKinetic Parameter *ProductRef
T,
°C
t,
h
kcatkmk
Waste PET bottleIsPETaseE. coli BL21402411.59 ± 0.16 s−10.19 ± 0.01 mmol.L−1NATPA, BHET,[122]
Waste PET bottleIsPETaseW159H/F229YE. coli BL2140249.64 ± 0.130.08 ± 0.01NATPA, BHET,[122]
PET sheetLCCICCGEscherichia coli504.1NA0.11 ± 0.02 µMNANA[123]
Post-consumer recycled PET flakesLCCICCGE. coli BL2165484.66 g/(µmol⋅h)5.39 g/L0.036 h−1TPA, EG, 1.3 Propanediol, 1.4 Butanediol[124]
Post-consumer PET bottleICCG-GS4-αSPE. coli Shuffle T7 and E. coli BL21 (DE3)60243.91 ± 0.81 s−15.8 ± 1.2 g/LNATPA, BHET, MHET[125]
PET FilmPETasePolyester hydrolase TfCut26010.31 ± 0.01 s−17.33 × 10−3 ± 3.62 × 10−4 mol L−1147.673 ± 4.928 min−1MHET, BHET[110]
Semi-crystalline PET powderHiC [AAE13316.1] andHumicola insolens and5050.088 s−10.27 g/LNABETEB, BHET, MHET[126]
Semi-crystalline PET powderTfC [AAZ54921.1]Thermobifida fusca5050.015 s−11.2 g/LNABETEB, BHET, MHET[126]
* kcat: specific activity, km: Michaelis constant, k: enzyme reaction stability coefficient.

4.5. IsPETase

IsPETase is a desirable option for a biocatalyst in PET recycling due to its unique catalytic characteristics. IsPETase was initially discovered in a novel microbial consortium named Ideonella sakaiensis 201-F6, which develops on PET film, as reported by Yoshida et al. [106]. IsPETase is a functional monomer which has a molecular mass of 30.1 kDa, consisting of 290 amino acid residues [127]. Due to its ease of cultivation and high yield efficiency, Escherichia coli (E. coli) is the most widely utilized host for the synthesis of PET hydrolases (such as BhrPETase, IsPETase, and their derivatives) [128] as shown in Table 4. According to structural studies, IsPETase and other previously identified PET-hydrolyzing enzymes, such as Cut190 [129]. Furthermore, the structural analyses of IsPETase have demonstrated that the enzyme’s mechanism of action is distinctive, exhibiting more activity on PET films compared to other hydrolases/esterases, and it can effectively use extremely massive and hydrophobic polymers [130]. It is believed that the interesting structural variations within homologous cutinases and IsPETase are responsible for this enhanced activity. The structural analyses indicate that IsPETase is a member of the α/β-hydrolase superfamily, consisting of a canonical α/β-hydrolase fold of a central β-sheet surrounded by α-helices [131]. Currently, it was recognized mechanistic hypothesis follows to the conventional serine hydrolase mechanism, which consists of a Ser–His–Asp (Ser160–His237–Asp206), catalytic triad stabilized by a two-residue oxyanion hole formed by the backbones of residues Tyr87 and Met161 [132]. Nevertheless, the wild-type enzyme is limited in terms of thermostability. Biotransformations must be conducted at ambient temperatures much below the Tg of PET, which causes problems with depolymerization rates [133]. In addition, IsPETase promotes the degradation of PET into the primary product MHET and the byproducts TPA, MEG, and BHET as shown in Figure 3 [134]. On the other hand, IsMHET can be hydrolyzed to MHET to become MEG and TPA [106].

4.6. IsMHETase

IsMHETase obtained from the MHETase Gram-negative bacteria from I. sakaiensis. It successfully depolymerized polymeric chains, yielding ecologically harmless substances such as carbon dioxide and water [100]. It was discovered that the catalytic machinery of IsMHETase and IsPETase was identical, as shown in Figure 3. There are four basic steps for both IsMHETase and IsPETase. (i) Nucleophilic attack introduced by Ser-His-Asp, (ii) The cleavage of the C−O bond, (iii) The nucleophilic assault by water molecules, and (iv) Deacylation of IsPETase/IsMHETase [135]. Furthermore, phylogenetic studies had identified IsMHETase as a member of the tannase enzyme family, consisting of fungal and bacterial tannases, along with feruloyl esterases [136]. Duan et al. demonstrated the IsMHETase’s lack of activity towards pNP-aliphatic esters and conventional aromatic ester compounds that can be catalyzed by enzymes belonging to the tannase family [137]. In addition to that, the initial stage of degradation involves the depolymerization of PET film, yielding the primary product MHET along with byproducts BHET, TPA, and MEG, facilitated by the IsPETase enzyme. IsMHETase afterwards continues to the second phase of degradation by hydrolyzing MHET to generate TPA and MEG [138].
Overall, Table 4 illustrates that achieving high depolymerization efficiency depends on balancing three factors: enzyme thermostability, substrate morphology, and controlled reaction chemistry. Systems that align these parameters not only exhibit superior kinetic constants but also demonstrate improved resource efficiency, as higher conversion at moderate energy input translates directly into reduced greenhouse-gas emissions and lower process costs. These findings reinforce that rational enzyme engineering and optimized pretreatment are complementary levers for advancing scalable, low-carbon PET recycling. Each enzyme has unique catalytic mechanisms and operational ranges. Cutinases demonstrate broad substrate tolerance and high activity against amorphous PET, while PETases are optimized for mild-temperature conditions. MHETase works with PETase by preventing the buildup of intermediates. Lipases and esterases assist in hydrolysis but are less effective alone. Comparative studies show that engineered cutinases and PETases surpass natural versions in both turnover rate and stability. These insights guide future enzyme development for sustainable PET recycling.

5. The Implementation of Biocatalytic PET for Large-Scale Industrial Applications

5.1. Process Engineering and Reactor Design

The process of biocatalytic PET with larger-scale production begins with melt-extrusion-based amorphization. For efficient biocatalytic depolymerization, the heated polymer extrudates must be quickly quenched in cold water or another solvent to achieve a low crystallinity, ideally less than 15% [139]. Nevertheless, the commercial PET has a large crystallinity of up to 30%. Therefore, it is crucial to have the pre-treatment process, such as micronization and extrusion, to enhance the amorphous region [140]. Furthermore, the process flow diagram for biocatalytic PET for large–scale industrial use is shown in Figure 4. After that, the amorphized PET generally undergoes a mechanical size-reduction process, which might include ambient temperature crushing or cryogenic grinding [48]. Brizendine et al. studied the effect of particle size reduction of PET during biocatalytic depolymerization. The reduction in the particle size of PET enhanced the maximum reaction rate for high crystallinity of PET. In contrast, the maximum hydrolysis rate for low crystallinity cryomilled PET demonstrated no significant changes across different particle sizes for every single substrate. Nonetheless, it indicates that the reduction in particle size has minimal impact on the total amount of conversion. The low-crystallinity cryomilled film achieved a mass loss of 99% within 48 h, whereas the high-crystallinity of PET powder attained only 23.5% conversion over 144 h [141].
The commercial potential of biocatalytic PET recycling has been previously explored. A dual enzyme reactor, combining a polyester hydrolase and a carboxyl esterase, could efficiently hydrolyze amorphous PET films. By converting the hydrolysis products of polyester hydrolase, MHET and BHET, into TPA and EG, the carboxyl esterase increases the rate of PET depolymerization. Additionally, a polyester hydrolase can effectively break down PET films in an enzyme reactor, with an ultrafiltration membrane continuously removing the hydrolysis byproducts [84,142]. Furthermore, Lu et al. studied the depolymerization of PET by using a rotating packed bed reactor. A rotating packed bed reactor may offer superior mass transfer and diffusion efficacy. Nevertheless, the shear strength intolerance of the enzyme prevents the direct application of a rotating packed bed reactor for the biocatalytic depolymerization process. A PET particle depolymerization rate up to 95% was attained under optimal circumstances after 24 h, representing a 30% enhancement compared to the rate obtained in a stirred tank reactor [143].
In addition, immobilizing the biocatalyst offers well-proven advantages, for instance, simplified product/biocatalyst separation, continuous processing, and, in some cases, enhanced stability [144]. Therefore, using immobilized biocatalysts in two-liquid-phase systems is an exciting method to combine two innovative technologies. Immobilization can increase the biocatalyst’s stability in these systems by protecting it from the phase-reducing effects of the organic solvent. However, issues may arise from unwanted interactions between the solvent and the immobilization support material. The type of support material influences both the activity and stability of the biocatalyst, as well as the hydrophilicity-hydrophobicity balance of natural or synthetic gels. The choice of solvent affects how substrates and products are distributed between the biocatalyst and the liquid phases [145,146].
Using the evaporative crystallizer process, several methods can enhance the separation of biocatalytic PET, including the TPA crystallizer. In this process, the solid product is TPA, and the remaining liquid product will continue to the membrane unit to remove the by-products. An ultrafiltration membrane can maintain insoluble substrate components and the enzyme within the reactor’s recycling loop, while allowing soluble reaction products to permeate the membrane and be separated from the reaction mixture [147,148]. Ultrafiltration technologies have effectively reduced product inhibition in enzymatic hydrolysis operations by eliminating soluble products and sustaining high hydrolysis yields [149]. Furthermore, the process will proceed in the salt crystallizer to separate the sodium sulfate and MEG. Then, the MEG and other byproducts will be separated at the distillation column.

5.2. Techno-Economic and Environmental Considerations

When assessing the viability of PET recycling strategies, it is crucial to not only analyze catalytic or mechanistic performance but also evaluate process-level economics, energy use, greenhouse gas emissions, and life-cycle environmental impacts. This section offers a comparative analysis of mechanical, chemical, and biocatalytic recycling pathways from techno-economic and environmental perspectives, highlights major cost or energy factors, and discusses the challenges and opportunities for improving resource efficiency in each approach.
(i)
Mechanical Recycling: Lowest Energy & Cost but Quality Loss and Limited Cycles
Mechanical recycling of PET is currently the most mature and widely implemented pathway. Because it avoids the need to break chemical bonds, its energy consumption and capital cost tend to be lower than those of more intensive methods. Several life-cycle assessment (LCA) studies indicate that mechanical recycling often yields lower environmental burdens (e.g., global warming potential, energy demand) than virgin PET production or more aggressive chemical routes, particularly when the recycled product can substitute for virgin resin [150].
However, mechanical recycling has inherent limitations: the recycled PET often suffers from degradation in molecular weight, coloration, contamination, and limited cycles of reuse. Over multiple reuse cycles, material properties deteriorate, constraining their application in high-value uses. Moreover, mechanical recycling is sensitive to feedstock purity; mixed, colored, or contaminated PET streams often cannot be recycled reliably by mechanical means and may be diverted to lower-value uses or incineration [40].
From a cost perspective, mechanical recycling typically features lower capital expenditures (CAPEX) and operating expenditures (OPEX), but its profitability is constrained by the market value of the recycled resin and the costs of sorting, washing, and quality control. Some techno-economic analyses suggest that in favorable local energy and logistics conditions, mechanical recycling can deliver recycled PET at a price comparable to virgin resin [151]. Mechanical recycling is currently the most resource-efficient route in many contexts (especially for clean PET streams), but its limitations in feedstock flexibility and property retention mean it cannot fully absorb all PET waste, particularly low-quality or mixed waste streams.
(ii)
Chemical Recycling: Flexibility and Quality at Cost of Energy Intensity
Chemical recycling (also called feedstock recycling or depolymerization) encompasses a variety of processes such as glycolysis, methanolysis, hydrolysis, pyrolysis, and solvolysis that break PET down into monomeric or oligomeric intermediates for repolymerization or chemical valorization. Its key advantage is that the recovered monomers can approach virgin-equivalent purity, offering higher substitutability and broader reuse scopes than mechanically recycled PET [150].
However, the trade-off lies in energy consumption, reagent costs, and emissions associated with breaking strong ester bonds. For example, Uekert et al. (2023) report that glycolysis-based chemical recycling and mechanical recycling show the best economic performance across many scenarios, with up to 9–73% lower cost than some alternative routes, and environmental impact reductions of 7–88% relative to baseline waste treatments (depending on system boundaries and assumptions) [152]. Nevertheless, in many LCAs chemical recycling still underperforms compared to mechanical recycling unless heat integration, renewable energy inputs, and efficient system design are implemented [153].
A further complication arises from the capital and operational costs associated with reactors, separation, purification, and catalyst recycling. The costs of chemical reagents (e.g., glycol, methanol, acids or bases), heating, and product recovery (distillation, separation) often dominate the cost structure, especially for smaller-scale plants or fragmented feedstocks. Moreover, the environmental footprint is sensitive to the quality of the feedstock (e.g., presence of additives, dyes, contaminants) and the efficiency of purification [26,154].
In comparison to mechanical recycling, chemical recycling offers higher flexibility and product quality, but at the expense of higher energy and capital demands. To make chemical recycling competitive, innovations in catalyst efficiency, heat recovery, process intensification, and integration with waste heat or renewable energy sources are essential.
(iii)
Biocatalytic Recycling: Promising Resource Efficiency, But High Uncertainty & Scale Challenges
Biocatalytic (enzymatic) PET recycling seeks to combine the resource attributes of chemical recycling (monomer-level recovery) with milder operating conditions and potentially lower energy input. Techno-economic and life-cycle studies to date suggest substantial upside potential, but also significant uncertainties and scaling challenges.
From the techno-economic perspective, Singh et al. (2021) built a process model for enzymatic PET depolymerization and estimated a minimum selling price for recycled TPA of about USD 1.93/kg under base-case assumptions, assuming effective enzyme reuse and optimized downstream purification [155]. Their model suggests that with further process optimization, especially reducing energy demand in pretreatment and downstream separation, enzymatic recycling can become cost-competitive with virgin production. On the environmental side, Singh et al. also report potential reductions in supply-chain energy consumption of 69–83% and greenhouse gas emissions of 17–43% per kg of TPA, compared to virgin TPA production [155].
However, more recent LCA work reveals a more nuanced picture. Uekert et al. (2022) conducted a detailed LCA of enzymatic PET recycling. They found that, in many impact categories and under current process assumptions, enzymatic recycling still performs worse than virgin PET production and chemical recycling—primarily due to high energy and chemical inputs in pretreatment, enzyme production, and product recovery steps [156]. Their sensitivity analysis shows that improved yields, elimination of harsh pre-treatment steps, and optimized pH control are crucial to bring enzymatic recycling’s environmental footprint down to parity with conventional routes [156].
Key cost and energy drivers include feedstock preparation (shredding, amorphization, micronization), buffer control and pH maintenance, enzyme production and recovery, purification of monomers (crystallization, filtration, distillation), and utilities (heating, cooling, electricity). In many process models, pretreatment and separation contribute the largest fraction of the energy and cost burden, limiting net benefits [156].
Despite these challenges, enzymatic recycling has attractive potential in certain niches: for low-quality, mixed, or contaminated PET streams not suitable for mechanical recycling; for modular or decentralized plants; and in systems with access to low-carbon electricity. Ongoing enzyme engineering (e.g., more thermostable and active variants) and process integration (e.g., combining biocatalysis with mild chemical steps) may help push the economics and environmental metrics toward competitiveness.

6. Limitations of Biocatalytic PET

Despite rapid progress, the scalability of biocatalytic PET recycling remains constrained by three key factors: (i) crystallinity-induced mass transfer barriers, (ii) enzyme stability under prolonged operation, and (iii) product inhibition by intermediates. Addressing these challenges requires both molecular-level and reactor-level strategies.
Overcoming the limitations of PET crystallinity and mass transfer is essential for large-scale industrial applications [157]. Table 5 shows the parameters and constraints of several biocatalysts for degrading PET. Solid porous supports may affect the efficacy of PET degradation, mainly because most proteins live within porous structures. Consequently, only enzymes exposed on the surface will initially engage with solid PET until the breakdown products are sufficiently small to pass through the pores. Moreover, enzymes attached to the carrier surface may exhibit inadequate interaction with the macromolecular PET substrate because of rigidification effects, which may result in considerable rate limitations related to the enzymes utilized in soluble form [85,158]. Nevertheless, the enzymatic depolymerization of amorphous PET has been significantly enhanced as the temperature increases, especially when approaching temperatures near the Tg of PET. The range of temperatures for depolymerization is typically 60–80 °C, depending on the substrate condition [123].
Table 5. The parameters and limitations of several biocatalysts for degrading PET.
Table 5. The parameters and limitations of several biocatalysts for degrading PET.
BiocatalystParametersLimitationsRef
Cutinase
  • The range of optimal temperature and pH for biocatalytic depolymerization is 60–70 °C and 8.0–9.0, respectively.
The smooth hydrophobic surface of PET inhibits the binding of cutinase to PET, consequently limiting the rate of the PET degradation process.[159]
Carboxylesterases
  • The optimal temperature and pH degradation of biocatalytic are 50–70 °C and 8.0–9.0, respectively.
For applications in industrial polyester and plastic recycling, the stability and hydrolytic activity of known natural esterases toward synthetic polyesters are often inadequate.[139,160,161]
Lipases
  • The desired temperature is usually lower than the glass transition temperature (Tg) of PET, and the ideal pH for the biocatalytic activity of PET is 7.
The challenges to depolymerization of PET by lipases in biocatalytic systems are the consistent product inhibition by hydrolysis byproducts, such as MHET, and another limitation is that it is tough to degrade the crystalline structure of PET [162,163]
PETase
  • The ideal temperature for PETase to degrade the PET depends on the kind of enzyme and its engineering, occurring within the temperature range of 40–80 °C, with the optimal pH around 8–9.
PETase is heat sensitive, and lower temperatures might reduce the specific volume of polymeric materials. Meanwhile, it potentially reduces its degrading activity at mild temperatures.[160,164]
IsPETase
  • IsPETase works efficiently to degrade the PET at temperatures and pH levels up to 30–60 °C and 7.5–8.0.
IsPETase’s poor catalytic effectiveness and inherent lack of thermostability are two of its drawbacks for PET degradation.[114,116]
IsMHETase
  • The ideal temperature for IsMHETase in biocatalytic PET breakdown varies from 50–70 °C, and the ideal pH is around 9.0.
Low catalytic efficiency, poor thermostability, possible product inhibition from degradation intermediates, and the intrinsic stability and crystallinity of the PET polymer itself are the drawbacks of biocatalytic PET depolymerization employing enzymes such as IsMHETase.[47,160]
Thomsen et al. studied the effect of substrate crystallinity and glass transition temperature on the biocatalytic activity of PET by utilizing DuraPETase, which is derived from Ideonella sakaiensis PETase, and LCCICCG, a variation of the leaf-branch compost cutinase. The degree of crystallinity (XC) influenced the rate of enzymatic product release, which stopped successfully around XC 22–27% for LCCICCG compared to the degree of crystallinity of approximately 17% for DuraPETase. The Tg of amorphous PET disks declined from 75 °C to 60 °C after 24 h of pre-soaking in water at 65 °C, whereas the XC remained constant. The enzymatic reaction on pre-soaked disks at 68 °C, more than the Tg, did not influence the removal rate of the enzymatic product catalyzed by LCCICCG [165]. Consequently, high-temperature resistant enzymes are required to facilitate the effective hydrolysis of high-crystalline PET. The inhibition caused by MHET or intermediate metabolites during the catalytic reaction stage is an additional constraint in enzymatic PET biodegradation [166].
Furthermore, the release of by-products from biodegradation or hydrolysis processes increases the acidity of the reaction mixture. Consequently, the wild-type enzyme is inactivated, slowing down the reaction rate. Native wild-type enzymes cannot operate effectively due to these limitations [67,167]. There are some limitations of biocatalytic PET because the enzymatic depolymerization of PET, a heterogeneous catalytic process at the solid–liquid interface, is significantly affected by mass transfer limitations, including insoluble substrates, enzyme adsorption and desorption, and the released products, especially at large solid loadings [157], whereas the mass transfer in solid–liquid heterogeneous processes during PET degradation is hindered by the poor diffusion rate of enzyme molecules across the liquid-solid boundary barrier, which blocks the effective transport of enzymes to the solid surface [143,168]. To eliminate mass transfer limitations and difficulties with accessibility for macromolecular enzymatic substrates, the agitation may significantly reduce the diffusional mass transfer boundary layer, enhancing mass transfer in the hydrolysis unit [169]. Continuous flow reactors are often much smaller than batch reactors; nonetheless, under optimal circumstances and perfect configurations, these bench-scale reactors can provide a greater quantity of product within specific periods compared to their batch counterparts [170]. In addition to that, an additional benefit is the direct transfer of flow systems to large-scale production without substantial further optimization, which sharply contrasts with the upscaling of batch operations [171]. The conventional upscaling of (bio)catalysis in large batch reactors has undergone substantial transformation in recent decades, with downsizing and numbering-up of catalytic processes in continuous flow emerging as a dominant framework [170].
Despite impressive progress in engineering PET-degrading enzymes, several knowledge gaps and practical challenges still hinder their industrial deployment [114,172,173]. Enzyme cost and stability remain critical: most high-performance PET hydrolases still require relatively high loadings, carefully controlled buffers, and stabilizing additives, and there is little long-term data on continuous operation under harsh industrial conditions (mixed plastics, dyes, additives, surfactants, shear, and fluctuating temperatures). Product inhibition and substrate heterogeneity are also not fully resolved; intermediates such as MHET and PET oligomers can accumulate on particle surfaces and slow further hydrolysis, while real waste streams contain multilayer packaging, pigments, and fillers that limit enzyme accessibility and complicate integration with downstream purification. In addition, mass-transfer and scalability issues present major bottlenecks: because PET is a solid, often semi-crystalline material, achieving efficient enzyme–substrate contact at the ton scale requires optimized pretreatment (e.g., grinding, crystallinity reduction) and reactor design, yet techno-economic and reactor-engineering data for slurry or packed-bed systems at high solids loading remain sparse. Finally, systems-level questions—including enzyme recycling, cofactor-free operation, life-cycle performance relative to mechanical and chemical recycling, and compatibility with existing mixed-polymer recycling infrastructure—are only beginning to be addressed and will be crucial for translating PET degradase enzymes from promising laboratory tools into truly competitive large-scale biorecycling technologies.
Recent advances in biocatalytic PET recycling have targeted overcoming the inherent limitations of native PET hydrolases—such as low thermostability, limited activity on crystalline PET, and quick activity loss—through enzyme engineering, directed evolution, and structure-guided design. Structure-based mutations of Ideonella sakaiensis PETase and cutinase-like enzymes have produced variants like DuraPETase, HotPETase, FAST-PETase, and ThermoPETase, which demonstrate significant increases in melting temperature (up to approximately 30 °C) and tens to hundreds of times higher PET hydrolysis rates compared to their wild types, allowing efficient depolymerization at 50–70 °C where PET chains are more mobile [107,163,174]. Directed evolution and semi-rational mutagenesis—often guided by molecular dynamics, free-energy calculations, or machine-learning models—have been used to optimize flexible loops, binding residues, and second-shell positions, balancing substrate affinity and catalytic turnover while enhancing expression and long-term stability under process conditions [175]. Together, these strategies are transforming PET hydrolases from fragile lab curiosities into robust biocatalysts capable of near-complete depolymerization of post-consumer PET, bringing enzymatic PET recycling much closer to industrial feasibility.
Despite remarkable progress in enzymatic PET depolymerization, several bottlenecks continue to limit translation from laboratory studies to full-scale operations. Future research should focus on three dimensions: (i) Enzyme innovation, utilizing machine-learning-guided protein engineering to enhance thermostability and reduce inhibition; (ii) Process integration, through hybrid chemo-enzymatic and continuous-flow systems that minimize energy loss; and (iii) System-level resource governance, promoting regulatory and logistic mechanisms to incorporate enzymatic recycling into national circular economy frameworks. Collectively, these efforts can transform PET waste management from end-of-pipe treatment to proactive resource recovery. Furthermore, Biocatalytic recycling of PET offers an emerging technological possibility. Focused on overcoming existing manufacturing sector difficulties through enzyme engineering and process optimization.

7. Conclusions

From the standpoint of sustainable resource utilization, biocatalytic recycling represents a paradigm shift in the management of PET waste. Rather than treating discarded plastics and textiles as pollutants, enzymatic depolymerization enables their conversion into valuable monomers, TPA and MEG, that can serve as renewable raw materials for new production cycles. This resource-centric approach contributes directly to a circular economy, in which carbon embedded in polymeric materials is conserved and continually reused, lowering dependence on fossil-based feedstocks.
Current research demonstrates that biocatalytic processes can achieve high conversion efficiency under moderate temperatures and without harsh reagents, reducing both energy consumption and environmental footprint compared to traditional chemical routes. Enzyme engineering has advanced rapidly, yielding variants and thermostable cutinases with improved catalytic efficiency and resilience to crystalline PET structures. When integrated with optimized pre-treatment and continuous-flow reactor systems, these enzymes can make closed-loop PET recycling technically and economically viable at an industrial scale.
However, several limitations remain—particularly regarding enzyme stability, mass transfer barriers, product inhibition, and cost-effective recovery of biocatalysts. Addressing these challenges will require interdisciplinary strategies that combine computational enzyme design, immobilization technologies, and life-cycle assessment–based optimization to maximize resource efficiency. Future work should also emphasize policy and infrastructure mechanisms to incorporate enzymatic recycling into broader resource management frameworks. Biocatalytic PET recycling embodies the convergence of biotechnology and resource science. By transforming waste polymers into renewable carbon feedstocks, it not only mitigates environmental pollution but also advances the global transition toward sustainable material circularity and responsible resource stewardship.

Author Contributions

Conceptualization, D.D., H.S.W. and D.S.S.M.; investigation, D.D., A.R.K.; A.S. and D.R.S.; writing—original draft preparation, D.D., D.S.S.M., A.R.K., L.R., Y.F., D.Q.A. and P.Z.S.; writing—review and editing, D.D., H.S.W., C.L.I. and D.S.S.M.; visualization, D.D., D.U.J. and P.Z.S.; supervision, D.D. and H.S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Institut Teknologi Sumatera and Lembaga Penelitian dan Pengabdian Kepada Masyarakat (LPPM) ITERA for financially supporting this research under the grant number 1998bn/IT9.2.1/PT.01.03/2025.

Institutional Review Board Statement

No applicable.

Informed Consent Statement

No applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The researcher would like to express sincere gratitude to all the participants who contributed to this research study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
BHETBis-(2-hydroxyethyl) terephthalate
MHETMono(2-hydroxyethyl) terephthalic acid
MEGMonoethylene glycol
PETPolyethylene terephthalate
TPATerephthalic acid
TgGlass transition temperature

References

  1. Damayanti, D.; Saputri, D.R.; Marpaung, D.S.S.; Yusupandi, F.; Sanjaya, A.; Simbolon, Y.M.; Asmarani, W.; Ulfa, M.; Wu, H.-S. Current Prospects for Plastic Waste Treatment. Polymers 2022, 14, 3133. [Google Scholar] [CrossRef]
  2. Gabriel, V.H.; Schaffernak, A.; Pfitzner, M.; Fellner, J.; Tacker, M.; Apprich, S. Rigid Polyethylene Terephthalate Packaging Waste: An Investigation of Waste Composition and Its Recycling Potential in Austria. Resources 2023, 12, 128. [Google Scholar] [CrossRef]
  3. Ding, Q.; Zhu, H. The Key to Solving Plastic Packaging Wastes: Design for Recycling and Recycling Technology. Polymers 2023, 15, 1485. [Google Scholar] [CrossRef] [PubMed]
  4. Lai, Y.-Y.; Lee, Y.-M. Management strategy of plastic wastes in Taiwan. Sustain. Environ. Res. 2022, 32, 11. [Google Scholar] [CrossRef]
  5. Evode, N.; Qamar, S.A.; Bilal, M.; Barceló, D.; Iqbal, H.M. Plastic waste and its management strategies for environmental sustainability. Case Stud. Chem. Environ. Eng. 2021, 4, 100142. [Google Scholar] [CrossRef]
  6. Oberoi, G.; Garg, A. Single-use plastics: A roadmap for sustainability? Supremo Amic. 2021, 24, 585. [Google Scholar]
  7. Masnadi, M.S.; El-Houjeiri, H.M.; Schunack, D.; Li, Y.; Englander, J.G.; Badahdah, A.; Monfort, J.-C.; Anderson, J.E.; Wallington, T.J.; Bergerson, J.A. Global carbon intensity of crude oil production. Science 2018, 361, 851–853. [Google Scholar] [CrossRef]
  8. Hamilton, L.A.; Feit, S.; Muffett, C.; Kelso, M.; Rubright, S.M.; Bernhardt, C.; Schaeffer, E.; Moon, D.; Morris, J.; Labbé-Bellas, R. Plastic and Climate: The Hidden Costs of a Plastic Planet; CIEL: Washington, DC, USA, 2019. [Google Scholar]
  9. Royer, S.-J.; Ferrón, S.; Wilson, S.T.; Karl, D.M. Production of methane and ethylene from plastic in the environment. PLoS ONE 2018, 13, e0200574. [Google Scholar] [CrossRef]
  10. Egan, J.; Salmon, S. Strategies and progress in synthetic textile fiber biodegradability. SN Appl. Sci. 2022, 4, 22. [Google Scholar] [CrossRef]
  11. Damayanti, D.; Wulandari, L.A.; Bagaskoro, A.; Rianjanu, A.; Wu, H.S. Possibility Routes for Textile Recycling Technology. Polymers 2021, 13, 3834. [Google Scholar] [CrossRef]
  12. Mäkelä, M.; Rissanen, M.; Sixta, H. Machine vision estimates the polyester content in recyclable waste textiles. Resour. Conserv. Recycl. 2020, 161, 105007. [Google Scholar] [CrossRef]
  13. MacArthur, E. A New Textiles Economy: Redesigning Fashion’s Future (2017); Ellen MacArthur Foundation: Cowes, UK, 2021. [Google Scholar]
  14. Koszewska, M. Circular economy—Challenges for the textile and clothing industry. Autex Res. J. 2018, 18, 337–347. [Google Scholar] [CrossRef]
  15. Pristiani, M.; Damayanti, D.; Wu, H.-S. Microwave-Assisted Acid Hydrolysis of PA6 Wastes in PA6 Process: Kinetics, Activation Energies, and Monomer Recovery. Processes 2025, 13, 3175. [Google Scholar] [CrossRef]
  16. Moroni, M.; Mei, A.; Leonardi, A.; Lupo, E.; La Marca, F. PET and PVC separation with hyperspectral imagery. Sensors 2015, 15, 2205–2227. [Google Scholar] [CrossRef] [PubMed]
  17. Esfandiari, A.; Kaghazchi, T.; Soleimani, M. Preparation and evaluation of activated carbons obtained by physical activation of polyethyleneterephthalate (PET) wastes. J. Taiwan Inst. Chem. Eng. 2012, 43, 631–637. [Google Scholar] [CrossRef]
  18. Exchange, T. Preferred Fiber & Materials Market Report 2021; Textile Exchange: Lamesa, TX, USA, 2021; Volume 4. [Google Scholar]
  19. Park, S.H.; Kim, S.H. Poly (ethylene terephthalate) recycling for high value added textiles. Fash. Text. 2014, 1, 1. [Google Scholar] [CrossRef]
  20. Shen, L.; Nieuwlaar, E.; Worrell, E.; Patel, M.K. Life cycle energy and GHG emissions of PET recycling: Change-oriented effects. Int. J. Life Cycle Assess. 2011, 16, 522–536. [Google Scholar] [CrossRef]
  21. Mazhandu, Z.S.; Muzenda, E.; Mamvura, T.A.; Belaid, M.; Nhubu, T. Integrated and consolidated review of plastic waste management and bio-based biodegradable plastics: Challenges and opportunities. Sustainability 2020, 12, 8360. [Google Scholar] [CrossRef]
  22. Jehanno, C.; Alty, J.W.; Roosen, M.; De Meester, S.; Dove, A.P.; Chen, E.Y.-X.; Leibfarth, F.A.; Sardon, H. Critical advances and future opportunities in upcycling commodity polymers. Nature 2022, 603, 803–814. [Google Scholar] [CrossRef]
  23. Raghuwanshi, V.S.; Garnier, G. Nanoparticle decorated cellulose nanocrystals (CNC) composites for energy, catalysis, and biomedical applications. Adv. Funct. Mater. 2025, 35, 2412869. [Google Scholar] [CrossRef]
  24. Mancini, S.D.; Schwartzman, J.A.S.; Nogueira, A.R.; Kagohara, D.A.; Zanin, M. Additional steps in mechanical recyling of PET. J. Clean. Prod. 2010, 18, 92–100. [Google Scholar] [CrossRef]
  25. Bohre, A.; Jadhao, P.R.; Tripathi, K.; Pant, K.K.; Likozar, B.; Saha, B. Chemical recycling processes of waste polyethylene terephthalate using solid catalysts. ChemSusChem 2023, 16, e202300142. [Google Scholar] [CrossRef]
  26. Damayanti; Wu, H.-S. Strategic possibility routes of recycled PET. Polymers 2021, 13, 1475. [Google Scholar] [CrossRef]
  27. Sevigné-Itoiz, E.; Gasol, C.M.; Rieradevall, J.; Gabarrell, X. Contribution of plastic waste recovery to greenhouse gas (GHG) savings in Spain. Waste Manag. 2015, 46, 557–567. [Google Scholar] [CrossRef]
  28. Vollmer, I.; Jenks, M.J.; Roelands, M.C.; White, R.J.; Van Harmelen, T.; De Wild, P.; van Der Laan, G.P.; Meirer, F.; Keurentjes, J.T.; Weckhuysen, B.M. Beyond mechanical recycling: Giving new life to plastic waste. Angew. Chem. Int. Ed. 2020, 59, 15402–15423. [Google Scholar] [CrossRef] [PubMed]
  29. García, J.L. Enzymatic recycling of polyethylene terephthalate through the lens of proprietary processes. Microb. Biotechnol. 2022, 15, 2699–2704. [Google Scholar] [CrossRef]
  30. Hiraga, K.; Taniguchi, I.; Yoshida, S.; Kimura, Y.; Oda, K. Biodegradation of waste PET: A sustainable solution for dealing with plastic pollution. EMBO Rep. 2019, 20, e49365. [Google Scholar] [CrossRef]
  31. Kawai, F. Emerging strategies in polyethylene terephthalate hydrolase research for biorecycling. ChemSusChem 2021, 14, 4115–4122. [Google Scholar] [CrossRef] [PubMed]
  32. Fatima, A.; Khan, F.U.; Hussain, M.; Malik, R.N. Sustainability assessment of home textiles made of recycled PET fibre using life cycle assessment and life cycle costing analyses. Sci. Total Environ. 2025, 982, 179652. [Google Scholar] [CrossRef]
  33. Benyathiar, P.; Kumar, P.; Carpenter, G.; Brace, J.; Mishra, D.K. Polyethylene terephthalate (PET) bottle-to-bottle recycling for the beverage industry: A review. Polymers 2022, 14, 2366. [Google Scholar] [CrossRef] [PubMed]
  34. Mohtaram, F.; Fojan, P. From Waste to Value: Advances in Recycling Textile-Based PET Fabrics. Textiles 2025, 5, 24. [Google Scholar] [CrossRef]
  35. Majumdar, A.; Shukla, S.; Singh, A.A.; Arora, S. Circular fashion: Properties of fabrics made from mechanically recycled poly-ethylene terephthalate (PET) bottles. Resour. Conserv. Recycl. 2020, 161, 104915. [Google Scholar] [CrossRef]
  36. Awaja, F.; Pavel, D. Recycling of PET. Eur. Polym. J. 2005, 41, 1453–1477. [Google Scholar] [CrossRef]
  37. Lerici, L.C.; Renzini, M.S.; Pierella, L.B. Chemical catalyzed recycling of polymers: Catalytic conversion of PE, PP and PS into fuels and chemicals over HY. Procedia Mater. Sci. 2015, 8, 297–303. [Google Scholar] [CrossRef]
  38. Valerio, O.; Muthuraj, R.; Codou, A. Strategies for polymer to polymer recycling from waste: Current trends and opportunities for improving the circular economy of polymers in South America. Curr. Opin. Green Sustain. Chem. 2020, 25, 100381. [Google Scholar] [CrossRef]
  39. Scheirs, J.; Kaminsky, W. Feedstock Recycling and Pyrolysis of Waste Plastics; John Wiley & Sons: Hoboken, NJ, USA, 2006. [Google Scholar]
  40. Muringayil Joseph, T.; Azat, S.; Ahmadi, Z.; Moini Jazani, O.; Esmaeili, A.; Kianfar, E.; Haponiuk, J.; Thomas, S. Polyethylene terephthalate (PET) recycling: A review. Case Stud. Chem. Environ. Eng. 2024, 9, 100673. [Google Scholar] [CrossRef]
  41. Roohi; Bano, K.; Kuddus, M.; Zaheer, M.R.; Zia, Q.; Khan, M.F.; Ashraf, G.M.; Gupta, A.; Aliev, G. Microbial enzymatic degradation of biodegradable plastics. Curr. Pharm. Biotechnol. 2017, 18, 429–440. [Google Scholar] [CrossRef] [PubMed]
  42. Sarda, P.; Hanan, J.C.; Lawrence, J.G.; Allahkarami, M. Sustainability performance of polyethylene terephthalate, clarifying challenges and opportunities. J. Polym. Sci. 2022, 60, 7–31. [Google Scholar] [CrossRef]
  43. Fiorillo, C.; Trossaert, L.; Bezeraj, E.; Debrie, S.; Ohnmacht, H.; Van Steenberge, P.H.; D’hooge, D.R.; Edeleva, M. Molecular and material property variations during the ideal degradation and mechanical recycling of PET. RSC Sustain. 2024, 2, 3596–3637. [Google Scholar] [CrossRef]
  44. Cusano, I.; Campagnolo, L.; Aurilia, M.; Costanzo, S.; Grizzuti, N. Rheology of recycled PET. Materials 2023, 16, 3358. [Google Scholar] [CrossRef] [PubMed]
  45. Elamri, A.; Zdiri, K.; Harzallah, O.; Lallam, A. Progress in polyethylene terephthalate recycling. In Polyethylene Terephthalate: Uses, Properties and Degradation; Nova Science: Hauppauge, NY, USA, 2017. [Google Scholar]
  46. Carta, D.; Cao, G.; D’Angeli, C. Chemical recycling of poly (ethylene terephthalate)(PET) by hydrolysis and glycolysis. Environ. Sci. Pollut. Res. 2003, 10, 390–394. [Google Scholar] [CrossRef]
  47. Wei, R.; von Haugwitz, G.; 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]
  48. Wei, R.; Westh, P.; Weber, G.; Blank, L.M.; Bornscheuer, U.T. Standardization guidelines and future trends for PET hydrolase research. Nat. Commun. 2025, 16, 4684. [Google Scholar] [CrossRef] [PubMed]
  49. Dubdub, I.; Alhulaybi, Z. Catalytic pyrolysis of PET polymer using nonisothermal thermogravimetric analysis data: Kinetics and artificial neural networks studies. Polymers 2022, 15, 70. [Google Scholar] [CrossRef]
  50. Daggubati, L.; Sobhani, Z.; Carbery, M.; Ramadass, K.; Palanisami, T. Fingerprinting risk from recycled plastic products using physical and chemical properties. J. Hazard. Mater. 2025, 488, 137507. [Google Scholar] [CrossRef] [PubMed]
  51. Arcenegui-Troya, J.; Sánchez-Jiménez, P.E.; Perejón, A.; Pérez-Maqueda, L.A. Determination of the activation energy under isothermal conditions: Revisited. J. Therm. Anal. Calorim. 2023, 148, 1679–1686. [Google Scholar] [CrossRef]
  52. Martín-Gullón, I.; Esperanza, M.; Font, R. Kinetic model for the pyrolysis and combustion of poly-(ethylene terephthalate)(PET). J. Anal. Appl. Pyrolysis 2001, 58, 635–650. [Google Scholar] [CrossRef]
  53. Davidson, M.G.; Furlong, R.A.; McManus, M.C. Developments in the life cycle assessment of chemical recycling of plastic waste–A review. J. Clean. Prod. 2021, 293, 126163. [Google Scholar] [CrossRef]
  54. Zimmermann, W. Biocatalytic recycling of plastics: Facts and fiction. Chem. Sci. 2025, 16, 6573–6582. [Google Scholar] [CrossRef]
  55. 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]
  56. Ragaert, K.; Delva, L.; Van Geem, K. Mechanical and chemical recycling of solid plastic waste. Waste Manag. 2017, 69, 24–58. [Google Scholar] [CrossRef] [PubMed]
  57. Sambyal, P.; Najmi, P.; Sharma, D.; Khoshbakhti, E.; Hosseini, H.; Milani, A.S.; Arjmand, M. Plastic recycling: Challenges and opportunities. Can. J. Chem. Eng. 2025, 103, 2462–2498. [Google Scholar] [CrossRef]
  58. Mwanza, B.G. Introduction to recycling. In Recent Developments in Plastic Recycling; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–13. [Google Scholar]
  59. Herrero Acero, E.; Ribitsch, D.; Steinkellner, G.; Gruber, K.; Greimel, K.; Eiteljoerg, I.; Trotscha, E.; Wei, R.; Zimmermann, W.; Zinn, M. Enzymatic surface hydrolysis of PET: Effect of structural diversity on kinetic properties of cutinases from Thermobifida. Macromolecules 2011, 44, 4632–4640. [Google Scholar] [CrossRef]
  60. Tournier, V.; Duquesne, S.; Guillamot, F.; Cramail, H.; Taton, D.; Marty, A.; André, I. Enzymes’ power for plastics degradation. Chem. Rev. 2023, 123, 5612–5701. [Google Scholar] [CrossRef]
  61. Ruginescu, R.; Purcarea, C. Plastic-Degrading Enzymes from Marine Microorganisms and Their Potential Value in Recycling Technologies. Mar. Drugs 2024, 22, 441. [Google Scholar] [CrossRef]
  62. Castro-Rodríguez, J.A.; Rodríguez-Sotres, R.; Farrés, A. Determinants for an efficient enzymatic catalysis in poly (ethylene terephthalate) degradation. Catalysts 2023, 13, 591. [Google Scholar] [CrossRef]
  63. Carniel, A.; de Abreu Waldow, V.; de Castro, A.M. A comprehensive and critical review on key elements to implement enzymatic PET depolymerization for recycling purposes. Biotechnol. Adv. 2021, 52, 107811. [Google Scholar] [CrossRef] [PubMed]
  64. Brackmann, R.; de Oliveira Veloso, C.; de Castro, A.M.; Langone, M.A.P. Enzymatic post-consumer poly (ethylene terephthalate)(PET) depolymerization using commercial enzymes. 3 Biotech 2023, 13, 135. [Google Scholar] [CrossRef]
  65. Amanna, R.; Rakshit, S.K. Review of nomenclature and methods of analysis of polyethylene terephthalic acid hydrolyzing enzymes activity. Biodegradation 2024, 35, 341–360. [Google Scholar] [CrossRef]
  66. Matsuzaki, T.; Saeki, T.; Yamazaki, F.; Koyama, N.; Okubo, T.; Hombe, D.; Ogura, Y.; Hashino, Y.; Tatsumi-Koga, R.; Koga, N. Development and Production of Moderate-Thermophilic PET Hydrolase for PET Bottle and Fiber Recycling. ACS Sustain. Chem. Eng. 2025, 13, 10404–10417. [Google Scholar] [CrossRef]
  67. Kawai, F.; Kawabata, T.; Oda, M. Current knowledge on enzymatic PET degradation and its possible application to waste stream management and other fields. Appl. Microbiol. Biotechnol. 2019, 103, 4253–4268. [Google Scholar] [CrossRef]
  68. Zhang, Z.; Huang, S.; Cai, D.; Shao, C.; Zhang, C.; Zhou, J.; Cui, Z.; He, T.; Chen, C.; Chen, B. Depolymerization of post-consumer PET bottles with engineered cutinase 1 from Thermobifida cellulosilytica. Green Chem. 2022, 24, 5998–6007. [Google Scholar] [CrossRef]
  69. Abokitse, K.; Grosse, S.; Leisch, H.; Corbeil, C.R.; Perrin-Sarazin, F.; Lau, P.C. A novel actinobacterial cutinase containing a noncatalytic polymer-binding domain. Appl. Environ. Microbiol. 2022, 88, e01522-21. [Google Scholar] [CrossRef] [PubMed]
  70. de Oliveira, M.V.D.; Calandrini, G.; da Costa, C.H.S.; da Silva de Souza, C.G.; Alves, C.N.; Silva, J.R.A.; Lima, A.H.; Lameira, J. Evaluating cutinase from Fusarium oxysporum as a biocatalyst for the degradation of nine synthetic polymer. Sci. Rep. 2025, 15, 2887. [Google Scholar] [CrossRef]
  71. Ellen MacArthur Found. Redesigning Fashion’s Future; Ellen MacArthur Found: Cowes, UK, 2017; pp. 1–150. [Google Scholar]
  72. Won, S.-J.; Yim, J.H.; Kim, H.-K. Functional production, characterization, and immobilization of a cold-adapted cutinase from Antarctic Rhodococcus sp. Protein Expr. Purif. 2022, 195, 106077. [Google Scholar] [CrossRef] [PubMed]
  73. Hwang, J.; Yoo, W.; Shin, S.C.; Kim, K.K.; Kim, H.-W.; Do, H.; Lee, J.H. Structural and biochemical insights into Bis (2-hydroxyethyl) terephthalate degrading carboxylesterase isolated from Psychrotrophic bacterium Exiguobacterium antarcticum. Int. J. Mol. Sci. 2023, 24, 12022. [Google Scholar] [CrossRef]
  74. Johan, U.U.M.; Rahman, R.N.Z.R.A.; Kamarudin, N.H.A.; Ali, M.S.M. An integrated overview of bacterial carboxylesterase: Structure, function and biocatalytic applications. Colloids Surf. B Biointerfaces 2021, 205, 111882. [Google Scholar] [CrossRef] [PubMed]
  75. Franklin, M.R. Phase I Biotransformation Reactions-Esterases, and Amidases; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar] [CrossRef]
  76. Hitch, T.C.; Clavel, T. A proposed update for the classification and description of bacterial lipolytic enzymes. PeerJ 2019, 7, e7249. [Google Scholar] [CrossRef]
  77. Wongsatit, T.; Srimora, T.; Kiattisewee, C.; Uttamapinant, C. Enzymes, auxiliaries and cells for the recycling and upcycling of polyethylene terephthalate (PET). Curr. Opin. Syst. Biol. 2024, 38, 100515. [Google Scholar] [CrossRef]
  78. Ghodke, V.M.; Punekar, N.S. Environmental role of aromatic carboxylesterases. Environ. Microbiol. 2022, 24, 2657–2668. [Google Scholar] [CrossRef]
  79. Wawer, J.; Panuszko, A.; Kozłowski, D.; Juniewicz, J.; Szymikowski, J.; Brodnicka, E. Sustainable Management of Microplastic Pollutions from PET Bottles: Overview and Mitigation Strategies. Appl. Sci. 2025, 15, 5322. [Google Scholar] [CrossRef]
  80. Sulaiman, S.; You, D.-J.; Kanaya, E.; Koga, Y.; Kanaya, S. Crystal structure and thermodynamic and kinetic stability of metagenome-derived LC-cutinase. Biochemistry 2014, 53, 1858–1869. [Google Scholar] [CrossRef] [PubMed]
  81. Barth, M.; Honak, A.; Oeser, T.; Wei, R.; Belisário-Ferrari, M.R.; Then, J.; Schmidt, J.; Zimmermann, W. A dual enzyme system composed of a polyester hydrolase and a carboxylesterase enhances the biocatalytic degradation of polyethylene terephthalate films. Biotechnol. J. 2016, 11, 1082–1087. [Google Scholar] [CrossRef] [PubMed]
  82. Tournier, V.; Topham, C.; Gilles, A.; David, B.; Folgoas, C.; Moya-Leclair, E.; Kamionka, E.; Desrousseaux, M.-L.; Texier, H.; Gavalda, S. An engineered PET depolymerase to break down and recycle plastic bottles. Nature 2020, 580, 216–219. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, Y.; Liu, Z.; Guo, X.; Tong, K.; Niu, Y.; Shen, Z.; Weng, H.; Zhang, F.; Wu, J. Enhanced degradation activity of PET plastics by fusion protein of anchor peptide LCI and Thermobifida fusca cutinase. Enzym. Microb. Technol. 2025, 184, 110562. [Google Scholar] [CrossRef]
  84. Wei, R.; Breite, D.; Song, C.; Gräsing, D.; Ploss, T.; Hille, P.; Schwerdtfeger, R.; Matysik, J.; Schulze, A.; Zimmermann, W. Biocatalytic degradation efficiency of postconsumer polyethylene terephthalate packaging determined by their polymer microstructures. Adv. Sci. 2019, 6, 1900491. [Google Scholar] [CrossRef]
  85. Fritzsche, S.; Popp, M.; Spälter, L.; Bonakdar, N.; Vogel, N.; Castiglione, K. Recycling the recyclers: Strategies for the immobilisation of a PET-degrading cutinase. Bioprocess Biosyst. Eng. 2025, 48, 605–619. [Google Scholar] [CrossRef] [PubMed]
  86. Oda, M.; Yamagami, Y.; Inaba, S.; Oida, T.; Yamamoto, M.; Kitajima, S.; Kawai, F. Enzymatic hydrolysis of PET: Functional roles of three Ca2+ ions bound to a cutinase-like enzyme, Cut190*, and its engineering for improved activity. Appl. Microbiol. Biotechnol. 2018, 102, 10067–10077. [Google Scholar] [CrossRef] [PubMed]
  87. Ronkvist, Å.M.; Xie, W.; Lu, W.; Gross, R.A. Cutinase-catalyzed hydrolysis of poly (ethylene terephthalate). Macromolecules 2009, 42, 5128–5138. [Google Scholar] [CrossRef]
  88. Fritzsche, S.; Hübner, H.; Oldiges, M.; Castiglione, K. Comparative evaluation of the extracellular production of a polyethylene terephthalate degrading cutinase by Corynebacterium glutamicum and leaky Escherichia coli in batch and fed-batch processes. Microb. Cell Factories 2024, 23, 274. [Google Scholar] [CrossRef] [PubMed]
  89. Lu, D.; Chen, Y.; Jin, S.; Wu, Q.; Wu, J.; Liu, J.; Wang, F.; Deng, L.; Nie, K. The evolution of cutinase Est1 based on the clustering strategy and its application for commercial PET bottles degradation. J. Environ. Manag. 2024, 368, 122217. [Google Scholar] [CrossRef]
  90. Silva, C.; Da, S.; Silva, N.; Matamá, T.; Araújo, R.; Martins, M.; Chen, S.; Chen, J.; Wu, J.; Casal, M. Engineered Thermobifida fusca cutinase with increased activity on polyester substrates. Biotechnol. J. 2011, 6, 1230–1239. [Google Scholar] [CrossRef]
  91. Han, Y.; Wang, R.; Wang, D.; Luan, Y. Enzymatic degradation of synthetic plastics by hydrolases/oxidoreductases. Int. Biodeterior. Biodegrad. 2024, 189, 105746. [Google Scholar] [CrossRef]
  92. Freije García, F.; García Liñares, G. Use of Lipases as a Sustainable and Efficient Method for the Synthesis and Degradation of Polymers. J. Polym. Environ. 2024, 32, 2484–2516. [Google Scholar] [CrossRef]
  93. Carniel, A.; Valoni, É.; Junior, J.N.; da Conceição Gomes, A.; De Castro, A.M. Lipase from Candida antarctica (CALB) and cutinase from Humicola insolens act synergistically for PET hydrolysis to terephthalic acid. Process Biochem. 2017, 59, 84–90. [Google Scholar] [CrossRef]
  94. Kushwaha, A.; Goswami, L.; Singhvi, M.; Kim, B.S. Biodegradation of poly (ethylene terephthalate): Mechanistic insights, advances, and future innovative strategies. Chem. Eng. J. 2023, 457, 141230. [Google Scholar] [CrossRef]
  95. Müller, R.J.; Schrader, H.; Profe, J.; Dresler, K.; Deckwer, W.D. Enzymatic degradation of poly (ethylene terephthalate): Rapid hydrolyse using a hydrolase from T. fusca. Macromol. Rapid Commun. 2005, 26, 1400–1405. [Google Scholar] [CrossRef]
  96. Eberl, A.; Heumann, S.; Brückner, T.; Araujo, R.; Cavaco-Paulo, A.; Kaufmann, F.; Kroutil, W.; Guebitz, G.M. Enzymatic surface hydrolysis of poly (ethylene terephthalate) and bis (benzoyloxyethyl) terephthalate by lipase and cutinase in the presence of surface active molecules. J. Biotechnol. 2009, 143, 207–212. [Google Scholar] [CrossRef]
  97. Boneta, S.; Arafet, K.; Moliner, V. QM/MM study of the enzymatic biodegradation mechanism of polyethylene terephthalate. J. Chem. Inf. Model. 2021, 61, 3041–3051. [Google Scholar] [CrossRef]
  98. Swiderek, K.; Velasco-Lozano, S.; Galmés, M.; Olazabal, I.; Sardon, H.; López-Gallego, F.; Moliner, V. Mechanistic studies of a lipase unveil a dramatic effect of pH on the selectivity toward the hydrolysis of PET oligomers. Nat. Commun. 2022, 14, 3556. [Google Scholar] [CrossRef] [PubMed]
  99. Liu, X.-h.; Jin, J.-l.; Sun, H.-t.; Li, S.; Zhang, F.-f.; Yu, X.-h.; Cao, Q.-z.; Song, Y.-x.; Li, N.; Lu, Z.-h. Perspectives on the microorganisms with the potentials of PET-degradation. Front. Microbiol. 2025, 16, 1541913. [Google Scholar] [CrossRef]
  100. De Jesus, R.; Alkendi, R. A minireview on the bioremediative potential of microbial enzymes as solution to emerging microplastic pollution. Front. Microbiol. 2023, 13, 1066133. [Google Scholar] [CrossRef]
  101. Guo, R.-T.; Li, X.; Yang, Y.; Huang, J.-W.; Shen, P.; Liew, R.K.; Chen, C.-C. Natural and engineered enzymes for polyester degradation: A review. Environ. Chem. Lett. 2024, 22, 1275–1296. [Google Scholar] [CrossRef]
  102. Safdar, A.; Ismail, F.; Imran, M. Biodegradation of synthetic plastics by the extracellular lipase of Aspergillus niger. Environ. Adv. 2024, 17, 100563. [Google Scholar] [CrossRef]
  103. Heydemann, P.; Guicking, H. Specific volume of polymers as a function of temperature and pressure. Kolloid-Z. Und Z. Für Polym. 1963, 193, 16–25. [Google Scholar] [CrossRef]
  104. Lee, S.H.; Seo, H.; Hong, H.; Park, J.; Ki, D.; Kim, M.; Kim, H.-J.; Kim, K.-J. Three-directional engineering of IsPETase with enhanced protein yield, activity, and durability. J. Hazard. Mater. 2023, 459, 132297. [Google Scholar] [CrossRef]
  105. Hu, Z.; Klupt, K.; Zechel, D.L.; Jia, Z.; Howe, G. Mining Thermophile Genomes for New PETases with Exceptional Thermostabilities Using Sequence Similarity Networks. Chembiochem 2025, 26, e202500065. [Google Scholar] [CrossRef]
  106. 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]
  107. García-Meseguer, R.; Ortí, E.; Tuñón, I.; Ruiz-Pernía, J.J.; Aragó, J. Insights into the enhancement of the poly (ethylene terephthalate) degradation by FAST-PETase from computational modeling. J. Am. Chem. Soc. 2023, 145, 19243–19255. [Google Scholar] [CrossRef] [PubMed]
  108. Eugenio, E.d.Q.; Campisano, I.S.P.; de Castro, A.M.; Coelho, M.A.Z.; Langone, M.A.P. Kinetic modeling of the post-consumer poly (ethylene terephthalate) hydrolysis catalyzed by cutinase from Humicola insolens. J. Polym. Environ. 2022, 30, 1627–1637. [Google Scholar] [CrossRef]
  109. Ribitsch, D.; Acero, E.H.; Greimel, K.; Eiteljoerg, I.; Trotscha, E.; Freddi, G.; Schwab, H.; Guebitz, G.M. Characterization of a new cutinase from Thermobifida alba for PET-surface hydrolysis. Biocatal. Biotransform. 2012, 30, 2–9. [Google Scholar] [CrossRef]
  110. Barth, M.; Oeser, T.; Wei, R.; Then, J.; Schmidt, J.; Zimmermann, W. Effect of hydrolysis products on the enzymatic degradation of polyethylene terephthalate nanoparticles by a polyester hydrolase from Thermobifida fusca. Biochem. Eng. J. 2015, 93, 222–228. [Google Scholar] [CrossRef]
  111. Kondo, F.; Kamiya, N.; Bekker, G.-J.; Nagao, S.; Numoto, N.; Sekiguchi, H.; Ito, N.; Oda, M. Structure-activity relationship of PET-degrading cutinase regulated by weak Ca2+ binding and temperature. Biophys. Physicobiol. 2025, 22, e220009. [Google Scholar] [CrossRef]
  112. Lee, Y.L.; Jaafar, N.R.; Ling, J.G.; Huyop, F.; Bakar, F.D.A.; Rahman, R.A.; Illias, R.M. Cross-linked enzyme aggregates of polyethylene terephthalate hydrolyse (PETase) from Ideonella sakaiensis for the improvement of plastic degradation. Int. J. Biol. Macromol. 2024, 263, 130284. [Google Scholar] [CrossRef]
  113. Chen, C.C.; Han, X.; Ko, T.P.; Liu, W.; Guo, R.T. Structural studies reveal the molecular mechanism of PET ase. FEBS J. 2018, 285, 3717–3723. [Google Scholar] [CrossRef]
  114. Ermis, H. A mini-review on the role of PETase in polyethylene terephthalate degradation. Rev. Environ. Sci. Bio/Technol. 2025, 24, 545–555. [Google Scholar] [CrossRef]
  115. Lu, H.; Diaz, D.J.; Czarnecki, N.J.; Zhu, C.; Kim, W.; Shroff, R.; Acosta, D.J.; Alexander, B.R.; Cole, H.O.; Zhang, Y. Machine learning-aided engineering of hydrolases for PET depolymerization. Nature 2022, 604, 662–667. [Google Scholar] [CrossRef]
  116. Stevensen, J.; Janatunaim, R.Z.; Ratnaputri, A.H.; Aldafa, S.H.; Rudjito, R.R.; Saputro, D.H.; Suhandono, S.; Putri, R.M.; Aditama, R.; Fibriani, A. Thermostability and Activity Improvements of PETase from Ideonella sakaiensis. ACS Omega 2025, 10, 6385–6395. [Google Scholar] [CrossRef] [PubMed]
  117. Britton, D.; Liu, C.; Xiao, Y.; Jia, S.; Legocki, J.; Kronenberg, J.; Montclare, J.K. Protein-engineered leaf and branch compost cutinase variants using computational screening and IsPETase homology. Catal. Today 2024, 433, 114659. [Google Scholar] [CrossRef]
  118. Dissanayake, L.; Jayakody, L.N. Engineering microbes to bio-upcycle polyethylene terephthalate. Front. Bioeng. Biotechnol. 2021, 9, 656465. [Google Scholar] [CrossRef] [PubMed]
  119. Palm, G.J.; Reisky, L.; Böttcher, D.; Müller, H.; Michels, E.A.; Walczak, M.C.; Berndt, L.; Weiss, M.S.; Bornscheuer, U.T.; Weber, G. Structure of the plastic-degrading Ideonella sakaiensis MHETase bound to a substrate. Nat. Commun. 2019, 10, 1717. [Google Scholar] [CrossRef]
  120. Burgin, T.; Pollard, B.C.; Knott, B.C.; Mayes, H.B.; Crowley, M.F.; McGeehan, J.E.; Beckham, G.T.; Woodcock, H.L. The reaction mechanism of the Ideonella sakaiensis PETase enzyme. Commun. Chem. 2024, 7, 65. [Google Scholar] [CrossRef]
  121. Erickson, E.; Shakespeare, T.J.; Bratti, F.; Buss, B.L.; Graham, R.; Hawkins, M.A.; König, G.; Michener, W.E.; Miscall, J.; Ramirez, K.J. Comparative performance of PETase as a function of reaction conditions, substrate properties, and product accumulation. ChemSusChem 2022, 15, e202101932. [Google Scholar] [CrossRef]
  122. Meng, X.; Yang, L.; Liu, H.; Li, Q.; Xu, G.; Zhang, Y.; Guan, F.; Zhang, Y.; Zhang, W.; Wu, N. Protein engineering of stable IsPETase for PET plastic degradation by Premuse. Int. J. Biol. Macromol. 2021, 180, 667–676. [Google Scholar] [CrossRef]
  123. Thomsen, T.B.; Radmer, T.S.; Meyer, A.S. Enzymatic degradation of poly (ethylene terephthalate)(PET): Identifying the cause of the hypersensitive enzyme kinetic response to increased PET crystallinity. Enzym. Microb. Technol. 2024, 173, 110353. [Google Scholar] [CrossRef] [PubMed]
  124. Abid, U.; Sun, G.; Soong, Y.-H.V.; Williams, A.; Chang, A.C.; Ayafor, C.; Patel, A.; Wong, H.-W.; Sobkowicz, M.J.; Xie, D. Evaluation of enzymatic depolymerization of PET, PTT, and PBT polyesters. Biochem. Eng. J. 2023, 199, 109074. [Google Scholar] [CrossRef]
  125. Jia, Q.; Zhang, Z.; Su, L.; Bai, S.; Cai, D.; Chen, C.; Yu, L.; Sun, Y. Enhanced degradation of post-consumer polyethylene terephthalate (PET) wastes by fusion cutinase: Effects of anchor and linker peptides. Process Biochem. 2024, 145, 1–12. [Google Scholar] [CrossRef]
  126. Bååth, J.A.; Borch, K.; Jensen, K.; Brask, J.; Westh, P. Comparative biochemistry of four polyester (PET) hydrolases. ChemBioChem 2021, 22, 1627–1637. [Google Scholar] [CrossRef] [PubMed]
  127. da Cunha, J.M.F. Computational Development of New Biocatalyst for Plastic Degradation Through QM/MM Methods. Master’s Thesis, Universidade do Minho (Portugal), Braga, Portugal, 2021. [Google Scholar]
  128. Chen, X.-Q.; Rao, D.-M.; Zhu, X.-Y.; Zhao, X.-M.; Huang, Q.-S.; Wu, J.; Yan, Z.-F. Current state and sustainable management of waste polyethylene terephthalate bio-disposal: Enzymatic degradation to upcycling. Bioresour. Technol. 2025, 429, 132492. [Google Scholar] [CrossRef]
  129. Kan, Y.; He, L.; Luo, Y.; Bao, R. IsPETase is a novel biocatalyst for poly (ethylene terephthalate)(PET) hydrolysis. ChemBioChem 2021, 22, 1706–1716. [Google Scholar] [CrossRef]
  130. Seo, H.; Kim, S.; Son, H.F.; Sagong, H.-Y.; Joo, S.; Kim, K.-J. Production of extracellular PETase from Ideonella sakaiensis using sec-dependent signal peptides in E. coli. Biochem. Biophys. Res. Commun. 2019, 508, 250–255. [Google Scholar] [CrossRef] [PubMed]
  131. Liu, C.; Shi, C.; Zhu, S.; Wei, R.; Yin, C.-C. Structural and functional characterization of polyethylene terephthalate hydrolase from Ideonella sakaiensis. Biochem. Biophys. Res. Commun. 2019, 508, 289–294. [Google Scholar] [CrossRef]
  132. Magalhães, R.P.; Fernandes, H.S.; Sousa, S.F. The critical role of Asp206 stabilizing residues on the catalytic mechanism of the Ideonella sakaiensis PETase. Catal. Sci. Technol. 2022, 12, 3474–3483. [Google Scholar] [CrossRef]
  133. Bell, E.L.; Smithson, R.; Kilbride, S.; Foster, J.; Hardy, F.J.; Ramachandran, S.; Tedstone, A.A.; Haigh, S.J.; Garforth, A.A.; Day, P.J. Directed evolution of an efficient and thermostable PET depolymerase. Nat. Catal. 2022, 5, 673–681. [Google Scholar] [CrossRef]
  134. Brott, S.; Pfaff, L.; Schuricht, J.; Schwarz, J.N.; Böttcher, D.; Badenhorst, C.P.; 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]
  135. Feng, S.; Yue, Y.; Zheng, M.; Li, Y.; Zhang, Q.; Wang, W. Is PETase-and Is MHETase-catalyzed cascade degradation mechanism toward polyethylene terephthalate. ACS Sustain. Chem. Eng. 2021, 9, 9823–9832. [Google Scholar] [CrossRef]
  136. Sagong, H.-Y.; Seo, H.; Kim, T.; Son, H.F.; Joo, S.; Lee, S.H.; Kim, S.; Woo, J.-S.; Hwang, S.Y.; Kim, K.-J. Decomposition of the PET film by MHETase using Exo-PETase function. ACS Catal. 2020, 10, 4805–4812. [Google Scholar] [CrossRef]
  137. Duan, S.; Chao, T.; Wu, Y.; Wei, Z.; Cao, S. Improvement in the Thermal Stability of Is MHETase by Sequence and Structure-Guided Calculation. Molecules 2025, 30, 988. [Google Scholar] [CrossRef]
  138. Buhari, S.B.; Nezhad, N.G.; Normi, Y.M.; Shariff, F.M.; Leow, T.C. Insight on recently discovered PET polyester-degrading enzymes, thermostability and activity analyses. 3 Biotech 2024, 14, 31. [Google Scholar] [CrossRef]
  139. Thomsen, T.B.; Almdal, K.; Meyer, A.S. Significance of poly (ethylene terephthalate)(PET) substrate crystallinity on enzymatic degradation. New Biotechnol. 2023, 78, 162–172. [Google Scholar] [CrossRef]
  140. Kim, D.H.; Han, D.O.; In Shim, K.; 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]
  141. Brizendine, R.K.; Erickson, E.; Haugen, S.J.; Ramirez, K.J.; Miscall, J.; Salvachúa, D.; Pickford, A.R.; Sobkowicz, M.J.; McGeehan, J.E.; Beckham, G.T. Particle size reduction of poly (ethylene terephthalate) increases the rate of enzymatic depolymerization but does not increase the overall conversion extent. ACS Sustain. Chem. Eng. 2022, 10, 9131–9140. [Google Scholar] [CrossRef]
  142. Zimmermann, W. Biocatalytic recycling of polyethylene terephthalate plastic. Philos. Trans. R. Soc. A 2020, 378, 20190273. [Google Scholar] [CrossRef]
  143. Lu, D.; Wu, J.; Jin, S.; Wu, Q.; Deng, L.; Wang, F.; Nie, K. The enhancement of waste PET particles enzymatic degradation with a rotating packed bed reactor. J. Clean. Prod. 2024, 434, 140088. [Google Scholar] [CrossRef]
  144. Rosevear, A. Immobilised biocatalysts—A critical review. J. Chem. Technol. Biotechnol. 1984, 34, 127–150. [Google Scholar] [CrossRef]
  145. Fukui, S.; Tanaka, A. Bioconversion of lipophilic compounds by immobilized microbial cells in organic solvents. Acta Biotechnol. 1981, 1, 339–350. [Google Scholar] [CrossRef]
  146. Brink, L.; Tramper, J.; Luyben, K.C.A.; Van’t Riet, K. Biocatalysis in organic media. Enzym. Microb. Technol. 1988, 10, 736–743. [Google Scholar] [CrossRef]
  147. Barth, M.; Wei, R.; Oeser, T.; Then, J.; Schmidt, J.; Wohlgemuth, F.; Zimmermann, W. Enzymatic hydrolysis of polyethylene terephthalate films in an ultrafiltration membrane reactor. J. Membr. Sci. 2015, 494, 182–187. [Google Scholar] [CrossRef]
  148. Yang, S.; Ding, W.; Chen, H. Enzymatic hydrolysis of corn stalk in a hollow fiber ultrafiltration membrane reactor. Biomass Bioenergy 2009, 33, 332–336. [Google Scholar] [CrossRef]
  149. Andrić, P.; Meyer, A.S.; Jensen, P.A.; Dam-Johansen, K. Reactor design for minimizing product inhibition during enzymatic lignocellulose hydrolysis: I. Significance and mechanism of cellobiose and glucose inhibition on cellulolytic enzymes. Biotechnol. Adv. 2010, 28, 308–324. [Google Scholar] [CrossRef]
  150. Allen, R.D.; James, M.I. Chemical Recycling of PET. In Circular Economy of Polymers: Topics in Recycling Technologies; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2021; Volume 1391, pp. 61–80. [Google Scholar]
  151. Bezeraj, E.; Debrie, S.; Arraez, F.J.; Reyes, P.; Van Steenberge, P.H.M.; D’Hooge, D.R.; Edeleva, M. State-of-the-art of industrial PET mechanical recycling: Technologies, impact of contamination and guidelines for decision-making. RSC Sustain. 2025, 3, 1996–2047. [Google Scholar] [CrossRef]
  152. Uekert, T.; Singh, A.; DesVeaux, J.S.; Ghosh, T.; Bhatt, A.; Yadav, G.; Afzal, S.; Walzberg, J.; Knauer, K.M.; Nicholson, S.R.; et al. Technical, Economic, and Environmental Comparison of Closed-Loop Recycling Technologies for Common Plastics. ACS Sustain. Chem. Eng. 2023, 11, 965–978. [Google Scholar] [CrossRef]
  153. Uusitalo, T.; Toukoniitty, E. Environmental Impacts of Plastic Recycling—A Literature Review; VTT Technical Research Centre: Espoo, Finland, 2024. [Google Scholar]
  154. Huang, P.; Ahamed, A.; Sun, R.; De Hoe, G.X.; Pitcher, J.; Mushing, A.; Lourenço, F.; Shaver, M.P. Circularizing PET-G Multimaterials: Life Cycle Assessment and Techno-Economic Analysis. ACS Sustain. Chem. Eng. 2023, 11, 15328–15337. [Google Scholar] [CrossRef] [PubMed]
  155. 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]
  156. Uekert, T.; Desveaux, J.S.; Singh, A.; Nicholson, S.R.; Lamers, P.; Ghosh, T.; McGeehan, J.E.; Carpenter, A.C.; Beckham, G.T. Life cycle assessment of enzymatic poly(ethylene terephthalate) recycling. Green Chem. 2022, 24, 6531–6543. [Google Scholar] [CrossRef]
  157. Zhou, J.; Cui, Z.; Wei, R.; Dong, W.; Jiang, M. Interfacial catalysis in enzymatic PET plastic depolymerization. Trends Chem. 2025, 7, 175–185. [Google Scholar] [CrossRef]
  158. Garcia-Galan, C.; Berenguer-Murcia, Á.; Fernandez-Lafuente, R.; Rodrigues, R.C. Potential of different enzyme immobilization strategies to improve enzyme performance. Adv. Synth. Catal. 2011, 353, 2885–2904. [Google Scholar] [CrossRef]
  159. Yuan, H.; Liu, G.; Chen, Y.; Yi, Z.; Jin, W.; Zhang, G. A versatile tag for simple preparation of cutinase towards enhanced biodegradation of polyethylene terephthalate. Int. J. Biol. Macromol. 2023, 225, 149–161. [Google Scholar] [CrossRef]
  160. Jiang, C.; Zhai, K.; Wright, R.C.; Chen, J. Engineered Yeasts Displaying PETase and MHETase as Whole-Cell Biocatalysts for the Degradation of Polyethylene Terephthalate (PET). ACS Synth. Biol. 2025, 14, 2810–2820. [Google Scholar] [CrossRef]
  161. Williams, G.B.; Ma, H.; Khusnutdinova, A.N.; Yakunin, A.F.; Golyshin, P.N. Harnessing extremophilic carboxylesterases for applications in polyester depolymerisation and plastic waste recycling. Essays Biochem. 2023, 67, 715–729. [Google Scholar] [CrossRef] [PubMed]
  162. Świderek, K.; Velasco-Lozano, S.; Galmés, M.À.; Olazabal, I.; Sardon, H.; López-Gallego, F.; Moliner, V. Mechanistic studies of a lipase unveil effect of pH on hydrolysis products of small PET modules. Nat. Commun. 2023, 14, 3556. [Google Scholar] [CrossRef]
  163. Arnal, G.; Anglade, J.; Gavalda, S.; Tournier, V.; Chabot, N.; Bornscheuer, U.T.; Weber, G.; Marty, A. Assessment of four engineered PET degrading enzymes considering large-scale industrial applications. ACS Catal. 2023, 13, 13156–13166. [Google Scholar] [CrossRef] [PubMed]
  164. Chen, Z.; Duan, R.; Xiao, Y.; Wei, Y.; Zhang, H.; Sun, X.; Wang, S.; Cheng, Y.; Wang, X.; Tong, S. Biodegradation of highly crystallized poly (ethylene terephthalate) through cell surface codisplay of bacterial PETase and hydrophobin. Nat. Commun. 2022, 13, 7138. [Google Scholar] [CrossRef]
  165. Thomsen, T.B.; Hunt, C.J.; Meyer, A.S. Influence of substrate crystallinity and glass transition temperature on enzymatic degradation of polyethylene terephthalate (PET). New Biotechnol. 2022, 69, 28–35. [Google Scholar] [CrossRef] [PubMed]
  166. Maurya, A.; Bhattacharya, A.; Khare, S.K. Enzymatic remediation of polyethylene terephthalate (PET)–based polymers for effective management of plastic wastes: An overview. Front. Bioeng. Biotechnol. 2020, 8, 602325. [Google Scholar] [CrossRef] [PubMed]
  167. Ahmaditabatabaei, S.; Kyazze, G.; Iqbal, H.M.; Keshavarz, T. Fungal enzymes as catalytic tools for polyethylene terephthalate (PET) degradation. J. Fungi 2021, 7, 931. [Google Scholar] [CrossRef]
  168. Faridkhou, A.; Tourvieille, J.-N.; Larachi, F. Reactions, hydrodynamics and mass transfer in micro-packed beds—Overview and new mass transfer data. Chem. Eng. Process. Process Intensif. 2016, 110, 80–96. [Google Scholar] [CrossRef]
  169. Sun, C.; Meng, X.; Sun, F.; Zhang, J.; Tu, M.; Chang, J.-S.; Reungsang, A.; Xia, A.; Ragauskas, A.J. Advances and perspectives on mass transfer and enzymatic hydrolysis in the enzyme-mediated lignocellulosic biorefinery: A review. Biotechnol. Adv. 2023, 62, 108059. [Google Scholar] [CrossRef]
  170. De Santis, P.; Meyer, L.-E.; Kara, S. The rise of continuous flow biocatalysis–fundamentals, very recent developments and future perspectives. React. Chem. Eng. 2020, 5, 2155–2184. [Google Scholar] [CrossRef]
  171. Wegner, J.; Ceylan, S.; Kirschning, A. Ten key issues in modern flow chemistry. Chem. Commun. 2011, 47, 4583–4592. [Google Scholar] [CrossRef]
  172. Zhang, S.; Wang, X.; Shen, H.; Zhang, J.; Dong, W.; Yu, Z. Scalable nanoplastic degradation in water with enzyme-functionalized porous hydrogels. J. Hazard. Mater. 2025, 487, 137196. [Google Scholar] [CrossRef] [PubMed]
  173. Yao, J.; Liu, Y.; Gu, Z.; Zhang, L.; Guo, Z. Deconstructing PET: Advances in enzyme engineering for sustainable plastic degradation. Chem. Eng. J. 2024, 497, 154183. [Google Scholar] [CrossRef]
  174. Liu, Y.-J.; Zhou, J.; Li, Y.; Yan, X.; Xu, A.; Zhou, X.; Liu, W.; Xu, Y.; Su, T.; Wang, S. State-of-the-art advances in biotechnology for polyethylene terephthalate bio-depolymerization. Green Carbon 2025, 3, 303–319. [Google Scholar] [CrossRef]
  175. Zhou, J.; Huang, M. Navigating the landscape of enzyme design: From molecular simulations to machine learning. Chem. Soc. Rev. 2024, 53, 8202–8239. [Google Scholar] [CrossRef]
Figure 1. Reaction mechanism of biocatalytic Polyethylene Terephthalate via a classical serine-hydrolase. Redrawn on the basis of [47].
Figure 1. Reaction mechanism of biocatalytic Polyethylene Terephthalate via a classical serine-hydrolase. Redrawn on the basis of [47].
Resources 14 00176 g001
Figure 2. Schematic illustration of the enzymatic degradation pathway of PET by cutinase, lipase, and carboxylesterase by author’s own work.
Figure 2. Schematic illustration of the enzymatic degradation pathway of PET by cutinase, lipase, and carboxylesterase by author’s own work.
Resources 14 00176 g002
Figure 3. Several enzymes are used in product distribution and recycling of waste PET: PETase, IsPETase, and IsMHETase by author’s own work.
Figure 3. Several enzymes are used in product distribution and recycling of waste PET: PETase, IsPETase, and IsMHETase by author’s own work.
Resources 14 00176 g003
Figure 4. The process flow diagram for biocatalytic PET for large–scale industrial application. Redrawn on the basis of [54].
Figure 4. The process flow diagram for biocatalytic PET for large–scale industrial application. Redrawn on the basis of [54].
Resources 14 00176 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Damayanti, D.; Marpaung, D.S.S.; Kodarif, A.R.; Sanjaya, A.; Saputri, D.R.; Fahni, Y.; Rahmiyati, L.; Silvia, P.Z.; A’yuni, D.Q.; Imalia, C.L.; et al. Biocatalytic Recycling of Polyethylene Terephthalate: From Conventional to Innovative Routes for Transforming Plastic and Textile Waste into Renewable Resources. Resources 2025, 14, 176. https://doi.org/10.3390/resources14110176

AMA Style

Damayanti D, Marpaung DSS, Kodarif AR, Sanjaya A, Saputri DR, Fahni Y, Rahmiyati L, Silvia PZ, A’yuni DQ, Imalia CL, et al. Biocatalytic Recycling of Polyethylene Terephthalate: From Conventional to Innovative Routes for Transforming Plastic and Textile Waste into Renewable Resources. Resources. 2025; 14(11):176. https://doi.org/10.3390/resources14110176

Chicago/Turabian Style

Damayanti, Damayanti, David Septian Sumanto Marpaung, Abdul Rozak Kodarif, Andri Sanjaya, Desi Riana Saputri, Yunita Fahni, Lutfia Rahmiyati, Putri Zulva Silvia, Dewi Qurrota A’yuni, Calaelma Logys Imalia, and et al. 2025. "Biocatalytic Recycling of Polyethylene Terephthalate: From Conventional to Innovative Routes for Transforming Plastic and Textile Waste into Renewable Resources" Resources 14, no. 11: 176. https://doi.org/10.3390/resources14110176

APA Style

Damayanti, D., Marpaung, D. S. S., Kodarif, A. R., Sanjaya, A., Saputri, D. R., Fahni, Y., Rahmiyati, L., Silvia, P. Z., A’yuni, D. Q., Imalia, C. L., Janah, D. U., & Wu, H. S. (2025). Biocatalytic Recycling of Polyethylene Terephthalate: From Conventional to Innovative Routes for Transforming Plastic and Textile Waste into Renewable Resources. Resources, 14(11), 176. https://doi.org/10.3390/resources14110176

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