Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 282 word(s) 282 2021-12-27 04:53:22 |
2 Revised + 2288 word(s) 2570 2021-12-30 02:30:56 | |
3 Done + 2288 word(s) 2570 2021-12-30 02:37:33 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Qi, X. Biodegradation of Polyethylene Terephthalate. Encyclopedia. Available online: https://encyclopedia.pub/entry/17639 (accessed on 25 July 2024).
Qi X. Biodegradation of Polyethylene Terephthalate. Encyclopedia. Available at: https://encyclopedia.pub/entry/17639. Accessed July 25, 2024.
Qi, Xinhua. "Biodegradation of Polyethylene Terephthalate" Encyclopedia, https://encyclopedia.pub/entry/17639 (accessed July 25, 2024).
Qi, X. (2021, December 30). Biodegradation of Polyethylene Terephthalate. In Encyclopedia. https://encyclopedia.pub/entry/17639
Qi, Xinhua. "Biodegradation of Polyethylene Terephthalate." Encyclopedia. Web. 30 December, 2021.
Biodegradation of Polyethylene Terephthalate
Edit

Polyethylene terephthalate (PET) is a widely used plastic that is polymerized by terephthalic acid (TPA) and ethylene glycol (EG).

Polyethylene terephthalate Biodegradation Hydrolases Chassis

1. Introduction

Polyethylene terephthalate (PET) is one of the most widely used synthetic plastics in people’s lives [1]. It is polymerized by terephthalic acid (TPA) and ethylene glycol (EG) through ester bonds [2]. Since PET was first used to produce disposable plastic bottles in the 20th century, it has been welcomed worldwide and has become an indispensable part of people’s lives [3]. As PET is highly resistant to natural degradation, the recycling of PET has been encouraged [4]. At present, the main methods for managing PET waste include landfilling, incineration, as well as physical and chemical recycling [5][6]. These methods usually cause secondary pollution to the environment and consume huge amounts of energy, which is not economical or environmentally friendly. Due to the improper recycling strategies and the strong mechanical properties of plastic products, serious environmental problems, such as soil pollution and the disturbance of marine ecosystems, have occurred [7]. Therefore, PET biodegradation has attracted attention as an environmentally friendly alternative, requiring milder temperatures and lower energy consumption than other recycling methods [8][9]. Additionally, the degradation monomers can easily be recycled, with the hope of converting PET into high value chemicals.
In 1977, several commercial lipases and an esterase were reported to hydrolyze various kinds of polyesters [10]. Since then, many PET hydrolases, such as lipases, cutinases and esterases, have been discovered and characterized by various microorganisms [1][11]. In 2016, Ideonella sakaiensis 201-F6 was isolated from a waste recycling station [12]. It was found to produce PET hydrolase (PETase) and monohydroxyethyl terephthalate (MHET) hydrolase (MHETase), which can degrade PET into intermediate products at 30 °C. Then, the structures of the two enzymes were analyzed and a series of effective enzyme modifications were carried out [13][14][15][16][17], efficiently improving the activity and stability of the two enzymes. The discovery and modification of PETase and MHETase has provided an important basis for the degradation of PET waste under ambient temperatures.
Synthetic biology and metabolic engineering strategies have been applied to the biodegradation and bioconversion of PET waste, especially in the modification of PET hydrolases, optimization of microbial chassis, and reconstruction of degradation pathways. At present, some bacteria, fungi, and marine microalgae have been reported as being good microbial chassis for PET biodegradation. The whole-cell biocatalysts have been able to achieve the initial degradation of PET. The bioconversion pathways of TPA and EG have been identified. Some microorganisms have been engineered to produce high value chemicals from PET monomers, which is an important development direction for PET upcycling. Based on these current advances, developing enhanced microbial chassis and constructing artificial microbial consortia to couple the biodegradation of PET by secreted PET hydrolases with the bioconversion of high value chemicals from monomers is a promising method to realize the circular economy of PET waste.

2. PET Biodegradation

During PET biodegradation, the microorganisms first adhere onto the surface of PET films and then secrete extracellular PET hydrolases, which bind to the PET films and initiate the biodegradation process [18][19]. PET hydrolases act on the ester bond of PET, hydrolyzing it into TPA and EG and generating incomplete hydrolysis products, such as MHET and Bis-(2-hydroxyethyl) terephthalate (BHET). In I. sakaiensis 201-F6, MHET can be further hydrolyzed into TPA and EG under the action of MHETase [12]. It was reported that MHETase has a hydrolysis activity against the termini-generated PET film, demonstrating the exo-PETase function of the enzyme [20]. PET hydrolases can further hydrolyze BHET to produce MHET, TPA, and EG [12]. The products TPA and EG can be used by different microorganisms and be further metabolized into the tricarboxylic acid cycle (TCA cycle) [21][22][23][24][25][26][27][28][29]. Additionally, these intermediate and final products of PET biodegradation have been identified as competitive inhibitors of PET hydrolases [30][31].

2.1. Engineered PET Hydrolases

The hydrolases, including lipases [31][32][33][34], cutinases [35][36][37][38][39][40][41][42], esterases [43][44][45][46], PETase [12] and MHETase [12], that can degrade PET have been identified. Among them, lipases have the lowest hydrolysis activity of PET mainly because their catalytic centers are covered by lid structures, which limits the hydrolases’ contact and catalysis with the substrate PET. Cutinases always have a strong PET hydrolysis ability due to their large substrate binding pockets without lid structures, which is conducive to the combination of PET with their active centers. However, cutinases usually degrade PET at high temperatures (50–70 °C), while PETase and MHETase can efficiently and specifically hydrolyze PET at 30 °C [12]. The discovery of PETase and MHETase is helpful in achieving the high efficiency biodegradation of PET at ambient temperatures. At present, the structures of these two enzymes have been studied extensively, and more high activity hydrolases variants have appeared.
The PET hydrolases identified in nature always have poor stability, low activity, and low expression levels, which limit their large-scale industrial application. A series of strategies that could enhance the catalytic activity of PET hydrolases have been proposed [13].
One strategy is to engineer the binding pocket, which can improve the specificity of the PET hydrolases and increase the effective adsorption of enzymes and substrates [15][47][48][49]. The laboratory previously focused on six key amino acids near the binding of PETase to the substrate and conducted site-directed mutations. The R61A, L88F, and I179F mutants were successfully screened, and the enzyme activity increased 1.4-fold, 2.1-fold, and 2.5-fold, respectively, in comparison to wild-type PETase [50]. Silva et al. [51] modified the cutinase from Thermobifida fusca_0883 by site-directed mutagenesis and constructed a single mutation Ile218Ala and a double mutation Q132A/T101A, which expanded the catalytic space and improved the efficiency of the PET biodegradation. Chen et al. [52] identified the unique amino acids S214 and I218 through the structural analysis of PETase and noted that they are associated with W185 wobbling and β6-β7 loop flexibility. This research is helpful in designing PETase mutants that increase the flexibility of the substrate binding pocket.
Some studies focused on using enzyme engineering strategies to improve the stability of the PET hydrolases to improve PET biodegradation efficiency [53][54]. Methods such as adding Ca2+ or Mg2+ [38][55], introducing a disulfide bond and salt bridge [56][57], and glycosylation have all been proven to improve the stability of PET hydrolases. Researchers added disulfide bonds to improve the thermal stability of leaf-branch compost cutinase (LCC) and performed site-directed mutations on hot amino acids near the substrate binding to obtain the combined mutation F243I/D238C/S283C/Y127G (ICCG) [53]. Finally, 90% of shredded PET plastic bottles were degraded at 72 °C for 10 h, which is by far the most efficient PET hydrolase [53].
Additionally, increasing the substrate accessibility for PET hydrolases by engineering the PET hydrolases has also been widely studied [58][59][60]. It is reported that the fusion expression of Thc_Cut1 from Thermobifida cellulosilytica and hydrophobins (HFB4 and HFB7) from Trichoderma reesei can increase the hydrolysis effect of PET by more than 16 times, while a mixture of the enzyme and the hydrophobins led to only a 4-fold increase at most [61].
The intermediate and final products of PET biodegradation, such as Bis-(2-hydroxyethyl) terephthalate (BHET), monohydroxyethyl terephthalate (MHET), TPA, and EG, are all competitive inhibitors of the PET hydrolases [30]. Therefore, the mixtures of hydrolases that act synergistically or protein engineering strategies that reduce the interaction between the enzymes and products are effective methods for solving the inhibition [62][63][64].
In addition, other strategies have been studied to increase the activity of the enzymes and enhance the biodegradation of PET [36][38][57][59][65].

2.2. Engineered PET Biodegradation Chassis

Most of the microorganisms identified that are capable of secreting PET hydrolases are non-model microorganisms and they are difficult to genetically engineer due to their complex genetic background. In addition, the expression level of the PET hydrolases from wild strains is insufficient to satisfy the demand for large-scale degradation. Therefore, it is necessary to develop recombinant expression systems using model microorganisms to express the PET hydrolases efficiently. PET is a high molecular polymer that is polymerized from TPA and EG and cannot enter cells, so in vitro enzymatic degradation of PET has been studied extensively. Owing to the purification and preparation process of PET hydrolases being time-consuming and cost-intensive, the efficient expression PET hydrolases extracellularly for practical applications is necessary [66][67].
At present, some microbial chassis such as bacteria, fungi, and marine microalgae have been applied to the secretion and expression of PET hydrolases, which have been studied and proven to be promising chassis to degrade PET. Several whole-cell biocatalysts have been designed to degrade PET, which are able to not only avoid the complicated steps of enzyme purification but also be reused in multi-step reactions, in comparison to the free enzyme-based approach [68][69]. Additionally, the difficulty of the reduced activity of the enzymes, or even enzymes being inactivated, under the influence of environmental factors has been solved. The following is a summary of several microbial chassis that are suitable for PET biodegradation.

2.2.1. Bacteria

Escherichia coli

E. coli is an important model microorganism for the production of recombinant proteins due to its clear genetic background, simple growth conditions, and its advantages in high density cultivation [70]. In recent years, with the continuous discovery of PET hydrolases, more and more enzymes have achieved the heterologous expression in E. coli [12][14][16][50][53][71]. PET hydrolases heterologously expressed in E. coli have been summarized [66] and it is helpful in further analyzing the crystal structures of these enzymes and explore the degradation mechanism for PET.
Recent studies have shown that engineered E. coli can be used as a whole-cell biocatalyst for PET biodegradation. Selecting the optimal signal peptide is a common strategy used to improve the section of heterologous PET hydrolases. A study tested the effects of Sec-dependent and SRP-dependent signal peptides from E. coli in secreting PETase, and successfully produced 6.2 mg/L PETase by fusing SPLamB and PETase [72]. Some other research improved the expression titer and enzymatic activity by modifying the signal peptide. An evolved signal peptide PelB (G58A) obtained by random mutagenesis was successfully used to express heterologous PETase in E. coli and enabled up to 1.7-fold higher PETase secretion [73]. An enhancer of signal peptides B1 (MERACVAV) was studied to mediate the excretion of PETase, and finally, the excretion efficiency of PETase mediated by B1PelB demonstrated a 62-fold increase over that of PelB [74].

Bacillus subtilis

Gram-positive B. subtilis has the advantages of high secretion capacity, fast growth, and the lack of an outer membrane, and it is regarded as an excellent microbial chassis for secreting heterologous proteins compared to E. coli, which usually forms an inclusion body [75][76]. Additionally, B. subtilis has a strong resistance to harsh environments and it has been used to secrete proteins that can degrade many pollutants, which is why it is considered to be a promising microbial chassis for biodegradation [77][78].
In terms of PET biodegradation, B. subtilis has been engineered to secrete PET hydrolases. It is reported that PETase was successfully secreted into the medium by B. subtilis 168 under the direction of its native signal peptide (SPPETase). SPPETase is predicted to be a twin-arginine signal peptide, and the inactivation of twin-arginine translocation (Tat) complexes improved the secretion amount of PETase 3.8-fold [79]. Another two PET hydrolases (BhrPETase and LCC) were also expressed in B. subtilis, and the expression titer of BhrPETase and LCC reached 0.66 g/L and 0.89 g/L in an engineered chaperone-overexpression of B. subtilis, respectively [42]. Additionally, the combinations of signal peptides and promoters were optimized to promote the expression of PETase in B. subtilis WB600, and the combination of the signal peptide SPamy and the weak promoter P43 was proved to be best [80].

Thermophilic Bacteria

Most of the hydrolases capable of degrading PET, including lipases, cutinases and esterases, have higher enzymatic activity at higher temperatures, while the optimal growth temperature of most model microorganisms that can produce heterologous PET hydrolases is usually 30–40 °C. Whole-cell biocatalyst is not compatible with some PET hydrolases that are only functional at high temperatures [81]. Therefore, a thermophilic expression system is necessary to improve the efficiency of PET biodegradation [36][82]. Most thermophilic microorganisms are usually difficult to genetically engineer except for Clostridium thermocellum, which has a mature genetic manipulation platform [81]. C. thermocellum has been engineered for lignocellulose bioconversion [83] and biofuel production [84], which is why it is regarded as a potential microbial chassis for the biodegradation of PET.
LCC has been successfully obtained from an engineered C. thermocellum. This engineered whole-cell biocatalyst realized a high level expression of LCC and more than 60% of a commercial PET film was converted into soluble monomers at 60 °C after 14 days [85]. This thermophilic whole-cell degradation system has the advantage of simultaneous enzyme production and PET degradation compared to only using free enzymes, which is why it is a promising strategy to degrade PET using other high temperature hydrolases [86][87]. In addition to the thermophilic whole-cell degradation system, an alkali-tolerant whole-cell catalytic system has also been reported [88][89][90].

2.2.2. Fungi

In addition to bacteria, the potential of some yeasts, including Pichia pastoris and Yarrowia lipolytica, being used in PET biodegradation has been studied. P. pastoris, with a great secretion expression system and scalable fermentation capability, has become a common strain for protein production in industrial applications. Researchers have expressed BurPL (H344S/F348I) and PETase in P. pastoris and E. coli and noted that both enzymes produced from P. pastoris showed higher activity than that expressed in E. coli because of the protein half-life protection mechanism of P. pastoris [52][91]. A whole-cell biocatalyst was developed by displaying PETase on the surface of P. pastoris and the enzymatic activity of PETase increased 36-fold towards a highly crystalline PET in comparison to that of purified PETase [92]. Additionally, this whole-cell biocatalyst can be reused seven times without obvious activity loss, which is helpful in developing other whole-cell biocatalysts for PET biodegradation [92]. Considering the ability of P. pastoris to perform N-linked glycosylation, some researchers studied the effects of glycosylation on the LCC expressed in P. pastoris and found that the kinetic stability and activity of LCC were both improved [35]. Y. lipolytica is also a great microbial chassis for bioremediation [93]. Researchers isolated Y. lipolytica IMUFRJ 50682 with the ability to convert PET into MHET and verified that the PET monomers may act as inducers in the process of lipase production [94], which showed that Y. lipolytica is a potential microbial chassis for PET biodegradation. Other research expressed PETase in Y. lipolytica Po1f with a signal peptide from lipase and confirmed that the engineered strain could hydrolyze BHET and PET powder into the monomers [95]. Surface display systems and whole-cell biocatalysts provide novel ideas and strategies for achieving the high efficiency expression of PET hydrolases and promoting PET biodegradation [67][96][97]. Yeasts, together with efficient genetic tools, have been used as great microbial chassis for biodegradation and bioconversion [98].

2.2.3. Marine Microalgae

At present, the existing native and engineered microbial chassis that are capable of producing PET hydrolases are usually difficult to adapt to the complexity of the marine environment and produce much PET waste. Recently, some marine microalgae have been used as chassis for PET biodegradation [99]. A photosynthetic microalga Phaeodactylum tricornutum has been reported as being engineered as a chassis capable of secreting a PETase mutant into the culture medium, and the recombinant PETase was able to efficiently degrade different substrates, including PET films, poly (ethylene terephthalateco-1,4-cylclohexylenedimethylene terephthalate) (PETG) film, and shredded PET, at 30 °C or even at mesophilic temperatures (21 °C) [100]. Additionally, Chlamydomonas reinhardtii, the green algae, was also successfully engineered to produce PETase with degrading activity, and the chemical and morphological changes appeared on the PET films after 4 weeks of culture [101]. As environmentally friendly chassis for the biodegradation of PET waste in a saltwater-based environment, marine microalgae have the potential for future biotechnological applications in the degradation of PET polluted seawater [100].

References

  1. Mohanan, N.; Montazer, Z.; Sharma, P.K.; Levin, D.B. Microbial and enzymatic degradation of synthetic plastics. Front. Microbiol. 2020, 11, 2837.
  2. Webb, H.K.; Arnott, J.; Crawford, R.J.; Ivanova, E.P. Plastic degradation and its environmental implications with special reference to Poly(ethylene terephthalate). Polymers 2013, 5, 1–18.
  3. Koshti, R.; Mehta, L.; Samarth, N. Biological recycling of polyethylene terephthalate: A mini-review. J. Polym. Environ. 2018, 26, 3520–3529.
  4. Taniguchi, I.; Yoshida, S.; Hiraga, K.; Miyamoto, K.; Kimura, Y.; Oda, K. Biodegradation of PET: Current status and application aspects. ACS Catal. 2019, 9, 4089–4105.
  5. Peng, R.; Xia, M.; Ru, J.; Huo, Y.; Yang, Y. Microbial degradation of polyurethane plastics. Chin. J. Biotechnol. 2018, 34, 1398–1409.
  6. Ciobanu, C.S.; Copae, R.; Bulgariu, D.; Bulgariu, L. Comparative study of Pb(II) ions adsorption on pet fibers and flakes: Isotherm, kinetic and mechanism considerations. Desalination Water Treat. 2021, 222, 375–385.
  7. Yang, H.; Chen, G.; Wang, J. Microplastics in the marine environment: Sources, fates, impacts and microbial degradation. Toxics 2021, 9, 41.
  8. Wei, R.; Zimmermann, W. Biocatalysis as a green route for recycling the recalcitrant plastic polyethylene terephthalate. Microb. Biotechnol. 2017, 10, 1302–1307.
  9. Zimmermann, W.; Billig, S. Enzymes for the biofunctionalization of Poly(Ethylene Terephthalate). In Biofunctionalization of Polymers and Their Applications; Advances in Biochemical Engineering—Biotechnology; Nyanhongo, G.S., Steiner, W., Gubitz, G.M., Eds.; Springer: Cham, Switzerland, 2011; Volume 125, pp. 97–120.
  10. Tokiwa, Y.; Suzuki, T. Hydrolysis of polyesters by lipases. Nature 1977, 270, 76–78.
  11. Magalhaes, R.P.; Cunha, J.M.; Sousa, S.F. Perspectives on the role of enzymatic biocatalysis for the degradation of plastic PET. Int. J. Mol. Sci. 2021, 22, 11257.
  12. 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.
  13. Biundo, A.; Ribitsch, D.; Guebitz, G.M. Surface engineering of polyester-degrading enzymes to improve efficiency and tune specificity. Appl. Microbiol. Biotechnol. 2018, 102, 3551–3559.
  14. Palm, G.J.; Reisky, L.; Boettcher, D.; Mueller, H.; Michels, E.A.P.; 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.
  15. Austin, H.P.; Allen, M.D.; Donohoe, B.S.; Rorrer, N.A.; Kearns, F.L.; Silveira, R.L.; Pollard, B.C.; Dominick, G.; Duman, R.; El Omari, K.; et al. Characterization and engineering of a plastic-degrading aromatic polyesterase. Proc. Natl. Acad. Sci. USA 2018, 115, E4350–E4357.
  16. Han, X.; Liu, W.; Huang, J.-W.; Ma, J.; Zheng, Y.; Ko, T.-P.; Xu, L.; Cheng, Y.-S.; Chen, C.-C.; Guo, R.-T. Structural insight into catalytic mechanism of PET hydrolase. Nat. Commun. 2017, 8, 382.
  17. Joo, S.; Cho, I.J.; Seo, H.; Son, H.F.; Sagong, H.-Y.; Shin, T.J.; Choi, S.Y.; Lee, S.Y.; Kim, K.-J. Structural insight into molecular mechanism of poly (ethylene terephthalate) degradation. Nat. Commun. 2018, 9, 2106.
  18. 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.
  19. Jaiswal, S.; Sharma, B.; Shukla, P. Integrated approaches in microbial degradation of plastics. Environ. Technol. Innov. 2020, 17, 100567.
  20. 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.
  21. Ru, J.; Huo, Y.; Yang, Y. Microbial degradation and valorization of plastic wastes. Front. Microbiol. 2020, 11, 442.
  22. Child, J.; Willetts, A. Microbial metabolism of aliphatic glycols. Bacterial metabolism of ethylene glycol. Biochim. Biophys. Acta 1978, 538, 316–327.
  23. Kataoka, M.; Sasaki, M.; Hidalgo, A.; Nakano, M.; Shimizu, S. Glycolic acid production using ethylene glycol-oxidizing microorganisms. Biosci. Biotechnol. Biochem. 2001, 65, 2265–2270.
  24. Wei, G.; Yang, X.; Zhou, W.; Lin, J.; Wei, D. Adsorptive bioconversion of ethylene glycol to glycolic acid by Gluconobacter oxydans DSM 2003. Biochem. Eng. J. 2009, 47, 127–131.
  25. Trifunovic, D.; Schuchmann, K.; Mueller, V. Ethylene glycol metabolism in the acetogen Acetobacterium woodii. J. Bacteriol. 2016, 198, 1058–1065.
  26. Panda, S.; Fung, V.Y.K.; Zhou, J.F.J.; Liang, H.; Zhou, K. Improving ethylene glycol utilization in Escherichia coli fermentation. Biochem. Eng. J. 2021, 168.
  27. Sasoh, M.; Masai, E.; Ishibashi, S.; Hara, H.; Kamimura, N.; Miyauchi, K.; Fukuda, M. Characterization of the terephthalate degradation genes of Comamonas sp. strain E6. Appl. Environ. Microbiol. 2006, 72, 1825–1832.
  28. Choi, K.Y.; Kim, D.; Sul, W.J.; Chae, J.C.; Zylstra, G.J.; Kim, Y.M.; Kim, E. Molecular and biochemical analysis of phthalate and terephthalate degradation by Rhodococcus sp strain DK17. Fems Microbiol. Lett. 2005, 252, 207–213.
  29. Hosaka, M.; Kamimura, N.; Toribami, S.; Mori, K.; Kasai, D.; Fukuda, M.; Masai, E. Novel Tripartite aromatic acid transporter essential for terephthalate uptake in Comamonas sp. Strain E6. Appl. Environ. Microbiol. 2013, 79, 6148–6155.
  30. 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.
  31. Vertommen, M.; Nierstrasz, V.A.; van der Veer, M.; Warmoeskerken, M. Enzymatic surface modification of poly(ethylene terephthalate). J. Biotechnol. 2005, 120, 376–386.
  32. Muller, 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.
  33. Eberl, A.; Heumann, S.; Brueckner, 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.
  34. Nechwatal, A.; Blokesch, A.; Nicolai, M.; Krieg, M.; Kolbe, A.; Wolf, M.; Gerhardt, M. A contribution to the investigation of enzyme-catalysed hydrolysis of poly(ethylene terephthalate) oligomers. Macromol. Mater. Eng. 2006, 291, 1486–1494.
  35. Shirke, A.N.; White, C.; Englaender, J.A.; Zwarycz, A.; Butterfoss, G.L.; Linhardt, R.J.; Gross, R.A. Stabilizing leaf and branch compost Cutinase (LCC) with glycosylation: Mechanism and effect on PET hydrolysis. Biochemistry 2018, 57, 1190–1200.
  36. Acero, E.H.; Ribitsch, D.; Steinkellner, G.; Gruber, K.; Greimel, K.; Eiteljoerg, I.; Trotscha, E.; Wei, R.; Zimmermann, W.; Zinn, M.; et al. Enzymatic surface hydrolysis of PET: Effect of structural diversity on kinetic properties of Cutinases from thermobifida. Macromolecules 2011, 44, 4632–4640.
  37. Kawai, F.; Oda, M.; Tamashiro, T.; Waku, T.; Tanaka, N.; Yamamoto, M.; Mizushima, H.; Miyakawa, T.; Tanokura, M. A novel Ca2+-activated, thermostabilized polyesterase capable of hydrolyzing polyethylene terephthalate from Saccharomonospora viridis AHK190. Appl. Microbiol. Biotechnol. 2014, 98, 10053–10064.
  38. Then, J.; Wei, R.; Oeser, T.; Barth, M.; Belisario-Ferrari, M.R.; Schmidt, J.; Zimmermann, W. Ca2+ and Mg2+ binding site engineering increases the degradation of polyethylene terephthalate films by polyester hydrolases from Thermobifida fusca. Biotechnol. J. 2015, 10, 592–598.
  39. Wei, R.; Breite, D.; Song, C.; Graesing, 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.
  40. Ronkvist, A.M.; Xie, W.; Lu, W.; Gross, R.A. Cutinase-catalyzed hydrolysis of Poly(ethylene terephthalate). Macromolecules 2009, 42, 5128–5138.
  41. Vazquez-Alcantara, L.; Oliart-Ros, R.M.; Garcia-Borquez, A.; Pena-Montes, C. Expression of a Cutinase of Moniliophthora roreri with Polyester and PET-plastic residues degradation activity. Microbiol. Spectr. 2021, 9, e0097621.
  42. Xi, X.; Ni, K.; Hao, H.; Shang, Y.; Zhao, B.; Qian, Z. Secretory expression in Bacillus subtilis and biochemical characterization of a highly thermostable polyethylene terephthalate hydrolase from bacterium HR29. Enzym. Microb. Technol. 2021, 143, 109715.
  43. Billig, S.; Oeser, T.; Birkemeyer, C.; Zimmermann, W. Hydrolysis of cyclic poly(ethylene terephthalate) trimers by a carboxylesterase from Thermobifida fusca KW3. Appl. Microbiol. Biotechnol. 2010, 87, 1753–1764.
  44. Ribitsch, D.; Acero, E.H.; Greimel, K.; Dellacher, A.; Zitzenbacher, S.; Marold, A.; Rodriguez, R.D.; Steinkellner, G.; Gruber, K.; Schwab, H.; et al. A new esterase from Thermobifida halotolerans hydrolyses polyethylene terephthalate (PET) and polylactic acid (PLA). Polymers 2012, 4, 617–629.
  45. Biundo, A.; Reich, J.; Ribitsch, D.; Guebitz, G.M. Synergistic effect of mutagenesis and truncation to improve a polyesterase from Clostridium botulinum for polyester hydrolysis. Sci. Rep. 2018, 8, 3745.
  46. Biundo, A.; Braunschmid, V.; Pretzler, M.; Kampatsikas, I.; Darnhofer, B.; Birner-Gruenberger, R.; Rompel, A.; Ribitsch, D.; Guebitz, G.M. Polyphenol oxidases exhibit promiscuous proteolytic activity. Commun. Chem. 2020, 3, 62.
  47. Wei, R.; Oeser, T.; Barth, M.; Weigl, N.; Luebs, A.; Schulz-Siegmund, M.; Hacker, M.C.; Zimmermann, W. Turbidimetric analysis of the enzymatic hydrolysis of polyethylene terephthalate nanoparticles. J. Mol. Catal. B-Enzym. 2014, 103, 72–78.
  48. 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.
  49. Liu, B.; He, L.; Wang, L.; Li, T.; Li, C.; Liu, H.; Luo, Y.; Bao, R. Protein crystallography and site-direct mutagenesis analysis of the Poly(ethylene terephthalate) hydrolase PETase from Ideonella sakaiensis. Chembiochem 2018, 19, 1471–1475.
  50. Ma, Y.; Yao, M.D.; Li, B.Z.; Ding, M.Z.; He, B.; Chen, S.; Zhou, X.; Yuan, Y.J. Enhanced Poly(ethylene terephthalate) hydrolase activity by protein engineering. Engineering 2018, 4, 888–893.
  51. Silva, C.; Da, S.; Silva, N.; Matama, T.; Araujo, R.; Martins, M.; Chen, S.; Chen, J.; Wu, J.; Casal, M.; et al. Engineered Thermobifida fusca cutinase with increased activity on polyester substrates. Biotechnol. J. 2011, 6, 1230–1239.
  52. Chen, C.C.; Han, X.; Li, X.; Jiang, P.C.; Niu, D.; Ma, L.X.; Liu, W.D.; Li, S.Y.; Qu, Y.Y.; Hu, H.B.; et al. General features to enhance enzymatic activity of poly(ethylene terephthalate) hydrolysis. Nat. Catal. 2021, 4, 425–430.
  53. Tournier, V.; Topham, C.M.; Gilles, A.; David, B.; Folgoas, C.; Moya-Leclair, E.; Kamionka, E.; Desrousseaux, M.L.; Texier, H.; Gavalda, S.; et al. An engineered PET depolymerase to break down and recycle plastic bottles. Nature 2020, 580, 216–219.
  54. Cui, Y.; Chen, Y.; Liu, X.; Dong, S.; Tian, Y.E.; Qiao, Y.; Mitra, R.; Han, J.; Li, C.; Han, X.; et al. Computational redesign of a PETase for plastic biodegradation under ambient condition by the GRAPE strategy. Acs Catal. 2021, 11, 1340–1350.
  55. Thumarat, U.; Nakamura, R.; Kawabata, T.; Suzuki, H.; Kawai, F. Biochemical and genetic analysis of a cutinase-type polyesterase from a thermophilic Thermobifida alba AHK119. Appl. Microbiol. Biotechnol. 2012, 95, 419–430.
  56. Then, J.; Wei, R.; Oeser, T.; Gerdts, A.; Schmidt, J.; Barth, M.; Zimmermann, W. A disulfide bridge in the calcium binding site of a polyester hydrolase increases its thermal stability and activity against polyethylene terephthalate. Febs Open Bio 2016, 6, 425–432.
  57. Shirke, A.N.; Basore, D.; Holton, S.; Su, A.; Baugh, E.; Butterfoss, G.L.; Makhatadze, G.; Bystroff, C.; Gross, R.A. Influence of surface charge, binding site residues and glycosylation on Thielavia terrestris cutinase biochemical characteristics. Appl. Microbiol. Biotechnol. 2016, 100, 4435–4446.
  58. Linder, M.B.; Szilvay, G.R.; Nakari-Setala, T.; Penttila, M.E. Hydrophobins: The protein-amphiphiles of filamentous fungi. Fems Microbiol. Rev. 2005, 29, 877–896.
  59. Puspitasari, N.; Tsai, S.-L.; Lee, C.-K. Fungal Hydrophobin RolA enhanced PETase hydrolysis of polyethylene terephthalate. Appl. Biochem. Biotechnol. 2021, 193, 1284–1295.
  60. Ribitsch, D.; Yebra, A.O.; Zitzenbacher, S.; Wu, J.; Nowitsch, S.; Steinkellner, G.; Greimel, K.; Doliska, A.; Oberdorfer, G.; Gruber, C.C.; et al. Fusion of binding domains to Thermobifida cellulosilytica Cutinase to tune sorption characteristics and enhancing PET hydrolysis. Biomacromolecules 2013, 14, 1769–1776.
  61. Ribitsch, D.; Acero, E.H.; Przylucka, A.; Zitzenbacher, S.; Marold, A.; Gamerith, C.; Tscheliessnig, R.; Jungbauer, A.; Rennhofer, H.; Lichtenegger, H.; et al. Enhanced cutinase-catalyzed hydrolysis of polyethylene terephthalate by covalent fusion to hydrophobins. Appl. Environ. Microbiol. 2015, 81, 3586–3592.
  62. Carniel, A.; Valoni, E.; Nicomedes Junior, J.; Gomes, A.d.C.; de Castro, A.M. Lipase from Candida antarctica (CALB) and cutinase from Humicola insolens act synergistically for PET hydrolysis to terephthalic acid. Process Biochem. 2017, 59, 84–90.
  63. Wei, R.; Oeser, T.; Schmidt, J.; Meier, R.; Barth, M.; Then, J.; Zimmermann, W. Engineered bacterial polyester hydrolases efficiently degrade polyethylene terephthalate due to relieved product inhibition. Biotechnol. Bioeng. 2016, 113, 1658–1665.
  64. Zhou, Y.J.J.; Gao, W.; Rong, Q.X.; Jin, G.J.; Chu, H.Y.; Liu, W.J.; Yang, W.; Zhu, Z.W.; Li, G.H.; Zhu, G.F.; et al. Modular Pathway Engineering of diterpenoid synthases and the mevalonic acid pathway for miltiradiene production. J. Am. Chem. Soc. 2012, 134, 3234–3241.
  65. Furukawa, M.; Kawakami, N.; Oda, K.; Miyamoto, K. Acceleration of enzymatic degradation of Poly(ethylene terephthalate) by surface coating with anionic surfactants. Chemsuschem 2018, 11, 4018–4025.
  66. Samak, N.A.; Jia, Y.; Sharshar, M.M.; Mu, T.; Yang, M.; Peh, S.; Xing, J. Recent advances in biocatalysts engineering for polyethylene terephthalate plastic waste green recycling. Environ. Int. 2020, 145, 106144.
  67. Gao, R.; Pan, H.; Lian, J. Recent advances in the discovery, characterization, and engineering of poly (ethylene terephthalate) (PET) hydrolases. Enzym. Microb. Technol. 2021, 150, 109868.
  68. Serra, S.; De Simeis, D. Use of Lactobacillus rhamnosus (ATCC 53103) as whole-cell biocatalyst for the regio- and stereoselective hydration of oleic, linoleic, and linolenic acid. Catalysts 2018, 8, 109.
  69. Polak, J.; Jarosz-Wilkolazka, A. Whole-cell fungal transformation of precursors into dyes. Microb. Cell Factories 2010, 9, 51.
  70. Touchon, M.; Hoede, C.; Tenaillon, O.; Barbe, V.; Baeriswyl, S.; Bidet, P.; Bingen, E.; Bonacorsi, S.; Bouchier, C.; Bouvet, O.; et al. Organised genome dynamics in the escherichia coli species results in highly diverse adaptive paths. PLoS Genet. 2009, 5, e1000344.
  71. Bollinger, A.; Thies, S.; Knieps-Gruenhagen, E.; Gertzen, C.; Kobus, S.; Hoeppner, A.; Ferrer, M.; Gohlke, H.; Smits, S.H.J.; Jaeger, K.-E. A novel polyester hydrolase from the marine bacterium Pseudomonas aestusnigri—Structural and functional insights. Front. Microbiol. 2020, 11, 114.
  72. 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.
  73. Shi, L.; Liu, H.; Gao, S.; Weng, Y.; Zhu, L. Enhanced extracellular production of IsPETase in Escherichia coli via engineering of the pelB signal peptide. J. Agric. Food Chem. 2021, 69, 2245–2252.
  74. Cui, L.; Qiu, Y.; Liang, Y.; Du, C.; Dong, W.; Cheng, C.; He, B. Excretory expression of IsPETase in E. coli by an enhancer of signal peptides and enhanced PET hydrolysis. Int. J. Biol. Macromol. 2021, 188, 568–575.
  75. Van Dijl, J.M.; Hecker, M. Bacillus subtilis: From soil bacterium to super-secreting cell factory. Microb. Cell Factories 2013, 12, 3.
  76. Westers, L.; Westers, H.; Quax, W.J. Bacillus subtilis as cell factory for pharmaceutical proteins: A biotechnological approach to optimize the host organism. Biochim. Biophys. Acta Mol. Cell Res. 2004, 1694, 299–310.
  77. Huang, K.; Chen, C.; Shen, Q.; Rosen, B.P.; Zhao, F.-J. Genetically Engineering bacillus subtilis with a heat-resistant arsenite methyltransferase for bioremediation of arsenic-contaminated organic waste. Appl. Environ. Microbiol. 2015, 81, 6718–6724.
  78. Yang, C.; Song, C.; Freudl, R.; Mulchandani, A.; Qiao, C. Twin-arginine translocation of methyl parathion hydrolase in Bacillus subtilis. Environ. Sci. Technol. 2010, 44, 7607–7612.
  79. Huang, X.; Cao, L.; Qin, Z.; Li, S.; Kong, W.; Liu, Y. Tat-independent secretion of polyethylene terephthalate hydrolase PETase in Bacillus subtilis 168 mediated by its native signal peptide. J. Agric. Food Chem. 2018, 66, 13217–13227.
  80. Wang, N.; Guan, F.; Lv, X.; Han, D.; Zhang, Y.; Wu, N.; Xia, X.; Tian, J. Enhancing secretion of polyethylene terephthalate hydrolase PETase in Bacillus subtilis WB600 mediated by the SP(amy)signal peptide. Lett. Appl. Microbiol. 2020, 71, 235–241.
  81. Ng, T.K.; Weimer, T.K.; Zeikus, J.G. Cellulolytic and physiological properties of Clostridium thermocellum. Arch. Microbiol. 1977, 114, 1–7.
  82. Guyot, S.; Pottier, L.; Hartmann, A.; Ragon, M.; Tiburski, J.H.; Molin, P.; Ferret, E.; Gervais, P. Extremely rapid acclimation of Escherichia coli to high temperature over a few generations of a fed-batch culture during slow warming. Microbiologyopen 2014, 3, 52–63.
  83. Lin, P.P.; Mi, L.; Moriok, A.H.; Yoshino, M.M.; Konishi, S.; Xu, S.C.; Papanek, B.A.; Riley, L.A.; Guss, A.M.; Liao, J.C. Consolidated bioprocessing of cellulose to isobutanol using Clostridium thermocellum. Metab. Eng. 2015, 31, 44–52.
  84. Argyros, D.A.; Tripathi, S.A.; Barrett, T.F.; Rogers, S.R.; Feinberg, L.F.; Olson, D.G.; Foden, J.M.; Miller, B.B.; Lynd, L.R.; Hogsett, D.A.; et al. High ethanol titers from cellulose by using metabolically engineered thermophilic, anaerobic microbes. Appl. Environ. Microbiol. 2011, 77, 8288–8294.
  85. Yan, F.; Wei, R.; Cui, Q.; Bornscheuer, U.T.; Liu, Y.-J. Thermophilic whole-cell degradation of polyethylene terephthalate using engineered Clostridium thermocellum. Microb. Biotechnol. 2021, 14, 374–385.
  86. Sooch, B.S.; Kauldhar, B.S.; Puri, M. Isolation and polyphasic characterization of a novel hyper catalase producing thermophilic bacterium for the degradation of hydrogen peroxide. Bioprocess Biosyst. Eng. 2016, 39, 1759–1773.
  87. Rudolph, B.; Gebendorfer, K.M.; Buchner, J.; Winter, J. Evolution of Escherichia coli for growth at high temperatures. J. Biol. Chem. 2010, 285, 19029–19034.
  88. Zhang, J.F.; Wang, X.C.; Gong, J.X.; Gu, Z.Y. A study on the biodegradability of polyethylene terephthalate fiber and diethylene glycol terephthalate. J. Appl. Polym. Sci. 2004, 93, 1089–1096.
  89. Gong, J.; Duan, N.; Zhao, X. Evolutionary engineering of Phaffia rhodozyma for astaxanthin-overproducing strain. Front. Chem. Sci. Eng. 2012, 6, 174–178.
  90. Gong, J.; Kong, T.; Li, Y.; Li, Q.; Li, Z.; Zhang, J. Biodegradation of microplastic derived from Poly(ethylene terephthalate) with bacterial whole-cell biocatalysts. Polymers 2018, 10, 1326.
  91. Xu, Z.; Cen, Y.-K.; Zou, S.-P.; Xue, Y.-P.; Zheng, Y.-G. Recent advances in the improvement of enzyme thermostability by structure modification. Crit. Rev. Biotechnol. 2020, 40, 83–98.
  92. Chen, Z.Z.; Wang, Y.Y.; Cheng, Y.Y.; Wang, X.; Tong, S.W.; Yang, H.T.; Wang, Z.F. Efficient biodegradation of highly crystallized polyethylene terephthalate through cell surface display of bacterial PETase. Sci. Total Environ. 2020, 709, 136138.
  93. Madzak, C. Yarrowia lipolytica: Recent achievements in heterologous protein expression and pathway engineering. Appl. Microbiol. Biotechnol. 2015, 99, 4559–4577.
  94. Da Costa, A.M.; de Oliveira Lopes, V.R.; Vidal, L.; Nicaud, J.-M.; de Castro, A.M.; Zarur Coelho, M.A. Poly(ethylene terephthalate) (PET) degradation by Yarrowia lipolytica: Investigations on cell growth, enzyme production and monomers consumption. Process Biochem. 2020, 95, 81–90.
  95. Liu, P.; Zhang, T.; Zheng, Y.; Li, Q.; Su, T.; Qi, Q. Potential one-step strategy for PET degradation and PHB biosynthesis through co-cultivation of two engineered microorganisms. Eng. Microbiol. 2021, 1, 100003.
  96. Lebreton, S.; Zurzolo, C.; Paladino, S. Organization of GPI-anchored proteins at the cell surface and its physiopathological relevance. Crit. Rev. Biochem. Mol. Biol. 2018, 53, 403–419.
  97. Yang, C.-E.; Chu, I.M.; Wei, Y.-H.; Tsai, S.-L. Surface display of synthetic phytochelatins on Saccharomyces cerevisiae for enhanced ethanol production in heavy metal-contaminated substrates. Bioresour. Technol. 2017, 245, 1455–1460.
  98. Beopoulos, A.; Cescut, J.; Haddouche, R.; Uribelarrea, J.-L.; Molina-Jouve, C.; Nicaud, J.-M. Yarrowia lipolytica as a model for bio-oil production. Progress Lipid Res. 2009, 48, 375–387.
  99. Barone, G.D.; Ferizovic, D.; Biundo, A.; Lindblad, P. Hints at the applicability of Microalgae and Cyanobacteria for the biodegradation of plastics. Sustainability 2020, 12, 449.
  100. Moog, D.; Schmitt, J.; Senger, J.; Zarzycki, J.; Rexer, K.-H.; Linne, U.; Erb, T.; Maier, U.G. Using a marine microalga as a chassis for polyethylene terephthalate (PET) degradation. Microb. Cell Factories 2019, 18, 171.
  101. Kim, J.W.; Park, S.-B.; Tran, Q.-G.; Cho, D.-H.; Choi, D.-Y.; Lee, Y.J.; Kim, H.-S. Functional expression of polyethylene terephthalate-degrading enzyme (PETase) in green microalgae. Microb. Cell Factories 2020, 19, 97.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 1.8K
Entry Collection: Environmental Sciences
Revisions: 3 times (View History)
Update Date: 30 Dec 2021
1000/1000
Video Production Service