Industrially viable and selective catalytic system: simple and sustainable pathway for efficient degradation of waste polyester textiles

Yu Zhou a, Jiaxing Zhang *ab, Bowen Shen a, Wenyan Ba f, Shengping You *ac, Mengfan Wang de, Rongxin Su abf and Wei Qi *ab
aChemical Engineering Research Centre, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, PR China. E-mail: zhangjiaxing7137@tju.edu.cn; ysp@tju.edu.cn; qiwei@tju.edu.cn
bState Key Laboratory of Chemical Engineering and Low-Carbon Technology, Tianjin University, Tianjin 300072, PR China
cBeijing Meihao Biotechnology Co., Ltd, China
dSchool of Life Sciences, Faculty of Medicine, Tianjin University, Tianjin 300072, P. R. China
eYuantian Biotechnology (Tianjin) Co., Ltd, China
fTianjin Key Laboratory for Marine Environmental Research and Service, School of Marine Science and Technology, Tianjin University, Tianjin 300072, China

Received 15th January 2025 , Accepted 7th April 2025

First published on 8th April 2025


Abstract

Waste textile recycling is hampered by their complexity and high crystallinity, leading to substantial environmental pollution and resource wastage. Although enzymatic polyester recycling has promising substrate specificity, natural enzymes lack stability and activity on high-crystallinity substrates, and thus are limited in industrial viability. Herein, in light of the degradation mechanism of polyethylene terephthalate (PET) hydrolyase, we established a simple ethylene glycol (EG) selective catalytic system, possessing a nucleophilic attack mechanism on ester bonds similar to that of the catalytic serine residue in the natural enzyme. The EG catalytic system achieved efficient PET degradation, with a PET conversion rate of 99.63% and a terephthalic acid (TPA) yield of 95.46% under mild conditions (90 °C for 1 h). Meanwhile, the insolubility of the degradation products in EG catalytic system facilitates their separation, allowing the reaction system to be recycled at least five times. Additionally, first-principles molecular dynamics simulations revealed that the EG catalytic system generates active species EG, which follows the same reaction mechanism as natural enzymes and has a lower energy barrier than alkaline hydrolysis. Notably, the process maintains effective hydrolysis capability at a 100 L scale, with selective degradation and decolorization, underscoring its industrial potential for PET degradation from colored composite textiles. Overall, this work offers a sustainable, efficient, and practical solution to the challenge of textile waste recycling.



Green foundation

1. This study established a simple EG catalytic system for PET degradation, achieving the rapid decomposition of waste PET fibers under mild conditions (90 °C, 1 h) with a high PET conversion rate of 99.63% and a TPA yield of 95.46%. The reaction system can be directly recycled for subsequent PET degradation five times and enables selective degradation and decolorization of colored composite fibers.

2. The process is green and scalable, free of toxic reagents or heavy metal ions, with its effectiveness comfirmed by both theoretical analysis (first-principles molecular dynamics simulation) and experimental validation. It exhibits excellent PET degradation performance at a 100 L scale and efficiently recycles real post-consumer PET waste, demonstrating significant potential for industrial applications.

3. EG produced during PET degradation is directly recycled within the EG catalytic reaction system, avoiding the high energy consumption of distillation separation. Meanwhile, TPA also has the potential to be upcycled into high-value products, and future research could focus on enhancing TPA valorization and expanding the applications of this process.


Introduction

The rapid expansion of the textile industry has led to a substantial accumulation of post-industrial and post-consumer textile waste.1,2 Polyester is the most commonly used fibre in the textile sector, with polyethylene terephthalate (PET) being particularly favoured for fibre production owing to its cost-effectiveness, high durability, and water resistance.3 Notably, although textile manufacturing generates over 1.2 billion tons of carbon emissions annually and causes substantial environmental harm,4,5 research on PET recycling has predominantly focused on the clean, high-purity, pigment-free, bottle-grade PET, even though most PET (approximately 60%) is used in textile production, with only a small portion (approximately 30%) used in bottle production.6–9

Currently, the primary technologies for PET recovery include the following: (1) mechanical recycling, which involves PET melting and respinning processes, but results in relatively low and heterogeneous intrinsic viscosity values, thereby reducing the quality of recycled PET products.7,10,11 Additionally, mechanical recycling is limited to well-separated, high-purity waste plastics and faces challenges in recycling blended textiles.12 (2) Chemical recycling, typically encompassing hydrolysis, aminolysis, methanolysis, and glycolysis,13,14 which is ineffective for degrading complex mixtures and requires time-consuming and energy-intensive separation and purification processes.15 To address this issue, the metal-catalyzed oxidation process has been used to depolymerize mixed PET into oxygenated compounds, which are then biologically converted into value-added chemicals, thereby eliminating the need for mixed plastic waste separation.16 Moreover, chemical recycling often involves harsh conditions, such as high temperatures, high pressure, strong acids, strong bases, or expensive industrial catalysts,17 leading to high energy costs and the generation of hazardous waste. Further, the PET acetolysis process has been developed, which enables the continuous precipitation of terephthalic acid (TPA) from acetic acid, thereby avoiding the additional consumption of strong acids and bases required for TPA acidification.18 (3) Biological recovery utilizing PET hydrolase, which presents an eco-friendly solution characterized by mild reaction conditions and substrate specificity. However, biocatalysts suffer from poor stability during recycling and struggle to degrade highly crystalline substrates,19 thus hindering their industrialization.

Meanwhile, highly selective transformation of original building blocks is essential for effectively constructing target high-value products.20 However, textile fabrics are typically designed for enhanced performance by incorporating multiple fibre composites. For example, interwoven cotton–polyester blends are widely used in clothing and textile materials because they combine the comfort and breathability of cotton with the strength and abrasion resistance of PET.3 Consequently, the heterogeneity of blended fabrics and the lack of efficient sorting and separation techniques hinder textile recycling,21,22 ultimately leading to downcycled products.23 Notably, while PET hydrolases offer advantages such as mild reaction conditions and high substrate specificity, their industrial application in PET polyester recycling is hindered by low stability, short shelf life, high production costs, and limited activity on highly crystalline substrates.7,24–27 To solve this problem, numerous artificial biocatalysts have been developed as enzyme mimics, aiming to replicate the catalytic function of natural enzymes, while mitigating their limitations.28 Enzyme mimics exhibit simpler structures, enhanced stability, and greater efficiency than native enzymes, making them robust and potentially more suitable alternatives for enzymatic catalysis in industrial applications.28,29 However, the residual metal catalysts in PET waste can accelerate transesterification and polycondensation reactions, leading to chemical inhomogeneity in the recycled PET and altering its melt flow properties.30

Moreover, balancing cost and efficiency is a critical consideration for the industrial application of PET recycling. Before the valorisation of PET into high-value products, PET must be depolymerized into its monomers, ethylene glycol (EG) and TPA.31 The TPA generated from PET degradation can be recovered from the reaction solution via acidification.31 However, the separation of EG is hindered by its high viscosity, water solubility, and boiling point of 197.6 °C, which results in elevated energy consumption during the separation process and, consequently, increases the overall production costs in PET recycling.32

Herein, considering the PET hydrolysis mechanism of natural PET hydrolase, we aim to develop a recycling process for highly crystalline PET fibre that enables selective degradation and demonstrates potential for industrial application (Fig. 1). Firstly, due to the high energy consumption associated with separating the PET degradation product EG, we sought to utilize activated EG as a nucleophile in the PET degradation reaction, thereby eliminating the need for a separate EG recovery step. Importantly, this EG catalytic system demonstrated efficient degradation without the need for heavy-metal catalysts or toxic organic reagents and could be directly separated from the PET degradation products for reuse, thereby reducing the consumption of strong acids and bases. Further, to elucidate the PET degradation mechanism facilitated by the EG catalytic system, we employed proton nuclear magnetic resonance (1H NMR) spectroscopy and first principles metadynamics simulations to gain deeper insight into the PET degradation pathway. Finally, we assessed the industrial application potential of the EG catalytic system through a 100 L scale-up and the investigation of real textile waste. Additionally, the system's substrate specificity was validated across polyester fibres, cotton, and nylon, thereby confirming its selective catalytic properties. Thus, this study proposes a high-efficiency, recyclable, eco-friendly, and practical process for the hydrolysis of PET-based blended fabrics, with the potential to accelerate the industrialization of textile waste recycling.


image file: d5gc00248f-f1.tif
Fig. 1 PET hydrolysis mechanism of the EG catalytic system, in connection with the enzymatic catalytic mechanism. (a) Construction of the EG catalytic system based on the PET hydrolase mechanism, incorporating BhrPETase33 (gray, PDB ID: 7EOA), LCC-ICCG (an engineered leaf-branch compost cutinase,34 purple, PDB ID: 7VVE), IsPETase-PA35 (orange, PDB ID: 8J17), and TfCUT2[thin space (1/6-em)]36 (pink, PDB ID: 4CG1); (b) catalytic mechanism of PET hydrolases;37 and (c) proposed PET degradation mechanism in the EG catalytic system.

Results and discussion

Establishing an EG catalytic system based on the mechanism of PET hydrolase

PETase is a member of the α/β hydrolase superfamily38 that features a nucleophilic serine within the conserved sequence (Gly-X1-Ser-X2-Gly). This serine serves as a covalent nucleophile to cleave the scissile ester bond.39 PET degradation begins when the catalytic serine attacks the carbonyl carbon of PET, forming a tetrahedral intermediate (TI). This intermediate is stabilized by histidine, aspartic acid, and the oxyanion hole, and subsequently transitions into an acyl–enzyme (AE) intermediate. A water molecule then attacks the carbonyl group of the AE, forming a second TI that undergoes deacylation to release carboxylic acid products (Fig. 1a and b).37

Considering the catalytic mechanism of PET hydrolase, we chose EG as the catalyst for two reasons: first, EG can function as a nucleophile after deprotonation, and second, it is both a component of PET and a degradation product, thus avoiding the introduction of impurities into the system. Therefore, we established an EG catalytic system with a NaOH component to explore the efficient degradation of PET. On the basis of the PET hydrolysis mechanism of the natural enzyme,40,41 the mechanism of PET degradation by the EG catalytic system is proposed in Fig. 1c. Initially, NaOH deprotonates EG to form EG, which then interacts with the PET surface, forming substrate complexes (SC). The EG then nucleophilically attacks the carbon atom of the carbonyl group in the PET unit (forming TI1), leading to the formation of bis(2-hydroxyethyl)terephthalate (BHET) and EG (AE). Subsequently, the free OH nucleophilically attacks the carbonyl group in BHET (forming TI2) and degrades the ester bonds, thereby finally degrading BHET into product complexes (PC).

Feasibility analysis of the EG catalytic system for PET degradation

First, we investigated the feasibility of PET degradation using the EG catalytic system. Notably, this system achieved a rapid degradation rate of 99.63% and a TPA yield of 95.46% under mild reaction conditions (90 °C, 1 h, Fig. S2). Meanwhile, without the use of additional heavy-metal catalysts or toxic organic reagents, the degradation efficiency of the EG catalytic system improved 5-fold (Fig. 2a) compared to the alkaline hydrolysis system (NaOH solution). The SEM images further demonstrated that, while the surface of the PET fibres showed slight etching in the alkaline hydrolysis system, many deep groove-like structures formed in the EG catalytic system (Fig. 2b), illustrating the substantial enhancement of PET degradation.
image file: d5gc00248f-f2.tif
Fig. 2 Feasibility analysis of the EG catalytic system for PET degradation. (a) Comparison of the PET degradation efficiency between the EG catalytic system (100 mL reaction system, molar ratio of EG to NaOH 6[thin space (1/6-em)]:[thin space (1/6-em)]1, substrate concentration 30 g L−1, 90 °C, 1 h) and alkaline hydrolysis system (100 mL reaction system, NaOH added in the same amount as in the EG catalytic system, under the same reaction conditions); (b) surface morphology of PET fibres after degradation; (c) steps in the EG catalytic system and the alkaline hydrolysis system; (d) material balance calculation of the EG catalytic system and alkaline hydrolysis system42 (for degrade 1 kg PET); (e) recycling of the EG catalytic system (100 mL reaction system, molar ratio of EG to NaOH 6[thin space (1/6-em)]:[thin space (1/6-em)]1, substrate concentration 30 g L−1, 90 °C, 1 h); and (f) degradation effect on PET fibres at different substrate concentrations (100 mL reaction system, molar ratio of EG to NaOH 6[thin space (1/6-em)]:[thin space (1/6-em)]1, substrate concentration 30–100 g L−1, 90 °C, 0.5–1.5 h). § EG can be directly recovered and recycled.

Considering the high energy consumption and cost associated with the separation of water-soluble EG, we investigated the feasibility of directly separating and recycling the EG catalytic system from the PET degradation product TPA (Fig. 2d). Interestingly, Na2TPA dissolved in EG (7.49 g L−1 at RT, Fig. S3b, blue line) but was almost insoluble in the EG catalytic system (Fig. S3a), thereby supporting the viability of direct separation of the reaction solution for reuse and TPA product recovery. This phenomenon could be attributed to the common-ion effect, in which the addition of a large amount of Na+ (from NaOH) substantially reduces the solubility of Na2TPA in EG catalytic system. Importantly, the excess NaOH in the EG catalytic system was directly recycled with EG, reducing the need for acid to neutralize the excess NaOH in the degradation system. This reduced the addition of strong acids and bases and minimized the generation of salt-containing wastewater (Fig. 2d). However, the traditional alkaline hydrolysis of PET presents difficulties in directly separating the reaction system from its degradation products, leading to excessive use of strong acids and bases and the generation of waste reaction solutions, which contribute to secondary pollution (Fig. 2d). As shown in Fig. 2e, the EG catalytic system reduced H2SO4 consumption by 12-fold, as the amount of H2SO4 required to neutralize excess alkali was minimized. Additionally, wastewater production was reduced by 10-fold compared to the alkaline hydrolysis system (see the calculation process in ESI), thereby reducing environmental impact and lowering degradation costs. Furthermore, the Na2TPA separated from the EG reaction system must be dissolved in water to separate it from the unreacted substrate. Since Na2TPA has high solubility in water (Fig. S3b, yellow line), only a small amount of water is required for complete dissolution. Simultaneously, the insolubility of TPA in water was confirmed, ensuring efficient recovery (Fig. S3a and S3b, black line). Notably, the EG produced during PET degradation was directly reused as the reaction system, thereby avoiding the energy-intensive process of EG distillation and recovery. The EG catalytic system was successfully recycled at least five times (Fig. 2c), while maintaining excellent PET conversion (>94%) and TPA yield (>93%). However, since the reaction between EG and NaOH generates a small amount of water, continuous recycling of the EG catalytic system may lead to gradual water accumulation, which could potentially affect PET degradation efficiency and TPA separation over time. Therefore, it is worth further investigation into effective and low-cost dehydration methods in the future.

Furthermore, previous studies have indicated that a low solid-to-liquid ratio leads to higher energy consumption during PET treatment and increases carbon footprints,43 which is detrimental to the economy and the environment in the context of PET textile recycling. Therefore, we also explored the influence of substrate concentration on PET textile degradation. According to Fig. 2f, under lower substrate concentrations (30 and 50 g L−1), the EG catalytic system achieved over 95% PET conversion within 1 h. However, PET degradation efficiency decreased as the substrate concentration increased. Nevertheless, when the reaction time was extended to 1.5 h, the PET conversion at higher substrate concentrations (70 and 100 g L−1) rose to 95%, with a high TPA yield (>90%). Therefore, the EG catalytic system demonstrates efficient PET degradation at high substrate concentrations without the need for additional heavy-metal catalysts or toxic organic reagents (Table S1), highlighting its sustainable and energy-saving features and making it suitable for industrial scale-up.

Mechanistic investigation of the EG catalytic system for polyester degradation

Based on previous assumptions, we propose that the EG catalytic system initially degrades PET to BHET through the EG active species, then degrades BHET to TPA by OH (Fig. 3a). To clarify the exact nucleophilic reagent in the EG catalytic system, we performed 1H NMR spectroscopy. The 1H NMR spectra (Fig. 3b and c) showed the disappearance of the signal assigned to the proton of EG (δ = 4.4 ppm), confirming its deprotonation into the active species. Neutral EG has two possible deprotonated species, namely the monoanion EG and the dianion EG2−. Since the highest occupied molecular orbital (HOMO) energy of EG, EG, and EG2− increases sequentially, it is hard for EG to be further deprotonated into EG2−,44 given the significant energy difference between EG and EG2− (Fig. S4). Therefore, the PET degradation ability of the EG catalytic system is attributed to EG. Acting as a general base nucleophilic group, EG attacks the carbonyl carbon in the PET chain and promotes PET degradation.45,46
image file: d5gc00248f-f3.tif
Fig. 3 Mechanistic analysis of the EG catalytic system for PET degradation. (a) PET degradation process via the EG catalytic system; (b and c) NMR analysis of the (b) EG/NaOH system and (c) pure EG; (d) relative reaction energy pathway of the EG catalytic system compared to the alkaline hydrolysis system; (e and f) PET degradation mechanism in the EG catalytic system, explored via meta-FPMD. The free energy landscape is depicted using colour gradients, where darker shades indicate higher population densities. The reaction path is shown by scatter points, starting at blue points, passing through white points, and finally reaching red points at 3000 fs; (inset) conformation of the TI state during the reaction. Similar methods for reaction pathways visualization have also been applied by Gao and co-workers.47,48

To further understand the PET degradation mechanism catalysed by the EG catalytic system, we performed FPMD simulations to elucidate the precise details of the PET-EG degradation process. As the catalytic process may be time-consuming, we introduced the metadynamics enhanced sampling method into FPMD (meta-FPMD) to accelerate the simulation progress. Utilizing two key distances d1 and d2 (Fig. 3e and f), as collective variables (CVs), we monitored the breaking of ester bond and formation of new bond during each reaction step. Specifically, d1 represents the distance between the carbonyl carbon and the oxygen of the ester bond in the substrate, reflecting bond cleavage, while d2 corresponds to the distance between the nucleophilic group (the hydroxyl oxygen of EG in step 1 and hydroxide ion oxygen in step 2) and the carbonyl carbon of the ester bond in the substrate, indicating the formation of a new bond. Based on these CVs, we analysed the free energy landscape (FEL) of the catalytic systems within a 3000 fs simulation. The FEL of the ester exchange reaction from the SC to the AE intermediate, reconstructed from the meta-FPMD simulation (Fig. 3e and Movie S1), revealed two energy minima and one key TI1 state. The energy path represents a reaction trajectory from SC to AE, passing through TI1 and terminating in the AE state. Similarly, the FEL transition from AE to PC proceeds through TI2, eventually forming the PC. This pathway aligns with that reported for natural enzymes degrading PET, such as LCC-ICCG41 and its mutant.49 The results demonstrate that the active species EG degrades PET ester bonds through a reaction mechanism similar to that of natural enzymes, consistent with experimental results (Fig. 2a). We also conducted meta-FPMD on the alkaline hydrolysis system and calculated the FEL for the two-step reactions (Fig. S5). Compared to the EG catalytic system, the alkaline hydrolysis system exhibited a higher energy barrier of the SC–AE reaction (Fig. 3d), resulting in lower PET degradation efficiency than that observed within the EG catalytic system.

Practical application of the EG catalytic system

Building on the systematic exploration of the EG catalytic system's excellent degradation capabilities described above, we further investigated its potential for industrialization by applying it to larger-scale PET fibre degradation under high substrate concentrations. Compared with the small-scale test (0.1 L with 30 g L−1 PET), the PET conversion at the 1 L scale with 100 g L−1 PET decreased slightly, but still maintained a high degradation efficiency (PET conversion >85%, Fig. 4a). The slight decrease in PET degradation efficiency may be attributed to the high concentration of fluffy PET fibres and the gradual accumulation of insoluble products, which could lead to incomplete reactions. We then further investigated PET degradation on a 100 L scale with 70 g L−1 PET, combining activated carbon (AC) decolorization to improve TPA quality. As shown in Fig. 4b, the EG catalytic system maintained efficient PET degradation even when scaled up to 100 L (87.4% PET conversion and 85.9% TPA yield), indicating its strong potential for industrial-scale PET recycling. To further scale up degradation at higher substrate concentrations, more suitable reactors will be required to achieve thorough mixing and maximise efficiency.
image file: d5gc00248f-f4.tif
Fig. 4 Practical application of the EG catalytic system. (a) 1 L and (b) 100 L scale-up tests of the EG catalytic system for PET fibres degradation.

Furthermore, the EG catalytic system was applied to degrade waste curtains and scarves recovered from waste stations, which mainly consist of PET polyester fibres with attached dye. As shown in Fig. 5a, the system achieved efficient degradation of PET textiles, with curtains and scarves reaching PET conversion rates of over 90%. However, the degradation efficiency for real textiles was slightly lower than that of pure PET polyester, prompting further analysis of scarf degradation residues using FTIR (Fig. S6a). Notably, the FTIR spectrum of the scarf degradation residues revealed distinct characteristic peaks compared to those of pure PET polyester, indicating that the undegraded components belonged to other materials present in the blended fabric. This result demonstrated the potential selective degradation capability of the EG catalytic system. Meanwhile, as a safe organic solvent, EG exhibits good affinity for organic dyes, allowing them to be dissolved and removed from the degradation products (Fig. 5b). Activated carbon can further eliminate residual colour, thereby enhancing the quality of the recovered TPA. Additionally, the chemical structure of the obtained TPA was verified using FTIR spectroscopy and XRD (Fig. 5c and d), which matched that of commercially purified TPA. Moreover, HPLC analysis (Fig. S6b) confirmed that the recovered TPA possessed high purity (>99%) and was free of oligomers. These results strongly demonstrate the efficiency of the EG catalytic system in degrading real waste textiles. However, as the EG catalytic system is continuously recycled, dyes may also accumulate over time; thus, more efficient and cost-effective decolorization methods should be considered in the future.


image file: d5gc00248f-f5.tif
Fig. 5 Practical application of the EG catalytic system. (a) Degradation performance on real PET waste fabrics (100 mL reaction system, molar ratio of EG to NaOH 6[thin space (1/6-em)]:[thin space (1/6-em)]1, substrate concentration 30 g L−1, 90 °C, 1 h); (b) degradation process of real PET waste fabric; (c) FTIR and (d) XRD analyses of PET degradation products, (e) selective degradation performance of the EG catalytic system (100 mL reaction system, molar ratio of EG to NaOH 6[thin space (1/6-em)]:[thin space (1/6-em)]1, substrate concentration 30 g L−1 (cotton/nylon to PET fibres mass ratio 1[thin space (1/6-em)]:[thin space (1/6-em)]1), 90 °C, 1 h); (f) FTIR analysis of raw cotton/nylon and residual cotton/nylon; and (g) SEM images of raw cotton/nylon and residual nylon/cotton.

Given that textiles often consist of mixed fibres, selective degradation plays a crucial role in obtaining a single degradation product, thereby simplifying further recycling. Therefore, the substrate specificity of the EG catalytic system was further explored using nylon, cotton, and mixtures of nylon/cotton with PET polyester as the degradation substrates (Fig. S7). Notably, Fig. 5e shows that the degradation of cotton and cotton/polyester fibre mixtures resulted in the recovery of more than 95% of cotton, while achieving 100% PET conversion. Similarly, the degradation of nylon and nylon/polyester fibre blends recovered 100% of the nylon and also achieved 100% PET conversion. Furthermore, FTIR analysis confirmed that raw and residual nylon/cotton fibres exhibited the same characteristic peaks (Fig. 5f). The degradation of nylon involves the cleavage of amide bonds, which exhibit low reactivity toward nucleophilic attack on the carbon due to the partial double bond character of the C–N bond, caused by delocalization between the nitrogen lone pair and the C[double bond, length as m-dash]O bond.50 SEM images in Fig. 5g also showed that raw and degraded nylon/cotton fibres and their residues exhibited similarly smooth surfaces, further confirming the EG catalytic system's remarkable selective degradation ability. Meanwhile, the degradation of glycosidic bonds in cotton requires a proton donor and a nucleophile/base group;51 however, the EG catalytic system lacks a proton donor, which explains why nylon and cotton were selectively resistant to degradation under mild reaction conditions. Moreover, the saline wastewater generated during TPA acidification was ingeniously repurposed as a buffer solution for cellulases that were used to degrade the remaining cotton after PET degradation in the mixed fabrics. After seven days of enzymatic hydrolysis of waste cotton, a glucose yield of 77.5% was achieved (Fig. S8), thereby further facilitating the optimal utilization of waste resources.

Overall, the EG catalytic system proposed in this study combines the dual functionalities of PET degradation and dye decolorization, enabling the selective degradation of PET fibres from blended textiles. Furthermore, efforts to reduce the use of strong acids and bases and to optimize the decolorization of products will help enhance the environmental sustainability and efficiency of the process. This process facilitates the effective recovery and value-added utilization of waste textiles, making it a promising candidate for industrial-scale textile recycling.

Conclusions

The heterogeneous nature of textile waste and the adherence of dyes present significant challenges for recycling. Herein, in light of the catalytic mechanism of PET hydrolases, we established an eco-friendly and effective EG catalytic system with dual capabilities of selective degradation and decolorization, facilitating the separation, degradation, and product recovery of PET-based mixed textile waste. The EG catalytic system efficiently degrades PET polyester under mild conditions (90 °C, 1 h), achieving a PET conversion of 99.63% and a TPA yield of 95.46%. Moreover, the reaction system can be directly separated and reused at least five times, effectively reducing the consumption of strong acids and bases. Importantly, its excellent degradation performance is consistently maintained, even in a large-scale system (100 L) with a high substrate amount (70 g L−1 PET). This capability successfully extends to real textile waste, highlighting its potential for industrial application. Overall, the EG catalytic system achieves efficient, eco-friendly, and selective PET recycling from blended fabrics, showcasing substantial potential for practical applications in the industry.

Author contributions

Yu Zhou: writing – original draft; conceptualization; investigation; methodology; data curation; formal analysis; and visualization; Jiaxing Zhang: formal analysis; methodology; software; writing – original draft; and writing – review & editing; Bowen Shen: investigation; Wenyan Ba: investigation; Shengping You: project administration; funding acquisition; and writing – review & editing; Mengfan Wang: supervision; Rongxin Su: resources; Wei Qi: supervision; writing – review & editing; and funding acquisition.

Data availability

The data supporting the findings of this article are included in the ESI.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (no. 22278314 and 22078239), Beijing-Tianjin-Hebei Basic Research Cooperation Project (B2021210008), and Beijing Meihao Biotechnology Co., Ltd (2023GKF-0249).

References

  1. H. S. Lee, S. Jung, K.-Y. A. Lin, E. E. Kwon and J. Lee, Sci. Total Environ., 2023, 859, 160393 CrossRef CAS PubMed.
  2. P. K. Mishra, A. M. D. Izrayeel, B. K. Mahur, A. Ahuja and V. K. Rastogi, Environ. Sci. Pollut. Res., 2022, 29, 65962–65977 CrossRef PubMed.
  3. E. J. Cho, Y. G. Lee, Y. Song, H. Y. Kim, D.-T. Nguyen and H.-J. Bae, Environ. Sci. Ecotechnology, 2023, 15, 100238 CrossRef CAS PubMed.
  4. S. Moazzem, L. Wang, F. Daver and E. Crossin, Resour., Conserv. Recycl., 2021, 166, 105338 CrossRef CAS.
  5. K. Subramanian, S. S. Chopra, E. Cakin, X. Li and C. S. K. Lin, Resour., Conserv. Recycl., 2020, 161, 104989 CrossRef.
  6. Y. Li, H. Yi, M. Li, M. Ge and D. Yao, J. Cleaner Prod., 2022, 366, 132985 CrossRef CAS.
  7. E. Barnard, J. J. Rubio Arias and W. Thielemans, Green Chem., 2021, 23, 3765–3789 RSC.
  8. A. F. Sousa, R. Patrício, Z. Terzopoulou, D. N. Bikiaris, T. Stern, J. Wenger, K. Loos, N. Lotti, V. Siracusa, A. Szymczyk, S. Paszkiewicz, K. S. Triantafyllidis, A. Zamboulis, M. S. Nikolic, P. Spasojevic, S. Thiyagarajan, D. S. van Es and N. Guigo, Green Chem., 2021, 23, 8795–8820 RSC.
  9. L. D. Poulikakos, C. Papadaskalopoulou, B. Hofko, F. Gschösser, A. Cannone Falchetto, M. Bueno, M. Arraigada, J. Sousa, R. Ruiz, C. Petit, M. Loizidou and M. N. Partl, Resour., Conserv. Recycl., 2017, 116, 32–44 CrossRef.
  10. N. George and T. Kurian, Ind. Eng. Chem. Res., 2014, 53, 14185–14198 CrossRef CAS.
  11. X. Fei, H. S. Freeman and D. Hinks, J. Cleaner Prod., 2020, 254, 120027 CrossRef CAS.
  12. S. Kumagai, S. Hirahashi, G. Grause, T. Kameda, H. Toyoda and T. Yoshioka, J. Mater. Cycles Waste Manage., 2018, 20, 439–449 CrossRef CAS.
  13. Z. Laldinpuii, S. Lalhmangaihzuala, Z. Pachuau and K. Vanlaldinpuia, Waste Manage., 2021, 126, 1–10 CrossRef CAS PubMed.
  14. A. Khan, M. Naveed, Z. Aayanifard and M. Rabnawaz, Resour., Conserv. Recycl., 2022, 187, 106639 CrossRef CAS.
  15. Q. Hou, M. Zhen, H. Qian, Y. Nie, X. Bai, T. Xia, M. L. U. Rehman, Q. Li and M. Ju, Cell Rep. Phys. Sci., 2021, 2, 100514 CrossRef CAS.
  16. K. P. Sullivan, A. Z. Werner, K. J. Ramirez, L. D. Ellis, J. R. Bussard, B. A. Black, D. G. Brandner, F. Bratti, B. L. Buss, X. Dong, S. J. Haugen, M. A. Ingraham, M. O. Konev, W. E. Michener, J. Miscall, I. Pardo, S. P. Woodworth, A. M. Guss, Y. Román-Leshkov, S. S. Stahl and G. T. Beckham, Science, 2022, 378, 207–211 CrossRef CAS PubMed.
  17. Z. Zhang, S. Huang, D. Cai, C. Shao, C. Zhang, J. Zhou, Z. Cui, T. He, C. Chen, B. Chen and T. Tan, Green Chem., 2022, 24, 5998–6007 RSC.
  18. Y. Peng, J. Yang, C. Deng, J. Deng, L. Shen and Y. Fu, Nat. Commun., 2023, 14, 3249 CrossRef CAS PubMed.
  19. D. Damayanti and H.-S. Wu, Polymers, 2021, 13(9), 1475 CrossRef PubMed.
  20. M. Zhang, Y. Yu, B. Yan, X. Song, Y. Liu, Y. Feng, W. Wu, B. Chen, B. Han and Q. Mei, Appl. Catal., B, 2024, 352, 124055 CrossRef CAS.
  21. S. Haslinger, M. Hummel, A. Anghelescu-Hakala, M. Määttänen and H. Sixta, Waste Manage., 2019, 97, 88–96 CrossRef CAS PubMed.
  22. X. Sun, X. Wang, F. Sun, M. Tian, L. Qu, P. Perry, H. Owens and X. Liu, Adv. Mater., 2021, 33, 2105174 CrossRef CAS PubMed.
  23. S. K. Ramamoorthy, M. Skrifvars, R. Alagar and N. Akhtar, J. Polym. Environ., 2018, 26, 487–498 CrossRef CAS.
  24. L. Amalia, C.-Y. Chang, S. S. S. Wang, Y.-C. Yeh and S.-L. Tsai, Curr. Opin. Biotechnol., 2024, 85, 103053 CrossRef CAS PubMed.
  25. A. Maurya, A. Bhattacharya and S. K. Khare, Front. Bioeng. Biotechnol., 2020, 8, 602325 CrossRef PubMed.
  26. R. Ragg, M. N. Tahir and W. Tremel, Eur. J. Inorg. Chem., 2016, 2016, 1906–1915 CrossRef CAS.
  27. R. K. Brizendine, E. Erickson, S. J. Haugen, K. J. Ramirez, J. Miscall, D. Salvachúa, A. R. Pickford, M. J. Sobkowicz, J. E. McGeehan and G. T. Beckham, ACS Sustainable Chem. Eng., 2022, 10, 9131–9140 CrossRef CAS.
  28. E. Kuah, S. Toh, J. Yee, Q. Ma and Z. Gao, Chem. – Eur. J., 2016, 22, 8404–8430 CrossRef CAS PubMed.
  29. X. Li, Y. Zhou, Z. Lu, R. Shan, D. Sun, J. Li and P. Li, J. Colloid Interface Sci., 2023, 646, 198–208 CrossRef CAS PubMed.
  30. C. Bharadwaj, R. Purbey, D. Bora, P. Chetia, R. U. Maheswari, R. Duarah, K. Dutta, E. R. Sadiku, K. Varaprasad and J. Jayaramudu, Mater. Today Sustainability, 2024, 27, 100936 CrossRef.
  31. H. T. Kim, J. K. Kim, H. G. Cha, M. J. Kang, H. S. Lee, T. U. Khang, E. J. Yun, D.-H. Lee, B. K. Song, S. J. Park, J. C. Joo and K. H. Kim, ACS Sustainable Chem. Eng., 2019, 7, 19396–19406 CrossRef CAS.
  32. M. Zhang, Y. Yu, Z. Wang, S. Zhang, X. Gao, J. Liu, J. Li, W. Wu and Q. Mei, Green Chem., 2024, 26, 11132–11139 RSC.
  33. Y. Cui, Y. Chen, J. Sun, T. Zhu, H. Pang, C. Li, W.-C. Geng and B. Wu, Nat. Commun., 2024, 15, 1417 CrossRef CAS PubMed.
  34. V. Tournier, C. M. Topham, A. Gilles, B. David, C. Folgoas, E. Moya-Leclair, E. Kamionka, M. L. Desrousseaux, H. Texier, S. Gavalda, M. Cot, E. Guémard, M. Dalibey, J. Nomme, G. Cioci, S. Barbe, M. Chateau, I. André, S. Duquesne and A. Marty, Nature, 2020, 580, 216–219 CrossRef CAS PubMed.
  35. Q. Yin, J. Zhang, S. Ma, T. Gu, M. Wang, S. You, S. Ye, R. Su, Y. Wang and W. Qi, Green Chem., 2024, 26, 2560–2570 RSC.
  36. C. Roth, R. Wei, T. Oeser, J. Then, C. Föllner, W. Zimmermann and N. Sträter, Appl. Microbiol. Biotechnol., 2014, 98, 7815–7823 CrossRef CAS PubMed.
  37. R. Wei, G. von Haugwitz, L. Pfaff, J. Mican, C. P. S. Badenhorst, W. Liu, G. Weber, H. P. Austin, D. Bednar, J. Damborsky and U. T. Bornscheuer, ACS Catal., 2022, 12, 3382–3396 CrossRef CAS PubMed.
  38. H. P. Austin, M. D. Allen, B. S. Donohoe, N. A. Rorrer, F. L. Kearns, R. L. Silveira, B. C. Pollard, G. Dominick, R. Duman, K. El Omari, V. Mykhaylyk, A. Wagner, W. E. Michener, A. Amore, M. S. Skaf, M. F. Crowley, A. W. Thorne, C. W. Johnson, H. L. Woodcock, J. E. McGeehan and G. T. Beckham, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, E4350–E4357 CrossRef CAS PubMed.
  39. S. Joo, I. J. Cho, H. Seo, H. F. Son, H.-Y. Sagong, T. J. Shin, S. Y. Choi, S. Y. Lee and K.-J. Kim, Nat. Commun., 2018, 9, 382 CrossRef PubMed.
  40. A. Kushwaha, L. Goswami, M. Singhvi and B. S. Kim, Chem. Eng. J., 2023, 457, 141230 CrossRef CAS.
  41. S. Boneta, K. Arafet and V. Moliner, J. Chem. Inf. Model., 2021, 61, 3041–3051 CrossRef CAS PubMed.
  42. J. Bengtsson, A. Peterson, A. Idström, H. de la Motte and K. Jedvert, Sustainability, 2022, 14, 6911 CrossRef CAS.
  43. S. Ügdüler, K. M. Van Geem, R. Denolf, M. Roosen, N. Mys, K. Ragaert and S. De Meester, Green Chem., 2020, 22, 5376–5394 RSC.
  44. T. Matsumoto, K. Takahashi, K. Kitagishi, K. Shinoda, J. L. Cuya Huaman, J.-Y. Piquemal and B. Jeyadevan, New J. Chem., 2015, 39, 5008–5018 RSC.
  45. L. Hao, R. Wang, Y. Zhao, K. Fang and Y. Cai, Cellulose, 2018, 25, 6759–6769 CrossRef CAS.
  46. C. Fan, L. Zhang, C. Zhu, J. Cao, Y. Xu, P. Sun, G. Zeng, W. Jiang and Q. Zhang, Green Chem., 2022, 24, 1294–1301 RSC.
  47. J. Zhang, Z. Zhang, Y. I. Yang, S. Liu, L. Yang and Y. Q. Gao, ACS Cent. Sci., 2017, 3(5), 407–414 CrossRef CAS PubMed.
  48. S. Yuan, X. Han, J. Zhang, Z. Xie, C. Fan, Y. Xiao, Y. Q. Gao and Y. I. Yang, J. Phys. Chem. A, 2024, 128, 4378–4390 CrossRef CAS PubMed.
  49. Y. Zheng, J. Zhang, S. You, W. Lin, R. Su and W. Qi, Bioresour. Technol., 2024, 406, 130929 CrossRef CAS PubMed.
  50. J. I. Mujika, J. M. Mercero and X. Lopez, J. Am. Chem. Soc., 2005, 127, 4445–4453 CrossRef CAS PubMed.
  51. G. Davies and B. Henrissat, Structure, 1995, 3, 853–859 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5gc00248f

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.